Skip to main content Accessibility help
×
Hostname: page-component-7c8c6479df-ws8qp Total loading time: 0 Render date: 2024-03-28T14:55:10.481Z Has data issue: false hasContentIssue false

2 - Suitability of (extra-)floral nectar, pollen, and honeydew as insect food sources

Published online by Cambridge University Press:  15 December 2009

Felix L. Wäckers
Affiliation:
Centre for Terrestrial Ecology Netherlands Institute of Ecology (NIOO-KNAW) The Netherlands
F. L. Wäckers
Affiliation:
Netherlands Institute of Ecology
P. C. J. van Rijn
Affiliation:
Netherlands Institute of Ecology
J. Bruin
Affiliation:
Universiteit van Amsterdam
Get access

Summary

Introduction

Although arthropod predators and parasitoids are usually associated with their carnivorous qualities, they often require plant-provided foods as well, at least during part of their life cycle. The level in which predators or parasitoids depend on these plant-provided foods varies (see Wäckers and van Rijn, Chapter 1). Temporal omnivores and permanent omnivores are facultative consumers of plant-derived food, using it as complement to their prey. This category includes mites (Bakker and Klein 1992b), spiders (Ruhren and Handel 1999), hemipterans (Bugg et al. 1991), beetles (Larochelle 1990; Pemberton and Vandenberg 1993; Pfannenstiel and Yeargan 2002), lacewings (Limburg and Rosenheim 2001), wasps (Beggs 2001), and ants (Porter 1989). Life-history omnivores, on the other hand, are obligatory consumers of plant-provided foods during certain stages (usually the adult stage). This means that they are entirely dependent on non-prey food for their survival and metabolic upkeep. Examples are syrphid flies (Lunau and Wacht 1994), some lacewings (Canard 2001), and many parasitoids (Jervis et al. 1996) and ants (Porter 1989; Tobin 1994).

Ants play a key role in the evolution of a range of food-mediated mutualisms, including extrafloral nectar, food bodies, elaiosomes, Lycaenid dorsal gland secretions and certain honeydews (see also Koptur, Chapter 3). The degree in which ants depend on these foods varies widely. The dietary requirements of ants range from species that are primarily predaceous, to species that rely almost entirely on honeydew and extrafloral nectar.

Type
Chapter
Information
Plant-Provided Food for Carnivorous Insects
A Protective Mutualism and its Applications
, pp. 17 - 74
Publisher: Cambridge University Press
Print publication year: 2005

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

Abe, T. and Kamo, K.. 2003. Seasonal changes of floral frequency and composition of flower in two cool temperate secondary forests in Japan. Forest Ecology and Management 175: 153–162.CrossRefGoogle Scholar
Adler, L. S. 2001. The ecological significance of toxic nectar. Oikos 91: 409–420.CrossRefGoogle Scholar
Allsopp, M. H., Nicolson, S. W., and Jackson, S.. 1998. Xylose as a nectar sugar: the response of Cape honeybees, Apis mellifera capensis Eschscholtz (Hymenoptera: Apidae). African Entomology 6: 317–323.Google Scholar
Alm, J., Ohnmeiss, T. E., Lanza, J., and Vriesenga, L.. 1990. Preference of cabbage white butterflies and honey bees for nectar that contains amino acids. Oecologia 84: 53–57.CrossRefGoogle ScholarPubMed
Alomar, O. and Wiedemann, R. N.. 1996. Zoophytophagous Heteroptera: Implications for Life History and Integrated Pest Management. Lanham, MD: Entomological Society of America.Google Scholar
,Anonymous. 1999. Subsidieregeling Agrarisch Natuurbeheer. The Hague, the Netherlands: Ministerie van Landbouw, Natuurbeheer en Visserij.Google Scholar
,Anonymous. 2001. Verordnung über die Direktzahlungen an die Landwirtschaft. Bern, Switzerland: Bundesamt für Landwirtschaft.Google Scholar
Arakaki, N. and Hattori, M.. 1998. Differences in the quality and quantity of honeydew from first instar soldier and ordinary morph nymphs of the bamboo aphid, Pseudoregma koshunensis (Takahashi) (Homoptera: Aphididae). Applied Entomology and Zoology 33: 357–361.CrossRefGoogle Scholar
Avidov, Z., Balshin, M., and Gerson, U.. 1970. Studies on Aphytis coheni, a parasite of the california red scale, Aonidiella aurantii, in Israel. Entomophaga 15: 191–207.CrossRefGoogle Scholar
Baggen, L. R., Gurr, G. M., and 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: 155–161.CrossRefGoogle Scholar
Baker, H. G., and I. Baker. 1975. Studies of nectar-constitution and pollinator–plant coevolution. In Gilbert, L. E. and Raven, P. H. (eds.) Co-Evolution of Animals and Plants. Austin, TX: University of Texas Press, pp. 100–140.Google Scholar
Baker, H. G., and I. Baker. 1982. Chemical constituents of nectar in relation to pollination mechanisms and phylogeny. In Nitecki, M. H. (ed.) Biochemical Aspects of Evolutionary Biology. Chicago, IL: University of Chicago Press, pp. 131–171.Google Scholar
Baker, H. G., and I. Baker. 1983a. Floral nectar sugar constituents in relation to pollinator type. In Jones, C. E. and Little, R. J. (eds.) Handbook of Experimental Pollination Biology. New York: Van Nostrand Reinhold, pp. 117–141.Google Scholar
Baker, H. G., and I. Baker. 1983b. Some evolutionary and taxonomic implications of variation in the chemical reserves of pollen. In Mulcahy, D. L. and Ottaviano, E. (eds.) Pollen: Biology and Implications for Plant Breeding. New York: Elsevier, pp. 43–52.Google Scholar
Baker, H. G., Opler, P. A., and Baker, I.. 1978. A comparison of the amino acid complements of floral and extrafloral nectar. Botanical Gazette 139: 322–332.CrossRefGoogle Scholar
Bakker, F. M. and Klein, M. E.. 1992a. How cassava plants enhance the efficacy of their phytoseiid bodyguards. Proc. 8th Int. Symp. Insect-Plant Relationship, Dordrecht, pp. 353–354.Google Scholar
Bakker, F. M. and Klein, M. E.. 1992b. Transtrophic interactions in cassava. Experimental and Applied Acarology 14: 293–311.CrossRefGoogle Scholar
Barker, R. J., 1990. Poisoning by plants. In Morse, R. A. and Nowogrodzki, R. (eds.) Honey Bee Pests, Predators and Diseases. Ithaca, NY: Cornell University Press, pp. 275–296.Google Scholar
Basden, R. 1965. The occurrence and composition of Manna in Eucalyptus and Angophora. Proceedings of the Linnean Society of New South Wales 90(2): 152–156.Google Scholar
Beach, J. P., Williams, L., Hendrix, D. L., and Price, L.. 2003. Different food sources affect the gustatory response of Anaphes iole, an egg parasitoid of Lygus spp. Journal of Chemical Ecology 29: 1203–1222.CrossRefGoogle ScholarPubMed
Beattie, A. J., 1985. The Evolutionary Ecology of Ant–Plant Mutualisms. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Beerwinkle, K. R., Shaver, T. N., and Lopez, J. D. Jr. 1993. Field observations of adult emergence and feeding behavior of Helicoverpa zea (Lepidoptera: Noctuidae) on dallisgrass ergot honeydew. Environmental Entomology 22: 554–558.CrossRefGoogle Scholar
Beggs, J. 2001. The ecological consequences of social wasps (Vespula spp.) invading an ecosystem that has an abundant carbohydrate resource. Biological Conservation 99: 17–28.CrossRefGoogle Scholar
Bentley, B. L. 1977a. Extrafloral nectaries and protection by pugnacious bodyguards. Annual Review of Ecology and Systematics 8: 407–427.CrossRefGoogle Scholar
Bentley, B. L. 1977b. The protective function of ants visiting the extrafloral nectaries of Bixa orellana (Bixaceae). Journal of Ecology 65: 27–38.CrossRefGoogle Scholar
Bequaert, J. 1924. Galls that secrete honeydew: a contribution to the problem as to whether galls are altruistic adaptations. Bulletin of the Brooklin Entomological Society 19(4): 101–124.Google Scholar
Beutler, R. 1935. Nectar. Bee World 24: 106–162.Google Scholar
Biere, A. and Honders, S.. 1996. Host adaptation in the anther smut fungus Ustilago violacea (Microbotryum violaceum): infection success, spore production and alteration of floral traits on two host species and their F1- hybrid. Oecologia 107: 307–320.CrossRefGoogle ScholarPubMed
Blüthgen, N. and Fiedler, K.. 2004. Preferences for sugars and amino acids and their conditionality in a diverse nectar-feeding ant community. Journal of Animal Ecology 73: 155–166.CrossRefGoogle Scholar
Blüthgen, N., Verhaagh, M., Goitia, W., et al. 2000. How plants shape the ant community in the Amazonian rainforest canopy: the key role of extrafloral nectaries and homopteran honeydew. Oecologia 125: 229–240.CrossRefGoogle ScholarPubMed
Boevé, J. L. and Wäckers, F. L.. 2003. Gustatory perception and metabolic utilization of sugars by Myrmica rubra ant workers. Oecologia 136: 508–514.CrossRefGoogle ScholarPubMed
Boggs, C. L., 1987. Ecology of nectar and pollen feeding in Lepidoptera. In Slansky, F. and Rodriguez, J. G. (eds.) Nutritional Ecology of Insects, Mites, Spiders, and Related Invertebrates. New York: John Wiley, pp. 369–391.Google Scholar
Bory, G. and Maczulajtys, D. Clair. 1986. Nectar composition and role of the extrafloral nectar in Ailanthus glandulosa. Canadian Journal of Botany 64: 247–253.CrossRefGoogle Scholar
Bouchard, Y. and Cloutier, C.. 1985. Role of olfaction in host finding by the aphid parasitoid Aphidius nigriceps (Hymenoptera: Aphidiidae). Journal of Chemical Ecology 11: 801–808.CrossRefGoogle Scholar
Bowden, B. 1970. The sugars in the extrafloral nectar of Andropogon gayanus var bisquamulatus. Phytochemistry 9: 2315–2318.CrossRefGoogle Scholar
Boyko, A. K. 1939. Larvae of Senotainia triguspis Meig. causing heavy losses of bees. Doklady Akademiia Nauk USSR 24: 306–309.Google Scholar
Bresinsky, A. 1963. Bau, Entwicklungsgeschichte und Inhaltstoffe der Elaiosomen. Bibliotheca Botanica 126: 1–54.Google Scholar
Briscoe, A. D. and Chittka, L.. 2001. The evolution of color vision in insects. Annual Review of Entomology 46: 471–510.CrossRefGoogle ScholarPubMed
Bristow, C. M., 1991. Why are so few aphids ant-tended? In Huxley, C. R. and Cutler, D. F. (eds.) Ant–Plant Interactions. Oxford, UK: Oxford University Press, pp. 104–119.Google Scholar
Brodbeck, B. V., J. Funderburk, J. Stavisky, P. C. Andersen, and J. Hulshof. 2002. Recent advances in the nutritional ecology of Thysanoptera, or the lack thereof. Proc. 7th Int. Symp. Thysanoptera, Canberra, pp. 145–153.
Bronstein, J. L. and Ziv, Y.. 1997. Costs of two non-mutualistic species in a yucca/yucca moth mutualism. Oecologia 112: 379–385.CrossRefGoogle Scholar
Broufas, G. D. and Koveos, D. S.. 2000. Effect of different pollens on development, survivorship and reproduction of Euseius finlandicus (Acari: Phytoseiidae). Environmental Entomology 29: 743–749.CrossRefGoogle Scholar
Budenberg, W. J. 1990. Honeydew as a contact kairomone for aphid parasitoids. Entomologia Experimentalis et Applicata 55: 139–148.CrossRefGoogle Scholar
Bugg, R. L., Ehler, L. E., and Wilson, L. T.. 1987. Effect of common knotweed (Polygonum aviculare) on abundance and efficiency of insect predators of crop pests. Hilgardia 55: 1–52.CrossRefGoogle Scholar
Bugg, R. L., Ellis, R. T., and Carlson, R. W.. 1989. Ichneumonidae (Hymenoptera) using extrafloral nectar of faba bean (Vicia faba L., Fabaceae) in Massachusetts. Biological Agriculture and Horticulture 6: 107–114.CrossRefGoogle Scholar
Bugg, R. L., Wäckers, F. L., Brunson, K. E., Dutcher, J. D., and Phatak, S. C.. 1991. Cool-season cover crops relay intercropped with cantaloupe: influence on a generalist predator Geocoris punctipes (Hemiptera: Lygaeidae). Journal of Economical Entomology 84: 408–416.CrossRefGoogle Scholar
Buller, A. H. R. 1950. Researches on Fungi, vol. 7, The Sexual Process in the Uredinales. Toronto, Ontario: University of Toronto Press.Google Scholar
Burquez, A. and Corbet, S. A.. 1991. Do flowers reabsorb nectar? Functional Ecology 5: 369–379.CrossRefGoogle Scholar
Butler, G. D., Loper, G. M., McGregor, S. E., Webster, J. L., and Margolis, H.. 1972. Amounts and kinds of sugars in the nectars of cotton (Gossypium spp.) and the time of their secretion. Agronomy Journal 64: 364–368.CrossRefGoogle Scholar
Byrne, D. N. and Miller, W. B.. 1990. Carbohydrate and amino acid composition of phloem sap and honeydew produced by Bemisia tabaci. Journal of Insect Physiology 36: 433–439.CrossRefGoogle Scholar
Calatayud, P. A., Rahbé, Y., Delobel, B., et al. 1994. Influence of secondary compounds in the phloem sap of cassava on expression of antibiosis towards the mealybug Phenacoccus manihoti. Entomologia Experimentalis et Applicata 72: 47–57.CrossRefGoogle Scholar
Caldwell, D. L. and Gerhardt, K. O.. 1986. Chemical analysis of peach extrafloral exudate. Phytochemistry 25: 411–413.CrossRefGoogle Scholar
Campana, B. J. and Moeller, F. E.. 1977. Honey bees: preference for and nutritive value of pollen from five plant sources. Journal of Economic Entomology 70: 39–41.CrossRefGoogle Scholar
Canard, M., 2001. Natural food and feeding habits of lacewings. In McEwen, P., New, T. R., and Whittington, A. E. (eds.) Lacewings in the Crop Environment. Cambridge, UK: Cambridge University Press, pp. 116–129.CrossRefGoogle Scholar
Carey, D. B. and Wink, M.. 1994. Elevational variation of quinolizidine alkaloid contents in a lupine (Lupinus argenteus) of the Rocky Mountains. Journal of Chemical Ecology 20: 849–857.CrossRefGoogle Scholar
Carisey, N. and Bauce, E.. 1997. Impact of balsam fir flowering on pollen and foliage biochemistry in relation to spruce budworm growth, development and food utilization. Entomologia Experimentalis et Applicata 85: 17–31.CrossRefGoogle Scholar
Carroll, C. R. and Janzen, D. H.. 1973. Ecology of foraging by ants. Annual Review of Ecology and Systematics 4: 231–257.CrossRefGoogle Scholar
Chapman, R. F., 1995. Chemosensory regulation of feeding. In Chapman, R. F. and Boer, G. (eds.) Regulatory Mechanisms in Insect Feeding. New York: Chapman and Hall, pp. 101–136.CrossRefGoogle Scholar
Chittka, L. 2001. Camouflage of predatory crab spiders on flowers and the colour perception of bees (Aranida: Thomisidae / Hymenoptera: Apidae). Entomologia Generalis 25: 181–187.CrossRefGoogle Scholar
Clark, T. B. 1978. Honey bee spiroplasmosis, a new problem for beekeepers. American Bee Journal 118: 18–23.Google Scholar
Cloutier, C. 1986. Amino acid utilization in the aphid Acyrthosiphon pisum infected by the parasitoid Aphidius smithi. Journal of Insect Physiology 32: 263–267.CrossRefGoogle Scholar
Collin, L. J. and Jones, C. E.. 1980. Pollen energetics and pollination modes. American Journal of Botany 67: 210–215.CrossRefGoogle Scholar
Comba, L., Corbet, S. A., Hunt, L., and Warren, B.. 1999. Flowers, nectar and insect visits: evaluating British plant species for pollinator-friendly gardens. Annals of Botany 83: 369–383.CrossRefGoogle Scholar
Cornelius, M. L., Grace, J. K., and Yates, J. R. III. 1996. Acceptability of different sugars and oils to three tropical ant species (Hymen., Formicidae). Anzeiger für Schädlingskunde, Pflanzenschutz und Umweltschutz 69: 41–43.CrossRefGoogle Scholar
Cottrell, T. E. and Yeargan, K. V.. 1998. Effect of pollen on Coleomegilla maculata (Coleoptera: Coccinellidae) population density, predation, and cannibalism in sweet corn. Environmental Entomology 27: 1402–1410.CrossRefGoogle Scholar
Crafts-Bradner, S. J. 2002. Plant nitrogen status rapidly alters amino acid metabolism and excretion in Bemisia tabaci. Journal of Insect Physiology 48: 33–41.CrossRefGoogle Scholar
Crailsheim, K., Schneider, L. H. W., Hrassnigg, N., et al. 1992. Pollen consumption and utilization in worker honeybees (Apis mellifera carnica): dependence on individual age and function. Journal of Insect Physiology 38: 409–419.CrossRefGoogle Scholar
Crane, E. 1978. Sugars poisonous to bees. Bee World 59: 37–38.CrossRefGoogle Scholar
Cruden, R. W., and S. M. Hermann. 1983. Studying nectar? Some observations on the art. In Bentley, B. and Elias, T. (eds.) The Biology of Nectaries. New York: Columbia University Press, pp. 223–241.Google Scholar
Cuautle, M. and Rico-Gray, V.. 2003. The effect of wasps and ants on the reproductive success of the extrafloral nectaried plant Turnera ulmifolia (Turneraceae). Functional Ecology 17: 417–423.CrossRefGoogle Scholar
Cury, R., 1951. Micoses das abelhas e fungos das colmeias. Biologico (Brazil) 17: 214–220.Google Scholar
Dafni, A., 1992. Pollination Ecology. Oxford, UK: Oxford University Press.Google Scholar
Dafni, A. and D. Firmage. 2000. Pollen viability and longevity: practical, ecological and evolutionary implications. In Dafni, A., Hesse, M., and Pacini, E. (eds.) Pollen and Pollination. Vienna: Springer-Verlag, pp. 113–133.CrossRefGoogle Scholar
Dafni, H., Lensky, Y., and Fahn, A.. 1988. Flower and nectar characteristics of nine species of Labiatae and their influence on honeybee visits. Journal of Apicultural Research 27: 103–114.CrossRefGoogle Scholar
Daumann, E. 1932. Über postflorale Nektarabscheidung. Beihefte zum botanischen Zentralblatt 49: 720–734.Google Scholar
Fuente, M., Penas, P. F., and Sols, A.. 1986. Mechanism of mannose toxicity. Biochemical and Biophysical Research Communications 140: 51–55.CrossRefGoogle ScholarPubMed
DeBach, P., Fischer, T. W., and Landi, J.. 1955. Some effects of meteorological factors on all stages of Aphytis lingnanensis, a parasite of the California red scale. Ecology 36: 743–753.CrossRefGoogle Scholar
Dethier, V. G., 1976. The Hungry Fly. Cambridge, MA: Harvard University Press.Google Scholar
Dethier, V. G., Evans, D. R., and Rhoades, M. V.. 1956. Some factors controlling the ingestion of carbohydrates by the blowfly. Biological Bulletin 111: 204–222.CrossRefGoogle Scholar
Detzel, A. and Wink, M.. 1993. Attraction, deterrence or intoxication of bees (Apis mellifera) by plant allelochemicals. Chemoecology 4: 8–18.CrossRefGoogle Scholar
DeVries, P. J., Walla, T. R., and Greeney, H. F.. 1999. Species diversity in spatial and temporal dimensions of fruit-feeding butterflies from two Ecuadorian rainforests. Biological Journal of the Linnean Society 68: 333–353.CrossRefGoogle Scholar
Dixon, A. F. G. 1971. The role of aphids in wood formation. II. The effect of the lime aphid Eucallipterus tiliae L. (Aphididae) on the growth of lime, Tilia × vulgaris Hayne. Journal of Applied Ecology 8: 393–399.CrossRefGoogle Scholar
Dixon, J. J. 1959. Studies on the oviposition behaviour of Syrphidae. Transactions of the Royal Entomological Society of London 11: 57–80.Google Scholar
Dobson, C. H., 1994. Floral volatiles in insect biology. In Bernays, E. (ed.) Insect–Plant Interactions. Boca Raton, FL: CRC Press, pp. 47–81.Google Scholar
Dobson, H. E. M. 1988. Survey of pollen and pollenkitt lipids: chemical cues to flower visitors? American Journal of Botany 75: 170–182.CrossRefGoogle Scholar
Dobson, H. E. M. and G. Bergström. 2000. Ecology and evolution of pollen odors. In Dafni, A., Hesse, M. and Pacini, E. (eds.) Pollen and Pollination. Vienna: Springer-Verlag, pp. 63–89.CrossRefGoogle Scholar
Dobson, H. E. M., Bergström, G., and Groth, I.. 1990. Differences in fragrance chemistry between flower parts of Rosa rugosa. Israel Journal of Botany 39: 143–156.Google Scholar
Domínguez, C. A., Dirzo, R., and Bullock, S. H.. 1989. On the function of floral nectar in Croton suberosus (Euphorbiaceae). Oikos 56: 109–114.CrossRefGoogle Scholar
Downes, W. L. and Dahlem, G. A.. 1987. Keys to the evolution of Diptera: role of Homoptera. Environmental Entomology 16: 847–854.CrossRefGoogle Scholar
Du, Y. J., Poppy, G. M., Powell, W., et al. 1998. Identification of semiochemicals released during aphid feeding that attract the parasitoid Aphidius ervi. Journal of Chemical Ecology 24: 1355–1368.CrossRefGoogle Scholar
Dukas, R. 2001. Effects of perceived danger on flower choice by bees. Ecology Letters 4: 327–333.CrossRefGoogle Scholar
Durrer, S. and Schmid-Hempel, P.. 1994. Shared use of flowers leads to horizontal pathogen transmission. Proceedings of the Royal Society of London Series B 258: 299–302.CrossRefGoogle Scholar
Dyer, L. E. and Landis, D. A.. 1996. Effects of habitat, temperature, and sugar availability on longevity of Eriborus terebrans (Hymenoptera: Ichneumonidae). Environmental Entomology 25: 1192–1201.CrossRefGoogle Scholar
Ehrlén, J. and Erickson, O.. 1993. Toxicity in fleshy fruits: a non-adaptive trait? Oikos 66: 107–113.CrossRefGoogle Scholar
Eijs, I. E. M., Ellers, J., and Duinen, G.-J.. 1998. Feeding strategies in drosophilid parasitoids: the impact of natural food resources on energy reserves in females. Ecological Entomology 23: 133–138.CrossRefGoogle Scholar
Elias, T. S., 1983. Extrafloral nectaries: their structure and distribution. In Bentley, B. L. and Elias, T. S. (eds.) The Biology of Nectaries. New York: Columbia University Press, pp. 174–203.Google Scholar
Elkinton, J. S., and R. T. Cardé. 1984. Odor dispersion. In Bell, W. J. and Cardé, R. T. (eds.) Chemical Ecology of Insects. London: Chapman and Hall, pp. 73–91.CrossRefGoogle Scholar
Engel, V., Fischer, M. K., Wäckers, F. L., and Völkl, W.. 2001. Interactions between extrafloral nectaries, aphids and ants: are there competition effects between plants and homopteran sugar sources? Oecologia 129: 577–584.CrossRefGoogle ScholarPubMed
Erdtman, G., 1952. Pollen Morphology and Plant Taxonomy. Stockholm: Almquist and Wicksell.Google Scholar
Erhardt, A. and Baker, I.. 1990. Pollen amino acids: an additional diet for a nectar feeding butterfly? Plant Systematics and Evolution 169: 111–121.CrossRefGoogle Scholar
Evans, E. W. and England, S.. 1996. Indirect interactions in biological control of insects: pests and natural enemies in alfalfa. Ecological Applications 6(3): 920–930.CrossRefGoogle Scholar
Faegri, K. and Pijl, L.. 1979. The Principles of Pollination Ecology. Oxford, UK: Pergamon Press.Google Scholar
Fahn, A. 1988. Secretory tissues in vascular plants. New Phytologist 108: 229–258.CrossRefGoogle Scholar
Feder, W. A. and Shrier, R.. 1990. Combination of UV-B and ozone reduces pollen-tube growth more than either stress alone. Environmental and Experimental Botany 30: 451–454.CrossRefGoogle Scholar
Feinsinger, P. and Swarm, L. A.. 1978. How common are ant-repellent nectars? Biotropica 10: 238–239.CrossRefGoogle Scholar
Fernandes, G. W., Fagundes, M., Woodman, R. L., and Price, P. W.. 1999. Ant effects on three-trophic level interactions: plant, galls, and parasitoids. Ecological Entomology 24: 411–415.CrossRefGoogle Scholar
Fiedler, K., Hölldobler, B., and Seufert, P.. 1996. Butterflies and ants: the communicative domain. Experientia 52: 14–24.CrossRefGoogle Scholar
Finch, S. 1974. Sugars available from flowers visited by the adult cabbage root fly, Erioischia brassicae (Bch.) (Diptera, Anthomyiidae). Bulletin of Entomological Research 64: 257–263.CrossRefGoogle Scholar
Finch, S. and Coaker, T. H.. 1969. Comparison of the nutritive values of carbohydrates and related compounds to Erioischia brassicae. Entomologia Experimentalis et Applicata 12: 441–453.CrossRefGoogle Scholar
Fischer, M. K. and Shingleton, A. W.. 2001. Host plant and ants influence the honeydew sugar composition of aphids. Functional Ecology 15: 544–550.CrossRefGoogle Scholar
Fischer, M. K., Hoffmann, K. H., and Volkl, W.. 2001. Competition for mutualists in an ant–homopteran interaction mediated by hierarchies of ant attendance. Oikos 92: 531–541.CrossRefGoogle Scholar
Fischer, M. K., Völkl, W., Schopf, R., and Hoffmann, K. H.. 2002. Age-specific patterns in honeydew production and honeydew composition in the aphid Metopeurum fuscoviride: implications for ant-attendance. Journal of Insect Physiology 48: 319–326.CrossRefGoogle ScholarPubMed
Fisher, B. L., Sternberg, S. da Silva Lobo, and Price, D.. 1990. Variation in the use of orchid extrafloral nectar by ants. Oecologia 83: 263–266.CrossRefGoogle ScholarPubMed
Flechtmann, C. H. W. and McMurtry, J. A.. 1992. Studies of cheliceral and deutosternal morphology of some Phytoseiidae (Acari: Mesostigmata) by scanning electron microscopy. International Journal of Acarology 18: 163–169.CrossRefGoogle Scholar
Fonta, C., Pham-Delègue, M., Marilleau, R., and Masson, C.. 1985. Rôle des nectars de tournesol dans le comportement des insectes pollinisateurs et analyse qualitative et quantitative des éléments glucidiques de ces sécrétions. Acta Oecologia 6: 175–186.Google Scholar
Ford, H. A. and Forde, N.. 1976. Birds as possible pollinators of Acacia pycnantha. Australian Journal of Botany 24: 793–795.CrossRefGoogle Scholar
Francis, F., Lognay, J., Wathelet, J. P., and Haubruge, E.. 2001. Effects of allelochemicals from first (Brassicaceae) and second (Myzus persicae and Brevicoryne brassicae) trophic levels on Adalia bipunctata. Journal of Chemical Ecology 27: 243–256.CrossRefGoogle ScholarPubMed
Freitas, A. V. L. and Oliveira, P. S.. 1996. Ants as selective agents on herbivore biology: effects on the behaviour of a non-myrmecophilous butterfly. Journal of Animal Ecology 65: 205–210.CrossRefGoogle Scholar
Gardener, M. C. and Gillman, M. P.. 2001. Analyzing variability in nectar amino acids: composition is less variable than concentration. Journal of Chemical Ecology 27: 2545–2558.CrossRefGoogle ScholarPubMed
Gaume, L. and McKey, D.. 1999. An ant–plant mutualism and its host-specific parasite: activity rhythms, young leaf patrolling, and effects on herbivores of two specialist plant-ants inhabiting the same myrmecophyte. Oikos 84: 130–144.CrossRefGoogle Scholar
Genissel, A., Aupinel, P., Bressac, C., Tasei, J. N., and Chevrier, C.. 2002. Influence of pollen origin on performance of Bombus terrestris micro-colonies. Entomologia Experimentalis et Applicata 104: 329–336.CrossRefGoogle Scholar
Ghazoul, J. 2001. Can floral repellents pre-empt potential ant–plant conflicts? Ecology Letters 4: 295–299.CrossRefGoogle Scholar
Gilbert, F. and Jervis, M. A.. 1998. Functional, evolutionary and ecological aspects of feeding-related mouthpart specializations in parasitoid flies. Biological Journal of the Linnean Society 63: 495–535.CrossRefGoogle Scholar
Giurfa, M., Vorobyev, M., Kevan, P., and Menzel, R.. 1996. Detection of coloured stimuli by honeybees: minimum visual angles and receptor specific contrasts. Journal of Comparative Physiology A 178: 699–709.CrossRefGoogle Scholar
Gleim, K. H., 1985. Die Blütentracht, Nahrungsquellen des Bienenvolkes. Germany: Sankt Augustin, Delta-Verlag.Google Scholar
Gori, D. F., 1983. Post-pollination phenomena and adaptive floral changes. In Jones, C. E. and Little, R. J. (eds.) Handbook of Experimental Pollination Biology. New York: Van Nostrand Reinhold, pp. 31–49.Google Scholar
Gottsberger, G., Arnold, T., and Linskens, H. F.. 1990. Variation in floral nectar amino acids with aging of flowers, pollen contamination, and flower damage. Israel Journal of Botany 39: 167–176.Google Scholar
Gottsberger, G., Schrauwen, J., and Linskens, H. F.. 1984. Amino acids and sugars in nectar, and their putative evolutionary siginificance. Plant Systematics and Evolution 145: 55–77.CrossRefGoogle Scholar
Gracie, C. 1991. Observation of dual function of nectaries in Ruellia radicans (Nees) Lindau (Acanthaceae). Bulletin of the Torrey Botanical Club 118: 1101–1106.CrossRefGoogle Scholar
Guerrant, E. O. and Fiedler, P. L.. 1981. Flower defenses against nectar pilferage by ants. Biotropica 13: 25–33.CrossRefGoogle Scholar
Györfi, J. 1945. Beobachtungen über die Ernährung der Schlupfwespenimagos. Erdészeti Kisérletek 45: 100–112.Google Scholar
Hagedorn, H. H. and Moeller, F. E.. 1968. Effect of the age of pollen used in pollen supplements on their nutritive value for the honey bee. I. Effect on thoracic weight, development of hypopharyngeal glands, and brood rearing. Journal of Apicultural Research 7: 89–95.CrossRefGoogle Scholar
Hagen, K. S., 1986. Ecosystem analysis: plant cultivars (HPR), entomophagous species and food supplements. In Boethel, D. J. and Eikenbary, R. D. (eds.) Interactions of Plant Resistance and Parasitoids and Predators of Insects. New York: John Wiley, pp. 153–197.Google Scholar
Hansen, D. M., Olesen, J. M., and Jones, C. G.. 2002. Trees, birds and bees in Mauritius: exploitative competition between introduced honey bees and endemic nectarivorous birds? Journal of Biogeography 29: 721–734.CrossRefGoogle Scholar
Harborne, J. B., and R. J. Grayer. 1993. Flavonoids and insects. In Harborne, J. B. (ed.) The Flavonoids: Advances in Research since 1986. London: Chapman and Hall, pp. 589–618.Google Scholar
Harder, L. D. 1998. Pollen-size comparisons among animal pollinated angiosperms with different pollination characteristics. Biological Journal of the Linnean Society 64: 513–525.CrossRefGoogle Scholar
Hartley, S. E., and C. G. Jones. 1997. Plant chemistry and herbivory, or why the world is green. In Crawley, M. J. (ed.) Plant Ecology. Oxford, UK: Blackwell Science, pp. 284–324.Google Scholar
Haskins, C. P. and Haskins, E. F.. 1950. Notes on the biology and social behavior of the archaic ponerine ants of the genera Myrmeca and Promyrmeca. Annals of the Entomological Society of America 43: 461–491.CrossRefGoogle Scholar
Haslett, J. R. 1989. Interpreting patterns of resource utilization: randomness and selectivity in pollen feeding by adult hoverflies. Oecologia 78: 433–442.CrossRefGoogle ScholarPubMed
Heil, M., Fiala, B., Baumann, B., and Linsenmair, K. E.. 2000. Temporal, spatial and biotic variations in extrafloral nectar secretion by Macaranga tanarius. Functional Ecology 14: 749–757.Google Scholar
Heil, M., Fiala, B., Kaiser, W., and Linsenmair, K. E.. 1998. Chemical contents of Macaranga food bodies: adaptations to their role in ant attraction and nutrition. Functional Ecology 12: 117–122.CrossRefGoogle Scholar
Heil, M., Koch, T., Hilpert, A., et al. 2001. Extrafloral nectar production of the ant-associated plant, Macaranga tanarius, is an induced, indirect, defensive response elicited by jasmonic acid. Proceedings of the National Academy of Sciences of the USA 98: 1083–1088.CrossRefGoogle ScholarPubMed
Heiling, A. M., Herberstein, M. E., and Chittka, L.. 2003. Pollinator attraction: crab-spiders manipulate flower signals. Nature 421: 334–334.CrossRefGoogle ScholarPubMed
Heimpel, G. E., Lee, J. C., Wu, Z., et al. 2004. Gut sugar analysis in field-caught parasitoids: adapting methods originally developed for biting flies. International Journal of Pest Management 50: 193–198.CrossRefGoogle Scholar
Hendrix, D. L. and Salvucci, M. E.. 2001. Isobemisiose: an unusual trisaccharide abundant in the silverleaf whitefly, Bemisia argentifolii. Journal of Insect Physiology 47: 423–432.CrossRefGoogle ScholarPubMed
Hendrix, D. L., Wei, Y., and Leggett, J. E.. 1992. Homopteran honeydew sugar composition is determined by both the insect and plant species. Comparative Biochemistry and Physiology B 101: 23–27.CrossRefGoogle Scholar
Henning, J. A., Peng, Y. S., Montague, M. A., and Teubler, L. R.. 1992. Honey-bee (Hymenoptera, Apidae) behavioral response to primary alfalfa (Rosales, Fabaceae) floral volatiles. Journal of Economic Entomology 85: 233–239.CrossRefGoogle Scholar
Hern, A. and Dorn, S.. 1999. Sexual dimorphism in the olfactory orientation of adult Cydia pomonella in response to alpha-farnesene. Entomologia Experimentalis et Applicata 92: 63–72.CrossRefGoogle Scholar
Hocking, H. 1966. The influence of food on longevity and oviposition in Rhyssa persuasoria (L.) (Hymenoptera: Ichneumonidae). Journal of the Australian Entomological Society 6: 83–88.CrossRefGoogle Scholar
Hölldobler, B. and Wilson, E. O.. 1990. The Ants. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Horvitz, C. C. and Schemske, D. W.. 1986. Seed dispersal of a neotropical myrmecochore: variation in removal rates and dispersal distance. Biotropica 16: 319–323.CrossRefGoogle Scholar
Hulshof, J. and I. Vänninen. 2002. Western flower thrips feeding on pollen, and its implications for control. In Thrips and Tospoviruses, Proc. 7th Int. Symp. Thysanoptera, Canberra, pp. 173–179.
Huxley, C. R. 1980. Symbiosis between ants and epiphytes. Biological Reviews 55: 321–340.CrossRefGoogle Scholar
Ingold, C. T., 1971. Fungal Spores: Their Liberation and Dispersal. Oxford, UK: Clarendon Press.Google Scholar
Inouye, D. W., 1983. The ecology of nectar robbing. In Bentley, B. and Elias, T. (eds.) The Biology of Nectaries. New York: Columbia University Press, pp. 153–173.Google Scholar
Inouye, D. W. and Taylor, O. R.. 1979. A temperate region plant–ant–seed predator system: consequences of extrafloral nectar secretion by Helianthella quinquenervis. Ecology 60: 1–7.CrossRefGoogle Scholar
Inouye, D. W. and Waller, G. D.. 1984. Responses of honey bees (Apis mellifera) to amino acid solutions mimicking floral nectars. Ecology 65: 618–625.CrossRefGoogle Scholar
Izzo, T. J. and Vasconcelos, H. L.. 2002. Cheating the cheater: domatia loss minimizes the effects of ant castration in an Amazonian ant-plant. Oecologia 133: 200–205.CrossRefGoogle Scholar
Jackson, S. and Nicolson, S. W.. 2002. Xylose as a nectar sugar: from biochemistry to ecology. Comparative Biochemistry and Physiology B 131: 613–620.CrossRefGoogle ScholarPubMed
Jander, R. 1998. Olfactory learning of fruit odors in the eastern yellow jacket, Vespula maculifrons (Hymenoptera: Vespidae). Journal of Insect Behavior 11: 879–888.CrossRefGoogle Scholar
Janzen, D. H. 1966. Coevolution of mutualism between ants and acacias in Central America. Evolution 20: 249–275.CrossRefGoogle ScholarPubMed
Janzen, D. H. 1977. Why don't ants visit flowers? Biotropica 9: 252.CrossRefGoogle Scholar
Jervis, M. A. 1990. Predation of Lissonota coracinus (Gmelin) (Hymenoptera: Ichneumonidae) by Dolichonabis limbatus (Dahlborn) (Hemiptera: Nabidae). Entomologist's Gazette 41: 231–233.Google Scholar
Jervis, M. A. 1998. Functional and evolutionary aspects of mouthpart stucture in parasitoid wasps. Biological Journal of the Linnean Society 63: 461–493.CrossRefGoogle Scholar
Jervis, M. A., Kidd, N. A. C., Fitton, M. G., Huddleston, T., and Dawah, H. A.. 1993. Flower-visiting by hymenopteran parasitoids. Journal of Natural History 27: 67–105.CrossRefGoogle Scholar
Jervis, M. A., Kidd, N. A. C., and Heimpel, G. E.. 1996. Parasitoid adult feeding behaviour and biocontrol: a review. Biocontrol News and Information 17: 11N–26N.Google Scholar
Johnson, M. W., Jones, V. P., and Toscano, N. C.. 1987. Diel activity patterns of tobacco budworm, Heliothis virescens (F.), and cabbage looper, Trichoplusia ni (Hübner) larvae. Environmental Entomology 16: 25–29.CrossRefGoogle Scholar
Jolivet, P., 1998. Myrmecophily and Ant-Plants. Boca Raton, FL: CRC Press.Google Scholar
Josens, R. B., Farina, W. M., and Roces, F.. 1998. Nectar feeding by the ant Camponotus mus as a function of sucrose concentration. Journal of Insect Physiology 44: 579–585.CrossRefGoogle Scholar
Keeler, K. H. 1981. Function of Mentzelia nuda (Loasaceae) postfloral nectaries in seed defense. American Journal of Botany 68: 295–299.CrossRefGoogle Scholar
Kevan, P. G. 1972. Floral colors in the high arctic with reference to insect–flower relations and pollination. Canadian Journal of Botany 50: 2289–2316.CrossRefGoogle Scholar
Kevan, P. G. 1973. Parasitoid wasps as flower visitors in the Canadian high arctic. Anzeiger für Schädlingskunde, Pflanzenschutz und Umweltschutz 46: 3–7.CrossRefGoogle Scholar
Kevan, P. G. and Baker, H. G.. 1983. Insects as flower visitors and pollinators. Annual Review of Entomology 28: 407–453.CrossRefGoogle Scholar
Kevan, P. G., and H. G. Baker. 1998. Insects on flowers. In Huffaker, C. B. and Gutierrez, A. P. (eds.) Ecological Entomology. New York: John Wiley, pp. 553–583.Google Scholar
Kikuchi, T. 1963. Studies on the coaction among insects visiting flowers. III. Dominance relationship among flower-visiting flies, bees and butterflies. Scientific Reports of the Tohoku University 29: 1–8.Google Scholar
Kirk, W. D. J. 1984. Pollen-feeding in thrips (Insecta: Thysanoptera). Journal of Zoology 204: 107–117.CrossRefGoogle Scholar
Kirk, W. D. J. 1985. Pollen-feeding and the host specificity and fecundity of flower thrips. Ecological Entomology 10: 281–289.CrossRefGoogle Scholar
Kiss, A. 1981. Melizitose, aphids and ants. Oikos 37: 382.CrossRefGoogle Scholar
Kleber, E. 1935. Hat das Zeitgedächtnis der Bienen biologische Bedeutung? Journal of Comparative Physiology 22: 221–262.Google Scholar
Knoll, F. 1930. Uber Pollenkitt und Bestäubungsart. Zeitschrift für Botanie 23: 609–675.Google Scholar
Knox, R. B., Kenrick, J., Bernhardt, P., et al. 1985. Extra-floral nectaries as adaptations for bird pollination in Acacia terminalis. American Journal of Botany 72: 1185–1196.CrossRefGoogle Scholar
Knudsen, J. T., Tollsten, L., and Bergström, L. G.. 1993. Floral scents: a checklist of volatile compounds isolated by head-space techniques. Phytochemistry 33: 253–280.CrossRefGoogle Scholar
Koptur, S., 1989. Is extrafloral nectar production an inducible defence? In Bock, J. and Linhart, Y. (eds.) Evolutionary Ecology of Plants. Boulder, CO: Westview Press, pp. 323–339.Google Scholar
Koptur, S. 1992. Extrafloral nectary-mediated interactions between insects and plants. In Bernays, E. (ed.) Insect–Plant Interactions. Boca Raton, FL: CRC Press, pp. 81–129.Google Scholar
Koptur, S. 1994. Floral and extrafloral nectars of Costa Rican Inga trees: a comparison of their constituents and composition. Biotropica 26: 276–284.CrossRefGoogle Scholar
Koptur, S. and Truong, N.. 1998. Facultative ant–plant interactions: nectar sugar preferences of introduced pest ant species in South Florida. Biotropica 30: 179–189.CrossRefGoogle Scholar
Kretschmar, J. A. and Baumann, T. W.. 1999. Caffeine in Citrus flowers. Phytochemistry 52: 19–23.CrossRefGoogle Scholar
Kretz, R. 1979. A behavioural analysis of colour vision in the ant Cataglyphis bicolor (Formicidae, Hymenoptera). Journal of Comparative Physiology A 131: 217–233.CrossRefGoogle Scholar
Krivan, V. and Sirot, E.. 1997. Searching for food or hosts: the influence of parasitoids behavior on host–parasitoid dynamics. Theoretical Population Biology 51: 201–209.CrossRefGoogle ScholarPubMed
Kugler, H., 1970. Blütenökologie. Stuttgart, Germany: Fischer Verlag.Google Scholar
Kunkel, H. and Kloft, W. J. (eds.). 1985. Waldtracht und Waldhonig in der Imkerei. Munich, Germany: Ehrenwirth Verlag.Google Scholar
Kunze, H. 1999. Pollination ecology in two species of Gonolobus (Asclepiadaceae). Flora 194: 309–316.CrossRefGoogle Scholar
Labandeira, C. C. 1997. Permian pollen eating. Science 277: 1422–1423.CrossRefGoogle Scholar
Lanza, J. 1988. Ant preferences for Passiflora nectar mimics that contain amino acids. Biotropica 20: 341–344.CrossRefGoogle Scholar
Lanza, J. and Krauss, B. R.. 1984. Detection of amino acids in artificial nectars by two tropical ants, Leptothorax and Monomorium. Oecologia 63: 423–425.CrossRefGoogle ScholarPubMed
Larochelle, A., 1990. The food of carabid beetles (Coleoptera: Carabidae, including Cicindelinae). Fabreries (suppl.) 5: 1–132.Google Scholar
Leatemia, J. A., Laing, J. E., and Corrigan, J. E.. 1995. Effects of adult nutrition on longevity, fecundity and offspring sex ratio of Trichogramma minutum Riley (Hymenoptera: Trichogrammatidae). Canadian Entomologist 127: 245–254.CrossRefGoogle Scholar
Leduc, N., Douglas, G. C., Monnier, M., and Connolly, V.. 1990. Pollination in vitro: effects on the growth of pollen tubes, seed set and gametophytic self-incompatibility in Trifolium pratense L. and Trifolium repens L. Theoretical and Applied Genetics 80: 657–664.CrossRefGoogle Scholar
Lee, J. C. and Heimpel, G. E.. 2003. Sugar feeding by parasitoids in cabbage fields and the consequences for pest control. Proc. 1st Int. Symp. Biological Control of Arthropods, Honolulu, pp. 220–225.Google Scholar
Leius, K. 1960. Attractiveness of different foods and flowers to the adults of some hymenopterous parasites. Canadian Entomologist 92: 369–376.CrossRefGoogle Scholar
Leius, K. 1961. Influence of various foods on fecundity and longevity of adults of Scambus buolianae (Htg.) (Hymenoptera: Ichneumonidae). Canadian Entomologist 93: 1079–1084.CrossRefGoogle Scholar
Letourneau, D. K. 1990. Code of ant–plant mutualism broken by a parasite. Science 248: 215–217.CrossRefGoogle ScholarPubMed
Leveau, J. H. J. and Lindow, S. E.. 2001. Appetite of an epiphyte: quantitative monitoring of bacterial sugar consumption in the phyllosphere. Proceedings of the National Academy of Sciences of the USA 98: 3446–3453.CrossRefGoogle ScholarPubMed
Lewis, A. C. 1986. Memory constraints and flower choice in Pieris rapae. Science 232: 863–865.CrossRefGoogle ScholarPubMed
Limburg, D. D. and Rosenheim, J. A.. 2001. Extrafloral nectar consumption and its influence on survival and development of an omnivorous predator, larval Chrysoperla plorabunda (Neuroptera: Chrysopidae). Environmental Entomology 30: 595–604.CrossRefGoogle Scholar
Lingren, P. D. and Lukefahr, M. J.. 1977. Effects of nectariless cotton on caged populations of Campoletis sonorensis. Environmental Entomology 6: 586–588.CrossRefGoogle Scholar
Lunau, K., 2000. The ecology and evolution of visual pollen signals. In Dafni, A., Hesse, M. and Pacini, E. (eds.) Pollen and Pollination. Vienna: Springer-Verlag, pp. 89–113.CrossRefGoogle Scholar
Lunau, K. and Wacht, S.. 1994. Optical releasers of the innate proboscis extension in the hoverfly Eristalis tenax L. (Syrphidae, Diptera). Journal of Comparative Physiology A 174: 575–579.CrossRefGoogle Scholar
Malcolm, S. B. 1990. Chemical defenses in chewing and sucking insect herbivores: plant-derived cardenolides in the monarch butterfly and oleander aphid. Chemoecology 1: 12–21.CrossRefGoogle Scholar
Markin, G. P. 1970. Food distribution within laboratory colonies of the Argentine ant, Iridomyrmex humilis (Mayr). Insectes Sociaux 17: 127–158.CrossRefGoogle Scholar
May, P. G. 1985. Nectar uptake rates and optimal nectar concentrations of two butterfly species. Oecologia 66: 381–386.CrossRefGoogle ScholarPubMed
Mendel, Z., Blumberg, D., Zehavi, A., and Weissenberg, M.. 1992. Some polyphagous Homoptera gain protection from their natural enemies by feeding on the toxic plants Spartium junceum and Erythrina corallodendrum (Leguminosa). Chemoecology 3: 118–124.CrossRefGoogle Scholar
Mercier, J. and Lindow, S. E.. 2000. Role of leaf surface sugars in colonization of plants by bacterial epiphytes. Applied and Environmental Microbiology 66: 369–374.CrossRefGoogle ScholarPubMed
Meurer, B., Wray, V., Wiermann, R., and Starck, D.. 1988. Hydroxy cinnamic acid-spermidine amides from pollen of Alnus glutinosa, Betula verrucosa and Pterocarya fraxinifolia. Phytochemistry 27: 839–843.CrossRefGoogle Scholar
Milewski, A. V., and W. J. Bond. 1982. Convergence of myrmecochory in mediterranean Australia and South Africa. In Buckley, R. C. (ed.) Ant–Plant Interactions in Australia. The Hague, the Netherlands: Junk, pp. 89–98.CrossRefGoogle Scholar
Milton, K. 1999. Nutritional characteristics of wild primate foods: do the diets of our closest living relatives have lessons for us? Nutrition 15: 488–498.CrossRefGoogle ScholarPubMed
Mittler, T. E. and Meikle, T.. 1991. Effects of dietary sucrose concentration on aphid honeydew carbohydrate levels and rates of excretion. Entomologia Experimentalis et Applicata 59: 1–7.CrossRefGoogle Scholar
Morse, D. H. 1986. Predation risk to insect foraging at flowers. Oikos 46: 223–228.CrossRefGoogle Scholar
Ne'eman, G. and Kevan, P. G.. 2001. The effect of shape parameters on maximal detection distance of model targets by honeybee workers. Journal of Comparative Physiology A 187: 653–660.CrossRefGoogle ScholarPubMed
Nepi, M. and G. G. Franchi. 2000. Cytochemistry of mature angiosperm pollen. In Dafni, A., Hesse, M., and Pacini, E. (eds.) Pollen and Pollination. Vienna: Springer-Verlag, pp. 45–62.CrossRefGoogle Scholar
Nettles, W. C. and Burks, M. L.. 1971. Absorption and metabolism of galactose and galactitol in Anthonomus grandis. Journal of Insect Physiology 17: 1615–1623.CrossRefGoogle Scholar
O'Dowd, D. J. and Catchpole, E. A.. 1983. Ants and extrafloral nectaries: no evidence for plant protection in Helichrysum spp.–ant interactions. Oecologia 59: 191–200.CrossRefGoogle ScholarPubMed
Oliveira, P. S. 1997. The ecological function of extrafloral nectaries: herbivore deterrence by visiting ants and reproductive output in Caryocar brasiliense (Caryocaraceae). Functional Ecology 11: 323–330.CrossRefGoogle Scholar
Olson, D. L. and Nechols, J. R.. 1995. Effects of squash leaf trichome exudates and honey on adult feeding, survival, and fecundity of the squash bug (Heteroptera: Coreidae) egg parasitoid Gryon pennsylvanicum (Hymenoptera: Scelionidae). Environmental Entomology 24: 454–458.CrossRefGoogle Scholar
Orivel, J. and Dejean, A.. 2002. Ant activity rhythms in a pioneer vegetal formation of French Guiana (Hymenoptera: Formicidae). Sociobiology 39: 65–76.Google Scholar
Osche, G. 1983. Optische Signale in der Coevolution von Pflanze und Tier. Berichte der deutschen botanischen Gesellschaft 96: 1–27.Google Scholar
Pascal, L. and Belin-Depoux, M.. 1991. La correlation entre les rhythmes biologiques de l'association plante–fourmis: le cas des nectaires extra-floraux de Malpighiaceae américaines. Comptes Rendus de l'Académie des Sciences Paris Series 3 312: 49–54.Google Scholar
Pate, J. S., Peoples, M. B., Storer, P. J., and Atkins, C. A.. 1985. The extrafloral nectaries of cowpea (Vigna unguiculata (L.) Walp.). II. Nectar composition, origin of nectar solutes, and nectary functioning. Plant 166: 28–38.CrossRefGoogle ScholarPubMed
Patt, J. M., Hamilton, G. C., and Lashomb, J. H.. 1997. Foraging success of parasitoid wasps on flowers: interplay of insect morphology, floral architecture and searching behavior. Entomologia Experimentalis et Applicata 83: 21–30.CrossRefGoogle Scholar
Pemberton, R. W. and Vandenberg, N. J.. 1993. Extrafloral nectar feeding by ladybird beetles (Coleoptera, Coccinellidae). Proceedings of the Entomological Society of Washington 95: 139–151.Google Scholar
Peng, Z. and Miles, P. W.. 1991. Oxidases in the gut of an aphid, Macrosiphum rosae (L.) and their relation to dietary phenolics. Journal of Insect Physiology 37: 779–787.CrossRefGoogle Scholar
Percival, M. S. 1955. The presentation of pollen in certain angiosperms and its collection by Apis mellifera. New Phytologist 54: 353–368.CrossRefGoogle Scholar
Percival, M. S. 1961. Types of nectar in angiosperms. New Phytologist 60: 235–281.CrossRefGoogle Scholar
Percival, M. S. 1965. Floral Biology. Oxford, UK: Pergamon Press.Google Scholar
Petanidou, T. and Vokou, D.. 1990. Pollination and pollen energetics in Mediterranean ecosystems. American Journal of Botany 77: 986–992.CrossRefGoogle Scholar
Peumans, W. J., Smeets, K., Nerum, K., Leuven, F., and Damme, E. J. M.. 1997. Lectin and alliinase are the predominant proteins in nectar from leek (Allium porrum L.) flowers. Planta 201: 298–302.CrossRefGoogle ScholarPubMed
Pfannenstiel, R. S. and Yeargan, K. V.. 2002. Identification and diel activity patterns of predators attacking Helicoverpa zea (Lepidoptera: Noctuidae) eggs in soybean and sweet corn. Environmental Entomology 31: 232–241.CrossRefGoogle Scholar
Pickett, C. H. and Clark, W. D.. 1979. The function of extrafloral nectaries in Opuntia acanthocarpa (Cactacea). American Journal of Botany 66: 618–625.CrossRefGoogle Scholar
Porter, S. D. 1989. Effects of diet on the growth of laboratory fire ant colonies (Hymenoptera: Formicidae). Journal of the Kansas Entomological Society 62: 288–291.Google Scholar
Potter, C. F. and Bertin, R. I.. 1988. Amino acids in artificial nectar: feeding preferences of the flesh fly Sarcophaga bullata. American Midland Naturalist 120: 156–162.CrossRefGoogle Scholar
Putman, W. L. 1958. Mortality of the European red mite (Acarina: Tetranychidae) from secretions of peach leaf nectaries. Canadian Entomologist 90: 720–721.CrossRefGoogle Scholar
Rahbé, Y., Sauvion, N., Febvay, G., Peumans, W. J., and Gatehouse, A. M. R.. 1995. Toxicity of lectins and processing of ingested proteins in the pea aphid Acyrthosiphon pisum. Entomologia Experimentalis et Applicata 76: 143–155.CrossRefGoogle Scholar
Retana, J., Bosch, J., Alsina, A., and Cerda, X.. 1987. Foraging ecology of the nectarivorous ant Camponotus foreli (Hymenoptera: Formicidae) in a savannah-like grassland. Miscellania Zoologica 11: 187–193.Google Scholar
Ricks, B. L. and Vinson, S. B.. 1970. Feeding acceptability of certain insects and various water-soluble compounds to varieties of the imported fire ant. Journal of Economic Entomology 63: 145–148.CrossRefGoogle Scholar
Rickson, F. R. 1980. Developmental anatomy and ultrastructure of the ant-food bodies (Beccarian bodies) of Macaranga triloba and M. hypoleuca (Euphorbiaceae). American Journal of Botany 67: 285–292.CrossRefGoogle Scholar
Rico-Gray, V. and Thien, L. B.. 1989. Effect of different ant species on reproductive fitness of Schomburgkia tibicinis (Orchidaceae). Oecologia 81: 487–489.CrossRefGoogle Scholar
Rico-Gray, V., Palacios-Rios, M., and Garcia-Franco, J. G.. 1998. Richness and seasonal variation of ant–plant associations as mediated by plant-derived food resources in the semiarid Zapotitlan Valley, Mexico. American Midland Naturalist 140: 21–26.CrossRefGoogle Scholar
Risch, S. J. and Rickson, F. R.. 1981. Mutualism in which ants must be present before plants produce food bodies. Nature 291: 149–150.CrossRefGoogle Scholar
Rogers, C. E. 1985. Extrafloral nectar: entomological implications. Bulletin of the Entomological Society of America 31: 15–20.CrossRefGoogle Scholar
Romeis, J. and Wäckers, F. L.. 2000. Feeding responses by female Pieris brassicae butterflies to carbohydrates and amino acids. Physiological Entomology 25: 247–253.CrossRefGoogle Scholar
Romeis, J. and Wäckers, F. L.. 2002. Nutritional suitability of individual carbohydrates and amino acids for adult Pieris brassicae. Physiological Entomology 27: 148–156.CrossRefGoogle Scholar
Romeis, J. and Zebitz, C. P. W.. 1997. Searching behaviour of Encarsia formosa as mediated by colour and honeydew. Entomologia Experimentalis et Applicata 82: 299–309.CrossRefGoogle Scholar
Romeis, J., Babendreier, D. and Wäckers, F. L.. 2003. Consumption of snowdrop lectin (Galanthus nivalis) agglutinin causes direct effects on adult parasitic wasps. Oecologia 34: 528–536.CrossRefGoogle Scholar
Rosenheim, J. A. 1998. Higher order predators and the regulation of insect herbivore populations. Annual Review of Entomology 43: 421–447.CrossRefGoogle ScholarPubMed
Roulston, T. H. and Buchmann, S. L.. 2000. A phylogenetic reconsideration of the pollen starch–pollination correlation. Evolutionary Ecology Research 2: 627–643.Google Scholar
Roulston, T. H., and J. H. Cane. 2000. Pollen nutritional content and digestibility for animals. In Dafni, A., Hesse, M., and Pacini, E. (eds.) Pollen and Pollination. Vienna: Springer-Verlag, pp. 187–211.CrossRefGoogle Scholar
Roy, B. A. 1993. Floral mimicry by a plant pathogen. Nature 362: 56–58.CrossRefGoogle Scholar
Ruhren, S. and Handel, S.. 1999. Jumping spiders (Salticidae) enhance the seed production of a plant with extrafloral nectaries. Oecologia 119: 227–230.CrossRefGoogle ScholarPubMed
Runions, C. J., Rensing, K. H., Takaso, T. and Owens, J.. 1999. Pollination of Picea orientalis (Pinaceae): saccus morphology governs pollen buoyancy. American Journal of Botany 86: 190–197.CrossRefGoogle Scholar
Rusterholz, H. P. and Erhardt, A.. 1997. Preferences for nectar sugars in the peacock butterfly Inachis io. Ecological Entomology 22: 220–224.CrossRefGoogle Scholar
Rusterholz, H. P. and Erhardt, A.. 1998. Effects of elevated CO2 on flowering phenology and nectar production of nectar plants important for butterflies of calcareous grasslands. Oecologia 113: 341–349.CrossRefGoogle ScholarPubMed
Salvucci, M. E., Rosell, R. C., and Brown, J. K.. 1998. Uptake and metabolism of leaf proteins by the silverleaf whitefly. Archives of Insect Biochemistry and Physiology 39: 155–165.3.0.CO;2-#>CrossRefGoogle Scholar
Sandström, J. P. and Moran, N.. 1999. How nutritionally imbalanced is phloem sap for aphids? Entomologia Experimentalis et Applicata 91: 203–210.CrossRefGoogle Scholar
Sandström, J. P. and Moran, N. A.. 2001. Amino acid budgets in three aphid species using the same host plant. Physiological Entomology 26: 202–211.CrossRefGoogle Scholar
Sasaki, T. and Ishikawa, H.. 1991. Amino acids and their metabolism in symbiotic and aposymbiotic pea aphids. Miscellaneous Publications of the Agricultural Experimental Station of Oklahoma State University 288.Google Scholar
Sasaki, T., Aoki, T., Hayashi, H., and Ishikawa, H.. 1990. Amino acid composition of the honeydew of symbiotic and aposymbiotic pea aphids Acyrthosiphon pisum. Journal of Insect Physiology 36: 35–40.CrossRefGoogle Scholar
Schneider, P. 1972. Versuche zur Frage der individuellen Futterverteilung bei der kleinen roten Waldameise (Formica polyctena). Insectes Sociaux 19: 279–299.CrossRefGoogle Scholar
Schoonhoven, L. M. 1972. Secondary plant substances and insects. Recent Advances in Phytochemistry 5: 197–224.CrossRefGoogle Scholar
Schubert, A. 1972. Bienenfeindliche Pflanzen. Bienenpflege 8: 171–172.Google Scholar
Schwarz, H. H. and Huck, K.. 1997. Phoretic mites use flowers to transfer between foraging bumblebees. Insectes Sociaux 44: 303–310.CrossRefGoogle Scholar
Schwörer, U. and Völkl, W.. 2001. Foraging behavior of Aphidius ervi (Haliday) (Hymenoptera: Braconidae: Aphidiinae) at different spatial scales: resource utilization and suboptimal weather conditions. Biological Control 21: 111–119.CrossRefGoogle Scholar
Scott, H. J. and Stojanovich, C. T.. 1963. Digestion of juniper pollen by Collembola. Florida Entomologist 46: 189–191.CrossRefGoogle Scholar
Seeman, O. D. 1996. Flower mites and phoresy: the biology of Hattena panopla Domrow and Hattena cometis Domrow (Acari: Mesostigmata: Ameroseiidae). Australian Journal of Zoology 44: 193–203.CrossRefGoogle Scholar
Seibert, T. F. 1993. A nectar-secreting gall wasp and ant mutualism: selection and counter-selection shaping gall wasp phenology, fecundity and persistence. Ecological Entomology 18: 247–253.CrossRefGoogle Scholar
Sheldon, J. K. and MacLeod, E. G.. 1971. Studies on the biology of Chrysopidae. II. The feeding behavior of the adult Chrysopa carnea (Neuroptera). Psyche 78: 107–121.CrossRefGoogle Scholar
Shiraishi, A. and Kuwabra, M.. 1970. The effects of amino acids on the labellar hair chemosensory cells of the fly. Journal of Genetic Physiology 56: 768–782.CrossRefGoogle Scholar
Shivanna, K. R., Linskens, H. F., and Cresti, M.. 1991. Responses of tobacco pollen to high humidity and heat-stress: viability and germinability in vitro and in vivo. Sexual Plant Reproduction 4: 104–109.CrossRefGoogle Scholar
Shykoff, J. A. and Bucheli, E.. 1995. Pollinator visitation patterns, floral rewards and the probability of transmission of Microbotryum violaceum, a venereal disease of plants. Journal of Ecology 83: 189–198.CrossRefGoogle Scholar
Siekmann, G., Tenhumberg, B., and Keller, M. A.. 2001. Feeding and survival in parasitic wasps: sugar concentration and timing matter. Oikos 95: 425–430.CrossRefGoogle Scholar
Simpson, S. J. and Raubenheimer, D.. 2001. The geometric analysis of nutrient–allelochemical interactions: a case study using locusts. Ecology 82: 422–439.Google Scholar
Smith, L. L., Lanza, J., and Smith, G. C.. 1990. Amino acid concentrations in extrafloral nectar of Impatiens sultani increase after simulated herbivory. Ecology 71: 107–115.CrossRefGoogle Scholar
Solberg, Y. and Remedios, G.. 1980. Chemical composition of pure and bee-collected pollen. Medlinger fra Norges Landbrukshoegskole 59: 2–12.Google Scholar
Sols, A., Cadenas, E., and Alvarado, F.. 1960. Enzymatic basis of mannose toxicity in honey bees. Science 131: 297–298.CrossRefGoogle ScholarPubMed
Stadler, B. and Dixon, A. F. G.. 1999. Ant attendance in aphids: why different degrees of myrmecophily? Ecological Entomology 24: 363–369.CrossRefGoogle Scholar
Stamp, N. E. and Bowers, M. D.. 1988. Direct and indirect effects of predatory wasps on gregarious larvae of the buckmoth, Hemileuca lucina (Saturnidae). Oecologia 75: 619–624.CrossRefGoogle Scholar
Stanley, R. G., and Linskens, H. F.. 1974. Pollen: Biology, Biochemistry, Management. Heidelberg, Germany: Springer-Verlag.CrossRefGoogle Scholar
Stanton, M. L. 2003. Interacting guilds: moving beyond the pairwise perspective on mutualisms. American Naturalist 162: S10–S23.CrossRefGoogle ScholarPubMed
Stapel, J. O., Cortesero, A. M., Moraes, C. M., Tumlinson, J. H., and Lewis, W. J.. 1997. Extrafloral nectar, honeydew and sucrose effects on searching behavior and efficiency of Microplitis croceipes (Hymenoptera: Braconidae) in cotton. Environmental Entomology 26: 617–623.CrossRefGoogle Scholar
Steinbauer, M. J. 1996. A note on manna feeding by ants (Hymenoptera: Formicidae). Journal of Natural History 30: 1185–1192.CrossRefGoogle Scholar
Stephens, M. J., France, C. M., Wratten, S. D., et al. 1998. Enhancing biological control of leafrollers (Lepidoptera: Tortricidae) by sowing buckwheat (Fagopyrum esculentum) in an orchard. Biocontrol Science and Technology 8: 547–558.CrossRefGoogle Scholar
Stephenson, A. G. 1982a. Iridoid glycosides in the nectar of Catalpa speciosa are unpalatable to nectar thieves. Journal of Chemical Ecology 8: 1025–1034.CrossRefGoogle Scholar
Stephenson, A. G. 1982b. The role of the extrafloral nectaries of Catalpa speciosa in limiting herbivory and increasing fruit production. Ecology 63: 663–669.CrossRefGoogle Scholar
Stoffolano, J. G., 1995. Regulation of a carbohydrate meal in the adult Diptera, Lepidoptera, and Hymenoptera. In Chapman, R. F. and Boer, G. (eds.) Regulatory Mechanisms in Insect Feeding. New York: Chapman and Hall, pp. 210–247.CrossRefGoogle Scholar
Sudd, J. H., and Franks, N. R.. 1987. The Behavioural Ecology of Ants. New York: Chapman and Hall.CrossRefGoogle Scholar
Swirski, E., Izhar, Y., Wysoki, M., Gurevitz, E., and Greenberg, S.. 1980. Integrated control of the long-tailed mealybug, Pseudococcus longispinus (Hom., Pseudococcidae), in avocado plantations in Israel. Entomophaga 25: 415–426.CrossRefGoogle Scholar
Takasu, K. and Lewis, W. J.. 1993. Host- and food-foraging of the parasitoid Microplitis croceipes: learning and physiological state effects. Biological Control 3: 70–74.CrossRefGoogle Scholar
Takasu, K. and Lewis, W. J.. 1995. Importance of adult food sources to host searching of the larval parasitoid Microplitis croceipes. Biological Control 5: 25–30.CrossRefGoogle Scholar
Takasu, K. and Lewis, W. J.. 1996. The role of learning in adult food location by the larval parasitoid, Microplitis croceipes. Journal of Insect Behavior 9: 265–281.CrossRefGoogle Scholar
Takeda, S., Kinomura, K. and Sakurai, H.. 1982. Effects of ant-attendance on the honeydew excretion and larviposition of the cowpea aphid, Aphis craccivora Koch. Applied Entomology and Zoology 17: 133–135.CrossRefGoogle Scholar
Tanowitz, B. D. and Koehler, D. L.. 1986. Carbohydrate analysis of floral and extrafloral nectars in selected taxa of Sansivieria (Agavaceae). Annals of Botany 58: 541.CrossRefGoogle Scholar
Thien, L. B., Bernhardt, P., Gibbs, G. W., et al. 1985. The pollination of Zygogynum (Winteraceae) by a moth, Sabatinca (Micropterigidae): an ancient association. Science 227: 540–543.CrossRefGoogle Scholar
Tilman, D. 1978. Cherries, ants, and tent caterpillars: timing of nectar production in relation to susceptibility of caterpillars to ant-predation. Ecology 59: 686–692.CrossRefGoogle Scholar
Tobin, J. E., 1994. Ants as primary consumers: diet and abundance in the Formicidae. In Hunt, J. H. and Nalepa, C. A. (eds.) Nourishment and Evolution in Insect Societies. Boulder, CO: Westview Press, pp. 279–307.Google Scholar
Triltsch, H. 1997. Contents in field sampled adults of Coccinella septempunctata (Col.: Coccinellidae). Entomophaga 42: 125–131.CrossRefGoogle Scholar
Tuckey, H. B., 1971. Leaching of substances from plants. In Peece, T. F. and Dickinson, C. H. (eds.) Ecology of Leaf Surface Organisms. New York: Academic Press, pp. 67–80.Google Scholar
Turlings, T. C. J. and F. L. Wäckers. 2004. Recruitment of predators and parasitoids by herbivore-injured plants. In Cardé, R. T. and Millar, J. (eds.) Advances in Insect Chemical Ecology. Cambridge, UK: Cambridge University Press, pp. 21–75.CrossRefGoogle Scholar
Turlings, T. C. J., Bernasconi, M., Bertossa, R., et al. 1998. The induction of volatile emissions in maize by three herbivore species with different feeding habits: possible consequences for their natural enemies. Biological Control 11: 122–129.CrossRefGoogle Scholar
Vaissière, B. E. and Vinson, S. B.. 1994. Pollen morphology and its effect on pollen collection by honey-bees, Apis mellifera L. (Hymenoptera, Apidae), with special reference to upland cotton, Gossypium hirsutum L. (Malvaceae). Grana 33: 128–138.CrossRefGoogle Scholar
Baalen, M., Krivan, V., Rijn, P. C. J., and Sabelis, M. W.. 2001. Alternative food, switching predators, and the persistence of predator–prey systems. American Naturalist 157: 512–524.Google ScholarPubMed
Dam, N. M., Harvey, J. A., Wäckers, F. L., et al. 2003. Interactions between aboveground and belowground induced responses against phytophages. Basic and Applied Ecology 4: 63–77.Google Scholar
Pijl, L. 1951. On the morphology of some tropical plants: Gloriosa, Bougainvillea, Honckenya and Rottboellia. Phytomorphology 1: 185–188.Google Scholar
Rijn, P. C. J. and Tanigoshi, L. K.. 1999a. The contribution of extrafloral nectar to survival and reproduction of the predatory mite Iphiseius degenerans on Ricinus communis. Experimental and Applied Acarology 23: 281–296.CrossRefGoogle Scholar
Rijn, P. C. J. and Tanigoshi, L. K.. 1999b. Pollen as food for the predatory mites Iphiseius degenerans and Neoseiulus cucumeris (Acari: Phytoseiidae): dietary range and life history. Experimental and Applied Acarology 23: 785–802.CrossRefGoogle Scholar
Rijn, P. C. J., Houten, Y. M., and Sabelis, M. W.. 2002. How plants benefit from providing food to predators even when it is also edible to herbivores. Ecology 83: 2664–2679.CrossRefGoogle Scholar
Vet, L. E. M., Wäckers, F. L., and Dicke, M.. 1991. How to hunt for hiding hosts: the reliability–detectability problem in foraging parasitoids. Netherlands Journal of Zoology 41: 202–213.CrossRefGoogle Scholar
Vinson, S. B. 1968. The distribution of an oil, carbohydrate and protein food source to members of the imported fire ant colony. Journal of Economic Entomology 61: 712–714.CrossRefGoogle Scholar
Vitzthum, H. G. 1930. Investigations of the causes of May sickness. Bee World 11: 14–15.Google Scholar
Völkl, W. 1992. Aphids or their parasitoids: who actually benefits from ant-attendance? Journal of Animal Ecology 61: 273–281.CrossRefGoogle Scholar
Völkl, W. 2001. Parasitoid learning during interactions with ants: how to deal with an aggressive antagonist. Behavioral Ecology and Sociobiology 49: 135–144.Google Scholar
Völkl, W. and Kraus, W.. 1996. Foraging behaviour and resource utilization of the aphid parasitoid Pauesia unilachni: adaptation to host distribution and mortality risks. Entomologia Experimentalis et Applicata 79: 101–109.CrossRefGoogle Scholar
Völkl, W. and Kroupa, A. S.. 1997. Effects of adult mortality risks on parasitoid foraging tactics. Animal Behaviour 54: 349–359.CrossRefGoogle Scholar
Völkl, W. and Mackauer, M.. 1993. Interactions between ants attending Aphis fabae ssp. cirsiiacanthoidis on thistles and foraging parasitoid wasps. Journal of Insect Behavior 6: 301–312.CrossRefGoogle Scholar
Völkl, W., Woodring, J., Fischer, M., Lorenz, M. W., and Hoffmann, K. H.. 1999. Ant–aphid mutualisms: the impact of honeydew production and honeydew sugar composition on ant preferences. Oecologia 118: 483–491.Google ScholarPubMed
Frisch, K. 1934. Über den Geschmackssinn der Biene: Ein Beitrag zur vergleichenden Physiologie des Geschmacks. Zeitschrift für vergleichenden Physiologie 21: 1–45.Google Scholar
Vrieling, K., Smit, W., and Meijden, E.. 1991. Tritrophic interactions between aphids (Aphis jacobaeae Schrank.), ant species, Tyria jacobaeae L., and Senecio jacobaea L. lead to maintenance of genetic variation in pyrrolizidine alkaloid concentration. Oecologia 86: 177–182.CrossRefGoogle Scholar
Wäckers, F. L. 1994. The effect of food deprivation on the innate visual and olfactory preferences in the parasitoid Cotesia rubecula. Journal of Insect Physiology 40: 641–649.CrossRefGoogle Scholar
Wäckers, F. L. 1999. Gustatory response by the hymenopteran parasitoid Cotesia glomerata to a range of nectar and honeydew sugars. Journal of Chemical Ecology 25: 2863–2877.CrossRefGoogle Scholar
Wäckers, F. L. 2000. Do oligosaccharides reduce the suitability of honeydew for predators and parasitoids? A further facet to the function of insect-synthesized honeydew sugars. Oikos 90: 197–201.CrossRefGoogle Scholar
Wäckers, F. L. 2001. A comparison of nectar and honeydew sugars with respect to their utilization by the hymenopteran parasitoid Cotesia glomerata. Journal of Insect Physiology 47: 1077–1084.CrossRefGoogle ScholarPubMed
Wäckers, F. L. 2003a. The effect of food supplements on parasitoid–host dynamics. Proc. 1st Int. Symp. Biological Control of Arthropods, Honolulu, pp. 226–231.
Wäckers, F. L. 2003b. The parasitoids' need for sweets: sugars in mass rearing and biological control. In Lenteren, J. C. (ed.) Quality Control of Natural Enemies. Wallingford, UK: CAB International, pp. 59–72.Google Scholar
Wäckers, F. L. 2004. Assessing the suitability of flowering herbs as parasitoid food sources: flower attractiveness and nectar accessibility. Biological Control 29: 307–314.CrossRefGoogle Scholar
Wäckers, F. L. and Bezemer, T. M.. 2003. Root herbivory induces an above-ground indirect defence. Ecology Letters 6: 9–12.CrossRefGoogle Scholar
Wäckers, F. L. and Bonifay, C.. 2004. How to be sweet? Extrafloral nectar allocation in Gossypium hirsutum fits optimal defense theory predictions. Ecology 85: 1512–1518.CrossRefGoogle Scholar
Wäckers, F. L. and Steppuhn, A.. 2003. Characterizing nutritional state and food source use of parasitoids collected in fields with high and low nectar availability. IOBC/WPRS Bulletin 26: 203–208.Google Scholar
Wäckers, F. L. and Swaans, C. P. M.. 1993. Finding floral nectar and honeydew in Cotesia rubecula: random or directed? Proceedings of the Section Experimental and Applied Entomology of the Netherlands Entomological Society 4: 67–72.Google Scholar
Wäckers, F. L. and Wunderlin, R.. 1999. Induction of cotton extrafloral nectar production in response to herbivory does not require a herbivore-specific elicitor. Entomologia Experimentalis et Applicata 91: 149–154.CrossRefGoogle Scholar
Wäckers, F. L., Björnsen, A. and Dorn, S.. 1996. A comparison of flowering herbs with respect to their nectar accessibility for the parasitoid Pimpla turionellae. Proceedings of the Section Experimental and Applied Entomology of the Netherlands Entomological Society 7: 177–182.Google Scholar
Wäckers, F. L., Bonifay, C., and Lewis, W. J.. 2002. Conditioning of appetitive behavior in the hymenopteran parasitoid Microplitis croceipes. Entomologia Experimentalis et Applicata 103: 135–138.CrossRefGoogle Scholar
Wäckers, F. L., Zuber, D., Wunderlin, R., and Keller, F.. 2001. The effect of herbivory on temporal and spatial dynamics of extrafloral nectar production in cotton and castor. Annals of Botany 87: 365–370.CrossRefGoogle Scholar
Wada, A., Isobe, Y., Yamaguchi, S., Yamaoka, R., and Ozaki, M.. 2001. Taste-enhancing effects of glycine on the sweetness of glucose: a gustatory aspect of symbiosis between the ant Camponotus japonicus and the larvae of the lycaenid butterfly Niphanda fusca. Chemical Senses 26: 983–992.CrossRefGoogle ScholarPubMed
Wagner, D. 1997. The influence of ant nests on Acacia seed production, herbivory and soil nutrients. Journal of Ecology 85: 83–93.CrossRefGoogle Scholar
Wagner, D. and Kay, A.. 2002. Do extrafloral nectaries distract ants from visiting flowers? An experimental test of an overlooked hypothesis. Evolutionary Ecology Research 4: 293–305.Google Scholar
Waller, G. D. 1972. Evaluating responses of honey bees to sugar solutions using an artificial-flower feeder. Annals of the Entomological Society of America 65: 857–862.CrossRefGoogle Scholar
Washburn, J. O. 1984. Mutualism between a cynipid gall wasp and ants. Ecology 65: 654–656.CrossRefGoogle Scholar
Wasserthal, L. T., 1993. Swing-hovering combined with long tongue in hovermoths, an antipredator adaptation during flower visits. In Barthlott, W., Naumann, C. M., Schmidt-Loske, K., and Schuchmann, K. L. (eds.) Animal–Plant Interactions in Tropical Environments. Bonn, Germany: Zoologische Forschungsinstitut und Museum Alexander Koenig, pp. 77–87.Google Scholar
Watt, W. B., Hoch, P. C., and Mills, S. G.. 1974. Nectar resource use by Colias butterflies: chemical and visual experiments. Oecologia 14: 353–374.CrossRefGoogle Scholar
Way, M. J. 1963. Mutualism between ants and honeydew producing Homoptera. Annual Review of Entomology 8: 307–344.CrossRefGoogle Scholar
Weber, G., Oswald, S., and Zöllner, U.. 1986. Die Wirtseignung von Rapssorten unterschiedlichen Glucosinolatsgehaltes für Brevicoryne brassicae (L.) und Myzus persicae. (Sulzer) (Hemiptera, Aphididae). Zeitschrift für Pflanzenkrankheiten und Pflanzenschutz 93: 113–124.Google Scholar
Weevers, T. 1952. Flower colours and their frequency. Acta Botanica Neerlandica 1: 81–92.CrossRefGoogle Scholar
Wehner, R. and M. V. Srinivasan. 1989. The world as the insect sees it. In Lewis, T. (ed.) Insect Communication. London: Academic Press, pp. 29–47.Google Scholar
Weisser, W. W., Houston, A. I., and Völkl, W.. 1994. Foraging strategies in solitary parasitoids: the trade-off between female and offspring mortality risks. Evolutionary Ecology 8: 587–597.CrossRefGoogle Scholar
Wellenstein, G. 1952. Zur Ernährungsbiologie der Roten Waldameise. Zeitschrift für Pflanzenkrankheiten und Pflanzenschutz: 430–451.Google Scholar
Whitman, D. W., 1996. Plant bodyguards: mutualistic interactions between plants and the third trophic level. In Ananthakrishnan, T. N. (ed.) Functional Dynamics of Phytophagous Insects. New Delhi: Oxford University Press / IBH Publishing, pp. 207–248.Google Scholar
Wilkinson, T. L., Ashford, D. A., Pritchard, J., and Douglas, A. E.. 1997. Honeydew sugars and osmoregulation in the pea aphid Acyrthosiphon pisum. Journal of Experimental Biology 200: 2137–2143.Google ScholarPubMed
Willmer, P. G. 1980. The effects of insect visitors on nectar constituents in temperate plants. Oecologia 47: 270–277.CrossRefGoogle ScholarPubMed
Willmer, P. G. and Stone, G. N.. 1997. How aggressive ant-guards assist seed-set in Acacia flowers. Nature 388: 165–167.CrossRefGoogle Scholar
Wink, M. and Römer, P.. 1986. Acquired toxicity: the advantages of specializing on alkaloid-rich lupins to Macrosiphon albifrons (Aphidae). Naturwissenschaften 73: 210–212.CrossRefGoogle Scholar
Wink, M. and Witte, L.. 1991. Storage of quinolizidine alkaloids in Macrosiphum albifrons and Aphis genistae (Homoptera: Aphididae). Entomologia Generalis 15: 237–254.CrossRefGoogle Scholar
Winkler, K., Wäckers, F. L., Valdivia, L. V., Larraz, V., and Lenteren, J. C.. 2003. Strategic use of nectar sources to boost biological control. IOBC/WPRS Bulletin 26: 209–214.Google Scholar
Wirth, R. and Leal, I. R.. 2001. Does rainfall affect temporal variability of ant protection in Passiflora coccinea? Ecoscience 8: 450–453.CrossRefGoogle Scholar
Wootton, J. T. and Sun, I. F.. 1990. Bract liquid as a herbivore defense mechanism for Heliconia wagneriana inflorescences. Biotropica 22: 155–159.CrossRefGoogle Scholar
Wratten, S. D., White, A. J., Bowie, M. H., Berry, N. A., and Weigmann, U.. 1995. Phenology and ecology of hoverflies (Diptera, Syrphidae) in New Zealand. Environmental Entomology 24: 595–600.CrossRefGoogle Scholar
Wunnachit, W., Jenner, C. F., and Sedgley, M.. 1992. Floral and extrafloral nectar production in Anacardium occidentale L (Anacardiaceae): an andromonoecious species. International Journal of Plant Sciences 153: 413–420.CrossRefGoogle Scholar
Yano, S. 1994. Flower nectar of an autogamous perennial Rorippa indica as an indirect defense mechanism against herbivorous insects. Researches in Population Ecology 36: 63–71.CrossRefGoogle Scholar
Yao, I. and Akimoto, S.. 2001. Ant attendance changes the sugar composition of the honeydew of the drepanosiphid aphid Tuberculatus quercicola. Oecologia 128: 36–43.CrossRefGoogle ScholarPubMed
Yao, I. and Akimoto, S.. 2002. Flexibility in the composition and concentration of amino acids in honeydew of the drepanosiphid aphid Tuberculatus quercicola. Ecological Entomology 27: 745–752.CrossRefGoogle Scholar
Yue, B. S. and Tsai, J. H.. 1996. Development, survivorship, and reproduction of Amblyseius largoensis (Acari: Phytoseiidae) on selected plant pollens and temperatures. Environmental Entomology 25: 488–494.CrossRefGoogle Scholar
Yue, B. S., Childers, C. C., and Fouly, A. H.. 1994. A comparison of selected plant pollens for rearing Euseius mesembrinus (Acari: Phytoseiidae). International Journal of Acarology 20: 103–108.CrossRefGoogle Scholar
Zimmerman, M. 1932. Über die extra floralen Nectarien der Angiospermen. Botanisches Zentralblatt Beihefte 49: 99–196.Google Scholar
Zimmerman, M. H., and H. Ziegler. 1975. List of sugars and sugar alcohols in sieve-tube exudates. In Zimmerman, M. H. and Milburn, J. A. (eds.) Encyclopedia of Plant Physiology vol. 1. New York: Springer-Verlag, pp. 480–503.Google Scholar
Zoebelein, G. 1956. Der Honigtau als Nahrung der Insekten. Zeitschrift für angewandte Entomologie 38: 369–416.CrossRefGoogle Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×