Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-28T21:00:12.824Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  20 April 2017

George E. Heimpel
Affiliation:
University of Minnesota
Nicholas J. Mills
Affiliation:
University of California, Berkeley
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
Biological Control
Ecology and Applications
, pp. 269 - 365
Publisher: Cambridge University Press
Print publication year: 2017

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abell, KJ & Van Driesche, RG (2012) Impact of latitude on synchrony of a scale (Fiorinia externa) (Hemiptera: Diaspididae) and its parasitoid (Encarsia citrina) (Hymenoptera: Aphelinidae) in the eastern United States. Biological Control 63: 339347.Google Scholar
Abrams, PA (2015) Why ratio dependence is (still) a bad model of predation. Biological Reviews 90: 794814.Google Scholar
Abrams, PA & Ginzburg, LR (2000) The nature of predation: prey dependent, ratio dependent, or neither? Trends in Ecology & Evolution 15: 337341.Google Scholar
Abrantes, J, Carmo, CR, Matthee, CA, Yamada, F, Loo, W & Esteves, P (2011) A shared unusual genetic change at the chemokine receptor type 5 between Oryctolagus, Bunolagus and Pentalagus. Conservation Genetics 12: 325330.Google Scholar
Abu-Dieyeh, MH & Watson, AK (2007) Grass overseeding and a fungus combine to control Taraxacum officinale. Journal of Applied Ecology 44: 115124.CrossRefGoogle Scholar
Acheampong, S & Stark, JD (2004) Can reduced rates of pymetrozine and natural enemies control the cabbage aphid, Brevicoryne brassicae (Homoptera: Aphididae) on broccoli? International Journal of Pest Management 50: 275279.Google Scholar
Acorn, J (2007) Ladybugs of Alberta. Edmonton, University of Alberta Press.Google Scholar
Adachi-Hagimori, T, Miura, K & Stouthamer, R (2008) A new cytogenetic mechanism for bacterial endosymbiont-induced parthenogenesis in Hymenoptera. Proceedings of the Royal Society B-Biological Sciences 275: 26672673.Google Scholar
Adams, JM, Fang, W, Callaway, RM, Cipollini, D & Newell, E (2009) A cross-continental test of the enemy release hypothesis: leaf herbivory on Acer platanoides (L.) is three times lower in North America than in its native Europe. Biological Invasions 11: 10051016.Google Scholar
Adams, PB (1990) The potential of mycoparasites for biological control of plant diseases. Annual Review of Phytopathology 28: 5972.Google Scholar
Agnello, AM, Reissig, WH, Kovach, J & Nyrop, JP (2003) Integrated apple pest management in New York State using predatory mites and selective pesticides. Agriculture Ecosystems & Environment 94: 183195.Google Scholar
Agrawal, AA & Kotanen, PM (2003) Herbivores and the success of exotic plants: a phylogenetically controlled experiment. Ecology Letters 6: 712715.CrossRefGoogle Scholar
Agrawal, AA, Kotanen, PM, Mitchell, CE, Power, AG, Godsoe, W & Klironomos, J (2005) Enemy release? An experiment with congeneric plant pairs and diverse above- and below ground enemies. Ecology 86: 29792989.Google Scholar
Agricola, U & Fisher, HU (1991) Hyperparasitism in two newly introduced parasitoids, Epidinocarsis lopezi and Gyranosoidea tebygi (Hymenoptera: Encyrtidae) after their establishment in Togo. Bulletin of Entomological Research 81: 127132.CrossRefGoogle Scholar
Ainsworth, N (2003) Integration of herbicides with arthropod biocontrol agents for weed control. Biocontrol Science and Technology 13: 547570.CrossRefGoogle Scholar
Alabouvette, C, Heilig, U & Cordier, C (2012) Microbial control of plant diseases. In Beneficial Microorganisms in Agriculture, Food and the Environment (Sundh, I, Wilcks, A & Goettel, M, eds.) Wallingford, UK, CABI Publishing, pp. 96111.Google Scholar
Alhmedi, A, Haubruge, E, D’Hoedt, S & Francis, F (2011) Quantitative food webs of herbivore and related beneficial community in non-crop and crop habitats. Biological Control 58: 103112.Google Scholar
Ali, AD & Reagan, TE (1985) Vegetation manipulation impact on predator and prey populations in Louisiana sugarcane ecosystems. Journal of Economic Entomology 78: 14091414.Google Scholar
Allen, SK, Thiery, RG & Hagstrom, NT (1986) Cytological evaluation of the likelihood that triploid grass carp will reproduce. Transactions of the American Fisheries Society 115: 841848.Google Scholar
Alphey, L (2014) Genetic control of mosquitoes. Annual Review of Entomology 59: 205224.Google Scholar
Alstad, DN & Edmonds, JGF (1983) Selection, outbreeding depression, and the sex ratio of scale insects. Science 220: 9395.Google Scholar
Altfeld, L & Stiling, P (2009) Effects of aphid-tending Argentine ants, nitrogen enrichment and early-season herbivory on insects hosted by a coastal shrub. Biological Invasions 11: 183191.CrossRefGoogle Scholar
Althoff, DM (2003) Does parasitoid attack strategy influence host specificity? A test with new world braconids. Ecological Entomology 28: 500502.CrossRefGoogle Scholar
Altieri, MA & Letourneau, DK (1982) Vegetation management and biological control in agroecosystems. Crop Protection 1: 405430.Google Scholar
Altieri, MA & Whitcomb, WH (1979) The potential use of weeds in the manipulation of beneficial insects. HortScience 14: 1218.Google Scholar
Altieri, MA & Whitcomb, WH (1980) Weed manipulation for insect pest management in corn. Environmental Management 4: 483489.CrossRefGoogle Scholar
Alyokhin, A & Sewell, G (2004) Changes in a lady beetle community following the establishment of three alien species. Biological Invasions 6: 463471.Google Scholar
Amaral, DS, Venzon, M, Pallini, A, Lima, PC & DeSouza, O (2010) Does vegetational diversification reduce coffee leaf miner Leucoptera coffeella (Guerin-Meneville) (Lepidoptera: Lyonetiidae) attack? Neotropical Entomology 39: 543548.CrossRefGoogle ScholarPubMed
Amsellem, Z, Cohen, BA & Gressel, J (2002) Engineering hypervirulence in a mycoherbicidal fungus for efficient weed control. Nature Biotechnology 20: 10351039.Google Scholar
Anagnostakis, SL (1982) Biological control of chestnut blight. Science 215: 466471.Google Scholar
Andersen, MC, Ewald, M & Northcott, J (2005) Risk analysis and management decisions for weed biological control agents: ecological theory and modeling results. Biological Control 35: 330337.CrossRefGoogle Scholar
Anderson, G, Delfosse, ES, Spencer, N, Prosser, C & Richard, R (2000) Biological control of leafy spurge: an emerging success story. In Proceedings of the X International Symposium on Biological Control of Weeds (Spencer, NR, ed.), Bozeman, Montana State University, pp. 1525.Google Scholar
Anderson, RM & May, RM (1978) Regulation and stability of host-parasite population interactions. 1. Regulatory processes. Journal of Animal Ecology 47: 219247.CrossRefGoogle Scholar
Anderson, RM & May, RM (1981) The population dynamics of microparasites and their invertebrate hosts. Philosophical Transactions of the Royal Society of London, Series B 291: 451524.Google Scholar
Anderson, RM & May, RM (1982) Coevolution of hosts and parasites. Parasitology 85: 411426.Google Scholar
Andow, DA (1990) Population dynamics of an insect herbivore in simple and diverse habitats. Ecology 71: 10061017.Google Scholar
Andow, DA (1991) Vegetational diversity and arthropod population response. Annual Review of Entomology 36: 561586.Google Scholar
Andow, DA & Hilbeck, A (2004) Science-based risk assessment for non-target effects of transgenic crops. BioScience 54: 637649.Google Scholar
Andow, DA & Imura, O (1994) Specialization of phytophagous arthropod communities on introduced plants. Ecology 75: 296300.CrossRefGoogle Scholar
Andow, DA, Lane, CP & Olson, DM (1995) Use of Trichogramma in maize – estimating environmental risks. In Biological Control: Benefits and Risks (Hokkanen, HMT & Lynch, JM, eds.), Cambridge, UK, Cambridge University Press, pp. 101118.Google Scholar
Andow, DA & Prokrym, DR (1990) Plant structural complexity and host-finding by a parasitoid. Oecologia 82: 162165.CrossRefGoogle ScholarPubMed
Andres, LA & Rees, NE (1995) Musk thistle. In Biological Control in the Western United States (Nechols, JR, ed.), Oakland, University of California Press, pp. 248251.Google Scholar
Andrewartha, HG & Birch, LC (1954) The Distribution and Abundance of Animals. Chicago, University of Chicago Press.Google Scholar
Angalet, GW & Fuester, R (1977) The Aphidius parasites of the pea aphid Acyrthosiphon pisum in the eastern half of the United States. Annals of the Entomological Society of America 70: 8796.Google Scholar
Antolin, MF (1999) A genetic perspective on mating systems and sex ratios of parasitoid wasps. Researches on Population Ecology 41: 2937.Google Scholar
Antolin, MF, Bjorksten, TA & Vaughn, TT (2006) Host-related fitness trade-offs in a presumed generalist parasitoid, Diaeretiella rapae (Hymenoptera: Aphidiidae). Ecological Entomology, 31: 242254.Google Scholar
Antolin, MF & Strand, MR (1992) Mating system of Bracon hebetor (Hymenoptera: Braconidae). Ecological Entomology 17: 17.CrossRefGoogle Scholar
Antonovics, J (1992) Toward community genetics. In Plant Resistance to Herbivores and Pathogens: Ecology, Evolution, and Genetics (Fritz, RS & Simms, EL, eds.), Chicago, University of Chicago Press, pp. 426449.Google Scholar
Araj, S-E, Wratten, S, Lister, A & Buckley, H (2009) Adding floral nectar resources to improve biological control: potential pitfalls of the fourth trophic level. Basic and Applied Ecology 10: 554562.Google Scholar
Arditi, R (1983) A unified model of the functional response of predators and parasitoids. Journal of Animal Ecology 52: 293303.Google Scholar
Arditi, R & Akcakaya, HR (1990) Underestimation of mutual interference in predators. Oecologia 83: 358361.Google Scholar
Arditi, R & Ginzburg, LR (1989) Coupling in predator–prey dynamics: ratio-dependence. Journal of Theoretical Biology 139: 311326.Google Scholar
Arditi, R & Ginzburg, LR (2012) How Species Interact: Altering the Standard View on Trophic Ecology. New York, Oxford University Press.Google Scholar
Argaud, O, Croizier, L, Lopez-Ferber, M & Croizier, G (1998) Two key mutations in the host-range specificity domain of the p143 gene of Autographa californica nucleopolyhedrovirus are required to kill Bombyx mori larvae. Journal of General Virology 79: 931935.Google Scholar
Armbrecht, I & Gallego, MC (2007) Testing ant predation on the coffee berry borer in shaded and sun coffee plantations in Colombia. Entomologia Experimentalis et Applicata 124: 261267.Google Scholar
Armbruster, P, Bradshaw, WE & Holzapfel, CM (1997) Evolution of the genetic architecture underlying fitness in the pitcher-plant mosquito, Wyeomyia smithii. Evolution 51: 451458.Google Scholar
Ashburner, M, Hoy, MA & Peloquin, JJ (1998) Prospects for the genetic transformation of arthropods. Insect Molecular Biology 7: 201213.Google Scholar
Asano, S & Miyamoto, K (2007) A laboratory method to evaluate the effectiveness of ultraviolet (UV) protectant for Bacillus thuringiensis product. Japanese Journal of Applied Entomology and Zoology 51: 121127.Google Scholar
Askew, RR (1994) Parasitoids of leaf-mining Lepidoptera: what determines their host ranges? Parasitoid Community Ecology (Hawkins, BA & Sheehan, W, eds.), Oxford, UK, Oxford University Press, pp. 177202.Google Scholar
Askew, RR & Shaw, MR (1986) Parasitoid communities: their size, structure and development. Insect Parasitoids (Waage, JK & Greathead, D, eds.), London, UK, Academic Press, pp. 225264.Google Scholar
Aspi, J (2000) Inbreeding and outbreeding depression in male courtship song characters in Drosophila montana. Heredity 84: 273282.Google Scholar
Asplen, MK, Chacon, JM & Heimpel, GE (2016) Divergent sex-specific dispersal by a parasitoid wasp in the field. Entomologia Experimentalis et Applicata 159: 252259.Google Scholar
Asplen, MK, Whitfield, JB, de Boer, JG & Heimpel, GE (2009) Is single-locus complementary sex determination the ancestral mechanism for hymenopteran haplodiploidy? Journal of Evolutionary Biology 22: 17621769.Google Scholar
Auerbach, M & Simberloff, D (1988) Rapid leaf-miner colonization of introduced trees and shifts in sources of herbivore mortality. Oikos 52: 4150.Google Scholar
Auslander, DM, Oster, GF & Huffaker, CB (1974) Dynamics of interacting populations. Journal of the Franklin Institute 297: 345376.CrossRefGoogle Scholar
Babendreier, D, Kuske, S & Bigler, F (2003) Non-target host acceptance and parasitism by Trichogramma brassicae Bezdenko (Hymenoptera: Trichogrammatidae) in the laboratory. Biological Control 26: 128138.Google Scholar
Babendreier, D, Bigler, F & Kuhlmann, U (2006) Current status and constraints in the assessment of non-target effects. Environmental Impact of Invertebrates for Biological Control of Arthropods: Methods and Risk Assessment (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 114.Google Scholar
Baek, J-M, Howell, CR & Kenerley, CM (1999) The role of an extracellular chitinase from Trichoderma virens Gv29–8 in the biocontrol of Rhizoctonia solani. Current Genetics 35: 4150.Google Scholar
Baer, CF, Tripp, DW, Bjorksten, TA & Antolin, MF (2004) Phylogeography of a parasitoid wasp (Diaeretiella rapae): no evidence of host-associated lineages. Molecular Ecology 13: 18591869.Google Scholar
Baggen, LR & Gurr, GM (1998) The influence of food on Copidosoma koehleri (Hymenoptera: Encyrtidae), on the use of flowering plants as a habitat management tool to enhance biological control of potato moth, Phythorimaea operculella (Lepidoptera: Gelechiidae). Biological Control 11: 917.Google Scholar
Baggen, LR, Gurr, GM & Meats, A (1999) Flowers in tri-trophic systems: mechanisms allowing selective exploitation by insect natural enemies for conservation biological control. Entomologia Experimentalis et Applicata 91: 155161.Google Scholar
Bai, C, Shapiro-Ilan, DI, Gaugler, R & Hopper, KR (2005) Stabilization of beneficial traits in Heterorhabditis bacteriophora through creation of inbred lines. Biological Control 32: 220227.Google Scholar
Bailey, EP (1992) Red foxes, Vulpes vulpes, as biological control agents for introduced arctic foxes, Alopex lagopus, on Alaskan islands. Canadian Field-Naturalist 106: 200205.Google Scholar
Baker, DA, Loxdale, HD & Edwards, OR (2003). Genetic variation and founder effects in the parasitoid wasp, Diaeretiella rapae (M’intosh) (Hymenoptera: Braconidae: Aphidiinae), affecting its potential as a biological control agent. Molecular Ecology 12: 33033311.Google Scholar
Baker, KF (1987) Evolving concepts of the biological control of plant pathogens. Annual Review of Plant Pathology 25:6785.Google Scholar
Baker, KF & Cook, RJ 1974. Biological Control of Plant Pathogens. San Francisco, CA, Freeman.Google Scholar
Bakker, PAHM, Pieterse, CMJ & Van Loon, LC (2007) Induced systemic resistance by fluorescent Pseudomonas spp. Phytopathology 97: 239243.Google Scholar
Bakker, PAHM, Ran, LX, Pieterse, CMJ & Van Loon, LC (2003) Understanding the involvement of rhizobacteria-mediated induction of systemic resistance in biocontrol of plant diseases. Canadian Journal of Plant Pathology 25: 59.Google Scholar
Balciunas, JK (2000) Code of best practices for classical biological control of weeds. In Proceedings of the X International Symposium on Biological Control of Weeds (Spencer, NR, ed.), Bozeman, Montana State University, p. 435.Google Scholar
Balciunas, JK (2004a) Are mono-specific agents necessarily safe? The need for pre-release assessment of probable impact of candidate biocontrol agents, with some examples. In Proceedings of the XI Symposium on Biological Control of Weeds (Cullen, JM, Briese, DT, Kriticos, DJ, Lonsdale, WM, Morin, L & Scott, JK, eds.), Canberra, Australia, CSIRO, pp. 252257.Google Scholar
Balciunas, JK (2004b) Four years of ‘Code of Best Practices’: has it had an impact? In Proceedings of the XI Symposium on Biological Control of Weeds (Cullen, JM, Briese, DT, Kriticos, DJ, Lonsdale, WM, Morin, L & Scott, JK, eds.), Canberra, Australia, CSIRO pp. 258260.Google Scholar
Bale, JS, Van Lenteren, JC & Bigler, F (2008) Biological control and sustainable food production. Philosophical Transactions of the Royal Society B 363: 761776.Google Scholar
Balmer, O, Geneau, CE, Belz, E, Weishaupt, B, Förderer, G, Moos, S, Ditner, N, Juric, I & Luka, H (2014) Wildflower companion plants increase pest parasitation and yield in cabbage fields: experimental demonstration and call for caution. Biological Control 76: 1927.Google Scholar
Banks, JE, Bommarco, R & Ekbom, B (2008) Population response to resource separation in conservation biological control. Biological Control 47: 141146.Google Scholar
Banks, JE, Stark, JD, Vargas, RI & Ackleh, AS (2011) Parasitoids and ecological risk assessment: can toxicity data developed for one species be used to protect an entire guild? Biological Control 59: 336339.Google Scholar
Barbercheck, ME & Millar, LC (2000) Environmental impacts of entomopathogenic nematodes used for biological control in soil. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 287308.Google Scholar
Barbosa, P, ed. (1998) Conservation Biological Control. San Diego, CA, Academic Press.Google Scholar
Barclay, HJ (1982) Models for pest control using predator release, habitat management and pesticide release in combination. Journal of Applied Ecology 19: 337348.Google Scholar
Barclay, HJ, Otvos, IS & Thomson, AJ (1985) Models of periodic inundative of parasitoids for pest control. Canadian Entomologist 117: 705716.Google Scholar
Barlow, ND (1994) Predicting the effect of a novel vertebrate biocontrol agent: a model for viral vectored immunocontraception of New Zealand opossums. Journal of Applied Ecology 31: 454462.Google Scholar
Barlow, ND (1998) Biological control in New Zealand: new models from real systems. In Proceedings of the VII International Congress of Ecology (Farina, A, Kennedy, J & Bossu, V, eds.), Cambridge, UK, Cambridge University Press, pp. 230259.Google Scholar
Barlow, ND (1999) Models in biological control: a field guide. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 4368.Google Scholar
Barlow, ND (2000) The ecological challenge of immunocontraception: editor’s introduction. Journal of Applied Ecology 37: 897902.Google Scholar
Barlow, ND, Barratt, BIP, Ferguson, CM & Barron, MC (2004) Using models to estimate parasitoid impact on nontarget host abundance. Environmental Entomology 33: 941948.Google Scholar
Barlow, ND, Barron, MC & Parkes, J (2002) Rabbit haemorrhagic disease in New Zealand: field test of a disease-host model. Wildlife Research 29: 649653.Google Scholar
Barlow, ND & Dixon, AFG (1980) Simulation of Lime Aphid Population Dynamics. Wageningen, The Netherlands, Centre for Agricultural Publishing.Google Scholar
Barlow, ND & Goldson, SL (1993) A modelling analysis of the successful biological control of Sitona discoideus (Coleoptera: Curculionidae) by Microctonus aethiopoides (Hymenoptera: Braconidae) in New Zealand. Journal of Applied Ecology 30: 165178.Google Scholar
Barlow, ND, Goldson, SL & McNeill, R (1994) A prospective model for the phenology of Microctonus hyperodae (Hymenoptera: Braconidae), a potential biological control agent of Argentine stem weevil in New Zealand. Biocontrol Science and Technology 4: 375386.Google Scholar
Barlow, ND & Kean, JM (1998) Simple models for the impact of rabbit calicivirus disease (RCD) on Australasian rabbits. Ecological Modelling 109: 225241.Google Scholar
Barlow, ND & Kean, JM (2004) Resource abundance and invasiveness: a simple model. Biological Invasions 6: 261268.Google Scholar
Barlow, ND, Moller, AP & Beggs, JR (1996) A model for the effect of Sphecophaga vesparum vesparum as a biological control agent of the common wasp in New Zealand. Journal of Applied Ecology 33: 3144.Google Scholar
Barratt, BIP (2004) Microctonus parasitoids and New Zealand weevils: comparing laboratory estimates of host ranges to realized host ranges. In Assessing Host Ranges for Parasitoids and Predators Used for Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, USDA Forest Service, pp. 103120.Google Scholar
Barratt, BIP, Blossey, B & Hokkanen, H (2006) Post-release evaluation of non-target effects of biological control agents. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 166186.Google Scholar
Barratt, BIP, Evans, AA, Ferguson, CM, Barker, GM, McNeill, MR & Phillips, CB (1997) Laboratory nontarget host range of the introduced parasitoids Microctonus aethiopoides and M. hyperodae (Hymenoptera: Braconidae) compared with field parasitism in New Zealand. Environmental Entomology 26: 694702.Google Scholar
Barratt, BIP, Ferguson, CM, Bixley, AS, Crook, KE, Barton, DM & Johnstone, PD (2007) Field parasitism of nontarget weevil species (Coleoptera: Curculionidae) by the introduced biological control agent Microctonus aethiopoides Loan (Hymenoptera: Braconidae) over an altitude gradient. Environmental Entomology 36: 826839.Google Scholar
Barratt, BIP, Howarth, FG, Withers, TM, Kean, JM & Ridley, GS (2010) Progress in risk assessment for classical biological control. Biological Control 52: 245254.Google Scholar
Barron, MC (2007) Retrospective modelling indicates minimal impact of non-target parasitism by Pteromalus puparum on red admiral butterfly (Bassaris gonerilla) abundance. Biological Control 41: 5363.Google Scholar
Barron, MC, Barlow, ND & Wratten, SD (2003) Non-target parasitism of the endemic New Zealand red admiral butterfly (Bassaris gonerilla) by the introduced biological control agent Pteromalus puparum. Biological Control 27: 329335.Google Scholar
Bartlett, BR (1964) Integration of chemical and biological control. In Biological Control of Insect Pests and Weeds (DeBach, P, ed.), New York, Reinhold, pp. 489511.Google Scholar
Barton, J (2004) How good are we at predicting the field host-range of fungal pathogens used for classical biological control of weeds? Biological Control, 31: 99122.Google Scholar
Barton Browne, L & Withers, TM (2002) Time-dependent changes in the host-acceptance threshold of insects: implications for host specificity testing of candidate biological control agents. Biocontrol Science and Technology 12: 677693.Google Scholar
Barve, N, Barve, V, Jimenez-Valverde, A, Lira-Noriega, A, Mahera, SP, Townsend Peterson, A, Soberóna, J & Villalobos, F (2011) The crucial role of the accessible area in ecological niche modeling and species distribution modeling. Ecological Modelling 222: 18101819.Google Scholar
Barzman, MS, Mills, NJ & Cuc, NTT (1996) Traditional knowledge and rationale for weaver ant husbandry in the Mekong Delta of Vietnam. Agriculture and Human Values 13: 29.Google Scholar
Bashan, Y & Holguin, G (1998) Proposal for the division of plant growth-promoting rhizobacteria into two classifications: Biocontrol-PGPB (Plant Growth-Promoting Bacteria) and PGPB. Soil Biology & Biochemistry 30: 12251228.Google Scholar
Bateman, R, Carey, M, Batt, D Prior, C, Abraham, Y, Moore, D, Jenkins, N & Fenlon, J (1996) Screening for virulent isolates of entomopathogenic fungi against the desert locust, Schistocerca gregaria (Forskal). Biocontrol Science and Technology 6: 549560.Google Scholar
Bathon, H (1996) Impact of entomopathogenic nematodes on non-target hosts. Biocontrol Science and Technology 6: 421434.Google Scholar
Batta, YA (2004) Postharvest biological control of apple gray mold by Trichoderma harzianum Rifai formulated in an invert emulsion. Crop Protection 23: 1926.Google Scholar
Baudoin, A, Abad, RG, Kok, LT & Bruckart, WL (1993) Field evaluation of Puccinia carduorum for biological control of musk thistle. Biological Control 3: 5360.Google Scholar
Baumhover, AH (2002) A personal account of developing the sterile insect technique to eradicate the screwworm from Curacao, Florida and the Southeastern United States. Florida Entomologist 85: 666673.Google Scholar
Bean, DW, Dalin, P & Dudley, TL (2012) Evolution of critical day length for diapause induction enables range expansion of Diorhabda carinulata, a biological control agent against tamarisk (Tamarix spp.). Evolutionary Applications 5: 511523.Google Scholar
Bean, D, Dudley, T & Hultine, K (2013) Bring on the beetles! In Tamarix: A Case Study of Ecological Change in the American West (Sher, A & Quigley, MF, eds.), New York, Oxford University Press, pp. 377403.Google Scholar
Bean, DW, Dudley, TL & Keller, JC (2007) Seasonal timing of diapause induction limits the effective range of Diorhabda elongata deserticola (Coleoptera: Chrysomelidae) as a biological control agent for tamarisk (Tamarix spp.). Environmental Entomology 36: 1525.Google Scholar
Beane, K & Bugg, RL (1998) Natural and artificial shelter to enhance arthropod biological control agents. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests (Pickett, CH & Bugg, RL, eds.), Berkeley, University of California Press, pp. 239254.Google Scholar
Beckendorf, SK & Hoy, MA (1985) Genetic improvement of arthropod natural enemies through selection, hybridization or genetic engineering techniques. In Biological Control in Agricultural IPM Systems (Hoy, MA & Herzog, DC, eds.), Orlando, FL, Academic Press, pp. 167187.Google Scholar
Becker, J, Eisenhauer, N, Scheu, S & Jousset, A (2012) Increasing antagonistic interactions cause bacterial communities to collapse at high diversity. Ecology Letters 15: 468474.Google Scholar
Beddington, JR (1974) Age distribution and the stability of simple discrete time population models. Journal of Theoretical Biology 47: 6574.CrossRefGoogle ScholarPubMed
Beddington, JR (1975) Mutual interference between parasites or predators and its effects on searching efficiency. Journal of Animal Ecology 44: 331340.Google Scholar
Beddington, JR, Free, CA & Lawton, JH (1975) Dynamic complexity in predator–prey models framed in difference equations. Nature 255: 5860.Google Scholar
Beddington, JR, Free, CA & Lawton, JH (1978) Characteristics of successful natural enemies in models of biological control of insect pests. Nature 273: 513519.Google Scholar
Beddington, JR & Hammond, PS (1977) On the dynamics of host-parasite-hyperparasite interactions. Journal of Animal Ecology 46: 811821.Google Scholar
Beddington, JR, Hassell, MP & Lawton, JH (1976) The components of arthropod predation II. The predator rate of increase. Journal of Animal Ecology 45: 165185.Google Scholar
Begum, M, Gurr, GM, Wratten, SD, Hedberg, PR & Nicol, HI (2006) Using selective food plants to maximize biological control of vineyard pests. Journal of Applied Ecology 43: 547554.Google Scholar
Behle, R & Birthisel, T (2014) Formulation of entomopathogens as bioinsecticides. In Mass Production of Beneficial Organisms (Morales-Ramos, JA, Guadalupe Rojas, M & Shapro-Ilan, DL, eds.), Amsterdam, The Netherlands, Elsevier, pp. 483517.CrossRefGoogle Scholar
Beirne, BP (1975) Biological control attempts by introductions against pest insects in the field in Canada. Canadian Entomologist 107: 225236.Google Scholar
Bellamy, DE & Byrne, DN (2001) Effects of gender and mating status on self-directed dispersal by the whitefly parasitoid Eretmocerus eremicus. Ecological Entomology 26: 571577.Google Scholar
Bellows, TS, Paine, TD, Gould, JR, Bezark, LG & Ball, JC (1992) Biological control of ash whitefly: a success in progress. California Agriculture 46: 2428.Google Scholar
Bellows, TS & Van Driesche, RG (1999) Life table construction and analysis for evaluating biological control agents. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 199223.Google Scholar
Belshaw, R (1994) Life history characteristics of Tachinidae (Diptera) and their effect of polyphagy. In Parasitoid Community Ecology (Hawkins, BA & Sheehan, W, eds.), Oxford, UK, Oxford University Press, pp. 145162.Google Scholar
Bence, SL, Stander, K & Griffiths, M (2003) Habitat characteristics of harvest mouse nests on arable farmland. Agriculture, Ecosystems and Environment 99: 179186.Google Scholar
Benson, J, Pasquale, A, Van Driesche, R & Elkinton, J (2003b) Assessment of risk posed by introduced braconid wasps to Pieris virginiensis, a native woodland butterfly in New England. Biological Control 26: 8393.Google Scholar
Benson, J, Van Driesche, RG, Pasquale, A & Elkinton, J (2003a) Introduced braconid parasitoids and range reduction of a native butterfly in New England. Biological Control 28: 197213.Google Scholar
Benvenuto, C, Tabone, E, Vercken, E, Sorbier, N, Colombel, E, Warot, S, Fauverge, X & Ris, N (2012) Intraspecific variability in the parasitoid wasp Trichogramma chilonis: can we predict the outcome of hybridization? Evolutionary Applications 5: 498510.Google Scholar
Berberet, RC, Zarrabi, AA, Payton, ME & Bisges, AD (2003) Reduction in effective parasitism of Hypera postica (Coleoptera: Curculionidae) by Bathyplectes curculionis (Hymenoptera: Ichneumonidae) due to encapsulation. Environmental Entomology 32: 11231130.Google Scholar
Bergelson, J & Kareiva, P (1987) Barriers to movement and the response of herbivores to alternative cropping patterns. Oecologia 71: 457460.CrossRefGoogle ScholarPubMed
Bergstedt, RA, McDonald, RB, Twohey, MB & Heinrich, JW (2003) Reduction in sea lamprey hatching success due to release of sterilized males. Journal of Great Lakes Research 29: 435444.Google Scholar
Berkelhamer, RC (1983) Intraspecific genetic variation and haplodiploidy, eusociality, and polygyny in the Hymenoptera. Evolution 37: 540545.Google Scholar
Bernal-Vicente, A, Ros, M & Antonio Pascual, J (2012) Inoculation of Trichoderma harzianum during maturation of vineyard waste compost to control muskmelon Fusarium wilt. Bioresources 7: 19481960.Google Scholar
Bernays, EA & Graham, M (1988) On the evolution of host specificity in phytophagous arthropods. Ecology 69: 886915.Google Scholar
Berryman, AA & Hawkins, BA (2006) The refuge as an integrating concept in ecology and evolution. Oikos 115: 192196.Google Scholar
Bess, HA, Van den Bosch, R & Haramoto, FH (1961) Fruit fly parasites and their activities in Hawaii. Proceedings of the Hawaiian Entomological Society 17: 367378.Google Scholar
Best, SM, Collins, SV & Kerr, PJ (2000) Coevolution of host and virus: cellular localization of virus in myxoma virus infection of resistant and susceptible European rabbits. Virology 277: 7691.Google Scholar
Best, SM & Kerr, PJ (2000) Coevolution of host and virus: the pathogenesis of virulent and attenuated strains of myxoma virus in resistant and susceptible European rabbits. Virology 267: 3648.Google Scholar
Beukeboom, LW (2001) Single-locus complementary sex determination in the ichneumonid Venturia canescens (Gravenhorst) (Hymenoptera). Netherlands Journal of Zoology 51: 115.Google Scholar
Bezemer, TM, Harvey, JA, Kamp, AFD, Wagenaar, R, Gols, R, Kostenko, O, Fortuna, T, Engelkes, T, Vet, LEM, Van der Putten, WH & Soler, R. (2010) Behaviour of male and female parasitoids in the field: influence of patch size, host density, and habitat complexity. Ecological Entomology 35: 341351.Google Scholar
Bianchi, FJJA, Booij, CJH & Tscharnkte, T (2006) Sustainable pest regulation in agricultural landscapes: a review on landscape composition, biodiversity and natural pest control. Proceedings of the Royal Society B-Biological Sciences 273: 17151727.Google Scholar
Bianchi, F, Goedhart, PW & Baveco, JM (2008) Enhanced pest control in cabbage crops near forest in The Netherlands. Landscape Ecology 23: 595602.Google Scholar
Bianchi, F & Van der Werf, W (2004) Model evaluation of the function of prey in non-crop habitats for biological control by ladybeetles in agricultural landscapes. Ecological Modelling 171: 177193.Google Scholar
Bigger, DS & Chaney, WE (1998) Effects of Iberis umbellata (Brassicaceae) on insect pests of cabbage and on potential biological control agents. Environmental Entomology 27: 161167.CrossRefGoogle Scholar
Bigler, F, Babendreier, D & Kuhlmann, U (2006) Environmental Impact of Invertebrates for Biological Control of Arthropods: Methods and Risk Assessment. Wallingford, UK, CABI Publishing.Google Scholar
Bigler, F & Kölliker-Ott, UM (2006) Balancing environmental risks and benefits: a basic approach. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 273286.Google Scholar
Bilgrami, AL, Gaugler, R, Shapiro-Ilan, DI & Adams, BJ (2006) Source of trait deterioration in entomopathogenic nematodes Heterorhabditis bacteriophora and Steinernema carpocapsae during in vivo culture. Nematology 8: 397409.Google Scholar
Biondi, A, Desneux, N, Siscaro, G & Zappala, L (2012a) Using organic-certified rather than synthetic pesticides may not be safer for biological control agents: selectivity and side effects of 14 pesticides on the predator Orius laevigatus. Chemosphere 87: 803812.Google Scholar
Biondi, A, Mommaerts, V, Smagghe, G, Viñuela, E, Zappalà, L & Desneux, N (2012b) The non-target impact of spinosyns on beneficial arthropods. Pest Management Science 68: 15231536.Google Scholar
Biondi, A, Zappala, L, Stark, JD & Desneux, N (2013) Do biopesticides affect the demographic traits of a parasitoid wasp and its biocontrol services through sublethal effects? PLoS One 8: e76548.Google Scholar
Birnbaum, MJ, Clem, RJ & Miller, LK (1994) An apoptosis-inhibiting gene from a nuclear polyhedrosis virus encoding a polypeptide with Cys/His sequence motifs. Journal of Virology 68: 25212528.Google Scholar
Bischoff, JF, Rehner, SA & Humber, RA (2009) A multilocus phylogeny of the Metarhizium anisopliae lineage. Mycologia 101: 512530.Google Scholar
Blackburn, TM, Essl, F, Evans, T, Hulme, PE, Jeschke, JM, Kühn, I, Kumschick, S, Marková, Z, Mrugała, A, Nentwig, W, Pergl, J, Pyšek, P, Rabitsch, W, Ricciardi, A, Richardson, DM, Sendek, A, Vilà, M, Wilson, JRU, Winter, M, Genovesi, P & Bacher, S (2014) Towards a unified classification of alien species based on the magnitude of their environmental impacts. PLoS Biology 12: e1001850.Google Scholar
Blackburn, TM, Prowse, TAA, Lockwood, JL & Cassey, P (2013) Propagule pressure as a driver of establishment success in deliberately introduced exotic species: fact or artefact? Biological Invasions 15: 14591469.Google Scholar
Blackburn, TM, Pysek, P, Bacher, S, Carlton, JT, Duncan, RP, Jarošík, V, Wilson, JR & Richardson, DM (2011) A proposed unified framework for biological invasions. Trends in Ecology & Evolution 26: 333339.Google Scholar
Blair, AC, Hanson, BD, Brunk, GR, Marrs, RA, Westra, P, Nissen, SJ, Hufbauer, RA (2005) New techniques and findings in the study of a candidate allelochemical implicated in invasion success. Ecology Letters 8: 10391047.Google Scholar
Blair, AC, Nissen, SJ, Brunk, GR & Hufbauer, RA (2006) A lack of evidence for an ecological role of the putative allelochemical (+/-)-catechin in spotted knapweed invasion success. Journal of Chemical Ecology 32: 23272331.Google Scholar
Blair, AC, Weston, LA, Nissen, SJ, Brunk, GR & Hufbauer, RA (2009) The importance of analytical techniques in allelopathy studies with the reported allelochemical catechin as an example. Biological Invasions 11: 325332.Google Scholar
Blair, AC & Wolfe, LM (2004) The evolution of an invasive plant: an experimental study with Silene latifolia. Ecology 85: 30353042.Google Scholar
Blanford, S, Thomas, MB & Langewald, J (1998) Behavioural fever in the Senegalese grasshopper, Oedaleus senegalensis, and its implications for biological control using pathogens. Ecological Entomology 23: 914.Google Scholar
Blossey, B (2004) Monitoring in weed biological control programs. Biological Control of Invasive Plants in the United States (Coombs, E, Clark, JK, Piper, GL & Cofrancesco, AF Jr., eds.), Corvallis, Oregon State University Press, pp. 95105.Google Scholar
Blossey, B, Casagrande, R, Tewksbury, L & Ellis, DR (2001) Nontarget feeding of leaf-beetles introduced to control purple loosestrife (Lythrum salicaria L.). Natural Areas Journal 21: 368377.Google Scholar
Blossey, B & Kamil, J (1996) What determines the increased competitive ability of invasive non-indigenous plants? In Proceedings of the IX Symposium on Biological Control (Moran, VC & Hoffmann, JH, eds.), Stellenbosch, S. Africa, University of Cape Town, pp. 39.Google Scholar
Blossey, B & Nötzold, R (1995) Evolution of increased competitive ability in invasive nonindigenous plants: a hypothesis. Journal of Ecology 83: 887889.Google Scholar
Blows, MW (1993) The genetics of central and marginal populations of Drosophila serrata. II Hybrid breakdown in fitness components as a correlated response to selection for dessication resistance. Evolution 47: 12711285.Google ScholarPubMed
Boag, B & Yeates, GW (2001) The potential impact of the New Zealand flatworm, a predator of earthworms, in western Europe. Ecological Applications 11: 12761286.Google Scholar
Boavida, C, Neuenschwander, P & Herren, HR (1995) Experimental assessment of the impact of the introduced parastoid Gyranusoidea tebygi Noyes on the mango mealybug Rastrococcus invadens Williams, by physical exclusion. Biological Control 5: 99103.Google Scholar
Bock, DG, Caseys, C, Cousens, RD, Hahn, MA, Heredia, SM, Hübner, S, Turner, KG, Whitney, K & Rieseberg, LH (2015) What we still don’t know about invasion genetics. Molecular Ecology 24: 22772297.CrossRefGoogle ScholarPubMed
Boettner, GH, Elkinton, JS & Boettner, CJ (2000) Effects of a biological control introduction on three nontarget native species of saturniid moths. Conservation Biology 14: 17981806.Google Scholar
Boivin, G, Kölliker-Ott, UM, Bale, J & Bigler, F (2006) Assessing the establishment potential of inundative biological control agents. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 98113.Google Scholar
Bokonon-Ganta, AH, De Groote, H & Neuenschwander, P (2002) Socio-economic impact of biological control of mango mealybug in Benin. Agriculture, Ecosystems and Environment 93: 367378.Google Scholar
Bokonon-Ganta, AH & Neuenschwander, P (1995) Impact of the biological control agent Gyranusoidea tebygi Noyes (Hymenoptera, Encyrtidae) on the mango mealybug, Rastrococcus invadens Williams (Homoptera, Pseudococcidae) in Benin. Biocontrol Science and Technology 5: 95107.Google Scholar
Bolwerk, A, Lagopodi, AL, Wijfjes, AHM Lamers, GE, Chin-A-Woeng, TF, Lugtenberg, BJ & Bloemberg, GV (2003) Interactions in the tomato rhizosphere of two Pseudomonas biocontrol strains with the phytopathogenic fungus Fusarium oxysporum f. sp radicis-lycopersici. Molecular Plant-Microbe Interactions 16: 983993.Google Scholar
Bommarco, R & Banks, JE (2003) Scale as modifier in vegetation diversity experiments: effects on herbivores and predators. Oikos 102: 440448.Google Scholar
Bonning, BC, Boughton, AJ & Harrison, RL (2002) Genetic enhancement of baculovirus insecticides. In Advances in Microbial Control of Insect Pests (Upadhyay, RK, ed.), New York, Plenum, pp. 109125.Google Scholar
Bonsall, MB, French, DR & Hassell, MP (2002) Metapopulation structures affect persistence of predator–prey interactions. Journal of Animal Ecology 71: 10751084.Google Scholar
Bonsall, MB & Hassell, MP (1999) Parasitoid-mediated effects: apparent competition and the persistence of host-parasitoid assemblages. Researches on Population Ecology 41: 5968.Google Scholar
Bonsall, MB, O’Reilly, DR, Cory, JS & Hails, RS (2005) Persistence and coexistence of engineered baculoviruses. Theoretical Population Biology 67: 217230.Google Scholar
Borer, ET, Hosseini, PR, Seabloom, EW & Dobson, AP (2007) Pathogen-induced reversal of native dominance in a grassland community. Proceedings of the National Academy of Sciences of the United States of America 104: 54735478.Google Scholar
Bornemissza, GF (1966) An attempt to control ragwort in Australia with the cinnabar moth, Callimorpha jacobaeae (L.) (Arctiidae: Lepidoptera). Australian Journal of Zoology 14: 201243.Google Scholar
Bossard, CC (1991) The role of habitat disturbance, seed predation and ant dispersal on establishment of the exotic shrub Cytisus scoparius in California. American Midland Naturalist 126: 113.Google Scholar
Bossdorf, O, Prati, D, Auge, H & Schmid, B (2004a) Reduced competitive ability in an invasive plant. Ecology Letters 7: 346353.Google Scholar
Bossdorf, O, Schroder, S, Prati, D & Auge, H (2004b) Palatability and tolerance to simulated herbivory in native and introduced populations of Alliaria petiolata (Brassicaceae). American Journal of Botany 91: 856862.Google Scholar
Boudouresque, CF & Verlaque, M (2002) Biological pollution in the Mediterranean Sea: invasive versus introduced macrophytes. Biological Invasions 44: 3238.Google Scholar
Boughton, AJ & Pemberton, RW (2008) Efforts to establish a foliage-feeding moth, Austromusotima camptozonale, against Lygodium microphyllum in Florida, considered in the light of a retrospective review of establishment success of weed biocontrol agents belonging to different arthropod taxa. Biological Control 47: 2836.Google Scholar
Boulter, JI, Boland, GJ & Trevors, JT (2000) Compost: a study of the development process and end-product potential for suppression of turfgrass disease. World Journal of Microbiology & Biotechnology 16: 115134.Google Scholar
Bourchier, RS and Smith, SM (1996) Influence of environmental conditions and parasitoid quality on field performance of Trichogramma minutum. Entomologia Experimentalis et Applicata 80: 461468.Google Scholar
Bourdot, GW, Baird, D, Hurrell, GA & de Jong, MD (2006) Safety zones for a Sclerotinia sclerotiorum-based mycoherbicide: accounting for regional and yearly variation in climate. Biocontrol Science and Technology 16: 345358.Google Scholar
Bourdot, GW, Hurrell, GA, Saville, DJ & de Jong, MD (2001) Risk analysis of Sclerotinia sclerotiorum for biological control of Cirsium arvense in pasture: ascospore dispersal. Biocontrol Science and Technology 11: 119139.Google Scholar
Bourdot, GW, Saville, DJ, Hurrell, GA, Harvey, IC & De Jong, MD (2000) Risk analysis of Sclerotinia sclerotiorum for biological control of Cirsium arvense in pasture: sclerotiorum survival. Biocontrol Science and Technology 10: 411425.Google Scholar
Braga, GUL, Flint, SD, Miller, CD, Anderson, AJ & Roberts, DW (2001) Variability in response to UV-B among species and strains of Metarhizium isolated from sites at latitudes from 61degreeN to 54degreeS. Journal of Invertebrate Pathology 78: 98108.Google Scholar
Brandle, JR, Hodges, L & Zhou, XH (2004) Windbreaks in North American agricultural systems. Agroforestry Systems 61–2: 6578.Google Scholar
Briese, DT (1986a) Factors affecting the establishment and survival of Anaitis efformata (Lepidoptera: Geometridae) introduced into Australia for the biological control of St. John’s Wort, Hypericum perforatum. II. Field trials. Journal of Applied Ecology 23: 821839.Google Scholar
Briese, DT (1986b) Host resistance to microbial control agents. Fortschritte der Zoologie 32: 233256.Google Scholar
Briese, DT (1986c) Insect resistance to baculoviruses. In The Biology of Baculoviruses Volume II: Practical Application for Insect Control (Granados, RR & Federici, BA, eds.), Boca Raton, FL, CRC Press, pp. 237264.Google Scholar
Briese, DT (1996) Biological control of weeds and fire management in protected natural areas: are they compatible strategies? Biological Conservation 77: 135141.Google Scholar
Briese, DT (1997) Biological control of St. John’s wort: past, present and future. Plant Protection Quarterly 12: 7380.Google Scholar
Briese, DT (2005) Translating host-specificity test results into the real world: the need to harmonize the yin and yang of current testing procedures. Biological Control 35: 208214.Google Scholar
Briese, DT & Walker, A (2002) A new perspective on the selection of test plants for evaluating the host-specificity of weed biological control agents: the case of Deuterocampta quadrijuga, a potential insect control agent of Heliotropium amplexicaule. Biological Control 25: 273287.Google Scholar
Briese, DT, Pettit, WJ & Walker, A (2004) Evaluation of the biological control agent, Lixus cardui, on Onopordum thistles: experimental studies on agent demography and impact. Biological Control 31: 165171.Google Scholar
Briese, DT, Walker, A, Pettit, WJ & Sagliocco, JL (2002a) Host-specificity of candidate agents for the biological control of Onopordum spp. thistles in Australia: an assessment of testing procedures. Biocontrol Science and Technology 12: 149163.Google Scholar
Briese, DT, Zapater, M, Andorno, A & Perez-Camargo, G (2002b) A two-phase open-field test to evaluate the host-specificity of candidate biological control agents for Heliotropium amplexicaule. Biological Control 25: 259272.Google Scholar
Briese, DT, Zapater, M & Walker, A (2005) Implementation of the blue heliotrope biological control strategy: host-specificity testing of Longitarsus sp. In A Report for the Rural Industries Research and Development Corporation, Barton, Australia, Rural Industries Research and Development Corporation.Google Scholar
Briggs, CJ (1993) Competition among parasitoid species on a stage-structured host and its effect on host suppression. American Naturalist 141: 372379.Google Scholar
Briggs, CJ (2009) Host-parasitoid interactions. In The Princeton Guide to Ecology (Levin, SA, ed.), Princeton NJ, Princeton University Press, pp. 213219.Google Scholar
Briggs, CJ & Borer, ET (2005) Why short-term experiments may not allow long-term predictions about intraguild predation. Ecological Applications 15: 11111117.Google Scholar
Briggs, CJ & Godfray, HCJ (1995a) Models of intermediate complexity in insect-pathogen interactions: population dynamics of the microsporidian pathogen, Nosema pyrausta, of the European corn borer, Ostrinia nubilalis. Parasitology 111: S71S89.Google Scholar
Briggs, CJ & Godfray, HCJ (1995b) The dynamics of insect-pathogen interactions in stage-structured populations. American Naturalist 145: 855887.Google Scholar
Briggs, CJ & Godfray, HCJ (1996) The dynamics of insect-pathogen interactions in seasonal environments. Theoretical Population Biology 50: 149177.Google Scholar
Briggs, CJ, Hails, ND, Barlow, ND & Godfray, HCJ (1995) The dynamics of insect-pathogen interactions. In Ecology of Infectious Diseases in Natural Populations (Grenfell, B & Dobson, A, eds.), Cambridge, UK, Cambridge University Press, pp. 295306.Google Scholar
Briggs, CJ & Hoopes, MF (2004) Stabilizing effects in spatial parasitoid–host and predator–prey models: a review. Theoretical Population Biology 65: 299315.Google Scholar
Briggs, CJ, Murdoch, WW & Nisbet, RM (1999) Recent developments in theory for biological control of insect pests by parasitoids. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 2242.Google Scholar
Brodeur, J (2000) Host specificity and trophic relationships of hyperparasitoids. In Parasitoid Population Biology (Hochberg, ME & Ives, AR, eds.), Princeton, NJ, Princeton University Press, pp. 163183.Google Scholar
Brodeur, J (2012) Host specificity in biological control: insights from opportunistic pathogens. Evolutionary Applications 5: 470480.Google Scholar
Brodeur, J & Boivin, G (2006) Trophic and Guild Interactions in Biological Control. Dordrecht, The Netherlands, Springer.Google Scholar
Brodeur, J, Geervliet, JBF & Vet, LEM (1996) The role of host species, age and defensive behavior on ovipositional decisions in a solitary specialist and gregarious generalist parasitoid (Cotesia species). Entomologia Experimentalis et Applicata 81: 125132.Google Scholar
Brodeur, J & Rosenheim, JA (2000) Intraguild interactions in aphid parasitoids. Entomologia Experimentalis et Applicata 97: 93108.Google Scholar
Brodmann, PA, Wilcox, CV & Harrison, S (1997) Mobile parasitoids may restrict the spatial spread of an insect outbreak. Journal of Animal Ecology 66: 6572.Google Scholar
Bronstein, JL (1998) The contribution of ant plant protection studies to our understanding of mutualism. Biotropica 30: 150161.Google Scholar
Brown, JH & Heske, EJ (1990) Control of a desert-grassland transition by a keystone rodent guild. Science 250: 17051707.Google Scholar
Brown, PMJ, Adriaens, T, Bathon, H, Cuppen, J, Goldarazena, A, Hägg, T, Klausnitzer, BEM, Kovář, I, Loomans, AJM, Majerus, MEN, Nedvěd, O, Pedersen, J, Rabitsch, W, Roy, HE, Ternois, V, Zakharov, IA, Roy, DB (2008) Harmonia axyridis in Europe: spread and distribution of a non-native coccinellid. BioControl 53: 521.Google Scholar
Brunner, K, Zeilinger, S, Ciliento, R, Woo, SL, Lorito, M, Kubicek, CP & Mach, RL (2005) Improvement of the fungal biocontrol agent Trichoderma atroviride to enhance both antagonism and induction of plant systemic disease resistance. Applied and Environmental Microbiology 71: 39593965.Google Scholar
Bucher, GE & Harris, P (1961) Food-plant spectrum and elimination of disease of cinnabar moth, Hypocrita jacobaeae (L.) (Lepidoptera: Arctiidae). Canadian Entomologist 93: 931936.Google Scholar
Buckingham, GR (2001) Quarantine host range studies with Lophyrotoma zonalis, an Australian sawfly of interest for biological control of melaleuca, Melaleuca quinquenervia, in Florida. BioControl 46: 363386.Google Scholar
Buckley, YM, Hinz, HL, Matthies, D & Rees, M (2001) Interactions between density-dependent processes, population dynamics and control of an invasive plant species, Tripleurospermum perforatum (scentless chamomile). Ecology Letters 4: 551558.Google Scholar
Buckley, YM, Rees, M, Paynter, Q & Lonsdale, M (2004) Modelling integrated weed management of an invasive shrub in tropical Australia. Journal of Applied Ecology 41: 547560.Google Scholar
Buckley, YM, Rees, M, Sheppard, AW & Smyth, MJ (2005) Stable coexistence of an invasive plant and biocontrol agent: a parameterized coupled plant-herbivore model. Journal of Applied Ecology 42: 7079.Google Scholar
Buerger, P, Hauxwell, C & Murray, D (2007) Nucleopolyhedrovirus introduction in Australia. Virologica Sinica 22: 173179.Google Scholar
Bugg, RL, Ehler, LE & Wilson, LT (1987) Effect of common Knotweed (Polygonum aviculare) on abundance and efficiency of insect predators of crop pests. Hilgardia 55: 153.Google Scholar
Bugg, RL & Waddington, C (1994) Using cover crops to manage arthropod pests of orchards: a review. Agriculture, Ecosystems and Environment 50: 1128.Google Scholar
Bukovinszky, T, Gols, R, Hemerik, L, Van Lenteren, JC & Vet, LEM (2007) Time allocation of a parasitoid foraging in heterogeneous vegetation: implications for host-parasitoid interactions. Journal of Animal Ecology 76: 845853.Google Scholar
Bukovinszky, T, Trefas, H, Van Lenteren, JC, Vet, LEM & Fremont, J (2004) Plant competition in pest-suppressive intercropping systems complicates evaluation of herbivore responses. Agriculture, Ecosystems & Environment 102: 185196.Google Scholar
Bull, JJ (1994) Virulence. Evolution 48: 14231437.Google Scholar
Burdon, JJ, Groves, RH & Cullen, JM (1981) The impact of biological control on the distribution and abundance of Chondrilla juncea in south-eastern Australia. Journal of Applied Ecology 18: 957966.Google Scholar
Burdon, JJ, Groves, RH, Kaye, PE & Speer, SS (1984) Competition in mixtures of susceptible and resistant genotypes of Chondrilla juncea differentially infected with rust. Oecologia 64: 199203.CrossRefGoogle ScholarPubMed
Burdon, JJ & Marshall, DR (1981) Biological control and the reproductive mode of weeds. Journal of Applied Ecology 18: 649658.Google Scholar
Burdon, JJ & Thrall, PH (2004) Genetic structure of natural plant and pathogen populations. In Genetics, Evolution and Biological Control (Ehler, LE, Sforza, R & Mateille, T, eds.), Wallingford, UK, CABI Publishing, pp. 117.Google Scholar
Burges, DH (1998) Formulation of Microbial Pesticides. Dordrecht, The Netherlands, Kluwer.Google Scholar
Burges, DH & Jones, KA (1998) Formulation of bacteria, viruses and Protozoa to control insects. In Formulation of Microbial Pesticides (Burges, DH, ed.), Dordrecht, The Netherlands, Kluwer, pp. 33127.Google Scholar
Bürgi, LP & Mills, NJ (2014) Lack of enemy release for an invasive leafroller in California: temporal patterns and influence of host plant origin. Biological Invasions 16: 10211034.Google Scholar
Bürgi, LP, Roltsch, WJ & Mills, NJ (2015) Allee effects and population regulation: a test for biotic resistance against an invasive leafroller by resident parasitoids. Population Ecology 57: 215225.Google Scholar
Burnett, T (1958) A model of host-parasite interactions. Proceedings of the 10th International Congress of Entomology 2: 679686.Google Scholar
Burrows, BE & Balciunas, JK (1997) Biology, distribution and host-range of the sawfly, Lophyrotoma zonalis (Hym., Pergidae), a potential biological control agent for the paperbark tree, Melaleuca quinquenervia. Entomophaga 42: 299313.Google Scholar
Bush, GL (1969) Sympatric host race formation and speciation in frugivorous flies of the genus Rhagoletis (Diptera, Tephritidae). Evolution 23: 237251.Google Scholar
Butcher, RDJ, Whitfield, WGF & Hubbard, SF (2000). Complementary sex determination in the genus Diadegma (Hymenoptera: Ichneumonidae). Journal of Evolutionary Biology 13: 593606.Google Scholar
Callaway, RM, DeLuca, TH & Belliveau, WM (1999) Biological-control herbivores may increase competitive ability of the noxious weed Centaurea maculosa. Ecology 80: 11961201.Google Scholar
Callaway, RM & Ridenour, WM (2004) Novel weapons: invasive success and the evolution of increased competitive ability. Frontiers in Ecology and the Environment 2: 436443.Google Scholar
Caltagirone, LE (1981) Landmark examples in classical biological control. Annual Review of Entomology 26: 213232.Google Scholar
Caltagirone, LE & Doutt, RL (1989) The history of the vedalia beetle importation to California and its impact on the development of biological control. Annual Review of Entomology 34:116.Google Scholar
Cameron, E (1935) A study of the natural control of ragwort (Senecio jacobaea) L. Journal of Ecology 23: 265322.Google Scholar
Cameron, PJ, Hill, RL, Bain, J & Thomas, WP (1993) Analysis of importations for biological control of insect pests and weeds in New Zealand. Biocontrol Science and Technology 3: 387404.Google Scholar
Campbell, MM (1976) Colonisation of Aphytis melinus DeBach (Hymenoptera, Aphelinidae) in Aonidiella aurantii (Mask.) (Hemiptera, Coccidae) on citrus in South Australia. Bulletin of Entomological Research 65: 659668.Google Scholar
Cao, YQ, Peng, GX, He, ZB, Wang, ZK, Yin, YP, Xia, YX (2007) Transformation of Metarhizium anisopliae with benomyl resistance and green fluorescent protein genes provides a tag for genetically engineered strains. Biotechnology Letters 29: 907911.Google Scholar
Carabajal Paladino, L, Muntaabski, I, Lanzavecchia, S, Le Bagousse-Pinguet, Y, Viscarret, M, Juri, M, Fueyo-Sánchez, L, Papeschi, A, Cladera, J, Bressa, MJ (2015) Complementary sex determination in the parasitic wasp Diachasmimorpha longicaudata. PLoS One 10: e0119619.Google Scholar
Carey, JR (1989) The multiple decrement life table: A unifying framework for cause-of-death analysis in ecology. Oecologia 78: 131137.Google Scholar
Carmona, DM & Landis, DA (1999) Influence of refuge habitats and cover crops on seasonal activity-density of ground beetles (Coleoptera: Carabidae) in field crops. Environmental Entomology 28: 11451153.Google Scholar
Carisse, O & Rolland, D (2004) Effect of timing of application of the biological control agent Microsphaeropsis ochracea on the production and ejection pattern of ascospores by Venturia inaequalis. Phytopathology 94: 13051314.Google Scholar
Carruthers, RI & D’Antonio, CM (2005) Science and decision making in biological control of weeds: benefits and risks of biological control. Biological Control 35: 181182.Google Scholar
Carson, WP, Hovick, SM, Baumert, AJ, Bunker, DE & Pendergast, TH (2008) Evaluating the post-release efficacy of invasive plant biocontrol by insects: a comprehensive approach. Arthropod-Plant Interactions 2: 7786.Google Scholar
Carter, N, Dixon, AFG & Rabbinge, R (1982) Cereal Aphid Populations: Biology, Simulation and Prediction. Wageningen, The Netherlands, Centre for Agricultural Publishing and Documentation.Google Scholar
Carton, Y & Kitano, H (1981) Evolutionary relationships to parasitism by seven species of the Drosophila melanogaster subgroup. Biological Journal of the Linnean Society 16: 227241.Google Scholar
Carvalheiro, LG, Buckley, YM, Ventim, R, Fowler, SV & Memmott, J (2008) Apparent competition can compromise the safety of highly specific biological control agents. Ecology Letters 11: 690700.Google Scholar
Case, CM & Crawley, MJ (2000) Effect of intraspecific competition and herbivory on the recruitment of an invasive alien plant: Conyza sumatrensis. Biological Invasions 2: 103110.Google Scholar
Cassani, JR & Caton, WE (1985) Induced triploidy in grass garp, Ctenopharyngodon idella Val. Aquaculture 46: 3744.Google Scholar
Castells, E, Morante, M, Blanco-Moreno, JM, Sans, FX, Vilatersana, R & Blasco, A (2013) Reduced seed predation after invasion supports enemy release in a broad biogeographical survey. Oecologia 173: 13971409.Google Scholar
Caughley, G & Lawton, JH (1981) Plant-herbivore systems. In Theoretical Ecology: Principles and Applications, 2nd edition (May, RM, ed.), Oxford, UK, Blackwell, pp. 132166.Google Scholar
Causton, CE (2009) Success in biological control: the scale and the ladybird. In Galapagos: Preserving Darwin’s Legacy (De Roy, T, ed.), Auckland, New Zealand, David Bateman, pp. 184190.Google Scholar
Cavallini, P & Serafini, P (1995) Winter diet of the small Indian mongoose, Herpestes auropunctatus, on an Adriatic island. Journal of Mammalogy 76: 569574.Google Scholar
Ceballo, FA, Walter, GH & Rochester, W (2010) The impact of climate on the biological control of citrus mealybug [Planococcus citri (Risso)] by the parasitoid Coccidoxenoides perminutus Girault as predicted by the climate-matching program CLIMEX. Philippine Agricultural Scientist 93: 317328.Google Scholar
Center, TD & Dray, FA (2010) Bottom-up control of water hyacinth weevil populations: do the plants regulate the insects? Journal of Applied Ecology 47: 329337.Google Scholar
Center, TD, Dray, FA, Jubinsky, GP & Grodowitz, MJ (1999) Biological control of water hyacinth under conditions of maintenance management: can herbicides and insects be integrated? Environmental Management 23: 241256.Google Scholar
Center, TD, Purcell, MF, Pratt, PD, Rayamajhi, M, Tipping, PW, Wright, SA & Dray, A (2012) Biological control of Melaleuca quinquenervia: an Everglades invader. BioControl 57: 151165.Google Scholar
Chaboudez, P & Burdon, JJ (1995) Frequency-dependent selection in a wild plant-pathogen system. Oecologia 102: 490493.Google Scholar
Chaboudez, P & Sheppard, AW (1995) Are particular weeds more amenable to biological control? A reanalysis of reproduction of life history. In Proceedings of the Eighth International Symposium on Biological Control of Weeds (Delfosse, ES & Scott, RR, eds.), Melbourne, Australia, DSIR/CSIRO, pp. 95102.Google Scholar
Chace, AB (1979) The Rhind Mathematical Papyrus. Oberlin, OH, Mathematical Association of America.Google Scholar
Chacon, J & Heimpel, GE (2010) Density-dependent intraguild predation of an aphid parasitoid. Oecologia 164: 213220.Google Scholar
Chaddick, PR & Leek, FF (1972) Further specimens of stored product insects found in an ancient Egyptian tomb. Journal of Stored Product Research 8: 8386.Google Scholar
Chailleux, A, Mohl, EK, Alves, MT, Messelink, GJ & Desneux, N (2014) Natural enemy-mediated indirect interactions among prey species: potential for enhancing biocontrol services in agroecosystems. Pest Management Science 70: 17691779.Google Scholar
Channer, AGdR & Gowen, SR (1992) Selection for increased host resistance and increased pathogen specificity in the Meloidogyne-Pasteuria penetrans interaction. Fundamental and Applied Nematology 15: 331339.Google Scholar
Chaplin-Kramer, R, O’Rourke, ME, Blitzer, EJ & Kremen, C (2011) A meta-analysis of crop pest and natural enemy response to landscape complexity. Ecology Letters 14: 922932.Google Scholar
Charleston, DS, Kfir, R, Dicke, M & Vet, LEM (2006) Impact of botanical extracts derived from Melia azedarach and Azadirachta indica on populations of Plutella xylostella and its natural enemies: a field test of laboratory findings. Biological Control 39: 105114.Google Scholar
Charudattan, R (2001) Biological control of weeds by means of plant pathogens: significance for integrated weed management in modern agro-ecology. BioControl 46: 229260.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.Google Scholar
Charudattan, R, Chandramohan, S & Wyss, GS (2002) Biological control. In Pesticides in Agriculture and the Environment (Wheeler, WB, ed.), New York, Marcel Dekker, pp. 2558.Google Scholar
Chaston, JM, Dillman, AR, Shapiro-Ilan, DI, Bilgrami, AL, Gaugler, R, Hopper, KR, Adams, BJ (2011) Outcrossing and crossbreeding recovers deteriorated traits in laboratory cultured Steinernema carpocapsae nematodes. International Journal for Parasitology 41: 801809.Google Scholar
Chater, KF, Biro, S, Lee, KJ, Palmer, T & Schrempf, H (2010) The complex extracellular biology of Streptomyces. Fems Microbiology Reviews 34: 171198.Google Scholar
Chavalle, S, Buhl, PN, Censier, F & De Proft, M (2015) Comparative emergence phenology of the orange wheat blossom midge, Sitodiplosis mosellana (Gehin) (Diptera: Cecidomyiidae) and its parasitoids (Hymenoptera: Pteromalidae and Platygastridae) under controlled conditions. Crop Protection 76: 114120.Google Scholar
Chen, C-J & Thiem, SM (1997) Differential infectivity of two Autographa californica nucleopolyhedrovirus mutants on three permissive cell lines is the result of lef-7 deletion. Virology 227: 8895.Google Scholar
Cherry, AJ, Jenkins, NE, Heviefo, G, Bateman, R & Lomer, CJ (1999) Operational and economic analysis of a West African pilot-scale production plant for aerial conidia of Metarhizium spp. for use as a mycoinsecticide against locusts and grasshoppers. Biocontrol Science and Technology 9: 3551.Google Scholar
Chesson, PL & Murdoch, WW (1986) Aggregation of risk relationships among host-parasitoid models. American Naturalist 127: 696715.Google Scholar
Chilcutt, CF & Tabashnik, BE (1999) Simulation of integration of Bacillus thuringiensis and the parasitoid Cotesia plutellae (Hymenoptera: Braconidae) for control of susceptible and resistant diamondback moth (Lepidoptera: Plutellidae). Environmental Entomology 28: 505512.Google Scholar
Childs, MR (2006) Comparison of gila topminnow and western mosquitofish as biological control agents of mosquitoes. Western North American Naturalist 66: 181190.Google Scholar
Chobot, V, Huber, C, Trettenhahn, G & Hadacek, F (2009) (+/−)-Catechin: chemical weapon, antioxidant, or stress regulator? Journal of Chemical Ecology 35: 980996.Google Scholar
Chun, YJ, Van Kleunen, M & Dawson, W (2010) The role of enemy release, tolerance and resistance in plant invasions: linking damage to performance. Ecology Letters 13: 937946.Google Scholar
Cilliers, CJ & Neser, S (1991) Biological control of Lantana camara (Verbenaceae) in South Africa. Agriculture, Ecosystems and Environment 37: 5775.Google Scholar
Civeyrel, L & Simberloff, D (1996) A tale of two snails: is the cure worse than the disease? Biodiversity and Conservation 5: 12311252.Google Scholar
Clarke, B, Murray, J & Johnson, MS (1984) The extinction of endemic species by a program of biological control. Pacific Science 38: 97104.Google Scholar
Clausen, CP (1978) Introduced Parasites and Predators of Arthropod Pests and Weeds: A World Review. Washington, DC, USDA Agriculture Research Service.Google Scholar
Clay, K (2014) Defensive symbionts: a microbial perspective. Functional Ecology 28: 293298.Google Scholar
Clay, K & Kover, PX (1996) The red queen hypothesis and plant/pathogen interactions. Annual Review of Phytopathology 34: 2950.Google Scholar
Clem, RJ, Fechheimer, M & Miller, LK (1991) Prevention of apoptosis by a baculovirus gene during infection of insect cells. Science 254: 13881390.Google Scholar
Clem, RJ & Miller, LK (1993) Apoptosis reduces both the in vitro replication and the in vivo infectivity of a baculovirus. Journal of Virology 67: 37303738.Google Scholar
Clewley, GD, Eschen, R, Shaw, RH & Wright, DJ (2012) The effectiveness of classical biological control of invasive plants. Journal of Applied Ecology 49: 12871295.Google Scholar
Coatzee, JA, Hill, MP, Byrne, MJ & Bownes, AA (2011) A review of the biological control programmes on Eichhornia crassipes (C.Mart.) Solms (Pontederiaceae), Salvinia molesta DS Mitch. (Salviniaceae), Pistia stratiotes L. (Araceae), Myriophyllum aquaticum (Vell.) Verdc. (Haloragaceae) and Azolla filiculoides Lam. (Azollaceae) in South Africa. African Entomology 19: 451468.Google Scholar
Cock, MJW (1985) A Review of Biological Control of Pests in the Commonwealth Caribbean and Bermuda up to 1982. Slough, UK, Commonwealth Agricultural Bureau.Google Scholar
Cock, MJW (1986) Requirements for biological control: an ecological perspective. Biocontrol News and Information 7: 717.Google Scholar
Cock, MJW, van Lenteren, JC, Brodeur, J, Barratt, BIP, Bigler, F, Bolckmans, K, Consoli, FL, Haas, F, Mason, PG & Parra, JRP (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
Cohen, AC (2004) Insect Diets: Science and Technology. Boca Raton, FL, CRC Press.Google Scholar
Cohen, AC & Smith, LK (1998) A new concept in artificial diets for Chrysoperla rufilabris: the efficacy of solid diets. Biological Control 13: 4954.Google Scholar
Cohen, MF & Mazzola, M (2006) Resident bacteria, nitric oxide emission and particle size modulate the effect of Brassica napus seed meal on disease incited by Rhizoctonia solani and Pythium spp. Plant and Soil 286: 7586.Google Scholar
Cohen, MF, Yamasaki, H & Mazzola, M (2005) Brassica napus seed meal soil amendment modifies microbial community structure, nitric oxide production and incidence of Rhizoctonia root rot. Soil Biology & Biochemistry 37: 12151227.Google Scholar
Colautti, RI, Ricciardi, A, Grigorovich, IA & MacIsaac, HJ (2004) Is invasion success explained by the enemy release hypothesis? Ecology Letters 7: 721733.Google Scholar
Coll, M (2004) Precision agriculture approaches in support of ecological engineering for pest management. In Ecological Engineering for Pest Management (Gurr, GM, ed.), Ithaca, NY, Cornell University Press, pp. 113142.Google Scholar
Coll, M & Abd-Rabou, S (1998) Effect of oil emulsion sprays on parasitoids of the black parlatoria, Parlatoria ziziphi, in grapefruit. BioControl 43: 2937.Google Scholar
Coll, M & Bottrell, DG (1994) Effects of nonhost plants on an insect herbivore in diverse habitats. Ecology 73: 723731.Google Scholar
Coll, M & Bottrell, DG (1995) Predator–prey association in monocultures and dicultures – effect of maize and bean vegetation. Agriculture Ecosystems & Environment 54: 115125.Google Scholar
Coll, M & Bottrell, DG (1996) Movement of an insect parasitoid in simple and divers plant assemblages. Ecological Entomology 21: 141149.Google Scholar
Coll, M, De Mendoza, LG & Roderick, GK (1994) Population structure of a predatory beetle: the importance of gene flow for intertrophic level interactions. Heredity 72: 228236.Google Scholar
Coll, M & Guershon, M (2002) Omnivory in terrestrial arthropods: mixing plant and prey diets. Annual Review of Entomology 47: 267298.Google Scholar
Collier, TR, Murdoch, WW & Nisbet, RM (1994) Egg load and the decision to host-feed in the parasitoid, Aphytis melinus. Journal of Animal Ecology 63: 299306.Google Scholar
Collier, T & Van Steenwyk, R (2004) A critical evaluation of augmentative biological control. Biological Control 31: 245256.Google Scholar
Collier, T & Van Steenwyk, R (2006) How to make a convincing case for augmentative biological control. Biological Control 39: 119120.Google Scholar
Collins, DP & Jacobsen, BJ (2003) Optimizing a Bacillus subtilis isolate for biological control of sugar beet cercospora leaf spot. Biological Control 26: 153161.Google Scholar
Collins, KL, Boatman, ND, Wilcox, A & Holland, JM (2003a) A 5-year comparison of overwintering polyphagous predator densities within a beetle bank and two conventional hedgebanks. Annals of Applied Biology 143: 6371.Google Scholar
Collins, KL, Boatman, ND, Wilcox, A & Holland, JM (2003b) Effects of different grass treatments used to create overwintering habitat for predatory arthropods on arable farmland. Agriculture Ecosystems & Environment 96: 5967.Google Scholar
Collins, KL, Boatman, ND, Wilcox, A, Holland, JM & Chaney, K (2002) Influence of beetle banks on cereal aphid predation in winter wheat. Agriculture Ecosystems & Environment 93: 337350.Google Scholar
Collyer, E (1964) The effect of an alternative food supply between two Typhlodromus species and Panonychus ulmi (Kock) (Acarina). Entomologia Experimentalis et Applicata 7: 120124.Google Scholar
Connor, EF, Faeth, SH, Simberloff, D & Opler, PA (1980) Taxonomic isolation and accumulation of herbivorous insects: a comparison of introduced and native trees. Ecological Entomology 5: 205211.Google Scholar
Conway, HE, Steinkraus, DC, Ruberson, JR & Kring, TJ (2006) Experimental treatment threshold for the cotton aphid (Homoptera: Aphididae) using natural enemies in Arkansas cotton. Journal of Entomological Science 41: 361373.Google Scholar
Cook, JM (1993). Inbred lines as reservoirs of sex alleles in parasitoid rearing programs. Environmental Entomology 22: 12131216.Google Scholar
Cook, RJ (2007) Take-all decline: model system in the science of biological control and clue to the success of intensive cropping. In Biological Control: A Global Perspective (Vincent, C, Goettel, M & Lazarovits, G, eds.), Wallingford, UK, CABI Publishing, pp. 399413.Google Scholar
Coombes, DS & Sotherton, NW (1986) The dispersal and distribution of polyphagous predatory Coleoptera in cereals. Annals of Applied Biology 108: 461474.Google Scholar
Coombs, MT (2000) Seasonal phenology, parasitism, and evaluation of mowing as a control measure for Nezara viridula (Hemiptera: Pentatomidae) in Australian pecans. Environmental Entomology 29: 10271033.Google Scholar
Coombs, M (2004) Estimating the host range of the tachinid Trichopoda giacomellii, introduced into Australia for biological control of the green vegetable bug. In Assessing Host Ranges of Parasitoids and Predators Used for Classical Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, USDA Forest Service, pp. 143151.Google Scholar
Coquillard, P, Thibaut, T, Hill, DRC, Gueugnot, J, Mazel, C & Coquillard, Y (2000) Simulation of the mollusc Ascoglossa Elysia subornata population dynamics: application to the potential biocontrol of Caulerpa taxifolia growth in the Mediterranean Sea. Ecological Modelling 135: 116.Google Scholar
Corbett, A (1998) The importance of movement in the response of natural enemies to habitat manipulation. In Enhancing Biological Control (Pickett, CH & Bugg, RL, eds.), Berkeley, University of California Press, pp. 2548.Google Scholar
Corbett, A, Leigh, TF & Wilson, LT (1991) Interplanting alfalfa as a source of Mataseiulus occidentalis (Acari: Phytoseiidae) for managing spider mites in cotton. Biological Control 1: 188196.Google Scholar
Corbett, A & Plant, RE (1993) Role of movement in the response of natural enemies to agroecosystem diversification: a theoretical evaluation. Environmental Entomology 22: 519531.Google Scholar
Corbett, A & Rosenheim, JA (1996) Impact of a natural enemy overwintering refuge and its interaction with the surrounding landscape. Ecological Entomology 21: 155164.Google Scholar
Cornell, HV & Hawkins, BA (1993) Accumulation of native parasitoid species on introduced herbivores: a comparison of hosts as natives and hosts as invaders. American Naturalist 141: 847865.Google Scholar
Cornwallis, LJ, Stewart, A, Bourdot, GW, Gaunt, RE, Harvey, IC & Saville, DJ (1999) Pathogenicity of Sclerotinia sclerotiorum on Ranunculus acris is in dairy pasture. Biocontrol Science and Technology 9: 365377.Google Scholar
Cory, JS (2003) Ecological impacts of virus insecticides: host range and non-target organisms. In Environmental Impact of Microbial Insecticides (Hokkanen, HMT & Hajek, AE, eds.), Dordrecht, The Netherlands, Kluwer, pp. 7392.Google Scholar
Cory, JS (2007) Field tests in the UK of a genetically modified virus. In Biological Control: A Global Perspective (Vincent, C, Goettel, MS & Lazarovotis, G, eds.), Wallingford, UK, CABI Publishing, pp. 362373.Google Scholar
Cory, JS & Franklin, MT (2012) Evolution and microbial control of insects. Evolutionary Applications 5: 455469.Google Scholar
Cory, JS, Hirst, ML, Williams, T, Hails, RS, Goulson, D, Green, BM, Carty, TM, Possee, RD, Cayley, PJ & Bishop, DHL (1994) Field trial of a genetically improved baculovirus insecticide. Nature 370: 138140.Google Scholar
Costa, A & Stary, P (1988) Lysiphlebus testaceipes, an introduced aphid parasitoid in Portugal (Hym.: Aphidiidae). Entomophaga 33: 403412.Google Scholar
Costamagna, AC, Menalled, FD & Landis, DA (2004) Host density influences parasitism of the armyworm Pseudaletia unipuncta in agricultural landscapes. Basic and Applied Ecology 5: 347355.Google Scholar
Costanza, R, d’Arge, R, de Groot, R, Farber, S, Grasso, M, Hannon, B, Limburg, K, O’Neill, RV, Paruelo, J, Raskin, RG, Naeem, S, Sutton, PC & Van Den Belt, M (1997) The value of the world’s ecosystem services and natural capital. Nature 387: 253260.Google Scholar
Costello, MJ & Daane, KM (2003) Spider and leafhopper (Erythroneura spp.) response to vineyard ground cover. Environmental Entomology 32: 10851098.Google Scholar
Cottrell, TE (2004) Suitability of exotic and native lady beetle eggs (Coleoptera: Coccinellidae) for development of lady beetle larvae. Biological Control 31: 362371.Google Scholar
Cottrell, TE (2005) Predation and cannibalism of lady beetle eggs by adult lady beetles. Biological Control 34: 159164.Google Scholar
Cottrell, TE & Yeargan, KV (1998) Intraguild predation between an introduced lady beetle, Harmonia axyridis(Coleoptera: Coccinellidae), and a native lady beetle, Coleomegilla maculata (Coleoptera: Coccinellidae). Journal of the Kansas Entomological Society 71: 159163.Google Scholar
Cotxarrera, L, Trillas-Gay, MI, Steinberg, C & Alabouvette, C (2002) Use of sewage sludge compost and Trichoderma asperellum isolates to suppress Fusarium wilt of tomato. Soil Biology & Biochemistry 34: 467476.Google Scholar
Coupland, R & Baker, G (2007) Search for biological control agents of invasive Mediterranean snails. In Biological Control: A Global Perspective (Vincent, C, Goettel, MS & Lazarovits, G, eds.), Wallingford, UK, CABI Publishing, pp. 712.Google Scholar
Courchamp, F, Berec, L & Gasciogne, J (2008) Allee Effects in Ecology and Conservation. Oxford, UK, Oxford University Press.Google Scholar
Courchamp, F, Chapuis, J-L & Pascal, M (2003) Mammal invaders on islands: impact, control and control impact. Biological Reviews 78: 347383.Google Scholar
Courchamp, F & Cornell, SJ (2000) Virus-vectored immunocontraception to control feral cats on islands: a mathematical model. Journal of Applied Ecology 37: 903913.Google Scholar
Courchamp, F & Sugihara, G (1999) Modeling the biological control of an alien predator to protect island species from extinction. Ecological Applications 9: 112123.Google Scholar
Courtenay, WR & Meffe, GK (1989) Small fishes in strange places: a review of introduced poeciliids. In Ecology and Evolution of Livebearing Fishes (Meffe, GK & Snelson, FF, eds.), Englewood Cliffs, NJ, Prentice Hall, pp. 319331.Google Scholar
Cousens, R & Croft, AM (2000) Weed populations and pathogens. Weed Research 40: 6382.Google Scholar
Cowan, DP & Stahlhut, JK (2004) Functionally reproductive diploid and haploid males in an inbreeding hymenopteran with complementary sex determination. Proceedings of the National Academy of Sciences, USA, 101: 1037410379.Google Scholar
Cowie, RH (2001) Can snails ever be effective and safe biocontrol agents? International Journal of Pest Management 47: 2340.Google Scholar
Crawley, MJ (1986) The population biology of invaders. Philosophical Transactions of the Royal Society of London, Series B 314: 711731.Google Scholar
Crawley, MJ (1987) What makes a community invasible? In Colonization, Succession and Stability (Gray, AJ, Crawley, MJ & Edwards, PJ, eds.), Oxford, UK, Blackwell, pp. 429453.Google Scholar
Crawley, MJ (1989a) Insect herbivores and plant population dynamics. Annual Review of Entomology 34: 531564.Google Scholar
Crawley, MJ (1989b) The successes and failures of weed biocontrol using insects. Biocontrol News and Information 10: 213223.Google Scholar
Crawley, MJ (1997) Plant-herbivore dynamics. In Plant Ecology (Crawley, MJ, ed.), Oxford, UK, Blackwell, pp. 401474.Google Scholar
Croft, BA (1990) Arthropod Biological Control Agents and Pesticides. New York, John Wiley & Sons.Google Scholar
Croizier, G, Croizier, L, Argaud, O & Poudevigne, D (1994) Extension of Autographa californica nuclear polyhedrosis virus host range by interspecific replacement of a short DNA sequence in the p143 helicase gene. Proceedings of the National Academy of Sciences of the United States of America 91: 4852.Google Scholar
Crone, EE, Menges, ES, Ellis, MM, Bell, T, Bierzychudek, P, Ehrlén, J, Kaye, TN, Knight, TM, Lesica, P, Morris, WF, Oostermeijer, G, Quintana-Ascensio, PF, Stanley, A, Ticktin, T, Valverde, T & Williams, JL (2011) How do plant ecologists use matrix population models? Ecology Letters 14: 18.Google Scholar
Cronin, JT & Reeve, JD (2005) Host-parasitoid spatial ecology: a plea for a landscape-level synthesis. Proceedings of the Royal Society B-Biological Sciences 272: 22252235.Google Scholar
Cronin, JT & Reeve, JD (2014) An integrative approach to understanding host-parasitoid population dynamics in real landscapes. Basic and Applied Ecology 15: 101113.Google Scholar
Cronin, JT, Reeve, JD, Wilkens, R & Turchin, P (2000) The pattern and range of movement of a checkered beetle predator relative to its bark beetle prey. Oikos 90: 127138.Google Scholar
Crook, NE, Clem, RJ & Miller, LK (1993) An apoptosis-inhibiting baculovirus gene with a zinc finger-like motif. Journal of Virology 67: 21682174.Google Scholar
Cross, JV, Burgess, CM & Hanks, GR (1996) Integrating insecticide use with biological control of two spotted spider mite (Tetranychus urticae) by Phytoseiulus persimilis on strawberry in the UK. In Brighton Crop Protection Conference: Pests and Diseases, Volume 3. Alton, UK, British Crop Protection Council, pp. 899906.Google Scholar
Crowder, DW (2007) Impact of release rates on the effectiveness of augmentative biological control agents. Journal of Insect Science 7: 15.Google Scholar
Crowley, PH & Martin, EK (1989) Functional responses and interference within and between year classes of a dragonfly population. Journal of the North American Benthological Society 8: 211221.Google Scholar
Crozier, RH, Newey, PS, Schluns, EA & Robson, SKA (2010) A masterpiece of evolution – Oecophylla weaver ants (Hymenoptera: Formicidae). Myrmecological News 13: 5771.Google Scholar
Crump, NS, Cother, EJ & Ash, GJ (1999) Clarifying the nomenclature in microbial weed control. Biocontrol Science and Technology 9: 8997.Google Scholar
Cullen, JM & Delfosse, ES (1985) Echium plantanineum: catalyst for conflict and change in Australia. In Proceedings of the VI International Symposium on Biological Control (Delfosse, ES, ed.), Ottawa, Canada, Agriculture Canada, pp. 249292.Google Scholar
Cullen, JM, Kable, PF & Catt, M (1973) Epidemic spread of a rust imported for biological control. Nature 244: 462464.Google Scholar
Cullen, JM, McFadyen, REC & Julien, MH (2011) One hundred years of biological control of weeds in Australia. Proceedings of the XIII International Symposium of Weed Biological Control (Wu, Y, Johnson, TD, Singh, S, Raghu, S, Wheeler, G, Pratt, P, Warner, K, Center, T, Goolsby, J & Reardon, R, eds.), Washington, DC, U.S. Forest Service, pp. 360367.Google Scholar
Cullen, J & Sheppard, AW (2012) Carduus nutans L. – nodding thistle. In Biological Control of Weeds in Australia (Julien, M, McFadyen, R & Cullen, J, eds.), Canberra, Australia, CSIRO, pp. 118130.Google Scholar
Culliney, TW (2005) Benefits of classical biological control for managing invasive plants. Critical Reviews in Plant Sciences 24: 131150.Google Scholar
Culver, JJ (1919) A study of Compsilura concinnata, an imported tachinid parasite of the gipsy moth and the brown-tail moth. U.S. Department of Agriculture Bulletin 766: 127.Google Scholar
Cunniffe, NJ & Gilligan, CA (2011) A theoretical framework for biological control of soil-borne plant pathogens: identifying effective strategies. Journal of Theoretical Biology 278: 3243.Google Scholar
Currie, CR, Scott, JA, Summerbell, RC & Malloch, D (1999) Fungus-growing ants use antibiotic-producing bacteria to control garden parasites. Nature 398: 701704.Google Scholar
Currie, GA & Fyfe, RV (1938) The fate of certain European insects introduced into Australia for the control of weeds. Journal of the Council for Scientific and Industrial Research 11: 289301.Google Scholar
Curtis, CF (2000) The case for deemphasizing genomics in malaria control. Science 230: 1508.Google Scholar
Cutting, KJ & Hough-Goldstein, J (2013) Integration of biological control and native seeding to restore invaded plant communities. Restoration Ecology 21: 648655.Google Scholar
Da Ros, N, Ostermayer, R, Roques, A & Raimbault, JP (1993) Insect damage to cones of exotic conifer species introduced in arboreta I. Interspecific variations within the genus Picea. Journal of Applied Entomology 115: 113133.Google Scholar
Daehler, CC & Strong, DR (1997) Reduced herbivore resistance in an introduced smooth cordgrass (Spartina alterniflora) after a century of herbivore-free growth. Oecologia 110: 99108.Google Scholar
Dambroski, HR, Lin, JC, Berlocher, SH, Forbes, AA, Roelofs, W & Feder, JL (2005) The genetic basis for fruit odor discrimination in Rhagoletis flies and its significance for sympatric host shifts. Evolution 59: 19531964.Google Scholar
D’Antonio, CM (1993) Mechanisms controlling invasion of coastal plant communities by the alien succulent Carpobrotus edulis. Ecology 74: 8395.Google Scholar
D’Antonio, CM & Dudley, TL (1995) Biological invasions as agents of change on islands versus mainlands. In Islands: Biological Diversity and Ecosystem Function (Vitousek, PM, Loope, LL & Adsersen, H, eds.), Berlin, Germany, Springer, pp. 103122.Google Scholar
D’Antonio, CM, Dudley, TL & Mack, M (1999) Disturbance and biological invasions: direct effects and feedbacks. In Ecosystems of the World: Ecosystems of Disturbed Ground (Walker, LR, ed.), Amsterdam, The Netherlands, Elsevier, pp. 413452.Google Scholar
Danyk, TP & Mackauer, M (1996) An extraserosal envelope in eggs of Praon pequodorum (Hymenoptera: Aphidiidae), a parasitoid of pea aphid. Biological Control 7: 6770.Google Scholar
Darwin, C (1859) On the Origin of Species. Middlesex, UK, Penguin Books.Google Scholar
Dauer, JT, McEvoy, PB & Van Sickle, J (2012) Controlling a plant invader by targeted disruption of its life cycle. Journal of Applied Ecology 49: 322330.Google Scholar
Daugherty, MP, Harmon, JP & Briggs, CJ (2007) Trophic supplements to intraguild predation. Oikos 11: 662677.Google Scholar
David, AS, Kaser, JM, Morey, AC, Roth, AM & Andow, DA (2013) Release of genetically engineered insects: a framework to identify potential ecological effects. Ecology and Evolution 3: 40004015.Google Scholar
Davies, AP, Takashino, K, Watanabe, M & Miura, K (2009) Parental genetic traits in offspring from inter-specific crosses between introduced and indigenous Diadegma Foerster (Hymenoptera: Ichneumonidae): possible effects on conservation genetics. Applied Entomology and Zoology 44: 535541.Google Scholar
Davis, MA (2009) Invasion Biology. Oxford, UK, Oxford University Press.Google Scholar
Davis, AS, Landis, DA, Nuzzo, V, Blossey, B, Hinz, H & Gerber, E (2006) Demographic models inform selection of biocontrol agents for garlic mustard (Alliaria petiolata). Ecological Applications 16: 23992410.Google Scholar
Dawkins, R & Krebs, JR (1979) Arms races between and within species. Proceedings of the Royal Society of London, Series B 205: 489511.Google Scholar
Day, WH (1970) The survival value of its jumping cocoons to Bathyplectes anurus, a parasite of the alfalfa weevil. Journal of Economic Entomology 63: 586589.Google Scholar
Day, WH (1981) Biological control of the alfalfa weevil in the northeastern United States. In Biological Control in Crop Production (Papavizas, GC, ed.), London, UK, Allanheld, Osmun, & Co, pp. 361374.Google Scholar
Day, WH (2005) Changes in abundance of native and introduced parasites (Hymenoptera: Braconidae), and of the target and non-target plant bug species (Hemiptera: Miridae), during two classical biological control programs in alfalfa. Biological Control 33: 368374.Google Scholar
Day, WH, Eaton, AT, Romig, RF, Tilmon, KJ, Mayer, M & Dorsey, T (2003) Peristenus digoneutis (Hymenoptera: Braconidae), a parasite of Lygus lineolaris (Hemiptera: Miridae) in northeastern United States alfalfa, and the need for research on other crops. Entomological News 114: 105111.Google Scholar
Day, WH, Prokrym, DR, Ellis, DR & Chianese, RJ (1994) The known distribution of the predator Propylea quattuordecimpunctata (Coleoptera, Coccinellidae) in the United States, and thoughts on the origin of this species and 5 other exotic lady beetles in eastern North America. Entomological News 105: 244256.Google Scholar
DeAngelis, DL, Goldstein, RA & O’Neill, RV (1975) A model for trophic interaction. Ecology 56: 881892.Google Scholar
de Boer, JG & Dicke, M (2005) Information use by the predatory mite Phytoseiulus persimilis (Acari: Phytoseiidae), a specialised natural enemy of herbivorous spider mites. Applied Entomology and Zoology 40: 112.Google Scholar
de Boer, JG, Groenen, MAM, Pannebakker, BA, Beukeboom, LW & Kraus, RHS (2015) Population-level consequences of complementary sex determination in a solitary parasitoid. BMC Evolutionary Biology 15: 98.Google Scholar
de Boer, JG, Kuijper, B, Heimpel, GE & Beukeboom, LW (2012) Sex determination meltdown upon biological control introduction of the parasitoid Cotesia rubecula? Evolutionary Applications 5: 444454.Google Scholar
de Boer, JG, Ode, PJ, Rendahl, AK, Vet, LEM, Whitfield, JB & Heimpel, GE (2008) Experimental support for multiple-locus complementary sex determination in the parasitoid Cotesia vestalis. Genetics 180: 15251535.Google Scholar
de Boer, Ode PJ, Vet, LEM, Whitfield, J & Heimpel, GE (2007) Diploid males sire triploid daughters and sons in the parasitoid wasp Cotesia vestalis. Heredity 99: 288294.Google Scholar
de Briano, AEE, Acciaresi, HA & Briano, JA (2013) Establishment, dispersal, and prevalence of Rhinocyllus conicus (Coleoptera: Curculionidae), a biological control agent of thistles, Carduus species (Asteraceae), in Argentina, with experimental information on its damage. Biological Control 67: 186193.Google Scholar
De Clercq, P, Mason, PG & Babendreier, D (2011) Benefits and risks of exotic biological control agents. BioControl 56: 681698.Google Scholar
de Jong, MD, Aylor, DE & Bourdot, GW (1999) A methodology for risk analysis of plurivorous fungi in biological weed control: Scleroinia sclerotiorum as a model. BioControl 43: 397419.Google Scholar
de Jong, MD, Bourdot, GW, Powell, J & Gourdiaan, J (2002) A model of the escape of Sclerotinia sclerotiorum ascospores from pasture. Ecological Modelling 150: 83105.Google Scholar
De La Fuente, L, Mavrodi, DV, Landa, BB, Thomashow, LS & Weller, DM (2006) phlD-based genetic diversity and detection of genotypes of 2,4-diacetylphloroglucinol-producing Pseudomonas fluorescens. FEMS Microbiology Ecology 56: 6478.Google Scholar
De la Mora, A, Livingston, G & Philpott, SM (2008) Arboreal ant abundance and leaf miner damage in coffee agroecosystems in Mexico. Biotropica 40: 742746.Google Scholar
De Nardo, EAB, Grewal, PS, McCartney, D & Stinner, BR (2006) Non-target effects of entomopathogenic nematodes on soil microbial community and nutrient cycling processes: a microcosm study. Applied Soil Ecology 34: 250257.Google Scholar
De Nardo, EAB & Hopper, KR (2004) Using the literature to evaluate parasitoid host ranges: a case study of Macrocentrus grandii (Hymenoptera: Braconidae) introduced into North America to control Ostrinia nubilalis (Lepidoptera: Crambidae). Biological Control 31: 280295.Google Scholar
de Oliveira, ACS, Martins, SGF & Zacarias, MS (2013) An individual-based model for the interaction of the mite Tetranychus urticae (Koch, 1836) with its predator Neoseiulus californicus (McGregor, 1954) (Acari: Tetranychidae, Phytoseiidae). Ecological Modelling 255: 1120.Google Scholar
de Roos, AM, McCauley, E & Wilson, WG (1991) Mobility versus density-limited predator–prey dynamics on different spatial scales. Proceedings of the Royal Society of London B 246: 117122.Google Scholar
de Roos, AM, McCauley, E & Wilson, WG (1998) Pattern formation and the spatial scale of interaction between predators and their prey. Theoretical Population Biology 53: 108130.Google Scholar
DeBach, P (1946) An insecticidal check method for measuring the efficacy of entomophagous insects. Journal of Economic Entomology 39: 695697.Google Scholar
DeBach, P (1955) Validity of the insecticidal check method as a measure of the effectiveness of natural enemies of diaspine scale insects. Journal of Economic Entomology 48: 584588.Google Scholar
DeBach, P (1964) Biological Control of Insect Pests and Weeds. New York, Reinhold.Google Scholar
DeBach, P (1965) Some biological and ecological phenomena associated with colonizing enomophagous insects. In The Genetics of Colonizing Species (Baker, HG & Stebbins, GL, eds.), New York, Academic Press, pp. 287303.Google Scholar
DeBach, P (1971) The use of imported natural enemies in insect pest management ecology. Proceedings of the Tall Timbers Conference on Ecological Animal Control by Habitat Management 3: 211233.Google Scholar
DeBach, P (1972) The use of imported natural enemies in insect pest management ecology. Proceedings of the Tall Timbers Conference on Ecological Animal Control by Habitat Management 3: 211233.Google Scholar
DeBach, P (1974) Biological Control by Natural Enemies. Cambridge, UK, Cambridge University Press.Google Scholar
DeBach, P & Hagen, KS (1964) Manipulation of entomophagous species. In Biological Control of Insect Pests and Weeds (DeBach, P, ed.), New York, Reinhold, pp. 429458.Google Scholar
DeBach, P & Landi, J (1961) The introduced purple scale parasite, Aphytis lepidosaphes Compere, and a method of integrating chemical with biological control. Hilgardia 31: 459496.Google Scholar
DeBach, P & Rosen, D (1991) Biological Control by Natural Enemies. Second edition. Cambridge, UK, Cambridge University Press.Google Scholar
DeBach, P & Sundby, RA (1963) Competitive displacement between ecological homologues. Hilgardia 34: 105166.Google Scholar
Decou, CG (1994) Biological control of the two-spotted spider mite (Acarina: Tetranychidae) on commercial strawberries in Florida with Phytoseiulus persimilis (Acarina: Phytoseiidae). Florida Entomologist 77: 3341.Google Scholar
Delfosse, ES (1985) Echium plantagineum in Australia: effects of a major conflict of interest. In Proceedings of the VI International Symposium on Biological Control (Delfosse, ES, ed.), Ottawa, Canada, Agriculture Canada, pp. 293299.Google Scholar
Delfosse, ES (2005) Risk and ethics in biological control. Biological Control 35: 319329.Google Scholar
Dempster, JP (1983) The natural control of populations of butterflies and moths. Biological Reviews 58: 461481.Google Scholar
Dennis, P & Fry, GLA (1992) Field margins – can they enhance natural enemy population densities and general arthropod diversity on farmland? Agriculture Ecosystems & Environment 40: 95115.Google Scholar
Dennis, P, Thomas, MB & Sotherton, NW (1994) Structural features of field boundaries which influence the overwintering densities of beneficial arthropod predators. Journal of Applied Ecology 31: 361370.Google Scholar
Denno, RF & Finke, DL (2006) Multiple predator interactions and foodweb connectance implications for biological control. In Trophic and Guild Interactions in Biological Control (Brodeur, J & Boivin, G, eds.), Dordrecht, The Netherlands, Springer, pp. 2144.Google Scholar
Denoth, M, Frid, L & Myers, JH (2002) Multiple agents in biological control: improving the odds? Biological Control 24: 2030.Google Scholar
Denoth, M & Myers, JH (2005) Variable success of biological control of Lythrum salicaria in British Columbia. Biological Control 32: 269279.Google Scholar
Denslow, JS (2003) Weeds in paradise: thoughts on the invasibility of tropical islands. Annals of the Missouri Botanical Garden 90: 119127.Google Scholar
Desneux, N, Blahnik, R, Delebecque, CJ & Heimpel, GE (2012) Host phylogeny and specialisation in parasitoids. Ecology Letters 15: 453460.Google Scholar
Desneux, N, Decourtye, A & Delpuech, JM (2007) The sublethal effects of pesticides on beneficial arthropods. Annual Review of Entomology 52: 81106.Google Scholar
DeWalt, SJ, Denslow, JS & Ickes, K (2004) Natural-enemy release facilitates habitat expansion of the invasive tropical shrub, Clidemia hirta. Ecology 85: 471483.Google Scholar
Dhileepan, K (2003) Evaluating the effectiveness of weed biocontrol at the local scale. In Improving the Selection, Testing and Evaluation of Weed Biological Control Agents (Jacob, HS & Briese, DT, eds.), Osmond, Australia, CRC, pp. 5160.Google Scholar
Dhileepan, K, Taylor, DBJ, McCarthy, J, King, A & Shabbir, A (2013) Development of cat’s claw creeper leaf-tying moth Hypocosmia pyrochroma (Lepidoptera: Pyralidae) at different temperatures: Implications for establishment as a biological control agent in Australia and South Africa. Biological Control 67: 194202.Google Scholar
Dhileepan, K, Trevino, M, Donnelly, GP & Raghu, S (2005) Risk to non-target plants from Charidotis auroguttata (Chrysomelidae: Coleoptera), a potential biocontrol agent for cat’s claw creeper Macfadyena unguis-cati (Bignoniaceae) in Australia. Biological Control 32: 450460.Google Scholar
Di Giallonardo, F & Holmes, EC (2015a) Exploring host–pathogen interactions through biological control. PLoS Pathogens 11: e1004865.Google Scholar
Di Giallonardo, F & Holmes, EC (2015b) Viral biocontrol: grand experiments in disease emergence and evolution. Trends in Microbiology 23: 8390.Google Scholar
Di Giusto, B, Anstett, MC, Dounias, E & McKey, DB (2001) Variation in the effectiveness of biotic defence: the case of an opportunistic ant-plant protection mutualism. Oecologia 129: 367375.Google Scholar
Diamond, JM (1997) Guns, Germs and Steel: The Fates of Human Societies. New York, Norton & Co.Google Scholar
Didham, RK, Tylianakis, JM, Hutchison, MA, Ewers, RM & Gemmell, NJ (2005) Are invasive species the drivers of ecological change? Trends in Ecology and Evolution 20: 470474.Google Scholar
Dieckhoff, C (2011) Host acceptance behavior in the soybean aphid parasitoid Binodoxys communis (Hymenoptera: Braconidae) – the role of physiological state in biological control. Thesis, University of Minnesota, Minneapolis, U.S.A.Google Scholar
Dieckhoff, C, Theobald, JC, Wäckers, FL & Heimpel, GE (2014) Egg load dynamics and the risk of egg and time limitation experienced by an aphid parasitoid in the field. Ecology and Evolution 4: 17391750.Google Scholar
Diehl, JK, Holliday, NJ, Lindgren, CJ & Roughley, RE (1997) Insects associated with purple loosestrife, Lythrum salicaria L., in southern Manitoba. Canadian Entomologist 129: 937948.Google Scholar
Dijkerman, HJ (1990) Suitability of 8 Yponomeuta species as hosts of Diadegma armillata. Entomologia Experimentalis et Applicata 54: 173180.Google Scholar
Dixon, AFG (2000) Insect Predator–Prey Dynamics: Ladybird Beetles & Biological Control. Cambridge, UK, Cambridge University Press.Google Scholar
Djonovic, S, Vittone, G, Mendoza-Herrera, A & Kenerley, CM (2007) Enhanced biocontrol activity of Trichoderma virens transformants constitutively coexpressing beta-1,3- and beta-1,6-glucanase genes. Molecular Plant Pathology 8: 469480.Google Scholar
Dobson, AP (1988) Restoring island ecosystems: the potential of parasites to control introduced mammals. Conservation Biology 2: 3139.Google Scholar
Dobson, SL (2003) Reversing Wolbachia-based population replacement. Trends in Parasitology 19: 128133.Google Scholar
Dobson, SL, Fox, CW & Jiggins, FM (2002) The effect of Wolbachia-induced cytoplasmic incompatibility on host population size in natural and manipulated systems. Proceedings of the Royal Society of London, B 269: 437445.Google Scholar
Dodd, AP (1940) The Biological Campaign Against the Prickly-Pear. Brisbane, Australia, Commonwealth Prickly Pear Board.Google Scholar
Domenech, J, Reddy, MS, Kloepper, JW, Ramos, B & Gutierrez-Manero, J (2006) Combined application of the biological product LS213 with Bacillus, Pseudomonas or Chryseobacterium for growth promotion and biological control of soil-borne diseases in pepper and tomato. BioControl 51: 245258.Google Scholar
Dormann, CF, Schymanski, SJ, Cabral, J, Chuine, I, Graham, C, Hartig, F, Kearney, M, Morin, X, Römermann, C & Singer, A (2012) Correlation and process in species distribution models: bridging a dichotomy. Journal of Biogeography 39: 21192131.Google Scholar
Doutt, RL (1964) The historical development of biological control. In Biological Control of Insect Pests and Weeds (DeBach, P, ed.), New York, Halsted, pp. 2144.Google Scholar
Doutt, RL & Nakata, J (1973) The rubus leafhopper and its egg parasitoid: an endemic biotic system useful in grape-pest management. Environmental Entomology 2: 381386.Google Scholar
Dowden, PB (1952) The importance of coordinating applied control and natural control of forest insects. Journal of Economic Entomology 45: 481483.Google Scholar
Driver, F, Milner, RJ & Trueman, JWH (2000) A taxonomic revision of Metarhizium based on a phylogenetic analysis of rDNA sequence data. Mycological Research 104: 134150.Google Scholar
Drukker, B, Janssen, A, Ravensberg, W & Sabelis, MW (1997) Improved control capacity of the mite predator Phytoseiulus persimilis (Acari: Phytoseiidae) on tomato. Experimental and Applied Acarology 21: 507518.Google Scholar
Drukker, B, Yaninek, JS & Herren, HR (1993) A packaging and delivery system for aerial release of Phytoseiidae for biological control. Experimental and Applied Acarology 17: 129143.Google Scholar
Duan, JJ, Bauer, LS, Abell, KJ, Ulyshen, MD & Van Driesche, RG (2015) Population dynamics of an invasive forest insect and associated natural enemies in the aftermath of invasion: implications for biological control. Journal of Applied Ecology 52: 12461254.Google Scholar
Duan, JJ, Lundgren, JG, Naranjo, S & Marvier, M (2010) Extrapolating non-target risk of Bt crops from laboratory to field. Biology Letters 6: 7477.Google Scholar
Duan, JJ & Messing, RH (2000) Evaluating nontarget effects of classical biological control: fruit fly parasitoids in Hawaii as a case study. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 95110.Google Scholar
Dubuffet, A, Dupas, S, Frey, F, Drezen, J-M, Poirié, M & Carton, Y (2007) Genetic interactions between the parasitoid wasp Leptopilina boulardi and its Drosophila host. Heredity 98: 2127.Google Scholar
Dudley, TL & DeLoach, CJ (2004) Saltcedar (Tamarix spp.), endangered species, and biological weed control – Can they mix? Weed Technology 18: 15421551.Google Scholar
Dudley, TL & Kazmer, DJ (2005) Field assessment of the risk posed by Diorhabda elongata, a biocontrol agent for control of saltcedar (Tamarix spp.), to a nontarget plant, Frankenia salina. Biological Control 35: 265275.Google Scholar
Duffy, BK, Simon, A & Weller, DM (1996) Combination of Trichoderma koningii with fluorescent pseudomonads for control of take-all on wheat. Phytopathology 86: 188194.Google Scholar
Dugaw, CJ, Hastings, A, Preisser, EL & Strong, DR (2004) Parasitoid-mediated effects on apparent competition and the persistence of host–parasitoid assemblages. Bulletin of Mathematical Biology 66: 583594.Google Scholar
Dunn, RR (2005) Modern insect extinctions, the neglected majority. Conservation Biology 19: 10301036.Google Scholar
Dupas, S & Carton, Y (1999) Two non-linked genes for specific virulence of Leptopilina boulardi against Drosophila melanogaster and D. yakuba. Evolutionary Ecology 13: 211220.Google Scholar
Durrant, WE & Dong, X (2004) Systemic acquired resistance. Annual Review of Phytopathology 42: 185209.Google Scholar
Dushoff, J & Dwyer, G (2001) Evaluating the risks of engineered viruses: modeling pathogen competition. Ecological Applications 11: 16021609.Google Scholar
Dutcher, JD (2007) A review of resurgence and replacement causing pest outbreaks in IPM. In General Concepts in Integrated Pest and Disease Management (Ciancio, A & Mukerji, KG, eds.), Dordrecht, The Netherlands, Springer, pp. 2743.Google Scholar
Dwyer, G, Levin, SA & Buttel, L (1990) A simulation model of the population dynamics and evolution of myxomatosis. Ecological Monographs 60: 423447.Google Scholar
Eckberg, JO, Tenhumberg, B & Louda, SM (2012) Insect herbivory and propagule pressure influence Cirsium vulgare invasiveness across the landscape. Ecology 93: 17871794.Google Scholar
Eckberg, JO, Tenhumberg, B & Louda, SM (2014) Native insect herbivory limits population growth rate of a non-native thistle. Oecologia 175: 129138.Google Scholar
Ehler, LE (1990) Environmental impact of introduced biological-control agents: implications for agricultural biotechnology. In Risk Assessment in Agricultural Biotechnology (Marois, JJ & Bruening, G, eds.), Wallingford, UK, CABI Publishing, pp. 8596.Google Scholar
Ehler, LE (1998) Invasion biology and biological control. Biological Control 13: 127133.Google Scholar
Ehler, LE (2000) Critical issues related to nontarget effects in classical biological control of insects. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 313.Google Scholar
Ehler, LE (2007) Impact of native predators and parasites on Spodoptera exigua, an introduced pest of alfalfa hay in northern California. Biocontrol 52: 323338.Google Scholar
Ehler, LE, Sforza, R & Mateille, T (2004) Genetics, Evolution and Biological Control. Wallingford, UK, CABI Publishing.Google Scholar
Ehlers, R-U (2003) Biocontrol nematodes. In Environmental Impacts of Microbial Insecticides: Need and Methods for Risk Assessment (Hokkannen, HMT & Hajek, AE, eds.), Dordrecht, The Netherlands, Kluwer, pp. 177220.Google Scholar
Eilenberg, J, Hajek, A & Lomer, C (2001) Suggestions for unifying the terminology in biological control. BioControl 46: 387400.Google Scholar
Elias, J, Mazzi, D & Dorn, S (2009) No need to discriminate? Reproductive diploid males in a parasitoid with complementary sex determination. PLoS One 4: e6024.Google Scholar
Elkinton, JS & Boettner, GH (2012) Benefits and harm caused by the introduced generalist tachinid, Compsilura concinnata, in North America. Biocontrol 57: 277288.Google Scholar
Elkinton, JS, Buonaccorsi, JP, Bellows, TS & Van Driesche, RG (1992) Marginal attack rate, k-values and density dependence in the analysis of contemporaneous mortality factors. Researches on Population Ecology 34: 2944.Google Scholar
Elkinton, JS, Liebhold, AM & Muzika, RM (2004) Effects of alternative prey on predation by small mammals on gypsy moth pupae. Population Ecology 46: 171178.Google Scholar
Elkinton, JS, Parry, D & Boettner, GH (2006) Implicating an introduced generalist parasitoid in the invasive browntail moth’s enigmatic demise. Ecology 87: 26642672.Google Scholar
Elliot, NC, Kieckhefer, R & Kauffman, W (1996) Effects of an invading coccinellid on native coccinellids in an agricultural landscape. Oecologia 105: 537544.Google Scholar
Ellison, CA, Evans, HC & Ineson, J (2004) The significance of intraspecific pathogenicity in the selection of a rust pathotype for the classical biological control of Mikania micrantha (mile-a-minute weed) in Southeast Asia. In Proceedings of the XI International Symposium on Biological Control of Weeds (Cullen, JM, Briese, TD, Kriticos, DJ Lonsdale, WM, Morin, L & Scott, JK, eds.), Canberra, Australia, CSIRO Entomology, pp. 102107.Google Scholar
Ellner, SP, McCauley, E, Kendall, BE, Briggs, CJ, Hosseini, PR, Wood, SN, Janssen, A, Sabelis, MW, Turchin, P, Nisbet, RM & Murdoch, WW (2001) Habitat structure and population persistence in an experimental community. Nature 412: 538543.Google Scholar
El-Tarabily, KA, Hardy, G, Sivasithamparam, K, Hussein, AM & Kurtboke, DI (1997) The potential for the biological control of cavity-spot disease of carrots, caused by Pythium coloratum, by streptomycete and non-streptomycete actinomycetes. New Phytologist 137: 495507.Google Scholar
Elton, CS (1958) The Ecology of Invasions by Animals and Plants. Chicago, IL, Chicago University Press.Google Scholar
Emge, RG, Melching, JS & Kingsolver, CH (1981) Epidemiology of Puccinia chondrillina, a rust pathogen for the biological control of rush skeletonweed in the United States. Phytopathology 71: 839843.Google Scholar
English-Loeb, GM, Rhainds, M, Martinson, T & Ugine, T (2003) Influence of flowering cover crops on Anagrus parasitoids (Hymenoptera: Mymaridae) and Erythroneura leafhoppers (Homoptera: Cicadellidae) in New York vineyards. Agricultural and Forest Entomology 5: 173181.Google Scholar
Epstein, AH & Hill, JH (1999) Status of rose rosette disease as a biological control for multiflora rose. Plant Disease 83: 92101.Google Scholar
Errakhi, R, Bouteau, F, Lebrihi, A & Barakate, M (2007) Evidences of biological control capacities of Streptomyces spp. against Sclerotium rolfsii responsible for damping-off disease in sugar beet (Beta vulgaris L.). World Journal of Microbiology & Biotechnology 23: 15031509.Google Scholar
Espiau, C, Riviere, D, Burdon, JJ Gartner, S, Daclinat, B, Hasan, S & Chaboudez, P (1998) Host-pathogen diversity in a wild system: Chondrilla juncea-Puccinea chondrillina. Oecologia 113: 133139.Google Scholar
Evans, EW (1991) Intra versus interspecific interactions of ladybeetles (Coleoptera: Coccinellidae) attacking aphids. Oecologia 87: 401408.Google Scholar
Evans, EW (2000) Morphology of invasion: body size patterns associated with establishment of Coccinella septempunctata (Coleoptera: Coccinellidae) in western North America. European Journal of Entomology 97: 469474.Google Scholar
Evans, EW (2004) Habitat displacement of North American ladybirds by an introduced species. Ecology 85: 637647.Google Scholar
Evans, EW (2008) Multitrophic interactions among plants, aphids, alternate prey and shared natural enemies – a review. European Journal of Entomology 105: 369380.Google Scholar
Evans, EW, Anderson, MR & Bowling, PD (2010) Targeted sugar provision promotes parasitism of the cereal leaf beetle Oulema melanopus. Agricultural and Forest Entomology 12: 4147.Google Scholar
Evans, EW, Carlile, NR, Innes, MB & Pitigala, N (2013) Warm springs reduce parasitism of the cereal leaf beetle through phenological mismatch. Journal of Applied Entomology 137: 383391.Google Scholar
Evans, EW, Karren, JB & Israelsen, CE (2006) Interactions over time between cereal leaf beetle (Coleoptera: Chrysomelidae) and larval parasitoid Tetrastichus julis (Hymenoptera: Eulophidae) in Utah. Journal of Economic Entomology 99: 19671973.Google Scholar
Evans, JA, Davis, AS, Raghu, S, Ragavendran, A, Landis, DA & Schemske, DW (2012) The importance of space, time, and stochasticity to the demography and management of Alliaria petiolata. Ecological Applications 22: 14971511.Google Scholar
Evans, KJ & Gomez, DR (2004) Genetic markers in rust fungi and their application to weed biocontrol. In Genetics, Evolution and Biological Control (Ehler, LE, Sforza, R & Mateille, T, eds.), Wallingford, UK, CABI Publishing, pp. 7396.Google Scholar
Evenhuis, NL (2002) Hawaii’s Extinct Species – Insects. Website: http://hbs.bishopmuseum.org/endangered/ext-insects.html.Google Scholar
Ewald, JA, Aebischer, NJ, Richardson, SM, Grice, PV & Cooke, AI (2010) The effect of agri-environment schemes on grey partridges at the farm level in England. Agriculture, Ecosystems and Environment 138: 5563.Google Scholar
Facon, B, Crespin, L, Loiseau, A, Lombaert, E, Magro, A & Estoup, A (2010) Can things get worse when an invasive species hybridizes? The harlequin ladybird Harmonia axyridis in France as a case study. Evolutionary Applications 4: 7188.Google Scholar
Facon, B, Hufbauer, RA, Tayeh, A, Loiseau, A, Lombaert, E, Vitalis, R, Guillemaud, T, Lundgren, JG & Estoup, A (2011) Inbreeding depression is purged in the invasive insect Harmonia axyridis. Current Biology 21: 424427.Google Scholar
Fagan, WF & Bishop, JG (2000) Trophic interactions during primary succession: herbivores slow a plant reinvasion at Mount St. Helens. American Naturalist 155: 238251.Google Scholar
Fagan, WF, Lewis, MA, Neubert, MG & Van den Driesche, P (2002) Invasion theory and biological control. Ecology Letters 5: 148157.Google Scholar
Fagan, WF, Lewis, M, Neubert, MG, Aumann, C, Apple, JL & Bishop, JG (2005) When can herbivores slow or reverse the spread of an invading plant? A test case from Mount St. Helens. American Naturalist 166: 669685.Google Scholar
Fan, Y, Fang, W, Guo, S, Pei, X, Zhang, Y, Xiao, Y, Li, D, Jin, K, Bidochka, MJ & Pei, Y (2007) Increased insect virulence in Beauveria bassiana strains overexpressing an engineered chitinase. Applied and Environmental Microbiology 73: 295302.Google Scholar
Farrar, RR, Shapiro, M & Javaid, I (2003) Photostabilized titanium dioxide and a fluorescent brightener as adjuvants for a nucleopolyhedrovirus. BioControl 48: 543560.Google Scholar
Fauvergue, X & Hopper, KR (2009) French wasps in the New World: experimental biological control introductions reveal a demographic Allee effect. Population Ecology 51: 385397.Google Scholar
Fauvergue, X, Malausa, J-C, Giuge, L & Courchamp, F (2007) Invading parasitoids suffer no Allee effect: a manipulative field experiment. Ecology 88: 23922403.Google Scholar
Fauvergue, X, Vercken, E, Malausa, T & Hufbauer, RA (2012) The biology of small, introduced populations, with special reference to biological control. Evolutionary Applications 5: 424443.Google Scholar
Federici, BA (2005) Insecticidal bacteria: an overwhelming success for invertebrate pathology. Journal of Invertebrate Pathology 89:3038.Google Scholar
Federici, BA & Maddox, JV (1996) Host specificity in microbe-insect interactions. BioScience 46: 410421.Google Scholar
Felker-Quinn, E, Schweitzer, JA & Bailey, JK (2013) Meta-analysis reveals evolution in invasive plant species but little support for evolution of increased competitive ability (EICA). Ecology and Evolution 3: 739751.Google Scholar
Fellowes, MDE & Godfray, HCJ (2000) The evolutionary ecology of resistance to parasitoids in Drosophila. Heredity 84: 18.Google Scholar
Fenner, F (1953) Changes in the mortality rate due to myxomatosis in the Australian wild rabbit. Nature 172: 228230.Google Scholar
Fenner, F & Ratcliffe, FN (1965) Myxomatosis. Cambridge, UK, Cambridge University Press.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
Fernanda Diaz, M, Ramirez, A & Poveda, K (2012) Efficiency of different egg parasitoids and increased floral diversity for the biological control of noctuid pests. Biological Control 60: 182191.Google Scholar
Fernandes, EKK, Rangel, DEN, Braga, GUL & Roberts, DW (2015) Tolerance of entomopathogenic fungi to ultraviolet radiation: a review on screening of strains and their formulation. Current Genetics 61: 427440.Google Scholar
Ferrari, J, Darby, AC, Daniel, TJ, Godfray, HCJ & Douglas, AE (2004) Linking the bacterial community in pea aphids with host-plant use and natural enemy resistance. Ecological Entomology 29: 6065.Google Scholar
Ferrari, J, Müller, CB, Kraaijeveld, AR & Godfray, HCJ (2001) Clonal variation and covariation in aphid resistance to parasitoids and a pathogen. Evolution 55: 18051814.Google Scholar
Ferrero-Serrano, A, Collier, TR, Hild, AL, Mealor, BA & Smith, T (2008) Combined impacts of native grass competition and introduced weevil herbivory on Canada thistle (Cirsium arvense). Rangeland Ecology & Management 61: 529534.Google Scholar
Fine, PVA (2002) The invasibility of tropical forests by exotic plants. Journal of Tropical Ecology 18: 687705.Google Scholar
Finlayson, CJ, Landry, KN & Alyokhin, AV (2008) Abundance of native and non-native lady beetles (Coleoptera: Coccinellidae) in different habitats in Maine. Annals of the Entomological Society of America 101: 10781087.Google Scholar
Fisher, AJ, Smith, L & Woods, DM (2011) Climatic analysis to determine where to collect and release Puccinia jaceae var. solstitialis for biological control of yellow starthistle. Biocontrol Science and Technology 21: 333351.Google Scholar
Flanagan, GJ, Hills, A & Wilson, CG (2000) The successful biological control of spineyhead sida, Sida acuta (Malvaceae), by Calligrapha pantherina (Col.: Chrysomelidae) in Australia’s Northern Territory. In Proceedings of the 10th International Symposium on Biological Control of Weeds (Spencer, NR, ed.), Bozeman, Montana State University, pp. 3541.Google Scholar
Flanders, SE (1930) Mass production of egg parasites of the genus Trichogramma. Hilgardia 4: 465501.Google Scholar
Flanders, SE (1951) Mass culture of California red scale and its golden chalcid parasites. Hilgardia 21: 142.Google Scholar
Fleischer, SJ, Blom, PE & Weisz, R (1999) Sampling in precision IPM: When the objective is a map. Phytopathology 89: 11121118.Google Scholar
Fletcher, JP, Hughes, JP & Harvey, I (1994) Life expectancy and egg load affects oviposition decisions of a solitary parasitoid. Proceedings of the Royal Society of London, B 258: 63167.Google Scholar
Flexner, JL, Lighthart, B & Croft, BA (1986) The effects of microbial pesticides on non-target beneficial arthropods. Agriculture, Ecosystems and Environment 16: 203254.Google Scholar
Flinn, PW & Hagstrum, D (1995) Simulation model of Cephalonomyia waterstoni (Hymenoptera: Bethylidae) parasitizing the rusty grain beetle (Coleoptera: Cucujidae). Environmental Entomology 24: 16081615.Google Scholar
Flint, ML & Dreistadt, SH (1998) Natural Enemies Handbook. Berkeley, University of California Press.Google Scholar
Flower, CE, Long, LC, Knight, KS, Rebbeck, J & Brown, JS (2014) Native bark-foraging birds preferentially forage in infected ash (Fraxinus spp.) and prove effective predators of the invasive emerald ash borer (Agrilus planipennis Fairmaire). Forest Ecology and Management 313: 300306.Google Scholar
Follett, PA & Duan, JJ (2000) Nontarget Effects of Biological Control. Dordrecht, The Netherlands, Kluwer.Google Scholar
Follett, PA, Johnson, MT & Jones, VP (2000) Parasitoid drift in Hawaiian pentatomids. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 7794.Google Scholar
Forbes, VE, Calow, P, Grimm, V, Hayashi, TI, Jager, T, Katholm, A, Palmqvist, A, Pastorok, R, Salvito, D, Sibly, R, Spromberg, J, Stark, J, Stillman, RA (2011) Adding value to ecological risk assessment with population modeling. Human and Ecological Risk Assessment 17: 287299.Google Scholar
Force, DC (1967) Genetics in the colonization of natural enemies for biological control. Annals of the Entomological Society of America 60: 722728.Google Scholar
Fowler, SV, Syrett, P & Hill, RL (2000) Success and safety in the biological control of environmental weeds in New Zealand. Australian Ecology 25: 553562.Google Scholar
Frank, SD (2010) Biological control of arthropod pests using banker plant systems: past progress and future directions. Biological Control 52: 816.Google Scholar
Frank, SD & Shrewsbury, PM (2004) Effect of conservation strips on the abundance and distribution of natural enemies and predation of Agrotis ipsilon (Lepidoptera: Noctuidae) on golf course fairways. Environmental Entomology 33: 16621672.Google Scholar
Franks, SJ, Pratt, PD & Tsutsui, ND (2011) The genetic consequences of a demographic bottleneck in an introduced biological control insect. Conservation Genetics 12: 201211.Google Scholar
Fraser, SM & Lawton, JH (1994) Host range expansion by British moths onto introduced conifers. Ecological Entomology 19: 127137.Google Scholar
Frazer, BD & Van den Bosch, R (1973) Biological control of the walnut aphid in California: the interrelationship of the aphid and its parasite. Environmental Entomology 2: 561568.Google Scholar
Freestone, AL, Ruiz, GM & Torchin, ME (2013) Stronger biotic resistance in tropics relative to temperate zone: effects of predation on marine invasion dynamics. Ecology 94: 13701377.Google Scholar
Frere, I, Fabry, J & Hance, T (2007) Apparent competition or apparent mutualism? An analysis of the influence of rose bush strip management on aphid population in wheat field. Journal of Applied Entomology 131: 275283.Google Scholar
Fridley, JD, Stachowicz, JJ, Naeem, S, Sax, DF, Seabloom, EW, Smith, MD, Stohlgren, TJ, Tilman, D & Von Holle, B (2007) The invasion paradox: reconciling pattern and process in species invasions. Ecology 88: 317.Google Scholar
Fritz, RS (1983) Ant protection of a host plant’s defoliator – consequence of an ant-membracid mutualism. Ecology 64: 789797.Google Scholar
Fuester, R, Kenis, M, Swain, KS, Kingsley, PC, Lopez-Vaamonde, C & Herard, F (2001) Host range of Aphantorhaphopsis samarensis (Diptera: Tachinidae), a larval parasite of the gypsy moth (Lepidoptera: Lymantriidae). Environmental Entomology 30: 605611.Google Scholar
Fuester, R, Swan, KS, Kenis, M & Herard, F (2004) Determining the host range of Aphantorhaphopsis samarensis, a specialized tachinid introduced against gypsy moth. In Assessing Host Ranges of Parasitoids and Predators Used in Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, USDA Forest Service, pp. 177194.Google Scholar
Fuller, EB, Elderd, D & Dwyer, G (2012) Persistence in the environment and insect-baculovirus interactions: disease-density thresholds, epidemic burnout, and insect outbreaks. American Naturalist 179: E70E96.Google Scholar
Funasaki, GY, Lai, P-Y, Nakahara, LM, Beardsley, JW & Ota, AK (1988) A review of biological control introductions in Hawaii: 1890–1985. Proceedings of the Hawaiian Entomological Society 28: 105160.Google Scholar
Futuyma, DJ & Moreno, G (1988) The evolution of ecological specialization. Annual Review of Ecology and Systematics 19: 207233.Google Scholar
Fuxa, JR (1998) Environmental manipulation for microbial control of insects. In Conservation Biological Control (Barbosa, P, ed.), San Diego, CA, Academic Press, pp. 255268.Google Scholar
Gagic, V, Tscharnkte, T, Dormann, CF, Gruber, B, Wilstermann, A & Thies, C (2011) Food web structure and biocontrol in a four-trophic level system across a landscape complexity gradient. Proceedings of the Royal Society B 278: 29462953.Google Scholar
Gagne, WC & Howarth, FG (1985) Conservation status of endemic Hawaiian Lepidoptera. In Proceedings of 3rd Congress of European Lepidoptera, Cambridge, UK, Societas Europaea Lepidopterologica, pp. 7484.Google Scholar
Gagnon, AE, Heimpel, GE & Brodeur, J (2011) The ubiquity of intraguild predation among predatory arthropods. PLoS One 6: e28061.Google Scholar
Galat, DL & Robertson, B (1992) Response of endangered Poeciliopsis occidentalis Sonoriensis in the Rio-Yaqui Drainage, Arizona, to introduced Gambusia affinis. Environmental Biology of Fishes 33: 249264.Google Scholar
Gamradt, SC & Kats, LB (1996) Effect of introduced crayfish and mosquitofish on California newts. Conservation Biology 10: 11551162.Google Scholar
Gan-Mor, S & Matthews, GA (2003) Recent developments in sprayers for application of biopesticides – an overview. Biosystems Engineering 84: 119125.Google Scholar
Gao, L, Sun, MH, Liu, XZ & Che, YS (2007) Effects of carbon concentration and carbon to nitrogen ratio on the growth and sporulation of several biocontrol fungi. Mycological Research 111: 8792.Google Scholar
Garcia, R, Caltagirone, LE & Gutierrez, AP (1988) Comments on a redefinition of biological control. BioScience 38: 692694.Google Scholar
Garcia, R & Legner, EF (1999) Biological control of medical and veterinary pests. In Handbook of Biological Control: Principles and Applications of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 935954.Google Scholar
Garcia-Rossi, D, Rank, N & Strong, DR (2003). Potential for self-defeating biological control? Variation in herbivore vulnerability among invasive Spartina genotypes. Ecological Applications, 13: 16401649.Google Scholar
Gardiner, MM, Landis, DA, Gratton, C, Schmidt, N, O’Neal, M, Mueller, E, Chacon, J, Heimpel, GE & DiFonzo, CD (2009a) Landscape composition influences patterns of native and exotic lady beetle abundance. Diversity and Distributions 15: 554564.Google Scholar
Gardiner, MM, Landis, DA, Gratton, C, DiFonzo, CD, O’Neal, M, Chacon, JM, Wayo, MT, Schmidt, NP, Mueller, EE & Heimpel, GE (2009b) Landscape diversity enhances biological control of an introduced crop pest in the north-central USA. Ecological Applications 19: 143154.Google Scholar
Gardner, DE, Smith, CW & Markin, GP (1995) Biological control of alien plants in natural areas of Hawaii. In Proceedings of the 8th International Symposium on Biological Control of Weeds (Delfosse, ES & Scott, RR, eds.), Canberra, Australia, CSIRO, pp. 27.Google Scholar
Gardner, J & Giles, K (1997) Mechanical distribution of Chrysoperla rufilabris and Trichogramma pretiosum: survival and uniformity of discharge after spray dispersal in an aqueous suspension. Biological Control 8: 138142.Google Scholar
Garrett, SD (1965) Toward biological control of soil-borne pathogens. In Ecology of Soil-Borne Pathogens (Baker, KF & Snyder, WC, eds.), Berkeley, University of California Press, pp. 416.Google Scholar
Gaskin, JF, Bon, MC, Cock, JW, Cristofaro, M, De Biase, A, De Clerck-Floate, R, Ellison, CA, Hinz, HL, Hufbauer, RA, Julien, MH & Sforza, R (2011) Applying molecular-based approaches to classical biological control of weeds. Biological Control 58: 121.Google Scholar
Gassmann, A & Louda, SM (2001) Rhynocillus conicus: Initial evaluation and subsequent ecological impacts in North America. In Evaluating Indirect Ecological Effects of Biological Control (Wajnberg, E, Scott, JK & Quimby, PC, eds.), Wallingford, UK, CABI Publishing, pp. 147184.Google Scholar
Gaugler, R (1993) Ecological genetics of entomopathogenic nematodes. In Nematodes and the Biological Control of Insect Pests (Bedding, R, Akhurst, R & Kaya, H, eds.), East Melbourne, Australia, CSIRO, pp. 8995.Google Scholar
Gaugler, R, Lewis, E & Stuart, RJ (1997a) Ecology in the service of biological control: the case of entomopathogenic nematodes. Oecologia 109: 483489.Google Scholar
Gaugler, R, Wilson, M & Shearer, P (1997b) Field release and environmental fate of a transgenic entomopathogenic nematode. Biological Control 9: 7580.Google Scholar
Geiger, F, Wäckers, FL & Bianchi, FJJA (2009) Hibernation of predatory arthropods in semi-natural habitats. BioControl 54: 529535.Google Scholar
Gelernter, WD & Lomer, CJ (2000) Success in biological control of above-ground insects by pathogens. In Biological Control: Measures of Success (Gurr, G & Wratten, S, eds.), Dordrecht, The Netherlands, Kluwer, pp. 297322.Google Scholar
Geneau, CE, Wäckers, FL, Luka, H, Daniel, C & Balmer, O (2012) Selective flowers to enhance biological control of cabbage pests by parasitoids. Basic and Applied Ecology 13: 8593.Google Scholar
Gentz, MC, Murdoch, G & King, GF (2010) Tandem use of selective insecticides and natural enemies for effective, reduced-risk pest management. Biological Control 52: 208215.Google Scholar
Gerlach, J (2001) Predator, prey and pathogen interactions in introduced snail populations. Animal Conservation 4: 203209.Google Scholar
Getz, WM (1984) Population dynamics: a resource per capita approach. Journal of Theoretical Biology 108: 623644.Google Scholar
Getz, WM (1996) A hypothesis regarding the abruptness of density dependence and the growth rate of populations. Ecology 77: 20142026.Google Scholar
Getz, WM & Mills, NJ (1996) Host-parasitoid coexistence and egg-limited encounter rates. The American Naturalist 148: 333347.Google Scholar
Ghorbani, R, Leifert, C & Seel, W (2005) Biological control of weeds with antagonistic plant pathogens. Advances in Agronomy 86: 191225.Google Scholar
Giles, DK & Wunderlich, LR (1998) Electronically-controlled delivery system for beneficial insect eggs in liquid suspensions. Transactions of the American Society of Agricultural Engineers 41: 839847.Google Scholar
Giles, DK, Gardner, J & Studer, HE (1995) Mechanical release of predacious mites for biological pest control in strawberries. Transactions of the American Society of Agricultural Engineers 38: 12891296.Google Scholar
Giles, KL, Jones, DB, Royer, TA, Elliot, NC & Kindler, SD (2003) Development of a sampling plant in winter wheat that estimates cereal aphid parasitism levels and predicts population suppression. Journal of Economic Entomology 96: 975982.Google Scholar
Gillespie, DR & Raworth, DA (2004) Biological control of two-spotted spider mites on greenhouse vegetable crops. In Biocontrol in Protected Culture (Heinz, KM, Van Driesche, RG & Parrella, MP, eds.), Batavia, IL, Ball, pp 201220Google Scholar
Gilligan, CA (2002) An epidemiological framework for disease management. Advances in Botanical Research 38: 164.Google Scholar
Giorgini, M, Monti, MM, Caprio, E, Stouthamer, R & Hunter, MS (2009) Feminization and the collapse of haplodiploidy in an asexual parasitoid wasp harboring the bacterial symbiont Cardinium. Heredity 102: 365371.Google Scholar
Glare, TJ, Caradus, W, Gelernter, T, 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
Glaser, RW & Farrell, CC (1935) Field experiments with the Japanese beetle and its nematode parasite. Journal of the New York Entomological Society 43: 345371.Google Scholar
Goddard, JH, Torchin, ME, Kuris, AM & Lafferty, KD (2005) Host specificity of Sacculina carcini, a potential biological control agent of introduced European green crab Carcinus maenas in California. Biological Invasions 7: 895912.Google Scholar
Godfray, HCJ (1994) Parasitoids: Behavioral and Evolutionary Ecology. Princeton, NJ, Princeton University Press.Google Scholar
Godfray, HCJ (1995) Field experiments with genetically manipulated insect viruses – ecological issues. Trends in Ecology & Evolution 10: 465469.Google Scholar
Godfray, HCJ, Briggs, CJ, Barlow, ND, O’Callaghan, MO, Glare, TR & Jackson, TA (1999) A model of insect-pathogen dynamics in which a pathogenic bacterium can also reproduce saprophytically. Proceedings of the Royal Society of London B 266: 233240.Google Scholar
Godfray, HCJ & Chan, MS (1990) How insecticides trigger single-stage outbreaks in tropical pests. Functional Ecology 4: 329337.Google Scholar
Godfray, HCJ & Hassell, MP (1987) Natural enemies may be a cause of discrete generations in tropical insects. Nature 327: 144147.Google Scholar
Godfray, HCJ & Hassell, MP (1989) Discrete and continuous insect populations in tropical environments. Journal of Animal Ecology 58: 153174.Google Scholar
Godfray, HCJ, Hassell, MP & Holt, RD (1994) The population dynamic consequences of phenological asynchrony between parasitoids and their hosts. Journal of Animal Ecology 63: 110.Google Scholar
Godfray, HCJ & Waage, JK (1991) Predictive modelling in biological control: the mango mealy bug (Rastrococcus invadens) and its parasitoids. Journal of Applied Ecology 28: 434453.Google Scholar
Godsoe, W, Murray, R & Plank, MJ (2015) Information on biotic interactions improves transferability of distribution models. American Naturalist 185: 281290.Google Scholar
Goeden, RD (1978) Biological control of weeds. In Introduced Parasites and Predators of Arthropod Pests and Weeds; USDA Handbook No. 480 (Clausen, CP, ed.), Washington, DC, USDA Agricultural Research Service, pp. 357413.Google Scholar
Goeden, RD & Kok, LT (1986) Comments on a proposed “new” approach for selecting agents for the biological control of weeds. Canadian Entomologist 118: 5158.Google Scholar
Goeden, RD & Louda, SM (1976) Biotic interference with insects imported for weed control. Annual Review of Entomology 21: 325341.Google Scholar
Goettel, MS & Hajek, A (2001) Evaluation of non-target effects of pathogens used for management of arthopods. In Evaluating Indirect Ecological Effects of Biological Control (Wajnberg, E, Scott, JK & Quimby, PC, eds.), Wallingford, UK, CABI Publishing, pp. 8197.Google Scholar
Goettel, MS, Hajek, AE, Siegel, JP & Evans, HC (2001) Safety of fungal biocontrol agents. In Fungi as Biocontrol Agents (Butt, TM, Jackson, CW & Magan, N, eds.), Wallingford, UK, CABI Publishing, pp. 347375.Google Scholar
Goettel, MS, Poprawski, TJ, Vandenberg, JD, Li, Z & Roberts, DW (1990) Safety to nontarget invertebrates of fungal biocontrol agents. In Safety of Microbial Insecticides (Laird, M, Lacey, LA & Davidson, EW, eds.), Boca Raton, FL, CRC Press, pp. 210231.Google Scholar
Gold, CS, Altieri, MA & Bellotti, AC (1989) The effects of intercropping and mized varieties of predators and parasitoids of cassava whiteflies (Hemiptera: Aleyrodidae) in Colombia. Bulletin of Entomological Research 79: 115121.Google Scholar
Goldson, SL, McNeill, MR & Proffitt, JR (2003) Negative effects of strain hybridisation on the biocontrol agent Microctonus aethiopoides. New Zealand Plant Protection, 56, 138142.Google Scholar
Goldson, SL, Proffitt, JR & McNeil, MR (1990) Seasonal biology and ecology in New Zealand of Microtonus aethiopoides (Hymenoptera: Braconidae), a parasitoid of Sitona spp. (Coleoptera: Curculionidae), with special emphasis on atypical behaviour. Journal of Applied Ecology 27: 703722.Google Scholar
Goldson, SL, Wratten, SD, Ferguson, CM, Gerard, PJ, Barratt, BIP, Hardwick, S, McNeill, MR, Phillips, CB, Popay, AJ, Tylianakis, JM & Tomasetto, F (2014) If and when successful classical biological control fails. Biological Control 72: 7679.Google Scholar
Goldsworthy, CA & Bettoli, PW (2006) Growth, body condition, reproduction and survival of stocked Barrens topminnows, Fundulus julisia (Fundulidae). American Midland Naturalist 156: 331343.Google Scholar
Gols, R, Bukovinszky, T, Hemerik, L, Harvey, JA, van Lenteren, JC & Vet, LEM (2005) Reduced foraging efficiency of a parasitoid under habitat complexity: implications for population stability and species coexistence. Journal of Animal Ecology 74: 10591068.Google Scholar
Gonthier, DJ, Ennis, KK, Philpott, SM, Vandermeer, J & Perfecto, I (2013) Ants defend coffee from berry borer colonization. BioControl 58: 815820.Google Scholar
Gonzalez, D & Wilson, LT (1982) A food-web approach to economic thresholds: a sequence of prests/predaceous arthropods in California cotton. Entomophaga 3143.Google Scholar
Gonzalez, D, Hagen, KS, Stary, P, Bishop, GW, Davis, DW & Pike, KS (1995) Pea aphid and blue alfalfa aphid. In Biological Control in the Western United States (Nechols, JR, ed.), Oakland, University of California Press, pp. 129135.Google Scholar
Goodsell, JA & Kats, LB (1999) Effect of introduced mosquitofish on pacific treefrogs and the role of alternative prey. Conservation Biology 13: 921924.Google Scholar
Goolsby, JA, De Barro, PJ, Kirk, AA, Sutherst, RW, Canas, L & Ciomperlik, MA (2005). Post-release evaluation of biological control of Bemisia tabaci biotype ‘B’ in the USA and the development of predictive tools to guide introductions for other countries. Biological Control 32: 7077.Google Scholar
Goolsby, JA, De Barro, PJ, Makinson, JR, Pemberton, RW, Hartley, DM & Frohlich, DR (2006a) Matching the origin of an invasive weed for selection of a herbivore haplotype for a biological control programme. Molecular Ecology 15: 287297.Google Scholar
Goolsby, JA, Van Klinken, RD & Palmer, WA (2006b) Maximising the contribution of native-range studies towards the identification and prioritisation of weed biocontrol agents. Australian Journal of Entomology 45: 276286.Google Scholar
Gordon, CE, McGill, B, Ibarra-Nunez, G, Greenberg, R & Perfecto, I (2009) Simplification of a coffee foliage-dwelling beetle community under low-shade management. Basic and Applied Ecology 10: 246254.Google Scholar
Gould, F (1994) Potential and problems with high-dose strategies for pesticidal engineered crops. Biocontrol Science and Technology 4: 451461.Google Scholar
Gould, F & Schliekelman, P (2004) Population genetics of autocidal control and strain replacement. Annual Review of Entomology 49: 193217.Google Scholar
Gould, JR, Bellows, TS & Paine, TD (1992) Population dynamics of Siphoninus phillyreae in California in the presence and absence of a parasitoid Encarsia partenopea. Ecological Entomology 17: 127134.Google Scholar
Gould, JR, Elkinton, JS & Wallner, WE (1990) Density-dependent suppression of experimentally created gypsy moth, Lymantria dispar (Lepidoptera, Lymantriidae), populations by natural enemies. Journal of Animal Ecology 59: 213233.Google Scholar
Graham, GL, Peng, G, Bailey, KL & Holm, FA (2006) Effect of dew temperature, post-inoculation condition, and pathogen does on suppression of scentless chamomile by Colletrichum truncatum. Biocontrol Science and Technology 16: 271280.Google Scholar
Granett, J, Dunbar, DM & Weseloh, RM (1976) Gypsy moth control with Dimilin sprays timed to minimize effects on the parasite Apanteles melanoscelus. Journal of Economic Entomology 69: 403404.Google Scholar
Grasman, J, Van Herwaarden, OA, Hemerik, L & Van Lenteren, JC (2001) A two-component model of host-parasitoid interactions: determination of the size of inundative releases of parasitoids in biological pest control. Mathematical Biosciences 169: 207216.Google Scholar
Graur, D (1985) Gene diversity in Hymenoptera. Evolution 39: 190199.Google Scholar
Gray, RH, Lorimer, CG, Tobin, PC & Raffa, KF (2008) Preoutbreak dynamics of a recently established invasive herbivore: roles of natural enemies and habitat structure in stage-specific performance of gypsy moth (Lepidoptera: Lymantriidae) populations in northeastern Wisconsin. Environmental Entomology 37: 11741184.Google Scholar
Greathead, D (1971) A Review of Biological Control in the Ethiopian Region. Commonwealth Institute of Biological Control, Technical Communication No. 5. Slough, UK, Commonwealth Institute of Biological Control.Google Scholar
Greathead, DJ (1986) Parasitoids in classical biological control. In Insect Parasitoids (Waage, J & Greathead, DJ, eds.), London, UK, Academic Press, pp. 290318.Google Scholar
Greathead, DJ & Greathead, AH (1992) Biological control of insects pests by insect parasitoids and predators: the BIOCAT database. Biocontrol News and Information 13: 61 N67 N.Google Scholar
Greaves, MP, Pring, RJ & Lawrie, J (2001) A proposed mode of action of oil-based formulations of a microbial herbicide. Biocontrol Science and Technology 11: 273281.Google Scholar
Greenberg, R, Bichier, P, Angon, AC, MacVean, C, Perez, R & Cano, E (2000) The impact of avian insectivory on arthropods and leaf damage in some Guatemalan coffee plantations. Ecology 81: 17501755.Google Scholar
Greenslade, PJM (1971) Interspecific competition and frequency changes among ants in Solomon Islands coconut plantations. Journal of Applied Ecology 8: 323352.Google Scholar
Gressel, J (2001) Potential failsafe mechanisms against the spread and introgression of transgenic hypervirulent biocontrol fungi. Trends in Biotechnology 19: 149154.Google Scholar
Gressel, J (2002) Molecular Biology of Weed Control. London, UK, Taylor & Francis.Google Scholar
Grevstad, FS (1999a) Experimental invasions using biological control introductions: the influence of release size on the chance of population establishment. Biological Invasions 1: 313323.Google Scholar
Grevstad, FS (1999b) Factors influencing the chance of population establishment: implications for release strategies in biocontrol. Ecological Applications 9: 14391447.Google Scholar
Grewal, PS, Bornstein-Forst, S, Burnell, AM, Glazer, I and Jagdale, GB (2006) Physiological, genetic, and molecular mechanisms of chemoreception, thermobiosis, and anhydrobiosis in entomopathogenic nematodes. Biological Control 38: 5465.Google Scholar
Grewal, PS & Peters, A (2005) Formulation and quality. In Nematodes as Biocontrol Agents (Grewal, PS, Ehlers, R-U & Shapiro-Ilan, DI, eds.), Wallingford, UK, CABI Publishing, pp. 7990.Google Scholar
Grewal, PS, Selvan, S & Gaugler, R (1994) Thermal adaptation of entomopathogenic nematodes – niche breadth for infection, establishment and reproduction. Journal of Thermal Biology 19: 245253.Google Scholar
Griffiths, GJK, Holland, JM, Bailey, A & Thomas, MB (2008) Efficacy and economics of shelter habitats for conservation biological control. Biological Control 45: 200209.Google Scholar
Griffiths, O, Cook, A & Wells, SM (1993) The diet of the introduced carnivorous snail Euglandina rosea in Mauritius and its implications for threatened island gastropod faunas. Journal of Zoology, London 229: 7989.Google Scholar
Groenteman, R, Kelly, D, Fowler, SV & Bourdot, GW (2011) Abundance, phenology and impact of biocontrol agents on nodding thistle (Carduus nutans) in Canterbury 35 years into a biocontrol programme. New Zealand Journal of Agricultural Research 54: 113.Google Scholar
Gruner, DS (2005) Biotic resistance to an invasive spider conferred by generalist insectivorous birds on Hawai’i Island. Biological Invasions 7: 547552.Google Scholar
Guetsky, R, Shtienberg, D, Elad, Y, Fischer, E & Dinoor, A (2002) Improving biological control by combining biocontrol agents each with several mechanisms of disease suppression. Phytopathology 92: 976985.Google Scholar
Guo, QF, Sax, DF, Qian, H & Early, R (2012) Latitudinal shifts of introduced species: possible causes and implications. Biological Invasions 14: 547556.Google Scholar
Guretzky, JA & Louda, SM (1997) Evidence for natural biological control: insects decrease survival and growth of a native thistle. Ecological Applications 7: 13301340.Google Scholar
Gurney, WSC, Nisbet, RM & Lawton, JH (1983) The systematic formulation of tractable single species population models incorporating age structure. Journal of Animal Ecology 52: 479496.Google Scholar
Gurr, GM & Wratten, S (1999) Integrated biological control: a proposal for enhancing success in biological control. International Journal of Pest Management 45: 8184.Google Scholar
Gurr, GM, Wratten, SD & Barbosa, P (2000) Success in conservation biological control of arthropods. In Biological Control: Measures of Success (Gurr, GM & Wratten, SD, eds.), Dordrecht, The Netherlands, Kluwer, pp. 105132.Google Scholar
Gurr, GM, Wratten, SD & Altieri, MA (2004) Ecological Engineering for Pest Management. Ithaca, NY, Comstock.Google Scholar
Gurr, GM, Wratten, SD & Luna, JM (2003) Multi-function agricultural biodiversity: pest management and other benefits. Basic and Applied Ecology 4: 107116.Google Scholar
Gurr, GM, Wratten, SD, Tylianakis, J, Kean, J & Keller, M (2005) Providing plant foods for natural enemies in farming systems: balancing practicalities and theory. In Plant-Provided Food for Carnivorous Insects: A Protective Mutualism and Its Applications (Wäckers, F, Van Rijn, PCJ & Bruin, J, eds.), Cambridge, UK, Cambridge University Press, pp. 326345.Google Scholar
Gutierrez, AP (1996) Applied Population Ecology: A Supply-Demand Approach. New York, Wiley.Google Scholar
Gutierrez, AP & Baumgärtner, JU (1984) Multitrophic models of predator–prey energetics: 1. Age specific energetics models – pea aphid Acyrthosiphon pisum (Harris) (Homoptera: Aphididae) as an example. Canadian Entomologist 116: 924932.Google Scholar
Gutierrez, AP, Caltagirone, LE & Meikle, W (1999) Evaluation of results, economics of biological control. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 243252.Google Scholar
Gutierrez, AP, Daane, KM, Ponti, L, Walton, VM & Ellis, CK (2008) Prospective evaluation of the biological control of vine mealybug: refuge effects and climate. Journal of Applied Ecology 45: 524536.Google Scholar
Gutierrez, AP, Mills, NJ, Schreiber, SJ & Ellis, CK (1994) A physiologically based tritrophic perspective on bottom-up-top-down regulation of populations. Ecology 75: 22272242.Google Scholar
Gutierrez, AP, Neuenschwander, P, Schulthess, F, Herren, HR, Baumgaertner, JU, Wermelinger, B, Lohr, B & Ellis, CK (1988a) Analysis of biological control of cassava pests in Africa. II. Cassava mealybug Phenacoccus manihoti. Journal of Applied Ecology 25: 921940.Google Scholar
Gutierrez, AP, Neuenschwander, P & Van Alphen, JJM (1993) Factors affecting biological control of cassava mealybug by exotic parasitoids: A ratio-dependent supply-demand driven model. Journal of Applied Ecology 30: 706721.Google Scholar
Gutierrez, AP, Pitcairn, MJ, Ellis, CK, Carruthers, N & Ghazelbash, R (2005) Evaluating biological control of yellow starthistle (Centaurea solstitialis) in California: a GIS based supply-demand demographic model. Biological Control 34: 115131.Google Scholar
Gutierrez, AP & Ponti, L (2013) Deconstructing the control of the spotted alfalfa aphid Therioaphis maculata. Agricultural and Forest Entomology 15: 272284.Google Scholar
Gutierrez, AP, Ponti, L, Hoddle, M, Almeida, RPP & Irvin, NA (2011) Geographic distribution and relative abundance of the invasive glassy-winged sharpshooter: effects of temperature and egg parasitoids. Environmental Entomology 40: 755769.Google Scholar
Gutierrez, AP, Villacorta, A, Cure, JR & Ellis, CK (1998a) Tritrophic analysis of the coffee (Coffee arabica): Coffee berry borer (Hypothenemus hampei (Ferrari)): parasitoid system. Anais da Sociedade Entomologica do Brasil 27: 357385.Google Scholar
Gutierrez, AP, Wemelinger, B, Schulthess, F, Baumgaertner, JU, Herren, HR, Ellis, CK & Yaninek, JS (1988b) Analysis of biological control of cassava pests in Africa. I. Simulation of carbon, nitrogen and water dynamics in cassava. Ecology 25: 901920.Google Scholar
Gutierrez, AP, Yaninek, JS, Neuenschwander, P & Ellis, CK (1999) A physiologically-based tritrophic metapopulation model of the African cassava food web. Ecological Modelling 123: 225242.Google Scholar
Gutierrez, AP, Yaninek, JS, Wermelinger, B, Herren, HR & Ellis, CK (1988c) Analysis of biological control of cassava pests in Africa. III. Cassava green mite Mononychellus tanajoa. Journal of Applied Ecology 25: 941950.Google Scholar
Gwynn, DM, Callaghan, A, Gorham, J, Walters, KFA & Fellowes, MDE (2005) Resistance is costly: trade-offs between immunity, fecundity and survival in the pea aphid. Proceedings of the Royal Society of London B 272: 18031808.Google Scholar
Haag, KH & Habeck, DH (1991) Enhanced biological control of water hyacinth following limited herbicide application. Journal of Aquatic Plant Management 29: 2428.Google Scholar
Haas, D & Défago, F (2005) Biological control of soil-borne pathogens by fluorescent pseudomonads. Nature Reviews Microbiology 3: 307319.Google Scholar
Haddad, NM, Crutsinger, GM, Gross, K, Haarstad, J, Knops, JMH & Tilman, D (2009) Plant species loss decreases arthropod diversity and shifts trophic structure. Ecology Letters 12: 10291039.Google Scholar
Haddad, NM, Crutsinger, GM, Gross, K, Haarstad, J & Tilman, D (2011) Plant diversity and the stability of food webs. Ecology Letters 14: 4246.Google Scholar
Haddad, NM, Haarstad, J & Tilman, D (2000) The effects of long-term nitrogen loading on grassland insect communities. Oecologia 124: 7384.Google Scholar
Haddad, NM, Tilman, D, Haarstad, J, Ritchie, M & Knops, JMH (2001) Contrasting effects of plant richness and composition on insect communities: a field experiment. American Naturalist 158: 1735.Google Scholar
Hadfield, MG & Mountain, BS (1980) A field study of a vanishing species, Achatinella mustelina (Gastropoda, Pulmonata), in the Waianae mountains of Oahu. Pacific Science 34: 345358.Google Scholar
Hadfield, MG, Miller, SE & Carwille, AH (1993) The decimation of endemic Hawai’ian tree snails by alien predators. American Zoologist 33: 610622.Google Scholar
Haenke, S, Scheid, B, Schaefer, M, Tscharntke, T & Thies, C (2009) Increasing syrphid fly diversity and density in sown flower strips within simple vs. complex landscapes. Journal of Applied Ecology 46: 11061114.Google Scholar
Hagen, KS, Mills, NJ, Gordh, G & McMurtry, JA (1999) Terrestrial arthropod predators of insect and mite pests. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 383504.Google Scholar
Hails, RS, Hernandez-Crespo, P, Sait, SM, Donnelly, CA, Green, BM & Cory, JS (2002) Transmission patterns of natural and recombinant baculoviruses. Ecology 83: 906916.Google Scholar
Hajek, AE (2004) Natural Enemies: An Introduction to Biological Control. Cambridge, UK, Cambridge University Press.Google Scholar
Hajek, AE & Butler, L (2000) Predicting the host range of entomopathogenic fungi. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 263276.Google Scholar
Hajek, AE, Butler, L & Wheeler, MM (1995) Laboratory bioassays testing the host range of the gypsy-moth fungal pathogen Entomophaga maimaiga. Biological Control 5: 530544.Google Scholar
Hajek, AE, Butler, L, Walsh, SRA, Silver, JC, Hain, FP, Hastings, FL, ODell, TM & Smitley, DR (1996) Host range of the gypsy moth (Lepidoptera: Lymantriidae) pathogen Entomophaga maimaiga (Zygomycetes: Entomophthorales) in the field versus laboratory. Environmental Entomology 25: 709721.Google Scholar
Hajek, AE, McManus, DP & Delalibera, I Jr. (2005) Catalogue of Introductions of Pathogens and Nematodes for Classical Biological Control of Insects and Mites. Washington, DC, USA, U.S. Forest Service.Google Scholar
Hajek, AE, McManus, DP & 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
Hall, RW & Ehler, LE (1979) Rate of establishment of natural enemies in classical biological control. Bulletin of the Entomological Society of America 25: 280282.Google Scholar
Hall, RW, Ehler, LE & Bisabri-Ershadi, B (1980) Rate of success in classical biological control of arthropods. Bulletin of the Entomological Society of America 26: 111114.Google Scholar
Hallett, RH, Bahlai, CA, Xue, YG & Schaafsma, AW (2014) Incorporating natural enemy units into a dynamic action threshold for the soybean aphid, Aphis glycines (Homoptera: Aphididae). Pest Management Science 70: 879888.Google Scholar
Hallett, SG (2005) Where are the bioherbicides? Weed Science 53: 404415.Google Scholar
Hamer, AJ, Lane, SJ & Mahony, MJ (2002) The role of introduced mosquitofish (Gambusia holbrooki) in excluding the native green and golden bell frog (Litoria aurea) from original habitats in south-eastern Australia. Oecologia 132: 445452.Google Scholar
Hamilton, WD (1967). Extraordinary sex ratios. Science 156: 477488.Google Scholar
Hammock, B (1992) Virus release evaluation. Nature 355: 119.Google Scholar
Hance, T (1988) The demographic parameters of Phytoseiulus persimilis Athias-Henriot (Acari: Phytoseiidae) in relation to the possibilities of using it to control populations of Tetranychus urticae (Acari: Tetranychidae). Annales de la Societe Royale Zoologique de Belgique 118: 161170.Google Scholar
Hance, T, Van Baaren, J, Vernon, P & Boivin, G (2007) Impact of extreme temperatures on parasitoids in a climate change perspective. Annual Review of Entomology 52: 107126.Google Scholar
Handley, LJL, Estoup, A, Evans, D, Thomas, CE, Lombaert, E, Facon, B, Aebi, A & Roy, HE (2011) Ecological genetics of invasive alien species. BioControl 56: 409428.Google Scholar
Hanlon, SG, Hoyer, MV, Cichra, CE & Canfield, DE Jr. (2000) Evaluation of macrophyte control in 38 Florida lakes using triploid grass carp. Journal of Aquatic Plant Management 39: 4854.Google Scholar
Hanna, R, Onzo, A, Lingeman, R, Yaninek, JS & Sabelis, MW (2005) Seasonal cycles and persistence in an acarine predator–prey system on cassava in Africa. Population Ecology 47: 107117.Google Scholar
Hanski, I (1999) Metapopulation Ecology. Oxford, UK, Oxford University Press.Google Scholar
Hanski, I & Simberloff, D (1997) The metapopulation approach, its history, conceptual domain, and application to conservation. In Metapopulation Biology: Ecology, Genetics and Evolution (Hanski, I & Gilpin, ME, eds.) San Diego, CA, Academic Press, pp. 526.Google Scholar
Hardin, MR, Benrey, B, Coll, M, Lamp, WO, Roderick, GK & Barbosa, P (1995) Arthropod pest resurgence: an overview of potential mechanisms. Crop Protection 14: 318.Google Scholar
Hardman, JM, Van der Werf, W & Blatt, SE (2013) Simulating effects of environmental factors on biological control of Tetranychus urticae by Typhlodromus pyri in apple orchards. Experimental and Applied Acaraology 60: 181203.Google Scholar
Hardy, CM, Hinds, LA, Kerr, PJ, Lloyd, M, Redwood, A, Shellam, G & Strive, T (2006) Biological control of vertebrate pests using virally vectored immunocontraception. Journal of Reproductive Immunology 71: 102111.Google Scholar
Hare, JD & Morgan, DJW (1997) Mass-priming Aphytis: behavioral improvement of insectary-reared biological control agents. Biological Control 10: 207214.Google Scholar
Hare, JD, Morgan, DJW & Nguyun, T (1997) Increased parasitization of California red scale in the field after exposing its parasitoid, Aphytis melinus, to a synthetic kairomone. Entomologia Experimentalis et Applicata 82: 7381.Google Scholar
Harley, KLS & Forno, IW (1992) Biological Control of Weeds: A Handbook for Practitioners and Students. Melbourne, Australia, Inkata Press.Google Scholar
Harman, GE (2006) Overview of mechanisms and uses of Trichoderma spp. Phytopathology 96: 190194.Google Scholar
Harman, GE, Howell, CR, Viterbo, A, Chet, I & Loritto, M (2004) Trichoderma species – opportunistic, avirulent plant symbionts. Nature Reviews Microbiology 2: 4356.Google Scholar
Harmon, JP & Andow, DA (2004) Indirect effects between shared prey: predictions for biological control. BioControl 49: 605626.Google Scholar
Harpur, BA, Subhani, M & Zayed, A (2012) A review of the consequences of complementary sex determination and diploid male production on mating failures in Hymenoptera. Entomologia Experimentalis et Applicata 146: 156164.Google Scholar
Harris, P (1981) Stress as a strategy in the biological control of weeds. In Biological Control and Crop Protection (Papavizas, GC, ed.), Totowa, Allanheld Osmun, pp. 333340.Google Scholar
Harris, P (1988) Environmental impact of weed-control insects. BioScience 38: 542548.Google Scholar
Harris, P (1991) Classical biocontrol of weeds: its definitions, selection of effective agents, and administrative-political problems. Canadian Entomologist 123: 827849.Google Scholar
Harris, P, Wilkinson, ATS, Neary, ME & Thompson, LS (1971) Senecio jacobaea L., tansy ragwort (Compositae). In Biological Control Programmes Against Insects and Weeds in Canada: 1959–1968 (Kelleher, JS & Hulme, MA, eds.), Slough, UK, Commonwealth Agricultural Bureau, pp. 97104.Google Scholar
Harris, P, Wilkinson, ATS, Neary, ME, Thompson, LS & Finnamore, D (1975) Establishment in Canada of the cinnabar moth, Tyria jacobaeae (Lepidoptera: Arctiidae). Canadian Entomologist 107: 913917.Google Scholar
Harris, P & Zwölfer, H (1968) Screening of phytophagous insects for biological control of weeds. Canadian Entomologist 100: 295303.Google Scholar
Harrison, S & Taylor, AD (1997) Empirical evidence for metapopulation dynamics. In Metapopulation Biology: Ecology, Genetics and Evolution (Hanski, I & Gilpin, ME, eds.), San Diego, CA, Academic Press, pp. 2744.Google Scholar
Harrison, RL & Bonning, BC (2000) Genetic engineering of biocontrol agents for insects. In Biological and Biotechnological Control of Insect Pests (Rechcigl, JE & Rechcigl, NA, eds.), Boca Raton, FL, CRC Press, pp. 243280.Google Scholar
Hart, AD (1978) The onslaught against Hawaii’s tree snails. Natural History 87: 4657.Google Scholar
Hart, AJ, Bale, JS, Tullett, AG, Worland, MR & Walters, KFA (2002a) Effects of temperature on the establishment potential of the predatory mite Amblyseius californicus McGregor (Acari: Phytoseiidae) in the UK. Journal of Insect Physiology 48: 593599.Google Scholar
Hart, AJ, Tullett, AG, Bale, JS & Walters, KFA (2002b) Effects of temperature on the establishment potential in the UK of the non-native glasshouse biocontrol agent Macrolophus caliginosus. Physiological Entomology 27: 112123.Google Scholar
Harvey, IC & Bourdot, GW (2001) Giant buttercup (Ranunculus acris L.) control in pasture using a mycoherbicide based on Sclerotinia sclerotiorum. New Zealand Plant Protection 54: 120124.Google Scholar
Harwood, RJW, Hickman, JM, MacLeod, A, Sherratt, T & Wratten, SD (1994) Managing field margin for hover flies. In Field Margins: Integrating Agriculture and Conservation (Boatman, N, ed.), Alton, UK, British Crop Protection Society, pp. 147152.Google Scholar
Hasan, S, Chaboudez, P & Espiau, C (1995) Isozyme patterns and susceptibility of North American forms of Chondrilla juncea to European strains of the rust fungus Puccinia chondrillina. In Proceedings of the Eighth International Symposium on Biological Control of Weeds (Delfosse, ES & Scott, RR, eds.), Canterbury, New Zealand, DSIR/CSIRO, pp. 367373.Google Scholar
Hasan, S & Delfosse, ES (1995) Susceptibility of the Australian native, Heliotropium crispatum, to the rust fungus Uromyces heliotropii introduced to control common heliotrope, Heliotropium europaeum. Biocontrol Science and Technology 5: 165174.Google Scholar
Hashmi, S, Hashmi, G, Glazer, I, Gaugler, R (1998) Thermal response of Heterorhabditis bacteriophora transformed with the Caenorhabditis elegans hsp70 encoding gene. Journal of Experimental Zoology 281: 164170.Google Scholar
Hassan, SA (1989) Testing methodology and the concept of the IOBC/WPRS working group. In Pesticides and Non-target Invertebrates (Jepson, PC, ed.), Andover, UK, Intercept, pp. 118.Google Scholar
Hassell, MP (1978) The Dynamics of Arthropod Predator–Prey Systems. Princeton, NJ, Princeton University Press.Google Scholar
Hassell, MP (1980) Foraging strategies, population models and biological control: a case study. Journal of Animal Ecology 49: 603628.Google Scholar
Hassell, MP (1984a) Insecticides in host parasitoid interactions. Theoretical Population Biology 26: 378386.Google Scholar
Hassell, MP (1984b) Parasitism in patchy environments: inverse density dependence can be stabilizing. IMA Journal of Mathematics Applied in Medicine and Biology 1: 123133.Google Scholar
Hassell, MP (2000) The Spatial and Temporal Dynamics of Host-Parasitoid Interactions. Oxford, UK, Oxford University Press.Google Scholar
Hassell, MP & Comins, HN (1978) Sigmoid functional responses and population stability. Theoretical Population Biology 14: 6267.Google Scholar
Hassell, MP, Lawton, JH & Beddington, JR (1976) The components of arthropod predation I. The prey death-rate. Journal of Animal Ecology 45: 135164.Google Scholar
Hassell, MP & May, RM (1973) Stability in insect host-parasite models. Journal of Animal Ecology 42: 693726.Google Scholar
Hassell, MP & Varley, GC (1969) New inductive population model for insect parasites and its bearing on biological control. Nature 223: 11331136.Google Scholar
Hastings, A & Higgins, K (1994) Persistence of transients in spatially structured ecological models. Science 263: 11331136.Google Scholar
Hatherly, IS, Bale, JS, Walters, KFA & Worland, MR (2004) Thermal biology of Typhlodromips montdorensis: implications for its introduction as a glasshouse biological control agent in the UK. Entomologia Experimentalis et Applicata 111: 97109.Google Scholar
Hatherly, IS, Hart, AJ, Tullett, AG & Bale, JS (2005) Use of thermal data as a screen for the establishment potential of non-native biological control agents in the UK. BioControl 50: 687698.Google Scholar
Hautier, L, Gregoire, JC, de Schauwers, J, San Martin, G, Callier, P, Jansen, J-P & de Biseau, J-C (2008) Intraguild predation by Harmonia axyridis on coccinellids revealed by exogenous alkaloid sequestration. Chemoecology 18: 191196.Google Scholar
Havill, NP, Davis, G, Mausel, DL, 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
Hawkes, RB (1973) Natural mortality of cinnabar moth in California. Annals of the Entomological Society of America 66: 137146.Google Scholar
Hawkins, BA & Cornell, HV (1994) Maximum parasitism rates and successful biological control. Science 266: 1886.Google Scholar
Hawkins, BA, Cornell, HV & Hochberg, ME (1997) Predators, parasitoids, and pathogens as mortality agents in phytophagous insect populations. Ecology 78: 21452152.Google Scholar
Hawkins, BA & Marino, PC (1997) The colonization of native phytophagous insects in North America by exotic parasitoids. Oecologia 112: 566571.Google Scholar
Hawkins, BA, Mills, NJ, Jervis, MA & Price, PW (1999) Is the biological control of insects a natural phenomenon? Oikos 86: 493506.Google Scholar
Hawkins, BA, Thomas, MB & Hochberg, ME (1993) Refuge theory and biological control. Science 262: 14291432.Google Scholar
Haye, T, Goulet, H, Mason, PG & Kuhlmann, U (2005) Does fundamental host range match ecological host range? A retrospective case study of a Lygus plant bug parasitoid. Biological Control 35: 5567.Google Scholar
Haye, T, Kuhlmann, U, Goulet, H & Mason, PG (2006) Controlling Lygus plant bugs (Heteroptera: Miridae) with European Peristenus relictus (Hymenoptera: Braconidae) in Canada – risky or not? Bulletin of Entomological Research 96: 187196.Google Scholar
Hayes, KR & Barry, SC (2008) Are there any consistent predictors of invasion success? Biological Invasions 10: 483506.Google Scholar
Hayes, L, Fowler, SV, Paynter, Q, Groenteman, R, Peterson, P, Dodd, S & Bellgard, S (2013) Biocontrol of weeds: achievements to date and future outlook. In Ecosystem Services in New Zealand: Conditions and Trends (Dymond, JR, ed.), Lincoln, New Zealand, Manaaki Whenua, pp. 375385.Google Scholar
Hector, A, Dobson, K, Minns, A, Bazely-White, E & Lawton, JH (2001) Community diversity and invasion resistance: an experimental test of a grassland ecosystem and a review of comparable studies. Ecological Research 16: 819831.Google Scholar
Heimpel, GE (2000) Effects of clutch size on host-parasitoid population dynamics. In Parasitoid Population Biology (Hochberg, ME & Ives, AR, eds.), Princeton, NJ, Princeton University Press, pp. 2740.Google Scholar
Heimpel, GE, Antolin, MF, Franqui, RA & Strand, MR (1997) Reproductive isolation and genetic variation between two ‘strains’ of Bracon hebetor (Hymenoptera: Braconidae). Biological Control 9: 149156.Google Scholar
Heimpel, GE, Antolin, MR & Strand, MR (1999). Diversity of sex-determining alleles in Bracon hebetor. Heredity 82: 282291.Google Scholar
Heimpel, GE & Asplen, MK (2011) A ‘goldilocks’ hypothesis for dispersal of biological control agents. BioControl 56: 441450.Google Scholar
Heimpel, GE & Casas, J (2008) Parasitoid foraging and oviposition behaviour in the field. In Behavioural Ecology of Insect Parasitoids (Wajnberg, E, Bernstein, C & Van Alphen, JJM, eds.), Oxford, UK, Blackwell, pp. 5170.Google Scholar
Heimpel, GE & Collier, TR (1996) The evolution of host-feeding behavior in insect parasitoids. Biological Reviews 71: 373400.Google Scholar
Heimpel, GE & de Boer, JG (2008) Sex determination in the Hymenoptera. Annual Review of Entomology 53: 209230.Google Scholar
Heimpel, GE, Frelich, LE, Landis, DA, Hopper, KR, Hoelmer, KA, Sezen, Z, Asplen, MK & Wu, K (2010) European buckthorn and Asian soybean aphid as components of an extensive invasional meltdown in North America. Biological Invasions 12: 29132931.Google Scholar
Heimpel, GE & Jervis, MA (2005) Does floral nectar improve biological control by parasitoids? In Plant-Provided Food and Plant-Carnivore Mutualism (Wäckers, F, Van Rijn, P & Bruin, J, eds.), Cambridge, UK, Cambridge University Press, pp. 267304.Google Scholar
Heimpel, GE, Lee, JC, Wu, Z, Weiser, L, Wackers, F & Jervis, MA (2004b) Gut sugar analysis in field-caught parasitoids: adapting methods originally developed for biting flies. International Journal of Pest Management 50: 193198.Google Scholar
Heimpel, GE & Lundgren, JG (2000) Sex ratios of commercially reared biological control agents. Biological Control 19: 7793.Google Scholar
Heimpel, GE, Mangel, M & Rosenheim, JA (1998) Effects of egg and time limitation on lifetime reproductive success of a parasitoid in the field. American Naturalist 152: 273289.Google Scholar
Heimpel, GE, Neuhauser, C & Andow, DA (2005) Natural enemies and the evolution of resistance to transgenic insecticidal crops by pest insects: the role of egg mortality. Environmental Entomology 34: 512526.Google Scholar
Heimpel, GE, Neuhauser, C & Hoogendoorn, M (2003) Effects of parasitoid fecundity and host resistance on indirect interactions among hosts sharing a parasitoid. Ecology Letters 6: 556566.Google Scholar
Heimpel, GE, Ragsdale, DW, Venette, R, Hopper, KR, O’Neil, RJ, Rutledge, CE & Wu, Z (2004a) Prospects for importation biological control of the soybean aphid: anticipating potential costs and benefits. Annals of the Entomological Society of America 97: 249258.Google Scholar
Heimpel, GE & Rosenheim, JA (1998) Egg limitation in parasitoids: a review of the evidence and a case study. Biological Control 11: 160168.Google Scholar
Heimpel, GE, Rosenheim, JA & Mangel, M (1996) Egg limitation, host quality, and dynamic behavior by a parasitoid in the field. Ecology 77: 24102420.Google Scholar
Heimpel, GE, Yang, Y, Hill, JD & Ragsdale, DW (2013) Environmental consequences of invasive species: greenhouse gas emissions of insecticide use and the role of biological control in reducing emissions. PLoS ONE 8: e72293.Google Scholar
Hein, S, Poethke, H-J & Dorn, S (2009) What stops the ‘diploid male vortex’? A simulation study for species with single locus complementary sex determination. Ecological Modelling 220: 16631669.Google Scholar
Heinz, KM, Nunney, L and Parrella, MP (1993) Toward predictable biological control of Liriomyza trifolii (Diptera: Agromyzidae) infesting greenhouse cut chrysanthemums. Environmental Entomology 22: 12171233.Google Scholar
Heinz, KM, Van Driesche, R & Parrella, MP (2005) Biocontrol in Protected Culture. Batavia, IL, Ball.Google Scholar
Hennemann, ML & Memmott, J (2001) Infiltration of a Hawaiian community by introduced biological control agents. Science 293: 13141316.Google Scholar
Henning, J, Meers, J, Davies, PR & Morris, RS (2005) Survival of rabbit haemorrhagic disease virus (RHDV) in the environment. Epidemiology and Infection 113: 719730.Google Scholar
Henry, LM, May, N, Acheampong, S, Gillespie, DR & Roitberg, BD (2010) Host-adapted parasitoids in biological control: does source matter? Ecological Applications 20: 242250.Google Scholar
Henry, LM, Roitberg, BD & Gillespie, DR (2008) Host-range evolution in Aphidius parasitoids: fidelity, virulence and fitness trade-offs on an ancestral host. Evolution 62: 689699.Google Scholar
Henter, HJ (1995). The potential for coevolution in a host-parasitoid system. II. Genetic variation within a population of wasps in the ability to parasitize an aphid host. Evolution 49: 439445.Google Scholar
Henter, HJ (2003). Inbreeding depression and haplodiploidy: experimental measures in a parasitoid and comparisons across diploid and haplodiploid insect taxa. Evolution 57: 17931803.Google Scholar
Henter, HJ & Via, S (1995) The potential for coevolution in a host-parasitoid system. I. Genetic variation within an aphid population in susceptibility to a parasitic wasp. Evolution 49: 427438.Google Scholar
Hernandez-Rodriguez, A, Heydrich-Perez, M, Acebo-Guerrero, Y, Velazquez-del Valle, MG & Hernandez-Lauzardo, AN (2008) Antagonistic activity of Cuban native rhizobacteria against Fusarium verticillioides (Sacc.) Nirenb. in maize (Zea mays L.). Applied Soil Ecology 39: 180186.Google Scholar
Herzog, DC & Funderburk, JE (1985) Ecological bases for habitat management and pest cultural control. In Ecological Theory and Integrated Pest Management Practice (Kogan, M, ed.), New York, Wiley, pp. 217250.Google Scholar
Hewa-Kapuge, S, McDougall, S & Hoffmann, AA (2003) Effects of methoxyfenozide, indoxacarb, and other insecticides on the beneficial egg parasitoid Trichogramma nr. brassicae (Hymenoptera: Trichogrammatidae) under laboratory and field conditions. Journal of Economic Entomology 96: 10831090.Google Scholar
Hickling, GJ (2000) Success in biological control of vertebrate pests. In Biological Control: Measures of Success (Gurr, GM & Wratten, SD, eds.), Dordrecht, The Netherlands, Kluwer, pp. 341368.Google Scholar
Hickman, JM & Wratten, SD (1996) Use of Phacelia tanacetifolia strips to enhance biological control of aphids by hoverfly larvae in cereal fields. Journal of Economic Entomology 89: 832840.Google Scholar
Hight, SD, Carpenter, JE, Bloem, S & Bloem, KA (2005) Developing a sterile insect release program for Cactoblastis cactorum (Berg) (Lepidoptera: Pyralidae): effective overflooding ratios and release-recapture field studies. Environmental Entomology 34: 850856.Google Scholar
Hilbeck, A, Eckel, C & Kennedy, GC (1998) Impact of Bacillus thuringiensis – insecticides on population dynamics and egg predation of the Colorado potato beetle in North Carolina potato plantings. BioControl 43: 6575.Google Scholar
Hilborn, R (1975) The effect of spatial heterogeneity on the persistence of predator–prey interactions. Theoretical Population Biology 8: 346355.Google Scholar
Hill, G & Greathead, D (2000) Economic evaluation in classical biological control. In The Economics of Biological Invasions (Perrings, C, Williamson, M & Dalmazzone, S, eds.), Cheltenham, UK, Edward Elgar, pp. 208223.Google Scholar
Hill, MP & Hulley, PE (1995) Host-range expansion by native parasitoids to weed biocontrol agents introduced into South Africa. Biological Control 5: 297302.Google Scholar
Hochberg, ME & Hawkins, BA (1992) Refuges as a predictor of parasitoid diversity. Science 255: 973976.Google Scholar
Hochberg, ME & Hawkins, BA (1993) Predicting parasitoid species richness. American Naturalist 142: 671693.Google Scholar
Hochberg, ME & Hawkins, BA (1994) The implications of population dynamics theory to parasitoid diversity and biological control. In Parasitoid Community Ecology (Hawkins, BA & Sheehan, W, eds.), Oxford, UK, Oxford University Press, pp. 451471.Google Scholar
Hochberg, ME & Holt, RD (1995) Refuge evolution and the population dynamics of coupled host-parasitoid associations. Evolutionary Ecology 9: 633661.Google Scholar
Hochberg, ME & Waage, JK (1991) A model for the biological control of Oryctes rhinoceros (Coleoptera: Scarabaeidae) by means of pathogens. Journal of Applied Ecology 28: 514531.Google Scholar
Hoddle, MS (1999) Biological control of vertebrates. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press pp. 955974.Google Scholar
Hoddle, M (2004b) Analysis of fauna in the receiving area for the purpose of identifying native species that exotic natural enemies may potentially attack. In Assessing Host Ranges of Parasitoids and Predators Used for Classical Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, U.S. Forest Service, pp. 2439.Google Scholar
Hoddle, MS (2004a) Restoring balance: using exotic species to control invasive exotic species. Conservation Biology 18: 3849.Google Scholar
Hoddle, MS (2006) Historical review of control programs for Levuana iridescens (Lepidoptera: Zygaenidae) in Fiji and examination of possible extinction of this moth by Bessa remota (Diptera: Tachinidae). Pacific Science 60: 439453.Google Scholar
Hoddle, MS, Van Driesche, RG & Sanderson, JP (1998) Biology and use of the whitefly parasitoid Encarsia formosa. Annual Review of Entomology 43: 645659.Google Scholar
Hoelmer, KA & Kirk, AA (2005) Selecting arthropod biological control agents against arthropod pests: can the science be improved to decrease the risk of releasing ineffective agents? Biological Control 34: 255264.Google Scholar
Hoffmann, AA (2014) Facilitating Wolbachia invasions. Austral Entomology 53: 125132.Google Scholar
Hoffmann, AA, Montgomery, BL, Popovici, J, Iturbe-Ormaetxe, I, Johnson, PH, Muzzi, F, Greenfield, M, Durkan, M, Leong, YS, Dong, Y, Cook, H, Axford, J, Callahan, AG, Kenny, N, Omodei, C, McGraw, EA, Ryan, PA, Ritchie, SA, Turelli, M & O’Neill, SL (2011) Successful establishment of Wolbachia in Aedes populations to suppress dengue transmission. Nature 476: 454457.Google Scholar
Hoffmann, JH (1995) Biological control of weeds: the way forward, a South African perspective. Proceedings of the BCPC Symposium. Weeds in a Changing World 64: 7789.Google Scholar
Hoffmann, JH, Impson, FAC & Volchansky, CR (2002) Biological control of cactus weeds: implications of hybridization between control agent biotypes. Journal of Applied Ecology 39: 900908.Google Scholar
Hoffmann, JH & Moran, VC (1995) Localized failure of a weed biological control agent attributed to insecticide drift. Agriculture Ecosystems & Environment 52: 197203.Google Scholar
Hoffmann, JH & Moran, VC (1998) The population dynamics of an introduced tree, Sesbania punicea, in South Africa, in response to long-term damage caused by different combinations of three species of biological control agents. Oecologia 114: 343348.Google Scholar
Hoffmann, MP, Wilson, LT, Zalom, FG & Hilton, RJ (1990) Parasitism of Heliothis zea (Lepidoptera: Noctuidae) eggs: effect on pest managment decision rules for processing tomatoes in the Sacramento Valley of California. Environmental Entomology 19: 753763.Google Scholar
Hoffmann, MP, Wilson, LT, Zalom, FG & Hilton, RJ (1991) Dynamic sequential sampling plan for Helicoverpa zea (Lepidoptera: Noctuidae) eggs in processing tomatoes: parasitism and temporal patterns. Environmental Entomology 20: 10051012.Google Scholar
Hoffmeister, T (1992) Factors determining the structure and diversity of parasitoid complexes in Tephritid fruit flies. Oecologia 89: 288297.Google Scholar
Hoffmeister, TS, Babendreier, D & Wajnberg, E (2006) Statistical tools to improve the quality of experiments and data analysis for assessing non-target effects. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 222240.Google Scholar
Hoitink, HAJ & Boehm, MJ (1999) Biocontrol within the context of soil microbial communities: a substrate-dependent phenomenon. Annual Review of Phytopathology 37: 427446.Google Scholar
Hoitink, HAJ & Fahy, PC (1986) Basis for the control of soilborne plant pathogens with composts. Annual Review of Phytopathology 24: 93114.Google Scholar
Hokkanen, H (1985a) Exploiter-victim relationships of major plant diseases: implications for biological weed control. Agriculture, Ecosystems & Environment 14: 6376.Google Scholar
Hokkanen, HMT (1985b) Success in classical biological control. CRC Critical Reviews in Plant Sciences 3: 3572.Google Scholar
Hokkanen, HMT & Hajek, A (2003) Environmental Impacts of Microbial Insecticides. Dordrecht, The Netherlands, Kluwer.Google Scholar
Hokkanen, HMT & Lynch, JM (1995) Biological Control: Benefits and Risks. Cambridge, UK, 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 (1989) New associations in biological control: theory and practice. Canadian Entomologist. 121: 829840.Google Scholar
Holling, CS (1959) The components of predation as revealed by a study of small-mammal predation of the European pine sawfly. The Canadian Entomologist 41: 293320.Google Scholar
Holloway, AK, Heimpel, GE, Strand, MR & Antolin, MF (1999) Survival of diploid males in Bracon sp. near hebetor (Hymenoptera: Braconidae). Annals of the Entomological Society of America 92: 110116.Google Scholar
Holmes, PM (1990) Dispersal and predation in alien Acacia. Oecologia 83: 288290.Google Scholar
Holst, N & Ruggle, P (1997) A physiologically based model of pest-natural enemy interactions. Experimental and Applied Acarology 62: 325341.Google Scholar
Holt, RD (1977) Predation, apparent competition and the structure of prey communities. Theoretical Population Biology 12: 197229.Google Scholar
Holt, RD & Hassell, MP (1993) Environmental heterogeneity and the stability of host-parasitoid interactions. Journal of Animal Ecology 62: 89100.Google Scholar
Holt, RD & Hochberg, ME (1997) When is biological control evolutionarily stable (or is it?). Ecology 78: 16731683.Google Scholar
Holt, RD & Hochberg, ME (2001) Indirect interactions: community modules and biological control: a theoretical perspective. In Evaluating Indirect Ecological Effects of Biological Control (Wajnberg, E, Scott, JK & Quimby, PC, eds.), Wallingford, UK, CABI Publishing, pp. 1338.Google Scholar
Holt, RD, Hochberg, ME & Barfield, M (1999) Population dynamics and the evolutionary stability of biological control. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 219230.Google Scholar
Honegger, RE (1981) List of amphibians and reptiles either known or thought to have become extinct since 1600. Biological Conservation 19: 141158.Google Scholar
Hood, GM, Chesson, P & Pech, RP (2000) Biological control using sterilizing viruses: host suppression and competition between viruses in non-spatial models. Journal of Applied Ecology 37: 914925.Google Scholar
Hoogendoorn, M & Heimpel, GE (2002) Indirect interactions between an introduced and a native ladybird species mediated by a shared parasitoid. Biological Control 25: 224230.Google Scholar
Hoogendoorn, M & Heimpel, GE (2004) Competitive interactions between an exotic and a native ladybeetle: a field cage study. Entomologia Experimentalis et Applicata 111: 1928.Google Scholar
Hoopes, MF, Holt, RD & Holyoak, M (2005) The effects of spatial processes on two species interactions. In Spatial Dynamics and Ecological Communities (Holyoak, M, Leibhold, MA & Holt, RD, eds.), Chicago, IL, University of Chicago Press, pp. 3567.Google Scholar
Hoover, K, Schultz, CM, Lane, SS, Bonning, BC, Duffey, SS, McCutchen, BF & Hammock, BD (1995) Reduction in damage to cotton plants by recombinant baculovirus that knocks moribund larvae of Heliothis virescens off the plant. Biological Control 5: 419426.Google Scholar
Hopper, KR, Britch, SC & Wajnberg, E (2006) Risks of interbreeding between species used in biological control and native species, and methods for evaluating their occurrence and impact. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 7897.Google Scholar
Hopper, KR, Prager, SM & Heimpel, GE (2013) Is parasitoid acceptance of different host species dynamic? Functional Ecology 27: 12011211.Google Scholar
Hopper, KR & Roush, RT (1993) Mate finding, dispersal, number released, and the success of biological control introductions. Ecological Entomology 18: 321331.Google Scholar
Hopper, KR, Roush, RT & Powell, W (1993). Management of genetics of biological control introductions. Annual Review of Entomology 38: 2751.Google Scholar
Hornby, D (1983) Suppressive soils. Annual Review of Phytopathology 21: 6585.Google Scholar
Hossain, Z, Gurr, GM, Wratten, SD & Raman, A (2002) Habitat manipulation in lucerne Medicago sativa: arthropod population dynamics in harvested ‘refuge’ crop strips. Journal of Applied Ecology 39: 445454.Google Scholar
Hougardy, E & Mills, NJ (2006) The influence of host deprivation and egg expenditure on the rate of dispersal of a parasitoid following field release. Biological Control 37: 206213.Google Scholar
House, HL (1967) The decreasing occurrence of diapause in the fly Pseudosarcophaga affinis through laboratory-reared generations. Canadian Journal of Zoology 45: 149153.Google Scholar
Howarth, FG (1983) Classical biocontrol: panacea or Pandora’s box? Proceedings of the Hawaiian Entomological Society 24: 239244.Google Scholar
Howarth, FG (1991) Environmental impacts of classical biological control. Annual Review of Entomology 36: 485509.Google Scholar
Hoy, MA (1985) Recent advances in genetics and genetic improvement of the Phytoseiidae. Annual Review of Entomology 30: 345370.Google Scholar
Hoy, MA (2000a) Deploying transgenic arthropods in pest management programs: risks and realities. In Insect Transgenesis: Methods and Applications (Handler, AM & James, AA, eds.), Boca Raton, FL, CRC Press, pp. 335368.Google Scholar
Hoy, MA (2000b) Transgenic arthropods for pest management programs: risks and realities. Experimental and Applied Acarology 24: 463495.Google Scholar
Hoy, MA (2003) Insect Molecular Genetics: An Introduction to Principles and Applications. Second edition. San Diego, CA, Academic Press.Google Scholar
Hoy, MA (2008) Augmentative biological control. In Encyclopedia of Entomology, Second Edition (Capinera, JL, ed.), Dordrecht, The Netherlands, Springer, pp. 327334.Google Scholar
Hsiao, TH (1996) Studies of interaction between alfalfa weevil strains, Wolbachia endosymbionts and parasitoids. In The Ecology of Agricultural Pests (Symondson, WOC & Liddell, JE, eds.) London, UK, Chapman & Hall, pp. 5171.Google Scholar
Hu, G & St. Leger, J (2002) Field studies using a recombinant mycoinsecticide (Metarhizium anisopliae) reveal that it is rhizosphere competent. Applied and Environmental Microbiology 68: 63836387.Google Scholar
Huang, JW & Kuhlman, EG (1991a) Formulation of a soil amendment to control damping-off of slash pine-seedlings. Phytopathology 81: 163170.Google Scholar
Huang, JW & Kuhlman, EG (1991b) Mechanisms inhibiting damping-off pathogens of slash pine seedlings with a formulated soil amendment. Phytopathology 81: 171177.Google Scholar
Huang, NX, Enkegaard, A, Osborne, LS, Ramakers, PMJ, Messelink, GJ, Pijnakker, J & Murphy, G (2011) The banker plant method in biological control. Critical Reviews in Plant Sciences 30: 259278.Google Scholar
Huang, Z, Bonsall, RF, Mavrodi, DV, Weller, DM & Thomashow, LS (2004) Transformation of Pseudomonas fluorescens with genes for biosynthesis of phenazine-1-carboxylic acid improves biocontrol of rhizoctonia root rot and in situ antibiotic production. FEMS Microbiology Ecology 49: 243251.Google Scholar
Hufbauer, RA (2001) Pea aphid-parasitoid interactions: have parasitoids adapted to differential resistance? Ecology 82: 717725.Google Scholar
Hufbauer, RA (2002a) Aphid population dynamics: does resistance to parasitism influence population size? Ecological Entomology 27: 2532.Google Scholar
Hufbauer, RA (2002b) Evidence for nonadaptive evolution in parasitoid virulence following a biological control introduction. Ecological Applications 12: 6678.Google Scholar
Hufbauer, RA, Bogdanowicz, SM & Harrison, RG (2004) The population genetics of biological control introduction: mitochondrial DNA and microsatellite variation in native and introduced populations of Aphidius ervi, a parasitoid wasp. Molecular Ecology 13: 337348.Google Scholar
Hufbauer, RA & Roderick, GK (2005) Microevolution in biological control: mechanisms, patterns, and processes. Biological Control 35: 227239.Google Scholar
Hufbauer, RA & Via, S (1999) Evolution of an aphid-parasitoid interaction: variation in resistance to parasitism among aphid populations specialized on different plants. Evolution 53: 14351445.Google Scholar
Huffaker, CB (1957) Fundamentals of biological control of weeds. Hilgardia 27: 101157.Google Scholar
Huffaker, CB & Kennett, CE (1956) Experimental studies on predation: predation and cyclamen-mite populations on strawberries in California. Hilgardia 26: 191222.Google Scholar
Huffaker, CB & Kennett, CE (1959) A ten-year study of vegetational changes associated with biological control of Klamath weed. Journal of Range Management 12: 6982.Google Scholar
Huffaker, CB & Kennett, CE (1966) Biological control of olive scale Parlatoria oleae (Colvee) through the compensatory action of two introduced parasites. Hilgardia 37: 283335.Google Scholar
Huffaker, CB, Simmonds, FJ & Laing, JE (1976) The theoretical and empirical basis of biological control. In Theory and Practice of Biological Control (Huffaker, CB & Messenger, PS, eds.) New York, Academic Press, pp. 4178.Google Scholar
Hughes, RD & Bryce, MA (1984) Biological characterization of two biotypes of pea aphid, one susceptible and the other resistant to fungal pathogens, coexisting on lucerne in Australia. Entomologia Experimentalis et Applicata 36: 225229.Google Scholar
Hull, LA & Beers, EH (1985) Ecological selectivity: modifying chemical control practices to preserve natural enemies. In Biological Control in Agricultural IPM Systems (Hoy, MA & Herzog, DC, eds.), Orlando, FL, Academic Press pp. 103122.Google Scholar
Hull, LA, Hickey, KD & Kanour, WW (1983) Pesticide usage patterns and associated pest damage in commercial apple orchards of Pennsylvania. Journal of Economic Entomology 76: 577583.Google Scholar
Hultine, KR, Bean, DW, Dudley, TL & Gehring, CA (2015) Species introductions and their cascading impacts on biotic interactions in desert riparian ecosystems. Integrative and Comparative Biology 55: 587601.Google Scholar
Hunter, DM (2005) Mycopesticides as part of integrated pest management of locusts and grasshoppers. Journal of Orthoptera Research 14: 197201.Google Scholar
Hunter, DM, Milner, RJ, Scanlan, JC & Spurgin, PA (1999) Aerial treatment of the migratory locust, Locusta migratoria (L.) (Orthoptera: Acrididae) with Metarhizium anisopliae (Deuteromycotina: Hyphomycetes) in Australia. Crop Protection 18: 699704.Google Scholar
Hunter, MD & Price, PW (1992) Playing chutes and ladders – heterogeneity and the relative roles of bottom-up and top-down forces in natural communities. Ecology 73: 724732.Google Scholar
Hunter, MS (1999) Genetic conflict in natural enemies: a review, and consequences for the biological control of arthropods. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 231258.Google Scholar
Hunter, MS & Woolley, JB (2001) Evolution and behavioral ecology of heteronomous aphelinid parasitoids. Annual Review of Entomology 46: 251290.Google Scholar
Hunt-Joshi, TR, Root, RB & Blossey, B (2005) Disruption of weed biological control by an opportunistic mirid predator. Ecological Applications 15: 861870.Google Scholar
Hussey, NW & Scopes, N (1985) Biological Pest Control: The Glasshouse Experience. Ithaca, NY, Cornell University Press.Google Scholar
Hutchinson, GE (1959) Homage to Santa Rosalia or why are there so many kinds of animals? American Naturalist 93: 145159.Google Scholar
Hwang, AS, Northrup, SL, Alexander, JK, Vo, T & Edmands, S (2011) Long-term experimental hybrid swarms between moderately incompatible Tigriopus californicus populations: hybrid inferiority in early generations yields to hybrid superiority in later generations. Conservation Genetics 12: 895909.Google Scholar
Hynes, RK & Boyetchko, SM (2006) Research initiatives in the art and science of biopesticide formulations. Soil Biology & Biochemistry 38: 845849.Google Scholar
Ikeda, M, Reimbold, EA & Thiem, SM (2005) Functional analysis of the baculovirus host range gene, hrf-1. Virology 332: 602613.Google Scholar
Imms, AD (1937) Recent Advances in Entomology, Second edition. Philadelphia, PA, P. Blackiston’s Son & Co.Google Scholar
Inceoglu, AB, Kamita, SG & Hammock, BD (2006) Genetically modified baculoviruses: a historical overview and future outlook. Advances in Virus Research 68: 323360.Google Scholar
IPPC (2006) Guidelines for the Export, Shipment, Import and Release of Biological Control Agents and Other Beneficial Organisms. Rome, Italy, International Plant Protection Convention, Food and Agriculture Organization.Google Scholar
Ireson, JE, Gourlay, AH, Holloway, RJ, Chatterton, WS, Foster, SD & Kwong, RM (2008) Host specificity, establishment and dispersal of the gorse thrips, Sericothrips staphylinus Haliday (Thysanoptera: Thripidae), a biological control agent for gorse, Ulex europaeus L. (Fabaceae), in Australia. Biological Control 45: 460471.Google Scholar
Irvin, NA, Hoddle, MS, O’Brochta, DA, Carey, B & Atkinson, PW (2004) Assessing fitness costs for transgenic Aedes aegypti expressing the GFP marker and transposase genes. Proceedings of the National Academy of Sciences 101: 891896.Google Scholar
Irvin, NA, Suarez Espinosa, J & Hoddle, MS (2014) Maximum realised lifetime parasitism and occurrence of time limitation in Gonatocerus ashmeadi (Hymenoptera: Mymaridae) foraging in citrus orchards. Biocontrol Science and Technology 24: 662679.Google Scholar
Isaacs, R, Tuell, J, Fiedler, A, Gardiner, M & Landis, D (2009) Maximizing arthropod-mediated ecosystem services in agricultural landscapes: the role of native plants. Frontiers in Ecology and the Environment 7: 196203.Google Scholar
Ives, WGH & Muldrew, JA (1981) Pristiphora erichsonii (Hartig), Larch Sawfly (Hymenoptera: Tenthredinidae). In Biological Control Programmes against Insects and Weeds in Canada, 1969–1980 (Kelleher, JS & Hulme, MA, eds.), Slough, UK, Commonwealth Agricultural Bureau, pp. 369380.Google Scholar
Ives, AR & Carpenter, SR (2007) Stability and diversity of ecosystems. Science 317: 5862.Google Scholar
Ivlev, VS (1961) Experimental Ecology of the Feeding of Fish. New Haven, CT, Yale University Press.Google Scholar
Jackson, TA, Alves, SB & Pereira, RM (2000) Success in biological control of soil-dwelling insects by pathogens and nematodes. In Biological Control: Measures of Success (Gurr, G & Wratten, S, eds.), Dordrecht, The Netherlands, Kluwer, pp. 271296.Google Scholar
Jackson, TA, Crawford, AM & Glare, TR (2005) Oryctes virus – time for a new look at a useful biocontrol agent. Journal of Invertebrate Pathology 89: 9194.Google Scholar
Jacobsen, BJ, Zidack, NK & Larson, BJ (2004) The role of Bacillus-based biological control agents in integrated pest management systems: plant diseases. Phytopathology 94: 12721275.Google Scholar
Jacobson, RJ, Croft, P & Fenlon, J (2001) Suppressing establishment of Frankliniella occidentalis Pergande (Thysanoptera: Thripidae) in cucumber crops by prophylactic release of Amblyseius cucumeris Oudemans (Acarina: Phytoseiidae). Biocontrol Science and Technology 11: 2734.Google Scholar
Jacometti, M, Jorgensen, N & Wratten, S (2010) Enhancing biological control by an omnivorous lacewing: floral resources reduce aphid numbers at low aphid densities. Biological Control 55: 159165.Google Scholar
Jaksić, FM & Yáñez, JL (1983) Rabbit and fox introductions in Tierra del Fuego: history and assessment of the attempts at biological control of the rabbit infestation. Biological Conservation 26: 367374.Google Scholar
James, RR, McEvoy, PB & Cox, CS (1992) Combining the cinnabar moth (Tyria jacobaeae) and the ragwort leaf beetle (Longitarsus jacobaeae) for control of ragwort (Senecio jacobaea): an experimental analysis. Journal of Applied Ecology 29: 589596.Google Scholar
Janisiewicz, WJ, Tworkoski, TJ & Sharer, C (2000) Characterizing the mechanism of biological control of postharvest diseases on fruits with a simple method to study competition for nutrients. Phytopathology 90: 11961200.Google Scholar
Janssen, A (1999) Plants with spider-mite prey attract more predatory mites than clean plants under greenhouse conditions. Entomologia Experimentalis et Applicata 90: 191198.Google Scholar
Janssen, A, Montserrat, M, HilleRisLambers, R, De Roos, AM, Pallini, A & Sabelis, MW (2006) Intraguild predation usually does not disrupt biological control. In Trophic and Guild Interactions in Biological Control (Brodeur, J & Boivin, G, eds.), Dordrecht, The Netherlands, Springer, pp. 2144.Google Scholar
Janssen, A & Sabelis, MW (1992) Phytoseiid life histories, local predator–prey dynamics, and strategies for control of tetranychid mites. Experimental and Applied Acarology 14: 233250.Google Scholar
Jansson, JK (2003) Marker and reporter genes: illuminating tools for environmental microbiologists. Current Opinion in Microbiology 6: 310316.Google Scholar
Jaronski, ST (2012) Microbial control of invertebrate pests. In Beneficial Microorganisms in Agriculture, Food and the Environment (Sundh, I, Wilcks, A & Goettel, M, eds.), Wallingford, UK, CABI Publishing, pp. 7295.Google Scholar
Jaronski, ST & Jackson, MA (2012) Mass production of entomopathogenic Hypocreales. In Manual of Techniques in Invertebrate Pathology (Lacey, LA, ed.), San Diego, CA, Academic Press, pp. 255284.Google Scholar
Jarvis, PJ, Fowler, SV, Paynter, Q & Syrett, P (2006) Predicting the economic benefits and costs of introducing new biological control agents for Scotch broom Cytisus scoparius into New Zealand. Biological Control 39: 135146.Google Scholar
Jedlicka, JA, Greenberg, R & Letourneau, DK (2011) Avian conservation practices strengthen ecosystem services in California vineyards. PLoS One 6: e27347Google Scholar
Jeffries, MJ & Lawton, JH (1984) Enemy free space and the structure of ecological communities. Biological Journal of Linnean Society 23: 269286.Google Scholar
Jeger, MJ, Jeffries, P, Elad, Y. & Xu, X.-M. (2009) A generic theoretical model for biological control of foliar plant diseases. Journal of Theoretical Biology 256: 201214.Google Scholar
Jenner, WH, Kuhlmann, U, Miall, JH, Cappuccino, N & Mason, PG (2014) Does parasitoid state affect host range expression? Biological Control 78: 1522.Google Scholar
Jenkins, NE & Grzywacz, D (2000) Quality control of fungal and viral biocontrol agents: assurance of product performance. Biocontrol Science and Technology 10: 753777.Google Scholar
Jenkins, NE & Grzywacz, D (2003) Towards the standardisation of quality control of fungal and viral biocontrol agents. In Quality Control and Production of Biological Control Agents: Theory and Testing Procedures (Van Lenteren, JC, ed.), Wallingford, UK, CABI Publishing, pp. 247263.Google Scholar
Jensen, GL, Shelton, WL, Yang, SL & Wilken, LO (1983) Sex reversal of gynogenetic grass carp by implantation of methyltestosterone. Transactions of the American Fisheries Society 112: 7985.Google Scholar
Jervis, MA & Heimpel, GE (2005) Phytophagy. In Insects as Natural Enemies: A Practical Perspective (Jervis, MA, ed.), Dordrecht, The Netherlands, Springer, pp. 525550.Google Scholar
Jervis, MA & Kidd, NAC (1986) Host-feeding strategies in hymenopteran parasitoids. Biological Reviews 61: 396434.Google Scholar
Jervis, MA, Kidd, NAC & Walton, M (1992) A review of methods for determining dietary range in adult parasitoids. Entomophaga 37: 565574.Google Scholar
Jervis, MA, Lee, JC & Heimpel, GE (2004) Use of behavioural and life-history studies to understand the effects of habitat manipulation. In Ecological Engineering for Pest Management (Gurr, GM & Wratten, SD, eds.), Collingwood, Australia, CSIRO, pp. 65100.Google Scholar
Jervis, MA, Moe, A & Heimpel, GE (2012) The evolution of parasitoid fecundity: a paradigm under scrutiny. Ecology Letters 15: 357364.Google Scholar
Jeschke, JM, Kopp, M & Tollrian, R (2002) Predator functional responses: discriminating between handling and digesting prey. Ecological Monographs 72: 95112.Google Scholar
Jetter, K (2005) Economic framework for decision making in biological control. Biological Control 35: 348357.Google Scholar
Jezorek, H, Stiling, P & Carpenter, J (2011) Ant predation on an invasive herbivore: can an extrafloral nectar-producing plant provide associational resistance to Opuntia individuals? Biological Invasions 13: 22612273.Google Scholar
Jobin, A, Schaffner, U & Nentwig, W (1996) The structure of the phyophagous insect fanua on the introduced weed Solidago altissima in Switzerland. Entomologia Experimentalis et Applicata 79: 3342.Google Scholar
Johnson, DM & Stiling, P (1996) Host specificity of Cactoblastis cactorum (Lepidoptera: Pyralidae), an exotic Opuntia-feeding moth, in Florida. Environmental Entomology 25: 743748.Google Scholar
Johnson, DM & Stiling, PD (1998) Distribution and dispersal of Cactoblastis cactorum (Lepidoptera: Pyralidae), an exotic Opuntia-feeding moth, in Florida. Florida Entomologist 81: 1222.Google Scholar
Johnson, KB (1994) Dose-response relationships and inundative biological control. Phytopathology 84: 780784.Google Scholar
Johnson, KB (1999) Dose-response relationships in biocontrol of plant disease and their use to define pathogen refuge size. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 385392.Google Scholar
Johnson, KB (2010) Pathogen refuge: a key to understanding biological control. Annual Review of Phytopathology 48: 141160.Google Scholar
Johnson, KB & DiLeone, JA (1999) Effect of antibiosis on antagonist dose-plant disease response relationships for the biological control of crown gall of tomato and cherry. Phytopathology 89: 974980.Google Scholar
Johnson, MD, Kellermann, JL & Stercho, AM (2010) Pest reduction services by birds in shade and sun coffee in Jamaica. Animal Conservation 13: 140147.Google Scholar
Johnson, MD, Levy, NJ, Kellermann, JL & Robinson, DE (2009) Effects of shade and bird exclusion on arthropods and leaf damage on coffee farms in Jamaica’s Blue Mountains. Agroforestry Systems 76: 139148.Google Scholar
Johnson, MT, Follett, PA, Taylor, AD & Jones, VP (2005) Impacts of biological control and invasive species on a non-target native Hawaiian insect. Oecologia 142: 529540.Google Scholar
Johnson, MW & Tabashnik, BE (1999) Enhanced biological control through pesticide selectivity. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 297317.Google Scholar
Johnston, J (2001) Pesticides and Wildlife. Washington, DC, American Chemical Society.Google Scholar
Jones, CD (2005) The genetics of adaptation in Drosophila seychellia. Genetica 123: 137145.Google Scholar
Jones, DA, Ryder, MH, Clare, BG, Farrand, SK & Kerr, A (1988) Construction of a Tra-deletion mutant of pAgK84 to safeguard the biological control of crown gall. Molecular and General Genetics 212: 207214.Google Scholar
Jones, KA & Burges, D (1998) Technology of formulation and application. In Formulation of Microbial Pesticides (Burges, DH, ed.), Dordrecht, The Netherlands, Kluwer, pp. 730.Google Scholar
Jones, VP (1995) Reassessment of the role of predators and Trissolcus basalis in biological control of southern green stink bug (Hemiptera, Pentatomidae) in Hawaii. Biological Control 5: 566572.Google Scholar
Jones, VP, Westcott, DM, Finson, NN & Nishimoto, RK (2001) Relationship between community structure and southern green stink bug (Heteroptera: Pentatomidae) damage in macadamia nuts. Environmental Entomology 30: 10281035.Google Scholar
Jonsen, ID, Bourchier, RS & Roland, J (2007) Influence of dispersal, stochasticity, and an Allee effect on the persistence of weed biocontrol introductions. Ecological Modelling 203: 521526.Google Scholar
Jonsson, M, Buckley, HL, Case, BS, Wratten, SD, Hale, RJ & Didham, RK (2012) Agricultural intensification drives landscape-context effects on host-parasitoid interactions in agroecosystems. Journal of Applied Ecology 49: 706714.Google Scholar
Jonsson, M, Straub, CS, Didham, RK, Buckley, HL, Case, BS, Hale, RJ, Gratton, C & Wratten, SD (2015) Experimental evidence that the effectiveness of conservation biological control depends on landscape complexity. Journal of Applied Ecology 52: 12741282.Google Scholar
Jonsson, M, Wratten, SD, Landis, DA & Gurr, GM (2008) Recent advances in conservation biological control of arthropods by arthropods. Biological Control 45: 172175.Google Scholar
Jonsson, M, Wratten, SD, Landis, DA, Tompkins, JML & Cullen, R (2010) Habitat manipulation to mitigate the impacts of invasive arthropod pests. Biological Invasions 12: 29332945.Google Scholar
Jonsson, M, Wratten, SD, Robinson, KA & Sam, SA (2009) The impact of floral resources and omnivory on a four trophic level food web. Bulletin of Entomological Research 99: 275285.Google Scholar
Joshi, J & Vrieling, K (2005) The enemy release and EICA hypothesis revisited: incorporating the fundamental difference between specialist and generalist herbivores. Ecology Letters 8: 704714.Google Scholar
Julien, MH (1982) Biological Control of Weeds: A World Catalogue of Agents and Their Target Weeds. Slough, UK, Commonwealth Agricultural Bureaux.Google Scholar
Julien, MH (1987) Biological Control of Weeds: A World Catalogue of Agents and Their Target Weeds, 2nd edition. Wallingford, UK, CABI Publishing.Google Scholar
Julien, MH (1992) Biological Control of Weeds: A World Catalogue of Agents and Their Target Weeds, 3rd edition. Wallingford, UK, CABI Publishing.Google Scholar
Julien, MH & Griffiths, MW (1998) Biological Control of Weeds: A World Catalogue of Agents and Their Target Weeds, 4th edition. Wallingford, UK, CABI Publishing.Google Scholar
Kabaluk, T & Gazdik, Z (2007) Directory of Microbial Pesticides for Agricultural Crops in OECD Countries. Agriculture and AgriFood Canada, http://publications.gc.ca/site/eng/359060/publication.htmlGoogle Scholar
Kadir, JB, Charudattan, R, Stall, WM & Brecke, BJ (2000) Field efficacy of Dactylaria higginsii as a bioherbicide for the control of purple nutsedge (Cyperus rotundus). Weed Technology 14: 16.Google Scholar
Kajita, Y, Takano, F, Yasuda, H & Agarwala, BK (2000) Effects of indigenous ladybird species (Coleoptera: Coccinellidae) on the survival of an exotic species in relation to prey abundance. Applied Entomology and Zoology 35: 473479.Google Scholar
Kajita, Y, Takano, F, Yasuda, H & Evans, EW (2006a) Interactions between introduced and native predatory ladybirds (Coleoptera, Coccinellidae): factors influencing the success of species introductions. Ecological Entomology 31: 5867.Google Scholar
Kajita, Y, Yasuda, H & Evans, EW (2006b) Effects of native ladybirds on oviposition of the exotic species, Adalia bipunctata (Coleoptera: Coccinellidae), in Japan. Applied Entomology and Zoology 41: 5761.Google Scholar
Kaplan, I & Thaler, JS (2010) Plant resistance attenuates the consumptive and non-consumptive impacts of predators on prey. Oikos 119: 11051113.Google Scholar
Kapranas, A, Morse, JG, Pacheco, P, Forster, LD & Luck, RF (2007) Survey of brown soft scale Coccus hesperidum L. parasitoids in southern California citrus. Biological Control 42: 288299.Google Scholar
Kapuscinski, AR & Patronski, TJ (2005) Genetic Methods for Biological Control of Non-native Fish in the Gila River Basin. St. Paul, University of Minnesota, Institute for Social Economic and Ecological Sustainability.Google Scholar
Karban, R & Baldwin, IT (2000) Induced Responses to Herbivory. Chicago, IL, University of Chicago Press.Google Scholar
Karban, R, English-Loeb, G & Hougen-Eitzman, D (1997) Mite vaccinations for sustainable management of spider mites in vineyards. Ecological Applications 7: 183193.Google Scholar
Kareiva, P (1990) Establishing a foothold for theory in biocontrol practice: using models to guide experimental design and release protocols. In New Directions in Biological Control: Alternatives for Suppressing Agricultural Pests and Diseases (Baker, RR & Dunn, PE, eds.), New York, Alan R. Liss, pp. 6581.Google Scholar
Kareiva, P (1996) Contributions of ecology to biological control. Ecology 77: 19631964.Google Scholar
Karimzadeh, R, Hejazi, MJ, Helali, H, Iranipour, S & Mohammadi, SA (2011) Assessing the impact of site-specific spraying on control of Eurygaster integriceps (Hemiptera: Scutelleridae) damage and natural enemies. Precision Agriculture 12: 576593.Google Scholar
Kaser, JM & Heimpel, GE (2015) Linking risk and efficacy in biological control host-parasitoid models. Biological Control 90: 4960.Google Scholar
Kaser, JM & Ode, PJ (2016) Hidden risks and benefits of natural enemy mediated indirect effects Current Opinion in Insect Science 14: 105111.Google Scholar
Kassa, A, Stephan, D, Vidal, S & Zimmermann, G (2004) Laboratory and field evaluation of different formulations of Metarhizium anisopliae var. acridum submerged spores and aerial conidia for the control of locusts and grasshoppers. BioControl 49: 6381.Google Scholar
Kaya, HK & Gaugler, R (1993) Entomopathogenic nematodes. Annual Review of Entomology 38: 181206.Google Scholar
Kean, JM & Barlow, ND (2000a) Can host-parasitoid metapopulations explain successful biological control? Ecology 81: 21882197.Google Scholar
Kean, JM & Barlow, ND (2000b) Effects of dispersal on local population increase. Ecology Letters 3: 479482.Google Scholar
Kean, JM & Barlow, ND (2000c) Long-term assessment of the biological control of Sitona discoideus by Microctonus aethiopoides and test of a model. Biocontrol Science and Technology 10: 215221.Google Scholar
Kean, JM & Barlow, ND (2001) A spatial model for the successful biological control of Sitona discoideus by Microctonus aethiopoides. Journal of Applied Ecology 38: 162169.Google Scholar
Kean, J, Wratten, S, Tylianakis, J & Barlow, N (2003) The population consequences of natural enemy enhancement, and implications for conservation biological control. Ecology Letters 6: 604612.Google Scholar
Keane, RM & Crawley, MJ (2002) Exotic plant invasions and the enemy release hypothesis. Trends in Ecology & Evolution 17: 164170.Google Scholar
Kearney, MR & Porter, WP (2009) Mechanistic niche modelling: combining physiological and spatial data to predict species’ ranges. Ecology Letters 12: 334350.Google Scholar
Kelch, DG & McClay, A (2004) Putting the phylogeny into the centrifugal phylogenetic method. In XI International Symposium on Biological Control of Weeds (Cullen, JM, Briese, DT, Kriticos, DJ, Lonsdale, WM, Morin, L & Scott, JK, eds.), Canberra, Australia, CSIRO, pp. 287291.Google Scholar
Keller, S, Schweizer, C & Shah, P (1999) Differential susceptibility of two Melolontha populations to infections by the fungus Beauveria brongniartii. Biocontrol Science and Technology 9: 441446.CrossRefGoogle Scholar
Kellogg, SK, Fink, LS & Brower, LP (2003) Parasitism of native luna moths, Actias luna (L.) (Lepidoptera: Saturniidae) by the introduced Compsilura concinnata (Meigen) (Diptera: Tachinidae) in central Virginia, and their hyperparasitism by trigonalid wasps (Hymenoptera: Trigonalidae). Environmental Entomology 32: 10191027.Google Scholar
Kemp, JC & Barrett, GW (1989) Spatial patterning – impact of uncultivated corridors on arthropod populations within soybean agroecosystems. Ecology 70: 114128.Google Scholar
Kennedy, TA, Naeem, S, Howe, KM, Knops, MH, Tilman, D & Reich, P (2002) Biodiversity as a barrier to ecological invasion. Nature 417: 636638.Google Scholar
Kermack, WO & McKendrick, AG (1927) A contribution to the mathematical theory of epidemics. Proceedings of the Royal Society, A 115: 700721.Google Scholar
Kerr, PJ, Ghedin, E, DePasse, JV, Fitch, A, Cattadori, IM, Hudson, PJ, Tscharke, DC, Read, AF & Holmes, EC (2012) Evolutionary history and attenuation of myxoma virus on two continents. PLoS Pathogens 8: e1002950.Google Scholar
Kerr, P & McFadden, G (2002) Immune responses to myxoma virus. Viral Immunology 15: 229246.Google Scholar
Kerr, PJ, Rogers, MB, Fitch, A, Depasse, JV, Cattadori, IM, Twaddle, AC, Hudson, PJ, Tscharke, DC, Read, AF, Holmes, EC & Ghedin, E (2013) Genome scale evolution of myxoma virus reveals host-pathogen adaptation and rapid geographic spread. Journal of Virology 87: 1290012915.Google Scholar
Kessel, GJT, De Haas, BH, Van der Werf, W & Kohl, J (2002) Competitive substrate colonisation by Botrytis cinerea and Ulocladium atrum in relation to biological control of B. cinerea in cyclamen. Mycological Research 106: 716728.Google Scholar
Kettenring, KM & Mock, KE (2012) Genetic diversity, reproductive mode, and dispersal differ between the cryptic invader, Phragmites australis, and its native conspecific. Biological Invasions 14: 24892504.Google Scholar
Kevan, PG, Sutton, J & Shipp, L (2007) Pollinators as vectors of biocontrol agents – the B52 story. In Biological Control: A Global Perspective (Vincent, C, Goettel, M & Lazarovits, G, eds.), Wallingford, UK, CABI Publishing, pp. 319327.Google Scholar
Khan, ZR, Ampong Nyarko, K, Chiliswa, P, Hassanali, A, Kimani, S, Lwande, W, Overholt, WA, Pickett, JA, Smart, LE, Wadhams, LJ & Woodcock, CM (1997) Intercropping increases parasitism of pests. Nature 388: 631632.Google Scholar
Kidd, D & Amaresekare, P (2012) The role of transient dynamics in biological pest control: insights from a host-parasitoid community. Journal of Animal Ecology 81: 4757.CrossRefGoogle ScholarPubMed
Kimberling, DN (2004) Lessons from history: predicting successes and risks of intentional introductions for arthropod biological control. Biological Invasions 6: 301318.CrossRefGoogle Scholar
Kimbro, DL, Cheng, BS & Grosholz, ED (2013) Biotic resistance in marine environments. Ecology Letters 16: 821833.Google Scholar
Kindlmann, P & Dixon, AFG (1999) Strategies of aphidophagous predators: lessons from modelling insect predator–prey dynamics. Journal of Applied Entomology 123: 397399.Google Scholar
Kindlmann, P & Dixon, AFG (2001) When and why top-down regulation fails in arthropod predator–prey systems. Basic and Applied Ecology 2: 333340.CrossRefGoogle Scholar
King, C & Rubinoff, D (2008) First record of fossorial behavior in Hawaiian leafroller moth larvae, Omiodes continuatalis (Lepidoptera: Crambidae). Pacific Science 62: 147150.Google Scholar
King, GF, Escoubas, P & Nicholson, GM (2008) Peptide toxins that selectively target insect Na-V and Ca-V channels. Channels 2: 100116.Google Scholar
King, JR & Tschinkel, WR (2008) Experimental evidence that human impacts drive fire ant invasions and ecological change. Proceedings of the National Academy of Sciences of the United States of America 105: 2033920343.Google Scholar
Kinkel, LL, Bakker, MG & Schlatter, DC (2011) A coevolutionary framework for managing disease-suppressive soils. Annual Review of Phytopathology 49: 4767.Google Scholar
Kinzie, RA (1992) Predation by the introduced carnivorous snail Euglandina rosea (Ferussac) on endemic aquatic lymnaeid snails in Hawaii. Biological Conservation 60: 149155.Google Scholar
Kirkpatrick, JF & Frank, KM (2005) Contraception in free-ranging wildlife. In Wildlife Contraception: Issues, Methods, and Applications (Asa, CS & Porton, IJ, eds.), Baltimore, MD, Johns Hopkins University Press, pp. 195221.Google Scholar
Klingen, I & Haukeland, S (2006) The soil as a reservoir for natural enemies of pest insects and mites with emphasis on fungi and nematodes. In An Ecological and Societal Approach to Biological Control (Eilenberg, J & Hokkanen, HMT, eds.), Dordrecht, The Netherlands, Springer, pp. 145211.Google Scholar
Kloepper, J, Tuzun, S & Kuć, J (1992) Proposed definitions related to induced disease resistance. Biocontrol Science and Technology 2: 347349.CrossRefGoogle Scholar
Knipling, EF (1992) Principles of Insect Parasitism Analyzed from New Perspectives: Practical Implications for Regulating Insect Populations by Biological Means. Washington, DC, USDA-ARS Agriculture Handbook No. 693.Google Scholar
Knops, JMH, Tilman, D, Haddad, NM, Naeem, S, Mitchell, CE, Haarstad, J, Ritchie, ME, Howe, KM, Reich, PB, Siemann, E & Groth, J (1999) Effects of plant species richness on invasion dynamics, disease outbreaks, insect abundances and diversity. Ecology Letters 2: 286293.Google Scholar
Knutson, AE (2003) Release of Trichogramma pretiosum in cotton with a novel ground sprayer. Southwestern Entomologist 28: 1117.Google Scholar
Kolbe, JJ, Glor, RE, Rodríguez Schettino, L, Lara, AC, Larson, A & Losos, JB (2004) Genetic variation increases during biological invasion by a Cuban lizard. Nature 431: 177181.Google Scholar
Konishi, M & Ito, Y (1973) Early entomology in East Asia. In History of Entomology (Smith, RF, Mittler, TE & Smith, CN, eds.), Palo Alto, CA, Annual Reviews Inc., pp. 120.Google Scholar
Koss, AM, Jensen, AS, Schreiber, A, Pike, KS & Snyder, WE (2005) Comparison of predator and pest communities in Washington potato fields treated with broad-spectrum, selective, or organic insecticides. Environmental Entomology 34: 8795.Google Scholar
Kovaliski, J (1998) Monitoring the spread of rabbit hemorrhagic disease virus as a new biological agent for control of wild European rabbits in Australia. Journal of Wildlife Diseases 34: 421428.Google Scholar
Kowalchuk, GA, Os, GJ, Aartrijk, J & Veen, JA (2003) Microbial community responses to disease management soil treatments used in flower bulb cultivation. Biology and Fertility of Soils 37: 5563.Google Scholar
Kraaijeveld, AR (2004) Experimental evolution in host-parasitoid interactions. In Genetics, Evolution and Biological Control (Ehler, LE, Sforza, R & Mateille, T, eds.), Wallingford, UK, CABI Publishing, pp. 163181.Google Scholar
Kraaijeveld, AR, Ferrari, J & Godfray, HCJ (2002) Costs of resistance in insect-parasite and insect-parasitoid interactions. Parasitology 125: S71S82.Google Scholar
Kraaijeveld, AR & Van Alphen, JJM (1995) Geographical variation in encapsulation ability of Drosophila melanogaster larvae and evidence for parasitoid-specific components. Evolutionary Ecology 9: 1017.Google Scholar
Krafsur, ES (1998) Sterile insect technique for suppressing and eradicating insect populations: 55 years and counting. Journal of Agricultural Entomology 15: 303317.Google Scholar
Kramer, AM, Dennis, B, Liebhold, AM & Drake, JM (2009) The evidence for Allee effects. Population Ecology 51: 341354.Google Scholar
Kremer, RJ & Li, JM (2003) Developing weed-suppressive soils through improved soil quality management. Soil & Tillage Research 72: 193202.Google Scholar
Krenek, L & Rudolph, VHW (2014) Allometric scaling of indirect effects: body size ratios predict non-consumptive effects in multi-predator systems. Journal of Animal Ecology 83: 14611468.Google Scholar
Krips, OE, Willems, PEL & Dicke, M (1999) Compatibility of host plant resistance and biological control of the two-spotted spider mite Tetranychus urticae in the ornamental crop gerbera. Biological Control 16: 155163.Google Scholar
Krischik, VA, Landmark, AL & Heimpel, GE (2007) Soil-applied imidacloprid is translocated to nectar and kills nectar-feeding Anagyrus pseudococci (Girault) (Hymenoptera: Encyrtidae). Environmental Entomology 36: 12381245.Google Scholar
Kriticos, DJ (2003) The roles of ecological models in evaluating weed biological control agents and projects. In Improving the Selection, Testing and Evaluation of Weed Biological Control Agents (Spafford, Jacob H & Briese, DT, eds.), Osmond, Australia, CRC, pp. 6974.Google Scholar
Kriticos, DJ, Stuart, RM & Ash, JE (2004) Exploring interactions between cultural and biological control techniques: modelling bitou bush (Chrysantehmoides monilifera spp. rotundata) and a seed fly (Mesoclanis polana). In Proceedings of the XI International Symposium on Biological Control of Weeds (Cullen, JM, Briese, DT, Kriticos, DJ, Lonsdale, WM, Morin, L & Scott, JK, eds.), Canberra, Australia., CSIRO, p. 559.Google Scholar
Kruess, A & Tscharntke, T (1994) Habitat fragmentation, species loss, and biological control. Science 264: 15811584.Google Scholar
Kuhlmann, U, Mason, PG, Hinz, HL, Blossey, B, De Clerck-Floate, RA, Dosdall, LM, McCaffrey, JP, Schwarzlaender, M, Olfert, O, Brodeur, J, Gassmann, A, McClay, AS & Wiedenmann, RN (2006a) Avoiding conflicts between insect and weed biological control: selection of non-target species to assess host specificity of cabbage seedpod weevil parasitoids. Journal of Applied Entomology 130: 129141.Google Scholar
Kuhlmann, U, Schaffner, U & Mason, PG (2006b) Selection of non-target species for host specificity testing. In Environmental Impact of Invertebrates for Biological Control of Arthropods: Methods and Risk Assessment (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 1537.Google Scholar
Kumschick, S, Gaertner, M, Vilá, M, Essl, F, Jeschke, JM, Pysek, P, Ricciardi, A, Bacher, S, Blackburn, TM, Dick, JTA, Evans, T, Hulme, PE, Kuhn, I, Mrugala, A, Pergl, J, Rabitsch, W, Richardson, DM, Sendek, A & Winter, M (2015) Ecological impacts of alien species: quantification, scope, caveats, and recommendations. BioScience 65: 5563.Google Scholar
Kuris, AM (2003) Did biological control cause extinction of the coconut moth, Levuana iridescens, in Fiji? Biological Invasions 5: 133141.Google Scholar
Kuske, S, Babendreier, D, Edwards, PJ, Turlings, TCJ & Bigler, F (2004) Parasitism of non-target lepidoptera by mass-released Trichogramma brassicae and its implication for the larval parasitoid Lydella thomposoni. BioControl 49: 119.Google Scholar
Kuske, S, Widmer, F, Edwards, JP, Turlings, TCJ, Babendreier, D & Bigler, F (2003) Dispersal and persistence of mass released Trichogramma brassicae (Hymenoptera: Trichogrammatidae) in non-target habitats. Biological Control 27: 181193.Google Scholar
Kwak, Y-S & Weller, DM (2013) Take-all of wheat and natural disease suppression: a review. Plant Pathology Journal 29: 125135.Google Scholar
Kwok, OCH, Fahy, PC, Hoitink, HAJ & Kuter, GA (1987) Interactions between bacteria and Trichoderma hamatum in suppression of Rhizoctonia damping-off in bark compost media. Phytopathology 77: 12061212.Google Scholar
Lacey, LA & Arthurs, SP (2005) New method for testing solar sensitivity of commercial formulations of the granulovirus of codling moth (Cydia pomonella, Tortricidae: Lepidoptera). Journal of Invertebrate Pathology 90: 8590.Google Scholar
Lacey, LA, Grzywacz, D, Shapiro-Ilan, DI, Frutos, R, Brownbridge, M, Goettel, MS (2015) Insect pathogens as biological control agents: back to the future. Journal of Invertebrate Pathology 132: 141.Google Scholar
Lacey, LA, Kaya, HK & Vail, P (2001) Insect pathogens as biological control agents: do they have a future? Biological Control 21: 230248.Google Scholar
Lacey, LA, Thomson, D, Vincent, C & Arthurs, SP (2008) Codling moth granulovirus: a comprehensive review. Biocontrol Science and Technology 18: 639663.Google Scholar
Lafferty, KD & Kuris, AM (1996) Biological control of marine pests. Ecology 77: 19892000.Google Scholar
Lagnaoui, A & Radcliffe, EB (1997) Interference of fungicides with entomopathogens: effects on entomophthoran pathogens of green peach aphids. In Ecological Interactions and Biological Control (Andow, DA, Ragsdale, DW & Nyvall, RF, eds.), Boulder, CO, Westview, pp. 301315.Google Scholar
Lagnaoui, A & Radcliffe, EB (1998) Potato fungicides interfere with entomopathogenic fungi impacting population dynamics of green peach aphid. American Journal of Potato Research 75: 1925.Google Scholar
Laha, M & Mattingly, HT (2007) Ex situ evaluation of impacts of invasive mosquitofish on the imperiled Barrens topminnow. Environmental Biology of Fishes 78: 111.Google Scholar
Lambrinos, JG (2004) How interactions between ecology and evolution influence contemporary invasion dynamics. Ecology 85: 20612070.Google Scholar
Landis, DA, Wratten, SD & Gurr, GM (2000) Habitat management to conserve natural enemies of arthropod pests in agriculture. Annual Review of Entomology 45: 175201.Google Scholar
Lane, SD, Mills, NJ & Getz, WM (1999) The effects of parasitoid fecundity and host taxon on the biological control of insect pests: the relationship between theory and data. Ecological Entomology 24: 181190.Google Scholar
Lane, SD, St. Mary, CM & Getz, WM (2006) Coexistence of attack-limited parasitoids sequentially exploiting the same resource and its implications for biological control. Annales Zoologici Fennici 43: 1734.Google Scholar
Langellotto, GA & Denno, RF (2004) Responses of invertebrate natural enemies to complex-structured habitats: a meta-analytical synthesis. Oecologia 139: 110.Google Scholar
Langer, A & Hance, T (2004) Enhancing parasitism of wheat aphids through apparent competition: a tool for biological control. Agriculture Ecosystems & Environment 102: 205212.Google Scholar
Langewald, J & Kooyman, C (2007) Green Muscle, a fungal biopesticide for control of locusts and grasshoppers in Africa. In Biological Control: A Global Perspective (Vincent, C, Goettel, M & Lazarovits, G, eds.), Wallingford, UK, CABI Publishing, pp 311318.Google Scholar
Larkin, RP & Fravel, DR (1999) Mechanisms of action and dose-response relationships governing biological control of fusarium wilt of tomato by nonpathogenic Fusarium spp. Phytopathology 89: 11521161.Google Scholar
Larsen, A & Philpott, SM (2010) Twig-nesting ants: the hidden predators of the coffee berry borer in Chiapas, Mexico. Biotropica 42: 342347.Google Scholar
Larson, DL, Grace, JB, Rabie, PA & Andersen, P (2007) Short-term disruption of a leafy spurge (Euphorbia esula) biocontrol program following herbicide application. Biological Control 40: 18.Google Scholar
Lau, JA, Puliafico, KP, Kopshever, JA, Steltzer, H, Jarvis, EP, Schwarzländer, M, Strauss, SY & Hufbauer, RA (2008) Inference of allelopathy is complicated by effects of activated carbon on plant growth. New Phytologist 178: 412423.Google Scholar
Lavandero, BI, Wratten, SD, Didham, RK & Gurr, GM (2006) Increasing floral diversity for selective enhancement of biological control agents: a double-edged sward? Basic and Applied Ecology 7: 236243.Google Scholar
Lavandero, B, Wratten, S, Shishehbor, P & Worner, S (2005) Enhancing the effectiveness of the parasitoid Diadegma semiclausum (Helen): movement after use of nectar in the field. Biological Control 34: 152158.Google Scholar
Laven, H (1967) Eradication of Culex pipiens fatigans through cytoplasmic incompatibility. Nature 216: 383384.Google Scholar
Lavergne, S & Molofsky, J (2007) Increased genetic variation and evolutionary potential drive the success of an invasive grass. Proceedings of the National Academy of Sciences, USA 104: 38833888.Google Scholar
Lawton, JH (1985) Ecological theory and choice of biological control agents. Proceedings VI International Symposium on the Biological Control of Weeds (Delfosse, S, ed.), Ottawa, Canada, Agriculture Canada, pp. 1326.Google Scholar
le Masurier, AD & Waage, JK (1993) A comparison of attack rates in a native and an introduced population of the parasitoid Cotesia glomerata. Biocontrol Science and Technology 3: 467474.Google Scholar
Lee, CE (2002) Evolutionary genetics of invasive species. Trends in Ecology & Evolution, 17: 386391.Google Scholar
Lee, JC, Andow, DA & Heimpel, GE (2006) Influence of floral resources on sugar feeding and nutrient dynamics of a parasitoid in the field. Ecological Entomology 31: 470480.Google Scholar
Lee, JC & Heimpel, GE (2005) Impact of flowering buckwheat on lepidopteran cabbage pests and their parasitoids at two spatial scales. Biological Control 34: 290301.Google Scholar
Lee, JC & Heimpel, GE (2008) Floral resources impact longevity and oviposition rate of a parasitoid in the field. Journal of Animal Ecology 77: 565572.Google Scholar
Lee, JC, Menalled, FD & Landis, DA (2001) Refuge habitats modify impact of insecticide disturbance on carabid beetle communities. Journal of Applied Ecology 38: 472483.Google Scholar
Legner, EF (1972) Observations on hybridization and heterosis in parasitoids of synanthropic flies. Annals of the Entomological Society of America 65: 254263.Google Scholar
Legner, EF (2008) Biological pest control: a history. In Encyclopedia of Pest Management (Pimentel, D, ed.), London, UK, Taylor & Francis.Google Scholar
Leland, JE, Mullins, DE, Vaughan, LJ and Warren, HL (2005) Effects of media composition on submerged culture spores of the entomopathogenic fungus, Metarhizium anisopliae var. acridum, Part 1: Comparison of cell wall characteristics and drying stability among three spore types. Biocontrol Science and Technology 15: 379392.Google Scholar
Lemke, A & Poehling, HM (2002) Sown weed strips in cereal fields: overwintering site and ‘source’ habitat for Oedothorax apicatus (Blackwall) and Erigone atra (Blackwall) (Araneae: Erigonidae). Agriculture Ecosystems & Environment 90: 6780.Google Scholar
Leonardo, TE (2004) Removal of a specialization-associated symbiont does not affect aphid fitness. Ecology Letters 7: 461468.Google Scholar
Lester, PJ, Thistlewood, HMA & Harmsen, R (1998) The effects of refuge size and number on acarine predator–prey dynamics in a pesticide-disturbed apple orchard. Journal of Applied Ecology 35: 323331.Google Scholar
Letourneau, DK & Bothwell, SG (2008) Comparison of organic and conventional farms: challenging ecologists to make biodiversity functional. Frontiers in Ecology and the Environment 6: 430438.Google Scholar
Lever, C (1985) Naturalized Mammals of the World. London, UK, Longman.Google Scholar
Levin, S & Pimentel, D (1981) Selection of intermediate rates of increase in parasite-host systems. American Naturalist 117: 308315.Google Scholar
Levine, JM, Adler, PB & Yelenik, SG (2004) A meta-analysis of biotic resistance to exotic plant invasions. Ecology Letters 7: 975989.Google Scholar
Levine, JM & D’Antonio, CM (1999) Elton revisited: a review of evidence linking diversity and invasibility. Oikos 87: 1526.Google Scholar
Levine, JM, Vila, M, D’Antonio, CM, Dukes, JS, Grigulis, K & Lavorel, S (2003) Mechanisms underlying the impacts of exotic plant invasions. Proceedings of the Royal Society B-Biological Sciences 270: 775781.Google Scholar
Levins, R (1969) Some demographic and genetic consequences of environmental heterogeneity for biological control. Bulletin of the Entomological Society of America 15: 237240.Google Scholar
Levins, R (1975) Evolution in communities near equilibrium. Ecology and Evolution of Communities (Cody, ML & Diamond, JM, eds.), Cambridge, MA, Harvard University Press, pp. 1650.Google Scholar
Lewis, EE, Campbell, JF & Gaugler, R (1998) A conservation approach to using entomopathogenic nematodes in turf and landscapes. In Conservation Biological Control (Barbosa, P, ed.), San Diego, CA, Academic Press, pp. 235254.Google Scholar
Lewis, MA & Kareiva, P (1993) Allee dynamics and the spread of invading organisms. Theoretical Population Biology 43: 141158.Google Scholar
Leyse, KE, Lawler, SP & Strange, T (2004) Effects of an alien fish, Gambusia affinis, on an endemic California fairy shrimp, Linderiella occidentalis: implications for conservation of diversity in fishless waters. Biological Conservation 118: 5765.Google Scholar
Li, S, Falabella, P, Giannanonio, S, Fanti, P, Battaglia, D, Digilio, MC, Völkl, W, Sloggett, JJ, Weisser, W & Pennacchio, F (2002) Pea aphid clonal resistance to the endophagous parasitoid Aphidius ervi. Journal of Insect Physiology 48: 971980.Google Scholar
Liebhold, AM & Elkinton, JS (1989) Elevated parasitism in artificially augmented populations of Lymantria dispar (Lepidoptera: Lymantriidae). Environmental Entomology 18: 986995.Google Scholar
Liebhold, AM & Tobin, PC (2008) Population ecology of insect invasions and their management. Annual Review of Entomology 53: 387408.Google Scholar
Liljesthroom, G & Rabinovich, J (2004) Modeling biological control: the population regulation of Nezara viridula by Trichopoda giacomellii. Ecological Applications 14: 254267.CrossRefGoogle Scholar
Lin, BB (2007) Agroforestry management as an adaptive strategy against potential microclimate extremes in coffee agriculture. Agricultural and Forest Meteorology 144: 8594.Google Scholar
Lincango, MP, Causton, CE, Alvarez, CC & Jimenez-Uzcategui, G (2011) Evaluating the safety of Rodolia cardinalis to two species of Galapagos finch; Camarhynchus parvulus and Geospiza fuliginosa. Biological Control 56: 145149.Google Scholar
Lindow, SE & Leveau, JE (2002) Phyllosphere microbiology. Current Opinions in Biotechnology 13: 238243.Google Scholar
Liu, D, Kinkel, LL, Eckwall, EC, Anderson, NA & Schottel, JL (1997) Biological control of plant disease using antagonistic Streptomyces. In Ecological Interactions and Biological Control (Andow, DA, Ragsdale, DW & Nyvall, RF, eds.), Boulder, CO, Westview, pp. 224239.Google Scholar
Liu, H & Stiling, P (2006) Testing the enemy release hypothesis: a review and meta-analysis. Biological Invasions 8: 15351545.Google Scholar
Liu, H, Stiling, P & Pemberton, RW (2007) Does enemy release matter for invasive plants? Evidence from a comparison of insect herbivore damage among invasive, non-invasive and native congeners. Biological Invasions 9: 773781.Google Scholar
Liu, XX, Chen, M, Collins, HL, Onstad, DW, Roush, RT, Zhang, QW, Earle, ED & Shelton, AM (2014) Natural enemies delay insect resistance to Bt crops. PLoS One 9: e90366.Google Scholar
Lloyd, CJ, Hufbauer, RA, Jackson, A, Nissen, SJ & Norton, AP (2005). Pre- and post-introduction patterns in neutral genetic diversity in the leafy spurge gall midge, Spurgia capitigena (Bremi) (Diptera: Cecidomyiidae). Biological Control 33: 153164.Google Scholar
Lockwood, JA (1993) Environmental issues involved in biological control of rangeland grasshoppers (Orthoptera: Acrididae) with exotic agents. Environmental Entomology 22: 503518.Google Scholar
Lockwood, JA, Cassey, P & Blackburn, T (2005) The role of propagule pressure in explaining species invasions. Trends in Ecology & Evolution 20: 223228.Google Scholar
Lockwood, JL, Hoopes, MF & Marchetti, MP (2013) Invasion Ecology, 2nd edition. Malden, MA, Blackwell.Google Scholar
Lockwood, JL, Simberloff, D, McKinney, ML & Von Holle, B (2001) How many, and which, plants will invade natural areas? Biological Invasions 3: 18.Google Scholar
Lombaert, E, Guillemaud, T, Lundgren, J, Koch, R, Facon, B, Grez, A, Loomans, A, Malausa, T, Nedved, O, Rhule, E, Staverlokk, A, Steenberg, T & Estoup, A (2014) Complementarity of statistical treatments to reconstruct worldwide routes of invasion: the case of the Asian ladybird Harmonia axyridis. Molecular Ecology 23: 59795997.Google Scholar
Lomer, CJ, Bateman, RJ, Dent, D, de Groote, H, Duoro-Kpindou, O-K, Kooyman, C, Langewald, J, Ouambama, Z, Peveling, R & Thomas, M (1999) Development of strategies for the incorporation of biological pesticides into the integrated pest management of locusts and grasshoppers. Agricultural and Forest Entomology 1: 7188.Google Scholar
Lomer, CJ, Bateman, RP, Johnson, DL, Langewald, J & Thomas, M (2001) Biological control of locusts and grasshoppers. Annual Review of Entomology 46: 667702.Google Scholar
Lonsdale, WM (1999) Global patterns of plant invasions and the concept of invasibility. Ecology 80: 15221536.Google Scholar
Lonsdale, WM, Briese, DT & Cullen, JM (2001) Risk analysis and weed biological control. In Evaluating Indirect Ecological Effects of Biological Control (Wajnberg, E, Scott, JK & Quimby, PC, eds.), Wallingford, UK, CABI Publishing, pp. 185210.Google Scholar
Lonsdale, WM, Farrell, G & Wilson, CG (1995) Biological control of a tropical weed: a population model and experiment for Sida acuta. Journal of Applied Ecology 32: 391399.Google Scholar
Loper, JE, Kobayashi, DY & Paulsen, IT (2007) The genomic sequence of Pseudomonas fluorescens Pf-5: insights into biological control. Phytopathology 97: 233238.Google Scholar
Loreau, M, Mouquet, N & Gonzalez, A (2003) Biodiversity as spatial insurance in heterogeneous landscapes. Proceedings of the National Academy of Sciences of the United States of America 100: 1276512770.Google Scholar
Losey, JE & Denno, RF (1998) Positive predator-predator interactions: enhanced predation rates and synergistic suppression of aphid populations. Ecology 79: 21432152.Google Scholar
Losey, JE & Vaughan, M (2006) The economic value of ecological services provided by insects. BioScience 56: 311323.Google Scholar
Lotka, AJ (1923) Contribution to quantitative parasitology. Journal of the Washington Academy of Sciences 13: 152158.Google Scholar
Louda, SM (1998) Population growth of Rhinocyllus conicus (Coleoptera: Curculionidae) on two species of native thistles in Prairie. Environmental Entomology 27: 834841.Google Scholar
Louda, SM & Arnett, AE (2000) Predicting non-target ecological effects of biological control agents: evidence from Rhinocyllus conicus. In Proceedings of the X International Symposium on Biological Control of Weeds (Spencer, NR, ed.), Bozeman, Montana State University, pp. 551567.Google Scholar
Louda, SM, 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, SM & O’Brien, CW (2002) Unexpected ecological effects of distributing the exotic weevil, Larinus planus (F.), for the biological control of Canada thistle. Conservation Biology 16: 717727.Google Scholar
Louda, SM, Pemberton, RW, Johnson, MT & Follett, PA (2003) Nontarget effects: the Achilles’ heel of biological control? Annual Review of Entomology 48: 365396.Google Scholar
Louda, SM & Potvin, MA (1995) Effect of inflorescence-feeding insects on the demography and lifetime firness of a native plant. Ecology 76: 229245.Google Scholar
Louda, SM, Potvin, MA & Collinge, SK (1990) Predispersal seed predation, postdispersal seed predation and competition in the recruitment of seedlings of a native thistle in sandhills prairie. The American Midland Naturalist 124: 105113.Google Scholar
Louda, SM, Rand, TA, Arnett, AE, McClay, AS, Shea, K & McEachern, AK (2005a) Evaluation of ecological risk to populations of a threatened plant from an invasive biocontrol insect. Ecological Applications 15: 234249.Google Scholar
Louda, SM, Rand, TA, Russell, FL & Arnett, AE (2005b) Assessment of ecological risks in weed biocontrol: input from retrospective ecological analyses. Biological Control 35: 253264.Google Scholar
Lovei, GL, Andow, DA & Arpaia, S (2009) Transgenic insecticidal crops and natural enemies: a detailed review of laboratory studies. Environmental Entomology 38: 293306.Google Scholar
Lozier, JD & Mills, NJ (2011) Predicting the potential invasive range of light brown apple moth (Epiphyas postvittana) using biologically informed and correlative species distribution models. Biological Invasions 13: 24092421.Google Scholar
Lu, A & Miller, LK (1995) Differential requirements for baculovirus late expression factor genes in two cell lines. Journal of Virology 69: 62656272.Google Scholar
Lu, YH, Wu, KM, Jiang, YY, Guo, YY & Desneux, N (2012) Widespread adoption of Bt cotton and insecticide decrease promotes biocontrol services. Nature 487: 362365.Google Scholar
Lucas, AM (1969) The effect of population structure on the success of insect introductions. Heredity 24: 151154.Google Scholar
Lucas, P & Sarniguet, A (1998) Biological control of soil-borne pathogens with resident versus introduced antagonists: should diverging approaches become strategic convergence? In Conservation Biological Control (Barbosa, P, ed.), San Diego, CA, Academic Press, pp. 351370.Google Scholar
Luck, RF & Forster, LD (2003) Quality of augmentative biological control agents: a historical perspective and lessons learned from evaluating Trichogramma. In Quality Control and Production of Biological Control Agents: Theory and Testing Procedures (Van Lenteren, JC, ed.), Wallingford, UK, CABI Publishing, pp. 231246.Google Scholar
Luck, RF, Messenger, PS & Barbieri, JF (1981) The influence of hyperparasitism on the performance of biological control agents. In The Role of Hyperparasitism in Biological Control (Rosen, D, ed.), Berkeley, University of California Press, pp. 3442.Google Scholar
Luck, RF & Podoler, H (1985) Competitive exclusion of Aphytis lingnanensis by A. melinus: potential role of host size. Ecology 66: 904913.Google Scholar
Luck, RF, Shepard, BM & Kenmore, PE (1988) Experimental methods for evaluating arthropod natural enemies. Annual Review of Entomology 33: 367391.Google Scholar
Luck, RF, Shepard, BM & Kenmore, PE (1999) Evaluation of biological control with experimental methods. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 225242.Google Scholar
Luck, RF, Van den Bosch, R & Garcia, R (1977) Chemical insect control – a troubled pest management strategy. BioScience 27: 606611.Google Scholar
Lundgren, JG (2009) Relationships of Natural Enemies and Non-prey Foods. Dordrecht, The Netherlands, Springer.Google Scholar
Lundgren, JG & Fergen, JK (2010) The effects of a winter cover crop on Diabrotica virgifera (Coleoptera: Chrysomelidae) populations and beneficial arthropod communities in no-till maize. Environmental Entomology 39: 18161828.Google Scholar
Lundgren, JG, Gassmann, AJ, Bernal, J, Duan, JJ & Ruberson, J (2009) Ecological compatibility of GM crops and biological control. Crop Protection 28: 10171030.Google Scholar
Lynch, M (1991) The genetic interpretation of inbreeding depression and outbreeding depression. Evolution 45: 622629.Google Scholar
Lys, J-A & Nentwig, W (1992) Augmentation of beneficial arthropods by strip-management. 4. Surface activity, movements and activity density of abundant carabid beetles in a cereal field. Oecologia 92: 373382.Google Scholar
Lys, J-A, Zimmermann, M & Nentwig, W (1994) Increase in activity density and species number of carabid beetles in cereals as a result of strip-management. Entomologia Experimentalis et Applicata 73: 19.Google Scholar
Ma, WJ, Kuijper, B, de Boer, JG, van de Zande, L, Beukeboom, LW, Wertheim, B & Pannebakker, BA (2013) Absence of complementary sex determination in the parasitoid wasp genus Asobara (Hymenoptera: Braconidae). PLoS One 8: e60459.Google Scholar
MacArthur, RH & Wilson, EO (1967) The Theory of Island Biogeography. Princeton, NJ, Princeton University Press.Google Scholar
MacDonald, IAW & Cooper, J (1995) Insular lessons for global biodiversity conservation with particular reference to alien invasions. In Islands: Biological Diversity and Ecosystem Function (Vitousek, PM, Loope, LL & Adsersen, H, eds.), Berlin, Germany, Springer, pp. 189203.Google Scholar
Macen, JL, Graham, KA, Lee, SF, Schreiber, M, Boshkov, LK & McFadden, G (1996) Expression of the myxoma virus tumor necrosis factor receptor homologue and M11 L genes is required to prevent virus-induced apoptosis in infected rabbit T lymphocytes. Virology 218: 232237.Google Scholar
Mack, RN (1996) Biotic barriers to plant naturalization. In Proceedings of the IX International Symposium on Biological Control of Weeds (Moran, VC & Hoffmann, JH, eds.), Stellenbosch, South Africa, University of Cape Town, pp. 3946.Google Scholar
Mack, RN (2002) Natural barriers to plant naturalizations and invasions in the Sonoran Desert. In Invasive Exotic Species in the Sonoral Region (Tellman, B, ed.), Tucson, University of Arizona Press, pp. 6376.Google Scholar
Mack, RN, Simberloff, D, Lonsdale, WM, Evans, H, Clout, M & Bazzaz, FA (2000) Biotic invasions: causes, epidemiology, global consequences, and control. Ecological Applications 10: 698710.Google Scholar
Mackauer, M (1976) Genetic problems in the production of biological control agents. Annual Review of Entomology 21: 369385.Google Scholar
Mackauer, M. & Volkl, W (1993) Regulation of aphid populations by aphidiid wasps: does parasitoid foraging behaviour or hyperparasitism limit impact? Oecologia 94: 339350.Google Scholar
MacLeod, A, Wratten, SD, Sotherton, NW & Thomas, MB (2004) ‘Beetle banks’ as refuges for beneficial arthropods in farmland: long-term changes in predator communities and habitat. Agricultural and Forest Entomology 6: 147154.Google Scholar
Maddox, DM (1982) Biological control of diffuse knapweed (Centaurea diffusa) and spotted knapweed (C. maculosa). Weed Science 30: 7682.Google Scholar
Maeda, S, Kamita, SG & Kondo, A (1993) Host-range expansion of Autographa californica nuclear polyhedrosis virus (Npv) following recombination of a 0.6-kilobase-pair DNA fragment originating from Bombyx mori Npv. Journal of Virology 67: 62346238.Google Scholar
Mafokoane, LD, Zimmermann, HG & Hill, MP (2007) Development of Cactoblastis cactorum (Berg) (Lepidoptera: pyralidae) on six north American Opuntia species. African Entomology 15: 295299.Google Scholar
Majerus, MEN (1994) Ladybirds. London, UK, Harper Collins.Google Scholar
Majerus, MEN & Hurst, GDD (1997) Ladybirds as a model system for the study of male-killing endosymbionts. Entomophaga 42: 1320.Google Scholar
Malecki, RA, Blossey, B, Hight, SD, Schroeder, D, Kok, LT & Coulson, JR (1993) Biological control of purple loosestrife. BioScience 43: 680686.Google Scholar
Malmstrom, CM, McCullough, AJ, Johnson, HA, Newton, LA & Borer, ET (2005) Invasive annual grasses indirectly increase virus incidence in California native perennial bunchgrasses. Oecologia 145: 153164.Google Scholar
Maniania, NK, Bugeme, DM, Wekesa, VW, Delalibera, I & Knapp, M (2008) Role of entomopathogenic fungi in the control of Tetranychus evansi and Tetranychus urticae (Acari: Tetranychidae), pests of horticultural crops. Experimental and Applied Acarology 46: 259274.Google Scholar
Manrique, V, Diaz, R, Erazo, L, Reddi, N, Wheeler, GS, Williams, D & Overholt, WA (2014) Comparison of two populations of Pseudophilothrips ichini (Thysanoptera: Phlaeothripidae) as candidates for biological control of the invasive weed Schinus terebinthifolia (Sapindales: Anacardiaceae). Biocontrol Science and Technology 24: 518535.Google Scholar
Mansfield, S & Mills, NJ (2004) A comparison of methodologies for the assessment of host preference of the gregarious egg parasitoid Trichogramma platneri. Biological Control 29: 332340.Google Scholar
Marchetto, KM, Shea, K, Kelly, D, Groenteman, R, Sezen, Z & Jongejans, E (2014) Unrecognized impact of a biocontrol agent on the spread rate of an invasive thistle. Ecological Applications 24: 11781187.Google Scholar
Marino, PC & Landis, DA (1996) Effect of landscape structure on parasitoid diversity and parasitism in agroecosystems. Ecological Applications 6: 276284.Google Scholar
Maron, JL & Harrison, S (1997) Spatial pattern formation in an insect host-parasitoid system. Science 278: 16191621.Google Scholar
Maron, JL & Marler, M (2008) Effects of native species diversity and resource additions on invader impact. American Naturalist 172: S18S33.Google Scholar
Maron, JL & Vilá, M (2001) When do herbivores affect plant invasion? Evidence for the natural enemies and biotic resistance hypotheses. Oikos 95: 361373.Google Scholar
Maron, JL, Vilá, M & Arnason, J (2004) Loss of enemy resistance among introduced populations of St. John’s Wort (Hypericum perforatum). Ecology 85: 32433253.Google Scholar
Marrone, PG (2002) An effective biofungicide with novel modes of action. Pesticide Outlook 13: 193194.Google Scholar
Marsden, JS, Martin, GE, Parham, DJ, Risdill-Smith, TJ & Johnson, BG (1980) Returns on Australian Agricultural Research. Canberra, Australian CSIRO Division of Entomology.Google Scholar
Marsh, FL (1937) Ecological observations upon the enemies of Cecropia, with particular reference to its hymenopterous parasites. Ecology 18: 106112.Google Scholar
Marsh, FL (1941) A few life-history details of Samia cecropia within the southwestern limits of Chicago. Ecology 22: 331337.Google Scholar
Marsh, PM (1977) Notes on the taxonomy and nomenclature of Aphidius species (Hym.: Aphidiidae) parasitic on the pea aphid in North America. Entomophaga 22: 365372.Google Scholar
Marshall, ID & Fenner, F (1958) Studies in the epidemiology of infectious myxomatosis of rabbits. V. Changes in the innate resistance of Australian wild rabbits exposed to myxomatosis. Journal of Hygiene 56: 288302.Google Scholar
Marsico, TD, Burt, JW, Espeland, EK, Gilchrist, GW, Jamies, MA, Linsstrom, L, Roderick, GK, Swope, S, Szucs, M & Tsutsui, ND (2010) Underutilized resources for studying the evolution of invasive species during their introduction, establishment, and lag phases. Evolutionary Applications 3: 203219.Google Scholar
Mason, PG & Hopper, KR (1997) Temperature dependence in locomotion of the parasitoid Aphelinus asychis (Hymenoptera: Aphelindae) from geographical regions with different climates. Environmental Entomology 26: 14161423.Google Scholar
Mason, RR & Torgerson, TR (1987) Dynamics of a nonoutbreak population of the douglas-fir tussock moth (Lepidoptera: Lymantriidae) in southern Oregon. Environmental Entomology 16: 12171227.Google Scholar
Massad, E (1987) Transmission rates and the evolution of pathogenicity. Evolution 41: 11271130.Google Scholar
Massei, G & Cowan, D (2014) Fertility control to mitigate human-wildlife conflicts: a review. Wildlife Research 41: 121.Google Scholar
Mathenge, CW, Holford, P, Hoffmann, JH, Zimmermann, HG, Spooner-Hart, R & Beattie, GAC (2010) Hybridization between Dactylopius tomentosus (Hemiptera: Dactylopiidae) biotypes and its effects on host specificity. Bulletin of Entomological Research 100: 331338.Google Scholar
Mathews, CR, Bottrell, DG & Brown, MW (2009) Extrafloral nectaries alter arthropod community structure and mediate peach (Prunus persica) plant defense. Ecological Applications 19: 722730.Google Scholar
Mathews, CR, Brown, MW & Bottrell, DG (2007) Leaf extrafloral nectaries enhance biological control of a key economic pest, Grapholita molesta (Lepidoptera: Tortricidae), in peach (Rosales: Rosaceae). Environmental Entomology 36: 383389.Google Scholar
Matsumoto, T, Itioka, T, Nishida, T & Inoue, T (2003) Introduction of parasitoids has maintained a stable population of arrowhead scales at extremely low levels. Entomologia Experimentalis et Applicata 106: 115125.Google Scholar
Matsuo, T, Sugaya, S, Yasukawa, J, Aigaki, T & Fuyama, Y (2007) Odorant-binding proteins OBP57d and OBP57e affect taste perception and host-plant preference in Drosophila seychellia. PLoS Biology 5: e118.Google Scholar
Matsuoka, T & Seno, H (2008) Ecological balance in the native population dynamics may cause the paradox of pest control with harvesting. Journal of Theoretical Biology 252: 8797.Google Scholar
Matteson, PC (2000) Insect pest management in tropical Asian irrigated rice. Annual Review of Entomology 45: 549574.Google Scholar
Mausel, DL, Salom, SM, Kok, LT & Fidgen, JG (2008) Propagation, synchrony, and impact of introduced and native Laricobius spp. (Coleoptera: Derodontidae) on hemlock woolly adelgid in Virginia. Environmental Entomology 37: 14981507.Google Scholar
Mausel, DL, Van Driesche, RG & Elkinton, JS (2011) Comparative cold tolerance and climate matching of coastal and inland Laricobius nigrinus (Coleoptera: Derodontidae), a biological control agent of hemlock woolly adelgid. Biological Control 58: 96102.Google Scholar
May, RM (1973) Stability and Complexity in Model Ecosystems. Princeton, NJ, Princeton University Press.Google Scholar
May, RM (1978) Host-parasitoid systems in patchy environments: a phenomenological model. Journal of Animal Ecology 47: 833844.Google Scholar
May, RM & Hassell, MP (1981) The dynamics of multiparasitoid-host interactions. American Naturalist 117: 234261.Google Scholar
May, RM, Hassell, MP, Anderson, RM & Tonkyn, DW (1981) Density dependence in host-parasitoid models. Journal of Animal Ecology 50: 855865.Google Scholar
Mayer, VE, Frederickson, ME, McKey, D & Blatrix, R (2014) Current issues in the evolutionary ecology of ant-plant symbioses. New Phytologist 202: 749764.Google Scholar
Maynard Smith, J, Smith, NH, O’Rourke, M & Spratt, BG (1993) How clonal are bacteria? Proceedings of the National Academy of Sciences, USA 90: 43844388.Google Scholar
Mazzola, M (2004) Assessment and management of soil microbial community structure for disease suppression. Annual Review of Phytopathology 42: 3559.Google Scholar
Mazzola, M, Brown, J, Izzo, AD & Cohen, MF (2007) Mechanism of action and efficacy of seed meal-induced pathogen suppression differ in a Brassicaceae species and time-dependent manner. Phytopathology 97: 454460.Google Scholar
Mazzola, M, Fujimoto, DK, Thomashow, LS & Cook, RJ (1995) Variation in sensitivity of Gaeumannomyces graminis to antibiotics produced by fluorescent Pseudomonas spp. and effect on biological control of take-all of wheat. Applied and Environmental Microbiology 61: 25542559.Google Scholar
Mazzola, M, Granatstein, DM, Elfving, DC & Mullinix, K (2001) Suppression of specific apple root pathogens by Brassica napus seed meal amendment regardless of glucosinolate content. Phytopathology 91: 673679.Google Scholar
Mazzola, M & Manici, LM (2012) Apple replant disease: role of microbial ecology in cause and control. Annual Review of Phytopathology 50: 4565.Google Scholar
Mazzola, M & Mullinix, K (2005) Comparative field efficacy of management strategies containing Brassica napus seed meal or green manure for the control of apple replant disease. Plant Disease 89: 12071213.Google Scholar
McBride, CS (2007) Rapid evolution of smell and taste receptor genes during host specialization in Drosophila seychellia. Proceedings of the National Academy of Sciences, USA 104: 49965001.Google Scholar
McCallum, H (1996) Immunocontraception for wildlife population control. Trends in Ecology & Evolution 11: 491493.Google Scholar
McClay, AS & Balciunas, JK (2005) The role of pre-release efficacy assessment in selecting classical biological control agents for weeds – applying the Anna Karenina principle. Biological Control 35: 197207.Google Scholar
McClure, M (1986) Population dynamics of Japanese hemlock scales: a comparison of endemic and exotic communities. Ecology 67: 14111421.Google Scholar
McConnachie, AJ, Hill, KM & Byrne, MJ (2004) Field assessment of a frond-feeding weevil, a successful biological control agent of red waterfern, Azolla filiculoides, in southern Africa. Biological Control 28: 2532.Google Scholar
McCullough, DG, Mercader, RJ & Siegert, NW (2015) Developing and integrating tactics to slow ash (Oleaceae) mortality caused by emerald ash borer (Coleoptera: Buprestidae). Canadian Entomologist 147: 349358.Google Scholar
McDonald, RC & Kok, LT (1991) Hyperparasites attacking Cotesia glomerata (L.) and Cotesia rubecula (Marshall) (Hymenoptera: Braconidae) in southwestern Virginia. Biological Control 1: 170175.Google Scholar
McEvoy, PB & Coombs, EM (1999) Biological control of plant invaders: regional patterns, field experiments, and structured population models. Ecological Applications 9: 387401.Google Scholar
McEvoy, PB, Higgs, KM, Coombs, EM, Karacetin, E & Starcevich, LA (2012) Evolving while invading: rapid adaptive evolution in juvenile development time for a biological control organism colonizing a high-elevation environment. Evolutionary Applications 5: 524536.Google Scholar
McEvoy, PB, Rudd, NT, Cox, CS & Huso, M (1993) Disturbance, competition, and herbivory effects on ragwort Senecio jacobaea populations. Ecological Monographs 63: 5575.Google Scholar
McFadden, G (2005) Poxvirus tropism. Nature Reviews Microbiology 3: 201213.Google Scholar
McFadyen, REC (1998) Biological control of weeds. Annual Review of Entomology 43: 369393.Google Scholar
McFadyen, REC (2000) Successes in biological control of weeds. In Proceedings of the X International Symposium on Biological Control of Weeds (Spencer, NR, ed.), Bozeman, Montana State University, pp. 314.Google Scholar
McFadyen, RE (2003) Does ecology help in the selection of biocontrol agents? In Improving the Selection, Testing and Evaluation of Weed Biological Control Agents (Spafford Jacob, H & Briese, DT, eds.), Glen Osmond, Australia, CRC, pp. 59.Google Scholar
McFadyen, R (2008) Return on investment: determining the economic impact of biological control programs. In Proceedings of the XII International Symposium on Biological Control of Weeds (Julien, MH, Sforza, R, Bon, MC, Evans, HC, Hatcher, PE, Hinz, HL & Rector, BG, eds.), Wallingford, UK, CABI Publishing, pp. 6774.Google Scholar
McFadyen, R & Spafford Jacob, H (2004) Insects for the biocontrol of weeds: predicting parasitism levels in the new country. In Proceedings of the XI International Symposium on Biological Control of Weeds (Cullen, JM, Briese, DT, Kriticos, DJ, Lonsdale, WM, Morin, L & Scott, JK, eds.), Canberra, Australia, CSIRO, pp. 135140.Google Scholar
McMurtry, JA (1991) Augmentative releases to control mites in agriculture. In Modern Acarology, Volume 1 (Dusbabeck, F & Bukva, V, eds.), The Hague, The Netherlands, SPB Academic, pp. 151157.Google Scholar
McMurtry, JA & Croft, BA (1997) Life-styles of phytoseiid mites and their roles in biological control. Annual Review of Entomology 42: 291321.Google Scholar
McMurtry, JA, Moraes, GJ & Famah Sourassou, N (2013) Revision of lifestyles of phytoseiid mites (Acari: Phytoseiidae) and implications for biological control strategies. Systematic and Applied Acarology 18: 297320.Google Scholar
McQuate, GT, Sylva, CD & Jang, EB (2005) Mediterranean fruit fly (Dipt., Tephritidae) suppression in persimmon through bait sprays in adjacent coffee plantings. Journal of Applied Entomology 129: 110117.Google Scholar
McSpadden Gardener, BB (2004) Ecology of Bacillus and Paenibacillus in agricultural systems. Phytopathology 94: 12521258.Google Scholar
McSpadden Gardener, BB (2007) Diversity and ecology of biocontrol Pseudomonas spp. in agricultural systems. Phytopathology 97: 221226.Google Scholar
Mead, AR (1979) Pulmonates Volume 2B, Economic Malacology with Particular Reference to Achatina fulica. London, UK, Academic Press.Google Scholar
Meffe, GK (1983) Attempted chemical renovation of an Arizona springbrook for management of the endangered Sonoran topminnow. North American Journal of Fisheries Management 3: 315321.Google Scholar
Meffe, GK (1984) Effects of abiotic disturbance on coexistence of predator–prey fish species. Ecology 65: 15251534.Google Scholar
Meffe, GK (1985) Predation and species replacement in American southwestern fishes: a case study. The Southwestern Naturalist 30: 173187.Google Scholar
Meffe, GK, Hendrickson, DA, Minckley, WL & Rinne, JN (1983) Factors resulting in decline of the endangered Sonoran topminnow Poeciliopsis occidentalis (Atheriniformes, Poeciliidae) in the United-States. Biological Conservation 25: 135159.Google Scholar
Memmott, J, Craze, PG, Harman, HM, Syrett, P & Fowler, SV (2005) The effect of propagule size on the invasion of an alien insect. Journal of Animal Ecology 74: 5062.Google Scholar
Memmott, J, Fowler, SV & Hill, RL (1998) The effect of the release size on the probability of establishment of biological control agents: gorse thrips (Sericothrips staphylinus) released against gorse (Ulex europaeus) in New Zealand. Biocontrol Science and Technology 8: 103115.Google Scholar
Mensah, RK (1999) Habitat diversity: implications for the conservation and use of predatory insects of Helicoverpa spp. in cotton systems in Australia. International Journal of Pest Management 45: 91100.CrossRefGoogle Scholar
Mensah, RK & Madden, JL (1993) Development and application of an integrated pest management program for the psyllid, Ctenarytaina thysanura on Boronia megastigma in Tasmania. Entomologia Experimentalis et Applicata 66: 5974.Google Scholar
Mensah, RK & Sequeira, RV (2004) Habitat manipulation for insect pest management in cotton cropping systems. In Ecological Engineering for Pest Management (Gurr, GM, Wratten, SD & Altieri, MA, eds.), Ithaca, NY, Comstock, pp. 187198.Google Scholar
Mercado-Blanco, J. 2015. Pseudomonas strains that exert biocontrol of plant pathogens. In Pseudomonas, Volume 7, New Aspects of Pseudomonas Biology (Ramos, J-L, Goldberg, JB & Filloux, A, eds.), Dordrecht, The Netherlands, Springer, pp. 121172.Google Scholar
Messelink, GJ, Bloemhard, CMJ, Cortes, JA, Sabelis, MW & Janssen, A (2011) Hyperpredation by generalist predatory mites disrupts biological control of aphids by the aphidophagous gall midge Aphidoletes aphidimyza. Biological Control 57: 246252.Google Scholar
Messelink, GJ & Janssen, A (2014) Increased control of thrips and aphids in greenhouses with two species of generalist predatory bugs involved in intraguild predation. Biological Control 79: 17.Google Scholar
Messenger, PS (1970) Bioclimatic inputs to biological control and pest management programs. In Concepts of Pest Management (Rabb, RL & Guthrie, FE, eds.), Raleigh, University of North Carolina Press.Google Scholar
Messenger, PS, Biliotti, E & Van den Bosch, R (1976a) The importance of natural enemies in integrated control. In Theory and Practice of Biological Control (Huffaker, CB & Messenger, PS, eds.), New York, Academic Press, pp. 543654.Google Scholar
Messenger, PS, Wilson, F & Whitten, MJ (1976b) Variation, fitness, and adaptability of natural enemies. In Theory and Practice in Biological Control (Huffaker, CB & Messenger, PS, eds.), New York, Academic Press, pp. 209232.Google Scholar
Messersmith, CG & Adkins, SW (1995) Integrating weed-feeding insects and herbicides for weed control. Weed Technology 9: 199208.Google Scholar
Messina, FJ (1981) Plant protection as a consequence of an ant-membracid mutualism – Interactions on Goldenrod (Solidago sp). Ecology 62: 14331440.Google Scholar
Messing, RH (2000) The impact of nontarget concerns on the practice of biological control. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.) Dordrecht, The Netherlands, Kluwer, pp. 4558.Google Scholar
Messing, RH (2001) Centrifugal phylogeny as a basis for non-target host testing in biological control: is it relevant for parasitoids? Phytoparasitica 29: 187190.Google Scholar
Messing, RH, Roitberg, BD & Brodeur, J (2006) Measuring and predicting indirect impacts of biological control: competition, displacement and secondary interactions. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 6477.Google Scholar
Messing, RH & Wright, MG (2006) Biological control of invasive species: solution or pollution? Frontiers in Ecology and the Environment 4: 132140.Google Scholar
Meyer, G, Clare, R & Weber, E (2005) An experimental test of the evolution of increased competitive ability hypothesis in goldenrod, Solidago gigantea. Oecologia 144: 299307.Google Scholar
Meyling, NV & Eilenberg, J (2007) Ecology of the entomopathogenic fungi Beauveria bassiana and Metarhizium anisopliae in temperate agroecosystems: potential for conservation biological control. Biological Control 43: 145155.Google Scholar
Meyling, NV & Hajek, AE (2010) Principles from community and metapopulation ecology: application to fungal entomopathogens. Biocontrol 55: 3954.Google Scholar
Meyling, NV, Lubeck, M, Buckley, EP, Eilenberg, J & Rehner, SA (2009) Community composition, host range and genetic structure of the fungal entomopathogen Beauveria in adjoining agricultural and seminatural habitats. Molecular Ecology 18: 12821293.Google Scholar
Michaud, JP (1999) Sources of mortality in colonies of brown citrus aphid, Toxoptera citricida. Biocontrol 44: 347367.Google Scholar
Michaud, JP (2002) Invasion of the Florida citrus ecosystem by Harmonia axyridis (Coleoptera: Coccinellidae) and asymmetric competition with a native species, Cycloneda sanguinea. Environmental Entomology 31: 827835.Google Scholar
Midgarden, D, Fleischer, SJ, Weisz, R & Smilowitz, Z (1997) Site-specific integrated pest management impact on development of esfenvalerate resistance in Colorado potato beetle (Coleoptera: Chrysomelidae)and on densities of natural enemies. Journal of Economic Entomology 90: 855867.Google Scholar
Milgroom, MG & Cortesi, P (2004) Biological control of chestnut blight with hypovirulence: a critical analysis. Annual Review of Phytopathology 42: 311338.Google Scholar
Millar, LC & Barbercheck, ME (2001) Interaction between endemic and introduced entomopathogenic nematodes in conventional-till and no-till corn. Biological Control 22: 235245.Google Scholar
Millard, WA & Taylor, CB (1927) Antagonism of micro-organisms as the controlling factor in the inhibition of scab by green-manuring. Annals of Applied Biology 14: 202216.Google Scholar
Miller, D (1936) Biological control of noxious weeds. New Zealand Journal of Science and Technology 18: 581584.Google Scholar
Miller, LK, Lingg, AJ & Bulla, LA (1983) Bacterial, viral and fungal insecticides. Science 219: 715721Google Scholar
Miller, TEX, Shaw, AK, Inouye, BD & Neubert, MG (2011) Sex-biased dispersal and the speed of two-sex invasions. American Naturalist 177: 549561.Google Scholar
Miller-Rushing, AJ, Hoye, TT, Inouye, DW & Post, E (2010) The effects of phenological mismatches on demography. Philosophical Transactions of the Royal Society B: Biological Sciences 365: 31773186.Google Scholar
Mills, LS, Doak, DF & Wisdom, MJ (1999) Reliability of conservation actions based on elasticity analysis of matrix models. Conservation Biology 13: 815829.Google Scholar
Mills, MD, Rader, RB & Belk, MC (2004) Complex interactions between native and invasive fish: the simultaneous effects of multiple negative interactions. Oecologia 141: 713721.Google Scholar
Mills, NJ (1982) Satiation and the functional response: a test of a new model. Ecological Entomology 7: 305315.Google Scholar
Mills, NJ (1992) Parasitoid guilds, life-styles, and host ranges in the parasitoid complexes of torticoid hosts (Lepidoptera: Torticoidea). Environmental Entomology 21: 230239.Google Scholar
Mills, NJ (1994) Biological control: some emerging trends. In Individuals, Populations and Patterns in Ecology (Leather, SR, Watt, AD, Mills, NJ & Walters, KFA, eds.), Andover, UK, Intercept, pp. 213222.Google Scholar
Mills, NJ (1998) Trichogramma: the field efficacy of inundative biological control of the codling moth in Californian orchards. In Proceedings of the 1st California Conference on Biological Control (Hoddle, MS, ed.), Berkeley, University of California, pp. 1011.Google Scholar
Mills, NJ (2000) Biological control: the need for realistic models and experimental approaches to parasitoid introductions. In Parasitoid Population Biology (Hochberg, ME & Ives, AR, eds.), Princeton, NJ, Princeton University Press, pp. 217234.Google Scholar
Mills, NJ (2003) Augmentation in orchards: improving the efficacy of Trichogramma inundation. In Proceedings of 1st International Symposium on Biological Control of Arthropods (Van Driesche, RG, ed.), Honolulu, USDA Forest Service, pp. 130135.Google Scholar
Mills, NJ (2005) Selecting effective parasitoids for biological control introductions: codling moth as a case study. Biological Control 34: 274282.Google Scholar
Mills, NJ (2006a) Accounting for differential success in the biological control of homopteran and lepidopteran pests. New Zealand Journal of Ecology 30: 6172.Google Scholar
Mills, NJ (2006b) Interspecific competition among natural enemies and single versus multiple introductions in biological control. In Trophic and Guild Interactions in Biological Control (Brodeur, J & Boivin, G, eds.), Dordrecht, The Netherlands, Springer, pp. 191220.Google Scholar
Mills, NJ (2008) Can matrix models guide the selection of parasitoids for biological control introductions? LBAM in California as a case study. In Proceedings of the 3rd International Symposium on Biological Control of Arthropods (Mason, PG, Gillespie, DR & Vincent, C, eds.), Christchurch, New Zealand, USDA Forest Service, pp. 8997.Google Scholar
Mills, NJ (2010) Egg parasitoids in biological control and integrated pest management. In Egg Parasitoids in Agroecosystems with Emphasis on Trichogramma (Consoli, FL, Parra, JRP & Zucchi, RA, eds.), Dordrecht, The Netherlands, Springer, pp. 389411.Google Scholar
Mills, NJ, Babendreier, D & Loomans, AJM (2006) Methods for monitoring the dispersal of natural enemies from point source releases associated with augmentative biological control. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 114131.Google Scholar
Mills, NJ, Beers, EH, Shearer, PW, Unruh, TR & Amarasekare, KG (2016) Comparative analysis of pesticide effects on natural enemies in western orchards: a synthesis of laboratory bioassay data. Biological Control 102: 1725.Google Scholar
Mills, NJ & Getz, WM (1996) Modelling the biological control of insect pests: a review of host-parasitoid models. Ecological Modelling 92: 121143.Google Scholar
Mills, NJ & Gutierrez, AP (1996) Prospective modelling in biological control: an analysis of the dynamics of heteronomous hyperparasitism in a cotton-whitefly-parasitoid system. Journal of Applied Ecology 33: 13791394.Google Scholar
Mills, NJ & Gutierrez, AP (1999) Biological control of insect pests: a tritrophic perspective. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 89102.Google Scholar
Mills, NJ & Lacan, I (2004) Ratio dependence in the functional response of insect parasitoids: evidence from Trichogramma minutum foraging for eggs in small host patches. Ecological Entomology 29: 208216.Google Scholar
Mills, NJ, Pickel, C, Mansfield, S, McDougall, S, Buchner, R, Caprile, J, Edstrom, J, Elkins, R, Hasey, J, Kelley, K, Krueger, W, Olson, W & Stocker, R (2000) Mass releases of Trichogramma wasps can reduce damage from codling moth. California Agriculture 54: 2225.Google Scholar
Mills, NJ & Wajnberg, E (2008) Optimal foraging behavior and efficient biological control methods. In Behavioral Ecology of Insect Parasitoids (Wajnberg, E, Bernstein, C & Van Alphen, JJM, eds.), Malden, MA, Blackwell Publishing, pp. 330.Google Scholar
Milner, RJ & Hunter, DM (2001) Recent developments in the use of fungi as biopesticides against locusts and grasshoppers in Australia. Journal of Orthoptera Research 10: 271276.Google Scholar
Minckley, WL (1999) Ecological review and management recommendations for recovery of the endangered Gila topminnow. Great Basin Naturalist 59: 230244.Google Scholar
Minckley, WL & Deacon, JE (1968) Southwestern fishes and the enigma or ‘endangered species’. Science 159: 14241432.Google Scholar
Minkenberg, OPJM, Tatar, M & Rosenheim, JA (1992) Egg load as a major source of variability in insect foraging and oviposition behavior. Oikos 65: 134142.Google Scholar
Mitchell, AJ & Kelly, AM (2006) The public sector role in establishment of grass carp in the United States. Fisheries 31: 113122.Google Scholar
Mitchell, CE & Power, AG (2003) Release of invasive plants from fungal and viral pathogens. Nature 421: 625627.Google Scholar
Mkoji, GMB, Hofkin, BV, Kuris, AM, Stewart-Oaten, A, Mungai, BN, Kihara, JH, Mungai, F, Yundu, J, Mbui, J, Rashid, JR, Kariuki, CH, Ouma, JH, Koech, DK & Loker, ES (1999) Impact of the crayfish Procambarus clarkii on Schistosoma haematobium transmission in Kenya. American Journal of Tropical Medicine and Hygiene 61: 751759.Google Scholar
Mochiah, MB, Ngi-Song, AJ, Overholt, WA & Stouthamer, R (2002) Wolbachia infection in Cotesia sesamiae (Hymenoptera: Braconidae) causes cytoplasmic incompatability: implications for biological control. Biological Control 25: 7480.Google Scholar
Moeed, A, Hickson, R & Barratt, BIP (2006) Principles of environmental risk assessment with emphasis on the New Zealand perspective. In Environmental Impact of Invertebrates for Biological Control of Arthropods (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 241253.Google Scholar
Mols, CMM & Visser, ME (2007) Great tits (Parus major) reduce caterpillar damage in commercial apple orchards. PLoS One 2: e202.Google Scholar
Momanyi, C, Lohr, B & Gitonga, L (2006) Biological impact of the exotic parasitoid, Diadegma semiclausum (Hellen), of diamondback moth, Plutella xylostella L., in Kenya. Biological Control 38: 254263.Google Scholar
Moon, RD (1980) Biological control through interspecific competition. Environmental Entomology 9: 723728Google Scholar
Moore, D, Douro-Kpindou, O-K, Jenkins, NE & Lomer, CJ (1996) Effects of moisture content and temperature on storage of Metarhizium flavoviride conidia. Biocontrol Science and Technology 6: 5161.Google Scholar
Moore, NW (1987) The Bird of Time. Cambridge, UK, Cambridge University Press.Google Scholar
Morales Ramos, JA, Summy, KR & King, EG (1996) ARCASIM, a model to evaluate augmentation strategies of the parasitoid Catolaccus grandis against boll weevil populations. Ecological Modelling 93: 221235.Google Scholar
Moran, VC, Hoffmann, JH & Hill, MP (2011) A context for the 2011 compilation of reviews on the biological control of invasive alien plants in South Africa. African Entomology 19: 177185.Google Scholar
Moreau, G, Eveleigh, ES, Lucarotti, CJ & Quiring, DT (2006) Stage-specific responses to ecosystem alteration in an eruptive herbivorous insect. Journal of Applied Ecology 43: 2834.Google Scholar
Morehead, SA & Feener, DH Jr. (2000) An experimental test of potential host range in the ant parasitoid Apocephalus paraponerae. Ecological Entomology 25: 332340.Google Scholar
Morin, L, Evans, KJ & Sheppard, AW (2006) Selection of pathogen agents in weed biological control: critical issues and peculiarities in relation to arthropod agents. Australian Journal of Entomology 45: 349365.Google Scholar
Morin, L, Reid, AM, Sims-Chilton, NM, Buckley, YM, Dhileepan, K, Hastwell, GT, Nordblom, TL & Raghu, S (2009) Review of approaches to evaluate the effectiveness of weed biological control agents. Biological Control 51: 115.Google Scholar
Morse, JG (1998) Agricultural implications of pesticide-induced hormesis of insects and mites. Human and Experimental Toxicology 17: 266269.Google Scholar
Mossman, K, Lee, SF, Barry, M, Boshkov, L & McFadden, G (1996) Disruption of M-T5, a novel myxoma virus gene member of the poxvirus host range superfamily, results in dramatic attenuation of myxomatosis in infected European rabbits. Journal of Virology 70: 43944410.Google Scholar
Mouquet, N, Thomas, JA, Elmes, GW, Clarke, RT & Hochberg, ME (2005) Population dynamics and conservation of a specialized predator: a case study of Maculinea arion. Ecological Monographs 75: 525542.Google Scholar
Muesebeck, CFW & Dohanian, SM (1927) A study in hyperparasitism, with particular reference to the parasites of Apanteles melanoscelus (Ratzeburg). USDA Bulletin 1487.Google Scholar
Mukherjee, A, Christman, MC, Overholt, WA & Cuda, JP (2011) Prioritizing areas in the native range of Hygrophila for surveys to collect biological control agents. Biological Control 56: 254262.Google Scholar
Muldrew, JA (1953) The natural immunity of the larch sawfly (Pristiphora erichsonii (Htg.)) to the introduced parasite Mesoleius tenthredinis Morley, in Manitoba and Saskatchewan. Canadian Journal of Zoology 31: 313332.Google Scholar
Müller, CB, Adriaanse, ICT, Belshaw, R & Godfray, HCJ (1999) The structure of an aphid-parasitoid community. Journal of Animal Ecology 68: 346370.Google Scholar
Müller-Schärer, H (1991) The impact of root herbivory as a function of plant density and competition: survival, growth and fecundity of Centaurea maculosa in field plots. Journal of Applied Ecology 28: 759776.Google Scholar
Müller-Schärer, H & Schaffner, U (2008) Classical biological control: exploiting enemy escape to manage plant invasions. Biological Invasions 10: 859874.Google Scholar
Mumm, R & Dicke, M (2010) Variation in natural plant products and the attraction of bodyguards involved in indirect plant defense. Canadian Journal of Zoology 88: 628667.Google Scholar
Munro, VMW & Henderson, IM (2002) Nontarget effects of entomophagous biocontrol: shared parasitism between native lepidopteran parasitism and the biocontrol agent Trigonospila brevifacies (Diptera: Tachinidae) in forest habitats. Environmental Entomology 31: 388396.Google Scholar
Münster-Swendsen, M & Nachman, G (1978) Asynchrony in insect host–parasite interaction and its effect on stability, studied by a simulation model. Journal of Animal Ecology 47: 159171.Google Scholar
Murdoch, WW (1990) The relevance of pest-enemy models to biological control. In Critical Issues in Biological Control (Mackauer, M, Ehler, LE & Roland, J, eds.), Andover, UK, Intercept, pp. 124.Google Scholar
Murdoch, WW (2009) Biological control: theory and practice. In The Princeton Guide to Ecology (Levin, SA, ed.), Princeton, NJ, Princeton University Press, pp. 683688.Google Scholar
Murdoch, WW & Bence, J (1987) General predators and unstable prey populations. In Predation: Direct and Indirect Impacts on Aquatic Communities (Kerfoot, WC & Sih, A, eds.), Hanover, NH, University Press of New England, pp. 1730.Google Scholar
Murdoch, WW & Briggs, CJ (1996) Theory for biological control: recent developments. Ecology 77: 20012013.Google Scholar
Murdoch, WW, Briggs, CJ & Nisbet, RM (1996a) Competitive displacement and biological control in parasitoids: a model. American Naturalist 148: 807826.Google Scholar
Murdoch, WW, Briggs, CJ & Nisbet, (2003) Consumer-Resource Dynamics. Princeton, NJ, Princeton University Press.Google Scholar
Murdoch, WW, Briggs, CJ & Swarbrick, S (2005) Host suppression and stability in a parasitoid-host system: experimental demonstration. Science 309: 610613.Google Scholar
Murdoch, WW, Chesson, J & Chesson, PL (1985) Biological control in theory and practice. American Naturalist 125: 344366.Google Scholar
Murdoch, WW, Evans, FC & Peterson, CH (1972) Diversity and pattern in plants and insects. Ecology 53: 819829.Google Scholar
Murdoch, WW, Luck, RF, Swarbrick, SL, Walde, S, Yu, DS & Reeve, JD (1995) Regulation of an insect population under biological control. Ecology 76:206217.Google Scholar
Murdoch, WW, Luck, RF, Walde, SJ, Reeve, JD & Yu, DS (1989) A refuge for red scale under control by Aphytis: structural aspects. Ecology 70: 17071714.Google Scholar
Murdoch, WW, Nisbet, RM, Blythe, SP, Gurney, WSC & Reeve, JD (1987) An invulnerable age class and stability in delay-differential parasitoid-host models. American Naturalist 129: 263282.Google Scholar
Murdoch, WW, Nisbet, RM, Luck, RF, Godfray, HCJ & Gurney, WSC (1992) Size-selective sex-allocation and host feeding in a parasitoid-host model. Journal of Animal Ecology 61: 533541.Google Scholar
Murdoch, WW & Oaten, A (1975) Predation and population stability. Advances in Ecological Research 9: 1125Google Scholar
Murdoch, WW & Oaten, A (1989) Aggregation by parasitoids and predators: effects on equilibrium and stability. American Naturalist 123: 371392.Google Scholar
Murdoch, WW, Swarbrick, SL & Briggs, CJ (2006b) Biological control: lessons from a study of California red scale. Population Ecology 48: 297305.Google Scholar
Murdoch, WW, Swarbrick, SL, Luck, RF, Walde, S & Yu, DS (1996b) Refuge dynamics and metapopulation dynamics: an experimental test. American Naturalist 147: 424444.Google Scholar
Murphy, SM (2004) Enemy-free space maintains swallowtail butterfly host shift. Proceedings of the National Academy of Sciences, USA 101: 1804818052.Google Scholar
Murphy, BC, Rosenheim, JA, Dowell, RV & Granett, J (1998) Habitat diversification for improving biological control: parasitism of the western grape leafhopper. Entomologia Experimentalis et Applicata 87: 225235.Google Scholar
Murray, E (1993) The sinister snail. Endeavour 17: 7883.Google Scholar
Murray, J, Murray, E, Johnson, MS & Clarke, B (1988) The extinction of Partula on Moorea. Pacific Science 42: 150153.Google Scholar
Murray, TJ, Withers, TM & Mansfield, S (2010) Choice versus no-choice test interpretation and the role of biology and behavior in parasitoid host specificity tests. Biological Control 52: 153159.Google Scholar
Myers, JH (1985) How many insect species are necessary for successful biocontrol of weeds? Proceedings of the 6th International Symposium on the Biological Control of Weeds (Delfosse, ES, ed.), Ottawa, Canada, Agriculture Canada, pp. 7782.Google Scholar
Myers, JH & Bazely, DR (2003) Ecology and Control of Introduced Plants. Cambridge, UK, Cambridge University Press.Google Scholar
Myers, JH & Harris, P (1980) Distribution of Urophora galls in flower heads of diffuse and spotted knapweed in British Columbia. Journal of Applied Ecology 17: 359367.Google Scholar
Myers, JH, Higgins, C & Kovacs, E (1989) How many insect species are necessary for the biological control of insects? Environmental Entomology 18: 541547.Google Scholar
Myers, JH, Risley, C & Eng, R (1989) The ability of plants to compensate for insect attack: why biological control of weeds with insects is so difficult. In Proceedings of the 7th International Symposium on the Biological Control of Weeds (Delfosse, ES, ed.), Rome, Italy, Istituto Sperimentale per la Patalogia Vegetale, pp. 6773.Google Scholar
Myers, JH & Sabath, MD (1981) Genetic and phenotypic variability, genetic variance, and the success of establishment of insect introductions for the biological control of weeds. In Proceedings of the V International Symposium on Biological Control of Weeds (Delfosse, ES, ed.), Brisbane, Australia, Commonwealth Scientific and Industrial Research Organization, pp. 91102.Google Scholar
Myers, K, Marshall, ID & Fenner, F (1954) Studies in the epidemiology of infectious myxomatosis in rabbits. Journal of Hygiene 52: 337360.Google Scholar
Naeem, S & Li, SB (1997) Biodiversity enhances ecosystem reliability. Nature 390: 507509.Google Scholar
Nachman, G (1987) Systems analysis of acarine predator–prey interactions. I. A stochastic simulation model of spatial processes. Journal of Animal Ecology 56: 247265.Google Scholar
Nachman, G & Zemek, R (2003) Interactions in a tritrophic acarine predator–prey metapopulation system V: within-plant dynamics of Phytoseiulus persimilis and Tetranychus urticae (Acari: Phytoseiidae, Tetranychidae). Experimental and Applied Acarology 29:3568.Google Scholar
Nagayama, K, Watanabe, S, Kumakura, K, Ichikawa, T & Makino, T (2007) Development and commercialization of Trichoderma asperellum SKT-1 (Ecohope (R)), a microbial pesticide. Journal of Pesticide Science 32: 141142.Google Scholar
Naik, PR, Raman, G, Narayanan, KB & Sakthivel, N (2008) Assessment of genetic and functional diversity of phosphate solubilizing fluorescent pseudomonads isolated from rhizospheric soil. BMC Microbiology 8: 230Google Scholar
Naka, H, Mitsunaga, T & Mochizuki, A (2005) Laboratory hybridization between the introduced and the indigenous green lacewings (Neuroptera: Chrysopidae: Chrysoperla) in Japan. Environmental Entomology 34: 727731.Google Scholar
Nakasuji, F, Yamanaka, H & Kiritani, K (1973) The disturbing effect of mycriphantid spiders on the larval aggregation of the tobacco cutworm, Spodoptera litura (Lepidoptera: Noctuidae). Kontyu 41: 220227.Google Scholar
Nakayama, S, Seko, T, Takatsuki, J-I, Miura, K & Miyatake, T (2010) Walking activity of flightless Harmonia axyridis (Coleoptera: Coccinellidae) as a biological control agent. Journal of Economic Entomology 103: 15641568.Google Scholar
Naranjo, SE (2001) Conservation and evaluation of natural enemies in IPM systems for Bemisia tabaci. Crop Protection 20: 835852.Google Scholar
Naranjo, SE & Ellsworth, PC (2009) The contribution of conservation biological control to integrated control of Bemisia tabaci in cotton. Biological Control 51: 458470.Google Scholar
Naranjo, SE, Ellsworth, PC & Frisvold, GB (2015) Economic value of biological control in integrated pest management of managed plant systems. Annual Review of Entomology 60: 621645.Google Scholar
Navarro-Llopis, V, Ayala, I, Sanchis, J, Primo, J & Moya, P (2015) Field efficacy of a Metarhizium anisopliae-based attractant-contaminant device to control Ceratitis capitata (Diptera: Tephritidae). Journal of Economic Entomology 108: 15701578.Google Scholar
Nechols, JR, Obrycki, JJ, Tauber, CA & Tauber, MJ (1996) Potential impact of native natural enemies on Galerucella spp (Coleoptera: Chrysomelidae) imported for biological control of purple loosestrife: a field evaluation. Biological Control 7: 6066.Google Scholar
Neeno-Eckwall, EC, Kinkel, LL & Schottel, JL (2001) Competition and antibiosis in the biological control of potato scab. Canadian Journal of Microbiology 47: 332340.Google Scholar
Nei, M, Taruyama, T & Chakraborty, R (1975) The bottleneck effect and genetic variability in populations. Evolution 29: 110.Google Scholar
Nelson, EB & Hoitink, HAJ (1983) The role of microorganisms in the suppression of Rhizoctonia solani in container media amended with composted hardwood bark. Phytopathology 73: 274278.Google Scholar
Nelson, EH, Matthews, CE & Rosenheim, JA (2004) Predators reduce prey population growth by inducing changes in prey behavior. Ecology 85: 18531858.Google Scholar
Nentwig, W, Frank, T & Lethmayer, C (1998) Sown weed strips: artificial ecological compensation areas as an important tool in conservation biological control. In Conservation Biological Control (Barbosa, P, ed.), San Diego, CA, Academic Press, pp. 133154.Google Scholar
Neuenschwander, P (2001) Biological control of the cassava mealybug in Africa: a review. Biological Control 21: 214229.Google Scholar
Neuenschwander, P, Schulthess, F & Madojemu, E (1986) Experimental evaluation of the efficiency of Epidinocarsis lopezi, a parasitoid introduced into Africa against the cassava mealybug Phenacoccus manihoti. Entomologia Experimentalis et Applicata 42: 133138.Google Scholar
Neuhauser, C, Andow, DA, Heimpel, GE, May, G, Shaw, RG & Wagenius, S (2003) Community genetics: expanding the synthesis of ecology and genetics. Ecology 84: 545558.Google Scholar
Newman, RM (2004) Biological control of Eurasian watermilfoil by aquatic insects: basic insights from an applied problem. Archiv Fur Hydrobiologie 159: 145184.Google Scholar
Newman, RM, Borman, ME & Castro, SW (1997) Developmental performance of the weevil Euhrychiopsis lecontei on native and exotic watermilfoil host plants. Journal of the American Benthological Society 16: 627634.Google Scholar
Newman, RM, Thompson, DC & Richman, DB (1998) Conservation strategies for the biological control of weeds. In Conservation Biological Control (Barbosa, P, ed.), New York, Academic Press, pp. 371396.Google Scholar
Newsom, LD, Smith, RF & Whitcomb, WH (1976) Selective pesticides and selective use of pesticides. In Theory and Practices of Biological Control (Huffaker, CB & Messenger, PS, eds.), New York, Academic Press, pp. 565692.Google Scholar
Nicholson, AJ (1933) The balance of animal populations. Journal of Animal Ecology 2: 132178.Google Scholar
Nicholson, AJ & Bailey, VA (1935) The balance of animal populations. Part 1. Proceedings of the Zoological Society of London 3: 551598.Google Scholar
Nickerson, JC, Rolphkay, CA, Buschman, LL & Whitcomb, WH (1977) Presence of Spissistilus festinus (Homoptera: Membracidae) as a factor affecting egg predation by ants (Hymenoptera: Formicidae) in soybeans. Florida Entomologist 60: 193199.Google Scholar
Nielsen, C, Keena, M & Hajek, AE (2005) Virulence and fitness of the fungal pathogen Entomophaga maimaiga in its host Lymantria dispar, for pathogen and host strains originating from Asia, Europe, and North America. Journal of Invertebrate Pathology 89: 232242.Google Scholar
Nilsson, U, Rannback, LM, Anderson, P, Eriksson, A & Ramert, B (2011) Comparison of nectar use and preference in the parasitoid Trybliographa rapae (Hymenoptera: Figitidae) and its host, the cabbage root fly, Delia radicum (Diptera: Anthomyiidae). Biocontrol Science and Technology 21: 11171132.Google Scholar
Nimkingrat, P, Khanam, S, Strauch, O & Ehlers, R-U (2013) Hybridisation and selective breeding for improvement of low temperature activity of the entomopathogenic nematode Steinernema feltiae. BioControl 58: 417426.Google Scholar
Nisbet, RM, Diehl, S, Wilson, WG, Cooper, SD, Donaldson, DD & Kratz, K (1997) Primary-productivity gradients and short-term population dynamics in open systems. Ecological Monographs 67: 535553.Google Scholar
Niyibigira, EI, Overholt, WA & Stouthamer, R (2004a) Cotesia flavipes Cameron and Cotesia sesamiae (Cameron) (Hymenoptera: Braconidae) do not exhibit complementary sex determination: evidence from field populations. Applied Entomology and Zoology, 39: 705715.Google Scholar
Niyibigira, EI, Overholt, WA & Stouthamer, R (2004b) Cotesia flavipes Cameron (Hymenoptera: Braconidae) does not exhibit complementary sex determination (ii) evidence from laboratory experiments. Applied Entomology and Zoology 39: 717725.Google Scholar
Noonburg, EG & Byers, JE (2005) More harm than good: when invader vulnerability to predators enhances impact on native species. Ecology 86: 25552560.Google Scholar
Nordblom, TL, Smyth, MJ, Swirepik, A, Sheppard, AW & Briese, DT (2002) Spatial economics of biological control: investing in new releases of insects for earlier limitation of Patterson’s curse in Australia. Agricultural Economics 27: 403424.Google Scholar
Norris, RF (1997) Impact of leaf mining on the growth of Portulaca oleracea (common purslane) and its competitive interaction with Beta vulgaris (sugarbeet). Journal of Applied Ecology 34: 349362.Google Scholar
Novotny, V, Miller, SE, Cizek, L, Leps, J, Janda, M, Basset, Y, Weiblen, GD & Darrow, K (2003) Colonising aliens: caterpillars (Lepidoptera) feeding on Piper aduncum and P-umbellatum in rainforests of Papua New Guinea. Ecological Entomology 28: 704716.Google Scholar
NRC (2002) Predicting Invasions of Nonindigenous Plants and Plant Pests. Washington, DC, National Academy of Sciences.Google Scholar
Nuñez, MA, Simberloff, D & Relva, MA (2008) Seed predation as a barrier to alien conifer invasions. Biological Invasions 10: 13891398.Google Scholar
Nunney, L (2003) Managing captive populations for release: a population-genetic perspective. In Quality Control and Production of Biological Control Agents: Theory and Testing Procedures (Van Lenteren, JC, ed.), Wallingford, UK, CABI Publishing, pp. 7387.Google Scholar
Oaten, A & Murdoch, WW (1975) Functional response and stability in predator–prey systems. American Naturalist 109: 289298.Google Scholar
Obrycki, JJ (2000) Coccinellid introductions: potential for and evaluation of nontarget effects. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 127146.Google Scholar
Obrycki, JJ, Giles, KL & Ormord, AM (1998) Interactions between an introduced and indigenous coccinellid species at different prey densities. Oecologia 117: 279285.Google Scholar
Ochanda, J (2010) Potential areas of collaboration and communication strategies between researchers and potential GMO vector user communities. In Progress and Prospects for the Use of Genetically Modified Mosquitoes to Inhibit Disease Transmission (Crawford, VL & Reza, JN, eds.), Geneva, Switzerland, WHO, pp. 3739.Google Scholar
Ode, PJ, Antolin, MF & Strand, MR (1995) Brood-mate avoidance in the parasitic wasp Bracon hebetor say. Animal Behaviour 49: 12391248.Google Scholar
Ode, PJ, Antolin, MF & Strand, MR (1998) Differential dispersal and female-biased sex allocation in a parasitic wasp. Ecological Entomology 23: 314318.Google Scholar
Oelrichs, PB, MacLeod, JK, Seawright, AA & Grace, PB (2001) Isolation and identification of the toxic peptides from Lophyrotoma zonalis (Pergidae) sawfly larvae. Toxicon 39: 19331936.Google Scholar
Offenberg, J (2011) Oecophylla smaragdina food conversion efficiency: prospects for ant farming. Journal of Applied Entomology 135: 575581.Google Scholar
Olden, JD, Kennard, MJ & Pusey, BJ (2008) Species invasions and the changing biogeography of Australian freshwater fishes. Global Ecology and Biogeography 17: 2537.Google Scholar
Olivain, C, Humbert, C, Nahalkova, J, Fatehi, J, L’Haridon, F & Alabouvette, C (2006) Colonization of tomato root by pathogenic and nonpathogenic Fusarium oxysporum strains inoculated together and separately into the soil. Applied and Environmental Microbiology 72: 15231531.Google Scholar
Oliver, KM, Campos, J, Moran, NA & Hunter, MS (2008) Population dynamics of defensive symbionts in aphids. Proceedings of the Royal Society B 275: 293299.Google Scholar
Oliver, KM, Moran, NA & Hunter, MS (2005) Variation in resistance to parasitism in aphids is due to symbionts not host genotype. Proceedings of the National Academy of Sciences, USA 102: 1279512800.Google Scholar
Oliver, KM, Moran, NA & Hunter, MS (2006) Costs and benefits of a superinfection of facultative symbionts in aphids. Proceedings of the Royal Society of London, B 273: 12731280.Google Scholar
Oliver, KM, Russell, JA, Moran, NA & Hunter, MS (2003) Facultative bacterial symbionts in aphids confer resistance to parasitic wasps. Proceedings of the National Academy of Sciences, USA 100: 18031807.Google Scholar
Oliver, KM, Smith, AH & Russell, JA (2014) Defensive symbiosis in the real world – advancing ecological studies of heritable, protective bacteria and beyond. Functional Ecology 28: 341355.Google Scholar
Olkowski, W & Zhang, A (1998) Habitat management for biological control, examples from China. In Enhancing Biological Control (Pickett, CH & Bugg, RL, eds.), Berkeley, University of California Press, pp. 255270.Google Scholar
O’Neil, RJ, Giles, KL, Obrycki, JJ, Mahr, DL, Legaspi, JC & Katovich, K (1998) Evaluation of the quality of four commercially available natural enemies. Biological Control 11: 18.Google Scholar
O’Neil, RJ, Nagarajan, K, Wiedenmann, RN & Legaspi, JC (1996) A simulation model of Podisus maculiventris (Say) (Heteroptera: Pentatomidae) and Mexican bean beetle, Epilachna varivestis (Mulsant) (Coleoptera: Coccinellidae), population dynamics in soybean, Glycine max (L). Biological Control 6: 330339.Google Scholar
Onstad, DW, Liu, XX, Chen, M, Roush, R & Shelton, AM (2013) Modeling the integration of parasitoid, insecticide, and transgenic insecticidal crop for the long-term control of an insect pest. Journal of Economic Entomology 106: 11031111.Google Scholar
Onstad, DW & McManus, DP (1996) Risks of host range expansion by parasites of insects. BioScience 46: 430435.Google Scholar
Opit, GP, Nechols, JR & Margolies, DC (2004) Biological control of twospotted spider mites, Tetranychus urticae Koch (Acari: Tetranychidae), using Phytoseiulus persimilis Athias-Henriot (Acari: Phytoseidae) on ivy geranium: assessment of predator release ratios. Biological Control 29: 445452.Google Scholar
Opit, GP, Nechols, JR, Margolies, DC & Williams, KA (2005) Survival, horizontal distribution, and economics of releasing predatory mites (Acari: Phytoseiidae) using mechanical blowers. Biological Control 33: 344351.Google Scholar
Oppenheim, SJ, Gould, F & Hopper, KR (2012) The genetic architecture of a complex ecological trait: host plant use in the specialist moth, Heliothis subflexa. Evolution 66: 33363351.Google Scholar
Orr, DB & Suh, CP-C (2000) Parasitoids and predators. In Biological and Biotechnological Control of Insect Pests (Rechcigl, JE & Rechcigl, NA, eds.), Boca Raton, FL, CRC Press, pp. 334.Google Scholar
Orr, DB, Garcia-Salazar, C & Landis, DA (2000) Trichogramma nontarget impacts: a method for biological control risk assessment. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 111126.Google Scholar
Orr, DB, Landis, DA, Mutch, DR, Manley, GV, Stuby, SA & King, RL (1997) Ground cover influence on microclimate and Trichogramma (Hymenoptera: Trichogramma tidae) augmentation in seed corn production. Environmental Entomology 26: 433438.Google Scholar
Orrock, JL, Preisser, EL, Grabowski, JH & Trussel, GC (2013) The cost of safety: refuges increase the impact of predation risk in aquatic systems. Ecology 94: 573579.Google Scholar
Ortega, YK & Pearson, DE (2005) Weak vs. strong invaders of natural plant communities: Assessing invasibility and impact. Ecological Applications 15: 651661.Google Scholar
Ortega, YK, Pearson, DE & McKelvey, KS (2004) Effects of biological control agents and exotic plant invasion on deer mouse populations. Ecological Applications 14: 241253.Google Scholar
Ortega, YK, Pearson, DE, Waller, LP, Sturdevant, NJ & Maron, JL (2012) Population-level compensation impedes biological control of an invasive forb and indirect release of a native grass. Ecology 93: 783792.Google Scholar
Ouedraoga, RM, Goettel, MS & Brodeur, J (2004) Behavioral thermoregulation in the migratory locust: a therapy to overcome fungal infection. Oecologia 138: 312319.Google Scholar
Pacala, SW, Hassell, MP & May, RM (1990) Host-parasitoid associations in patchy environments. Nature 344: 150153.Google Scholar
Pachepsky, E, Nisbet, RM & Murdoch, WW (2008) Between discrete and continuous: consumer-resource dynamics with synchronized reproduction. Ecology 89: 280288.Google Scholar
Panaccione, DG, Beaulieu, WT & Cook, D (2014) Bioactive alkaloids in vertically transmitted fungal endophytes. Functional Ecology 28: 299314.Google Scholar
Paraiso, O, Hight, SD, Kairo, MTK & Bloem, S (2013a) Host specificity and risk assessment of Trichogramma fuentesi (Hymenoptera: Trichogrammatidae), a potential biological control agent of Cactoblastis cactorum (Lepidoptera: Pyralidae). Florida Entomologist 96: 13051310.Google Scholar
Paraiso, O, Kairo, MTK, Hight, SD, Leppla, NC, Cuda, JP, Owens, M & Olexa, MT (2013b) Opportunities for improving risk communication during the permitting process for entomophagous biological control agents: a review of current systems. BioControl 58: 115.Google Scholar
Parker, IM (2000) Invasion dynamics of Cytisus scoparius: a matrix model approach. Ecological Applications 10: 726743.Google Scholar
Parker, JD, Burkepile, DE & Hay, ME (2006) Opposing effects of native and exotic herbivores on plant invasions. Science 311: 14591461.Google Scholar
Parker, JD & Hay, ME (2005) Biotic resistance to plant invasions? Native herbivores prefer non-native plants. Ecology Letters 8: 959967.Google Scholar
Parker, IM, Simberloff, D, Lonsdale, WM, Goodell, K, Wonham, M, Kareiva, PM, Williamson, MH, Von Holle, B, Moyle, PB, Byers, JE & Goldwasser, L (1999) Impact: toward a framework for understanding the ecological effects of invaders. Biological Invasions 1: 319.Google Scholar
Parkes, JP, Norbury, GL, Heyward, RP & Sullivan, G (2002) Epidemiology of rabbit haemorrhagic disease (RHD) in the South Island, New Zealand, 1997–2001. Wildlife Research 29: 543555.Google Scholar
Parolin, P, Bresch, C, Poncet, C & Desneux, N (2014) Introducing the term ‘biocontrol plants’ for integrated pest management. Scientia Agricola 71: 7780.Google Scholar
Parry, D (2009) Beyond Pandora’s box: quantitatively evaluating non-target effects of parasitoids in classical biological control. Biological Invasions 11: 4758.Google Scholar
Parry, D, Spence, JR & Volney, JA (1997) Responses of natural enemies to experimentally increased populations of the forest tent caterpillar, Malacosoma distra. Ecological Entomology 22: 97108.Google Scholar
Pasteur, L (1874) Observations sure la coescistence der phyllocera et du mycelium constate a cully. Comptes Rendus de l’Académie des Sciences Paris 79: 1233.Google Scholar
Paynter, Q, Fowler, SV, Gourlay, AH, Groenteman, R, Peterson, PG, Smith, L & Winks, CJ (2010) Predicting parasitoid accumulation on biological control agents of weeds. Journal of Applied Ecology 47: 575582.Google Scholar
Paynter, Q, Fowler, SV, Gourlay, AH, Peterson, PG, Smith, LA & Winks, CJ (2015) Relative performance on test and target plants in laboratory tests predicts the risk of non-target attack in the field for arthropod weed biocontrol agents. Biological Control 80: 133142.Google Scholar
Paynter, Q, Gourlay, AH, Oboyski, PT, Fowler, SV, Hill, RL, Withers, TM, Parish, H & Hona, S (2008a) Why did specificity testing fail to predict the field host-range of the gorse pod moth in New Zealand? Biological Control 46: 453462.Google Scholar
Paynter, Q, Martin, N, Berry, J, Hona, S, Peterson, P, Gourlay, AH, Wilson-Davey, J, Smith, L, Winks, C & Fowler, SV (2008b) Non-target impacts of Phytomyza vitalbae a biological control agent of the European weed Clematis vitalba in New Zealand. Biological Control 44: 248258.Google Scholar
Paynter, Q, Overton, JM, Hill, RL, Bellgard, SE & Dawson, MI (2012) Plant traits predict the success of weed biocontrol. Journal of Applied Ecology 49: 11401148.Google Scholar
Paz, A, Jareno, D, Arroyo, L, Viñuela, J, Arroyo, B, Mougeot, F, Luque-Larenac, JJ & Fargallo, JA (2012) Avian predators as a biological control system of common vole (Microtus arvalis) populations in north-western Spain: experimental set-up and preliminary results. Pest Management Science 69: 444450.Google Scholar
Pearson, DE & Callaway, RM (2003) Indirect effects of host-specific biological control agents. Trends in Ecology & Evolution 18: 456461.Google Scholar
Pearson, DE & Callaway, RM (2004) Response to Thomas et al.: biocontrol and indirect effects. Trends in Ecology & Evolution 19: 6263.Google Scholar
Pearson, DE & Callaway, RM (2005) Indirect nontarget effects of host-specific biological control agents: implications for biological control. Biological Control 35: 288298.Google Scholar
Pearson, DE & Callaway, RM (2006) Biological control agents elevate hantavirus by subsidizing deer mouse populations. Ecology Letters 9: 443450.Google Scholar
Pearson, DE, McKelvey, KS & Ruggiero, LF (2000) Non-target effects of an introduced biological control agent on deer mouse ecology. Oecologia 122: 121128.Google Scholar
Pearson, DE, Potter, T & Maron, JL (2012) Biotic resistance: exclusion of native rodent consumers releases populations of a weak invader. Journal of Ecology 100: 13831390.Google Scholar
Pech, P, Fric, Z & Konvicka, M (2007) Species-specificity of the Phengaris (Maculinea)-Myrmica host system: fact or myth? (Lepidoptera: Lycaenidae; Hymenoptera: Formicidae). Sociobiology 50: 9831003.Google Scholar
Pedersen, BS & Mills, NJ (2004) Single vs. multiple introduction in biological control: the roles of parasitoid efficiency, antagonism and niche overlap. Journal of Applied Ecology 41: 973984.Google Scholar
Peever, TL, Liu, YC, Cortesi, P & Milgroom, MG (2000) Variation in tolerance and virulence in the chestnut blight fungus-hypovirus interaction. Applied and Environmental Microbiology 66: 48634869.Google Scholar
Pekar, S & Zd’Arkova, E (2004) A model of the biological control of Acarus siro by Cheyletus eruditus (Acari: Acaridae, Cheyletidae) on grain. Journal of Pest Science 77: 110.Google Scholar
Pels, B & Sabelis, MW (1999) Local dynamics, overexploitation and predator dispersal in an acarine predator–prey system. Oikos 86: 573583.Google Scholar
Pemberton, RW (2000) Predictable risks to native plants in weed biological control. Oecologia 125: 489494.Google Scholar
Pemberton, RW & Cordo, HA (2001) Potential and risks of biological control of Cactoblastis cactorum (Lepidoptera: Pyralidae) in North America. Florida Entomologist 84: 513526.Google Scholar
Pemberton, RW & Liu, H (2007) Control and persistence of native Opuntia on Nevis and St. Kitts 50 years after the introduction of Cactoblastis cactorum. Biological Control 41: 272282.Google Scholar
Peng, RK & Christian, K (2006) Effective control of Jarvis’s fruit fly, Bactrocera jarvisi (Diptera: Tephritidae), by the weaver ant, Oecophylla smaragdina (Hymenoptera: Formicidae), in mango orchards in the Northern Territory of Australia. International Journal of Pest Management 52: 275282.Google Scholar
Peng, RK, Christian, K & Reilly, D (2011) The effect of weaver ants Oecophylla smaragdina on the shoot borer Hypsipyla robusta on African mahoganies in Australia. Agricultural and Forest Entomology 13: 165171.Google Scholar
Pennacchio, F & Strand, MR (2006) Evolution of developmental strategies in parasitic Hymenoptera. Annual Review of Entomology 51: 233258.Google Scholar
Perez, C, Dill-Macky, R & Kinkel, LL (2008) Management of soil microbial communities to enhance populations of Fusarium graminearum antagonists in soil. Plant and Soil 302: 5369.Google Scholar
Perfecto, I & Castiñeras, A (1998) Deployment of the predaceous ants and their conservation in agroecosystems. In Conservation Biological Control (Barbosa, P, ed.), San Diego, CA, Academic Press, pp. 269290.Google Scholar
Perfecto, I & Snelling, R (1995) Biodiversity and the transformation of a tropical agroecosystem – ants in coffee plantations. Ecological Applications 5: 10841097.Google Scholar
Perfecto, I, Vandermeer, JH, Bautista, GL, Nunez, GI, Greenberg, R, Bichier, P & Langridge, S (2004) Greater predation in shaded coffee farms: the role of resident neotropical birds. Ecology 85: 26772681.Google Scholar
Periquet, G, Hedderwick, MP, El Agoze, M & Poirie, M (1993) Sex determination in the hymenopteran Diadromus pulchellus (Ichneumonoidea): validation of the one-locus multi-allele model. Heredity 70, 420427.Google Scholar
Perkins, RCL (1897) The introduction of beneficial insects into the Hawaiian islands. Nature 55: 499500.Google Scholar
Perlman, SJ, Kelly, SE, Zchori-Fein, E & Hunter, MS (2006) Cytoplasmic incompatibility and multiple symbiont infection in the ash whitefly parasitoid, Encarsia inaron. Biological Control 39: 474480.Google Scholar
Perrin, RM (1975) Role of perennial stinging nettle, Urtica dioica, as a reservoir of beneficial natural enemies. Annals of Applied Biology 81: 289297.Google Scholar
Persello-Cartieaux, F, Nussaume, L & Robaglia, C (2003) Tales from the underground: molecular plant-rhizobacteria interactions. Plant Cell and Environment 26: 189199.Google Scholar
Peschken, DP & McClay, A (1995) Picking the target: a revision of McClay’s scoring system to determine the suitability of a weed for classical biological control. In Proceedings of the VIII International Symposium on Biological Control of Weeds (Delfosse, ES & Scott, RR, eds.), Melbourne, Australia, DSIR/CSIRO, pp. 137143.Google Scholar
Peterson, AT, Soberon, J, Pearson, RG, Anderson, RP, Martínez-Meyer, E, Nakamura, M & Araújo, MB (2011) Ecological Niches and Geographic Distributions. Princeton, NJ, Princeton University Press.Google Scholar
Petitpierre, B, Kueffer, C, Broennimann, O, Randin, C, Daehler, C & Guisan, A (2012) Climatic niche shifts are rare among terrestrial plant invaders. Science 335: 13441348.Google Scholar
Pfeiffer, DG (1986) Effects of field applications of paraquat on densities of Panonychus ulmi (Koch) and Neoseiulus fallacis (Garman). Journal of Agricultural Entomology 3: 322325.Google Scholar
Pfeiffer, DG (2000) Selective insecticides. In Insect Pest Management: Techniques for Environmental Protection (Rechcigl, JE & Rechcigl, NA, eds.), Boca Raton, FL, CRC Press, pp. 131144.Google Scholar
Phatak, S, Callaway, MB & Vavrina, CS (1987) Biological control and its integration in weed management systems for purple and yellow nutsedge Cyperus rotundus and Cyperus esculentus. Weed Technology 1: 8491.Google Scholar
Phillips, CB, Baird, DB, Iline, II, McNeill, MR, Proffitt, JR, Goldson, SL & Kean, JM (2008) East meets west: adaptive evolution of an insect introduced for biological control. Journal of Applied Ecology 45: 948956.Google Scholar
Philpott, SM, Perfecto, I & Vandermeer, J (2008) Effects of predatory ants on lower trophic levels across a gradient of coffee management complexity. Journal of Animal Ecology 77: 505511.Google Scholar
Philpott, SM, Soong, O, Lowenstein, JH, Pulido, AL, Lopez, DT, Flynn, DFB & DeClerck, F (2009) Functional richness and ecosystem services: bird predation on arthropods in tropical agroecosystems. Ecological Applications 19: 18581867.Google Scholar
Pickett, CH & Bugg, RL (1998) Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests. Berkeley, University of California Press.Google Scholar
Pickett, CH, Roltsch, W & Corbett, A (2004) The role of a rubidium marked natural enemy refuge in the establishment and movement of Bemisia parasitoids. International Journal of Pest Management 50: 183191.Google Scholar
Pierson, EA & Mack, RN (1990) The population biology of Bromus tectorum in forests: effect of disturbance, grazing, and litter on seedling establishment and reproduction. Oecologia 84: 526533.Google Scholar
Pierson, EA & Weller, DM (1994) Use of mixtures of fluorescent pseudomonads to suppress take-all and improve the growth of wheat. Phytopathology 84: 940947.Google Scholar
Pike, KS & Burkhardt, CC (1974) Hyperparasites of Bathyplectes curculionis in Wyoming. Environmental Entomology 3: 953956.Google Scholar
Pilz, C, Wegensteiner, R & Keller, S (2007) Selection of entomopathogenic fungi for the control of the western corn rootworm Diabrotica virgifera virgifera. Journal of Applied Entomology 131: 426431.Google Scholar
Pimentel, D (1980) Environmental risks associated with biological controls. In Environmental Protection and Biological Forms of Control of Pest Organisms (Lundholm, B & Stackerud, M, eds.), Stockholm, Sweden, Swedish Natural Science Research Council, pp. 1124.Google Scholar
Pimentel, D, Glenister, C, Fast, S & Gallahan, D (1984) Environmental risks of biological pest controls. Oikos 42: 283290.Google Scholar
Pimentel, D, Lach, L, Zuniga, R & Morrison, D (2000) Environmental and economic costs of nonindigenous species in the United States. BioScience 50: 5365.Google Scholar
Pineda, A & Marcos-Garcia, MA (2008) Use of selected flowering plants in greenhouses to enhance aphidophagous hoverfly populations (Diptera: Syrphidae). Annales De La Societe Entomologique De France 44: 487492.Google Scholar
Piper, GL, Coombs, EM, Markin, GP & Joley, DB (2004) Rush skeletonweed. In Biological Control of Invasive Plants in the United States (Coombs, EM, Coombs, JK, Clark, GL & Cofrancesco, AF Jr., eds.), Corvallis, Oregon State University Press, pp. 293303.Google Scholar
Pizzatto, L & Shine, R (2012) Typhoid Mary in the frogpond: can we use native frogs to disseminate a lungworm biocontrol for invasive cane toads? Animal Conservation 15: 545552.Google Scholar
Plećaš, M, Gagic, V, Jankovic, M, Petrovic-Obradovic, O., Kavallieratos, NG, Tomanovic, Z, Thies, C, Tscharntke, T & Cetkovic, A (2014) Landscape composition and configuration influence cereal aphid-parasitoid-hyperparasitoid interactions and biological control differentially across years. Agriculture Ecosystems & Environment 183: 110.Google Scholar
Poehling, HM (1989) Selective application strategies for insecticides in agricultural crops. In Pesticides and Non-target Invertebrates (Jepson, PC, ed.), Andover, UK, Intercept, pp. 151175.Google Scholar
Pointier, J-P, David, P & Jarne, P (2011) The biological control of the snail hosts of schistosomes: the role of competitor snails and biological invasions. Biomphalaria Snails and Larval Trematodes (Toledo, R & Fried, B, eds.), New York, Springer, pp. 215238.Google Scholar
Polis, GA & Strong, DR (1996) Food web complexity and community dynamics. American Naturalist 147: 813846.Google Scholar
Pomerinke, MA, Thompson, DC & Clason, DL (1995) Bionomics of Cleonidius trivittatus (Coleoptera: Curculionidae): native biological control of purple locoweed (Rosales: Fabaceae). Environmental Entomology 24: 16961702.Google Scholar
Port, GR, Glen, DM & Symondson, WOC (2000) Success in biological control of terrestrial molluscs. In Biological Control: Measures of Success (Gurr, GM & Wratten, S, eds.), Dordrecht, The Netherlands, Kluwer, pp. 133158.Google Scholar
Postma, J, Montanari, M & Van den Boogert, P (2003) Microbial enrichment to enhance the disease suppressive activity of compost. European Journal of Soil Biology 39: 157163.Google Scholar
Powell, W (1989) Enhancing parasitoid activity in crops. In Insect Parasitoids (Waage, J & Greathead, D, eds.), London, UK, Academic Press, pp. 319340.Google Scholar
Power, M & McCarty, LS (2002) Trends in the development of ecological risk assessment and management frameworks. Human and Ecological Risk Assessment 8: 718.Google Scholar
Power, AG & Mitchell, CE (2004) Pathogen spillover in disease epidemics. American Naturalist 164: S79S89.Google Scholar
Pozo, MJ, Baek, J-M, Garcia, JM & Kenerley, CM (2004) Functional analysis of tvsp1, a serine protease-encoding gene in the biocontrol agent Trichoderma virens. Fungal Genetics and Biology 41: 336348.Google Scholar
Prasad, RP & Snyder, WE (2006) Polyphagy complicates conservation biological control that targets generalist predators. Journal of Applied Ecology 43: 343352.Google Scholar
Pratt, PD, Rayamajhi, MB, Van, TK & Center, TD (2004) Modeling the influence of resource availability on population densities of Oxyops vitiosa (Coleoptera: Curculionidae), a biological control agent of the invasive tree Melaleuca quinquenervia. Biocontrol Science and Technology 14: 5161.Google Scholar
Pratt, PD, Slone, DH, Rayamajhi, MB, Van, TK & Center, TD (2003) Geographic distribution and dispersal rate of Oxyops vitiosa (Coleoptera: Curculionidae), a biological control agent of the invasive tree Melaleuca quinquenervia in south Florida. Environmental Entomology 32: 397406.Google Scholar
Preisser, EL, Bolnick, DI & Benard, MF (2005) Scared to death? The effects of intimidation and consumption in predator–prey interactions. Ecology 86: 501509.Google Scholar
Price, RE, Müller, EJ, Brown, HD, D’Uamba, P & Jone, AA (1999) The first trial of a Metarhizium anisopliae var. acridum mycoinsecticide for the control of the red locust in a recognised outbreak area. Insect Science and Its Application 19: 323331.Google Scholar
Prior, C, Jollands, P & le Patourel, G (1988) Infectivity of oil and water formulations of Beauveria bassiana (Deuteromycotina: Hyphomycetes) to the cotton weevil pest Pantorhytes plutus (Coleoptera: Curculionidae). Journal of Invertebrate Pathology 52: 6672.Google Scholar
Prischmann, DA, James, DG, Storm, CP, Wright, LC & Snyder, WE (2007) Identity, abundance, and phenology of Anagrus spp. (Hymenoptera: Mymaridae) and leafhoppers (Homoptera: Cicadellidae) associated with grape, blackberry, and wild rose in Washington State. Annals of the Entomological Society of America 100: 4152.Google Scholar
Pyke, GH & White, AW (2000) Factors influencing predation on eggs and tadpoles of the endangered green and golden bell from Litora aurea by the introduced plague minnow Gambusia holbrooki. Australian Zoologist 31: 496505.Google Scholar
Pysek, P & Richardson, DM (2006) The biogeography of naturalization in alien plants. Journal of Biogeography 33: 20402050.Google Scholar
Quicke, DLJ (1997) Parasitic Wasps. London, UK, Chapman & Hall.Google Scholar
Quimby, PC, Gras, S, Widmer, T, Meikle, W & Sands, D (2004) Formulation of Selerotinia sclerotiorum for use against Cirsium arvense. Zeitschrift Fur Pflanzenkrankheiten Und Pflanzenschutz 19: 491495.Google Scholar
Raaijmakers, JM, Vlami, M & de Souza, JT (2002) Antibiotic production by bacterial biocontrol agents. Antonie van Leeuwenhoek International Journal of General and Molecular Microbiology 81: 537547.Google Scholar
Raaijmakers, JM & Weller, DM (1998) Natural plant protection by 2,4-diacetylphloroglucinol – producing Pseudomonas spp. in take-all decline soils. Molecular Plant-Microbe Interactions 11: 144152.Google Scholar
Raak-van den Berg, CL, De Lange, HJ & Van Lenteren, JC (2012) Intraguild predation behaviour of ladybirds in semi-field experiments explains invasion success of Harmonia axyridis. PLoS One 7: e40681.Google Scholar
Rabb, RL (1971) Naturally-occurring biological control in the eastern United States, with particular reference to tobacco insects. In Biological Control (Huffaker, CB, eds.), New York, Plenum, pp. 294311.Google Scholar
Radcliffe, EB & Flanders, KL (1998) Biological control of alfalfa weevil in North America. Integrated Pest Management Reviews 3: 225242.Google Scholar
Rafter, MA, Wilson, AJ, Senaratne, KADW & Dhileepan, K (2008) Climatic-requirements models of cat’s claw creeper Macfadyena unguis-cati (Bignoniaceae) to prioritise areas for exploration and release of biological control agents. Biological Control 44: 169179.Google Scholar
Raghu, S, Dhileepan, K & Scanlan, JC (2007) Predicting risk and benefit a priori in biological control of invasive plant species: a systems modelling approach. Ecological Modelling 208: 247262.Google Scholar
Raghu, S, Purcell, MF & Wright, AD (2013) Catchment context and the bottom-up regulation of the abundance of specialist semi-aquatic weevils on water hyacinth. Ecological Entomology 38: 117122.Google Scholar
Raghu, S & Van Klinken, RD (2006) Refining the ecological basis for agent selection in weed biological control. Australian Journal of Entomology 45: 251252.Google Scholar
Raghu, S & Walton, C (2007) Understanding the ghost of Cactoblastis past: historical clarifications on a poster child of classical biological control. BioScience 57: 699705.Google Scholar
Ramsell, J, Malloch, AJC & Whittaker, JB (1993) When grazed by Tipula paludosa, Lolium perenne is a stronger competitor of Rumex obtusifolius. Journal of Ecology 81: 777786.Google Scholar
Ramula, S, Knight, TM, Burns, JH & Buckley, YM (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
Rand, TA & Louda, SM (2004) Exotic weed invasion increases the susceptibility of native plants to attack by a biocontrol herbivore. Ecology 86: 15481554.Google Scholar
Rand, TA & Louda, SM (2006) Invasive insect abundance varies across the biogeographic distribution of a native host plant. Ecological Applications 16: 877890.Google Scholar
Rand, TA, Van Veen, FJF & Tscharnkte, T (2012) Landscape complexity differentially benefits generalized fourth, over specialized third, trophic level natural enemies. Ecography 35: 97104.Google Scholar
Randlkofer, B, Obermaier, E, Casas, J & Meiners, T (2010) Connectivity counts: disentangling effects of vegetation structure elements on the searching movement of a parasitoid. Ecological Entomology 35: 446455.Google Scholar
Randlkofer, B, Obermaier, E & Meiners, T (2007) Mother’s choice of the oviposition site: balancing risk of egg parasitism and need of food supply for the progeny with an infochemical shelter? Chemoecology 17: 177186.Google Scholar
Raoult, D, Audic, S, Robert, C, Abergel, C, Renesto, P, Ogata, H, La Scola, B, Suzan, M & Claverie, JM (2004) The 1.2 megabase genome sequence of mimivirus. Science 306: 13441350.Google Scholar
Rapp, G & Salum, MS (1995) Ant fauna, pest damage and yield in relation to the density of weeds in coconut sites in Zanzibar, Tanzania. Journal of Applied Entomology 119: 4548.Google Scholar
Rasgon, JL, Stryer, LM & Scott, TW (2003) Wolbachia-induced mortality as a mechanism to modulate pathogen transmission by vector arthropods. Journal of Medical Entomology 40: 125132.Google Scholar
Ravensberg, WJ (2011) A Roadmap to the Successful Development and Commercialization of Microbial Pest Control Products for Control of Arthropods. Dordrecht, The Netherlands, Springer.Google Scholar
Rechcigl, JE & Rechcigl, NA (1998) Biological and Biotechnological Control of Insect Pests. Boca Raton, FL, Lewis.Google Scholar
Rees, M & Hill, RL (2001) Large-scale disturbances, biological control and the dynamics of gorse populations. Journal of Applied Ecology 38: 364377.Google Scholar
Rees, M & Paynter, Q (1997) Biological control of Scotch broom: modelling the determinants of abundance and the potential impact of introduced insect herbivores. Journal of Applied Ecology 34: 12031221.Google Scholar
Reeve, JD (1988) Environmental variability, migration, and persistence in host-parasitoid systems. American Naturalist 132: 810836.Google Scholar
Reeves, JL & Lorch, PD (2012) Biological control of invasive aquatic and wetland plants by arthropods: a meta-analysis of data from the last three decades. BioControl 57: 103116.Google Scholar
Rehage, JS & Sih, A (2004) Dispersal behavior, boldness, and the link to invasiveness: a comparison of four Gambusia species. Biological Invasions 6: 379391.Google Scholar
Reid, AM, Morin, L, Downey, PO, French, K & Virtue, JG (2009) Does invasive plant management aid the restoration of natural ecosystems? Biological Conservation 142: 23422349.Google Scholar
Reitz, SR & Trumble, JT (2002) Competitive displacement among insects and arachnids. Annual Review of Entomology 47: 435466.Google Scholar
Rejmanek, M (1996) Species richness and resistance to invasions. In Biodiversity and Ecosystem Processes in Tropical Forests (Orians, GH, Dirzo, R & Cushman, JH, eds.), Heidelberg, Germany, Springer, pp. 153172.Google Scholar
Remington, CL (1968) The population genetics of insect introduction. Annual Review of Entomology 13: 415426.Google Scholar
Rethwisch, MD & Manglitz, GR (1986) Parasitoids of Bathyplectes curculionis (Hymenoptera: Ichneumonidae) in southeastern Nebraska. Journal of the Kansas Entomological Society 59: 648652.Google Scholar
Reznick, DN & Ghalambor, CK (2001) The population ecology of contemporary adaptations: what empirical studies reveal about the conditions that promote adaptive evolution. Genetica 112–113: 183198.Google Scholar
Rhodes, EM & Liburd, OE (2006) Evaluation of predatory mites and Acramite for control of twospotted spider mites in strawberries in north central Florida. Journal of Economic Entomology 99: 12911298.Google Scholar
Ricciardi, A & Cohen, J (2007) The invasiveness of an introduced species does not predict its impact. Biological Invasions 9: 309315.Google Scholar
Ricciardi, A, Hoopes, MF, Marchetti, E & Lockwood, JL (2013) Progress toward understanding the ecological impacts of nonnative species. Ecological Monographs 83: 263282.Google Scholar
Ricciardi, A & Simberloff, D (2009) Assisted colonization is not a viable conservation strategy. Trends in Ecology & Evolution 24: 248253.Google Scholar
Richardson, DM (2000) Naturalization and invasion of alien plants: concepts and definitions. Diversity and Distributions 6: 93107.Google Scholar
Richardson, DM (2011) Fifty Years of Invasion Ecology. The Legacy of Charles Elton. Oxford, UK, Wiley-Blackwell.Google Scholar
Richardson, DM, Allsopp, N, D’Antonio, CM, Milton, SJ & Rejmanek, M (2000) Plant invasions – the role of mutualisms. Biological Reviews 75: 6593.Google Scholar
Richardson, DM & Pyšek, P (2012) Naturalization of introduced plants: ecological drivers of biogeographical patterns. New Phytologist 196: 383396.Google Scholar
Ridenour, WM & Callaway, RM (2001) The relative importance of allelopathy in interference: the effects of an invasive weed on a native bunchgrass. Oecologia 126: 444450.Google Scholar
Ridenour, WL & Callaway, RM (2003) Root herbivores, pathogenic fungi, and competition between Centaurea maculosa and Festuca idahoensis. Plant Ecology 169: 161170.Google Scholar
Riechert, SE (1998) The role of spiders and their conservation in agroecosystems. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests (Pickett, CH & Bugg, RL, eds.), Berkeley, University of California Press, pp. 211237.Google Scholar
Rincon, C, Bordat, D, Lohr, B & Dupas, S (2006) Reproductive isolation and differentiation between five populations of Cotesia plutellae (Hymenoptera: Braconidae), parasitoid of Plutella xylostella (Lepidoptera: Plutellidae). Biological Control 36: 171182.Google Scholar
Ripper, WE, Greenslade, RM & Hartley, GS (1951) Selective insecticides and biological control. Journal of Economic Entomology 44: 448459.Google Scholar
Risch, SJ, Andow, DA & Altieri, MA (1983) Agroecosystem diversity and pest control: data, tentative conclusions, and new research directions. Environmental Entomology 12: 625629.Google Scholar
Rishbeth, J (1963) Stump protection against Fomes annosus: III. Inoculation with Peniophora gigantea. Annals of Applied Biology 52: 6377.Google Scholar
R’Kha, S, Capy, P & David, JR (1991) Host-plant specialization in the Drosophila melanogaster species complex: a physiological, behavioral and genetical analysis. Proceedings of the National Academy of Sciences, USA 88: 18351939.Google Scholar
Robinson, AS, Franz, G & Atkinson, PW (2004) Insect transgenesis and its potential role in agriculture and human health. Insect Biochemistry and Molecular Biology 34: 113120.Google Scholar
Robinson, KA, Jonsson, M, Wratten, SD, Wade, MR & Buckley, HL (2008) Implications of floral resources for predation by an omnivorous lacewing. Basic and Applied Ecology 9: 172181.Google Scholar
Rochat, J & Gutierrez, AP (2001) Weather-mediated regulation of olive scale by two parasitoids. Journal of Animal Ecology 70: 476490.Google Scholar
Roderick, GK (1992) Postcolonization evolution of natural enemies. In Selection Criteria and Ecological Consequences of Importing Natural Enemies (Kauffman, WC & Nechols, JR, eds.), Lanham, MD, Entomological Society of America, pp. 7186.Google Scholar
Roderick, GK (2004) Tracing the origins of pests and natural enemies: genetic and statistical approaches. In Genetics, Evolution and Biological Control (Ehler, LE, Sforza, R & Mateille, T, eds.), Wallingford, UK, CABI Publishing, pp. 97112.Google Scholar
Roderick, GK, Hufbauer, R & Navajas, M (2012) Evolution and biological control. Evolutionary Applications 5: 419423.Google Scholar
Roderick, GK & Navajas, M (2003) Genes in new environments: genetics and evolution in biological control. Nature Reviews Genetics 4: 889899.Google Scholar
Roderick, GK & Navajas, M (2008) The primacy of evolution in biological control. In Proceedings of the XII International Symposium on Weed Biological Control (Julien, M, ed.), 403409. Wallingford, UK, CABI Publishing, pp. 403409.Google Scholar
Rogers, DJ (1972) Random search and insect population models. Journal of Animal Ecology 41: 369383.Google Scholar
Rogers, DJ & Hassell, MP (1974) General models for insect parasite and predator searching behaviour: interference. Journal of Animal Ecology 43: 239253.Google Scholar
Rogers, MA, Krischik, VA & Martin, LA (2007) Effect of soil application of imidacloprid on survival of adult green lacewing, Chrysoperla carnea (Neuroptera: Chrysopidae), used for biological control in greenhouses. Biological Control 42: 172177.Google Scholar
Rogers, WE & Siemann, E (2004) Invasive ecotypes tolerate herbivory more effectively than native ecotypes of the Chinese tallow tree Sapium sebiferum. Journal of Applied Ecology 41: 561570.Google Scholar
Rohani, P, Godfray, HJC & Hassell, MP (1994) Aggregation and the dynamics of host-parasitoid systems: a discrete-generation model with within-generation redistribution. American Naturalist 144: 491509.Google Scholar
Roitberg, BD, Sircom, J, Roitberg, C, Van Alphen, JJM & Mangel, M (1992) Seasonal dynamic shifs in patch exploitation by parasitic wasps. Behavioral Ecology 3: 156165.Google Scholar
Roitberg, BD, Sircom, J, Roitberg, C, Van Alphen, JJM & Mangel, M (1993) Life expectancy and reproduction. Nature 364: 108.Google Scholar
Roland, J & Taylor, PD (1997) Insect parasitoid species respond to forest structure at different spatial scales. Nature 386: 710713.Google Scholar
Romeis, J, Babendreier, D, Wäckers, FL & Shanower, TG (2005) Habitat and plant specificity of Trichogramma egg parasitoids – underlying mechanisms and implications. Basic and Applied Ecology 6: 215236.Google Scholar
Romeis, J & Wäckers, FL (2000) Feeding responses by female Pieris brassicae butterflies to carbohydrates and amino acids. Physiological Entomology 25: 247253.Google Scholar
Room, PM (1990) Ecology of a simple plant-herbivore system: biological control of Salvinia. Trends in Ecology & Evolution 5: 7479.Google Scholar
Room, PM, Harley, KLS, Forno, IW & Sands, DPA (1981) Successful biological control of the floating weed salvinia. Nature 394: 7880.Google Scholar
Rondelaud, D, Vingoles, P, Dreyfuss, G & Mage, M (2006) The control of Galba trunculata (Gastropoda: Lymnaeidae) by the terrestrial snail Zonitoides nitidus on acid soils. Biological Control 39: 290299.Google Scholar
Root, RB (1973) Organization of a plant-arthropod association in simple and diverse habitats: the fauna of collards (Brassica oleracea). Ecological Monographs 43: 95124.Google Scholar
Rose, KE, Louda, SM & Rees, M (2005) Demographic and evolutionary impacts of native and invasive insect herbivores on Cirsium canescens. Ecology 86: 453465.Google Scholar
Rose, M & DeBach, P (1992) Biological control of Parabemisia myricae (Kuwana) (Homoptera: Aleyrodidae) in California. Israel Journal of Entomology 25–26: 7395.Google Scholar
Rose, MR, Nusbaum, TJ & Chippendale, AK (1996) Laboratory evolution: the experimental wonderland and the Cheshire Cat Syndrome. In Adaptation (Rose, MR & Lauder, GV, eds.), San Diego, CA, Academic Press, pp. 221224.Google Scholar
Rosen, D (1994) Advances in the Study of Aphytis. Andover, UK, Intercept.Google Scholar
Rosen, D & DeBach, P (1977) Use of scale-insect parasites in Coccoidea systematics. Virginia Polytechnic Institute and State University Research Division Bulletin 127: 521.Google Scholar
Rosen, D & DeBach, P (1979) Species of Aphytis of the World. The Hague, The Netherlands, W. Junk.Google Scholar
Rosenheim, JA (1998) Higher-order predators and the regulation of insect herbivore populations. Annual Review of Entomology 43: 421448.Google Scholar
Rosenheim, JA & Corbett, A (2003) Omnivory and the indeterminacy of predator function: can a knowledge of foraging behavior help? Ecology 84: 25382548.Google Scholar
Rosenheim, JA & Harmon, J (2006) The influence of intraguild predation on the suppression of a shared prey population: an empirical assessment. In Trophic and Guild Interactions in Biological Control (Brodeur, J & Boivin, G, eds.), Dordrecht, The Netherlands, Springer, pp. 120.Google Scholar
Rosenheim, JA, Kaya, HK, Ehler, LE, Marois, JJ & Jaffee, BA (1995) Intraguild predation among biological-control agents: theory and evidence. Biological Control 5: 303335.Google Scholar
Rosenheim, JA, Wilhoit, LR & Armer, CA (1993) Influence of intraguild predation among generalist insect predators on the suppression of an herbivore population. Oecologia 96: 439449.Google Scholar
Rosenheim, JA, Wilhoit, LR, Goodell, PB, Grafton-Cardwell, EE & Leigh, TF (1997) Plant compensation, natural biological control, and herbivory by Aphis gossypii on pre-reproductive cotton: the anatomy of a non-pest. Entomologia Experimentalis et Applicata 85: 4563.Google Scholar
Rouchet, R & Vorburger, C (2012) Strong specificity in the interaction between parasitoids and symbiont-protected hosts. Journal of Evolutionary Biology 25: 23692375.Google Scholar
Roush, RT (1990) Genetic variation in natural enemies: critical issues for colonization in biological control. In Critical Issues in Biological Control (Mackauer, M, Ehler, LE & Roland, J, eds.), Andover, UK, Intercept, pp. 265288.Google Scholar
Roush, RT & Hopper, KR (1995) Use of single family lines to preserve genetic variation in laboratory colonies. Annals of the Entomological Society of America 88: 713717.Google Scholar
Roy, HE, Adriaens, T, Isaac, NJB, Kenis, M, Onkelinx, T, San Martin, G, Brown, PMJ, Hautier, L, Poland, R, Roy, DB, Comont, R, Eschen, R, Frost, R, Zindel, R, Van Vlaenderen, J, Nedved, O, Ravn, HP, Gregoire, JC, de Biseau, J-C & Maes, D (2012) Invasive alien predator causes rapid declines of native European ladybirds. Diversity and Distributions 18: 717725.Google Scholar
Roy, HE, Lawson-Handley, L-J, Schonrogge, K, Poland, RL & Purse, BV (2011) Can the enemy release hypothesis explain the success of invasive alien predators and parasitoids? BioControl 56: 451468.Google Scholar
Roy, HE, Pell, JK, Clark, SJ & Alderson, PG (1998) Implications of predator foraging on aphid pathogen dynamics. Journal of Invertebrate Pathology 71: 236247.Google Scholar
Royama, T (1984) Population dynamics of the spruce budworm Choristoneura fumiferana. Ecological Monographs 54: 429462.Google Scholar
Ruiz, GM & Carlton, JT (2003) Invasive Species: Vectors and Management Practices. Washington, DC, Island Press.Google Scholar
Ruiz, L, Flores, S, Cancino, J, Arredondo, J, Valle, J, Diaz-Fleischer, F & Williams, T (2008) Lethal and sublethal effects of spinosad-based GF-120 bait on the tephritid parasitoid Diachasmimorpha longicaudata (Hymenoptera: Braconidae). Biological Control 44: 296304.Google Scholar
Russell, EP (1989) Enemies hypothesis: a review of the effect of vegetational diversity on predatory insects and parasitoids. Environmental Entomology 18: 590599.Google Scholar
Russell, K & Bessin, R (2009) Integration of Trichogramma ostriniae releases and habitat modification for suppression of European corn borer (Ostrinia nubilalis Hubner) in bell peppers. Renewable Agriculture and Food Systems 24: 1924.Google Scholar
Russell, M (2015) A meta-analysis of physiological and behavioral responses of parasitoid wasps to flowers of individual plant species. Biological Control 82: 96103.Google Scholar
Ruxton, GD, Gurney, WSC & de Roos, AM (1992) Interference and generation cycles. Theoretical Population Biology 42: 235253.Google Scholar
Ryan, PR, Dessaux, Y, Thomashow, LS & Weller, DM (2009) Rhizosphere engineering and management for sustainable agriculture. Plant and Soil 321: 363383.Google Scholar
Ryan, RB (1990) Evaluation of biological control: introduced parasites of larch casebearer (Lepidoptera: Coleophoridae) in Oregon. Environmental Entomology 19: 18731881.Google Scholar
Sailer, RI (1978) Our immigrant insect fauna. ESA Bulletin 24: 311.Google Scholar
Saint, KM, French, N & Kerr, P (2001). Genetic variation in Australian isolates of myxoma virus: an evolutionary and epidemiological study. Archives of Virology 146: 11051123.Google Scholar
Saito, Y, Urano, S, Nakao, H, Amimoto, K & Mori, H (1996) A simulation model for predicting the efficiency of biological control of spider mites by phytoseiid predators 2. Validity tests and data necessary for practical usage. Japanese Journal of Applied Entomology and Zoology 40: 113120.Google Scholar
Sakai, AK, Allendorf, FW, Holt, JS, Lodge, DM, Molofsky, J, With, KA, Baughman, S, Cabin, RJ, Cohen, JE, Ellstrand, NC, McCauley, DE, O’Neil, P, Parker, IM, Thompson, JN & Weller, SG (2001) The population biology of invasive species. Annual Review of Ecology and Systematics 32: 305332.Google Scholar
Sakuratani, Y, Matsumoto, Y, Oka, M, Kubo, T, Fujii, A, Uotani, M & Teraguchi, T (2000) Life history of Adalia bipunctata (Coleoptera: Coccinellidae) in Japan. European Journal of Entomology 97: 555558.Google Scholar
Salin, C, Deprez, B, Van Bockstaele, DR, Mahillon, J & Hance, T (2004) Sex determination mechanism in the hymenopteran parasitoid Aphidius rhopalosiphi De Stefani-Peres (Braconidae: Aphidiinae). Belgian Journal of Zoology 134: 1521.Google Scholar
Salt, G & Van den Bosch, R (1967) The defense reactions of three species of Hypera (Coleoptera: Curculionidae) to an ichneumon wasp. Journal of Invertebrate Pathology 9: 164177.Google Scholar
Samac, DA & Kinkel, LL (2001) Suppression of the root-lesion nematode (Pratylenchus penetrans) in alfalfa (Medicago sativa) by Streptomyces spp. Plant and Soil 235: 3544.Google Scholar
Samac, DA, Willert, AM, McBride, MJ & Kinkel, LL (2003) Effect of antibiotic-producing Streptomyces on nodulation and leaf spot in alfalfa. Applied Soil Ecology 22: 5566.Google Scholar
Sandrock, C & Vorburger, C (2011) Single-locus recessive inheritance of asexual reproduction in a parasitoid wasp. Current Biology 21: 433437.Google Scholar
Sands, DPA (1997) The ‘safety’ of biological control agents: assessing their impact on beneficial and other non-target hosts. Memoirs of the Museum of Victoria 56: 611615.Google Scholar
Sands, DPA & Van Driesche, RG (2004) Using the scientific literature to estimate the host range of a biological control agent. In Assessing Host Ranges for Parasitoids and Predators Used for Classical Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, U.S. Forest Service, pp. 1523.Google Scholar
Sant, D, Casanova, E, Segarra, G, Aviles, M, Reis, M & Trillas, MI (2010) Effect of Trichoderma asperellum strain T34 on Fusarium wilt and water usage in carnation grown on compost-based growth medium. Biological Control 53: 291296.Google Scholar
Santoyo, G, Orozco-Mosqueda, MdC & Govindappa, M (2012) Mechanisms of biocontrol and plant-growth promoting activity in soil bacterial species of Bacillus and Pseudomonas: a review. Biocontrol Science and Technology 22: 855872.Google Scholar
Sato, S & Dixon, AFG (2004) Effect of intraguild predation on the survival and development of three species of aphidophagous ladybirds: consequences for invasive species. Agricultural and Forest Entomology 6: 2124.Google Scholar
Sato, Y & Ohsaki, N (2004) Response of the wasp (Cotesia glomerata) to larvae of the large white butterfly. Ecological Research 19: 445449.Google Scholar
Sato, S, Yasuda, H, Evans, EW & Dixon, AFG (2009) Vulnerability of larvae of two species of aphidophagous ladybirds, Adalia bipunctata Linnaeus and Harmonia axyridis Pallas, to cannibalism and intraguild predation. Entomological Science 12: 111115.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
Sawyer, RC (1996) To Make a Spotless Orange. Ames, University of Iowa Press.Google Scholar
Sax, DF (2001) Latitudinal gradients and geographic ranges of exotic species: implications for biogeography. Journal of Biogeography 28: 139150.Google Scholar
Scanlan, JC, Grant, WE, Hunter, DM & Milner, RJ (2001) Habitat and environmental factors influencing the control of migratory locusts (Locusta migratoria) with an entomopathogenic fungus (Metarhizium anisopliae). Ecological Modelling 136: 223236.Google Scholar
Scarborough, CL, Ferrari, J & Godfray, HCJ (2005) Aphid protected from pathogen by endosymbiont. Science 310: 1781.Google Scholar
Schellhorn, NA, Bianchi, FJJA & Hsu, CL (2014) Movement of entomophagous arthropods in agricultural landscapes: links to pest suppression. Annual Review of Entomology 59: 559581.Google Scholar
Schellhorn, NA, Glatz, RV & Wood, GM (2010) The risk of exotic and native plants as hosts for four pest thrips (Thysanoptera: Thripinae). Bulletin of Entomological Research 100: 501510.Google Scholar
Schellhorn, NA, Kuhman, TR, Olson, AC & Ives, AR (2002) Competition between native and introduced parasitoids of aphids: nontarget effects and biological control. Ecology 83: 27452757.Google Scholar
Schemske, DW, Mittelbach, GG, Cornell, HV, Sobel, JM & Roy, K (2009) Is there a latitudinal gradient in the importance of biotic interactions? Annual Review of Ecology, Evolution and Systematics 40: 245269.Google Scholar
Schisler, DA, Slininger, PJ, Behle, RW & Jackson, MA (2004) Formulation of Bacillus spp. for biological control of plant diseases. Phytopathology 94: 12671271.Google Scholar
Schlatter, D, Fubuh, A, Xiao, K, Hernandez, D, Hobbie, S & Kinkel, L (2009) Resource amendments influence density and competitive phenotypes of Streptomyces in soil. Microbial Ecology 57: 413420.Google Scholar
Schmidl, L (1972) Studies on the control of ragwort, Senecio jacobaea L. with the cinnabar moth, Callimorpha jacobaeae (L.) (Arctiidae: Lepidoptera), in Victoria. Weed Research 12: 4657.Google Scholar
Schmidt, MH, Thies, C, Nentwig, W & Tscharntke, T (2008) Contrasting responses of arable spiders to the landscape matrix at different spatial scales. Journal of Biogeography 35: 157166.Google Scholar
Schmidt, MH & Tscharntke, T (2005) Landscape context of sheetweb spider (Araneae: Linyphiidae) abundance in cereal fields. Journal of Biogeography 32: 467473.Google Scholar
Schmidt, NP, O’Neal, ME & Singer, JW (2007) Alfalfa living mulch advances biological control of soybean aphid. Environmental Entomology 36: 416424.Google Scholar
Schmitz, OJ, Krival, V & Ovadia, O (2004) Trophic cascades: the primacy of trait-mediated indirect interactions. Ecology Letters 7: 153163.Google Scholar
Schoener, TW (1983) Field experiments on interspecific competition. American Naturalist 122: 240285.Google Scholar
Schooler, SS, De Barro, P & Ives, AR (2011) The potential for hyperparasitism to compromise biological control: why don’t hyperparasitoids drive their primary parasitoid hosts extinct? Biological Control 58: 167173.Google Scholar
Schreiner, I (1989) Biological control introductions in the Caroline and Marshall Islands. Proceedings of the Hawaiian Entomological Society 29: 5769.Google Scholar
Schrey, SD & Tarkka, MT (2008) Friends and foes: streptomycetes as modulators of plant disease and symbiosis. Antonie van Leeuwenhoek International Journal of General and Molecular Microbiology 94: 1119.Google Scholar
Seaman, GA & Randall, JE (1962) The mongoose as a predator in the Virgin Islands. Journal of Mammalogy 43: 544546.Google Scholar
Seamark, RF (2001) Biotech prospects for the control of introduced mammals in Australia. Reproduction, Fertility and Development 13: 705711.Google Scholar
Seastedt, TR (2015) Biological control of invasive plant species: a reassessment for the Anthropocene. New Phytologist 205: 490502.Google Scholar
Sebolt, DC & Landis, DA (2002) Neonate Galerucella calmariensis (Coleoptera: Chrysomelidae) behavior on purple loosestrife (Lythrum salicaria) contributes to reduced predation. Environmental Entomology 31: 880886.Google Scholar
Sebolt, DC & Landis, DA (2004) Arthropod predators of Galerucella calmariensis L. (Coleoptera: Chrysomelidae): an assessment of biotic interference. Environmental Entomology 33: 356461.Google Scholar
Secord, D (2003) Biological control of marine invasive species: cautionary tales and land-based lessons. Biological Invasions 5: 117131.Google Scholar
Segal, D & Glazer, I (2000) Genetics for improving biological control agents: the case of entomopathogenic nematodes. Crop Protection 19: 685689.Google Scholar
Segoli, M & Rosenheim, JA (2013) The link between host density and egg production in a parasitoid insect: comparison between agricultural and natural habitats. Functional Ecology 27: 12241232.Google Scholar
Seguni, ZSK, Way, MJ & Van Mele, P (2011) The effect of ground vegetation management on competition between the ants Oecophylla longinoda and Pheidole megacephala and implications for conservation biological control. Crop Protection 30: 713717.Google Scholar
Settle, WH, Ariawan, H, Astuti, ET, Cahyana, W, Hakim, AL, Hindayana, D, Lestari, AS & Pajarningsih, (1996) Managing tropical rice pests through conservation of generalist natural enemies and alternative prey. Ecology 77: 19751988.Google Scholar
Shabana, YM (2005) The use of oil emulsions for improving the efficacy of Alternaria eichhorniae as a mycoherbicide for waterhyacinth (Eichhornia crassipes). Biological Control 32: 7889.Google Scholar
Shah, FA, Wang, CS & Butt, TM (2005) Nutrition influences growth and virulence of the insect-pathogenic fungus Metarhizium anisopliae. FEMS Microbiology Letters 251: 259266.Google Scholar
Shah, PA, Gbongboui, C, Godonou, I, Hossou, A & Lomer, CJ (1998) Natural incidence of Metarhizium flavoviride infection in two grasshopper communities in northern Benin. Biocontrol Science and Technology 8: 335344.Google Scholar
Shanmuganathan, T, Pallister, J, Doody, S, McCallum, H, Robinson, T, Sheppard, A, Hardy, C, Halliday, D, Venables, D, Voysey, R, Strive, T, Hinds, L & Hyatt, A (2010) Biological control of the cane toad in Australia: a review. Animal Conservation 13: 1623.Google Scholar
Shapiro, DI, Glazer, I & Segal, D (1997) Genetic improvement of heat tolerance in Heterorhabditis bacteriophora through hybridization. Biological Control 8: 153159.Google Scholar
Shapiro-Ilan, DI, Stuart, RJ & McCoy, CW (2005) Targeted improvement of Steinernema carpocapsae for control of pecan weevil, Curculio caryae (Horn) (Coleoptera: Curculionidae) through hybridization and bacterial transfer. Biological Control 34: 215221.Google Scholar
Shaw, MR (1994) Parasitoid host ranges. In Parasitoid Community Ecology (Hawkins, BA & Sheehan, W, eds.), Oxford, UK, Oxford University Press, pp. 111144.Google Scholar
Shaw, PB (1984) Simulation model of a predator–prey system comprised of Phytoseiulus persimilis (Acari: Phytoseiidae) and Tetranychus urticae (Acari: Tetranychidae). 1. Structure and validation of the model. Researches on Population Ecology 26: 235260.Google Scholar
Shea, K & Kelly, D (1998) Estimating biocontrol agent impact with matrix models: Carduus nutans in New Zealand. Ecological Applications 8: 824832.Google Scholar
Shea, K, Jongejans, E, Skarpaas, O, Kelly, D & Sheppard, AW (2010) Optimal management strategies to control local population growth or population spread may not be the same. Ecological Applications 20: 11481161.Google Scholar
Shea, K, Kelly, D, Sheppard, AW & Woodburn, TL (2005) Context-dependent biological control of an invasive thistle. Ecology 86: 31743181.Google Scholar
Shea, K & Possingham, HP (2000) Optimal release strategies for biological control agents: an application of stochastic dynamic programming to population management. Journal of Applied Ecology 37: 7786.Google Scholar
Shea, K, Possingham, HP, Murdoch, WW & Roush, R (2002) Active adaptive management in insect pest and weed control: intervention with a plan for learning. Ecological Applications 12: 927936.Google Scholar
Shepherd, RL, Otvos, IS, Chorney, RJ & Cunningham, JC (1984) Pest management of Douglas-fir tussock moth (Lepidoptera: Lymantriidae): prevention of an outbreak through early treatment with a nuclear polyhedrosis virus by ground and aerial applications. Canadian Entomologist 116: 15331542.Google Scholar
Shearer, PW & Jones, VP (1998) Suitability of selected weeds and ground covers as host plants of Nezara viridula (L.) (Hemiptera: Pentatomidae). Proceedings of the Hawaiian Entomological Society 33: 7582.Google Scholar
Sheehan, W (1986) Response by specialist and generalist natural enemies to agroecosystem diversification: a selective review. Environmental Entomology 15: 456461.Google Scholar
Sheldon, SP & Creed, RP (2003) The effect of a native biological control agent for Eurasian watermilfoil on six North American watermilfoils. Aquatic Botany 76: 259265.Google Scholar
Sheppard, AW, Hill, R, DeClerck-Floate, RA, McClay, A, Olckers, T, Quimby, PCJ & Zimmermann, HG (2003) A global review of risk-benefit-cost analysis for the introcuction of classical biological control agents against weeds: a crisis in the making? Biocontrol News and Information 24: 91 N108 N.Google Scholar
Sheppard, AW & Raghu, S (2005) Working at the interface of art and science: how best to select an agent for classical biological control? Biological Control 34: 233235.Google Scholar
Sheppard, AW, Rees, M, Smyth, M, Grigulis, K & Buckley, YM (2002) Modelling the population interactions between Mogulones larvatus and its host plant Echium plantagineum. In Procceedings of the 13th Australian Weed Conference (Spafford Jacob, H, Dodd, J & Moore, JH, eds.), Perth, Australia, Plant Protectin Society of Western Australia, Pages 248251.Google Scholar
Sheppard, AW, Smyth, MJ & Swirepik, A (2001) The impact of a root-crown weevil and pasture competition on the winter annual Echium plantagineum. Journal of Applied Ecology 38: 291300.Google Scholar
Sheppard, AW, Van Klinken, RD & Heard, TA (2005) Scientific advances in the analysis of direct risks of weed biological control agents to nontarget plants. Biological Control 35: 215226.Google Scholar
Sherratt, TN & Jepson, PC (1993) A metapopulation approach to modeling the long-term impact of pesticides on invertebrates. Journal of Applied Ecology 30: 696705.Google Scholar
Shigesada, N & Kawasaki, K (1997) Biological Invasions: Theory and Practice. Oxford, UK, Oxford University Press.Google Scholar
Shishido, M, Miwa, C, Usami, T, Amemiya, Y & Johnson, KB (2005) Biological control efficiency of Fusarium wilt of tomato by nonpathogenic Fusarium oxysporum Fo-B2 in different environments. Phytopathology 95: 10721080.Google Scholar
Siddiqui, IA & Shaukat, SS (2002) Mixtures of plant disease suppressive bacteria enhance biological control of multiple tomato pathogens. Biology and Fertility of Soils 36: 260268.Google Scholar
Siddiqui, Y, Meon, S, Ismail, MR & Ali, A (2008) Trichoderma-fortified compost extracts for the control of choanephora wet rot in okra production. Crop Protection 27: 385390.Google Scholar
Siemann, E, Haarstad, J & Tilman, D (1999) Dynamics of plant and arthropod diversity during old field succession. Ecography 22: 406414.Google Scholar
Siemann, E & Rogers, WE (2003a) Increased competitive ability of an invasive tree may be limited by an invasive beetle. Ecological Applications 13: 15031507.Google Scholar
Siemann, E & Rogers, WE (2003b) Reduced resistance of invasive varieties of the alien tree Sapium sebiferum to a generalist herbivore. Oecologia 135: 451457.Google Scholar
Siemann, E, Tilman, D, Haarstad, J & Ritchie, M (1998) Experimental tests of the dependence of arthropod diversity on plant diversity. American Naturalist 152: 738750.Google Scholar
Silva, I, Van Meer, MMM, Roskam, MM, Hoogenboom, A, Gort, G & Stouthamer, R (2000) Biological control potential of Wolbachia-infected versus uninfected wasps: laboratory and greenhouse evaluation of Trichogramma cordubensis and T. deion strains. Biocontrol Science and Technology 10: 223238.Google Scholar
Simberloff, D (1986) Introduced insects: a biogeographic and systematic perspective. In Ecology of Biological Invasions of North America and Hawaii (Mooney, HA & Drake, JA, ed.), New York, Springer, pp. 326.Google Scholar
Simberloff, D (1995) Why do introduced species appear to devastate islands more than mainland areas? Pacific Science 49: 8797.Google Scholar
Simberloff, D (2009) The role of propagule pressure in biological invasions. Annual Review of Ecology Evolution and Systematics 40: 81102.Google Scholar
Simberloff, D (2012) Risks of biological control for conservation purposes. Biocontrol 57: 263276.Google Scholar
Simberloff, D (2013) Invasive Species: What Everyone Needs to Know. Oxford, UK, Oxford University Press.Google Scholar
Simberloff, D & Gibbons, L (2004) Now you see them, now you don’t! – population crashes of established introduced species. Biological Invasions 6: 161172.Google Scholar
Simberloff, D & Stiling, P (1996a) How risky is biological control? Ecology 77: 19651974.Google Scholar
Simberloff, D & Stiling, P (1996b) The risks of species introduced for biological control. Biological Conservation 78: 185192.Google Scholar
Simberloff, D & Von Holle, B (1999) Positive interactions of nonindigenous species: invasional meltdown? Biological Invasions 1: 2132.Google Scholar
Simmonds, FJ (1958) The effect of lizards on the biological control of scale insects in Bermuda. Bulletin of Entomological Research 49: 601612.Google Scholar
Simmonds, FJ (1963) Genetics and biological control. Canadian Entomologist 95: 561567.Google Scholar
Simmonds, FJ & Bennett, FD (1966) Biological control of Opuntia spp. by Cactoblastis cactorum in the Leeward Islands (West Indies). Entomophaga 11: 183189.Google Scholar
Simmonds, FJ, Franz, JM & Sailer, RI (1976) History of biological control. In Theory and Practice of Biological Control (Huffaker, CB & Messenger, PS, eds.), New York, Academic Press, pp. 1741.Google Scholar
Simoes, N & Rosa, JS (1996) Pathogenicity and host specificity of entomopathogenic nematodes. Biocontrol Science and Technology 6: 403411.Google Scholar
Simon, JC, Carre, S, Boutin, M, Prunier-Leterne, N, Sabater-Munoz, B, Latorre, A & Bournoville, R (2003) Host-based divergence in populations of the pea aphid: insights from nuclear markers and the prevalence of facultative symbionts. Proceedings of the Royal Society of London B 270: 17031712.Google Scholar
Simpson, RG, Cross, JE, Best, RL, Ellsbury, MM & Coseglia, AF (1979) Effects of hyperparasites on population levels of Bathyplectes curculionis in Colorado. Environmental Entomology 8: 96100.Google Scholar
Singh, A & Nisbet, RM (2007) Semi-discrete host-parasitoid models. Journal of Theoretical Biology 247: 733742.Google Scholar
Sinzogan, AAC, Van Mele, P & Vayssieres, J-F (2008) Implications of on-farm research for local knowledge related to fruit flies and the weaver ant Oecophylla longinoda in mango production. International Journal of Pest Management 54: 241246.Google Scholar
Sisterson, MS, Biggs, RW, Manhardt, NM, Carriere, Y, Dennehy, T & Tabashnik, BE (2007) Effects of transgenic Bt cotton on insecticide use and abundance of two generalist predators. Entomologia Experimentalis et Applicata 124: 305311.Google Scholar
Sisterson, M & Tabashnik, BE (2005) Effects of transgenic Bt crops on specialist parasitoids of target pests. Environmental Entomology 34: 733742.Google Scholar
Sivakoff, FS, Rosenheim, JA & Hagler, JR (2012) Relative dispersal ability of a key agricultural pest and its predators in an annual agroecosystem. Biological Control 63: 296303.Google Scholar
Skalski, GT & Gilliam, JF (2001) Functional responses with predator interference: viable alternatives to the Holling Type II model. Ecology 82: 30833092.Google Scholar
Skirvin, DJ & Fenlon, JS (2003) The effect of temperature on the functional response of Phytoseiulus persimilis (Acari: Phytoseiidae). Experimental and Applied Acarology 31: 3749.Google Scholar
Skirvin, DJ, Perry, JN & Harrington, R (1997) A model describing the population dynamics of Sitobion avenae and Coccinella septempunctata. Ecological Modelling 96: 2939.Google Scholar
Skirvin, DJ, Williams, MED, Fenlon, JS & Sunderland, KD (2002) Modelling the effects of plant species on biocontrol effectiveness in ornamental nursery crops. Journal of Applied Ecology 39: 469480.Google Scholar
Skovgard, H & Nachma, G (2015a) Temperature dependent functional response of Spalangia cameroni (Perkins) (Hymenoptera: Pteromalidae), a parasitoid of Stomoxys calcitrans (L.) (Diptera: Muscidae). Environmental Entomology 44: 9099.Google Scholar
Skovgard, H & Nachma, G (2015b) Effect of mutual interference on the ability of Spalangia cameroni (Hymenoptera: Pteromalidae) to attack and parasitize pupae of Stomoxys calcitrans (Diptera: Muscidae). Environmental Entomology 44: 10761084.Google Scholar
Smith, CA & Gardiner, MM (2013) Biodiversity loss following the introduction of exotic competitors: does intraguild predation explain the decline of native lady beetles? PLoS One 8: e84448.Google Scholar
Smith, KP, Handelsman, J & Goodman, RM (1997) Modeling dose-response relationships in biological control: partitioning host responses to the pathogen and biocontrol agent. Phytopathology 87: 720729.Google Scholar
Smith, L (2014) Prediction of the geographic distribution of the psyllid, Arytinnis hakani (Homoptera: Psyllidae), a prospective biological control agent of Genista monspessulana, based on the effect of temperature on development, fecundity, and survival. Environmental Entomology 43: 13891398.Google Scholar
Smith, MC, Reeder, RH & Thomas, MB (1997) A model to determine the potential for biological control of Rottboellia cochinchinensis with the head smut Sporisorium ophiuri. Journal of Applied Ecology 34:388398.Google Scholar
Smith, RF & Van den Bosch, R (1967) Integrated control. In Pest Control: Biological, Physical, and Selected Chemical Methods (Kilgore, WW & Doutt, RL, eds.), New York, Academic Press, pp. 295340.Google Scholar
Smith, RH & Mead, R (1974) Age structure and stability in models of prey predator systems. Theoretical Population Biology 6: 308322.Google Scholar
Smith, SF & Krischik, VA (1999) Effects of systemic imidacloprid on Coleomegilla maculata (Coleoptera: Coccinellidae). Environmental Entomology 28: 11891195.Google Scholar
Smith, SM (1996) Biological control with Trichogramma: Advances, successes, and potential of their use. Annual Review of Entomology 41: 375406.Google Scholar
Smith, SM & You, M (1990) A life system simulation model for improving inundative releases of the egg parasite, Trichogramma minutum against the spruce budworm. Ecological Modelling 51: 123142.Google Scholar
Snow, AA, Andow, DA, Gepts, P, Hallerman, EM, Power, A, Tiedje, JM & Wolfenbarger, LL (2005) Genetically engineered organisms and the environment: current status and recommendations. Ecological Applications 15: 377404.Google Scholar
Snyder, WE, Chang, GC & Prasad, RP (2005) Conservation biological control: biodiversity influences the effectiveness of predators. In Ecology of Predator–Prey Interactions (Barbosa, P & Castellanos, I, eds.), Oxford, UK, Oxford University Press, pp. 324343.Google Scholar
Snyder, WE, Clevenger, GM & Eigenbrode, SD (2004) Intraguild predation and successful invasion by introduced ladybird beetles. Oecologia 140: 559565.Google Scholar
Sobey, WR (1969) Selection for resistance to myxomatosis in domestic rabbits (Oryctolagus cuniculus). Journal of Hygiene 67: 743754.Google Scholar
Sobey, WR & Conolly, D (1986) Myxomatosis: non-genetic aspects of resistance to myxomatosis in rabbits, Oryctolagus cuniculus. Australian Wildlife Research 13: 177187.Google Scholar
Solarz, SL & Newman, RM (2001) Variation in host plant preference and performance by the milfoil weevil, Euhrychiopsis lecontei Dietz, exposed to native and exotic watermilfoils. Oecologia 126: 6675.Google Scholar
Soldaat, LL & Auge, H (1998) Interactions between an invasive plant, Mahonia aquifolium, and a native phytophagous insect, Rhagoletis meigenii. In Plant Invasions: Ecological Mechanisms and Human Responses (Starfinger, U, Edwards, K, Kowarik, I & Williamson, M, eds.), Leiden, The Netherlands, Backhuys, pp. 347360.Google Scholar
Sotherton, NW (1985) The distribution and abundance of predatory Coleoptera overwintering in field boundaries. Annals of Applied Biology 106: 1721.Google Scholar
Sotherton, NW (1995) Beetle banks – helping nature to control pests. Pesticide Outlook 6: 1317.Google Scholar
Soti, PG & Volin, JC (2010) Does water hyacinth (Eichhornia crassipes) compensate for simulated defoliation? Biological Control 54: 3540.Google Scholar
Spadaro, D & Gullino, ML (2005) Improving the efficacy of biocontrol agents against soilborne pathogens. Crop Protection 24: 601613.Google Scholar
Speyer, ER (1927) An important parasite of the greenhouse whitefly (Trialeurodes vaporariorum Westwood). Bulletin of Entomological Research 17: 301308.Google Scholar
St. Leger, RJ (2007) Genetic modification for improvement of virulence of Metarhizium anisopliae as a microbial insecticide. In Biological Control: A Global Perspective (Vincent, C, Goettel, MS & Lazarovotis, G, eds.), Wallingford, UK, CABI Publishing, pp. 328335.Google Scholar
St. Leger, RJ, Joshi, L, Bidochka, MJ & Roberts, DW (1996) Construction of an improved mycoinsecticide overexpressing a toxic protease. Proceedings of the National Academy of Science USA 93: 63496354.Google Scholar
St. Leger, RJ & Wang, C (2010) Genetic engineering of fungal biocontrol agents to achieve greater efficacy against insect pests. Applied Microbiology and Biotechnology 85: 901907.Google Scholar
St. Martin, CCG & Brathwaite, RAI (2012) Compost and compost tea: principles and prospects as substrates and soil-borne disease management strategies in soil-less vegetable production. Biological Agriculture & Horticulture 28: 133.Google Scholar
Stadler, B & Dixon, AFG (2005) Ecology and evolution of aphid-ant interactions. Annual Review of Ecology Evolution and Systematics 36: 345372.Google Scholar
Stamp, NE & Bowers, MD (1990) Parasitism of New England buckmoth caterpillars (Hemileuca lucina: Saturniidae) by tachinid flies. Journal of the Lepidopterists’ Society 44: 199200.Google Scholar
Stanford, MM, Werden, SJ & McFadden, G (2007) Myxoma virus in the European rabbit: interactions between the virus and its susceptible host. Veterinary Research 38: 299318.Google Scholar
Stanley, JG (1976) Production of hybrid, androgenetic, and gynogenetic grass carp and carp. Transactions of the American Fisheries Society 105: 1016.Google Scholar
Stapel, JO, Cortesero, AM, De Moraes, CM, Tumlinson, JH & Lewis, WJ (1997) Extrafloral nectar, honeydew, and sucrose effects on searching behavior and efficiency of Microplitis. Environmental Entomology 26: 617623.Google Scholar
Stark, JD, Vargas, R & Banks, JE (2007) Incorporating ecologically relevant measures of pesticide effect for estimating the compatibility of pesticides and biocontrol agents. Journal of Economic Entomology 100: 10271032.Google Scholar
Stark, JD, Vargas, R & Miller, N (2004) Toxicity of spinosad in protein bait to three economically important tephritid fruit fly species (Diptera: Tephritidae) and their parasitoids (Hymenoptera: Braconidae). Journal of Economic Entomology 97: 911915.Google Scholar
Stary, P (1974) Taxonomy, origin, distribution and host range of Aphidius species (Hym., Aphidiidae) in relation to biological control of the pea aphid in Europe and North America. Zeitschrift fuer Angewandte Entomologie 77: 141171.Google Scholar
Stary, P, Lyon, JP & Leclant, F (1988) Biocontrol of aphids by the introduced Lysiphlebus testaceipes (Cress.)(Hym., Aphidiidae) in Mediterranean France. Journal of Applied Entomology 105: 7487.Google Scholar
Steiner, WWM & Teig, DA (1989) Microplitis croceipes (Cresson): genetic characterization and developing insecticide resistant biotypes. Southwestern Entomologist 12: 7187.Google Scholar
Steinger, T & Müller-Schärer, H (1992) Physiological and growth responses of Centaurea maculosa (Asteraceae) to root herbivory under varying levels of interspecific plant competition and soil nitrogen availability. Oecologia 91: 141149.Google Scholar
Steinhaus, EA (1956) Microbial control – the emergence of an idea: a brief history of insect pathology through the nineteenth century. Hilgardia 26: 107160.Google Scholar
Stelzl, M & Devetak, D (1999) Neuroptera in agricultural ecosystems. Agriculture Ecosystems & Environment 74: 305321.Google Scholar
Stephens, AE, Srivastava, DS & Myers, JH (2013) Strength in numbers? Effects of multiple natural enemy species on plant performance. Proceedings of the Royal Society of London B 280: 20122756.Google Scholar
Stephens, MJ, France, CM, Wratten, SD & Frampton, C (1998) Enhancing biological control of leafrollers (Lepidoptera: Tortricidae) by sowing buckwheat (Fagopyrum esculentum) in an orchard. Biocontrol Science and Technology 8: 547558.Google Scholar
Sterk, G, Hassan, SA, Baillod, M, Bakker, F, Bigler, F, Blumel, S, Bogenschutz, H, Boller, E, Bromand, B, Brun, J, Calis, JNM, Coremans-Pelseneer, J, Duso, C, Garrido, A, Grove, A, Heimbach, U, Hokkanen, H, Jacas, J, Lewis, G, Moreth, L, Polgar, L, Roversti, L, Samsoe-Peterson, L, Sauphanor, B, Schaub, L, Staubli, A, Tuset, JJ, Vainio, A, Van de Veire, M, Viggiani, G, Vinuela, E & Vogt, H (1999) Results of the seventh joint pesticide testing programme carried out by the IOBC/WPRS-Working Group ‘Pesticides and Beneficial Organisms’. BioControl 44: 99117.Google Scholar
Stern, VM & Van den Bosch, R (1959) Field experiments on the effects of insecticides. Hilgardia 29: 103130.Google Scholar
Steyaert, JM, Ridgway, HJ, Elad, Y & Stewart, A (2003) Genetic basis of mycoparasitism: a mechanism of biological control by species of Trichoderma. New Zealand Journal of Crop and Horticultural Science 31: 281291.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 (2002) Potential non-target impacts of a biological control agent, prickly pear moth, Cactoblastis cactorum (Berg) (Lepidoptera: Pyralidae) in North America, and possible management actions. Biological Invasions 4: 273281.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
Stiling, P, Moon, D & Gordon, D (2004) Endangered cactus restoration: mitigating the non-target effects of a biological control agent (Cactoblastis cactorum) in Florida. Restoration Ecology 12: 605610.Google Scholar
Stiling, P & Simberloff, D (2000) The frequency and strength of nontarget effects of invertebrate biological control agents of plant pests and weeds. In Nontarget Effects of Biological Control (Follett, PA & Duan, JJ, eds.), Dordrecht, The Netherlands, Kluwer, pp. 3144.Google Scholar
Stinner, RE (1977) Efficacy of inundative releases. Annual Review of Entomology 22: 515531.Google Scholar
Stireman, JO, Dyer, LA, Janzen, DH, Singer, MS, Lill, JT, Marquis, RJ, Ricklefs, RE, Hallwachs, W, Coley, PD, Barone, JA, Greeney, HF, Connahs, H, Barbosa, P & Morais, HC (2005) Climatic unpredictability and parasitism of caterpillars: implications of global warming. Proceedings of the National Academy of Sciences of the United States of America 102: 1738417387.Google Scholar
Stohlgren, TJ, Barnett, DT & Kartesz, J (2003) The rich get richer: patterns of plant invasions in the United States. Frontiers in Ecology and the Environment 1: 1114.Google Scholar
Stokes, JFG (1917) Notes on the Hawaiian rat. Occasional Papers of the Bishop Museum 3: 1121.Google Scholar
Story, G, Berman, D, Palmer, R & Scanlan, J (2004) The impact of rabbit haemorrhagic disease on wild rabbit (Oryctolagus cuniculus) populations in Queensland. Wildlife Research 31: 183193.Google Scholar
Story, JM (1995) Spotted knapweed. In Biological Control in the Western United States (Nechols, JR, ed.), Oakland, University of California Press, pp. 258263.Google Scholar
Story, JM, Good, KM, Harris, P & Nowierskyi, WR (1991) Metzneria paucipunctella Zeller Lepidoptera: Gelechidae), a moth introduced against spotted knapweed: its feeding strategy and impact on two introduced Urophora spp. (Diptera: Tephritidae). Canadian Entomologist 123: 10011007.Google Scholar
Story, JM & Nowierski, RM (1984) Increase and dispersal of Urophora affinis (Diptera: Tephritidae) on spotted knapweed in western Montana. Environmental Entomology 13: 11511156.Google Scholar
Story, JM, Smith, L, Corn, JG & White, LJ (2008) Influence of seed head-attacking biological control agents on spotted knapweed reproductive potential in western Montana over a 30-year period. Environmental Entomology 37: 510519.Google Scholar
Story, JM & Stougaard, RN (2006) Compatibility of two herbicides with Cyphocleonus achates (Coleoptera: Curculionidae) and Agapeta zoegana (Lepidoptera: Tortricidae), two root insects introduced for biological control of spotted knapweed. Environmental Entomology 35: 373378.Google Scholar
Stouthamer, R (1993) The use of sexual versus asexual wasps in biological control. Entomophaga 38: 36.Google Scholar
Stouthamer, R (2004) Sex-ratio distorters and other selfish genetic elements: implications for biological control. In Genetics, Evolution and Biological Control (Ehler, LE, Sforza, R & Mateille, T, eds.), Wallingford, UK, CABI Publishing, pp. 235252.Google Scholar
Stouthamer, R, Jochemsen, P, Platner, GR & Pinto, JD (2000) Crossing incompatibility between Trichogramma minutum and T. platneri (Hymenoptera: Trichogrammatidae): implications for application in biological control. Environmental Entomology 29: 832837.Google Scholar
Stouthamer, R, Luck, RF & Werren, JH (1992). Genetics of sex determination and improvement of biological control using parasitoids. Environmental Entomology 21: 427435.Google Scholar
Stouthamer, R, Russell, JE, Vavre, F & Nunney, L (2010) Intragenomic conflict in populations infected by parthenogenesis inducing Wolbachia ends with irreversible loss of sexual reproduction. BMC Evolutionary Biology 10: 229.Google Scholar
Strand, MR & Vinson, SB (1983) Factors affecting host recognition and acceptance in the egg parasitoid Telenomus heliothidis (Hymenoptera, Scelionidae). Environmental Entomology 12: 11141119.Google Scholar
Straub, CS, Finke, DL & Snyder, WE (2008) Are the conservation of natural enemy biodiversity and biological control compatible goals? Biological Control 45: 225237.Google Scholar
Strauss, SY & Agrawal, AA (1999) The ecology and evolution of plant tolerance to herbivory. Trends in Ecology and Evolution 14: 179185.Google Scholar
Strecker, RL, Marshall, JT Jr., Jackson, WB, Barbehenn, KR & Johnson, DH (1962) Pacific Island Rat Ecology. Honolulu, HI, Bishop Museum.Google Scholar
Strong, DR (1997) Fear no weevil? Science 277: 10581059.Google Scholar
Strong, DR, Lawton, JH & Southwood, TRE (1984) Insects on Plants. Cambridge, MA, Harvard University Press.Google Scholar
Strong, DR & Pemberton, RW (2000) Biological control of invading species-risk and reform. Science 288: 19691970.Google Scholar
Strong, DR & Pemberton, RW (2001) Food webs, risks of alien enemies and reform of biological control. In Evaluating Indirect Ecological Effects of Biological Control (Wajnberg, E, Scott, JK & Quimby, PC, eds.), Wallingford, UK, CABI Publishing, pp. 5780.Google Scholar
Stuart, RJ, Barbercheck, ME, Grewal, PS, Taylor, RAJ & Hoy, CW (2006) Population biology of entomopathogenic nematodes: concepts, issues, and models. Biological Control 38: 80102.Google Scholar
Stuart, RJ & Gaugler, R (1996) Genetic adaptation and founder effect in laboratory populations of the entomopathogenic nematode Steinernema glaseri. Canadian Journal of Zoology 74: 164170.Google Scholar
Styrsky, JD & Eubanks, MD (2010) A facultative mutualism between aphids and an invasive ant increases plant reproduction. Ecological Entomology 35: 190199.Google Scholar
Subramaniam, TSS, Lee, HL, Ahmad, NW & Murad, S (2012) Genetically modified mosquito: the Malaysian public engagement experience. Biotechnology Journal 7: 13231327.Google Scholar
Suckling, DM (2013) Benefits from biological control of weeds in New Zealand range from negligible to massive: a retrospective analysis. Biological Control 66: 2732.Google Scholar
Sullivan, DJ & Völkl, W (1999) Hyperparasitism: multitrophic ecology and behaviour. Annual Review of Entomology 44: 291315.Google Scholar
Summers, CG (1976) Population fluctuations of selected arthropods in alfalfa: influence of two harvesting practices. Environmental Entomology 5: 103110.Google Scholar
Sun, XL, W. Van der Werf, W, Bianchi, FJJA, Hu, ZH & Vlak, JM (2006) Modelling biological control with wild-type and genetically modified baculoviruses in the Helicoverpa armigera-cotton system. Ecological Modelling 198: 387398.Google Scholar
Sun, XL, Wang, HL, Sun, XC, Chen, XW, Peng, CM, Pan, DM, Jehle, JA, Van der Werf, W, Vlak, JM & Hu, ZH (2004) Biological activity and field efficacy of a genetically modified Helicoverpa armigera SNPV expressing an insect-selective toxin from a chimeric promoter. Biological Control 29: 124137.Google Scholar
Sutherst, RW (2005) Prediction of species geographical ranges. Journal of Biogeography 30: 805816.Google Scholar
Sutton, JC, Li, D-W, Peng, G, Yu, H, Zhang, P & Valdebenito-Sanhueza, RM (1997) Gliocladium roseum: a versatile adversary of Botrytis cinerea in crops. Plant Disease 81: 316328.Google Scholar
Swift, MJ, Vandermeer, J, Ramakrishnan, JM, Anderson, C, Ong, CK & Hawkins, BA (1996) Biodiversity and agroecosystem function. In Functional Roles of Biodiversity: A Global Perspective (Mooney, HA, Cushman, JH, Medina, E, Sala, OE & Schulze, E-D, eds.), Chichester, UK, Wiley, pp. 261298.Google Scholar
Swope, SM (2014) Biocontrol attack increases pollen limitation under some circumstances in the invasive plant Centaurea solstitialis. Oecologia 174: 205215.Google Scholar
Swope, SM & Satterthwaite, WH (2012) Variable effects of a generalist parasitoid on a biocontrol seed predator and its target weed. Ecological Applications 20: 2234.Google Scholar
Swope, SM & Stein, IR (2012) Soil type mediates indirect interactions between Centaurea solstitialis and its biocontrol agents. Biological Invasions 14: 16971710.Google Scholar
Symondson, WOC, Sunderland, KD & Greenstone, MH (2002) Can generalist predators be effective biological control agents? Annual Review of Entomology 47: 561594.Google Scholar
Symstad, AJ, Siemann, E & Haarstad, J (2000) An experimental test of the effect of plant functional group diversity on arthropod diversity. Oikos 89: 243253.Google Scholar
Szendrei, Z & Weber, DC (2009) Response of predators to habitat manipulation in potato fields. Biological Control 50: 123128.Google Scholar
Szűcs, M, Eigenbrode, SD, Schwarzlander, M & Schaffner, U (2012a) Hybrid vigor in the biological control agent, Longitarsus jacobaeae. Evolutionary Applications 5: 489497.Google Scholar
Szűcs, M, Schaffner, U, Price, WJ & Schwarzlaender, M (2012b) Post-introduction evolution in the biological control agent Longitarsus jacobaeae (Coleoptera: Chrysomelidae). Evolutionary Applications 5: 858868.Google Scholar
Tabashnik, BE (1994) Evolution of resistance to Bacillus thuringiensis. Annual Review of Entomology 39: 4779.Google Scholar
Tabashnik, BE & Johnson, MW (1999) Evolution of pesticide resistance in natural enemies. In Handbook of Biological Control (Bellows, TS & Fisher, TW, eds.), San Diego, CA, Academic Press, pp. 673689.Google Scholar
Takabayashi, J & Dicke, M (1996) Plant-carnivore mutualism through herbivore-induced carnivore attractants. Trends in Plant Science 1: 109113.Google Scholar
Takano-Lee, M & Hoddle, MS (2001) Biological control of Oligonychus perseae (Acari: Tetranychidae) on avocado: IV. Evaluating the efficacy of a modified mistblower to mechanically dispense Neoseiulus californicus (Acari: Phytoseiidae). International Journal of Acarology 27: 157169.Google Scholar
Tallamy, DW (2004) Do alien plants reduce insect biomass? Conservation Biology 18: 16891692.Google Scholar
Tanada, Y & Kaya, HK (1993) Insect Pathology. San Diego, CA, Academic Press.Google Scholar
Tanaka, S, Nishida, T & Ohsaki, N (2007) Sequential rapid adaptation of indigenous parasitoid wasps to the invasive butterfly Pieris brassicae. Evolution 61: 17911802.Google Scholar
Tarmann, GM (2004) Zygaenid Moths of Australia. Collingwood, Australia, CSIRO.Google Scholar
Tate, CD, Carpenter, JE & Bloem, S (2007) Influence of radiation dose on the level of F-1 sterility in the cactus moth, Cactoblastis cactorum (Lepidoptera: Pyralidae). Florida Entomologist 90: 537544.Google Scholar
Tauber, MJ, Tauber, CA, Daane, KM & Hagen, KS (2000) Commercialization of predators: recent lessons from green lacewings (Neuroptera: Chrysopidae: Chrysoperla). American Entomologist 46: 2638.Google Scholar
Taylor, LAV & Cruzon, MB (2015) Propagule pressure and disturbance drive the invasion of perennial false-brome (Brachypodium sylvaticum). Invasive Plant Science and Management 8: 169180.Google Scholar
Taylor, THC (1955) Biological control of insect pests. Annals of Applied Biology 42: 190196.Google Scholar
TeBeest, DO & Templeton, GE (1985) Mycoherbicides: progress in the biological control of weeds. Plant Disease 69: 610.Google Scholar
Tedders, WL & Schaefer, PW (1994) Release and establishment of Harmonia axyridis (Coleoptera: Coccinellidae) in the southeastern United States. Entomological News 105: 228243.Google Scholar
Templeton, GE, TeBeest, DO & Smith, RJ (1979) Biological weed control with mycoherbicides. Annual Review of Phytopathology 17: 301310.Google Scholar
Teplitski, M & Ritchie, K (2009) How feasible is the biological control of coral diseases? Trends in Ecology & Evolution 24: 378385.Google Scholar
Thébaud, C & Simberloff, D (2001) Are plants really larger in their introduced ranges? American Naturalist 157: 231236.Google Scholar
Thiel, A & Weeda, AC (2014) Haploid, diploid, and triploid-discrimination ability against polyploid mating partner in the parasitic wasp, Bracon brevicornis (Hymenoptera: Braconidae). Journal of Insect Science 14: 291.Google Scholar
Thiem, SM (1997) Prospects for altering host range for baculovirus bioinsecticides. Current Opinion in Biotechnology 8: 317322.Google Scholar
Thiem, SM, Du, X, Quentin, ME & Berner, MM (1996) Identification of a baculovirus gene that promotes Autographa californica nuclear polyhedrosis virus replication in a nonpermissive insect cell line. Journal of Virology 70: 22212229.Google Scholar
Thies, C, Steffan-Dewenter, I & Tscharntke, T (2003) Effects of landscape context on herbivory and parasitism at different spatial scales. Oikos 101: 1825.Google Scholar
Thies, C & Tscharntke, T (1999) Landscape structure and biological control in agroecosystems. Science 285: 893895.Google Scholar
Thomas, CD & Kunin, WE (1999) The spatial structure of populations. Journal of Animal Ecology 57: 536546.Google Scholar
Thomas, CFG, Parkinson, L, Griffiths, GJK, Garcia, AF & Marshall, EJP (2001a) Aggregation and temporal stability of carabid beetle distributions in field and hedgerow habitats. Journal of Applied Ecology 38: 100116.Google Scholar
Thomas, J (1980) Why did the large blue become extinct in Britain? Oryx 15 243247.Google Scholar
Thomas, JA (1995) The ecology and conservation of Maculinea arion and other European species of large blue butterfly. In Ecology and Conservation of Butterflies (Pullin, AS, ed.) London, UK, Chapman & Hall, pp. 180197.Google Scholar
Thomas, MB, Arthurs, SP & Watson, EL. 2006. Trophic and guild interactions and the influence of multiple species on disease. In Trophic and Guild Interactions in Biological Control (Brodeur, J & Boivin, G, eds.), Dordrecht, The Netherlands, Springer, pp. 101122.Google Scholar
Thomas, MB, Casula, P & Wilby, A (2004) Biological control and indirect effects. Trends in Ecology & Evolution 19: 61.Google Scholar
Thomas, MB & Reid, AM (2007) Are exotic natural enemies an effective way of controlling invasive plants? Trends in Ecology & Evolution 22: 447453.Google Scholar
Thomas, MB, Wratten, SD & Sotherton, NW (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, MB, Wratten, SD & Sotherton, NW (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, SR, Goulson, D & Holland, JM (2001b) Resource provision for farmland gamebirds: the value of beetle banks. Annals of Applied Biology 139: 111118.Google Scholar
Thomashow, LS & Weller, DM (1988) Role of a phenazine antibiotic from Pseudomonas fluorescens in biological control of Gaeumannomyces graminis var. tritici. Journal of Bacteriology 170: 34993508.Google Scholar
Thompson, JN (1998) Rapid evolution as an ecological process. Trends in Ecology & Evolution 13: 329332.Google Scholar
Thompson, JN & Burdon, JJ (1992) Gene-for-gene coevolution between plants and parasites. Nature 360: 121125.Google Scholar
Thomson, LJ & Hoffmann, AA (2007) Ecologically sustainable chemical recommendations for agricultural pest control? Journal of Economic Entomology 100: 17411750.Google Scholar
Thorbek, P & Bilde, T (2004) Reduced numbers of generalist arthropod predators after crop management. Journal of Applied Ecology 41: 526538.Google Scholar
Thorpe, KW (1985) Effects of height and habitat type on egg parasitism by Trichgramma minutum and T. pretiosum (Hymenoptera: Trichogrammatidae). Agriculture, Ecosystems & Environment 12: 117126.Google Scholar
Thrall, PH & Burdon, JJ (2004) Host-pathogen life history interactions affect biological control success. Weed Technology (Supplement) 18: 12691274.Google Scholar
Tilman, D (1997) Community invasibility, recruitment limitation, and grassland biodiversity. Ecology 78: 8192.Google Scholar
Tilman, D (2004) Niche tradeoffs, neutrality, and community structure: a stochastic theory of resource competition, invasion, and community assembly. Proceedings of the National Academy of Sciences, USA 101: 1085410861.Google Scholar
Tilman, D, Hill, J & Lehman, C (2006) Carbon-negative biofuels from low-input high-diversity grassland biomass. Science 314: 15981600.Google Scholar
Tilman, D, Knops, J, Wedin, D, Reich, P, Ritchie, M & Siemann, E (1997) The influence of functional diversity and composition on ecosystem processes. Science 277: 13001302.Google Scholar
Tilman, D, Reich, PB, Knops, J, Wedin, D, Mielke, T & Lehman, C (2001) Diversity and productivity in a long-term grassland experiment. Science 294: 843845.Google Scholar
Tilman, D, Wedin, D & Knops, J (1996) Productivity and sustainability influenced by biodiversity in grassland ecosystems. Nature 379: 718720.Google Scholar
Timper, P (2014) Conserving and enhancing biological control of nematodes. Journal of Nematology 46: 7589.Google Scholar
Tipping, PW, Martin, MR, Nimmo, KR, Pierce, RM, Smart, MD, White, E, Madeira, PT & Center, TD (2009) Invasion of a West Everglades wetland by Melaleuca quinquenervia countered by classical biological control. Biological Control 48: 7378.Google Scholar
Tisdell, CA (1990) Economic impact of biological control of weeds and insects. In Critical Issues in Biological Control (Mackauer, M, Ehler, LE & Roland, J, eds.), Andover, UK, Intercept, pp. 301316.Google Scholar
Tisdell, CA & Auld, BA (1990) Evaluation of biological control projects. In Proceedings of the VII International Symposium on Biological Control of Weeds (Delfosse, ES, ed.), Rome, Italy, Istituto Sperimentale per la Patalogia Vegetale, pp. 93100.Google Scholar
Toda, Y & Sakuratani, Y (2006) Expansion of the geographical distribution of an exotic ladybird beetle, Adalia bipunctata (Coleoptera: Coccinellidae), and its interspecific relationships with native ladybird beetles in Japan. Ecological Research 21: 292300.Google Scholar
Tomasino, SF, Leister, RT, Dimock, MB, Beach, RM & Kelly, JL (1995) Field performance of Clavibacter xyli subsp. cynodontis expressing the insecticidal protein gene cryia(c) of Bacillus thuringiensis against European corn borer in field corn. Biological Control 5: 442448.Google Scholar
Tonhasca, A & Byrne, DN (1994) The effects of crop diversification on herbivorous insects: a meta-analysis approach. Ecological Entomology 19: 239244.Google Scholar
Torchin, ME, Lafferty, KD, Dobson, AP, McKenzie, VJ & Kuris, AM (2003) Introduced species and their missing parasites. Nature 421: 628630.Google Scholar
Tothill, JD, Taylor, THC & Paine, RW (1930) The Coconut Moth in Fiji: A History of Its Control by Means of Parasites. London, UK, Imperial Bureau of Entomology.Google Scholar
Toyoshima, S & Amano, H (1998) Effect of prey density on sex ratio of two predacious mites, Phytoseiulus persimilis and Amblyseius womersleyi (Acari: Phytoseiidae). Experimental and Applied Acarology 22: 709723.Google Scholar
Trenbath, BR (1993) Intercropping for the management of pests and diseases. Field Crops Research 34: 381405.Google Scholar
Trillas, MI, Casanova, E, Cotxarrera, L, Ordovas, J, Borrero, C & Aviles, M (2006) Composts from agricultural waste and the Trichoderma asperellum strain T-34 suppress Rhizoctonia solani in cucumber seedlings. Biological Control 39: 3238.Google Scholar
Trumble, JT & Kok, LT (1982) Integrated pest management techniques in thistle suppression in pastures of North America. Weed Research 22: 345359.Google Scholar
Trumble, JT & Morse, JP (1993) Economics of integrating the predaceous mite Phytoseiulus persimilis (Acari: Phytoseiidae) with pesticides in strawberries. Journal of Economic Entomology 86: 879885.Google Scholar
Trumper, EV & Holt, J (1998) Modelling pest population resurgence due to recolonization of fields following an insecticide application. Journal of Applied Ecology 35: 273285.Google Scholar
Tscharntke, T, Bommarco, R, Clough, Y, Crist, TO, Kleijn, D, Rand, TA, Tylianakis, J, van Nouhuys, S & Vidal, S (2007) Conservation biological control and enemy diversity on a landscape scale. Biological Control 43: 294309.Google Scholar
Tscharntke, T, Klein, AM, Kruess, A, Steffan-Dewenter, I & Thies, C (2005a) Landscape perspectives on agricultural intensification and biodiversity – ecosystem service management. Ecology Letters 8: 857874.Google Scholar
Tscharntke, T, Rand, TA & Bianchi, FJJA (2005b) The landscape context of trophic interactions: insect spillover across the crop-noncrop interface. Annales Zoologici Fennici 42: 421432.Google Scholar
Tscharntke, T, Sekercioglu, CH, Dietsch, TV, Sodhi, NS, Hoehn, P & Tylianakis, JM (2008) Landscape constraints on functional diversity of birds and insects in tropical agroecosystems. Ecology 89: 944951.Google Scholar
Tsuchida, T, Koga, R & Fukatsu, T (2004) Host plant specialization governed by facultative symbiont. Science 303: 1989.Google Scholar
Tullett, AG, Hart, AJ, Worland, MR & Bale, JS (2004) Assessing the effects of low temperature on the establishment potential in Britain of the non-native biological control agent Eretmocerus eremicus. Physiological Entomology 29: 363371.Google Scholar
Turchin, P (1995) Population regulation: old arguments and a new synthesis. In Population Dynamics: New Approaches and Synthesis (Cappuccino, N. & Price, PW, eds.), San Diego, CA, Academic Press, pp. 1940.Google Scholar
Turelli, M (2010) Cytoplasmic incompatibility in populations with overlapping generations. Evolution 64: 232241.Google Scholar
Turnbull, LA, Crawley, MJ & Rees, M (2000) Are plant populations seed limited? A review of seed sowing experiments. Oikos 88: 225238.Google Scholar
Turner, CE (1985) Conflicting interests in biological control of weeds. In Proceedings of the VI International Symposium of Biological Control of Weeds (Delfosse, ES, ed.), Ottawa, Canada, Agriculture Canada, pp. 203225.Google Scholar
Turnock, WJ, Wise, IL & Matheson, FO (2003) Abundance of some native coccinellids (Coleoptera: Coccinellidae) before and after the appearance of Coccinella septempunctata. Canadian Entomologist 135: 391404.Google Scholar
Twohey, MB, Heinrich, JW, Seelye, JG, Fredricks, KT, Bergstedt, RA, Kaye, CA, Scholefield, RJ, McDonald, RB & Christie, CG (2003) The sterile-male-release technique in Great Lakes sea lamprey management. Journal of Great Lakes Research 29: 410423.Google Scholar
Tylianakis, JM, Didham, R & Wratten, SD (2004) Improved fitness of aphid parasitoids receiving resource subsidies. Ecology 85: 658666.Google Scholar
Tyndale-Biscoe, CH (1994) Virus-vectored immunocontraception of feral mammals. Reproduction Fertility and Development 6: 281287.Google Scholar
Tyndale-Biscoe, M & Vogt, WG (1996) Population status of the bush fly, Musca vetustissima (Diptera: Muscidae), and native dung beetles (Coleoptera: Scarabaeinae) in south-eastern Australia in relation to establishment of exotic dung beetles. Bulletin of Entomological Research 86: 183192.Google Scholar
Unruh, T, White, W, Gonzalez, D, Gordh, G & Luck, RF (1983) Heterozygosity and effective size in laboratory populations of Aphidius ervi (Hym.: Aphidiidae). Entomophaga 28: 245258.Google Scholar
Urano, S, Shima, K, Hongo, K & Suzuki, Y (2003) A simple criterion for successful biological control on annual crops. Population Ecology 45: 97103.Google Scholar
USDA (2003) Reviewer’s Manual for the Technical Advisory Group for Biological Control Agents of Weeds: Guidelines for Evaluating the Safety of Candidate Biological Control Agents. Washington, DC, USDA APHIS.Google Scholar
US EPA (1998) Guidelines for Ecological Risk Assessment. Washington, DC, Science Advisory Board.Google Scholar
US National Academy of Sciences (1988) Report of the Research Briefing on Biological Control in Managed Ecosystems, Washington, DC, National Academy of Sciences.Google Scholar
Usher, MB (1988) Biological invasions of nature reserves – a search for generalization. Biological Conservation 44: 119135.Google Scholar
Van Bael, SA, Philpott, SM, Greenberg, R, Bichier, P, Barber, NA, Mooney, KA & Gruner, DS (2008) Birds as predators in tropical agroforestry systems. Ecology 89: 928934.Google Scholar
Van den Bosch, R (1971) Biological control of insects. Annual Review of Ecology and Systematics 2: 4566.Google Scholar
Van den Bosch, R & Dietrick, EJ (1959) The interrelationships of Hypera brunneipennis (Coleoptera: Curculionidae) and Bathyplectes curculionis Hymenoptera: Ichneumonidae) in Southern California. Annals of the Entomological Society of America 52: 609616.Google Scholar
Van den Bosch, R, Hom, R, Matteson, P, Frazer, BD, Messenger, PS & Davis, CS (1979) Biological control of the walnut aphid in California: impact of the parasite, Trioxys pallidus. Hilgardia 47: 113.Google Scholar
Van den Bosch, R, Messenger, PS & Gutierrez, AP (1982) An Introduction to Biological Control. New York, Plenum.Google Scholar
Van den Bosch, R & Stern, VM (1969) The effect of harvesting practices on insect populations in alfalfa. Proceedings of the Tall Timbers Conference on Ecological Animal Control by Habitat Management 3: 4754.Google Scholar
Van den Bosch, R & Telford, AD (1964) Environmental modification and biological control. In Biological Control of Insect Pests and Weeds (DeBach, P, ed.), London, UK, Chapman & Hall, pp. 459488.Google Scholar
Van Driesche, RG & Bellows, TS (1993) Steps in Classical Arthropod Biological Control. Annapolis, MD, Entomological Society of America.Google Scholar
Van Driesche, RG & Bellows, TS (1996) Biological Control. New York, Chapman & Hall.Google Scholar
Van Driesche, RG, Carruthers, RI, Center, T, Hoddle, MS, Hough-Goldstein, J, Morin, L, Smith, L, Wagner, DL, Blossey, B, Brancatini, V, Casagrande, R, Causton, CE, Coetzee, JA, Cudam, J, Ding, J, Fowler, SV, Frankm, JH, Fuester, R, Goolsby, J, Grodowitz, M, Heard, TA, Hill, MP, Hoffmann, JH, Huber, J, Julien, M, Kairo, MTK, Kenis, M, Mason, P, Medalm, J, Messing, R, Miller, R, Moore, A, Neuenschwander, P, Newman, R, Norambuena, H, Palmer, WA, Pemberton, R, Perez Panduro, A, Pratt, PD, Rayamajhi, M, Salom, S, Sands, D, Schooler, S, Schwarzländer, M, Sheppard, A, Shaw, R, Tipping, PW & van Klinken, RD (2010) Classical biological control for the protection of natural ecosystems. Biological Control 54: S2S33.Google Scholar
Van Driesche, RG & Hoddle, M (1997) Should arthropod parasitoids and predators be subject of host range testing when used as biological control agents? Agriculture and Human Values 14: 211226.Google Scholar
Van Driesche, RG, Hoddle, M & Center, TD (2008) Control of Pests and Weeds by Natural Enemies: An Introduction to Biological Control. Hoboken, NJ, Wiley-Blackwell.Google Scholar
Van Driesche, RG & Murray, TJ (2004) Overview of testing schemes and designs used to estimate host ranges. In Assessing Host Ranges for Parasitoids and Predators Used for Classical Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, USDA Forest Service, pp. 6889.Google Scholar
Van Driesche, RG, Nunn, C, Kreke, N, Goldstein, B & Benson, J (2003) Laboratory and field host preferences of introduced Cotesia spp. parasitoids (Hymenoptera: Braconidae) between native and invasive Pieris butterflies. Biological Control 28: 214221.Google Scholar
Van Driesche, RG, Nunn, C & Pasquale, A (2004) Life history pattern, host plants, and habitat as determinants of population survival of Pieris napi oleracea interacting with an introduced braconid parasitoid. Biological Control 29: 278287.Google Scholar
Van Driesche, R & Reardon, R (2004) Assessing Host Ranges for Parasitoids and Predators Used for Classical Biological Control: A Guide to Best Practice. Morgantown, WV, USDA Forest Service.Google Scholar
Van Eenennaam, JP, Stocker, RK, Thiery, RG, Hagstrom, NT & Doroshov, SI (1990) Egg fertility, early development and survival from crosses of diploid female x triploid male grass carp (Ctenopharyngodon idella). Aquaculture 86: 111125.Google Scholar
Van Emden, HF (1965) The role of uncultivated land in the biology of crop pests and beneficial insects. Scientific Horticulture 17: 121136.Google Scholar
Van Emden, HF (1990) Plant diversity and natural enemy efficiency in agroecosystems. In Critical Issues in Biological Control (Mackauer, M, Ehler, LE & Roland, J, eds.), Andover, UK, Intercept, pp. 6380.Google Scholar
Van Emden, HF & Williams, GF (1974) Insect stability and diversity in agro-ecosystems. Annual Review of Entomology 19: 455475.Google Scholar
Van Hamburg, H & Hassell, MP (1984) Density dependence and the augmentative release of egg parasitoids against graminaceous stalk borers. Ecological Entomology 9: 101108.Google Scholar
Van Kleunen, M & Schmid, B (2003) No evidence for an evolutionary increased competitive ability in an invasive plant. Ecology 84: 28162823.Google Scholar
Van Klinken, RD & Edwards, OR (2002) Is host-specificity of weed biological control agents likely to evolve rapidly following establishment? Ecology Letters 5: 590596.Google Scholar
Van Klinken, RD, Fichera, G & Cordo, H (2003) Targeting biological control across diverse landscapes: the release, establishment, and early success of two insects on mesquite (Prosopis spp.) insects in Australian rangelands. Biological Control 26: 820.Google Scholar
Van Lenteren, JC (2000) Success in biological control of arthropods by augmentation of natural enemies. In Biological Control: Measures of Success (Gurr, G & Wratten, SD, eds.), Dordrecht, The Netherlands, Kluwer, pp. 77103.Google Scholar
Van Lenteren, JC (2003) Quality Control and Production of Biological Control Agents: Theory and Testing Procedures. Wallingford, UK, CABI Publishing.Google Scholar
Van Lenteren, JC (2006) How not to evaluate augmentative biological control. Biological Control 39: 115118.Google Scholar
Van Lenteren, JC (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, JC, Babendreier, D, Bigler, F, Burgio, G, Hokkanen, HMT, Kuske, S, Loomans, AJM, Menzler-Hokkanen, I, Van Rijn, PCJ, Thomas, MB, Tommasini, MG & Zeng, QQ (2003) Environmental risk assessment of exotic natural enemies used in inundative biological control. BioControl 48: 338.Google Scholar
Van Lenteren, JC, Bale, J, Bigler, F & Hokkanen, HMT (2006a) Assessing risks of releasing exotic biological control agents of arthropod pests. Annual Review of Entomology 51: 609634.Google Scholar
Van Lenteren, JC, Cock, MJW, Hoffmeister, TS & Sands, DPA (2006b) Host specificity in arthropod biological control, methods for testing and interpretation of the data. In Environmental Impact of Invertebrates for Biological Control of Arthropods: Methods and Risk Assessment (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 3863.Google Scholar
Van Lenteren, JC & Loomans, AJM (2006) Environmental risk assessment: methods for comprehensive evaluation and quick scan. In Environmental Impact of Invertebrates for Biological Control of Arthropods: Methods and Risk Assessment (Bigler, F, Babendreier, D & Kuhlmann, U, eds.), Wallingford, UK, CABI Publishing, pp. 254272.Google Scholar
Van Lenteren, JC & Tommasini, MG (2003) Mass production, storage, shipment and release of natural enemies. In Quality Control and Production of Biological Control Agents: Theory and Testing Procedures (van Lenteren, JC, ed.), Wallingford, UK, CABI Publishing, pp. 181189.Google Scholar
Van Lenteren, JC & Woets, J (1988) Biological and integrated pest control in greenhouses. Annual Review of Entomology 33: 239270.Google Scholar
Van Mele, P (2008) A historical review of research on the weaver ant Oecophylla in biological control. Agricultural and Forest Entomology 10: 1322.Google Scholar
Van Mele, P & Chien, HV (2004) Farmers, biodiversity and plant protection: developing a learning environment for sustainable tree cropping systems. International Journal of Agricultural Sustainability 2: 6776.Google Scholar
Van Mele, P & Cuc, NTT (2000) Evolution and status of Oecophylla smaragdina (Fabricius) as a pest control agent in citrus in the Mekong Delta, Vietnam. International Journal of Pest Management 46: 295301.Google Scholar
Van Mele, P, Vayssieres, J-F, Van Tellingen, E & Vrolijks, J (2007) Effects of an African weaver ant, Oecophylla longinoda, in controlling mango fruit flies (Diptera: Tephritidae) in Benin. Journal of Economic Entomology 100: 695701.Google Scholar
Van Rijn, PCJ & Sabelis, MW (2005) Impact of plant-provided food on herbivore-carnivore dynamics. In Plant-Provided Food for Carnivorous Insects: A Protective Mutualism and Its Applications (Wäckers, FL, Van Rijn, PCJ & Bruin, J, eds.), Cambridge, UK, Cambridge University Press, pp. 223266.Google Scholar
Van Valen, L (1973) A new evolutionary law. Evolutionary Theory 1: 130.Google Scholar
Van Veen, FJF, Memmott, J & Godfray, HCJ (2006) Indirect effects, apparent competition and biological control. In Trophic and Guild Interactions in Biological Control (Brodeur, J & Boivin, G, eds.), Dordrecht, The Netherlands, Springer, pp. 145170.Google Scholar
Van Wilgenberg, E, Driessen, G & Beukeboom, LW (2006) Single locus complementary sex determination in Hymenoptera: an “unintelligent” design? Frontiers in Zoology 3:1.Google Scholar
Vandergheynst, J, Scher, H, Guo, H-Y & Schultz, D (2007) Water-in-oil emulsions that improve the storage and delivery of the biolarvacide Lagenidium giganteum. BioControl 52: 207229.Google Scholar
Vanderplank, FL (1960) The bionomics and ecology of the red tree ant, Oecophylla sp. and its relationship to the coconut bug, Pseudotheraptus wayi (Coreidae). Journal of Animal Ecology 29: 1533.Google Scholar
Vargas, RI, Miller, NW & Prokopy, RJ (2002) Attraction and feeding responses of Mediterranean fruit fly and a natural enemy to protein baits laced with two novel toxins, phloxine B and spinosad. Entomologia Experimentalis et Applicata 102: 273282.Google Scholar
Vargas, RI, Peck, SL, McQuate, GT, Jackson, CG, Stark, JD & Armstrong, JW (2001) Potential for areawide integrated management of Mediterranean fruit fly (Diptera: Tephritidae) with a braconid parasitoid and a novel bait spray. Journal of Economic Entomology 94: 817825.Google Scholar
Varley, GC (1959) The biological control of agricultural pests. Journal of the Royal Society of Arts 107: 475490.Google Scholar
Varley, GC, Gradwell, GR & Hassel, MP (1973) Insect Population Ecology: An Analytical Approach. Berkeley, University of California Press.Google Scholar
Vasquez, CJ, Stouthamer, R, Jeong, G & Morse, JG (2011) Discovery of a CI-inducing Wolbachia and its associated fitness costs in the biological control agent Aphytis melinus DeBach (Hymenoptera: Aphelinidae). Biological Control 58: 192198.Google Scholar
Vega, FE, Dowd, PF, Lacey, LA, Pell, JK, Jackson, DM & Klein, MG (2007) Dissemination of beneficial microbial agents by insects. In Field Manual of Techniques in Invertebrate Pathology: Application and Evaluation of Pathogens for Control of Insects and Other Invertebrate Pests, 2nd edition (Lacey, LA & Kaya, HK, eds.), Dordrecht, The Netherlands, Springer, pp. 127146.Google Scholar
Veldtman, R, Lado, TF, Botes, A, Proches, S, Timm, AE, Geertsema, H & Chown, SL (2011) Creating novel food webs on introduced Australian acacias: indirect effects of galling biological control agents. Diversity and Distributions 17: 958967.Google Scholar
Venette, RC, Kriticos, DJ, Magarey, RD, Koch, FH, Baker, RHA, Worner, SP, Raboteaux, NNG, McKenney, DW, Dobesberger, EJ, Yemshanov, D, De Barro, PJ, Hutchison, WD, Fowler, G, Kalaris, TM & Pedlar, J (2010) Pest risk maps for invasive alien species: a roadmap for improvement. BioScience 60: 349362.Google Scholar
Verhulst, PF (1838) Notice sur la loi que la population suit dans son accroissement. Correspondance Mathématique et Physique 10: 113121.Google Scholar
Vetukhiv, M (1956) Fecundity of hybrids between geographic populations of Drosophila pseudoobscura. Evolution 10: 139146.Google Scholar
Vetukhiv, M (1957) Longevity of hybrids between geographic populations of Drosophila subobscura. Evolution 11: 348360.Google Scholar
Vietch, CR & Clout, MN (2001) Human dimensions in the management of invasive species in New Zealand. In The Great Reshuffling: Human Dimensions of Invasive Alien Species (McNeely, JA, ed.), Gland, Switzerland, pp. 6374.Google Scholar
Vigueras, AL & Portillo, L (2001) Uses of Opuntia species and the potential impact of Cactoblastis cactorum (Lepidoptera: Pyralidae) in Mexico. Florida Entomologist 84: 493498.Google Scholar
Vilá, M & Gimeno, C (2003) Seed predation of two alien Opuntia species invading Mediterranean communities. Plant Ecology 167: 18.Google Scholar
Vilá, M, Gomez, A & Maron, JL (2003) Are alien plants more competitive than their native conspecifics? A test using Hypericum perforatum L. Oecologia 137: 211215.Google Scholar
Villareal, LP (2004) Are viruses alive? Scientific American 291: 100105.Google Scholar
Vitousek, PM (1990) Biological invasions and ecosystem processes: towards an integration of population biology and ecosystem studies. Oikos 57: 713.Google Scholar
Volterra, V (1926) Fluctuations in the abundance of a species considered mathematically. Nature 118: 558560.Google Scholar
Vorley, WT & Wratten, SD (1987) Migration of parasitoids (Hymenoptera: Braconidae) of cereal aphids (Hemiptera: Aphididae) between grassland, early-sown cereals and late-sown cereals in southern England. Bulletin of Entomological Research 77: 555568.Google Scholar
Vorsino, AE, Wieczorek, AM, Wright, MG & Messing, RH (2012) Using evolutionary tools to facilitate the prediction and prevention of host-based differentiation in biological control: a review and perspective. Annals of Applied Biology 160: 204216.Google Scholar
Vos, M & Hemerik, L (2003) Linking foraging behavior to lifetime reproductive succes for an insect parasitoid: adaptation to host distributions. Behavioral Ecology 14: 236245.Google Scholar
Vos, M & Vet, LEM (2004) Geographic variation in host acceptance by an insect parasitoid: genotype versus experience. Evolutionary Ecology Research 6: 10211035.Google Scholar
Waage, JK (1990) Ecological theory and the selection of biological control agents. Critical Issues in Biological Control (Ehler, LE & Roland, J, eds.), Andover, UK, Intercept, pp. 135157.Google Scholar
Waage, JK & Greathead, DJ (1988) Biological control: challenges and opportunities. Philosophical Transactions of the Royal Society B 318: 111128.Google Scholar
Waage, JK, Hassell, MP & Godfray, HCJ (1985) The dynamics of pest parasitoid insecticide interactions. Journal of Applied Ecology 22: 825838.Google Scholar
Wäckers, FL, Van Rijn, PCJ & Bruin, J (2005) Plant-Provided Food for Carnivorous Insects. Cambridge, UK, Cambridge University Press.Google Scholar
Wäckers, F, Van Rijn, PCJ & Heimpel, GE (2008) Honeydew as a food source for natural enemies: making the best of a bad meal? Biological Control 45: 176184.Google Scholar
Wade, MR, Gurr, GM & Wratten, SD (2008a) Ecological restoration of farmland: progress and prospects. Philosophical Transactions of the Royal Society B 363: 831847.Google Scholar
Wade, MR, Zalucki, MP, Wratten, SD & Robinson, KA (2008b) Conservation biological control of arthropods using artificial food sprays: current status and future challenges. Biological Control 45: 185199.Google Scholar
Waite, GK (1988) Integrated control of Tetranychus urticae in strawberries in southeast Queensland, Australia. Experimental and Applied Acarology 5: 2332.Google Scholar
Wajnberg, E (2004) Measuring genetic variation in natural enemies used for biological control: why and how? In Genetics, Evolution and Biological Control (Ehler, LE, Sforza, R & Mateille, T, eds.), Wallingford, UK, CABI Publishing, pp. 1938.Google Scholar
Wajnberg, E, Roitberg, BD & Boivin, G (2016) Using optimality models to improve the efficacy of parasitoids in biological control programmes. Entomologia Experimentalis et Applicata 158: 216.Google Scholar
Wajnberg, E, Scott, JK & Quimby, PC (2001) Evaluating Indirect Ecological Effects of Biological Control. Wallingford, UK, CABI Publishing.Google Scholar
Walde, SJ & Nachman, G (1999) Dynamics of spatially structured spider mite populations. In Theoretical Approaches to Biological Control (Hawkins, BA & Cornell, HV, eds.), Cambridge, UK, Cambridge University Press, pp. 163189.Google Scholar
Walker, GP, Herman, TJB, Kale, AJ & Wallace, AR (2010) An adjustable action threshold using larval parasitism of Helicoverpa armigera (Lepidoptera: Noctuidae) in IPM for processing tomatoes. Biological Control 52: 3036.Google Scholar
Wang, B, Ferro, DN & Hosmer, DW (1999) Effectiveness of Trichogramma ostriniae and T. nubilale for controlling the European corn borer Ostrinia nubilalis in corn. Entomologia Experimentalis et Applicata 91: 297303.Google Scholar
Wang, CS & St. Leger, RJ (2007) A scorpion neurotoxin increases the potency of a fungal insecticide. Nature Biotechnology 25: 14551456.Google Scholar
Wang, X & Grewal, PS (2002) Rapid genetic deterioration of environmental tolerance and reproductive potential of an entomopathogenic nematode during laboratory maintenance. Biological Control, 23: 7178.Google Scholar
Wang, X-G, Duff, J, Keller, MA, Zalucki, MP, Liu, S-S & Bailey, P (2004) Role of Diadegma semiclausum (Hymenoptera: Ichneumonidae) in controlling Plutella xylostella (Lepidoptera: Plutellidae): Cage exclusion experiments and direct observation. Biocontrol Science and Technology 14: 571586.Google Scholar
Wang, Y, Van Oers, MM, Crawford, AM, Vlak, JM & Jehle, JA (2007) Genomic analysis of Oryctes rhinocerus virus reveals genetic relatedness to Heliothis zea virus 1. Archives of Virology 152: 519531.Google Scholar
Wang, YH & Gutierrez, AP (1980) An assessment of the use of stability analyses in population ecology. Journal of Animal Ecology 49: 435452.Google Scholar
Wanger, TC, Wielgoss, AC, Motzke, I, Clough, Y, Brook, BW, Sodhi, NS & Tscharnkte, T (2011) Endemic predators, invasive prey and native diversity. Proceedings of the Royal Society B-Biological Sciences 278: 690694.Google Scholar
Wanner, H, Gu, H, Hattendorf, B, Gunther, D & Dorn, S (2006) Using the stable isotope marker 44Ca to study dispersal and host-foraging activity in parasitoids. Journal of Applied Ecology 43: 10311039.Google Scholar
Wapshere, AJ (1974) A strategy for evaluating the safety of organisms for biological weed control. Annals of Applied Biology 77: 201211.Google Scholar
Wapshere, AJ (1989) A testing sequence for reducing rejection of potential biological control agents for weeds. Annals of Applied Biology 114: 515526.Google Scholar
Ward-Fear, G, Brown, GP & Shine, R (2010) Using a native predator (the meat ant, Iridomyrmex reburrus) to reduce the abundance of an invasive species (the cane toad, Bufo marinus) in tropical Australia. Journal of Applied Ecology 47: 273280.Google Scholar
Wardle, DA (1987) The ecology of ragwort (Senecio jacobaea L.) – A review. New Zealand Journal of Ecology 10: 6776.Google Scholar
Warner, KD (2012) Fighting pathophobia: how to construct constructive public engagement with biocontrol for nature without augmenting public fears. BioControl 57: 307317.Google Scholar
Watanabe, S, Kato, H, Kumakura, K, Ishibashi, E & Nagayama, K (2006) Properties and biological control activities of aerial and submerged spores in Trichoderma asperellum SKT-1. Journal of Pesticide Science 31: 375379.Google Scholar
Waterhouse, DF (1974) The biological control of dung. Scientific American 230: 100109.Google Scholar
Waterhouse, DF & Norris, KR (1987) Biological Control: Pacific Prospects. Melbourne, Australia, Inkata.Google Scholar
Watkinson, AR (1980) Density dependence in single species populations of plants. Journal of Theoretical Biology 83: 345358.Google Scholar
Watt, KEF (1959) A mathematical model for the effect of densities of attacked and attacking species on the number attacked. Canadian Entomologist 91: 129144.Google Scholar
Watt, MS, Whitehead, D, Kriticos, DJ, Gous, SF & Richardson, B (2007) Using a process-based model to analyse compensatory growth in response to defoliation: simulating herbivory by a biological control agent. Biological Control 43: 119129.Google Scholar
Way, MJ (1951) An insect pest of coconuts and its relationship to certain ant species. Nature 168: 302.Google Scholar
Way, MJ (1966) The natural environment and integrated methods of pest control. Journal of Applied Ecology 3 (supplement): 2932.Google Scholar
Way, MJ, Cammell, ME & Paiva, MS (1992) Studies on egg predation by ants (Hymenoptera: Formicidae) especially on the eucalyptus borer Phoracantha semipunctata (Coleoptera: Cerambycidae). Bulletin of Entomological Research 82: 425432.Google Scholar
Way, MJ, Paiva, MR & Cammell, ME (1999) Natural biological control of the pine processionary moth Thaumetopoea pityocampa (Den. & Schiff.) in the Argentine ant Linepithema humile (Mayr) in Portugal. Agricultural and Forest Entomology 1: 2731.Google Scholar
Weed, AS & Schwarzlander, M (2014) Density dependence, precipitation and biological control agent herbivory influence landscape-scale dynamics of the invasive Eurasian plant Linaria dalmatica. Journal of Applied Ecology 51: 825834.Google Scholar
Weisser, WW, Jansen, VAA & Hassell, MP (1997) The effects of a pool of dispersers on host-parasitoid systems. Journal of Theoretical Biology 189: 413425.Google Scholar
Welch, KD & Harwood, JD (2014) Temporal dynamics of natural enemy–pest interactions in a changing environment. Biological Control 75: 1827.Google Scholar
Weller, DM (1988) Biological control of soilborne pathogens in the rhizosphere with bacteria. Annual Review of Phytopathology 26: 379407.Google Scholar
Weller, DM & Cook, RJ (1983) Suppression of take-all of wheat by seed treatments with fluorescent pseudomonads. Phytopathology 73: 463469.Google Scholar
Weller, DM, Landa, BB, Mavrodi, OV, Schroeder, KL, De La Fuente, L, Bankhead, SB, Molar, RA, Bonsall, RF, Mavrodi, DV & Thomashow, LS (2007) Role of 2,4-diacetylphloroglucinol-producing fluorescent Pseudomonas spp. in the defense of plant roots. Plant Biology 9: 420.Google Scholar
Weller, DM, Raaijmakers, JM, Gardener, BBM & Thomashow, LS (2002) Microbial populations responsible for specific soil suppressiveness to plant pathogens. Annual Review of Phytopathology 40: 309348.Google Scholar
Werling, BP & Gratton, C (2008) Influence of field margins and landscape context on ground beetle diversity in Wisconsin (USA) potato fields. Agriculture Ecosystems & Environment 128: 104108.Google Scholar
Weseloh, RM (1982) Implications of tree microhabitat preferences of Compsilura concinnata (Diptera: Tachinidae) for its effectiveness as a gypsy moth parasitoid. The Canadian Entomologist 114: 617622.Google Scholar
Weseloh, RM (1984) Effect of size, stress, and ligation of gypsy moth (Lepidoptera: Lymantriidae) larvae on development of the tachinid parasite Compsilura concinnata Meigen (Diptera: Tachinidae). Annals of the Entomological Society of America 77: 423428.Google Scholar
Wheeler, AG & Stoops, CA (1996) Status and spread of the Palearctic lady beetles Hippodamia variegate and Propylea quatordecimpunctata (Coleoptera: Coccinellidae) in Pennsylvania, 1993–1995. Entomological News 107: 291298.Google Scholar
Whipps, JM (2001) Microbial interactions and biocontrol in the rhizosphere. Journal of Experimental Botany 52: 487511.Google Scholar
Whipps, JM (2004) Prospects and limitations for mycorrhizas in biocontrol of root pathogens. Canadian Journal of Botany 82: 11981227.Google Scholar
Whipps, JM & McQuilken, MP (2009) Biological control agents in plant disease control. In Disease Control in Crops: Biological and Environmentally Friendly Approaches (Walters, D, ed.), Oxford, UK, Wiley, pp. 2761.Google Scholar
White, PJ, Norman, RA & Hudson, PJ (2002) Epidemiological consequences of a pathogen having both virulent and avirulent modes of transmission: the case of rabbit haemorrhagic disease virus. Epidemiology and Infection 129: 665677.Google Scholar
Whitham, TG, Maschinski, J, Larson, KC & Paige, KN (1991) Plant responses to herbivory: the continuum from negative to positive and underlying physiological mechanisms. In Plant-Animal Interactions: Evolutionary Ecology in Tropican and Temperate Regions (Price, PW, Lewinsohn, TM, Fernandes, GW & Benson, WW, eds.), Hoboken, NJ, Wiley, pp. 227255.Google Scholar
Whitham, TG, Young, WP, Martinsen, GD, Gehring, CA, Schweitzer, JA, Shuster, SM, Wimp, GM, Fischer, DG, Bailey, JK, Lindroth, RL, Woolbright, S & Kuske, CR (2003) Community and ecosystem genetics: a consequence of the extended phenotype. Ecology 84: 559573.Google Scholar
Wiebe, AP & Obrycki, JJ (2002) Prey suitability of Galerucella pusilla eggs for two generalist predators, Coleomegilla maculata and Chrysoperla carnea. Biological Control 23: 143148.Google Scholar
Wiebe, AP & Obrycki, JJ (2004) Quantitative assessment of predation of eggs and larvae of Galerucella pusilla in Iowa. Biological Control 31: 1628.Google Scholar
Wiedenmann, RN & Smith, JW Jr. (1997) Novel assocations and importation biological control: the need for ecological and physiological equivalencies. International Journal of Tropical Insect Science 17: 5160.Google Scholar
Wiens, JJ & Graham, CH (2005) Niche conservatism: integrating evolution, ecology, and conservation biology. Annual Review of Ecology, Evolution and Systematics 36: 519539.Google Scholar
Wiggins, BE & Kinkel, LL (2005a) Green manures and crop sequences influence alfalfa root rot and pathogen inhibitory activity among soil-borne Streptomycetes. Plant and Soil 268: 271283.Google Scholar
Wiggins, BE & Kinkel, LL (2005b) Green manures and crop sequences influence potato diseases and pathogen inhibitory activity of indigenous streptomycetes. Phytopathology 95: 178185.Google Scholar
Wiklund, C & Friberg, M (2008) Enemy-free space and habitat-specific host specialization. Oecologia 157: 287294.Google Scholar
Wilby, A & Thomas, MB (2002) Natural enemy diversity and pest control: patterns of pest emergence with agricultural intensification. Ecology Letters 5: 353360.Google Scholar
Wilkinson, ATS (1965) Releases of cinnabar moth, Hypocrita jacobaeae (L.) (Lepidoptera: Arctiidae) on tansy ragwort in British Columbia. Proceedings of the Entomological Society of British Columbia 62: 1015.Google Scholar
Wilkinson, TK & Landis, DA (2005) Habitat diversification in biological control: the role of plant resources. In Plant-Provided Food for Carnivorous Insects: A Protective Mutualism and Its Application (Wäckers, FL, Van Rijn, PCJ & Bruin, J, eds.), Cambridge, UK, Cambridge University Press, pp. 305325.Google Scholar
Williams, L & Martinson, TE (2000) Colonization of New York vineyards by Anagrus spp. (Hymenoptera: Mymaridae): Overwintering biology, within-vineyard distribution of wasps, and parasitism of grape leafhopper, Erythroneura spp. (Homoptera: Cicadellidae), eggs. Biological Control 18: 136146.Google Scholar
Williams, RT, Fullagar, PJ, Kogon, C & Davey, C (1973) Observations on a naturally-occurring winter epizootic of myxomatosis at Canberra, Australia, in the presences of rabbit fleas (Spilopsyllus cuniculi Dale) and virulent myxoma virus. Journal of Applied Ecology 10: 417427.Google Scholar
Williamson, M (1991) Biocontrol risks. Nature 353: 394.Google Scholar
Williamson, M (1996) Biological Invasions. London, UK, Chapman & Hall.Google Scholar
Williamson, M & Fitter, A (1996) The varying success of invaders. Ecology 77: 16611666.Google Scholar
Willis, AJ & Blossey, B (1999) Benign environments do not explain the increased vigour of non-indigenous plants: a cross-continental transplant experiment. Biocontrol Science and Technology 9: 567577.Google Scholar
Willis, AJ & Memmott, J (2005) The potential for indirect effects between a weed, one of its biocontrol agents and native herbivores: a food web approach. Biological Control 35: 299306.Google Scholar
Willis, AJ, Memmott, J & Forrester, RI (2000) Is there evidence for the post-invasion evolution of increased size among invasive plant species? Ecology Letters 3: 275283.Google Scholar
Willis, AJ, Thomas, MB & Lawton, JH (1999) Is the increased vigour of invasive weeds explained by a trade-off between growth and herbivore resistance? Oecologia 120: 632640.Google Scholar
Willoquet, JMR (2010) Collaboration and communication strategies between researchers and potential GMO vector user communities. In Progress and Prospects for the Use of Genetically Modified Mosquitoes to Inhibit Disease Transmission (Crawford, VL & Reza, JN, eds.), Geneva, Switzerland, WHO, pp. 3436.Google Scholar
Wilson, CL (1969) Use of plant pathogens in weed control. Annual Review of Phytopathology 7: 411434.Google Scholar
Wilson, CL (1998) Conserving epiphytic microorganisms on fruits and vegetables for biological control. In Conservation Biological Control (Barbosa, P, ed.), San Diego, CA, Academic Press, pp. 335350.Google Scholar
Wilson, F (1965) Biological control and the genetics of colonizing species. In The Genetics of Colonizing Species (Baker, HG & Stebbins, GL, eds.), New York, Academic Press, pp. 307325.Google Scholar
Wilson, JRU, Ajuonu, O, Center, TD, Hill, MP, Julien, MH, Katagira, FF, Neuenschwander, P, Njoka, SW, Ogwang, J, Reeder, RH & Van, T (2007) The decline of water hyacinth on Lake Victoria was due to biological control by Neochetina spp. Aquatic Botany 87: 9093.Google Scholar
Wilson, K, Thomas, MB, Blanford, S, Doggett, M, Simpson, SJ & Moore, SL (2002) Coping with crowds: density-dependent disease resistance in desert locusts. Proceedings of the National Academy of Sciences USA 99: 54715475.Google Scholar
Wilson, M (1997) Biocontrol of aerial plant diseases in agriculture and horticulture: current approaches and future prospects. Journal of Industrial Microbiology & Biotechnology 19: 188191.Google Scholar
Wilson, M, Xin, WM, Hashmi, S & Gaugler, R (1999) Risk assessment and fitness of a transgenic entomopathogenic nematode. Biological Control 15: 8187.Google Scholar
Wilson, MJ, Glen, DM, Hamacher, GM & Smith, JU (2004) A model to optimise biological control of slugs using nematode parasites. Applied Soil Ecology 26: 179191.Google Scholar
Windels, CE (1997) Altering community balance: organic ammendments, selection pressures, and biocontrol. In Ecological Interactions and Biological Control (Andow, DA, Ragsdale, DW & Nyvall, RF, eds.), Boulder, CO, Westview, pp. 282300.Google Scholar
Winkler, K, Wäckers, FL, Kaufman, LV, Larraz, V & Van Lenteren, JC (2009a) Nectar exploitation by herbivores and their parasitoids is a function of flower species and relative humidity. Biological Control 50: 299306.Google Scholar
Winkler, K, Wäckers, F & Pinto, DM (2009b) Nectar-providing plants enhance the energetic state of herbivores as well as their parasitoids under field conditions. Ecological Entomology 34: 221227.Google Scholar
Winston, RL, Schwarzländer, M, Hinz, HL, Day, MD, Cock, MJW & Julien, MH (2014) Biological Control of Weeds: A World Catalogue of Agents and Their Target Weeds, 5th edition. USDA Forest Service, FHTET-2014-04, Morgantown, WV.Google Scholar
Wissinger, SA (1997) Cyclic colonization in predictably ephemeral habitats: a template for biological control in annual crop systems. Biological Control 10: 415.Google Scholar
Withers, TM & Browne, BL (2004) Behavioral and physiological processes affecting outcomes of host range testing. In Assessing Host Ranges for Parasitoids and Predators Used for Classical Biological Control: A Guide to Best Practice (Van Driesche, RG & Reardon, R, eds.), Morgantown, WV, USDA Forest Service, pp. 4055.Google Scholar
Withers, TM, Carlson, CA & Gresham, BA (2013) Statistical tools to interpret risks that arise from rare events in host specificity testing. Biological Control 64: 177185.Google Scholar
Wodzicki, K (1981) Some nature conservation problems in the South Pacific. Biological Conservation 21: 518.Google Scholar
Woets, J & Van Lenteren, JC (1976) The parasite-host relationship between Encarsia formosa (Hym., Aphelinidae) and Trialeurodes vaporariorum (Hom., Aleyrodidae). VI. Influence of the host plant on the greenhouse whitefly and its parasite Encarsia formosa. IOBC/WPRS Bulletin 4: 151164.Google Scholar
Wolfe, LM (2002) Why alien invaders succeed: support for the escape-from-enemy hypothesis. American Naturalist 160: 705711.Google Scholar
Wolfe, LM, Elzinga, JA & Biere, A (2004) Increased susceptibility to enemies following introduction in the invasive plant Silene latifolia. Ecology Letters 7: 813820.Google Scholar
Wolfenbarger, LL, Naranjo, SE, Lundgren, JG, Bitzer, RJ & Watrud, LS (2008) Bt crop effects on functional guilds of non-target arthropods: a meta-analysis. PLoS One 3: e2118.Google Scholar
Wood, AR, Crous, PW & Lennox, CL (2004) Predicting the distribution of Endophyllum osteospermi (Uredinales, Pucciniaceae) in Australia based on its climatic requirements and distribution in South Africa. Australian Plant Pathology 33: 549558.Google Scholar
Wooton, JT (1994) The nature and consequences of indirect effects in ecological communities. Annual Review of Ecology and Systematics 25: 443466.Google Scholar
Wraight, SP & Ramos, ME (2005) Synergistic interaction between Beauveria bassiana- and Bacillus thuringiensis tenebrionis-based biopesticides applied against field populations of Colorado potato beetle larvae. Journal of Invertebrate Pathology 90: 139150.Google Scholar
Wratten, SD, Bowie, MH, Hickman, JM, Evans, AM, Sedcole, JR & Tylianakis, JM (2003) Field boundaries as barriers to movement of hover flies (Diptera: Syrphidae) in cultivated land. Oecologia 134: 605611.Google Scholar
Wratten, SD, Van Emden, HF & Thomas, MB (1998) Within-field and border refugia for the enhancement of natural enemies. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests (Pickett, CH & Bugg, RL, eds.), Berkeley, University of California Press, pp. 375403.Google Scholar
Wright, DJ & Verkerk, RHJ (1995) Integration of chemical and biological control systems for arthropods – evaluation in a multitrophic context. Pesticide Science 44: 207218.Google Scholar
Wright, MG, Hoffmann, MP, Kuhar, TP, Gardner, J & Pitcher, SA (2005) Evaluating risks of biological control introductions: a probabilistic risk-assessment approach. Biological Control 35: 338347.Google Scholar
Wu, Z, Hopper, KR, Ode, PJ, Fuester, R, Tuda, M & Heimpel, GE (2005). Single-locus complementary sex determination absent in Heterospilus prosopidis (Hymenoptera: Braconidae). Heredity 95: 228234.Google Scholar
Wu, Z, Hopper, KR, O’Neil, RJ, Voegtlin, DJ, Prokrym, DR & Heimpel, GE (2004) Reproductive compatibility and genetic variation between two strains of Aphelinus albipodus (Hymenoptera: Aphelinidae), a parasitoid of the soybean aphid, Aphis glycines (Homoptera: Aphididae). Biological Control 31: 311319.Google Scholar
Wyckhuys, KAG & Heimpel, GE (2007) Response of the soybean aphid parasitoid Binodoxys communis (Gahan) (Hymenoptera: Braconidae) to olfactory cues from target and non-target host-plant complexes. Entomologia Experimentalis et Applicata 123: 149158.Google Scholar
Wyckhuys, KAG, Hopper, KR, Wu, K-M, Straub, C, Gratton, C & Heimpel, GE (2007a) Predicting potential ecological impact of soybean aphid biological control introductions. Biocontrol News and Information 28: 30 N34 N.Google Scholar
Wyckhuys, KAG, Koch, RL & Heimpel, GE (2007b) Physical and ant-mediated refuges from parasitism: implications for non-target effects in biological control. Biological Control 40: 306313.Google Scholar
Wyckhuys, KAG, Koch, RL, Kula, RR & Heimpel, GE (2009) Potential exposure of a classical biological control agent of the soybean aphid, Aphis glycines, on non-target aphids in North America. Biological Invasions 11: 857871.Google Scholar
Wyckhuys, KAG, Lu, Y, Morales, H, Vazquez, LL, Legaspi, JC, Eliopoulos, PA & Hernandez, LM (2013) Current status and potential of conservation biological control for agriculture in the developing world. Biological Control 65: 152167.Google Scholar
Wyss, E, Niggli, U & Nentwig, W (1995) The impact of spiders on aphid populations in a strip-managed apple orchard. Journal of Applied Entomology 119: 473478.Google Scholar
Xi, Z, Khoo, CCH & Dobson, SL (2005) Wolbachia establishment and invasion in an Aedes aegypti laboratory population. Science 310: 326328.Google Scholar
Xiao, K, Kinkel, LL & Samac, DA (2002) Biological control of Phytophthora root rots on alfalfa and soybean with Streptomyces. Biological Control 23: 285295.Google Scholar
Xu, X-M & Jeger, MJ (2013) Theoretical modeling suggests that synergy may result from combined use of two biocontrol agents for controlling foliar pathogens under spatial heterogeneous conditions. Phytopathology 103: 768775.Google Scholar
Xu, X-M, Jeffries, P, Pautasso, M & Jeger, MJ (2011) Combined use of biocontrol agents to manage plant diseases in theory and practice. Phytopathology 101: 10241031.Google Scholar
Xu, X-M, Salama, N, Jeffries, P & Jeger, MJ (2010) Numerical studies of biocontrol efficacies of foliar plant pathogens in relation to the characteristics of a biocontrol agent. Phytopathology 100: 814821.Google Scholar
Yachi, S & Loreau, M (1999) Biodiversity and ecosystem productivity in a fluctuating environment: the insurance hypothesis. Proceedings of the National Academy of Sciences, USA 96: 14631468.Google Scholar
Yano, E (1989) A simulation study of population interaction between the greenhouse whitefly Trialeurodes vaporariorum Westwood (Homoptera: Aleyrodidae) and the parasitoid Encarsia formosa Gahan (Hymenoptera: Aphelinidae). II. Simulation analysis of population dynamics and strategy of biological control. Researches on Population Ecology 31: 89104.Google Scholar
Yano, E (2006) Ecological considerations for biological control of aphids in protected culture. Population Ecology 48: 333339.Google Scholar
Yara, K, Sasawaki, T & Kunimi, Y (2010) Hybridization between introduced Torymus sinensis (Hymenoptera: Torymidae) and indigenous T. beneficus (late-spring strain), parasitoids of the Asian chestnut gall wasp Dryocosmus kuriphilus (Hymenoptera: Cynipidae). Biological Control 54: 1418.Google Scholar
Yasuda, H, Evans, EW, Kajita, Y, Urakawa, K & Takizawa, T (2004) Asymmetric larval interactions between introduced and indigenous ladybirds in North America. Oecologia 141: 722731.Google Scholar
Ye, SD, Ying, SH, Chen, C, and M. G. & Feng, MG (2006) New solid-state fermentation chamber for bulk production of aerial conidia of fungal biocontrol agents on rice. Biotechnology Letters 28: 799804.Google Scholar
Yeates, AG, Schooler, SS, Garono, RJ & Buckley, YM (2012) Biological control as an invasion process: disturbance and propagule pressure affect the invasion success of Lythrum salicaria biological control agents. Biological Invasions 14: 255271.Google Scholar
Yeates, GW & Wardle, DA (1996) Nematodes and predators and prey: relationships to biological and soil processes. Pedobiologia 40: 4350.Google Scholar
Yong, TH & Hoffmann, MP (2006) Habitat selection by the introduced biological control agent Trichogramma ostriniae (Hymenoptera: Trichogrammatidae) and implications for nontarget effects. Environmental Entomology 35: 725732.Google Scholar
Yong, TH, Pitcher, S, Gardner, J & Hoffmann, MP (2007) Odor specificity testing in the assessment of efficacy and non-target risk for Trichogramma ostriniae (Hymenoptera: Trichogrammatidae). Biocontrol Science and Technology 17: 135153.Google Scholar
Yu, H & Sutton, JC (1997) Effectiveness of bumblebees and honeybees for delivering inoculum of Gliocladium roseum to raspberry flowers to control Botrytis cinerea. Biological Control 10: 113122.Google Scholar
Zayed, A & Packer, L (2005) Complementary sex determination substantially increases extinction proneness of haplodiploid populations. Proceedings of the National Academy of Sciences, USA 102: 1074210746.Google Scholar
Zchori-Fein, E, Perlman, SJ, Kelly, SE, Katzir, N & Hunter, MS (2004) Characterization of a ‘Bacteroidetes’ symbiont in Encarsia wasps (Hymenoptera: Aphelinidae): proposal of ‘Candidatus Cardinium hertigii. International Journal of Systematic and Evolutionary Microbiology 54: 961968.Google Scholar
Zelazny, B, Alfiler, AR & Lolong, A (1989) Possibility of resistance to a baculovirus in populations of the coconut beetle (Oryctes rhinoceros). FAO Plant Protection Bulletin 37: 7782.Google Scholar
Zera, AJ & Harshman, LG (2001) The physiology of life history trade-offs in animals. Annual Review of Ecology and Systematics 32: 95126.Google Scholar
Zhou, Y, Gu, H & Dorn, S (2006) Single-locus sex determination in the parasitoid wasp Cotesia glomerata(Hymenoptera: Braconidae). Heredity 96: 487492.Google Scholar
Zimmerman, EC (1948) Insects of Hawaii. Volume 1: Introduction. Honolulu, University of Hawaii Press.Google Scholar
Zimmermann, HG, Moran, VC & Hoffmann, JH (2000) The renowned cactus moth, Cactoblastis cactorum: its natural history and threat to native Opuntia floras in Mexico and the United States. Diversity and Distributions 6: 259269.Google Scholar
Zuniga, MC (2002). A pox on thee! Manipulation of the host immune system by myxoma virus and implications for viral-host co-adaptation. Virus Research 88: 1733.Google Scholar
Zwölfer, H & Harris, P (1984) Biology and host specificity of Rhinocyllys conicus (Froel.) (Col., Curculionidae), a successful agent for biocontrol of the thistle, Carduus nutans L. Zeitschrift für Angewandte Entomologie 97: 3662.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.

  • References
  • George E. Heimpel, University of Minnesota, Nicholas J. Mills, University of California, Berkeley
  • Book: Biological Control
  • Online publication: 20 April 2017
  • Chapter DOI: https://doi.org/10.1017/9781139029117.013
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
  • George E. Heimpel, University of Minnesota, Nicholas J. Mills, University of California, Berkeley
  • Book: Biological Control
  • Online publication: 20 April 2017
  • Chapter DOI: https://doi.org/10.1017/9781139029117.013
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
  • George E. Heimpel, University of Minnesota, Nicholas J. Mills, University of California, Berkeley
  • Book: Biological Control
  • Online publication: 20 April 2017
  • Chapter DOI: https://doi.org/10.1017/9781139029117.013
Available formats
×