Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-n9wrp Total loading time: 0 Render date: 2024-07-16T16:11:02.700Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  08 January 2010

M. J. Lehane
Affiliation:
Liverpool School of Tropical Medicine
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
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

Aboul-Nasr, A. E. (1967) On the behaviour and sensory physiology of the bed bug (Cimex lectularius). I. Temperature reactions. Bull. Soc. Ent. Egypte, 51, 43–54Google Scholar
Adams, T. S. (1999) Hematophagy and hormone release. Ann. Ent. Soc. Am., 92, 1–13CrossRefGoogle Scholar
Adlington, D., Randolph, S. E. and Rogers, D. J. (1996) Flying to feed or flying to mate: gender differences in the flight activity of tsetse (Glossina palpalis). Phys. Ent., 21, 85–92CrossRefGoogle Scholar
Ahmadi, A. and McClelland, G. A. H. (1985) Mosquito-mediated attraction of female mosquitoes to a host. Phys. Ent., 10, 251–5CrossRefGoogle Scholar
Ahmed, A. M., Baggott, S. L., Maingon, R. and Hurd, H. (2002) The costs of mounting an immune response are reflected in the reproductive fitness of the mosquito Anopheles gambiae. Oikos, 97, 371–7CrossRefGoogle Scholar
Ahmed, A. M., Maingon, R., Romans, P. and Hurd, H. (2001) Effects of malaria infection on vitellogenesis in Anopheles gambiae during two gonotrophic cycles. Insect Mol. Biol., 10, 347–56CrossRefGoogle ScholarPubMed
Ahmed, A. M., Maingon, R. D., Taylor, P. J. and Hurd, H. (1999) The effects of infection with Plasmodium yoelii nigeriensis on the reproductive fitness of the mosquito Anopheles gambiae. Invertebrate Reproduction and Development, 36, 217–22CrossRefGoogle Scholar
Aikawa, M., Suzuki, M. and Gutierez, Y. (1980) Pathology of malaria. In Kreier, J. P. (ed.), Malaria. New York:Academic PressGoogle Scholar
Akai, H. and Sato, S. (1973) Ultrastructure of the larval hemocytes of the silkworm, Bombyx mori L. (Lepidoptera: Bombycidae). International Journal of Insect Morphology and Embryology, 2, 207–31CrossRefGoogle Scholar
Akman, L., Yamashita, A., Watanabe, H.et al. (2002) Genome sequence of the endocellular obligate symbiont of tsetse flies, Wigglesworthia glossinidia. Nature Genetics, 32, 402–7CrossRefGoogle ScholarPubMed
Aksoy, S. (2000) Tsetse – a haven for microorganisms. Parasitology Today, 16, 114–18CrossRefGoogle ScholarPubMed
Aksoy, S., Gibson, W. C. and Lehane, M. J. (2003) Perspectives on the interactions between tsetse and trypanosomes with implications for the control of trypanosomiasis. Advances in Parasitology, 53, 1–84CrossRefGoogle Scholar
Albritton, A. (1952) Standard Values in Blood. Phiadelphia: Saunders
Alekseev, A., Rasnityn, S. and Vitlin, L. (1977) On group behaviour of females of blood-sucking mosquitoes. Communication I. Discovery of the ‘effect’ of invitation. [In Russian]Parazitologiyai Parazitarnye Bolezni, 46, 23–4Google Scholar
Alekseyev, A. N., Abdullayev, I. T., Rasnitsyn, S. P. and Martsinovskiy, M. (1984) Comparison of flight capability of Aedes aegypti that are infected and not infected with plasmodia. Med. Parazitol., 1, 11–13Google Scholar
Allan, S. A., Day, J. F. and Edman, J. D. (1987) Visual ecology of biting flies. Ann. Rev. Ent., 32, 297–316CrossRefGoogle ScholarPubMed
Allan, S. A. and Stoffolano, J. G. (1986a) Effects of background contrast on visual attraction and orientation of Tabanus nigrovittatus (Diptera: Tabanidae). Environ. Entomol., 15, 689–94CrossRefGoogle Scholar
Allan, S. A. and Stoffolano, J. G. (1986b) The effects of hue and intensity on visual attraction of adult Tabanus nigrovittatus (Diptera: Tabanidae). J. Med. Ent., 23, 83–91CrossRefGoogle Scholar
Altman, P. L. and Dittmer, D. S.(1971) Blood and other body fluids. In Respiration and Circulation. Bethesda, MDI: Federation of American Societies of Experimental Biology, p. 540
Altner, H. and Loftus, R. (1985) Ultrastructure and function of insect thermo and hygroreceptors. Ann. Rev. Ent., 30, 273–95CrossRefGoogle Scholar
Amin, O. M. and Wagner, M. E. (1983) Further notes on the function of pronotal combs in fleas (Siphonaptera). Ann. Ent. Soc. Am., 76, 232–4CrossRefGoogle Scholar
Anderson, J. R. and Ayala, S. C. (1968) Trypanosome transmitted by a Phlebotomus: first report from the Americas. Science, 161, 1023–5CrossRefGoogle ScholarPubMed
Anderson, J. R. and Hoy, J. B. (1972) Relationship between host attack rates and CO2 baited insect flight trap catches of certain Symphoromyia spp. J. Med. Ent., 9, 373–92CrossRefGoogle Scholar
Anderson, R. A. and Brust, R. A. (1996) Blood feeding success of Aedes aegypti and Culex nigripalpus (Diptera: Culicidae) in relation to defensive behavior of Japanese quail (Coturnix japonica) in the laboratory. Journal of Vector Ecology, 21, 94–104Google Scholar
Anderson, R. A., Koella, J. C. and Hurd, H. (1999) The effect of Plasmodium yoelii nigeriensis infection on the feeding persistence of Anopheles stephensi Liston throughout the sporogonic cycle. Proc. R. Soc. Lond. B Biol. Sci., 266, 1729–33CrossRefGoogle ScholarPubMed
Anderson, R. C. (1957) The life cycles of Dipetalonematid nematodes (Filarioidea, Dipetalonematidae): the problem of their evolution. J. Helminth., 31, 203–24CrossRefGoogle ScholarPubMed
Anderson, R. M. (1986) Genetic variability in resistance to parasitic invasion: population implications for invertebrate host species. In Lackie, A. M. (ed.), Immune Mechanisms in Invertebrate Vectors, Vol. 56, Oxford: Oxford University PressGoogle Scholar
Anderson, R. M. and May, R. M. (1978) Regulation and stability of host–parasite population interactions. Journal of Animal Ecology, 47, 219–47CrossRefGoogle Scholar
Ansell, J., Hamilton, K. A., Pinder, M., Walraven, G. E. L. and Lindsay, S. W. (2002) Short-range attractiveness of pregnant women to Anopheles gambiae mosquitoes. Trans. R. Soc. of Trop. Med. and Hyg., 96, 113–16CrossRefGoogle ScholarPubMed
Arlian, L. (2002) Arthropod allergens and human health. In Annual Review of Entomology. Annual Reviews Ic., Palo Alto, Vol. 47, 395–434
Arrese, E. L., Canavoso, L. E., Jouni, Z. E.et al. (2001) Lipid storage and mobilization in insects: current status and future directions. Insect Biochemistry and Molecular Biology, 31, 7–17CrossRefGoogle ScholarPubMed
Aschner, M. (1932) Experimentelle unter Suchungen über die Symbiose der Kleiderlaus. Naturwiss., 20, 501–5CrossRefGoogle Scholar
Aschner, M. (1934) Studies on the symbiosis of the body louse. I. Elimination of the symbionts by centrifugation of the eggs. Parasitol., 26, 309–14CrossRefGoogle Scholar
Aschner, M. (1946) The symbiosis of Eucampsipoda aegyptica Mcq. Bull. Soc. Ent. Egypte, 30, 1–6Google Scholar
Ashford, R. W. (2001) The Leishmania ses. In M. W. Service (ed.), Encyclopedia of Arthropod-transmitted Infections of Man and Domesticated Animals. CABI, 269–79
Ashida, M., Ochiai, M. and Niki, T. (1988) Immunolocalization of prophenoloxidase among hemocytes of the Silkworm, Bombyx mori. Tissue and Cell, 20, 599–610Google ScholarPubMed
Audy, J. R., Radovsky, F. J. and Vercammen-Grandjean, P. H. (1972) Neosomy: radical intrastadial metamorphosis associated with arthropod symbioses. J. Med. Ent., 9, 487–94CrossRefGoogle ScholarPubMed
Avila, A., Silverman, N., Diaz-Meco, M. T. and Moscat, J. (2002) The Drosophila atypical protein kinase C-ref(2)p complex constitutes a conserved module for signaling in the toll pathway. Mol. Cell. Biol. 22, 8787–95CrossRefGoogle ScholarPubMed
Ayala, S. C. and Lee, D. (1970) Saurian malaria: development of sporozoites in two species of phelebotomine sandflies. Science, 167, 891–2CrossRefGoogle Scholar
Bacot, A. M. and Martin, C. J. (1914) Observations on the mechanism of the transmission of the plague by fleas. J. Hyg., 13, 423–39Google ScholarPubMed
Bailey, L. (1952) The action of the proventriculus of the worker honeybee. J. Exp. Biol., 29, 310–27Google Scholar
Baker, J. R. (1965) The evolution of parasitic protozoa. In Taylor, A. E. R. (ed.), of the British Society for Parasitology. Oxford: Blackwell, Vol. 3Google Scholar
Baker, R. C. (1986) Pheromane-modulated movement of flying moths. In Payne, T. L., Birch, M. C. and Kennedy, C. E. J. (eds.), Mechanisms in Insect Olfaction. Oxford: Clarendon PressGoogle Scholar
Baker, T. C. (1990) Upwind flight and casting flight: complementary phasic and tonic systems used for location of sex pheromone sources by male moths. In K. B. Doving (ed.), Symposium on Olfaction and Taste, Oslo, Vol. X
Balashov, Y. (1984) Interaction between blood-sucking arthropods and their hosts, and its influence on vector potential. Ann. Rev. Ent., 29, 137–56CrossRefGoogle ScholarPubMed
Ball, G. H. (1943) Parasitism and evolution. Am. Nat., 77, 345–64CrossRefGoogle Scholar
Banziger, H. (1971) Blood-sucking moths of Malaya. Fauna, 1, 5–16Google Scholar
Baranov, N. (1935) New information on the Golubatz fly, S. columbaczense. Rev. Appl. Ent. B, 23, 275–6Google Scholar
Barrera, A. (1966) Hallazgo de Amblyopinus tiptoni Barrera, 1966 en Costa Rica, A. C. (Col.: Staph.). Acta zool. Mex., 8, 1–3Google Scholar
Barrera, A. and Machado-Allison, C. E. (1965) Coleopteros ectoparasiticos de Mamiferos. Ciencia Mexico, 23, 201–8Google Scholar
Bar-Zeev, M., Maibach, H. I. and Khan, A. A. (1977) Studies on the attraction of Aedes aegypti (Diptera: Culicidae) to man. J. Med. Ent., 14, 113–20CrossRefGoogle Scholar
Baudisch, K. (1958) Beitraäge zur Zytologie und Embryologie einiger Insektensymbiosen. Zeit. Morph. Okol. Tiere, 47, 436–88CrossRefGoogle Scholar
Baylis, M. and Mbwabi, A. L. (1995) Feeding-behavior of tsetse-flies (Glossina pallidipes Austen) on Trypanosoma-infected oxen in Kenya. Parasitology, 110, 297–305CrossRefGoogle ScholarPubMed
Beach, R., Kiilu, G. and Leeuwenburg, J. (1985) Modification of sandfly biting behaviour by Leishmania leads to increased parasite transmission. Am. J. Trop. Med. Hyg., 34, 279–83CrossRefGoogle ScholarPubMed
Beard, C. B., Mason, P. W., Aksoy, S., Tesh, R. B. and Richards, F. F. (1992) Transformation of an insect symbiont and expression of a foreign gene in the Chagas-disease vector Rhodnius prolixus. Am. J. Trop. Med. and Hyg., 46, 195–200CrossRefGoogle Scholar
Beckenbach, A. T. and Borkent, A. (2003) Molecular analysis of the biting midges (Diptera: Ceratopogonidae), based on mitochondrial cytochrome oxidase subunit 2. Mol. Phylogenet. Evol. 27, 21–35CrossRefGoogle Scholar
Beckett, E. B. and Macdonald, W. W. (1971) The development and survival of subperiodic Brugia malayi and B. pahangi larvae in a selected strain of Aedes aegypti. Trans. R. Soc. Trop. Med. Hyg., 65, 339–46CrossRefGoogle Scholar
Beenakkers, A. M. T., Horst, D. J. and Marrewijk, W. J. A. (1984) Insect flight muscle metabolism. Insect Biochemistry, 14, 243–60CrossRefGoogle Scholar
Beerntsen, B. T., James, A. A. and Christensen, B. M. (2000) Genetics of mosquito vector competence. Microbiology and Molecular Biology Reviews, 64, 115–37CrossRefGoogle ScholarPubMed
Beerntsen, B. T., Severson, D. W., Klinkhammer, J. A., Kassner, V. A. and Christensen, B. M. (1995) Aedes aegypti: A quantitative trait locus (QTL) influencing filarial worm intensity is linked to QTL for susceptibility to other mosquito-borne pathogens. Experimental Parasitology, 81, 355–62CrossRefGoogle ScholarPubMed
Beklemishev, V. N. (1957) Some general questions on the biology of bloodsucking lower flies. Meditsinskaya parazitologiya i parazitarnye bolezni, 5, 562–6Google Scholar
Belkaid, Y., Valenzuela, J. G., Kamhawi, S.et al. (2000) Delayed-type hypersensitivity to Phlebotomus papatasi sand fly bite: An adaptive response induced by the fly?Proceedings of the National Academy of Sciences of the United States of America, 97, 6704–9CrossRefGoogle Scholar
Bell, J. F., Stewart, S. J. and Nelson, W. A. (1982) Transplant of acquired resistance to Polyplax serrata (Phthiraptera: Hoplopleuridae) in skin allografts to athymic mice. J. Med. Ent., 19, 164–8CrossRefGoogle ScholarPubMed
Belzer, W. R. (1978a) Factors conducive to increased protein feeding by the blowfly Phormia regina. Phys. Ent. 3, 251–7CrossRefGoogle Scholar
Belzer, W. R. (1978b) Patterns of selective protein ingestion by the blowfly Phormia regina. Phys. Ent., 3, 169–257CrossRefGoogle Scholar
Belzer, W. R.(1978c) Recurrent nerve inhibition of protein feeding in the blowfly Phormia regina. Phys. Ent., 3, 259–63CrossRefGoogle Scholar
Belzer, W. R. (1979) Abdominal stretch receptors in the regulation of protein ingestion by the black blowfly, Phormia regina. Phys. Ent., 4, 7–14CrossRefGoogle Scholar
Bennet-Clark, H. C. (1963a) The control of meal size in the blood sucking bug, Rhodnius prolixus. J. Exp. Biol., 40, 741–50Google Scholar
Bennet-Clark, H. C. (1963b) Negative pressures produced in the pharyngeal pump of the blood-sucking bug, Rhodnius prolixus. J. Exp. Biol., 40, 223–9Google Scholar
Bennett, G. F., Fallis, A. M. and Campbell, A. G. (1972) The response of Simulium (EuSimulium) euryadminiculum Davies (Diptera: Simuliidae) to some olfactory and visual stimuli. Can J. Zool., 50, 793–800CrossRefGoogle Scholar
Bequaert, J. C. (1953) The Hippoboscidae or house-flies (Diptera) of mammals and birds. Part 1. Structure, physiology and natural history. Entomologia Americana, 32, 1–209Google Scholar
Bergman, D. K. (1996) Mouthparts and feeding mechanisms of haematophagous arthropods. In Wikel, S. K. (ed.), Immunology of Host-Ectoparasitic Arthropod Relationships. Wallingford: CABInternational, 30–61Google Scholar
Besansky, N. J. (1999) Complexities in the analysis of cryptic taxa within the genus Anopheles. Parasitologia, 41, 97–100Google ScholarPubMed
Bidlingmayer, W. L. (1994) How mosquitos see traps – role of visual responses. J. Am. Mosq. Control Assoc., 10, 272–9Google ScholarPubMed
Bidlingmayer, W. L., Day, J. F. and Evans, D. G. (1995) Effect of wind velocity on suction trap catches of some Florida mosquitoes. J. Am. Mosq. Control Assoc., 11, 295–301Google ScholarPubMed
Bidlingmayer, W. L. and Hem, D. G. (1979) Mosquito (Diptera: Culicidae) flight behaviour near conspicuous objects. Bull. Ent. Res., 69, 691–700CrossRefGoogle Scholar
Bidlingmayer, W. L. and Hem, D. G. (1980) The range of visual attraction and the effect of competitive visual attractants upon mosquito (Diptera: Culicidae) flight. Bull. Ent. Res., 70, 321–42CrossRefGoogle Scholar
Biessmann, H., Walter, M. F., Dimitratos, S. and Woods, D. (2002) Isolation of cDNA clones encoding putative odourant binding proteins from the antennae of the malaria-transmitting mosquito, Anopheles gambiae. Insect Mol. Biol., 11, 123–32CrossRefGoogle ScholarPubMed
Billingsley, P. B. (1990) The midgut ultrastructure of haematophagous arthropods. Ann. Rev. Ent., 35, 219–48CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1983) Ultrastructural changes in posterior midgut cells associated with blood-feeding in adult female Rhodnius prolixus Stal (Hemiptera: Reduviidae). Can J. Zool., 61, 1175–87CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1985) Cellular localisation of aminopeptidase in the midgut of Rhodnius prolixus Stal (Hemiptera: Reduviidae) during blood digestion. Cell and Tissue Research, 241, 421–8CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1986) The surface morphology of the midgut cells of Rhodnius prolixus Stal (Hemiptera: Reduviidae) during blood digestion. Acta Trop., 43, 355–66Google ScholarPubMed
Billingsley, P. F. and Downe, A. E. R. (1988) Ultrastructural localisation of cathepsin B in the midgut of Rhodnius prolixusStal (Hemiptera: Reduviidae) during blood digestion. International Journal of Insect Morphology and Embryology, 17, 295–302CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1989) Changes in the anterior midgut cells of adult female Rhodnius prolixus Stal. (Hemiptera: Reduviidae) after feeding. J. Med. Ent., 26, 104–8CrossRefGoogle Scholar
Billingsley, P. F. and Rudin, W. (1992) The role of the mosquito peritrophic membrane in bloodmeal digestion and infectivity of Plasmodium species. J. Parasit., 78, 430–40CrossRefGoogle ScholarPubMed
Bissonnette, E. Y., Rossignol, P. A. and Befus, A. D. (1993) Extracts of mosquito salivary gland inhibit tumour necrosis factor alpha. Parasite Immunology, 15, 27–33CrossRefGoogle ScholarPubMed
Bitkowska, E., Dzbenski, T. H., Szadziewska, M. and Wegner, Z. (1982) Inhibition of xenograft rejection reaction in the bug Triatoma infestans during infection with a protozoan, Tryapnosoma cruzi. J. Invert. Path., 40, 186–9CrossRefGoogle Scholar
Bize, P., Roulin, A. and Richner, H. (2003) Adoption as an offspring strategy to reduce ectoparasite exposure. Proc. R. Soc. Lond. B. Biol. Sci., 270 Suppl 1, 114–16CrossRefGoogle ScholarPubMed
Blackwell, A. (2000) Scottish biting midges: tourist attraction or deterrent?Antenna, 24, 144–50Google Scholar
Blackwell, A. and Page, S. (2003) Managing tourist health and Safety in the new millennium: global perspectives. In Managing Tourist Health and Safety in the New Millennium. Pergamon. 177–96CrossRefGoogle Scholar
Blandin, S., Moita, L. F., Kocher, T.et al. (2002) Reverse genetics in the mosquito Anopheles gambiae: targeted disruption of the Defensin gene. EMBO Report, 3, 852–6CrossRefGoogle ScholarPubMed
Boatin, B. A. (2003) The current state of the Onchocerciasis Control Programme in West Africa. Trop. Doct., 33, 209–14CrossRefGoogle ScholarPubMed
Boete, C., Paul, R. E. L. and Koella, J. C. (2002) Reduced efficacy of the immune melanization response in mosquitoes infected by malaria parasites. Parasitology, 125, 93–8CrossRefGoogle ScholarPubMed
Bos, H. J. and Laarman, J. J. (1975) Guinea pig lysine, cadaverine, and estradiol as attractants for the malaria mosquito Anopheles stephensi. Ent. Exp. Appl., 18, 161–72CrossRefGoogle Scholar
Bosch, O. J., Geier, M. and Boeckh, J. (2000) Contribution of fatty acids to olfactory host finding of female Aedes aegypti. Chemical Senses, 25, 323–30CrossRefGoogle ScholarPubMed
Bosio, C. F., Beaty, B. J. and Black, W. C. (1998) Quantitative genetics of vector competence for Dengue-2 virus in Aedes aegypti. Am. J. Trop. Med. Hyg., 59, 965–70CrossRefGoogle ScholarPubMed
Bosio, C. F., Fulton, R. E., Salasek, M. L., Beaty, B. J. and Black, W. C. T. (2000) Quantitative trait loci that control vector competence for dengue-2 virus in the mosquito Aedes aegypti. Genetics, 156, 687–98Google ScholarPubMed
Bossard, R. L. (2002) Speed and Reynolds number of jumping cat fleas (Siphonaptera: Pulicidae). Journal of the Kansas Entomological Society, 75, 52–4Google Scholar
Boulanger, N., Munks, R. J., Hamilton, J. V.et al. (2002) Epithelial innate immunity. A novel antimicrobial peptide with antiparasitic activity in the blood-sucking insect Stomoxys calcitrans. J. Biol. Chem., 277, 49921–6CrossRefGoogle ScholarPubMed
Boutros, M., Agaisse, H. and Perrimon, N. (2002) Sequential activation of signaling pathways during innate immune responses in Drosophila. Dev. Cell, 3, 711–22CrossRefGoogle ScholarPubMed
Bozza, M., Soares, M. B., Bozza, P. T.et al. (1998) The PACAP-type I receptor agonist maxadilan from sand fly saliva protects mice against lethal endotoxemia by a mechanism partially dependent on IL-10. Eur. J. Immunol., 28, 3120–73.0.CO;2-3>CrossRefGoogle ScholarPubMed
Bracken, G. K., Hanec, W. and Thorsteinson, A. J. (1962) The orientation of horseflies and deerflies (Tabanidae: Diptera). II. The role of some visual factors in the attractiveness of decoy silhouettes. Can. J. Zool., 40, 685–95CrossRefGoogle Scholar
Bracken, G. K. and Thorsteinson, A. J. (1965) The orientation behaviour of horse flies and deer flies (Tabanidae: Diptera). IV. The influence of some physical modifications of visual decoys on orientation of horse flies. Ent. Exp. Appl., 8, 314–18CrossRefGoogle Scholar
Bradbury, W. C. and Bennett, G. F. (1974) Behaviour of adult Simuliidae (Diptera). I. Response to colour and shape. Can. J. Zool., 52, 251–9CrossRefGoogle Scholar
Bradley, G. H. (1935) Notes on the southern buffalo gnat Eusimulium pecuarum (Riley) (Diptera: Simuliidae). Proc. Ent. Soc. Washington, 37, 60–4Google Scholar
Bradshaw, W. E. (1980) Blood-feeding and capacity for increase in the pitcher plant mosquito, Wyeomyia smithii. Environ. Entomol, 9, 86–9CrossRefGoogle Scholar
Bradshaw, W. E. and Holtzapfel, C. M. (1983) Life cycle strategies in Wyeomyia smithii: seasonal and geographic adaptations. In Brown, V. K. and Hodek, I. (eds.), Diapause and Life Cycle Strategies in Insects. The Hague: JunkGoogle Scholar
Brady, J. (1972) The visual responsiveness of the tsetse fly Glossina morsitans Westw. (Glossinidae) to moving objects: the effects of hunger, sex, host odour and stimulus characteristics. Bull. Ent. Res., 62, 257–79CrossRefGoogle Scholar
Brady, J. (1973) Changes in the probing responsiveness of starving tsetse flies (Glossina morsitans Westw.) (Diptera, Glossinidae). Bull. Ent. Res., 63, 247–55CrossRefGoogle Scholar
Brady, J. (1975) 'Hunger' in the tsetse fly: the nutritional correlates of behaviour. J. Insect Physiol., 21, 807–29CrossRefGoogle ScholarPubMed
Brady, J., Costantini, C., Sagnon, N., Gibson, G. and Coluzzi, M. (1997) The role of body odours in the relative attractiveness of different men to malarial vectors in Burkina Faso. Ann. Trop. Med. Parasit., 91, S121–S122CrossRefGoogle Scholar
Brady, J., Gibson, G. and Packer, M. J. (1989) Odour movement, wind direction and the problem of host finding by tsetse flies. Phys. Ent., 14, 369–380CrossRefGoogle Scholar
Brady, J., Griffiths, N. and Paynter, Q. (1995) Wind speed effects on odour source location by tsetse flies (Glossina). Phys. Ent., 20, 293–302CrossRefGoogle Scholar
Brady, J. and Shereni, A. (1988) Landing responses of the tsetse fly Glossina morsitans morsitans Weidemann and the stablefly Stomoxys calcitrans (L.). (Diptera: Glossinidae & Muscidae) to black and white patterns: a laboratory study. Bull. Ent. Res., 78, 301–11CrossRefGoogle Scholar
Braks, M. A. H., Scholte, E. J., Takken, W. and Dekker, T. (2000) Microbial growth enhances the attractiveness of human sweat for the malaria mosquito, Anopheles gambiae sensu stricto (Diptera: Culicidae). Chemoecology, 10, 129–34CrossRefGoogle Scholar
Braun, A., Hoffmann, J. A. and Meister, M. (1998) Analysis of the Drosophila host defense in domino mutant larvae, which are devoid of hemocytes. Proceedings of the National Academy of Sciences of the United States of America, 95, 14337–42CrossRefGoogle ScholarPubMed
Bray, R. S. (1963) The exo-erythrocytic phase of malaria parasites. Int. Rev. Trop. Med., 2, 41Google Scholar
Bray, R. S., McCrae, A. W. R. and Smalley, M. E. (1976) Lack of a circadian rhythm in the ability of the gametocytes of Plasmodium falciparum to infect Anopheles gambiae. Int. J. Parasit., 6, 399–401CrossRefGoogle ScholarPubMed
Brecher, G. and Wigglesworth, V. B. (1944) The transmission of Actinomyces rhodnii Erkison in Rhodnius prolixus Stal. (Hemiptera) as its influence on the growth of the host. Parasitol., 35, 220–4CrossRefGoogle Scholar
Breev, K. V. (1950) The behaviour of blood-sucking Diptera and warble flies when attacking reindeer and the responsive reactions of reindeer. [In Russian] Parasitologicheskii Sbornik, 12, 167–89
Brehelin, M. (1982) Comparative study of structure and function of blood cells from two Drosophila species. Cell and Tissue Research, 221, 607–15CrossRefGoogle ScholarPubMed
Brehelin, M. (1986) Insect haemocytes: a new classification to rule out the controversy. In Brehelin, M. (ed.), in Invertebrates. Berlin, Heidelberg, New York, Tokyo: Springer Verlag Press, 36–49Google Scholar
Brehelin, M., Zachary, D. and Hoffmann, J. A. (1978) A comparative ultrastructural study of blood cells from nine insect orders. Cell and Tissue Research, 195, 45–57CrossRefGoogle ScholarPubMed
Briegel, H., Hefti, A. and DiMarco, E. (2002) Lipid metabolism during sequential gonotrophic cycles in large and small female Aedes aegypti. J. Insect Physiol., 48, 547–54CrossRefGoogle ScholarPubMed
Briegel, H., Knusel, I. and Timmermann, S. E. (2001) Aedes aegypti: size, reserves, survival, and flight potential. Journal of Vector Ecology, 26, 21–31Google ScholarPubMed
Brook, M. L. (1985) The effect of allopreening on tick burdens of moulting eudyptid penguins. Auk, 102, 893Google Scholar
Browne, S. M. and Bennett, G. F. (1980) Colour and shape as mediators of host-seeking responses of simuliids and tabanids (Diptera) in the Tantramar marshes, New Brunswick, Canada. Can. J. Med. Entomol., 17, 58–62CrossRefGoogle Scholar
Browne, S. M. and Bennett, G. F. (1981) Response of mosquitoes (Diptera: Culicidae) to visual stimuli. J. Med. Ent., 6, 505–21CrossRefGoogle Scholar
Bruce, D. (1895) Preliminary Report on the Tsetse Fly Disease or Nagana, in Zululand. Durban: Bennett and Davis
Bruce, D. and Nabarro, D. (1903) Progress report on sleeping sickness in Uganda
Buchner, P. (1965) Endosymbiosis of Animals with Plant Microorganisms. New York: Wiley
Budd, L. T. (1999) DFID-Funded Tsetse and Trypanosomiasis Research and Development since 1980 (V. 2. Economic Analysis). London: Department for International Development
Bulet, P., Hetru, C., Dimarcq, J. L. and Hoffmann, D. (1999) Antimicrobial peptides in insects; structure and function. Developmental and Comparative Immunology, 23, 329–44CrossRefGoogle ScholarPubMed
Bungener, W. and Muller, G. (1976) Adharenz-Phänomene bei Trypanosoma congolense. Tropenmed. Parasitol., 27, 307–71Google Scholar
Burg, J. G., Knapp, F. W. and Silapanuntakul, S. (1993) Feeding Haematobia irritans (Diptera, Muscidae) adults through a nylon-reinforced silicone membrane. J. Med. Ent., 30, 462–6CrossRefGoogle ScholarPubMed
Burgess, I., Maunder, J. W. and Myint, T. T. (1983) Maintenance of the crab louse, Pthirus pubis, in the laboratory and behavioural studies using volunteers. Community Med., 5, 238–41Google Scholar
Burkett, D. A., Butler, J. F. and Kline, D. L. (1998) Field evaluation of colored light-emitting diodes as attractants for woodland mosquitoes and other diptera in north central Florida. J. Am. Mosq. Control Assoc., 14, 186–95Google ScholarPubMed
Burkhart, C. N., Stankiewicz, B. A., Pchalek, I., Kruge, M. A. and Burkhart, C. G. (1999) Molecular composition of the louse sheath. J. Parasit., 85, 559–61CrossRefGoogle ScholarPubMed
Burkot, T. R. (1988) Non-random host selection by anopheline mosquitoes. Parasitology Today, 4, 156–62CrossRefGoogle ScholarPubMed
Burkot, T. R., Narara, A., Paru, R., Graves, P. M. and Garner, P. (1989) Human host selection by anophelines: no evidence for preferential selection of malaria or microfilariae-infected individuals in a hyperendemic area. Parasitology, 98, 337–42CrossRefGoogle ScholarPubMed
Bursell, E. (1961) The behaviour of tsetse flies (Glossina swynnertoni Austen) in relation to problems of sampling. Proc. R. Ent. Soc. Lond. (A), 36, 9–20Google Scholar
Bursell, E. (1975) Substrates of oxidative metabolism in dipteran flight muscle. Comp. Biochem. Physiol., 52, 235–8Google ScholarPubMed
Bursell, E. (1977) Synthesis of proline by fat body of the tsetse fly (Glossina morsitans). Metabolic pathways. Insect Biochem., 7, 427–34CrossRefGoogle Scholar
Bursell, E. (1981) Energetics of haematophagous arthropods: influence of parasites. Parasitology, 82, 107–10Google Scholar
Bursell, E. (1984) Effects of host odour on the behaviour of tsetse. Insect Sci. Applic., 5, 345–9Google Scholar
Bursell, E. (1987) The effect of wind-borne odours on the direction of flight in tsetse flies, Glossina spp. Phys. Ent., 12, 149–56CrossRefGoogle Scholar
Bursell, E. and Taylor, P. (1980) An energy budget for Glossina (Diptera: Glossinidae). Bull. Ent. Res., 70, 187–96CrossRefGoogle Scholar
Burton, R. (1860) The Lake Regions of Central Africa. London: Longman
Butt, T. M. and Shields, K. S. (1996) The structure and behavior of gypsy moth (Lymantria dispar) hemocytes. J. Invert. Path., 68, 1–14CrossRefGoogle ScholarPubMed
Buxton, P. A. (1930) The biology of a blood-sucking bug, Rhodnius prolixus. Trans. R. Soc. Trop. Med. Hyg., 78, 227–36Google Scholar
Buxton, P. A. (1947) The Louse. London: Edward Arnold
Buxton, P. A. (1948) Experiments with lice and fleas. I. The baby mouse. Parasitol., 39, 119–24CrossRefGoogle Scholar
Canyon, D. V., Hii, J. L. K. and Muller, R. (1998) Multiple host-feeding and biting persistence of Aedes aegypti. Ann. Trop. Med. Parasit., 92, 311–16CrossRefGoogle ScholarPubMed
Cappello, M., Li, S., Chen, X. O.et al. (1998) Tsetse thrombin inhibitor: bloodmeal-induced expression of an anticoagulant in salivary glands and gut tissue of Glossina morsitans morsitans. Proceedings of the National Academy of Sciences of the United States of America, 95, 14290–5CrossRefGoogle ScholarPubMed
Carde, R. T. (1996) Odour plumes and odour-mediated flight in insects. In Bock, G. R. and G. Cardew (eds.), Olfaction in Mosquito-Host Interactions. Chichester: Wiley Ciba Foundation Symposium 200, 54–70Google Scholar
Carwardine, S. L. and Hurd, H. (1997) Effects of Plasmodium yoelii nigeriensis infection on Anopheles stephensi egg development and resorption. Med. Vet. Entomol, 11, 265–9CrossRefGoogle ScholarPubMed
Castanera, M. B., Aparicio, J. P. and Gurtler, R. E. (2003) A stage-structured stochastic model of the population dynamics of Triatoma infestans, the main vector of Chagas disease. Ecological Modelling, 162, 33–53CrossRefGoogle Scholar
Caterino, M. S., Cho, S. and Sperling, F. A. H. (2000) The current state of insect molecular systematics: a thriving Tower of Babel. Ann. Rev. Ent., 45, 1–54CrossRefGoogle Scholar
Cavanaugh, D. C. (1971) Specific effect of temperature upon transmission of the plague bacillus by the oriental rat flea, Xenopsylla cheopis. Am. J. Trop. Med. Hyg., 20, 264–73CrossRefGoogle ScholarPubMed
Chadee, D. D. and Beier, J. C. (1997) Factors influencing the duration of blood-feeding by laboratory-reared and wild Aedes aegypti (Diptera: Culicidae) from Trinidad, West Indies. Ann. of Trop. Med. Parasit., 91, 199–207CrossRefGoogle ScholarPubMed
Chadee, D. D., Beier, J. C. and Martinez, R. (1996) The effect of the cibarial armature on blood meal haemolysis of four anopheline mosquitoes. Bull. Ent. Res., 86, 351–4CrossRefGoogle Scholar
Chagas, C. (1909) Ueber eine neue Trypanosomiasis des Menschen. Memorias Institut Oswaldo Cruz, 1, 159–218CrossRef
Challier, A., Eyraud, M., Lafaye, A. and Laveissiere, C. (1977) Amelioration due rendement du piege biconique pour glossines (Diptera, Glossinidae) par l'emploi d'un cone inférieur bleu. Cah. ORSTOM, Ser. Entomol. Med. Parasitol., 15, 283–6Google Scholar
Champagne, D. E., Nussenzveig, R. H. and Ribeiro, J. M. C. (1995a) purification, partial characterization, and cloning of nitric oxide-carrying heme-proteins (nitrophorins) from salivary-glands of the bloodsucking insect Rhodnius prolixus. J. Biol. Chem., 270, 8691–5CrossRefGoogle Scholar
Champagne, D. E. and Ribeiro, J. M. C. (1994) Sialokinin-I and Sialokinin-Ii – Vasodilatory Tachykinins from the Yellow-Fever Mosquito Aedes aegypti. Proceedings of the National Academy of Sciences of the United States of America, 91, 138–42CrossRefGoogle ScholarPubMed
Champagne, D. E., Smartt, C. T., Ribeiro, J. M. C. and James, A. A. (1995b) The salivary gland-specific apyrase of the mosquito Aedes aegypti is a member of the 5ʹ-nucleotidase family. Proceedings of the National Academy of Sciences of the United States of America, 92, 694–8CrossRefGoogle Scholar
Chang, Y. H. and Judson, C. L. (1977) The role of isoleucine in differential egg production by the mosquito Aedes aegypti Linnaeus (Diptera: Culicidae) following feeding on human or guinea pig blood. Comp. Biochem. Physiol. A., 57, 23–8CrossRefGoogle Scholar
Chapman, R. F. (1961) Some experiments to determine the methods used in host finding by tsetse flies, Glossina medicorum. Bull. Ent. Res., 52, 83–97CrossRefGoogle Scholar
Chapman, R. F. (1982) Chemoreception: the significance of receptor numbers. Adv. Insect Physiol., 16, 247–356CrossRefGoogle Scholar
Charlab, R., Rowton, E. D. and Ribeiro, J. M. C. (2000) The salivary adenosine deaminase from the sand fly Lutzomyia longipalpis. Experimental Parasitology, 95, 45–53CrossRefGoogle ScholarPubMed
Charlwood, J. D., Billingsley, P. F. and Hoc, T. Q. (1995a) Mosquito-mediated attraction of female European but not African mosquitos to hosts. Ann. Trop. Med. Parasit., 89, 327–9CrossRefGoogle Scholar
Charlwood, J. D., Smith, T., Kihonda, J. (1995b) Density-independent feeding success of malaria vectors (Diptera, Culicidae) in Tanzania. Bull. Ent. Res., 85, 29–35CrossRefGoogle Scholar
Chen, C. C. and Chen, C. S. (1995) Brugia pahangi: effects of melanization on the uptake of nutrients by microfilariae in vitro. Experimental Parasitology, 81, 72–8CrossRefGoogle ScholarPubMed
Chen, C. C. and Laurence, B. R. (1985) An ultrastructural study on the encapsulation of microfilariae of Brugia pahangi in the haemocoel of Anopheles quadrimaculatus. Int. J. Parasitol, 15, 421–8CrossRefGoogle ScholarPubMed
Chikilian, M. L., Bradley, T. J., Nayar, J. K., Cashclark, C. E. and Knight, J. W. (1995) Ultrastructure of the intracellular melanization of Brugia malayi (Buckley) (Nematoda, Filarioidea) in the thoracic muscles of Anopheles quadrimaculatus (Say) (Diptera, Culicidae). International Journal of Insect Morphology and Embryology, 24, 83–92CrossRefGoogle Scholar
Chikilian, M. L., Bradley, T. J., Nayar, J. K. and Knight, J. W. (1994) Ultrastructural comparison of extracellular and intracellular encapsulation of Brugia malayi in Anopheles quadrimaculatus. Journal of Parasitology, 80, 133–40CrossRefGoogle ScholarPubMed
Choe, K. M., Werner, T., Stoven, S., Hultmark, D. and Anderson, K. V. (2002) Requirement for a peptidoglycan recognition protein (PGRP) in relish activation and antibacterial immune responses in Drosophila. Science, 296, 359–62CrossRefGoogle ScholarPubMed
Christensen, B. M. (1978) Dirofilaria immitis: effects on the longevity of Aedes trivittatus. Experimental Parasitology, 44, 116–23CrossRefGoogle Scholar
Christensen, B. M., Forton, K. F., Lafond, M. M. and Grieve, R. B. (1987) Surface changes on Brugia pahangi microfilariae and their association with immune evasion in Aedes aegypti. J. Invert. Pathol., 49, 14–18CrossRefGoogle ScholarPubMed
Christensen, B. M. and LaFond, M. M. (1986) Parasite induced suppression of the immune response in Aedes aegypti by Brugia pahangi. J. Parasit., 72, 216–19CrossRefGoogle ScholarPubMed
Christophides, G. K., Zdobnov, E., Barillas-Mury, C.et al. (2002) Immunity-related genes and gene families in Anopheles gambiae. Science, 298, 159–65CrossRefGoogle ScholarPubMed
Ciurea, I. and Dinulescu, G. (1924) Ravages causes par la mouchede Goloubatz en Roumanie; ses attaques contre les animaux et contre l'homme. Ann. Trop. Med. Parasit., 18, 323–42CrossRefGoogle Scholar
Clay, T. (1963) A new species of Haematomyzus Piaget (Phthiraptera, Insecta). Proc. Zool. Soc. Lond., 141, 153–61CrossRefGoogle Scholar
Clifford, C. M., Bell, J. F., Moore, G. J. and Raymond, C. (1967) Effects of limb disability of lousiness in mice. IV. Evidence of genetic factors in susceptibility to Polyplax serrata. Experimental Parasitology, 20, 56–67CrossRefGoogle ScholarPubMed
Coatney, G. R., Collins, W. E., McWilson, W. and Contacos, P. G. (1971) The Primate Malarias. Bethesda, MD: NIAID
Cockerell, T. D. A. (1918) New species of North American fossil beetles, cockroaches and tsetse flies. Proc. U. S. Natn. Mus., 54, 301–11CrossRefGoogle Scholar
Coetzee, M., Craig, M. and Sueur, D. (2000) Distribution of African malaria mosquitoes belonging to the Anopheles gambiae complex. Parasitology Today, 16, 74–7CrossRefGoogle ScholarPubMed
Colless, D. H. and Chellapah, W. T. (1960) Effects of body weight and size of blood meal upon egg production in Aedes aegypti (Linnaeus) (Diptera, Culicidae). Ann. Trop. Med. Parasit., 54, 475–82CrossRefGoogle Scholar
Collins, F. H., Sakai, R. K., Vernick, K. D.et al. (1986) Genetic selection of a refractory strain of the malaria vector Anopheles gambiae. Science, 234, 607–10CrossRefGoogle ScholarPubMed
Colman, R. W. (2001) Hemostasis and Thrombosis: Basic Principles and Clinical Practice. London: Lippincott, Williams and Wilkins
Coluzzi, M., Concetti, A. and Ascoli, F. (1982) Effect of cibarial armature of mosquitoes (Diptera: Culicidae) on blood-meal haemolysis. J. Insect Physiol., 28, 885–8CrossRefGoogle Scholar
Coluzzi, M., Sabatini, A., Torre, A., Di Deco, M. A. and Petrarca, V. (2002) A polygene chromosome analysis of the Anopheles gambiae species complex. Science, 298, 1415–18CrossRefGoogle Scholar
Colvin, J., Brady, J. and Gibson, G. (1989) Visually-guided, upwind turning behaviour of free-flying tsetse flies in odour-laden wind: a wind-tunnel study. Phys. Ent., 14, 31–9CrossRefGoogle Scholar
Colyer, C. N. and Hammond, C. O. (1968) Flies of the British Isles. London: Frederick Warne
Compton-Knox, P. and Hayes, K. L. (1972) Attraction of Tabanus spp. (Diptera: Tabanidae) to traps baited with carbon dioxide and other chemicals. Environ. Entomal., 1, 323–6CrossRefGoogle Scholar
Cook, S. P. and McCleskey, E. W. (2002) Cell damage excites nociceptors through release of cytosolic ATP. Pain, 95, 41–7CrossRefGoogle ScholarPubMed
Cornford, E. M., Freeman, B. J. and MacInnis, A. J. (1976) Physiological relationships and circadian periodicities in rodent trypanosomes. Trans. R. Soc. Trop. Med. Hyg., 70, 238–43CrossRefGoogle ScholarPubMed
Croft, S. L., East, J. S. and Molyneux, D. H. (1982) Antitrypanosomal factor in the haemolymph of Glossina. Acta Trop., 39, 293–302Google Scholar
Cross, M. L., Cupp, E. W. and Enriquez, F. J. (1994) Modulation of murine cellular immune-responses and cytokines by salivary-gland extract of the black fly Simulium vittatum. Tropical Medicine and Parasitology, 45, 119–124Google ScholarPubMed
Cupp, E. W., Cupp, M. S., Ribeiro, J. M. C. and Kunz, S. E. (1998a) Blood-feeding strategy of Haematobia irritans (Diptera: Muscidae). J. Med. Ent., 35, 591–5CrossRefGoogle Scholar
Cupp, E. W. and Stokes, G. M. (1976) Feeding patterns of Culex salinarius Coquillett in Jefferson parish, Louisiana. Mosq. News, 36, 332–5Google Scholar
Cupp, M. S., Ribeiro, J. M. C., Champagne, D. E. and Cupp, E. W. (1998b) Analyses of cDNA and recombinant protein for a potent vasoactive protein in saliva of a blood-feeding black fly, Simulium vittatum. J. Exp. Biol., 201, 1553–61Google Scholar
Dale, C., Young, S. A., Haydon, D. T. and Welburn, S. C. (2001) The insect endosymbiont Sodalis glossinidius utilizes a type III secretion system for cell invasion. Proceedings of the National Academy of Sciences of the United States of America, 98, 1883–8CrossRefGoogle ScholarPubMed
Dan, A., Pereira, M. H., Pesquero, J. L., Diotaiuti, L. and Beirao, P. S. L. (1999) Action of the saliva of Triatoma infestans (Heteroptera: Reduviidae) on sodium channels. J. Med. Entomol., 36, 875–9CrossRefGoogle ScholarPubMed
Daniel, T. L. and Kingsolver, J. G. (1983) Feeding strategy and the mechanics of blood sucking in insects. J. Theor. Biol., 105, 661–72CrossRefGoogle ScholarPubMed
David, C. T., Kennedy, J. S., Ludlow, A. R., Perry, J. N. and Wall, C. (1982) A reappraisal of insect flight towards a distant point source of wind-borne odor. J. Chem. Ecol., 8, 1207–15CrossRefGoogle ScholarPubMed
Davidson, G. and Draper, C. C. (1953) Field studies of some of the basic factors concerned in the transmission of malaria. Trans. R. Soc. Trop. Med. Hyg., 47, 522–35CrossRefGoogle ScholarPubMed
Davis, E. E. (1984) Regulation of sensitivity in the peripheral chemoreceptor systems for host-seeking behaviour by a haemolymph-borne factor in Aedes aegypti. J. Insect Physiol., 30, 179–83CrossRefGoogle Scholar
Davis, E. E. and Sokolove, P. G. (1975) Temperature response of the antennal receptors in the mosquito, Aedes aegypti. J. Comp. Physiol., 96, 223–36CrossRefGoogle Scholar
Day, J. F. and Edman, J. D. (1983) Malaria renders mice susceptible to mosquito feeding when gametocytes are most infective. J. Parasitol., 69, 163–70CrossRefGoogle ScholarPubMed
Day, J. F. and Edman, J. D. (1984a) The importance of disease induced changes in mammalian body temperature to mosquito blood feeding. Comp. Biochem. Physiol., 77, 447–52CrossRefGoogle Scholar
Day, J. F. and Edman, J. D. (1984b) Mosquito engorgement on normally defensive hosts depends on host activity patterns. J. Med. Ent., 21, 732–40CrossRefGoogle Scholar
Azambuja, P., Guimares, J. A. and Garcia, E. S. (1983) Haemolytic factor from the crop of Rhodnius prolixus: evidence and partial characterisation. J. Insect Physiol., 29, 833–7CrossRefGoogle Scholar
Gregorio, E., Spellman, P. T., Tzou, P., Rubin, G. M. and Lemaitre, B. (2002) The Toll and Imd pathways are the major regulators of the immune response in Drosophila. EMBOJ., 21, 2568–79CrossRefGoogle ScholarPubMed
Jong, R. and Knols, B. G. J. (1995) Selection of biting sites on man by 2 malaria mosquito species. Experientia, 51, 80–4CrossRefGoogle Scholar
DeFoliart, G. R., Grimstad, P. R. and Watts, D. M. (1987) Advances in mosquito-borne arbovirus/vector research. Ann. Rev. Ent., 32, 479–505CrossRefGoogle ScholarPubMed
Dei Cas, E., Maurois, P., Landau, I.et al. (1980) Morphologie et infectivite des gametocytes de Plasmodium inui. Annales Parasit., 55, 621–33Google Scholar
Dekker, T. and Takken, W. (1998) Differential responses of mosquito sibling species Anopheles arabiensis and An. quadriannulatus to carbon dioxide, a man or a calf. Med. Vet. Entomol., 12, 136–40CrossRefGoogle ScholarPubMed
Dekker, T., Takken, W. and Braks, M. A. H. (2001) Innate preference for host-odor blends modulates degree of anthropophagy of Anopheles gambiae sensu lato (Diptera: Culicidae). J. Med. Ent., 38, 868–71CrossRefGoogle Scholar
Dekker, T., Takken, W., Knols, B. G. J.et al. (1998) Selection of biting sites on a human host by Anopheles gambiae s. s., An. arabiensis and An. quadriannulatus. Entomologia Experimentalis et Applicata, 87, 295–300CrossRefGoogle Scholar
Denotter, C. J., Tchicaya, T. and Schutte, A. M. (1991) Effects of age, sex and hunger on the antennal olfactory sensitivity of tsetse-flies. Phys. Ent., 16, 173–82Google Scholar
Desquesnes, M. and Dia, M. L. (2003) Trypanosoma vivax: mechanical transmission in cattle by one of the most common African tabanids, Atylotus agrestis. Experimental Parasitology, 103, 35–43CrossRefGoogle ScholarPubMed
Dethier, V. G. (1954) Notes on the biting response of tsetse flies. Am. J. Trop. Med. Hyg., 3, 160–71CrossRefGoogle ScholarPubMed
Detinova, T. S. (1962) Age Grouping Methods in Diptera of Medical Importance. Geneva: World Health Organization
Dias, J. C., Silveira, A. C. and Schofield, C. J. (2002) The impact of Chagas disease control in Latin America: a review. Mem. Inst. Oswaldo Cruz, 97, 603–12CrossRefGoogle ScholarPubMed
Diatta, M., Spiegel, A., Lochouarn, L. and Fontenille, D. (1998) Similar feeding preferences of Anopheles gambiae and An. arabiensis in Senegal. Trans. R. Soc. Trop. Med. Hyg., 92, 270–2CrossRefGoogle ScholarPubMed
Dickerson, G. and Lavoipierre, M. M. J. (1959) Studies on the methods of feeding blood-sucking arthropods. II. The method of feeding adopted by the bed-bug (Cimex lectularius) when obtaining a blood meal from the mammalian host. Ann. Trop. Med. Parasit., 53, 347–57CrossRefGoogle Scholar
Dimopoulos, G. (2003) Insect immunity and its implication in mosquito-malaria interactions. Cell Microbiol., 5, 3–14CrossRefGoogle ScholarPubMed
Dimopoulos, G., Christophides, G. K., Meister, S.et al. (2002) Genome expression analysis of Anopheles gambiae: responses to injury, bacterial challenge, and malaria infection. Proceedings of the National Academy of Sciences of the United States of America, 99, 8814–19CrossRefGoogle ScholarPubMed
Dimopoulos, G., Seeley, D., Wolf, A. and Kafatos, F. C. (1998) Malaria infection of the mosquito Anopheles gambiae activates immune-responsive genes during critical transition stages of the parasite life cycle. EMBO J., 17, 6115–23CrossRefGoogle ScholarPubMed
Dobson, A. P. (1988) Parasite-induced changes in host behaviour, Q. Rev. Biol., 63(4), 140–65
Downes, J. A. (1970) The ecology of blood-sucking diptera: an evolutionary perspective. In Fallis, A. M. (ed.), and Physiology of Parasites. Toronto: University of Toronto PressGoogle Scholar
Downs, C. M., Theberge, J. B. and Smith, S. M. (1986) The influence of insects on the distribution, microhabitat choice, and behaviour of the Burwash caribou herd. Can. J. Zool., 64, 622–9CrossRefGoogle Scholar
Dryden, M. W. (1989) Host association, on-host longevity and egg production of Ctenocephalides felis felis. Vet Parasitol., 34, 117–22CrossRefGoogle ScholarPubMed
Dujardin, J. C., Banuls, A. L., Llanos-Cuentas, A.et al. (1995) Putative Leishmania hybrids in the Eastern Andean valley of Huanuco, Peru. Acta Trop., 59, 293–307CrossRefGoogle ScholarPubMed
Duncan, P. and Vigne, N. (1979) The effect of group size in horses on the rate of attacks by blood-sucking flies. Animal Behaviour, 27, 623–5CrossRefGoogle Scholar
Dunnet, G. M. (1970) Siphonaptera (Fleas). In CSIRO (ed.) The Insects of Australia. Melbourne: Melbourne University Press
Durvasula, R. V., Gumbs, A., Panackal, A.et al. (1997) Prevention of insect-borne disease: an approach using transgenic symbiotic bacteria. Proceedings of the National Academy of Sciences of the United States of America, 94, 3274–8CrossRefGoogle ScholarPubMed
Duval, J., Rajaonarivelo, E. and Rabenirainy, L. (1974) Ecologie de Styloconops spinosifrons (Carter, 1921) (Diptera, Ceratopogonidae) sur les plages de la côte Est de Madagascar. Cahiers ORSTOM, 12, 245–58Google Scholar
Dye, C. (1992) The analysis of parasite transmission by bloodsucking insects. Ann. Rev. Ent., 37, 1–19CrossRefGoogle ScholarPubMed
East, J., Molyneux, D. H., Maudlin, I. and Dukes, P. (1983) Effect of Glossina haemolymph on salivarian trypanosomes in vitro. Annals of Tropical Medicine and Parasitology, 77, 97–9CrossRefGoogle ScholarPubMed
Edman, J. D. (1974) Host-feeding patterns of Florida mosquitoes III Culex (Culex) and Culex (NeoCulex). J. Med. Ent., 11, 95–104CrossRefGoogle Scholar
Edman, J. D., Day, J. F. and Walker, E. (1985) Vector-host interplay: factors affecting disease transmission. In Lounibos, L. P., Rey, R. and Frank, J. H. (eds.), of Mosquitoes: Proceedings of a Workshop. New Beach, Florida: Florida Medical Entomology LaboratoryGoogle Scholar
Edman, J. D. and Kale, H. W. (1971) Host behaviour: its influence on the feeding success of mosquitoes. Ann. Ent. Soc. Am., 64, 513–16CrossRefGoogle Scholar
Edman, J. D. and Spielman, A. (1988) Blood feeding by vectors: physiology, ecology, behaviour and vertebrate defence. In Monath, T. P. (ed.), The Arboviruses: Epidemiology and Ecology. Baton Rouge: C.R.C. Press, Vol. 1Google Scholar
Edman, J. D. and Taylor, D. J. (1968) Culex nigripalpus: seasonal shift in the bird-mammal feeding ratio in a mosquito vector of human encephalitis. Science, 161, 67–8CrossRefGoogle Scholar
Edman, J. D., Webber, L. A. and Kale, H. W. (1972) Effect of mosquito density on the interrelationship of host behavior and mosquito feeding success. Am. J. Trop. Med. Hyg., 21, 487–91CrossRefGoogle ScholarPubMed
Eichler, D. A. (1973) Studies on Onchocerca gutterosa (Neumann, 1910) and its development in Simulium ornatum (Meigen, 1818). 3. Factors affecting the development of the parasite in its vector. J. Helm., 47, 73–88CrossRefGoogle Scholar
Eiras, A. E. and Jepson, P. C. (1994) Responses of female Aedes aegypti (Diptera, Culicidae) to host odors and convection currents using an olfactometer bioassay. Bull. Ent. Res., 84, 207–11CrossRefGoogle Scholar
Elkinton, J. S., Schal, C., Ono, T. and Carde, R. T. (1987) Pheromone puff trajectory and upwind flight of male gypsy moths in a forest. Phys. Ent., 12, 399–406CrossRefGoogle Scholar
Ellis, D. S. and Evans, D. A. (1977) Passage of Trypanosoma brucei rhodesiense through the peritrophic membrane of Glossina morsitans morsitans. Nature, 267, 834–5CrossRefGoogle ScholarPubMed
Elrod-Erickson, M., Mishra, S. and Schneider, D. (2000) Interactions between the cellular and humoral immune responses in Drosophila. Current Biology, 10, 781–4CrossRefGoogle ScholarPubMed
Elsen, P., Amoudi, M. A. and Leclercq, M. (1990) 1st record of Glossina fuscipes fuscipes Newstead, 1910 and Glossina morsitans submorsitans Newstead, 1910 in Southwestern Saudi-Arabia. Annales De La Societe Belge De Medecine Tropicale, 70, 281–7Google Scholar
Emmerson, K. C., Kim, K. C. and Price, R. D. (1973) Lice. In Flynn, R. J (ed.), Parasites of Laboratory Animals. Ames, Iowa: Iowa State University PressGoogle Scholar
Escalante, A. A. and Ayala, F. J. (1995) Evolutionary origin of Plasmodium and other apicomplexa based on ribosomal-RNAgenes. Proceedings of the National Academy of Sciences of the United States of America, 92, 5793–7CrossRefGoogle Scholar
Esseghir, S., Ready, P. D., KillickKendrick, R. and BenIsmail, R. (1997) Mitochondrial haplotypes and phylogeography of Phlebotomus vectors of Leishmania major. Insect Mol. Biol., 6, 211–25CrossRefGoogle ScholarPubMed
Evans, G. O. (1950) Studies on the bionomics of the sheep ked, Melophagus ovinus L., in West Wales. Bull. Ent. Res., 40, 459–78CrossRefGoogle Scholar
Ewert, A. (1965) Comparative migration of microfilariae and development of Brugia pahangi in various mosquitoes. Am. J. Trop. Med. Hyg., 14, 254–9CrossRefGoogle ScholarPubMed
Falleroni, D. (1927) Per la soluzione del problema malarico italiano. Riv. Malariol., 6, 344–409Google Scholar
Fallis, A. M., Bennett, G. F., Griggs, G. and Allen, T. (1967) Collecting Simulium venustum female in fan traps and on silhouettes with the aid of carbon dioxide. Can. J. Zool., 45, 1011–17CrossRefGoogle ScholarPubMed
Fallis, A. M. and Raybould, J. N. (1975) Response of two African simuliids to silhouettes and carbon dioxide. J. Med. Entomol., 12, 349–51CrossRefGoogle ScholarPubMed
Fallis, A. M. and Smith, S. M. (1964) Ether extract from birds and CO2 as attractants for some ornithophilic simuliids. Can. J. Zool., 42, 723–30CrossRefGoogle Scholar
Farkas, S. R. and Shorey, H. H. (1972) Chemical trail-following by flying insects: a mechanism for orientation to a distant odor source. Science (Washington, D.C), 178, 67–8CrossRefGoogle ScholarPubMed
Farmer, J., Maddrell, S. H. P. and Spring, J. H. (1981) Absorption of fluid by the midgut of Rhodnius. J. Exp. Biol., 94, 301–16Google Scholar
Faust, E. C., Russel, P. F. and Jung, R. C. (1977) Craig and Fausts Clinical Parasitology. Philadelphia: Lea and Febiger
Favia, G., Torre, A., Bagayoko, M.et al. (1997) Molecular identification of sympatric chromosomal forms of Anopheles gambiae and further evidence of their reproductive isolation. Insect Mol. Biol., 6, 377–83CrossRefGoogle ScholarPubMed
Feingold, B. F. and Benjamini, E. (1961) Allergy to flea bites: clinical and experimental observations. Ann. Allerg., 19, 1274–89Google ScholarPubMed
Ferdig, M. T., Beerntsen, B. T., Spray, F. J., Li, J. and Christensen, B. M. (1993) Reproductive costs associated with resistance in a mosquito-filarial worm system. Am. J. Trop. Med. Hyg., 49, 756–62CrossRefGoogle Scholar
Ferguson, H. M. and Read, A. F. (2002a) Genetic and environmental determinants of malaria parasite virulence in mosquitoes. Proc. R. Soc. Lond. B Biol. Sciences, 269, 1217–24CrossRefGoogle Scholar
Ferguson, H. M. and Read, A. F. (2002b) Why is the effect of malaria parasites on mosquito survival still unresolved?Trends in Parasitology, 18, 256–61CrossRefGoogle Scholar
Ferrari, J., Muller, C. B., Kraaijeveld, A. R. and Godfray, H. C. (2001) Clonal variation and covariation in aphid resistance to parasitoids and a pathogen. Evolution, 55, 1805–1814CrossRefGoogle Scholar
Ferris, G. R. (1931) The louse of elephants Haematomyzus elephantis Piaget (Mallophaga: Haematomyzidae). Parasitol., 23, 112–27CrossRefGoogle Scholar
Flores, G. B. and Lazzari, C. R. (1996) The role of the antennae in Triatoma infestans: Orientation towards thermal sources. J. Insect Physiol., 42, 433–40CrossRefGoogle Scholar
Foley, E. and Farrell, P. H. (2003) Nitric oxide contributes to induction of innate immune responses to gram-negative bacteria in Drosophila. Genes Dev., 17, 115–25CrossRefGoogle ScholarPubMed
Foster, W. A. (1976) Male sexual maturation of the tsetse flies Glossina morsitans Westwood and G. austeni Newstead (Diptera: Glossinidae) in relation to feeding. Bull. Ent. Res., 66, 389–99CrossRefGoogle Scholar
Fox, A. N., Pitts, R. J., Robertson, H. M., Carlson, J. R. and Zwiebel, L. J. (2001) Candidate odorant receptors from the malaria vector mosquito Anopheles gambiae and evidence of down-regulation in response to blood feeding. Proceedings of the National Academy of Sciences of the United States of America, 98, 14693–7CrossRefGoogle ScholarPubMed
Francischetti, I. M. B., Ribeiro, J. M. C., Champagne, D. and Andersen, J. (2000) Purification, cloning, expression, and mechanism of action of a novel platelet aggregation inhibitor from the salivary gland of the blood-sucking bug, Rhodnius prolixus. J. Biol. Chem., 275, 12639–50CrossRefGoogle ScholarPubMed
Francischetti, I. M. B., Valenzuela, J. G. and Ribeiro, J. M. C. (1999) Anophelin: kinetics and mechanism of thrombin inhibition. Biochemistry, 38, 16678–85CrossRefGoogle ScholarPubMed
Fredeen, F. J. H. (1961) A trap for studying the attacking behaviour of black flies Simulium articum Mall. Can. Entomol., 93, 73–8CrossRefGoogle Scholar
Freier, J. E. and Friedman, S. (1976) Effect of host infection with Plasmodium gallinaceum on the reproductive capacity of Aedes aegypti. J. Invert. Pathol., 28, 161–6CrossRefGoogle ScholarPubMed
Friend, W. G. (1978) Physical factors affecting the feeding responses of Culiseta inornata to ATP, sucrose and blood. Ann. Ent. Soc. Am., 71, 935–40CrossRefGoogle Scholar
Friend, W. G. and Smith, J. J. B. (1971) Feeding in Rhodnius prolixus: mouthpart activity and salivation and their correlation with changes of electrical resistance. J. Insect Physiol., 17, 233–43CrossRefGoogle Scholar
Friend, W. G. and Smith, J. J. B. (1975) Feeding in Rhodnius prolixus: increasing sensitivity to ATP during prolonged food deprivation. J. Insect Physiol., 21, 1081–4CrossRefGoogle ScholarPubMed
Friend, W. G. and Smith, J. J. B. (1977) Factors affecting feeding by blood-sucking insects. Ann. Rev. Ent., 22, 309–31CrossRefGoogle Scholar
Friend, W. G. and Stoffolano, J. G. (1984) Feeding responses of the horsefly, Tabanus nigrovittatus, to physical factors, ATP analogues and blood fractions. Phys. Ent., 9, 395–402CrossRefGoogle Scholar
Fu, H., Leake, C. J., Mertens, P. P. and Mellor, P. S. (1999) The barriers to bluetongue virus infection, dissemination and transmission in the vector, Culicoides variipennis (Diptera: Ceratopogonidae). Arch. Virol., 144, 747–61CrossRefGoogle Scholar
Gad, A. M., Maier, W. A. and Piekorski, G. (1979) Pathology of Anopheles stephensi after infection with Plasmodium berghei berghei. I. Mortality rate. Z. Parasit., 60, 249–61CrossRefGoogle ScholarPubMed
Gade, G. and Auerswald, L. (2002) Beetles' choice: proline for energy output: control by AKHs. Comp. Biochem. Physiol. B., 132, 117–29CrossRefGoogle ScholarPubMed
Galun, R. (1966) Feeding stimulants of the rat flea Xenopsylla cheopis Roth. Life Sci., 5, 1335–42CrossRefGoogle Scholar
Galun, R. (1986) Diversity of phagostimulants used for recognition of blood meal by haematophagous arthropods. In Borovsky, D. and Spielman, A. (eds.), Host-Regulated Development Mechanisms in Vector Arthropods. Florida: IFAS, University of FloridaGoogle Scholar
Galun, R. (1987) The evolution of purinergic receptors involved in recognition of a blood meal by haematophagous insects. Mem. Inst. Oswaldo Cruz, 82, 5–9CrossRefGoogle Scholar
Galun, R., Avidor, Y. and Bar-Zeev, M. (1963) Feeding response in Aedes aegypti: stimulation by adnosine triphosphate. Science, 124, 1674–5CrossRefGoogle Scholar
Galun, R., Friend, W. G. and Nudelman, S. (1988) Purinergic reception by culicine mosquitoes. J. Comp. Physiol. A., 163, 665–70CrossRefGoogle ScholarPubMed
Galun, R. and Kabayo, J. P. (1988) Gorging response of Glossina palpalis palpalis to ATP analogues. Phys. Ent., 13, 419–23CrossRefGoogle Scholar
Galun, R., Koontz, L. C. and Gwadz, R. W. (1985) Engorgement response of anopheline mosquitoes to blood fractions and artificial solutions. Phys. Ent., 10, 145–9CrossRefGoogle Scholar
Galun, R., Vardimonfriedman, H. and Frankenburg, S. (1993) Gorging response of culicine mosquitos (Diptera, Culicidae) to blood fractions. J. Med. Ent., 30, 513–17CrossRefGoogle Scholar
Garcia, R. and Radovsky, F. J. (1962) Haematophagy by two non-biting myscid flies and its relationship to tabanid feeding. Can. Entomol., 94, 1110–16CrossRefGoogle Scholar
Gardiner, E. M. and Strand, M. R. (1999) Monoclonal antibodies bind distinct classes of hemocytes in the moth Pseudoplusia includens. J Insect Physiol., 45, 113–26CrossRefGoogle ScholarPubMed
Garms, R., Walsh, J. F. and Davies, J. B. (1979) Studies on the reinvasion of the Onchocerciasis Control Programme in the Volta River Basin by Simulium damnosum s.l. with emphasis on the south-western areas. Tropenmed. Parasitol., 30, 345–62Google Scholar
Garrett-Jones, C. and Shidrawi, G. R. (1969) Malaria vectorial capacity of a population of Anopheles gambiae. Bulletin of the World Health Organization, 40, 531–45Google ScholarPubMed
Gaston, K. A. and Randolph, S. E. (1993) Reproductive under-performance of tsetse-flies in the laboratory, related to feeding frequency. Phys. Ent., 18, 130–6CrossRefGoogle Scholar
Gatehouse, A. G. (1970) The probing response of Stomoxys calcitrans to certain physical and olfactory stimuli. J. Insect Physiol., 16, 61–74CrossRefGoogle ScholarPubMed
Gatehouse, A. G. (1972) Some responses of tsetse flies to visual and olfactory stimuli. Nature New Biol., 236, 63–4CrossRefGoogle ScholarPubMed
Gaunt, M. W. and Miles, M. A. (2002) An insect molecular clock dates the origin of the insects and accords with palaeontological and biogeographic landmarks. Molecular Biology and Evolution, 19, 748–61CrossRefGoogle ScholarPubMed
Gaunt, M. W., Yeo, M., Frame, I. A. (2003) Mechanism of genetic exchange in American trypanosomes. Nature, 421, 936–9CrossRefGoogle ScholarPubMed
Gautret, P. (2001) Plasmodium falciparum gametocyte periodicity. Acta Trop., 78, 1–2CrossRefGoogle ScholarPubMed
Gautret, P. and Motard, A. (1999) Periodic infectivity of Plasmodium gametocytes to the vector: a review. Parasite-Journal De La Societe Francaise De Parasitologie, 6, 103–11Google ScholarPubMed
Geden, C. J. and Hogsette, J. A. (1994) Research and extension needs for integrated pest management for arthropods of veterinary importance. Proceedings of a Workshop in Lincoln, Nebraska, 12–14 April, Lincoln, NebraskaGoogle Scholar
Gee, J. C. (1975) Diuresis in the tsetse fly Glossina austeni. J. Exp. Biol., 63, 381–90Google ScholarPubMed
Geier, M., Bosch, O. J. and Boeckh, J. (1999) Ammonia as an attractive component of host odour for the yellow fever mosquito, Aedes aegypti. Chemical Senses, 24, 647–53CrossRefGoogle ScholarPubMed
Gentile, G., Della Torre, A., Maegga, B., Powell, J. R. and Caccone, A. (2002) Genetic differentiation in the African malaria vector, Anopheles gambiae s.s., and the problem of taxonomic status. Genetics, 161, 1561–78Google ScholarPubMed
Ghosh, K. N. and Mukhopadhyay, J. (1998) The effect of anti-sandfly saliva antibodies on Phlebotomus argentipes and Leishmania donovani. Int. J. Parasit., 28, 275–81CrossRefGoogle ScholarPubMed
Gibson, G. (1992) Do tsetse-flies see zebras: a field-study of the visual response of tsetse to striped targets. Phys. Ent., 17, 141–7CrossRefGoogle Scholar
Gibson, G. and Brady, J. (1985) Anemotactic flight paths of tsetse flies in relation to host odour: preliminary video study in nature. Phys. Ent., 10, 395–406CrossRefGoogle Scholar
Gibson, G. and Brady, J. (1988) Flight behaviour of tsetse flies in host odour plumes: the initial response to leaving or entering odour. Phys. Ent., 13, 29–42CrossRefGoogle Scholar
Gibson, G. and Torr, S. J. (1999) Visual and olfactory responses of haematophagous Diptera to host stimuli. Med. Vet. Entomol., 13, 2–23CrossRefGoogle ScholarPubMed
Gibson, G. and Young, S. (1991) The optics of tsetse-fly eyes in relation to their behavior and ecology. Phys. Ent., 16, 273–82CrossRefGoogle Scholar
Gikonyo, N. K., Hassanali, A., Njagi, P. G. N. and Saini, R. K. (2000) Behaviour of Glossina morsitans morsitans Westwood (Diptera: Glossinidae) on waterbuck Kobus defassa Ruppel and feeding membranes smeared with waterbuck sebum indicates the presence of allomones. Acta Trop., 77, 295–303CrossRefGoogle ScholarPubMed
Gillespie, R. D., Mbow, M. L. and Titus, R. G. (2000) The immunomodulatory factors of bloodfeeding arthropod saliva. Parasite Immunology, 22, 319–31CrossRefGoogle ScholarPubMed
Gillett, J. D. (1967) Natural selection and feeding speed in a blood sucking insect. Proc. R. Soc. London Ser. B., 167, 316–29CrossRefGoogle Scholar
Gillett, J. D. and Connor, J. (1976) Host temperature and the transmission of arboviruses by mosquitoes. Mosq. News, 36, 472–7Google Scholar
Gillies, M. T. (1980) The role of carbon dioxide in host-finding by mosquitoes (Diptera: Culicidae): a review. Bull. Ent. Res., 70, 525–32CrossRefGoogle Scholar
Gillies, M. T. and Wilkes, T. J. (1969) A comparison of the range of attraction of animal baits and carbon dioxide for some West African mosquitoes. Bull. Ent. Res., 59, 441–56CrossRefGoogle ScholarPubMed
Gillies, M. T. and Wilkes, T. J. (1970) The range of attraction of single baits for some West African mosquitoes. Bull. Ent. Res., 60, 225–35CrossRefGoogle ScholarPubMed
Gillies, M. T. and Wilkes, T. J. (1972) The range of attraction of animal baits and carbon dioxide for mosquitoes. Studies in a freshwater area of West Africa. Bull. Ent. Res., 61, 389–404CrossRefGoogle Scholar
Glasgow, J. P. (1961) The feeding habits of Glossina swynnertoni. J. An. Ecol., 30, 77–85CrossRefGoogle Scholar
Goodchild, A. J. P. (1955) Some observations on growth and egg production of the blood-sucking Reduviids, Rhodnius prolixus and Triatoma infestans. Proc. R. Ent. Soc., 30, 137–44Google Scholar
Gooding, R. H. (1968) A note on the relationship between feeding and insemination in Pediculus humanus. J. Med. Ent., 5, 265–6CrossRefGoogle ScholarPubMed
Gooding, R. H. (1972) Digestive processes of haematophagous insects. I. A literature review. Quaest. Ent., 8, 5–60Google Scholar
Gooding, R. H.(1974) Digestive processes in haematophagous insects. Control of trypsin secretion in Glossina morsitans morsitans. J. Insect Physiol., 20, 957–64CrossRefGoogle Scholar
Gooding, R. H. (1975) Inhibition of diuresis in the tsetse fly (Glossina morsitans) by ouabain or acetazolamide. Experientia, 31, 938–9CrossRefGoogle ScholarPubMed
Gooding, R. H. (1977) Digestive processes of haematophagous insects. XIV Haemolytic activity in the midgut of Glossina morsitans morsitans Westwood (Diptera: Glossinidae). Can. J. Zool., 55, 1899–1905CrossRefGoogle Scholar
Gordon, R. M., Crewe, W. and Willett, K. C. (1956) Studies on the deposition, migration and development to the blood forms of tryapnosomes belonging to the Trypanosoma brucei group. I. An account of the process of feeding adopted by the tsetse fly when obtaining a blood meal from the mammalian host, with special reference to the ejection of saliva and the relationship of the feeding process to the deposition of the metacyclic trypanosomes. Ann. Trop. Med., 50, 426–37CrossRefGoogle Scholar
Gore, T. C. and Pittman-Noblet, G. (1978) The effect of photoperiod on the deep body temperature of domestic turkeys and its relationship to the diurnal periodicity of Leucocytozoon smithi gametocytes in the peripheral blood of turkeys. Poultry Science, 57, 603–7CrossRefGoogle ScholarPubMed
Gorman, M. J., Cornel, A. J., Collins, F. H. and Paskewitz, S. M. (1996) A shared genetic mechanism for melanotic encapsulation of CM- Sephadex beads and a malaria parasite, Plasmodium cynomolgi B, in the mosquito, Anopheles gambiae. Experimental Parasitology, 84, 380–6CrossRefGoogle Scholar
Gottar, M., Gobert, V., Michel, T.et al. (2002) The Drosophila immune response against Gram-negative bacteria is mediated by a peptidoglycan recognition protein. Nature, 416, 640–4CrossRefGoogle ScholarPubMed
Gotz, P. (1986) Encapsulation in arthropods. In Brehelin, M. (ed.), Immunity in Invertebrates. Springer Verlag, Berlin Heidelberg, New York, Tokyo, 153–70CrossRefGoogle Scholar
Graf, R., Raikhel, A. S., Brown, M. R., Lea, A. O. and Briegel, H. (1986) Mosquito trypsin: immunocytochemical localisation in the midgut of blood-fed Aedes aegypti (L.). Cell, 245, 19–27Google Scholar
Graham, H. (1902) Dengue: a study of its mode of propagation and pathology. Medical Record, 61, 204–7
Grant, A. J., Wigton, B. E., Aghajanian, J. G. and O' Connell, R. J. (1995) Electrophysiological responses of receptor neurons in mosquito maxillary palp sensilla to carbon-dioxide. J. Comp. Physiol., 177, 389–96CrossRefGoogle ScholarPubMed
Grassi, B., Bignami, A. E. and Bastianelli, G. (1899) Ciclo evolutivo delle semilune nell' Anopheles claviger ed altri studi sulla malaria dall' ottobre 1898 all maggio 1899. Atti. Soc. Studi Malaria, 1, 143–27
Green, C. H. (1986) effects of colors and synthetic odors on the attraction of Glossina pallidipes and Glossina morsitans morsitans to traps and screens. Phys. Ent., 11, 411–21CrossRefGoogle Scholar
Green, C. H. (1989) The use of two-coloured screens for catching Glossina palpalis palpalis (Robineau-Desvoidy) (Diptera: Glossinidae). Bull. Ent. Res., 79, 81–93CrossRefGoogle Scholar
Green, C. H. and Cosens, D. (1983) Spectral responses of the tsetse fly, Glossina morsitans morsitans. J. Insect Physiol., 29, 795–800CrossRefGoogle Scholar
Griffiths, N. and Brady, J. (1995) Wind structure in relation to odour plumes in tsetse fly habitats. Phys. Ent., 20, 286–92CrossRefGoogle Scholar
Griffiths, N., Paynter, Q. and Brady, J. (1995) Rates of progress up odour trails by tsetse flies: a mark-release video study of the timing of odour source location by Glossina pallidipes. Phys. Ent., 20, 100–8CrossRefGoogle Scholar
Grimstad, P. R., Paulson, S. L. and Craig, G. B. Jr (1985) Vector competence of Aedes hendersoni (Diptera: Culicidae) for La Crosse virus and evidence of a salivary gland escape barrier. J. Med. Ent., 22, 447–53CrossRefGoogle ScholarPubMed
Grimstad, P. R., Ross, Q. E. and Craig, G. B. Jr (1980) Aedes triseriatus (Diptera: Culicidae) and La Crosse virus. II. Modification of mosquito feeding behaviour by virus infection. J. Med. Ent., 17, 1–7CrossRefGoogle ScholarPubMed
Grossman, G. L. and Pappas, L. G. (1991) Human skin temperature and mosquito (Diptera, Culicidae) blood feeding rate. J. Med. Ent., 28, 456–60CrossRefGoogle ScholarPubMed
Grubhoffer, L., Hypsa, V. and Volf, P. (1997) Lectins (hemagglutinins) in the gut of the important disease vectors. Parasite-Journal De La Societe Francaise De Parasitologie, 4, 203–16Google ScholarPubMed
Guarneri, A. A., Diotaiuti, L., Gontijo, N. F., Gontijo, A. F. and Pereira, M. H. (2000) Comparison of feeding behaviour of Triatoma infestans, Triatoma brasiliensis and Triatoma pseudomaculata in different hosts by electronic monitoring of the cibarial pump. J. Insect Physiol., 46, 1121–7CrossRefGoogle ScholarPubMed
Guerenstein, P. G. and Nunez, J. A. (1994) Feeding response of the hematophagous bugs Rhodnius prolixus and Triatoma infestans to saline solutions: a comparative-study. J. Insect Physiol., 40, 747–52CrossRefGoogle Scholar
Gwadz, R. W. (1969) Regulation of blood meal size in the mosquito. J. Insect Physiol., 15, 2039–44CrossRefGoogle ScholarPubMed
Haarlov, N. (1964) Life cycle and distribution pattern of Lipoptena cervi (Dipt., Hippobosc.) on Danish deer. Oikos, 15, 93–129CrossRefGoogle Scholar
Hacker, C. S. and Kilama, W. L. (1974) The relationship between Plasmodium gallinaceum density and the fecundity of Aedes aegypti. J. Invert. Pathol., 23, 101–5CrossRefGoogle ScholarPubMed
Hackett, L. W. and Missiroli, A. (1931) The natural disappearence of malaria in certain regions of Europe. Am. J. Hyg., 13, 57–78Google Scholar
Hafner, M. S., Sudman, P. D., Villablanca, F. X.et al. (1994) Disparate rates of molecular evolution in cospeciating hosts and parasites. Science, 265, 1087–90CrossRefGoogle ScholarPubMed
Hailman, J. P. (1979) Environmental light and conspicuous colours. In Burtt, E. H. (ed.), The Behavioural Significance of Colour. New York: Garland STPM PressGoogle Scholar
Hall, D. R., Beevor, P. S., Cork, A., Nesbitt, B. F. and Vale, G. A. (1984) 1-Octen-3-ol: a potent olfactory stimulant and attractant for tsetse isolated from cattle odours. Insect Sci. Applic., 5, 335–9Google Scholar
Hall, L. R. and Titus, R. G. (1995) Sand fly vector saliva selectively modulates macrophage functions that inhibit killing of Leishmania major and nitric oxide production. Journal of Immunology, 155, 3501–6Google ScholarPubMed
Halstead, S. B. (1990) Dengue. In Warren, K. S. and Mahmoud, A. A. F. (eds.), Tropical and Geographical Medicine.New York: McGraw Hill, 675–85Google Scholar
Handman, E. (2000) Cell biology of Leishmania. In Advances in Parasitology, Vol. 44, 1–39Google Scholar
Hansens, E. J., Bosler, E. M. and Robinson, J. W. (1971) Use of traps for study and control of salt marsh flies. J. Econ. Entomol., 64, 1481–6CrossRefGoogle Scholar
Hao, Z., Kasumba, I., Lehane, M. J.et al. (2001) Tsetse immune responses and trypanosome transmission: implications for the development of tsetse-based strategies to reduce trypanosomiasis. Proc. Natl. Acad. Sci. USA, 98, 12648–53CrossRefGoogle ScholarPubMed
Haque, A. and Capron, A. (1982) Transplacental transfer of rodent microfilariae induces antigen-specific tolerance in rats. Nature, 299, 361–3CrossRefGoogle ScholarPubMed
Hargrove, J. W. (1980a) The effect of ambient-temperature on the flight performance of the mature male tsetse-fly, Glossina morsitans. Phys. Ent., 5, 397–400CrossRefGoogle Scholar
Hargrove, J. W. (1980b) The importance of model size and ox odour on the alighting response of Glossina morsitans Westwood and Glossina pallidipes Austen (Diptera: Glossinidae). Bull. Ent. Res., 70, 229–34CrossRefGoogle Scholar
Hargrove, J. W., Holloway, M. T. P., Vale, G. A., Gough, A. J. E. and Hall, D. R. (1995) Catches of tsetse (Glossina spp) (Diptera, Glossinidae) from traps and targets baited with large doses of natural and synthetic host odor. Bull. Ent. Res., 85, 215–27CrossRefGoogle Scholar
Hargrove, J. W. and Williams, B. G. (1995) A cost-benefit-analysis of feeding in female tsetse. Med. Vet. Ent., 9, 109–19CrossRefGoogle ScholarPubMed
Harrington, L. C., Edman, J. D. and Scott, T. W. (2001) Why do female Aedes aegypti (Diptera: Culicidae) feed preferentially and frequently on human blood?J. Med. Ent., 38, 411–22CrossRefGoogle ScholarPubMed
Harris, J. A., Hillerton, J. E. and Morant, S. V. (1987) Effect on milk production of controlling muscid flies, and reducing fly-avoidance behaviour by the use of Fenvalerate ear tags during the dry period. Journal of Dairy Research, 54, 165–71CrossRefGoogle ScholarPubMed
Harris, P., Riordan, D. F. and Cooke, D. (1969) Mosquitoes feeding on insect larvae. Science, 164, 184–5CrossRefGoogle ScholarPubMed
Harrison, G. (1978) Mosquitoes, Malaria and Man: A History of the Hostilities since 1880. London: Murray
Hart, B. L. and Hart, L. A. (1994 ) Fly switching by Asian elephants: tool use to control parasites. Animal Behaviour, 48, 35–45CrossRefGoogle Scholar
Hawking, F. (1962) Microfilaria infestation as an instance of periodic phenomena seen in host-parasite relationships. Ann. N. Y. Acad. Sci., 98, 940–53CrossRefGoogle ScholarPubMed
Hawking, F. (1976) Circadian rhythms in Trypanosoma congolense. Trans. R. Soc. Trop. Med. Hyg., 70, 170CrossRefGoogle ScholarPubMed
Hawking, F., Gammage, K. and Worms, M. J. (1972) The asexual and sexual circadian rhythms of Plasmodium vinckei chabaudi, P. berghei and P. gallinaceum. Parasitology, 65, 189–201CrossRefGoogle ScholarPubMed
Hawking, F., Wilson, M. E. and Gammage, K. (1971) Guidance for cyclic development and short-lived maturity in in the gametocytes of Plasmodium falciparum. Parasitology, 65, 549–59Google Scholar
Hawking, F., Worms, M. J. and Gammage, K. (1968) 24- and 48-hour cycles of malaria parasites in the blood: their purpose, production and control. Trans. R. Soc. Trop. Med. Hyg., 62, 731–60CrossRefGoogle ScholarPubMed
Hawking, F., Worms, M. J., Gammage, K. and Goddard, P. A. (1966) The biological purpose of the blood-cycle of the malaria parasite Plasmodium cynomolgi. Lancet, 2, 422–4CrossRefGoogle Scholar
Hecker, H. and Rudin, W. (1981) Morphometric parameters of the midgut cells of Aedes aegypti L. (Insecta, Diptera) under various conditions. Cell, 219, 619–27Google ScholarPubMed
Helle, T. and Aspi, J. (1983) Does herd formation reduce insect harassment among reindeer? A field experiment with animal traps. Acta Zool. Fernica, 175, 129–31Google Scholar
Hendry, G. and Godwin, G. (1988) Biting midges in Scottish forestry: a costly irritant or a trivial nuisance?Scottish Forestry, 42, 113–19Google Scholar
Henry, V. G. and Conley, R. H. (1970) Some parasites of European wild hogs in the southern Appalachians. J. Wildl. Mgmt., 34, 913–17CrossRefGoogle Scholar
Hill, C. A., Fox, A. N., Pitts, R. J.et al. (2002) G protein coupled receptors in Anopheles gambiae. Science, 298, 176–8CrossRefGoogle ScholarPubMed
Hill, P., Saunders, D. S. and Campbell, J. A. (1973) The production of ‘symbiont-free’ Glossina morsitans and an associated loss of female fertility. Trans. R. Soc. Trop. Med. Hyg., 67, 727–8CrossRefGoogle ScholarPubMed
Hillyer, J. F. and Christensen, B. M. (2002) Characterization of hemocytes from the yellow fever mosquito, Aedes aegypti. Histochemistry and Cell Biology, 117, 431–40CrossRefGoogle ScholarPubMed
Hinnebusch, B. J., Fischer, E. R. and Schwan, T. G. (1998) Evaluation of the role of the Yersinia pestis plasminogen activator and other plasmid-encoded factors in temperature-dependent blockage of the flea. Journal of Infectious Diseases, 178, 1406–15CrossRefGoogle ScholarPubMed
Hinton, H. E. (1958) The phylogeny of the panorpoid orders. Ann. Rev. Ent., 3, 181–206CrossRefGoogle Scholar
Hoc, T. Q. and Schaub, G. A. (1996) Improvement of techniques for age grading hematophagous insects: ovarian oil-injection and ovariolar separation techniques. J. Med. Ent., 33, 286–9CrossRefGoogle ScholarPubMed
Hocking, B. (1953) The intrinsic range and speed of flight of insects. Trans. R. Soc. Lond., 104, 223–345Google Scholar
Hocking, B. (1957) Louse control through textile fibre size. Bull. Ent. Res., 48, 507–14CrossRefGoogle Scholar
Hocking, B. (1971) Blood-sucking behaviour of terrestrial arthropods. Ann. Rev. Ent., 16, 1–26CrossRefGoogle ScholarPubMed
Hockmeyer, W. T., Schieffer, B. A., Redington, B. C. and Eldridge, B. V. (1975) Brugia pahangi effects upon the flight capability of Aedes aegypti. Experimental Parasitology, 38, 1–5CrossRefGoogle ScholarPubMed
Hoffmann, J. A. (2003) The immune response of Drosophila. Nature, 426, 33–8CrossRefGoogle ScholarPubMed
Hoffmann, J. A. and Reichhart, J. M. (2002) Drosophila innate immunity: an evolutionary perspective. Nature Immunology, 3, 121–6CrossRefGoogle Scholar
Hogg, J. C. and Hurd, H. (1995) Plasmodium yoelli nigeriensis – the effect of high and low intensity of infection upon the egg production and bloodmeal size of Anopheles stephensi during three gonotrophic cycles. Parasitol., 111, 555–62CrossRefGoogle ScholarPubMed
Holloway, M. T. P. and Phelps, R. J. (1991) The responses of Stomoxys spp. (Diptera, Muscidae) to traps and artificial host odors in the field. Bull. Ent. Res., 81, 51–5CrossRefGoogle Scholar
Hooke, R. (1664) Micrographia.
Hopkins, F. H. E. (1949) The host associations of the lice of mammals. Proc. Zool. Soc. Lond., 119, 387–604CrossRefGoogle Scholar
Hopwood, J. A., Ahmed, A. M., Polwart, A., Williams, G. T. and Hurd, H. (2001) Malaria-induced apoptosis in mosquito ovaries: a mechanism to control vector egg production. J. Exp. Biol., 204, 2773–2780Google ScholarPubMed
Hosoi, T. (1958) Adenosine 5' phosphates as the stimulating agent in blood for inducing gorging of the mosquito. Nature, 181, 1664–5CrossRefGoogle ScholarPubMed
Hosoi, T. (1959) Identification of blood components which induce gorging in the mosquito. J. Insect Phys., 3, 191–218CrossRefGoogle Scholar
Houseman, J. G., Downe, A. E. R. and Morrison, P. E. (1985a) Similarities in digestive proteinase production in Rhodnius prolixus (Hemiptera: Reduviidae) and Stomoxys calcitrans (Diptera: Muscidae). Insect Biochem., 15, 471–4CrossRefGoogle Scholar
Houseman, J. G., Morrison, P. E. and Downe, A. E. R. (1985b) Cathepsin B and aminopeptidase in the posterior midgut of Euschistus euschistoides (Hemiptera: Phymatidae). Can. J. Zool., 63, 1288–91CrossRefGoogle Scholar
Howe, M. A. and Lehane, M. J. (1986) Post-feed buzzing in the tsetse, Glossina morsitans morsitans, is an endothermic mechanism. Phys. Ent., 11, 279–86CrossRefGoogle Scholar
Huang, C. T. (1971) Vertebrate serum inhibitors of Aedes aegypti trypsin. Insect Biochem., 1, 27–38CrossRefGoogle Scholar
Hudson, A. (1970) Notes on the piercing mouthparts of three species of mosquitoes (Diptera: Culicidae) viewed with the scanning electron miscroscope. Can. Entomal., 102, 501–9CrossRefGoogle Scholar
Hudson, B. W., Feingold, B. F. and Kartman, L. (1960) Allergy to flea bites. II. Investigations of flea bite sensitivity in humans. Experimental Parasitology, 9, 264–70CrossRefGoogle Scholar
Huff, C. (1929) Ovulation requirements of Culex pipiens Linn. Biol. Bull. (Woods Hole), 56, 347–50CrossRefGoogle Scholar
Huff, C. (1931) A proposed classification of disease transmission by arthropods. Science, 74, 456–7CrossRefGoogle Scholar
Hughes, A. L. and Piontkivska, H. (2003) Phylogeny of trypanosomatidae and bodonidae (Kinetoplastida) based on 18S rRNA: evidence for paraphyly of Trypanosoma and six other genera. Molecular Biology and Evolution, 20, 644–52CrossRefGoogle ScholarPubMed
Humphries, D. A. (1967) Function of combs in ectoparasites. Nature, 215, 319CrossRefGoogle ScholarPubMed
Hunter, D. M. and Moorhouse, D. W. (1976) The effects of Austrosimulium pestilens on the milk production of dairy cattle. Aust. Vet. J., 52, 97–9CrossRefGoogle ScholarPubMed
Huq, M. (1961) African horse sickness. Veterinary Record, 73, 123Google Scholar
Hurd, H. (1998) Parasite manipulation of insect reproduction: who benefits? Parasitology, 116, S13–21CrossRefGoogle ScholarPubMed
Hurd, H. (2003) Manipulation of medically important insect vectors by their parasites. Ann. Rev. Ent., 48, 141–61CrossRefGoogle ScholarPubMed
Hurd, H., Hogg, J. C. and Renshaw, M. (1995) Interactions between bloodfeeding, fecundity and infection in mosquitoes. Parasitology Today, 11, 411–6CrossRefGoogle Scholar
Hursey, B. S. (2001) The programme against African trypanosomiasis: aims, objectives and achievements. Trends in Parasitology, 17, 2–3CrossRefGoogle ScholarPubMed
Ibrahim, E. A. R., Ingram, G. A. and Molyneux, D. H. (1984) Haemagglutinins and parasite agglutinins in haemolymph and gut of Glossina. Tropenmed. Parasit., 35, 151–6Google ScholarPubMed
ICZN (1999) International Code of Zoological Nomenclature. London: ICZN
Irving, P., Troxler, L., Heuer, T. S.et al. (2001) A genome-wide analysis of immune responses in Drosophila. Proceedings of the National Academy of Sciences of the United States of America, 98, 15119–24CrossRefGoogle ScholarPubMed
Isawa, H., Yuda, M., Orito, Y. and Chinzei, Y. (2002) A mosquito salivary protein inhibits activation of the plasma contact system by binding to factor Ⅻ and high molecular weight kininogen. J. Biol. Chem., 277, 27651–8CrossRefGoogle ScholarPubMed
Isawa, H., Yuda, M., Yoneda, K. and Chinzei, Y. (2000) The insect salivary protein, prolixin-S, inhibits factor IXa generation and Xase complex formation in the blood coagulation pathway. J. Biol. Chem., 275, 6636–41CrossRefGoogle ScholarPubMed
Iwanaga, S. (2002) The molecular basis of innate immunity in the horseshoe crab. Current Opinion in Immunology, 14, 87–95CrossRefGoogle ScholarPubMed
James, M. T. and Harwood, R. F. (1969) Herm's Medical Entomology. London: Macmillan
Janeway, C. A., Travers, P., Walport, M. and Schlomchik, M. (2001) Immunology. Edinburgh: Churchill Livingstone
Janse, C. J., Rouwenhorst, R. J., Klooster, P. F. J., Kaay, H. J. and Overdulve, J. P. (1985) Development of Plasmodium berghei ookinetes in the midgut of Anopheles atroparvus mosquitoes and in vitro. Parasitol., 91, 219–25CrossRefGoogle ScholarPubMed
Jefferies, D. (1984) Transmission of disease by haematophagous arthropods. Unpublished Ph.D. thesis, University of Salford
Jenkins, D. W. (1964) Advances in medical entomology using radio-isotopes. Experimental Parasitology, 3, 474–90CrossRefGoogle Scholar
Jenni, L., Molyneux, D. H., Livesey, J. L. and Galun, R. (1980) Feeding behaviour of tsetse flies infected with salivarian trypanosomes. Nature, 283, 383–5CrossRefGoogle ScholarPubMed
Jobling, B. (1976) On the fascicle of blood-sucking Diptera. In addition a description of the maxillary glands in Phlebotomus papatasi, together with the musculature of the labium and pulsatory organ of both the latter species and also of some other Diptera. J. Nat. Hist., 10(4), 457–61
Johnson, K. P., Adams, R. J. and Clayton, D. H. (2002a) The phylogeny of the louse genus Brueelia does not reflect host phylogeny. Biological Journal of the Linnaean Society, 77, 233–47CrossRefGoogle Scholar
Johnson, K. P., Weckstein, J. D., Witt, C. C., Faucett, R. C. and Moyle, R. G. (2002b) The perils of using host relationships in parasite taxonomy: phylogeny of the Degeeriella complex. Mol. Phylogenet. Evol., 23, 150–7CrossRefGoogle Scholar
Jones, C. J. (1996) Immune responses to fleas, bugs and sucking lice. In Wikel, S. K. (ed.), The Immunology of Host–Ectoparasitic Arthropod Interactions. Wallingford: CAB International, 150–74Google Scholar
Jones, J. C. and Pillitt, D. R. (1973) Blood-feeding behavior of adult Aedes aegypti mosquitoes. Biol. Bull., 145, 127–39CrossRefGoogle ScholarPubMed
Jordan, A. M. (1974) Recent development in the ecology and methods of control of tsetse flies (Glossina spp.), a review. Bull. Ent. Res., 63(4), 361–99CrossRef
Jordan, A. M. (1986) Trypanosomiasis Control and African Rural Development. London: Longman
Jordan, A. M. and Curtis, C. F. (1968) The performance of Glossina austeni when fed on lop-eared rabbits and goats. Trans. R. Soc. Trop. Med. Hyg., 62, 123–4CrossRefGoogle Scholar
Jordan, A. M. and Curtis, C. F. (1972) Productivity of Glossina morsitans morsitans Westwood maintained in the laboratory, with particular reference to the sterile-insect release method. Bull. W.H.O., 46, 33–8Google Scholar
Jordan, K. (1962) Notes on the Tunga caecigena (Siphonaptera: Tungidae). Bull. Br. Mus. Nat. Hist. (Ent.), 12(4), 353–64CrossRef
Julius, D. and Basbaum, A. I. (2001) Molecular mechanisms of nociception. Nature, 413, 203–10CrossRefGoogle ScholarPubMed
Kaaya, G. P. and Ratcliffe, N. A. (1982) Comparative study of haemocytes and associated cells of some medically important dipterans. J. Morph., 173, 351–65CrossRefGoogle Scholar
Kamhawi, S. (2000) The biological and immunomodulatory properties of sand fly saliva and its role in the establishment of Leishmania infections. Microbes Infect, 2, 1765–73CrossRefGoogle ScholarPubMed
Kamhawi, S., Belkaid, Y., Modi, G., Rowton, E. and Sacks, D. (2000) Protection against cutaneous Leishmaniasis resulting from bites of uninfected sand flies. Science, 290, 1351–4CrossRefGoogle ScholarPubMed
Kangwangye, T. N. (1977) Reactions of large mammals to biting flies in Rwenzori National Park, Uganda. In C. P. F. Lima (ed.), Proceedings of the First East African Conference on Entomological Pest Control.
Kartman, L. (1953) Factors influencing infection of the mosquito with Dirofilaria immitis (Leidy, 1856). Experimental Parasitology, 2, 27–78CrossRefGoogle Scholar
Kathirithamby, J., Ross, L. D. and Johnston, J. S. (2003) Masquerading as self? Endoparasitic Strepsiptera (Insecta) enclose themselves in host-derived epidermal bag. Proc. Natl. Acad. Sci. USA, 100, 7655–9CrossRefGoogle ScholarPubMed
Katz, O., Waitumbi, J. N., Zer, R. and Warburg, A. (2000) Adenosine, AMP, and protein phosphatase activity in sandfly saliva. Am. J. Trop. Med. Hyg., 62, 145–50CrossRefGoogle ScholarPubMed
Kavaliers, M., Choleris, E. and Colwell, D. D. (2001) Learning from others to cope with biting flies: social learning of fear-induced conditioned analgesia and active avoidance. Behavioral Neuroscience, 115, 661–74CrossRefGoogle ScholarPubMed
Keiper, R. R. and Berger, J. (1982) Refuge-seeking and pest avoidance by feral horses in desert and island environments. Applied Animal Ethology, 9, 111–20CrossRefGoogle Scholar
Kellogg, F. E. (1970) Water vapour and carbon dioxide receptors in Aedes aegypti. J. Insect Physiol., 16, 99–108CrossRefGoogle ScholarPubMed
Kellogg, F. E. and Wright, R. H. (1962) The guidance of flying insects. V. Mosquito attraction. Can. Entomol., 94, 1009–16CrossRefGoogle Scholar
Kelly, D. W. (2001) Why are some people bitten more than others?Trends in Parasitology, 17, 578–81CrossRefGoogle ScholarPubMed
Kelly, D. W., Mustafa, Z. and Dye, C. (1996) Density-dependent feeding success in a field population of the sandfly, Lutzomyia longipalpis. Journal of Animal Ecology, 65, 517–27CrossRefGoogle Scholar
Kelly, D. W. and Thompson, C. E. (2000) Epidemiology and optimal foraging: modelling the ideal free distribution of insect vectors. Parasitology, 120, 319–27CrossRefGoogle ScholarPubMed
Kennedy, J. S. (1940) The visual responses of flying mosquitoes. Proc. Zool. Soc. Lond., 109, 221–42Google Scholar
Kennedy, J. S. (1983) Zigzagging and casting as a programmed response to wind-borne odour: a review. Phys. Ent., 8, 109–20CrossRefGoogle Scholar
Kettle, D. S. (1984) Medical and Veterinary Entomology. London: Croom Helm
Khan, A. A. and Maibach, H. I. (1970) A study of the probing response of Aedes aegypti. I. Effect of nutrition on probing. J. Econ. Ent., 63, 974–6CrossRefGoogle ScholarPubMed
Khan, A. A. and Maibach, H. I. (1971) A study of the probing response of Aedes aegypti. 2. Effect of desiccation and blood feeding on probing to skin and an artificial target. J. Econ. Entomol., 64, 439–42CrossRefGoogle ScholarPubMed
Kilama, W. L. and Craig, G. B. (1969) Monofactorial inheritance of susceptibility to Plasmodium gallinaceum in Aedes aegypti. Ann. Trop. Med. Parasit., 63, 419–32CrossRefGoogle ScholarPubMed
Killeen, G. F., McKenzie, F. E., Foy, B. D., Bogh, C. and Beier, J. C. (2001) The availability of potential hosts as a determinant of feeding behaviours and malaria transmission by African mosquito populations. Trans. R. Soc. Trop. Med. Hyg., 95, 469–76CrossRefGoogle ScholarPubMed
Kim, K. C. (1985) Evolution and host association of Anoplura. In Kim, K. C. (ed.), Coevolution of Parasitic Arthropods and Mammals. New York: WileyGoogle Scholar
Kim, K. D. and Adler, P. H. (1985) Evolution and host association of Anoplura. In Kim, K. C. (ed.), Coevolution of Parasitic Arthropods and Mammals.New York: WileyGoogle Scholar
Kingsolver, J. G. (1987) Mosquito host choice and the epidemiology of malaria. Am. Nat., 130, 811–27CrossRefGoogle Scholar
Kirch, H. J., Spates, G., Droleskey, R., Kloft, W. J. and Deloach, J. R. (1991a) Mechanism of hemolysis of erythrocytes by hemolytic factors from Stomoxys calcitrans (L) (Diptera, Muscidae). J. Insect Phys., 37, 851–61CrossRefGoogle Scholar
Kirch, H. J., Spates, G., Kloft, W. J. and Deloach, J. R. (1991b) The relationship of membrane-lipids to species-specific hemolysis by hemolytic factors from Stomoxys calcitrans (L) (Diptera, Muscidae). Insect Biochem., 21, 113CrossRefGoogle Scholar
Klein, T. A., Harrison, B. A., Andre, R. G., Whitmire, R. E. and Inlao, I. (1982) Detrimental effects of Plasmodium cynomolgi infections on the longevity of Anopheles dirus. Mosq. News, 42, 265–71Google Scholar
Kline, D. L. and Lemire, G. F. (1995) Field evaluation of heat as an added attractant to traps baited with carbon dioxide and octenol for Aedes taeniorhynchus. J. Am. Mosq. Control Assoc., 11, 454–6Google ScholarPubMed
Klowden, M. J. (1993) Mating and nutritional state affect the reproduction of Aedes albopictus mosquitoes. J. Am. Mosq. Control Assoc., 9, 169–73Google ScholarPubMed
Klowden, M. J., Davis, E. E. and Bowen, M. F. (1987) Role of the fat body in the regulation of host-seeking behaviour in the mosquito, Aedes aegypti. J. Insect Physiol., 33, 643–6CrossRefGoogle Scholar
Klowden, M. J., Kline, D. L., Takken, W., Wood, J. R. and Carlson, D. A. (1990) Field studies on the potential of butanone, carbon-dioxide, honey extract, 1-octen-3-ol, L-lactic acid and phenols as attractants for mosquitos. Med. Vet. Entomol., 4, 383–91Google Scholar
Klowden, M. J. and Lea, A. O. (1979) Abdominal distention terminates subsequent host-seeking behavior of Aedes aegypti following a blood meal. J. Insect Physiol., 25, 583–5CrossRefGoogle ScholarPubMed
Knols, B. G. J., Jong, R. and Takken, W. (1995) Differential attractiveness of isolated humans to mosquitoes in Tanzania. Trans. R. Soc. Trop. Med. Hyg., 89, 604–6CrossRefGoogle ScholarPubMed
Knols, B. G. J., Mboera, L. E. G. and Takken, W. (1998) Electric nets for studying odour-mediated host-seeking behaviour of mosquitoes. Med. Vet. Entomol., 12, 116–20CrossRefGoogle ScholarPubMed
Knols, B. G. J., Loon, J. J. A., Cork, A.et al. (1997) Behavioural and electrophysiological responses of the female malaria mosquito Anopheles gambiae (Diptera: Culicidae) to Limburger cheese volatiles. Bull. Ent. Res., 87, 151–9CrossRefGoogle Scholar
Koella, J. C., Agnew, P. and Michalakis, Y. (1998a) Coevolutionary interactions between host life histories and parasite life cycles. Parasitology, 116, S47–S55CrossRefGoogle Scholar
Koella, J. C. and Boete, C. (2002) A genetic correlation between age at pupation and melanization immune response of the yellow fever mosquito Aedes aegypti. Evolution Int. J. Org. Evolution, 56, 1074–9CrossRefGoogle ScholarPubMed
Koella, J. C., Sorensen, F. L. and Anderson, R. A. (1998b) The malaria parasite, Plasmodium falciparum, increases the frequency of multiple feeding of its mosquito vector, Anopheles gambiae. Proc. R. Soc. Lond. B Biol. Sci., 265, 763–8CrossRefGoogle Scholar
Komano, H., Mizuno, D. and Natori, S. (1980) Purification of a lectin induced in the haemolymph of Sarcophaga peregrina larvae on injury. J. Biol. Chem., 255, 2919–24Google Scholar
Kramer, L. D., Hardy, J. L., Presser, S. B. and Houk, E. G. (1981) Dissemination barriers for western equine encephalomyelitis virus in Culex tarsalis infected after ingestion of low viral doses. Am. J. Trop. Med. Hyg., 30, 190–7CrossRefGoogle ScholarPubMed
Krasnov, B. R., Khokhlova, I. S. and Shenbrot, G. I. (2003a) Density-dependent host selection in ectoparasites: an application of isodar theory to fleas parasitizing rodents. Oecologia, 134, 365–72CrossRefGoogle Scholar
Krasnov, B. R., Sarfati, M., Arakelyan, M. S.et al. (2003b) Host specificity and foraging efficiency in blood-sucking parasite: feeding patterns of the flea Parapulex chephrenis on two species of desert rodents. Parasitology Research, 90, 393–9CrossRefGoogle Scholar
Krynski, S., Kuchta, A. and Becla, E. (1952) Research on the nature of the noxious action of guinea-pig blood on the body louse (in Polish). Bull. Inst. Mar. Med. Gdansk, 4, 104–7Google Scholar
Ksiazkiewicz-Ilijewa, M. and Rosciszewska, E. (1979) Ultrastructure of the haemocytes of Tetrodontophora bielanensis Waga (Collembola). Cytobios, 26, 113–21Google Scholar
Kunz, S. E., Murrell, K. D., Lambert, G., James, L. F. and Terrill, C. E. (1991) Estimated losses of livestock to pests. In Pimentel, D. (ed.), CRCHandbook of Pest Management in Agriculture. Boca Raton, Florida: CRC Press, Vol. 1, pp. 69–98Google Scholar
Kurata, S. (2004) Recognition of infectious non-self and activation of immune responses by peptidoglycan recognition protein (PGRP)-family members in Drosophila. Dev. Comp. Immunol., 28, 89–95CrossRefGoogle ScholarPubMed
Kurtz, J. and Franz, K. (2003) Evidence for memory in invertebrate immunity. Nature, 425, 37–8CrossRefGoogle ScholarPubMed
Breque, G. C., Meifert, D. W. and Rye, J. (1972) Experimental control of stable flies, Stomoxys calcitrans (Diptera: Muscidae), by the release of chemosterilized adults. Can. Entomol., 104, 885–7Google Scholar
Laarman, J. J. (1958) The host-seeking behaviour of anopheline mosquitoes. Trop. Geogr. Med., 10, 293–305Google ScholarPubMed
Lackie, A. M. (1986) Evasion of insect immunity by helminth larvae. In Lackie, A. M. (ed.), Immune Mechanisms in Invertebrate Vectors, Oxford: Oxford University Press, Vol. 56, 161–78Google Scholar
Lafond, M. M., Christensen, B. M. and Lasee, B. A. (1985) Defense reactions of mosquitoes to filarial worms: potential mechanisms for avoidance of the response by Brugia pahangi microfilaria. J. Invert. Pathol., 46, 26–30CrossRefGoogle Scholar
Lainson, R. and Shaw, J. J. (1987) Evolution, classification and geographical distribution. In Peters, W. and Killick-Kendrick, R. (eds.), The Leishmaniases in Biology and Medicine. New York: Academic PressGoogle Scholar
Lall, S. B. (1969) Phagostimulants of haematophagous tabanids. Entomol. Exp. Appl., 12, 325–36CrossRefGoogle Scholar
Land, M. F., Gibson, G. and Horwood, J. (1997) Mosquito eye design: conical rhabdoms are matched to wide aperture lenses. Proc. R. Soc. Lond. B. Biol. Sci., 264, 1183–7CrossRefGoogle Scholar
Land, M. F., Gibson, G., Horwood, J. and Zeil, J. (1999) Fundamental differences in the optical structure of the eyes of nocturnal and diurnal mosquitoes. J. Comp. Physiol., 185, 91–103CrossRefGoogle Scholar
Lane, N. J. and Harrison, J. B. (1979) An unusual cell surface modification: a double plasma membrane. J. Cell Sci., 39, 355–72Google ScholarPubMed
Langley, P. A. (1970) Post-teneral development of thoracic flight musculature in the tsetse flies Glossina austeni and G. morsitans. Entomologia Exp. Appl., 13, 133–40CrossRefGoogle Scholar
Langley, P. A. and Maly, H. (1969) Membrane feeding technique for tsetse flies (Glossina spp.). Nature, 221, 855–6CrossRefGoogle Scholar
Lanzaro, G. C., Toure, Y. T., Carnahan, J.et al. (1998) Complexities in the genetic structure of Anopheles gambiae populations in west Africa as revealed by microsatellite DNA analysis. Proceedings of the National Academy of Sciences of the United States of America, 95, 14260–5CrossRefGoogle ScholarPubMed
Larrivee, D. H., Benjamini, E., Feingold, B. F. and Shimuzu, M. (1964) Histologic studies of guinea pig skin: different stages of allergic reactivity to flea bites. Exp. Parasitol., 15, 491–502CrossRefGoogle ScholarPubMed
Laurence, B. R. (1966) Intake and migration of the microfilariae of Onchocerca volvulus (Leukart) in Simulium damnosum Theobald. J. Helm., 40, 337–42CrossRefGoogle Scholar
Laurence, B. R. and Pester, F. R. N. (1961) The ability of Anopheles gambiae Giles to transmit Brugia patei (Buckley, Nelson and Heisch). J. Trop. Med. Hyg., 64, 169–71Google Scholar
Laurence, B. R. and Pester, F. R. N. (1967) Adaptation of a filarial worm, Brugia patei, to a new mosquito host, Aedes togoi. J. Helminth., 41, 365–92CrossRefGoogle Scholar
Lavine, M. D. and Strand, M. R. (2002) Insect hemocytes and their role in immunity. Insect Biochem. Mol. Biol., 32, 1295–309CrossRefGoogle ScholarPubMed
Lavoipierre, M. M. J., Dickerson, G. and Gordon, R. M. (1959) Studies on the methods of feeding of blood-sucking arthropods. I. The manner in which triatomine bugs obtain their blood meal as observed in the tissues of the living rodent, with some remarks on the effects of the bite on human volunteers. Ann. Trop. Med. Parasitol., 53, 235–50CrossRefGoogle Scholar
Lavoipierre, M. M. J. and Hamachi, M. (1961) An apparatus for observations on the feeding mechanism of the flea. Nature, 192, 998–9CrossRefGoogle Scholar
Lazzari, C. R., Reiseman, C. E. and Insausti, T. C. (1998) The role of the ocelli in the phototactic behaviour of the haematophagous bug Triatoma infestans. J. Insect Physiol., 44, 1159–62CrossRefGoogle ScholarPubMed
Lebestky, T., Chang, T., Hartenstein, V. and Banerjee, U. (2000) Specification of Drosophila hematopoietic lineage by conserved transcription factors. Science, 288, 146–9CrossRefGoogle ScholarPubMed
Lee, J. H., Rowley, W. A. and Platt, K. B. (2000) Longevity and spontaneous flight activity of Culex tarsalis (Diptera: Culicidae) infected with western equine encephalomyelitis virus. J. Med. Ent., 37, 187–93CrossRefGoogle ScholarPubMed
Lee, R. (1974) Structure and function of the fascicular stylets, and the labral and cibarial sense organs of male and female Aedes aegypti (L>). Quest. Entomol., 10, 187–215Google Scholar
Lehane, M. J. (1976a) Digestive enzyme secretion in Stomoxys calcitrans (Diptera: Muscidae). Tissue and Cell, 170, 275–87Google Scholar
Lehane, M. J. (1976b) The formation and histochemical structure of the peritrophic membrane in the stable fly, Stomoxys calcitrans. J. Insect Physiol., 22, 1551–7CrossRefGoogle Scholar
Lehane, M. J. (1977a) An hypothesis of the mechanism controlling proteolytic digestive enzyme production levels in Stomoxys calcitrans. J. Insect Physiol., 23, 713–15CrossRefGoogle Scholar
Lehane, M. J. (1977b) Transcellular absorption of lipids in the midgut of the stablefly, Stomoxys calcitrans. J. Insect Physiol., 23, 945–54CrossRefGoogle Scholar
Lehane, M. J. (1985) Determining the age of an insect. Parasitology Today, 1, 81–5CrossRefGoogle ScholarPubMed
Lehane, M. J. (1987) Quantitative evidence for merocrine secretion in an insect midgut cell. Tissue and Cell, 19, 451–561CrossRefGoogle Scholar
Lehane, M. J. (1988) Evidence for secretion by the release of cytoplasmic extrusions from midgut cells of Stomoxys calcitrans. J. Insect Physiol., 34, 949–53CrossRefGoogle Scholar
Lehane, M. J. (1989) The intracellular pathway and kinetics of digestive enzyme secretion in an insect midgut cell. Tissue and Cell, 21, 101–11CrossRefGoogle Scholar
Lehane, M. J. (1997) Peritrophic matrix structure and function. Ann. Rev. Ent., 42, 525–50CrossRefGoogle ScholarPubMed
Lehane, M. J., Aksoy, S., Gibson, W. (2003) Adult midgut EST from the tsetse fly Glossina morsitans morsitans and expression analysis of putative immune response genes. Genome Biology, 4 (10), R63CrossRefGoogle ScholarPubMed
Lehane, M. J., Aksoy, S. and Levashina, E. A. (2004) Blood-sucking insect immune responses and parasite transmission. Trends in Parasitology, in press
Lehane, M. J., Allingham, P. G. and Weglicki, P. (1996) Peritrophic matrix composition of the tsetse fly, Glossina morsitans morsitans. Cell and Tissue Research, 272, 158–62Google Scholar
Lehane, M. J. and Billingsley, P. A. (eds.) (1996) The Biology of the Insect Midgut.London: Chapman and HallCrossRefGoogle Scholar
Lehane, M. J., Crisanti, A. and Mueller, H. M. (1996) Mechanisms controlling the synthesis and secretion of digestive enzymes in insects. In Lehane, M. J. (ed.), The Insect Midgut. London: Chapman and HallCrossRefGoogle Scholar
Lehane, M. J. and Hargrove, J. (1988) Field experiments on a new method for determining age in tsetse flies (Diptera, Glossinidae). Ecol. Entomol., 13, 319–22CrossRefGoogle Scholar
Lehane, M. J. and Laurence, B. R. (1977) Flight muscle ultrastructure of susceptible and refractory mosquitoes parasitized by larval Brugia pahangi. Parasitology, 74, 87–92CrossRefGoogle ScholarPubMed
Lehane, M. J. and Mail, T. S. (1985) Determining the age of adult male and female Glossina morsitans morsitans using a new technique. Ecol. Entomol., 10, 219–24CrossRefGoogle Scholar
Lehane, M. J. and Schofield, C. J. (1981) Field experiments of dispersive flight by Triatoma infestans. Trans. R. Soc. Trop. Med. Hyg., 75, 399–400CrossRefGoogle ScholarPubMed
Lehane, M. J. and Schofield, C. J. (1982) Flight initiation in Triatoma infestans (Klug) (Hemiptera: Reduviidae). Bull. Ent. Res., 72, 497–510CrossRefGoogle Scholar
Lehane, M. J., Wu, D. and Lehane, S. M. (1997) Midgut-specific immune molecules are produced by the blood-sucking insect Stomoxys calcitrans. Proceedings of the National Academy of Sciences of the United States of America, 94, 11502–7CrossRefGoogle ScholarPubMed
Lehane, S. M., Assinder, S. J. and Lehane, M. J. (1998) Cloning, sequencing, temporal expression and tissue-specificity of two serine proteases from the midgut of the blood-feeding fly Stomoxys calcitrans. European Journal of Biochemistry, 254, 290–6CrossRefGoogle ScholarPubMed
Lemaitre, B., Reichhart, J. M. and Hoffmann, J. A. (1997) Drosophila host defense: differential induction of antimicrobial peptide genes after infection by various classes of microorganisms. Proceedings of the National Academy of Sciences of the United States of America, 94, 14614–19CrossRefGoogle ScholarPubMed
Lemos, F. J. A., Cornel, A. J. and Jacobs Lorena, M. (1996) Trypsin and aminopeptidase gene expression is affected by age and food composition in Anopheles gambiae. Insect Biochemistry and Molecular Biology, 26, 651–8CrossRefGoogle ScholarPubMed
Lester, H. M. O. and Lloyd, L. (1929) Notes on the process of digestion in tsetse flies. Bull. Ent. Res., 19, 39–60CrossRefGoogle Scholar
Leulier, F., Parquet, C., Pili-Floury, S. et al. (2003) The Drosophila immune system detects bacteria through specific peptidoglycan recognition. Nature Immunology, 4, 478–84CrossRefGoogle ScholarPubMed
Levashina, E. A., Langley, E., Green, C. et al. (1999) Constitutive activation of toll-mediated antifungal defense in serpin-deficient Drosophila. Science, 285, 1917–19CrossRefGoogle ScholarPubMed
Levashina, E. A., Moita, L. F., Blandin, S.et al. (2001) Conserved role of a complement-like protein in phagocytosis revealed by dsRNA knockout in cultured cells of the mosquito Anopheles gambiae. Cell, 104, 709–18CrossRefGoogle ScholarPubMed
Lewis, D. J. (1953) Simulium damnosum and its relation to Onchocerciasis in the Anglo-Egyptian Sudan. Bull. Ent. Res., 43, 597–644CrossRefGoogle Scholar
Lewis, L. F., Christenson, D. M. and Eddy, G. W. (1967) Rearing the long-nosed cattle louse and cattle-biting louse on host animals in Oregon. J. Econ. Entomol., 60, 755–7CrossRefGoogle Scholar
Lewis, T. and Taylor, L. R. (1965) Diurnal periodicity of flight by insects. Trans. R. Ent. Soc. Lond., 116, 393–479CrossRefGoogle Scholar
Li, X., Sina, B. and Rossignol, P. A. (1992) Probing behaviour and sporozoite delivery by Anopheles stephensi infected with Plasmodium berghei. Med. Vet. Entomol., 6, 57–61CrossRefGoogle ScholarPubMed
Ligoxygakis, P., Pelte, N., Hoffmann, J. A. and Reichart, J. M. (2002) Activation of Drosophila toll during fungal infection by a blood serine protease. Nature Reviews Immunology, 2, 545Google Scholar
Lindsay, L. B. and Galloway, T. D. (1998) Reproductive status of four species of fleas (Insecta: Siphonaptera) on Richardson's ground squirrels (Rodentia: Sciuridae) in Manitoba, Canada. J. Med. Ent., 35, 423–30CrossRefGoogle Scholar
Lindsay, S. W., Adiamah, J. H., Miller, J. E., Pleass, R. J. and Armstrong, J. R. M. (1993) Variation in attractiveness of human subjects to malaria mosquitoes (Diptera, Culicidae) in the Gambia. J. Med. Ent., 30, 368–73CrossRefGoogle Scholar
Linley, J. R. and Davies, J. B. (1971) Sandflies and tourism in Florida and the Bahamas and Caribbean area. J. Econ. Entomol., 64, 264–78CrossRefGoogle Scholar
Linsenmair, K. E. (1973) Die Windorientierurung laufender Insekten. Fortschr. Zool., 21, 59–79Google Scholar
Liu, C. T., Hou, R. F. and Chen, C. C. (1998) Formation of basement membrane-like structure terminates the cellular encapsulation of microfilariae in the haemocoel of Anopheles quadrimaculatus. Parasitology, 116 (Pt 6), 511–18CrossRefGoogle ScholarPubMed
Lochmiller, R. L. and Deerenberg, C. (2000) Trade-offs in evolutionary immunology: just what is the cost of immunity?Oikos, 88, 87–98CrossRefGoogle Scholar
Loder, P. M. J., Hargrove, J. W. and Randolph, S. E. (1998) A model for blood meal digestion and fat metabolism in male tsetse flies (Glossinidae). Phys. Ent., 23, 43–52CrossRefGoogle Scholar
Lodmell, D. L., Bell, J. F., Clifford, C. M., Moore, G. J. and Raymond, G. (1970) Effects of limb disability on lousiness in mice. V. Hierarchy disturbance on mutual grooming and reproductive capacities. Expl. Parasit., 27, 184–92CrossRefGoogle Scholar
Loke, H. and Randolph, S. E. (1995) Reciprocal regulation of fat-content and flight activity in male tsetse-flies (Glossina palpalis). Phys. Ent., 20, 243–7CrossRefGoogle Scholar
Lord, W. D., DiZinno, J. A., Wilson, M. R. et al. (1998) Isolation, amplification, and sequencing of human mitochondrial DNA obtained from human crab louse, Pthirus pubis (L.), blood meals. Journal of Forensic Sciences, 43, 1097–100CrossRefGoogle ScholarPubMed
Loudon, C. and McCulloh, K. (1999) Application of the Hagen-Poiseuille equation to fluid feeding through short tubes. Ann. Ent. Soc. Am., 92, 153–8CrossRefGoogle Scholar
Lowenberger, C. A., Ferdig, M. T., Bulet, P. et al. (1996) Aedes aegypti – induced antibacterial proteins reduce the establishment and development of Brugia malayi. Experimental Parasitology, 83, 191–201CrossRefGoogle ScholarPubMed
Lowther, J. K. and Wood, D. M. (1964) Specificity of a black fly, Simulium euryadmiculum Davies, towards its host, the common loon. Can. Entomol., 96, 911–13CrossRefGoogle Scholar
Luckhart, S. and Rosenberg, R. (1999) Gene structure and polymorphism of an invertebrate nitric oxide synthase gene. Gene, 232, 25–34CrossRefGoogle ScholarPubMed
Luckhart, S., Vodovotz, Y., Cui, L. W. and Rosenberg, R. (1998) The mosquito Anopheles stephensi limits malaria parasite development with inducible synthesis of nitric oxide. Proceedings of the National Academy of Sciences of the United States of America, 95, 5700–5CrossRefGoogle ScholarPubMed
Lyman, D. E., Monteiro, F. A., Escalante, A. A. et al. (1999) Mitochondrial DNA sequence variation among triatomine vectors of Chagas' disease. Am. J. Trop. Med. Hyg., 60, 377–86CrossRefGoogle ScholarPubMed
Lythgoe, K. A. (2000) The coevolution of parasites with host-acquired immunity and the evolution of sex. Evolution Int. J. Org. Evolution, 54, 1142–56CrossRefGoogle ScholarPubMed
Maa, T. C. and Marshall, A. G. (1981) Diptera Pupipara of the New Hebrides (South Pacific): taxonomy, zoogeography, host association and ecology. Q. Jl. Taiwan Mus., 34, 213–32Google Scholar
MacCormack, C. P. (1984) Human ecology and behaviour in malaria control in tropical Africa. Bull. WHO, 62, 81–7Google ScholarPubMed
Macdonald, W. W. (1962a) The genetic basis of susceptibility to infection with semi-periodic Brugia malayi in Aedes aegypti. Ann. Trop. Med. Parasit., 56, 373–82CrossRefGoogle Scholar
Macdonald, W. W. (1962b) The selection of a strain of Aedes aegypti susceptible to infection with semi-periodic Brugia malayi. Ann. Trop. Med. Parasit., 56, 368–72CrossRefGoogle Scholar
Macdonald, W. W. (1963) Further studies on a strain of Aedes aegypti susceptible to infection with sub-periodic Brugia malayi. Ann. Trop. Med. Parasit., 57, 452–60CrossRefGoogle ScholarPubMed
Macdonald, W. W. and Ramachandran, A. (1965) The influence of the gene fm (filarial susceptibility, Brugia malayi) on the susceptibility of Aedes aegypti to several strains of Brugia, Wuchereria and Dirofilaria. Annals of Tropical Medicine and Parasitology, 59, 64–73CrossRefGoogle ScholarPubMed
Mackie, F. P. (1907) The part played by Pediculus corporis in the transmission of relapsing fever. British Medical Journal, 2, 1706–9CrossRef
Macvicker, J. A. K., Billingsley, P. F., Djamgoz, M. B. A. and Harrow, I. D. (1994) Ouabain-sensitive Na+/K+-ATPase activity in the reservoir zone of the midgut of Stomoxys calcitrans (Diptera, Muscidae). Insect Biochemistry and Molecular Biology, 24, 151–9CrossRefGoogle Scholar
Maddrell, S. H. P. (1963) Control of ingestion in Rhodnius prolixus Stal. Nature, 198, 210CrossRefGoogle Scholar
Maddrell, S. H. P. (1980) Characteristics of epithelial transport in insect Malpighian tubules. Curr. Topics Memb. Transport, 14, 427–63CrossRefGoogle Scholar
Magesa, S. M., Mdira, Y. K., Akida, J. A., Bygbjerg, I. C. and Jakobsen, P. H. (2000) Observations on the periodicity of Plasmodium falciparum gametocytes in natural human infections. Acta Trop., 76, 239–46CrossRefGoogle ScholarPubMed
Mahmood, F. (2000) Susceptibility of geographically distinct Aedes aegypti L. from Florida to Dirofilaria immitis (Leidy) infection. J. Vector Ecol., 25, 36–47Google ScholarPubMed
Mahon, R. and Gibbs, A. (1982) Arbovirus-infected hens attract more mosquitoes. In MacKenzie, J. S. (ed.), Viral Diseases in South Esat Asia and the Western Pacific. New York: Academic PressGoogle Scholar
Maier, W. A. and Omer, O. (1973) Der einfluss von Plasmodium cathemerium auf den Aminosauregehalt und die eizahl von Culex pipiens fatigans. Z. Parasit., 42, 265–78CrossRefGoogle Scholar
Malhotra, I., Ouma, J. H., Wamachi, A. et al. (2003) Influence of maternal filariasis on childhood infection and immunity to Wuchereria bancrofti in Kenya. Infection and Immunity, 71, 5231–7CrossRefGoogle ScholarPubMed
Mallon, E. B., Loosli, R. and Schmid-Hempel, P. (2003) Specific versus nonspecific immune defense in the bumblebee, Bombus terrestris L. Evolution, 57, 1444–7Google ScholarPubMed
Mans, B. J., Louw, A. I. and Neitz, A. W. H. (2002) Evolution of hematophagy in ticks: common origins for blood coagulation and platelet aggregation inhibitors from soft ticks of the genus Ornithodoros. Molecular Biology and Evolution, 19, 1695–705CrossRefGoogle ScholarPubMed
Manson, P. (1878) On the development of Filaria sanguinis hominis, and on the mosquito considered as a nurse. J. Linn. Soc. Zool. London, 14, 304–11CrossRefGoogle Scholar
Marchoux, E. and Salinberi, A. (1903) La spirillose des poules. Annales de la Institut Pasteur, 17, 569–80
Margalit, J., Galun, R. and Rice, M. J. (1972) Mouthpart sensilla of the tsetse fly and their function. I. Feeding patterns. Ann. Trop. Med. Parasit., 66, 525–36CrossRefGoogle ScholarPubMed
Marshall, A. G. (1981) The Ecology of Ectoparasitic Insects. New York: Academic Press
Marx, R. (1955) Über die wirtsfindung und die Bedeutung de artspezifischen duftstoffes bei Cimex lectularius Linne. Z. Parasit., 17, 41–72CrossRefGoogle Scholar
Masaninga, F. and Mihok, S. (1999) Host influence on adaptation of Trypanosoma congolense metacyclics to vertebrate hosts. Med. Vet. Entomol., 13, 330–2CrossRefGoogle ScholarPubMed
Matsumoto, Y., Oda, Y., Uryu, M. and Hayakawa, Y. (2003) Insect cytokine growth-blocking peptide triggers a termination system of cellular immunity by inducing its binding protein. J. Biol. Chem., 278, 38579–85CrossRefGoogle ScholarPubMed
Matthysse, J. G. (1946) Cattle lice: their biology and control. Cornell Agr. Exp. Sta. Bull, 832, 1–67Google Scholar
Mattingley, P. F. (1965) The evolution of parasite–arthropod vector systems. In Taylor, A. E. R. (ed.), Symposium of the British Society for Parasitology. Oxford:Blackwell, Vol. 3Google Scholar
Maudlin, I. and Dukes, P. (1985) Extrachromosomal inheritance of susceptibility to trypanosome infection in tsetse flies. I. Selection of susceptible and refractory strains of Glossina morsitans morsitans. Ann. Trop. Med. Parasit., 79, 317–24CrossRefGoogle Scholar
Maudlin, I. and Ellis, D. (1985) Association between intracellular rickettsia-like infections of midgut cells and susceptibility to trypanosome infections in Glossina species. Z. Parasit., 71, 683–7CrossRefGoogle Scholar
Maudlin, I., Kabayo, J. P., Flood, M. E. T. and Evans, D. A. (1984) Serum factors and the maturation of Trypanosoma congolense infections in Glossina morsitans. Z. Parasit., 70, 11–19CrossRefGoogle ScholarPubMed
Maudlin, I. and Welburn, S. C. (1987) Lectin-mediated establishment of midgut infections of Trypanosoma congolense and Trypanosoma bruce in Glossina morsitans. Tropical Medicine and Parasitology, 38, 167–70Google Scholar
Maudlin, I., Welburn, S. C. and Milligan, P. J. M. (1998) Trypanosome infections and survival in tsetse. Parasitology, 116, S23–S28CrossRefGoogle ScholarPubMed
Mayer, M. S. and James, J. D. (1969) Attraction of Aedes aegypti (L.): responses to human arms, carbon dioxide, and air currents in a new type of olfactometer. Bull. Ent. Res., 58, 629–42CrossRefGoogle Scholar
Mayer, M. S. and James, J. D. (1970) Attraction of Aedes aegypti. II. Velocity of reaction to host with and without additional carbon dioxide. Ent. Exp. Appl., 13, 47–53CrossRefGoogle Scholar
Mbow, M. L., Bleyenberg, J. A., Hall, L. R. and Titus, R. G. (1998) Phlebotomus papatasi sand fly salivary gland lysate down-regulates a Th1, but up-regulates a Th2, response in mice infected with Leishmania major. Journal of Immunology, 161, 5571–7Google ScholarPubMed
McCabe, C. T. and Bursell, E. (1975a) Interrelationships between amino acid and lipid metabolism in the tsetse fly, Glossina morsitans. Insect Biochem., 5, 781–9CrossRefGoogle Scholar
McCabe, C. T. and Bursell, E. (1975b) Metabolism of digestive products in the tsetse fly, Glossina morsitans. Insect Biochem., 5, 769–79CrossRefGoogle Scholar
McCall, P. J. and Kelly, D. W. (2002) Learning and memory in disease vectors. Trends in Parasitology, 18, 429–33CrossRefGoogle ScholarPubMed
McCall, P. J. and Lemoh, P. A. (1997) Evidence for the ‘invitation effect’ during bloodfeeding by blackflies of the Simulium damnosum complex (Diptera, Simuliidae). Journal of Insect Behavior, 10, 299–303CrossRefGoogle Scholar
McCall, P. J., Mosha, F. W., Njunwa, K. J. and Sherlock, K. (2001) Evidence for memorized site-fidelity in Anopheles arabiensis. Trans. R. Soc. of Trop. Med. Hyg., 95, 587–90CrossRefGoogle ScholarPubMed
McDermott, M. J., Weber, E., Hunter, S. et al. (2000) Identification, cloning, and characterization of a major cat flea salivary allergen (Cte f 1). Molecular Immunology, 37, 361–75CrossRefGoogle Scholar
McGavin, G. C. (2001) Essential Entomology: An Order by Order Introduction. Oxford: Oxford University Press
McGreevy, P. B., Bryan, J. H., Oothuman, P. and Kolstrup, N. (1978) The lethal effects of the cibarial and pharyngeal armatures of mosquitoes on microfilariae. Trans. R. Soc. Trop. Med. Hyg., 74, 361–8CrossRefGoogle Scholar
McGreevy, P. B., McClelland, G. A. H. and Lavoipierre, M. M. J. (1974) Inheritance of susceptibility to Dirofilaria immitis infection in Aedes aegypti. Ann. Trop. Med. Parasit., 68, 97–109CrossRefGoogle ScholarPubMed
McKeever, S. (1977) Observations of Corethrella feeding on tree frogs (Hyla). Mosq. News, 37, 522Google Scholar
McKeever, S. and French, F. E. (1991) Corethrella (Diptera, Corethrellidae) of Eastern North-America – Laboratory Life-History and Field Responses to Anuran Calls. Ann. Ent. Soc. Am., 84, 493–7CrossRefGoogle Scholar
McKelvey, J. J. (1973) Man against Tsetse: Struggle for Africa. Ithaca: Cornell University Press
Mead-Briggs, A. R. (1964) The reproductive biology of the rabbit flea Spilopsyllus cuniculi (Dale) and the dependance of this species on the upon the breeding of its host. J. Exp. Biol., 41, 371–402Google Scholar
Medvedev, S. I. and Skylar, V. Y. (1974) Beetles (Coleoptera) from nests of small mammals in Donotsk Province (in Russian). Entomologicheskoe Obozrenie, 53, 561–71Google Scholar
Meijerink, J., Braks, M. A. H., Brack, A. A. et al. (2000) Identification of olfactory stimulants for Anopheles gambiae from human sweat samples. Journal of Chemical Ecology, 26, 1367–82CrossRefGoogle Scholar
Meister, M. and Lagueux, M. (2003) Drosophila blood cells. Cell Microbiol., 5, 573–80CrossRefGoogle ScholarPubMed
Mellink, J. J. (1981) Selections for blood-feeding efficiency in colonized Aedes aegypti. Mosq. News, 41, 119–25Google Scholar
Mellink, J. J. and Bovenkamp, W. (1981) Functional aspects of mosquito salivation in blood feeding in Aedes aegypti. Mosq. News, 41, 110–15Google Scholar
Mellor, P. S. and Boorman, J. (1980) Multiplication of the bluetongue virus in Culicoides nubeculosus (Meigen) simultaneously infected with the virus and microfilaria of Onchocerca cervicalis (Railliet and Henry). Ann. Trop. Med. Parasit., 74, 463–9CrossRefGoogle Scholar
Menezes, H. and Jared, C. (2002) Immunity in plants and animals: common ends through different means using similar tools. Comp. Biochem. Physiol. C Toxicol. Pharmacol., 132, 1–7CrossRefGoogle ScholarPubMed
Mews, A. R., Baumgartner, H., Luger, D. and Offori, E. D. (1976) Colonisation of Glossina morsitans morsitans Westw. in the laboratory using in vitro feeding techniques. Bull. Ent. Res., 65, 631–42CrossRefGoogle Scholar
Miall, R. C. (1978) The flicker fusion frequencies of six laboratory insects, and the response of the compound eye to mains fluorescent ‘ripple’. Phys. Ent., 3, 99–106CrossRefGoogle Scholar
Michel, T., Reichhart, J. M., Hoffmann, J. A. and Royet, J. (2001) Drosophila toll is activated by Gram-positive bacteria through a circulating peptidoglycan recognition protein. Nature, 414, 756–9CrossRefGoogle ScholarPubMed
Miller, N. and Lehane, M. J. (1990) In vitro perfusion studies on the peritrophic membrane of the tsetse fly Glossina morsitans morsitans (Diptera: Glossinidae). J. Insect Phys., 36, 813–18CrossRefGoogle Scholar
Minchella, D. J. (1985) Host life-history variation in response to parasitism. Parasitol., 90, 205–16CrossRefGoogle Scholar
Minchella, D. J. and Loverde, P. T. (1983) Laboratory comparison of the relative success of Biomphalaria glabrata stocks which are susceptible and insusceptible to infection with Schistosoma mansoni. Parasitology, 86, 335–44CrossRefGoogle ScholarPubMed
Mitchell, B. K. and Reinouts van Haga-Kelker, H. A. (1976) A comparison of the feeding behaviour in teneral and post-teneral Glossina morsitans (Diptera, Glossinidae) using an artificial membrane. Ent. Exp. Appl., 20, 105–12CrossRefGoogle Scholar
Mockford, E. L. (1967) Some Psocoptera from the plumage of birds. Proc. Ent. Soc. Washington, 69, 307–9Google Scholar
Mockford, E. L. (1971) Psocoptera from the dusky-footed wood rat in southern California (Psocoptera: Atropidae, Psoguillidae, Liposcelidae). Pan-Pacific Entomologist, 47, 127–40Google Scholar
Moffatt, M. R., Blakemore, D. and Lehane, M. J. (1995) Studies on the synthesis and secretion of digestive trypsin in Stomoxys calcitrans (Insecta-Diptera). Comp. Biochem. Phys. B, 110B, 291–300CrossRefGoogle Scholar
Mohr, C. O. (1943) Cattle droppings as ecological units. Ecol. Monographs, 13, 275CrossRefGoogle Scholar
Moloo, S. K. (1983) Feeding behaviour of Glossina morsitans morsitans infected with Trypanosoma vivax, T. congolense or T. brucei. Parasit., 86, 51–6CrossRefGoogle ScholarPubMed
Moloo, S. K. and Dar, F. (1985) Probing by Glossina morsitans centralis infected with pathogenic Trypanosoma species. Trans. R. Soc. of Trop. Med. Hyg., 79, 119CrossRefGoogle ScholarPubMed
Moloo, S. K. and Kutuza, S. B. (1970) Feeding and crop-emptying in Glossina brevipalpis Newstead. Acta Trop., 27, 356–77Google ScholarPubMed
Moloo, S. K., Sabwa, C. L. and Baylis, M. (2000) Feeding behaviour of Glossina pallidipes and G. morsitans centralis on Boran cattle infected with Trypanosoma congolense or T. vivax under laboratory conditions. Med. Vet. Entomol., 14., 290–9CrossRefGoogle Scholar
Molyneux, D. H. (1984) Evolution of the Trypanosomatidae: considerations of polyphyletic origins of mammalian parasites. CNRS/INSERM, 1986, 231–40Google Scholar
Molyneux, D. H., Bradley, M., Hoerauf, A., Kyelem, D. and Taylor, M. J. (2003) Mass drug treatment for lymphatic filariasis and onchocerciasis. Trends in Parasitology, 19, 516–22CrossRefGoogle ScholarPubMed
Molyneux, D. H. and Killick-Kendrick, R. (1987) Morphology, ultrastructure and life cycles. In Peters, W. and Killick-Kendrick, R. (eds.), The Leishmaniases in Biology and Medicine. New York: Academic PressGoogle Scholar
Molyneux, D. H., Killick-Kendrick, R. and Ashford, R. W. (1975) Leishmania in phlebotomid sandflies. III. The ultrastructure of Leishmania mexicana amazonensis in the midgut and pharynx of Lutzomyia longipalpis. Proc. R. Soc. B, 190, 341–57CrossRefGoogle ScholarPubMed
Montfort, W. R., Weichsel, A. and Andersen, J. F. (2000) Nitrophorins and related antihemostatic lipocalins from Rhodnius prolixus and other blood-sucking arthropods. Biochimica et Biophysica Acta – Protein Structure and Molecular Enzymology, 1482, 110–18CrossRefGoogle ScholarPubMed
Mooring, M. S., Benjamin, J. E., Harte, C. R. and Herzog, N. B. (2000) Testing the interspecific body size principle in ungulates: the smaller they come, the harder they groom. Anim. Behav., 60, 35–45CrossRefGoogle ScholarPubMed
Mooring, M. S. and Hart, B. L. (1992) Animal grouping for protection from parasites – selfish herd and encounter-dilution effects. Behaviour, 123, 173–93CrossRefGoogle Scholar
Morand, S. and Poulin, R. (1998) Density, body mass and parasite species richness of terrestrial mammals. Evolutionary Ecology, 12, 717–27CrossRefGoogle Scholar
Moret, Y. and Schmid-Hempel, P. (2000) Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 1166–8CrossRefGoogle Scholar
Moro, O. and Lerner, E. A. (1997) Maxadilan, the vasodilator from sand flies, is a specific pituitary adenylate cyclase activating peptide type I receptor agonist. J. Biol. Chem., 272, 966–70CrossRefGoogle ScholarPubMed
Morris, R. V., Shoemaker, C. B., David, J. R., Lanzaro, G. C. and Titus, R. G. (2001) Sandfly maxadilan exacerbates infection with Leishmania major and vaccinating against it protects against L. major infection. J. Immunol., 167, 5226–30CrossRefGoogle ScholarPubMed
Moskalyk, L. A. and Friend, W. G. (1994) Feeding behavior of female Aedes aegypti – effects of diet, temperature, bicarbonate and feeding technique on the response to ATP. Phys. Ent., 19, 223–9CrossRefGoogle Scholar
Moyer, B. R., Gardiner, D. W. and Clayton, D. H. (2002) Impact of feather molt on ectoparasites: looks can be deceiving. Oecologia, 131, 203–10CrossRefGoogle ScholarPubMed
Msangi, A. R., Whitaker, C. J. and Lehane, M. J. (1998) Factors influencing the prevalence of trypanosome infection of Glossina pallidipes on the Ruvu flood plain of Eastern Tanzania. Acta Trop., 70, 143–55CrossRefGoogle ScholarPubMed
Muir, L. E., Thorne, M. J. and Kay, B. H. (1992) Aedes aegypti (Diptera, Culicidae) vision – spectral sensitivity and other perceptual parameters of the female eye. J. Med. Ent., 29, 278–81CrossRefGoogle ScholarPubMed
Mukabana, W. R., Takken, W. and Knols, B. G. J. (2002a) Analysis of arthropod bloodmeals using molecular genetic markers. Trends in Parasitology, 18, 505–9CrossRefGoogle Scholar
Mukabana, W. R., Takken, W., Seda, P. et al. (2002b) Extent of digestion affects the success of amplifying human DNA from blood meals of Anopheles gambiae (Diptera: Culicidae). Bull. Ent. Res., 92, 233–9CrossRefGoogle Scholar
Mukerji, D. and Sen-Sarma, P. (1955) Anatomy and affinity of the elephant louse, Haematomyzus elephantis Piaget (Insecta: Rhyncophthiraptera). Parasitol., 45, 5–30CrossRefGoogle Scholar
Mukwaya, L. G. (1977) Genetic control of feeding preference in the mosquitoes Aedes (Stegomyia) simpsoni and aegypti. Phys. Ent., 2, 133–45CrossRefGoogle Scholar
Mullens, B. A. and Gerhardt, R. R. (1979) Feeding behaviour of some Tennessee Tabanidae. Environ. Ent., 8, 1047–51CrossRefGoogle Scholar
Muller, H. M., Catteruccia, F., Vizioli, J., DellaTorre, A. and Crisanti, A. (1995) Constitutive and blood meal-induced trypsin genes in Anopheles gambiae. Experimental Parasitology, 81, 371–85CrossRefGoogle ScholarPubMed
Muller, H. M., Crampton, J. M., Dellatorre, A., Sinden, R. and Crisanti, A. (1993) Members of a trypsin gene family in Anopheles gambiae are induced in the gut by blood meal. EMBO J., 12, 2891–900Google ScholarPubMed
Mumcuoglu, Y. and Galun, R. (1987) Engorgement response of human body lice Pediculus humanus (Insecta: Anoplura) to blood fractions and their components. Phys. Ent., 12, 171–4CrossRefGoogle Scholar
Munstermann, L. E. and Conn, J. E. (1997) Systematics of mosquito disease vectors (Diptera, Culicidae): impact of molecular biology and cladistic analysis. Ann. Rev. Ent., 42, 351–69CrossRefGoogle ScholarPubMed
Murlis, J., Willis, M. A. and Carde, R. T. (2000) Spatial and temporal structures of pheromone plumes in fields and forests. Phys. Ent., 25, 211–22CrossRefGoogle Scholar
Murray, M. D. (1957) The distribution of the eggs of mammalian lice on their hosts. II. Analysis of the oviposition behaviour of Damalinia ovis. Aust. J. Zool, 5, 19–29CrossRefGoogle Scholar
Murray, M. D. (1963) Influence of temperature on the reproduction of Damalinia equi (Denny). Aust. J. Zool., 11, 183–9CrossRefGoogle Scholar
Murray, M. D. (1987) Effects of host grooming on louse populations. Parasitology Today, 3, 276–8CrossRefGoogle ScholarPubMed
Murray, M. D. and Nicholls, D. G. (1965) Studies on the ectoparasites of seals and penguins. I. The ecology of the louse Lepidophthirus macrorhini Enderlein on the southern elephant seal, Mirounga leonina (L.). Aust. J. Zool., 13, 437–54CrossRefGoogle Scholar
Mwandawiro, C., Boots, M., Tuno, N. et al. (2000) Heterogeneity in the host preference of Japanese encephalitis vectors in Chiang Mai, northern Thailand. Trans. R. Soc. Trop. Med. Hyg., 94, 238–42CrossRefGoogle Scholar
Naksathit, A. T., Edman, J. D. and Scott, T. W. (1999) Utilization of human blood and sugar as nutrients by female Aedes aegypti (Diptera: Culicidae). J. Med. Ent, 36, 13–17CrossRefGoogle Scholar
Naksathit, A. T. and Scott, T. W. (1998) Effect of female size on fecundity and survivorship of Aedes aegypti fed only human blood versus human blood plus sugar. J. Am. Mosq. Control Assoc., 14, 148–52Google Scholar
Napier Bax, S. (1937) The senses of smell and sight in Glossina swynnertoni. Bull. Ent. Res., 28, 539–82CrossRefGoogle Scholar
Nappi, A. J., Vass, E., Frey, F. and Carton, Y. (2000) Nitric oxide involvement in Drosophila immunity. Nitric Oxide Biology and Chemistry, 4, 423–30CrossRefGoogle ScholarPubMed
Nasci, R. S. (1982) Differences in host choice between the sibling species of treehole mosquitoes Aedes triseriatus and Aedes hendersoni. Am. J. Trop. Med. Hyg., 31, 411–15CrossRefGoogle ScholarPubMed
Nelson, R. L. (1965) Carbon dioxide as an attractant for Culicoides. J. Med. Ent., 2, 56–7CrossRefGoogle ScholarPubMed
Nelson, W. A. (1987) Other blood-sucking and myiasis-producing arthropods. In Soulsby, E. J. L. (ed.), Immune Responses in Parasitic Infections: Immunology, Immunopathology and Immunoprophylaxis.Boca Raton, Florida: CRC Press, Vol. IVGoogle Scholar
Nelson, W. A., Bell, J. F., Clifford, C. M. and Keirans, A. J. (1977) Interaction of ectoparasites and their hosts. J. Med. Ent., 13, 389–428CrossRefGoogle ScholarPubMed
Nelson, W. A., Keirans, J. E., Bell, J. F. and Clifford, C. M. (1975) Host–ectoparasite relationships. J. Med. Ent., 12, 143–66CrossRefGoogle ScholarPubMed
Nelson, W. A. and Kozub, G. C. (1980) Melophagus ovinus (Diptera: Hippoboscidae): evidence of local mediation in acquired resistance of sheep to keds. J. Med. Ent., 17, 291–7CrossRefGoogle Scholar
Newson, R. M. and Holmes, R. G. (1968) Some ectoparasites of the coypu (Myocastor coypus) in eastern England. J. Anim. Ecol., 37, 471–81CrossRefGoogle Scholar
Nguu, E. K., Osir, E. O., Imbuga, M. O. and Olembo, N. K. (1996) The effect of host blood in the in vitro transformation of bloodstream trypanosomes by tsetse midgut homogenates. Med. Vet. Entomol., 10, 317–22CrossRefGoogle ScholarPubMed
Niare, O., Markianos, K., Volz, J. et al. (2002) Genetic loci affecting resistance to human malaria parasites in a west African mosquito vector population. Science, 298, 213–16CrossRefGoogle Scholar
Nieves, E. and Pimenta, P. F. P. (2002) Influence of vertebrate blood meals on the development of Leishmania (Viannia) braziliensis and Leishmania (Leishmania) amazonensis in the sand fly Lutzomyia migonei (Diptera: Psychodidae). Am. J. Trop. Med. Hyg., 67, 640–7CrossRefGoogle Scholar
Nigam, Y. and Ward, R. D. (1991) The effect of male sandfly pheromone and host factors as attractants for female Lutzomyia longipalpis (Diptera, Psychodidae). Phys. Ent., 16, 305–12CrossRefGoogle Scholar
Nogge, G. (1978) Aposymbiotic tsetse flies, Glossina morsitans morsitans obtained by feeding on rabbits immunized specifically with symbionts. J. Insect Physiol., 24, 299–304CrossRefGoogle ScholarPubMed
Nogge, G. (1981) Significance of symbionts for the maintenance of an optimal nutritional state for successful reproduction in haematophagous arthropods. Parasitology, 82, 101–4Google Scholar
Nogge, G. and Ritz, R. (1982) Number of symbionts and its regulation in tsetse flies, Glossina spp. Ent. Exp. Appl., 31, 249–54CrossRefGoogle Scholar
Noriega, F. G., Edgar, K. A., Bechet, R. and Wells, M. A. (2002) Midgut exopeptidase activities in Aedes aegypti are induced by blood feeding. J. Insect Physiol., 48, 205–12CrossRefGoogle ScholarPubMed
Noriega, F. G. and Wells, M. A. (1999) A molecular view of trypsin synthesis in the midgut of Aedes aegypti. J. Insect Physiol., 45, 613–20CrossRefGoogle ScholarPubMed
Nuttal, G. H. F. (1899) On the role of insects, arachnids, and myriapods as carriers in the spread of bacterial and parasitic disease of man and animals. A critical and historical study. Johns Hopkins Hospital Reports, 8, 1–154Google Scholar
Obiamiwe, B. A. and Macdonald, W. W. (1973) 1. The effect of heparin on the migration of Brugia pahangi microfilariae Culex pipiens. 2. The uptake of B. pahangi microfilariae in C. pipiens and the infectivity of C. pipiens in relation to microfilarial densities. 3. Evidence of a sex-linked recessive gene, sb, controlling susceptibility of C. pipiens to B. pahangi. Trans. R. Soc. Trop. Med. Hyg., 67, 32–3CrossRefGoogle Scholar
Ochiai, M., Niki, T. and Ashida, M. (1992) Immunocytochemical localization of beta-1,3-glucan recognition protein in the silkworm, Bombyx mori. Cell and Tissue Research, 268, 431–7CrossRefGoogle ScholarPubMed
Ogston, C. W. and London, W. T. (1980) Excretion of hepatitis B surface antigen by the bedbug Cimex hemipterus Fabr. Trans. R. Soc. Trop. Med. Hyg., 74, 823–5CrossRefGoogle ScholarPubMed
Olubayo, R. O., Mihok, S., Munyoki, E. and Otieno, L. H. (1994) Dynamics of host blood effects in Glossina morsitans spp. infected with Trypanosoma congolense and Trypanosoma brucei. Parasitology Research, 80, 177–81CrossRefGoogle Scholar
Meara, G. F. (1979) Variable expressions of autogeny in three mosquito species. Int. J. Invert. Reprod., 1, 253–61CrossRefGoogle Scholar
O'Meara, G. F. (1985) Ecology and autogeny in mosquitoes. In Lounibos, L. P., Rey, J. R. and Frank, J. H.,(eds.), Ecology of Mosquitoes. Florida: Florida Medical LaboratoryGoogle Scholar
O'Meara, G. F. (1987) Nutritional ecology of blood feeding diptera. In Slansky, F. and Rodriguez, J. G. (eds.),Nutritional Ecology of Insects, Mites, Spiders and Related Invertebrates. New York: WileyGoogle Scholar
Meara, G. F. and Edman, J. D. (1975) Autogenous egg production in the salt marsh mosquito, Aedes taeniorrhynchus. Biol. Bull., 149, 384–96CrossRefGoogle Scholar
Meara, G. F. and Evans, D. G. (1973) Blood-feeding requirements of the mosquito: geographical variation in Aedes taeniorhynchus. Science, 180, 1291–3CrossRefGoogle Scholar
Meara, G. F. and Evans, D. G. (1976) The influence of mating on autogenous egg development in the mosquito, Aedes taeniorrhynchus. J. Insect Physiol., 22, 613–17CrossRefGoogle Scholar
Omer, S. M. and Gillies, M. T. (1971) Loss of response to carbon dioxide in palpectomized female mosquitoes. Ent. Exp. Appl., 14, 251–2Google Scholar
Osbrink, L. A. and Rust, M. A. (1985) Cat flea (Siphonaptera: Pulicidae): factors influencing host-finding behaviour in the laboratory. Ann. Ent. Soc. Am., 78, 29–34CrossRefGoogle Scholar
Shea, B., Rebollar-Tellez, E., Ward, R. D. et al. (2002) Enhanced sandfly attraction to Leishmania-infected hosts. Trans. R. Soc. Trop. Med. Hyg., 96, 117–18CrossRefGoogle Scholar
Overal, W. L. (1980) Biology and behaviour of North American Trichobius bat flies (Diptera: Streblidae). Ph.D. thesis, University of Kansas
Overal, W. L. and Wingate, L. R. (1976) The biology of the batbug Strictimex antennatus (Hemiptera: Cimicidae) in South Africa. Ann. Natal Mus., 22, 821–8Google Scholar
Owaga, M. L. and Challier, A. (1985) Catch composition of the tsetse Glossina pallidipes Austen in revolving and stationary traps with respect to age, sex ratio and hunger stage. Insect Sci. Applic., 6, 711–18Google Scholar
Page, R. D. M., Clayton, D. H. and Paterson, A. M. (1996) Lice and cospeciation: a response to Barker. Int. J. Parasit., 26, 213–18CrossRefGoogle ScholarPubMed
Pagel, M. and Bodmer, W. (2003) A naked ape would have fewer parasites. Proc. R. Soc. Lond. B Biol. Sci., 270, Suppl 1, 117–19CrossRefGoogle ScholarPubMed
Pant, C. P., Houba, V. and Engers, H. D. (1987) Bloodmeal identification in vectors. Parasitology Today, 3, 324–6Google Scholar
Panton, L. J., Tesh, R. B., Nadeau, K. C. and Beverley, S. M. (1991) A test for genetic exchange in mixed infections of Leishmania major in the sand fly Phlebotomus papatasi. J. Protozool, 38, 224–8CrossRefGoogle ScholarPubMed
Pappas, L. G., Pappas, C. D. and Grossman, G. L. (1986) Hemodynamics of human skin during mosquito (Diptera: Culicidae) blood feeding. J. Med. Ent., 23, 581–7CrossRefGoogle ScholarPubMed
Parker, K. R. and Gooding, R. H. (1979) Effects of host anaemia, local skin factors and circulating antibodies upon biology of laboratory reared Glossina morsitans morsitans (Diptera: Glossinidae). Can. J. Zool., 57, 2393–401CrossRefGoogle Scholar
Paskewitz, S. M., Brown, M. R., Lea, A. O. and Collins, F. H. (1988) Ultrastructure of the encapsulation of Plasmodium cynomolgi (B-strain) on the midgut of a refractory strain of Anopheles gambiae. Journal of Parasitology, 74, 432–9CrossRefGoogle ScholarPubMed
Patton, W. S. and Craig, F. W. (1913) On certain haematophagous species of the genus Musca, with descriptions of two new species. Indian Journal of Medical Research, 1, 13–25Google Scholar
Peacock, A. J. (1981) Distribution of (Na+ K+)-ATPase activity in the mid-guts and hind-guts of adult Glossina morsitans and Sarcophaga nodosa and the hind-gut of Bombyx mori larvae. Comp. Biochem. Physiol. A, 69, 133–6CrossRefGoogle Scholar
Peacock, A. J. (1982) Effects of sodium transport inhibitors on diuresis and midgut (Na+ and K+) ATPase in the tsetse fly Glossina morsitans. J. Insect Physiol., 28, 553–8CrossRefGoogle Scholar
Pearman, J. V. (1960) Some African psocoptera found on rats. Entomologist, 93, 246–50Google Scholar
Pearson, T. W., Beecroft, R. P., Welburn, S. C. et al. (2000) The major cell surface glycoprotein procyclin is a receptor for induction of a novel form of cell death in African trypanosomes in vitro. Molecular and Biochemical Parasitology, 111, 333–49CrossRefGoogle ScholarPubMed
Pell, P. E. and Southern, D. I. (1976) Effect of the coccidiostat, sulphaquinoxline, on symbiosis in the tsetse fly, Glossina species. Microbios Letters, 2, 203–11Google Scholar
Pereira, H., Penido, C. M., Martins, M. S. and Diotaiuti, L. (1998) Comparative kinetics of bloodmeal intake by Triatoma infestans and Rhodnius prolixus, the two principal vectors of Chagas disease. Med. Vet. Entomol., 12, 84–8CrossRefGoogle ScholarPubMed
Pereira, M. E. A., Andrade, A. F. B. and Ribeiro, J. M. C. (1981) Lectins of distinct specificity in Rhodnius prolixus interact selectively with Trypanosoma cruzi. Science, 211, 597–9CrossRefGoogle ScholarPubMed
Perrin, N., Christe, P. and Richner, H. (1996) On host life-history response to parasitism. Oikos, 75, 317–20CrossRefGoogle Scholar
Peschken, D. P. and Thorsteinson, A. J. (1965) Visual orientation of black flies (Simuliidae: Diptera) to colour, shape and movement of targets. Ent. Exp. Appl., 8, 282–8CrossRefGoogle Scholar
Peters, W. (1968) Vorkommen, Zusammensetzung und feinstruktur peritrophischer membranen im tierreich. Zeit. Morph. Okol. Tiere., 64, 21–58CrossRefGoogle Scholar
Peters, W., Kolb, H. and Kolb-Bachofen, V. (1983) Evidence for a sugar receptor (lectin) in the peritrophic membrane of the blowfly larva, Calliphora erythrocephala Mg. (Diptera). J. Insect Physiol., 29, 275–80CrossRefGoogle Scholar
Peters, W., Zimmermann, U. and Becker, B. (1973) Investigations on the transport function and structure of peritrophic membranes. IV. Anisotropic cross bands in peritrophic membranes of Diptera. J. Insect Physiol., 19, 1067–77CrossRefGoogle Scholar
Peterson, D. G. and Brown, A. W. A. (1951) Studies of the responses of female Aedes mosquito. III. The response of Aedes aegypti (L.) to a warm body and its radiation. Bull. Ent. Res., 42, 535–41CrossRefGoogle Scholar
Phelps, R. J. and Vale, G. A. (1976) Studies on the local distribution and on the methods of host location of some Rhodesian Tabanidae (Diptera). J. Ent. Soc. S. Afr., 39, 67–81Google Scholar
Pichon, G., Awono-Ambene, H. P. and Robert, V. (2000) High heterogeneity in the number of Plasmodium falciparum gametocytes in the bloodmeal of mosquitoes fed on the same host. Parasitology, 121, 115–20CrossRefGoogle ScholarPubMed
Piot, P. and Schofield, C. J. (1986) No evidence for arthropod transmission of AIDS. Parasitology Today, 2, 294–5CrossRefGoogle Scholar
Platt, K. B., Linthicum, K. J., Myint, K. S. A. et al. (1997) Impact of dengue virus infection on feeding behavior of Aedes aegypti. Am. J. Trop. Med. Hyg., 57, 119–25CrossRefGoogle ScholarPubMed
Politzar, H. and Merot, P. (1984) Attraction of the tsetse fly Glossina morsitans submorsitans to acetone, 1 octen-3-ol, and the combination of these compounds in west Africa. Rev. Elev. Med. Vet. Pays Trop., 37, 468–73Google ScholarPubMed
Ponnudurai, T., Billingsley, P. F. and Rudin, W. (1988) Differential infectivity of Plasmodium for mosquitoes. Parasitology Today, 4, 319–21CrossRefGoogle ScholarPubMed
Port, G. R., Bateham, P. F. L. and Bryan, J. H. (1980) The relationship of host size to feeding by mosquitoes of the Anopheles gambiae complex (Diptera: Culicidae). Bull. Ent. Res., 70, 133–44CrossRefGoogle Scholar
Pospisil, J. and Zdarek, J. (1965) On the visual orientation of the stable fly (Stomoxys calcitrans L.) to colours. Acta Entomol. Bohemoslov., 62, 85–91Google Scholar
Powell, J. R., Petrarca, V., Torre, A., Caccone, A. and Coluzzi, M. (1999) Population structure, speciation, and introgression in the Anopheles gambiae complex. Parasitologia, 41, 101–13Google ScholarPubMed
Price, G. D., Smith, N. and Carlson, D. A. (1979) The attraction of female mosquitoes (Anopheles quadrimaculatus Say) to stored human emanations in conjunction with adjusted levels of relative humidity, temperature and carbon dioxide. J. Chem. Ecol., 5, 383–95CrossRefGoogle Scholar
Price, R. D. (1975) The Menacanthus eurysternus complex (Mallophaga: Menoponidae) of the Passeriformes and Piciformes (Aves). Ann. Ent. Soc. Am., 68, 617–22CrossRefGoogle Scholar
Prior, A. and Torr, S. J. (2002) Host selection by Anopheles arabiensis and An. quadriannulatus feeding on cattle in Zimbabwe. Med. Vet. Entomol., 16, 207–13CrossRefGoogle ScholarPubMed
Raikhel, A. S., Kokoza, V. A., Zhu, J. et al. (2002) Molecular biology of mosquito vitellogenesis: from basic studies to genetic engineering of antipathogen immunity. Insect Biochem. Mol. Biol., 32, 1275–86CrossRefGoogle ScholarPubMed
Ramet, M., Lanot, R., Zachary, D. and Manfruelli, P. (2002a) JNK signaling pathway is required for efficient wound healing in Drosophila. Dev. Biol., 241, 145–56CrossRefGoogle Scholar
Ramet, M., Manfruelli, P., Pearson, A., Mathey-Prevot, B. and Ezekowitz, R. A. B. (2002b) Functional genomic analysis of phagocytosis and identification of a Drosophila receptor for E-coli. Nature, 416, 644–8CrossRefGoogle Scholar
Ratcliffe, N. A. and Rowley, A. F. (1979) Role of hemocytes in defence against biological agents. In Gupta, A. P. (ed.),Insect Hemocytes. Cambridge: Cambridge University Press, 331–422CrossRefGoogle Scholar
Ratzlaff, R. E. and Wikel, S. K. (1990) Murine immune responses and immunization against Polyplax serrata (Anoplura: Polyplacidae). J. Med. Ent., 27, 1002–7CrossRefGoogle Scholar
Ready, P. D. (1978) The feeding habits of laboratory bred Lutzomyia longipalpis (Diptera: Psychodidae). J. Med. Ent., 14, 545–552CrossRefGoogle Scholar
Reddy, V. B., Kounga, K., Mariano, F. and Lerner, E. A. (2000) Chrysoptin is a potent glycoprotein IIb/IIIa fibrinogen receptor antagonist present in salivary gland extracts of the deerfly. Journal of Biological Chemistry, 275, 15861–7CrossRefGoogle ScholarPubMed
Read, W., Carrall, J., Agramonte, A. and Lazear, J. (1900) The etiology of yellow fever: a preliminary note. The Philadelphia Medical Journal, 6, 790–3
Reichardt, T. R. and Galloway, T. D. (1994) Seasonal occurrence and reproductive status of Opisocrostis bruneri (Siphonaptera, Ceratophyllidae), a flea on franklin ground-squirrel, Spermophilus franklinii (Rodentia, Sciuridae) near Birds Hill Park, Manitoba. J. Med. Ent., 31, 105–13CrossRefGoogle Scholar
Reid, G. D. F. and Lehane, M. J. (1984) Peritrophic membrane formation in three temperate simuliids, Simulium ornatum, S. equinum and S. lineatum with respect to the migration of Onchocercal microfilariae. Ann. Trop. Med. Parasit., 78, 527–39CrossRefGoogle ScholarPubMed
Reinouts van Haga, H. A. and Mitchell, B. K. (1975) Temperature receptors on tarsi of the tsetse fly Glossina morsitans West. Nature, 255, 225–6CrossRefGoogle ScholarPubMed
Reunala, T., Brummer-Korvenkontio, H., Lappalainen, P., Rasanen, L. and Palosuo, T. (1990) Immunology and treatment of mosquito bites. Clin. Exp. Allergy, 20, Suppl 4, 19–24CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. (1982) The anti-serotonin and antihistamine activities of salivary secretion of Rhodnius prolixus. J. Insect Physiol., 28, 69–75CrossRefGoogle Scholar
Ribeiro, J. M. C. (1987) Role of saliva in blood-feeding by arthropods. Ann. Rev. Ent., 32, 463–78CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. (1995) Blood-feeding arthropods – live syringes or invertebrate pharmacologists. Infectious Agents and Disease – Reviews Issues and Commentary, 4, 143–52Google ScholarPubMed
Ribeiro, J. M. C. (1998) Rhodnius prolixus salivary nitrophorins display heme-peroxidase activity. Insect Biochemistry and Molecular Biology, 28, 1051–7CrossRefGoogle Scholar
Ribeiro, J. M. C., Charlab, R., Rowton, E. D. and Cupp, E. W. (2000) Simulium vittatum (Diptera: Simuliidae) and Lutzomyia longipalpis (Diptera: Psychodidae) salivary gland hyaluronidase activity. J. Med. Ent., 37, 743–7CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C., Charlab, R. and Valenzuela, J. G. (2001) The salivary adenosine deaminase activity of the mosquitoes Culex quinquefasciatus and Aedes aegypti. J. Exp. Biol., 204, 2001–10Google ScholarPubMed
Ribeiro, J. M. C. and Francischetti, I. M. B. (2003) Role of arthropod saliva in blood feeding: sialome and post-sialome perspectives. Ann. Rev. Ent., Vol. 48, 73–88CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. and Garcia, E. S. (1981a) Platelet antiaggregating activity in the salivary secretion of the blood-sucking bug Rhodnius prolixus. Experientia, 37, 384–5CrossRefGoogle Scholar
Ribeiro, J. M. C. and Garcia, E. S. (1981b) The role of saliva in feeding in Rhodnius prolixus. J. Exp. Biol., 94, 219–30Google Scholar
Ribeiro, J. M. C., Katz, O., Pannell, L. K., Waitumbi, J. and Warburg, A. (1999) Salivary glands of the sand fly Phlebotomus papatasi contain pharmacologically active amounts of adenosine and 5-AMP. J. Exp. Biol., 202, 1551–9Google ScholarPubMed
Ribeiro, J. M. C., Rossignol, P. A. and Spielman, A. (1985) Salivary gland apyrase determines probing time in anopheline mosquitoes. J. Insect Physiol., 9, 689–92CrossRefGoogle Scholar
Ribeiro, J. M. C., Schneider, M. and Guimaraes, J. A. (1995) Purification and characterization of prolixin-S (nitrophorin-2), the salivary anticoagulant of the bloodsucking bug Rhodnius prolixus. Biochemical Journal, 308, 243–9CrossRefGoogle Scholar
Ribeiro, J. M. C. and Valenzuela, J. G. (1999) Purification and cloning of the salivary peroxidase/catechol oxidase of the mosquito Anopheles albimanus. J. Exp. Biol., 202, 809–16Google ScholarPubMed
Rice, M. J., Galun, R. and Margalit, J. (1973) Mouthpart sensilla of the tsetse fly and their function. II. Labial sensilla. Ann. Trop. Med. Parasit., 67, 101–7CrossRefGoogle ScholarPubMed
Richman, A. M., Dimopoulos, G., Seeley, D. and Kafatos, F. C. (1997) Plasmodium activates the innate immune response of Anopheles gambiae mosquitoes. EMBO J., 16, 6114–9CrossRefGoogle ScholarPubMed
Roberts, L. W. (1981) Probing by Glossina morsitans morsitans and transmission of Trypanosoma (Nannomonas) congolense. Am. J. Trop. Med. Hyg., 30, 948–51CrossRefGoogle ScholarPubMed
Roberts, R. H. (1972) Relative attractiveness of CO2 and a steer to Tabanidae, Culicidae, and Stomoxys calcitrans. Mosq. News, 32, 208–11Google Scholar
Roberts, R. H. (1977) Attractancy of two black decoys and CO2 to tabanids (Diptera: Tabanidae). Mosq. News, 37, 169–72Google Scholar
Robinson, A. (1939) The mouthparts and their function in the female mosquito, Anopheles maculipennis. Parasitol., 31, 212–42CrossRefGoogle Scholar
Rogers, K. A. and Titus, R. G. (2003) Immunomodulatory effects of Maxadilan and Phlebotomus papatasi sand fly salivary gland lysates on human primary in vitro immune responses. Parasite Immunol., 25, 127–34CrossRefGoogle ScholarPubMed
Rogers, M. E., Chance, M. L. and Bates, P. A. (2002) The role of promastigote secretory gel in the origin and transmission of the infective stage of Leishmania mexicana by the sandfly Lutzomyia longipalpis. Parasitology, 124, 495–507CrossRefGoogle ScholarPubMed
Rosenfeld, A. and Vanderberg, J. P. (1998) Identification of electrophoretically separated proteases from midgut and hemolymph of adult Anopheles stephensi mosquitoes. Journal of Parasitology, 84, 361–5CrossRefGoogle ScholarPubMed
Ross, R. (1897) On same peculier pigmented cells found in two mosquitoes fed on malaria blood. British Medical Journal, 2, 1786–8CrossRefGoogle Scholar
Ross, R. (1898) Report on the cultivation of protessoma, Labb, in grey mosquitoes. Indian Med. Gaz., 33, 401–8Google Scholar
Rossignol, P. A., Ribeiro, J. M. C., Jungery, M., Turell, M. J., Spielman, A. and Bailey, C. L. (1985) Enhanced mosquito blood-finding success on parasitaemic hosts: Evidence for vector-parasite mutualism. Proc. Nat. Acad. Sci., 82, 7725–7CrossRefGoogle Scholar
Rossignol, P. A., Ribeiro, J. M. C. and Spielman, A. (1984) Increased intradermal probing time in sporozoite-infected mosquitoes. Am. J. Trop. Med. Hyg., 33, 17–20CrossRefGoogle ScholarPubMed
Rossignol, P. A., Ribeiro, J. M. C. and Spielman, A. (1986) Increased biting rate and reduced fertility in sporozoite-infected mosquitoes. Am. J. Trop. Med. Hyg., 35, 277–9CrossRefGoogle ScholarPubMed
Rossignol, P. A. and Rossignol, A. M. (1988) Simulations of enhanced malaria transmission and host bias induced by modified vector blood location behaviour. Parasitol., 97, 363–72CrossRefGoogle ScholarPubMed
Rothschild, M. (1975) Recent advances in our knowledge of the Siphonoptera. Ann. Rev. Ent., 20, 241–59CrossRefGoogle Scholar
Rothschild, M. and Clay, T. (1952) Fleas, Flukes and Cuckoos. New York: Philosophical Library
Rothschild, M. and Ford, B. (1973) Factors influencing the breeding of the rabbit flea (Spilopsyllus cuniculi): a spring-time accelerator and a kairomone in nestling rabbit urine (with notes on Cediopsylla simplex, another ‘hormone bound’ species). J. Zool., 170, 87–137CrossRefGoogle Scholar
Rothschild, M., Schlein, Y., Parker, K. and Sternberg, S. (1972) Jump of the oriental rat flea Xenopsylla cheopis (Roths.). Nature, 239, 45–8CrossRefGoogle Scholar
Rowland, M. and Boersma, E. (1988) Changes in the spontaneous flight activity of the mosquito Anopheles stephensi by parasitization with the rodent malaria Plasmodium yoelii. Parasitology, 97, 221–7CrossRefGoogle ScholarPubMed
Rowland, M. W. and Lindsay, S. L. (1986) The circadian flight activity of Aedes aegypti parasitized with the filarial nematode Brugia pahangi. Phys. Ent., 11, 325–34CrossRefGoogle Scholar
Roy, D. N. (1936) On the role of blood in ovulation in Aedes aegypti, Linn. Bull. Ent. Res., 27, 423–9CrossRefGoogle Scholar
Royet, J., Meister, M. and Ferrandon, D. (2003) Humoral and cellular responses in Drosophilainnate immunity. In Ezekowitz, R. A. and Hoffman, J. A. (eds.), Infectious Disease: Innate Immunity. Totowa, NJ: Humana Press 137–53Google ScholarPubMed
Rubenstein, D. I. and Hohmann, M. E. (1989) Parasites and social-behavior of island feral horses. Oikos, 55, 312–20CrossRefGoogle Scholar
Rudin, W. and Hecker, H. (1979) Functional morphology of the midgut of Aedes aegypti L. (Insecta; Diptera) during blood digestion. Cell, 200, 193–203Google ScholarPubMed
Rudin, W. and Hecker, H. (1982) Functional morphology of the midgut of a sandfly as compared to other haematophagous nematocera. Tissue and Cell, 14, 751–8CrossRefGoogle Scholar
Rutberg, A. T. (1987) Horse fly harassment and the social-behavior of feral ponies. Ethology, 75, 145–54CrossRefGoogle Scholar
Sabelis, M. W. and Schippers, P. (1984) Variable wind direction and anemotactic strategies of searching for an odour plume. Oecologia, 63, 225–8CrossRefGoogle ScholarPubMed
Sacks, D. L. (1989) Metacyclogenesis in Leishmania promastigotes. Exp. Parasitol., 69, 100–3CrossRefGoogle ScholarPubMed
Sacks, D. L. and Kamhawi, S. (2001) Molecular aspects of parasite–vector and vector–host interactions in Leishmaniasis. Ann. Rev. Microbiol., 55, 453–83CrossRefGoogle ScholarPubMed
Sallum, M. A. M., Schultz, T. R., Foster, P. G. et al. (2002) Phylogeny of Anophelinae (Diptera: Culicidae) based on nuclear ribosomal and mitochondrial DNA sequences. Systematic Entomology, 27, 361–82CrossRefGoogle Scholar
Samuel, W. M. and Trainer, D. O. (1972) Lipoptena mazamae Rondani, 1878 (Diptera: Hippoboscidae) on white-tailed deer in southern Texas. J. Med. Ent., 9, 104–6CrossRefGoogle ScholarPubMed
Sandeman, R. M. (1996) Immune rsponses to mosquitoes and flies. In Wikel, S. K. (ed.), The Immunology of Host–Ectoparasitic Arthropod Interactions. Wallingford: CAB International 175–203Google Scholar
Sangiorgi, G. and Frosini, D. (1940) Di un principio emolitico (‘Cimicina’) nella saliva del Cimex lectularius. Pathologica, 32, 189–91Google Scholar
Saraiva, E. M., Pimenta, P. F., Brodin, T. N. (1995) Changes in lipophosphoglycan and gene expression associated with the development of Leishmania major in Phlebotomus papatasi. Parasitology, 111, (Pt 3), 275–87CrossRefGoogle ScholarPubMed
Sarkis, J. J. F., Guimaraes, J. A. and Ribeiro, J. M. C. (1986) Salivary apyrase of Rhodnius prolixus: kinetics and purification. Biochem. J., 233, 885–91CrossRefGoogle ScholarPubMed
Scaraffia, P. Y. and Wells, M. A. (2003) Proline can be utilized as an energy substrate during flight of Aedes aegypti females. J. Insect Physiol., 49, 591–601CrossRefGoogle ScholarPubMed
Schall, J. J. (2002) Parasite virulence. In Lewis, E. E., Campbell, J. F. and Sukdheo, M. D. K. (eds.), The Behavioural Ecology of Parasites. Wallingford: CAB InternationalCrossRefGoogle Scholar
Schiefer, B. A. et al. (1977) Plasmodium cynomolgi: effects of malaria infection on laboratory flight performance of Anopheles stephensi mosquiotoes. Exp. Parasitol., 41(2), 397–404CrossRef
Schlein, Y. (1977) Lethal effect of tetracycline on tsetse flies following damage to bacteroid symbionts. Experimentia, 33, 450–1CrossRefGoogle Scholar
Schlein, Y. and Jacobson, R. L. (1998) Resistance of Phlebotomus papatasi to infection with Leishmania donovani is modulated by components of the infective bloodmeal. Parasitology, 117, 467–73CrossRefGoogle ScholarPubMed
Schlein, Y., Warburg, A., Schnur, L. F. and Shlomai, J. (1983) Vector compatibility of Phlebotomus papatasi dependent on differentially induced digestion. Acta Trop., 40, 65–70Google ScholarPubMed
Schlein, Y., Yuval, B. and Warburg, A. (1984) Aggregation pheromone released from the palps of feeding female Phlebotomus papatasi (Psychodidae). J. Insect Physiol., 30, 153–6CrossRefGoogle Scholar
Schmid-Hempel, P. (2003) Immunology and evolution of infectious disease. Science, 300, 254CrossRefGoogle Scholar
Schmitz, H., Trenner, S., Hofmann, M. H. and Bleckmann, H. (2000) The ability of Rhodnius prolixus (Hemiptera; Reduviidae) to approach a thermal source solely by its infrared radiation. J. Insect Physiol., 46, 745–51CrossRefGoogle ScholarPubMed
Schoeler, G. B. and Wikel, S. K. (2001) Modulation of host immunity by haematophagous arthropods. Annals of Tropical Medicine and Parasitology, 95, 755–71CrossRefGoogle ScholarPubMed
Schofield, C. J. (1981) Chagas disease, triatomine bugs, and blood loss. Lancet, 1, 1316Google Scholar
Schofield, C. J. (1982) The role of blood intake in density regulation of populations of Triatoma infestans (Klug) (Hemiptera: Reduviidae). Bull. Ent. Res., 72, 617–29CrossRefGoogle Scholar
Schofield, C. J. (1985) Population dynamics and control of Triatoma infestans. Ann. Soc. Belge, Med. Trop., 65, 149–64Google ScholarPubMed
Schofield, C. J. (1988) Biosystematics of the triatominae. In Service, M. W. (ed.), Biosystematics of Haematophagous Insects. Oxford: Clarendon PressGoogle Scholar
Schofield, S. and Sutcliffe, J. F. (1996) Human individuals vary in attractiveness for host-seeking black flies (Diptera: Simuliidae) based on exhaled carbon dioxide. J. Med. Ent., 33, 102–8CrossRefGoogle ScholarPubMed
Schofield, S. and Sutcliffe, J. F. (1997) Humans vary in their ability to elicit biting responses from Simulium venustum (Diptera: Simuliidae). J. Med. Ent., 34, 64–7CrossRefGoogle Scholar
Schofield, S. and Torr, S. J. (2002) A comparison of the feeding behaviour of tsetse and stable flies. Med. Vet. Entomol., 16, 177–85CrossRefGoogle ScholarPubMed
Senghor, J. E. and Samba, E. M. (1988) Onchocerciasis control program – the human perspective. Parasitology Today, 4, 332–3CrossRefGoogle Scholar
Severson, D. W., Brown, S. E. and Knudson, D. L. (2001) Genetic and physical mapping in mosquitoes: molecular approaches. Ann. Rev. Ent., 46, 183–219CrossRefGoogle ScholarPubMed
Severson, D. W., Mori, A., Zhang, Y. and Christensen, B. M. (1994) Chromosomal mapping of two loci affecting filarial worm susceptibility in Aedes aegypti. Insect Mol. Biol., 3, 67–72CrossRefGoogle ScholarPubMed
Severson, D. W., Thathy, V., Mori, A., Zhang, Y. and Christensen, B. M. (1995) Restriction-fragment-length-polymorphism mapping of quantitative trait loci for malaria parasite susceptibility in the mosquito Aedes aegypti. Genetics, 139, 1711–17Google ScholarPubMed
Shahabuddin, M. (1998) Plasmodium ookinete development in the mosquito midgut: a case of reciprocal manipulation. Parasitology, 116, S83–S93CrossRefGoogle ScholarPubMed
Shahabuddin, M., Fields, I., Bulet, P., Hoffmann, J. A. and Miller, L. H. (1998) Plasmodium gallinaceum: differential killing of some mosquito stages of the parasite by insect defensin. Experimental Parasitology, 89, 103–12CrossRefGoogle ScholarPubMed
Shahan, M. S. and Giltner, L. T. (1945) A review of the epizootiology of equine encephalomyelitis in the United States. J. Am. Vet. Med. Assoc., 107, 279–88Google ScholarPubMed
Shin, S. W., Kokoza, V., Lobkov, I. and Raikhel, A. S. (2003) Relish-mediated immune deficiency in the transgenic mosquito Aedes aegypti. Proc. Natl. Acad. Sci. USA, 100, 2616–21CrossRefGoogle ScholarPubMed
Sieber, K. P., Huber, M., Kaslow, D. et al. (1991) The peritrophic membrane as a barrier. Its penetration by Plasmodium gallinaceum and the effect of a monoclonal antibody to ookinetes. Experimental Parasitology, 72, 145–56CrossRefGoogle ScholarPubMed
Silva, C. P., Ribeiro, A. F., Gulbenkian, S. and Terra, W. R. (1995) Organization, origin and function of the outer microvillar (perimicrovillar) membranes of Dysdercus peruvianus (Hemiptera) midgut cells. J. Insect Physiol., 41, 1093–103CrossRefGoogle Scholar
Silverman, N., Zhou, R., Stoven, S., Pandey, N., Hultmark, D. and Maniatis, T. (2000) A Drosophila IkappaB kinase complex required for Relish cleavage and antibacterial immunity. Genes Dev., 14, 2461–71CrossRefGoogle ScholarPubMed
Simond, P. L. (1898) La propagation de la peste. Annales de la Institut Pasteur, 12, 625
Sippel, W. L. and Brown, A. W. A. (1953) Studies on the responses of the female Aedes mosquito. Part V. The role of visual factors. Bull. Ent. Res., 43, 567–74CrossRefGoogle Scholar
Smit, F. G. A. M. (1972) On some adaptive structures in Siphonaptera. Folia Parasit., 19, 5–17Google ScholarPubMed
Smith, C. N., Smith, N., Gouck, H. K. et al. (1970) L-lactic acid as a factor in the attraction of Aedes aegypti to human hosts. Ann. Ent. Soc. Am., 63, 760–70CrossRefGoogle ScholarPubMed
Smith, H. V. and Titchener, R. N. (1980) Mouthparts of ectoparasites and host damage. Proc. R. Soc. Edin. B-Biol. Sci., 79, 139CrossRefGoogle Scholar
Smith, J. J. B. (1979) Effect of diet viscosity on the operation of the pharyngeal pump in the blood-feeding bug Rhodnius prolixus. J. Exp. Biol., 82, 93–104Google ScholarPubMed
Smith, J. J. B. (1984) Feeding mechanisms. In Kerkut, G. A. and Gilbert, L. I. (eds.), Comprehensive Insect Physiology, Biochemistry and Pharmacology. Oxford: PergamonGoogle Scholar
Smith, J. J. B. and Friend, W. G. (1970) Feeding in Rhodnius prolixus: responses to artificial diets as revealed by changes in electrical resistance. J. Insect Physiol., 16, 1709–20CrossRefGoogle Scholar
Smith, J. J. B. and Friend, W. G. (1982) Feeding behaviour in response to blood fractions and chemical phagostimulants in the blackfly, Simulium venustum. Phys. Ent., 7, 219–26CrossRefGoogle Scholar
Smith, K. G. V. (ed.) (1973) Insects and Other Arthropods of Medical Importance. London: British Museum (Natural History)Google Scholar
Smith, T. and Kilbourne, F. L. (1893) Investigations into the nature, causation and prevention of Texas or Southern cattle fever. U.S. Dept. Agric. Bur. Animal. Indust. Bull., Vol. 1, 301
Snodgrass, R. E. (1944) The anatomy of the Mallophaga. Occ. Pap. Calif. Acad. Sci., 6, 145–229Google Scholar
Soares, M. B., Titus, R. G., Shoemaker, C. B., David, J. R. and Bozza, M. (1998) The vasoactive peptide maxadilan from sand fly saliva inhibits TNF-alpha and induces IL-6 by mouse macrophages through interaction with the pituitary adenylate cyclase-activating polypeptide (PACAP) receptor. J. Immunol., 160, 1811–16Google ScholarPubMed
Soderhall, K. and Cerenius, L. (1998) Role of the prophenoloxidase-activating system in invertebrate immunity. Current Opinion in Immunology, 10, 23–8CrossRefGoogle ScholarPubMed
Soldatos, A. N., Metheniti, A., Mamali, I., Lambropoulou, M. and Marmaras, V. J. (2003) Distinct LPS-induced signals regulate LPS uptake and morphological changes in medfly hemocytes. Insect Biochem. Mol. Biol., 33, 1075–84CrossRefGoogle ScholarPubMed
Sorci, G., Fraipont, M. and Clobert, J. (1997) Host density and ectoparasite avoidance in the common lizard (Lacerta vivipara). Oecologia, 111, 183–8CrossRefGoogle Scholar
Soulsby, E. J. L. (1982) Helminths, Arthropods and Protozoa of Domesticated Animals. London: Bailliere Tindall
Southwood, T. R. E., Khalaf, S. and Sinden, R. E. (1975) The micro-organisms of tsetse flies. Acta Trop., 32, 259–66Google ScholarPubMed
Southworth, G. C., Mason, G. and Seed, J. R. (1968) Studies in frog trypanosomiasis. I. A 24-hour cycle in the parasitaemia level of Trypanosoma rotatorium in Rana clamitans from Louisiana. J. Parasit., 54, 255–8CrossRefGoogle Scholar
Spates, G. E. (1981) Proteolytic and haemolytic activity in the midgut of the stablefly Stomoxys calcitrans (L.): partial purification of the haemolysin. Insect Biochem., 11, 143–7CrossRefGoogle Scholar
Spates, G. E., Stipanovic, R. D., Williams, H. and Holman, G. M. (1982) Mechanisms of haemolysis in a blood-sucking dipteran, Stomoxys calcitrans. Insect Biochem., 12, 707–12CrossRefGoogle Scholar
Spindler, K. (2001) The man in the ice under special consideration of paleo-pathological evidence [in German]. Verhandlungen der Deutschen Gesellschaft für Pathologie, 85, 229–36Google Scholar
Stange, G. (1981) The ocellar component of flight equilibrium control in dragonflies. J. Comp. Physiol., 141, 335–47CrossRefGoogle Scholar
Stanko, M., Miklisova, D., Bellocq, J. G. and Morand, S. (2002) Mammal density and patterns of ectoparasite species richness and abundance. Oecologia, 131, 289–95CrossRefGoogle ScholarPubMed
Stark, K. R. and James, A. A. (1995) A factor Xa-directed anticoagulant from the salivary glands of the yellow fever mosquito Aedes aegypti. Experimental Parasitology, 81, 321–31CrossRefGoogle ScholarPubMed
Steelman, C. D. (1976) Effects of external and internal arthropod parasites on domestic livestock production. Ann. Rev. Ent, 21, 155–78CrossRefGoogle ScholarPubMed
Stevens, J. R., Noyes, H. A., Schofield, C. J. and Gibson, W. (2001) The molecular evolution of Trypanosomatidae. Advances in Parasitology, Vol. 48, 1–56CrossRefGoogle ScholarPubMed
Stierhof, Y. D., Bates, P. A., Jacobson, R. L. (1999) Filamentous proteophosphoglycan secreted by Leishmania promastigotes forms gel-like three-dimensional networks that obstruct the digestive tract of infected sandfly vectors. European Journal of Cell Biology, 78, 675–89CrossRefGoogle ScholarPubMed
Stojanovich, C. J. (1945) The head and mouthparts of the sucking lice (Insecta: Anoplura). Microentomology, 10, 1–46Google Scholar
Stoven, S., Silverman, N., Junell, A. et al. (2003) Caspase-mediated processing of the Drosophila NF-kappaB factor Relish. Proc. Natl. Acad. Sci. USA, 100, 5991–6CrossRefGoogle ScholarPubMed
Strand, M. R. and Clark, K. D. (1999) Plasmatocyte spreading peptide induces spreading of plasmatocytes but represses spreading of granulocytes. Arch. Insect Biochem. Physiol., 42, 213–233.0.CO;2-4>CrossRefGoogle ScholarPubMed
Strand, M. R. and Pech, L. L. (1995) Immunological basis for compatibility in parasitoid host relationships. Ann. Rev. Ent., 40, 31–56CrossRefGoogle ScholarPubMed
Stys, P. and Daniel, M. (1957) Lyctocoris compestris F. (Heteroptera: Anthocoridae) as a human facultative ectoparasite. Acta Societatis Entomologicae Cechoslovenicae, 54, 1–10Google Scholar
Sutcliffe, J. F. (1986) Black fly host location: a review, Can J. Zool, 64(4), 1041–53
Sutcliffe, J. F. (1987) Distance orientation of biting flies to their hosts. Insect Sci. Applic., 8, 611–16Google Scholar
Sutcliffe, J. F. and McIver, S. B. (1975) Artificial feeding of simuliids (Simulium venustum), factors associated with probing and gorging. Experientia, 31, 694–5CrossRefGoogle Scholar
Sutcliffe, J. F., Steer, D. J. and Beardsall, D. (1995) Studies of host location behavior in the black fly Simulium arcticum (Iis-10.11) (Diptera, Simuliidae) – aspects of close range trap orientation. Bull. Ent. Res., 85, 415–24CrossRefGoogle Scholar
Sutherland, D. R., Christensen, B. M. and Lasee, B. A. (1986) Midgut barrier as a possible factor in filarial worm vector competency in Aedes trivittatus. J. Invert. Path., 47, 1–7CrossRefGoogle ScholarPubMed
Sutherst, R. W., Ingram, J. S. I. and Scherm, H. (1998) Global change and vector-borne diseases. Parasitology Today, 14, 297–9CrossRefGoogle ScholarPubMed
Sutton, O. G. (1947) The problem of diffusion in the lower atmosphere. Quart. J. Roy. Meteorol. Soc., 73, 257–81CrossRefGoogle Scholar
Swellengrebel, N. H. (1929) La dissociation des fonctions sexuelles de nutritives (dissociation gonotrophique) d'Anopheles maculipennis comme cause du paludisme dans les Pays-Bas et ses rapports avec ‘l'infection domiciliare’. Ann. Inst. Pasteur, 43, 1370–89Google Scholar
Takehana, A., Katsuyama, T., Yano, T. et al. (2002) Overexpression of a pattern-recognition receptor, peptidoglycan-recognition protein-LE, activates imd/relish-mediated antibacterial defense and the prophenoloxidase cascade in Drosophila larvae. Proc. Natl. Acad. Sci. USA, 99, 13705–10CrossRefGoogle ScholarPubMed
Takken, W. (1996) Synthesis and future challenges: the response of mosquitoes to host odours. In Cardew, G. (ed.), Olfaction in Mosquito–Host Interactions. Chichester:Wiley 302–20Google Scholar
Takken, W., Kager, P. A., and Kaay, H. J., (1999) Endemische malaria terug in Nederland?Nederlands Tijdschrift voor Geneeskunde, 143, 836–8Google Scholar
Takken, W., Klowden, M. J. and Chambers, G. M. (1998) Effect of body size on host seeking and blood meal utilization in Anopheles gambiae sensu stricto (Diptera: Culicidae): the disadvantage of being small. J. Med. Ent., 35, 639–45CrossRefGoogle ScholarPubMed
Takken, W. and Knols, B. G. J. (1999) Odor-mediated behaviour of afrotropical malaria mosquitoes. Ann. Rev. Ent., 44, 131–57CrossRefGoogle ScholarPubMed
Takken, W., Loon, J. J. A. and Adam, W. (2001) Inhibition of host-seeking response and olfactory responsiveness in Anopheles gambiae following blood feeding. J. Insect Physiol., 47, 303–10CrossRefGoogle ScholarPubMed
Tashiro, H. and Schwardt, H. H. (1953) Biological studies of horse flies in New York. J. Econ. Ent., 46, 813–22CrossRefGoogle Scholar
Tawfik, M. S. (1968) Feeding mechanisms and the forces involved in some blood-sucking insects. Quaes. Ent., 4, 92–111Google Scholar
Taylor, P. J. and Hurd, H. (2001) The influence of host haematocrit on the blood feeding success of Anopheles stephensi: implications for enhanced malaria transmission. Parasitology, 122, 491–6CrossRefGoogle ScholarPubMed
Teesdale, C. (1955) Studies on the bionomics of Aedes aegypti L. in its natural habitats in a coastal region of Kenya. Bull. Ent. Res., 46, 711–42CrossRefGoogle Scholar
Tempelis, C. H. and Washino, R. K. (1967) Host feeding patterns of Culex tarsalis in the Sacramento Valley, California, with notes on other species. J. Med. Ent., 4, 315–18CrossRefGoogle ScholarPubMed
Terra, W. R. (1988a) Physiology and biochemistry of insect digestion: an evolutionary perspective. Braz. J. Med. Biol. Res., 21, 675–734Google Scholar
Terra, W. R. (2001) The origin and functions of the insect peritrophic membrane and peritrophic gel. Arch. Insect Biochem. Physiol., 47, 47–61CrossRefGoogle ScholarPubMed
Terra, W. R. and Ferreira, C. (1994) Insect digestive enzymes – properties, compartmentalization and function. Comp. Biochem. Physiol. B, 109, 1–62CrossRefGoogle Scholar
Terra, W. R., Ferreira, C . and Garcia, E. S. (1988b) Origin, distribution, properties and functions of the major Rhodnius prolixus midgut hydrolases. Insect Biochem., 18, 423–34CrossRefGoogle Scholar
Thathy, V., Severson, D. W. and Christensen, B. M. (1994) Reinterpretation of the genetics of susceptibility of Aedes aegypti to Plasmodium gallinaceum. J. Parasitol., 80, 705–12CrossRefGoogle ScholarPubMed
Theodor, O. (1967) An Illustrated Catalogue of the Rothschild Collection of Nycteribiidae (Diptera) in the British Museum (Natural History). London: British Museum
Theodos, C. M., Ribeiro, J. M. and Titus, R. G. (1991) Analysis of enhancing effect of sand fly saliva on Leishmania infection in mice. Infection and Immunity, 59, 1592–8Google ScholarPubMed
Theodos, C. M. and Titus, R. G. (1993) Salivary gland material from the sand fly Lutzomyia longipalpis has an inhibitory effect on macrophage function in vitro. Parasite Immunology, 15, 481–7CrossRefGoogle ScholarPubMed
Thompson, B. H. (1976) Studies on the attraction of Simulium damnosum s.l. (Diptera: Simuliidae) to its hosts. I. The relative importance of sight, exhaled breath and smell. Tropenmed. Parasitol., 27, 455–73Google ScholarPubMed
Thompson, W. H. and Beattey, B. J. (1977) Veneral transmission of La Crosse (California encephalitis) arbovirus in Aedes triseriatus mosquitoes. Science, 196, 530–1CrossRefGoogle ScholarPubMed
Thorsteinson, A. J. and Bracken, G. K. (1965) The orientation behavior of horse flies and deer flies (Tabanidae: Diptera). III. The use of traps in the study of orientation of tabanids in the field. Ent. Exp. Appl., 8, 189–92CrossRefGoogle Scholar
Thorsteinson, A. J., Bracken, G. K. and Tostawaryk, W. (1966) The orientation behaviour of horse flies and deer flies (Tabanidae: Diptera). VI. The influence of the number of reflecting surfaces on attractiveness to tabanids of glossy black polyhedra. Can. J. Zool., 44, 275–9CrossRefGoogle Scholar
Tillyard, R. J. (1935) The evolution of scorpion flies and their derivatives (order Mecoptera). Ann. Ent. Soc. Am., 28, 37–45CrossRefGoogle Scholar
Titus, R. G. (1998) Salivary gland lysate from the sand fly Lutzomyia longipalpis suppresses the immune response of mice to sheep red blood cells in vivo and concanavalin A in vitro. Exp. Parasitol., 89, 133–6CrossRefGoogle ScholarPubMed
Titus, R. G. and Ribeiro, J. M. C. (1990) The role of vector saliva in transmission of arthropod-borne disease. Parasitology Today, 6, 157–60CrossRefGoogle ScholarPubMed
Tobe, S. S. and Davey, K. G. (1972) Volume relationships during the pregnancy cycle of the tsetse fly Glossina austeni. Can. J. Zool., 50, 999–1010CrossRefGoogle ScholarPubMed
Torr, S. J. (1989) The host-orientated behaviour of tsetse flies (Glossina): the interaction of visual and olfactory stimuli. Phys. Ent., 14, 325–40CrossRefGoogle Scholar
Torr, S. J., Hall, D. R. and Smith, J. L. (1995) Responses of tsetse-flies (Diptera, Glossinidae) to natural and synthetic ox odors. Bull. Ent. Res., 85, 157–66CrossRefGoogle Scholar
Torr, S. J. and Mangwiro, T. N. C. (2000) Interactions between cattle and biting flies: effects on the feeding rate of tsetse. Med. Vet. Ent., 14, 400–9CrossRefGoogle ScholarPubMed
Torr, S. J., Wilson, P. J., Schofield, S. et al. (2001) Application of DNA markers to identify the individual-specific hosts of tsetse feeding on cattle. Med. Vet. Ent., 15, 78–86CrossRefGoogle ScholarPubMed
Traub, R. (1985) Coevolution of fleas and mammals. In Kim, K. C. (ed.), Coevolution of Parasitic Arthropods and Mammals. New York: WileyGoogle Scholar
Trpis, M., Duhrkopf, R. E. and Parker, K. L. (1981) Non-Mendelian inheritance of mosquito susceptibility to infection with Brugia malayi and Brugia pahangi. Science, 211, 1435–7CrossRefGoogle ScholarPubMed
Trudeau, W. L., Fernandez-Caldas, E., Fox, R. W. (1993) Allergenicity of the cat flea (Ctenocephalides felis felis). Clinical and Experimental Allergy: Journal of the British Society for Allergy and Clinical Immunology, 23, 377–83CrossRefGoogle Scholar
Turell, M. J., Bailey, C. L. and Rossi, C. A. (1984a) Increased mosquito feeding on Rift Valley fever virus-infected lambs. Am. J. Trop. Med. Hyg., 33, 1232–8CrossRefGoogle Scholar
Turell, M. J., Rossignol, P. A., Spielman, A., Rossi, C. A. and Bailey, C. L. (1984b) Enhanced arboviral transmission by mosquitoes that concurrently ingested microfilaria. Science, 225, 1039–41CrossRefGoogle Scholar
Turner, D. A. and Invest, J. F. (1973) Laboratory analyses of vision in tsetse flies (Dipt., Glossinidae). Bull. Ent. Res., 62, 343–57CrossRefGoogle Scholar
Tzou, P., Gregorio, E. and Lemaitre, B. (2002) How Drosophila combats microbial infection: a model to study innate immunity and host-pathogen interactions. Curr. Opin. Microbiol., 5, 102–10CrossRefGoogle ScholarPubMed
Tzou, P., Ohresser, S., Ferrandon, D. et al. (2000) Tissue-specific inducible expression of antimicrobial peptide genes in Drosophila surface epithelia. Immunity, 13, 737–48CrossRefGoogle ScholarPubMed
Underhill, G. W. (1940) Some factors influencing feeding activity of simuliids in the field. J. Econ. Entomol., 33, 915–17CrossRefGoogle Scholar
Vale, G. A. (1974a) New field methods for studying the response of tsetse flies (Diptera, Glossinidae) to hosts. Bull. Ent. Res., 64, 199–208CrossRefGoogle Scholar
Vale, G. A. (1974b) The response of tsetse flies (Diptera, Glossinidae) to mobile and stationary baits. Bull. Ent. Res., 64, 545–88CrossRefGoogle Scholar
Vale, G. A. (1977) Feeding responses of tsetse flies (Diptera: Glossinidae) to stationary hosts. Bull. Ent. Res., 67, 635–49CrossRefGoogle Scholar
Vale, G. A. (1980) Flight as a factor in host-finding behaviour of tsetse flies (Diptera: Glossinidae). Bull. Ent. Res., 70, 299–307CrossRefGoogle Scholar
Vale, G. A. (1982) The trap-orientated behaviour of tsetse flies (Glossinidae) and other Diptera. Bull. Ent. Res., 72, 71–93CrossRefGoogle Scholar
Vale, G. A. (1983) The effects of odours, wind direction and wind speeds on the distribution of Glossina (Diptera: Glossinidae) and other insects near stationary targets. Bull. Ent. Res., 73, 53–64CrossRefGoogle Scholar
Vale, G. A. and Hall, D. R. (1985a) The role of 1-octen-3-ol, acetone and carbon dioxide in the attraction of tsetse flies, Glossina spp. (Diptera: Glossinidae), to ox odour. Bull. Ent. Res., 75, 209–17CrossRefGoogle Scholar
Vale, G. A. and Hall, D. R. (1985b) The use of 1-octen-3-ol, acetone and carbon dioxide to improve baits for tsetse flies, Glossina spp. (Diptera: Glossinidae). Bull. Ent. Res., 75, 219–31CrossRefGoogle Scholar
Vale, G. A., Hall, D. R. and Gough, A. J. E. (1988) The olfactory responses of tsetse flies, Glossina spp. (Diptera: Glossinidae), to phenols and urine in the field. Bull. Ent. Res., 78, 293–300CrossRefGoogle Scholar
Valenzuela, J. G., Belkaid, Y., Rowton, E. and Ribeiro, J. M. C. (2001) The salivary apyrase of the blood-sucking sand fly Phlebotomus papatasi belongs to the novel Cimex family of apyrases. J. Exp. Biol., 204, 229–37Google ScholarPubMed
Valenzuela, J. G., Francischetti, I. M. B. and Ribeiro, J. M. C. (1999) Purification, cloning, and synthesis of a novel salivary anti-thrombin from the mosquito Anopheles albimanus. Biochemistry, 38, 11209–15CrossRefGoogle ScholarPubMed
Valenzuela, J. G., Pham, V. M., Garfield, M. K., Francischetti, I. M. B. and Ribeiro, J. M. C. (2002) Toward a description of the sialome of the adult female mosquito Aedes aegypti. Insect Biochemistry and Molecular Biology, 32, 1101–22CrossRefGoogle Scholar
Valenzuela, J. G. and Ribeiro, J. M. C. (1998) Purification and cloning of the salivary nitrophorin from the hemipteran Cimex lectularius. J. Exp. Biol., 201, 2659–64Google ScholarPubMed
Handel, E. (1984) Metabolism of nutrients in the adult mosquito. Mosq. News, 44, 573–9Google Scholar
Naters, W. M. V., Otter, C. J. and Cuisance, D. (1998) The interaction of taste and heat on the biting response of the tsetse fly Glossina fuscipes fuscipes. Phys. Ent., 23, 285–8Google Scholar
Vargaftig, B. B., Chignard, M. and Benveniste, J. (1981) Present concepts on the mechanism of platelet aggregation. Biochem. Pharmacol., 30, 263–71
Vaughan, J. A. and Turell, M. J. (1996) Facilitation of Rift Valley fever virus transmission by Plasmodium berghei sporozoites in Anopheles stephensi mosquitoes. Am. J. Trop. Med. Hyg., 55, 407–9CrossRefGoogle ScholarPubMed
Venkatesh, K. and Morrison, P. E. (1982) Blood meal as a regulator of triacylglycerol synthesis in the haematophagous stable fly, Stomoxys calcitrans. J. Comp. Physiol., 147, 49–52CrossRefGoogle Scholar
Venkatesh, K., Morrison, P. E. and Kallapur, V. L. (1981) Influence of blood meals on the conversion of D-(U-14C)-glucose to lipid in the fat body of the haematophagous stablefly, Stomoxys calcitrans. Comp. Biochem. Physiol., 68, 425–9Google Scholar
Vernick, K. D., Fujioka, H., Seeley, D. C. et al. (1995) Plasmodium gallinaceum – a refractory mechanism of ookinete killing in the mosquito, Anopheles gambiae. Experimental Parasitology, 80, 583–95CrossRefGoogle ScholarPubMed
Victoir, K. and Dujardin, J. C. (2002) How to succeed in parasitic life without sex? Asking Leishmania. Trends Parasitol., 18, 81–5CrossRefGoogle ScholarPubMed
Voskamp, K. E., Otter, C. J. and Noorman, N. (1998) Electroantennogram responses of tsetse flies (Glossina pallidipes) to host odours in an open field and riverine woodland. Phys. Ent., 23, 176–83CrossRefGoogle Scholar
Waage, J. K. (1979) The evolution of insect/vertebrate associations. Biological Journal of the Linnaeon Society, 12, 187–224CrossRefGoogle Scholar
Waage, J. K. (1981) How the zebra got its stripes – biting flies as selective agents in the evolution of zebra coloration. J. Ent. Soc. Sth. Afr., 44, 351–8Google Scholar
Waage, J. K. and Davies, C. R. (1986) Host-mediated competition in a bloodsucking insect community. Journal of Animal Ecology, 55, 171–80CrossRefGoogle Scholar
Waage, J. K. and Nondo, J. (1982) Host behaviour and mosquito feeding success: an experimental study. Trans. R. Soc. Trop. Med. Hyg., 76, 119–22CrossRefGoogle Scholar
Wahid, I., Sunahara, T. and Mogi, M. (2003) Maxillae and mandibles of male mosquitoes and female autogenous mosquitoes (Diptera: Culicidae). J. Med. Ent., 40, 150–8CrossRefGoogle Scholar
Ward, R. A. (1963) Genetic aspects of the susceptibility of mosquitoes to malaria infections. Exp. Parasit., 13, 328–41CrossRefGoogle Scholar
Warnes, M. L. (1995) Field studies on the effect of cattle skin secretion on the behavior of tsetse. Med. Vet. Ent., 9, 284–8CrossRefGoogle ScholarPubMed
Warnes, M. L. and Finlayson, L. H. (1985) Responses of the stable fly, Stomoxys calcitrans (L.) (Diptera: Muscidae), to carbon dioxide and host odours. I. Activation. Bull. Ent. Res., 75, 519–27CrossRefGoogle Scholar
Warnes, M. L. and Finlayson, L. H. (1986) Electroantennogram responses of the stable fly, Stomoxys calcitrans, to carbon dioxide and other odours. Phys. Ent., 11, 469–73CrossRefGoogle Scholar
Warnes, M. L. and Finlayson, L. H. (1987) Effect of host behaviour on host preference in Stomoxys calcitrans. Med. Vet. Ent., 1, 53–7CrossRefGoogle ScholarPubMed
Waterhouse, D. F. (1953) The occurrence and significance of the peritrophic membrane, with special reference to adult Lepidoptera and Diptera. Aust. J. Zool., 1, 299–318CrossRefGoogle Scholar
Watts, D. M., Pantuwatana, S., Defoliart, G. S., Yuill, T. M. and Thompson, W. H. (1973) Transovarial transmission of La Crosse virus (California encephalitis group) in the mosquito, Aedes triseriatus. Science, 182, 1140–1CrossRefGoogle Scholar
Webb, P. A., Happ, C. M., Maupin, G. O. et al. (1989) Potential for insect transmission of HIV: experimental exposure of Cimex hemipterus and Toxorhynchites amboinensis to human immunodeficiency virus. J. Infect Dis., 160, 970–7CrossRefGoogle ScholarPubMed
Webber, L. A. and Edman, J. D. (1972) Anti-mosquito behaviour of ciconiiform birds. Animal Behaviour, 20, 228–32CrossRefGoogle Scholar
Webster, J. P. and Woolhouse, M. E. J. (1999) Cost of resistance: relationship between reduced fertility and increased resistance in a snail-schistosome host-parasite system. Proc. R. Soc. Lond. B Sci., 266, 391–6CrossRefGoogle Scholar
Wee, W. L. and Anderson, J. R. (1995) Tethered flight capabilities and survival of Lambornella clarki-infected, blood-fed, and gravid Aedes sierrensis (Diptera, Culicidae). J. Med. Ent., 32, 153–60Google Scholar
Weitz, B. (1963) The feeding habits of Glossina. Bull. WHO, 28, 711–29Google ScholarPubMed
Wekesa, J. W., Copeland, R. S. and Mwangi, R. W. (1992) Effect of Plasmodium falciparum on blood feeding behavior of naturally infected Anopheles mosquitoes in western Kenya. Am. J. Trop. Med. Hyg., 47, 484–8CrossRefGoogle ScholarPubMed
Welburn, S. C., et al. (1987) In vitro cultivation of rickettsia-like organisms from Glossina spp. Ann. Trop. Med. Parasit., 81(4), 331–5CrossRef
Welburn, S. C., Arnold, K., Maudlin, I. and Gooday, G. W. (1993) Rickettsia-like organisms and chitinase production in relation to transmission of trypanosomes by tsetse-flies. Parasitology, 107, 141–5CrossRefGoogle ScholarPubMed
Welburn, S. C. and Murphy, N. B. (1998) Prohibitin and RACK homologues are up-regulated in trypanosomes induced to undergo apoptosis and in naturally occurring terminally differentiated forms. Cell Death and Differentiation, 5 (7), 615–22CrossRef
Wells, E. A. (1982) Trypanosomiasis in the absence of tsetse. In Baker, J. R. (ed.), Perspectives in Trypansosomiasis Research. Chichester:Research Studies PressGoogle Scholar
Wen, Y., Muir, L. E. and Kay, B. H. (1997) Response of Culex quinquefasciatus to visual stimuli. J. Am. Mosq. Control Assoc., 13, 150–2Google ScholarPubMed
Wenk, P. (1962) Anatomie des Kopfes von Wilhelmia equina (Simuliidae syn. Melusinidae, Diptera). Zool. Jahrb. Abt. Ontog. Tiere, 80, 81–134Google Scholar
Wenk, P. and Schlorer, G. (1963) Wirtsorientierung und Kopulation bei blutsaugenden Simuliiden (Diptera). Z. Tropenmed. Parasitol., 14, 177–91Google Scholar
Werner, T., Liu, G., Kang, D. et al. (2000) A family of peptidoglycan recognition proteins in the fruit fly Drosophila melanogaster. Proc. Natl. Acad. Sci. USA, 97, 13772–7CrossRefGoogle ScholarPubMed
Werner-Reiss, U., Galun, R., Crnjar, R. and Liscia, A. (1999) Factors modulating the blood feeding behavior and the electrophysiological responses of labral apical chemoreceptors to adenine nucleotides in the mosquito Aedes aegypti (Culicidae). J. Insect Physiol., 45, 801–8CrossRefGoogle Scholar
Weyer, F. (1960) Biological relationships between lice (Anoplura) and microbial agents. Ann. Rev. Ent., 5, 405–20CrossRefGoogle Scholar
Wharton, R. H. (1957) Studies on filariasis in Malaya: observations on the development of Wuchereria malayi in Mansonia (Mansonioides) longipalpis. Ann. Trop. Med. Parasit., 51, 278–96CrossRefGoogle ScholarPubMed
White, G. B. (1974) Anopheles gambiae complex and disease transmission in Africa. Trans. R. Soc. Trop. Med. Hyg., 4, 278–98CrossRefGoogle Scholar
White, G. B., Magayuka, S. A. and Boreham, P. F. L. (1972) Comparative studies on sibling species of the Anopheles gambiae Giles complex (Dipt. Culicidae): bionomics and vectorial capacity of species A and species B at Segera, Tanzania. Bull. Ent. Res., 62, 295–317CrossRefGoogle Scholar
White, G. B. and Rosen, B. (1973) Comparative studies on sibling species of the Anopheles gambiae Giles complex (Dipt. Culicidae). II. Ecology of species A and B in savanna around Kaduna, Nigeria, during transition from wet to dry season. Bull. Ent. Res., 62, 613–25CrossRefGoogle Scholar
Whiting, M. F. (2002) Mecoptera is paraphyletic: multiple genes and phylogeny of Mecoptera and Siphonaptera. Zoologica Scripta, 31, 93–104CrossRefGoogle Scholar
Whitten, M. M. and Ratcliffe, N. A. (1999) In vitro superoxide activity in the haemolymph of the West Indian leaf cockroach, Blaberus discoidalis. J. Insect Physiol., 45, 667–75CrossRefGoogle ScholarPubMed
Wigglesworth, V. B. (1941) The sensory physiology of the human louse Pediculus humanus corporis de Greer (Anoplura). Parasitology, 33, 67–109CrossRefGoogle Scholar
Wigglesworth, V. B. (1979) Secretory activities of plasmatocytes and oenocytoids during the moulting cycle in an insect (Rhodnius). Tissue and Cell, 11, 69–78CrossRefGoogle Scholar
Wigglesworth, V. B. and Gillett, J. D. (1934) The function of antennae in Rhodnius prolixus and the mechanism of orientation of the host. J. Exp Biol., 11, 120–39Google Scholar
Williams, B. (1994) Models of trap seeking by tsetse-flies – anemotaxis, klinokinesis and edge-detection. Journal of Theoretical Biology, 168, 105–15CrossRefGoogle Scholar
Williams, P. D. and Day, T. (2001) Interactions between sources of mortality and the evolution of parasite virulence. Proc. R. Soc. Lond. B Sci., 268, 2331–7CrossRefGoogle ScholarPubMed
Wilson, J. J., Neame, P. V. and Kelton, J. G. (1982) Infection induced thrombocytopaenia. Seminars in Thrombosis and Haemostasis, 8, 217–33CrossRefGoogle Scholar
Wilson, M. (1978) The functional organisation of locust ocelli. J. Comp. Physiol., 124, 297–316CrossRefGoogle Scholar
Wilson, R., Chen, C. W. and Ratcliffe, N. A. (1999) Innate immunity in insects: the role of multiple, endogenous serum lectins in the recognition of foreign invaders in the cockroach, Blaberus discoidalis. Journal of Immunology, 162, 1590–6Google ScholarPubMed
Woke, P. A. (1937) Comparative effects of the blood of man and of canary on egg production of Culex pipiens Linn. J. Parasit., 23, 311–13CrossRefGoogle Scholar
Wood, D. M. (1964) Studies on the beetles Leptinillus validus (Horn) and Platypsyllus castoris Rissema (Coleoptera: Leptinidae) from beaver. Proceedings of the Entomological Society of Ontario, 95, 33–63Google Scholar
Wood, S. F. (1942) Observations on vectors of Chagas' disease in the United States. I. California. Bull. Calif. Acad. Sci., 41, 61–9Google Scholar
Worms, M. J. (1972) Circadian and seasonal rhythms in blood parasites. In Canning, E. U. and Wright, C. A. (eds.), Behavioural Aspects of Parasite Transmission. London: Linnean SocietyGoogle Scholar
Wright, R. H. (1958) The olfactory guidance of flying insects. Can. Entomol., 90, 81–9CrossRefGoogle Scholar
Wright, R. H. (1968) Tunes to which mosquitoes dance. New Sci., 37, 694–7Google Scholar
Wright, R. H. and Kellogg, F. E. (1962) Response of Aedes aegypti to moist convection currents. Nature, 194, 402–3CrossRefGoogle ScholarPubMed
Xie, H., Bain, O. and Williams, S. A. (1994) Molecular phylogenetic studies on filarial parasites based on 5s ribosomal spacer sequences. Parasite-Journal de la Societe Française de Parasitologie, 1, 141–51Google ScholarPubMed
Xu, P. X., Zwiebel, L. J. and Smith, D. P. (2003) Identification of a distinct family of genes encoding atypical odorant-binding proteins in the malaria vector mosquito, Anopheles gambiae. Insect Mol. Biol., 12, 549–60CrossRefGoogle ScholarPubMed
Yajima, M., Takada, M., Takahashi, N. et al. (2003) A newly established in vitro culture using transgenic Drosophila reveals functional coupling between the phospholipase A2-generated fatty acid cascade and lipopolysaccharide-dependent activation of the immune deficiency (imd) pathway in insect immunity. Biochem. J., 371, 205–10CrossRefGoogle ScholarPubMed
Yu, X. Q. and Kanost, M. R. (2000) Immulectin-2, a lipopolysaccharide specific lectin from an insect, Manduca sexta, is induced in response to gram-negative bacteria. J. Biol. Chem., 275, 37373–81CrossRefGoogle ScholarPubMed
Yuill, T. M. (1983) The role of mammals in the maintainence and dissemination of La Crosse virus. In Calisher, C. H. and Thompson, W. H. (eds.), California Serogroup Viruses. New York: Alan R. LissGoogle Scholar
Zahedi, M. (1994) The fate of Brugia pahangi microfilariae in Armigeres subalbatus during the first 48 hours post ingestion. Tropical Medicine and Parasitology, 45, 33–5Google ScholarPubMed
Zhang, D., Cupp, M. S. and Cupp, E. W. (2002) Thrombostasin: purification, molecular cloning and expression of a novel anti-thrombin protein from horn fly saliva. Insect Biochemistry and Molecular Biology, 32, 321–30CrossRefGoogle ScholarPubMed
Zhang, Y., Ribeiro, J. M. C., Guimaraes, J. A. and Walsh, P. N. (1998) Nitrophorin-2: a novel mixed-type reversible specific inhibitor of the intrinsic factor-X activating complex. Biochemistry, 37, 10681–90CrossRefGoogle ScholarPubMed
Zheng, L. (1999) Genetic basis of encapsulation response in Anopheles gambiae. Parasitologia, 41, 181–4Google ScholarPubMed
Zheng, L., Cornel, A. J., Wang, R. (1997) Quantitative trait loci for refractoriness of Anopheles gambiae to Plasmodium cynomolgi B. Science, 276, 425–8CrossRefGoogle ScholarPubMed
Zheng, L., Wang, S., Romans, P., et al. (2003) Quantitative trait loci in Anopheles gambiae controlling the encapsulation response against Plasmodium cynomolgi Ceylon. BMC Genet, 4, 16CrossRefGoogle ScholarPubMed
Zieler, H., Garon, C. F., Fischer, E. R. and Shahabuddin, M. (2000) A tubular network associated with the brush-border surface of the Aedes aegypti midgut: implications for pathogen transmission by mosquitoes. J. Exp. Biol., 203, 1599–611Google ScholarPubMed
Aboul-Nasr, A. E. (1967) On the behaviour and sensory physiology of the bed bug (Cimex lectularius). I. Temperature reactions. Bull. Soc. Ent. Egypte, 51, 43–54Google Scholar
Adams, T. S. (1999) Hematophagy and hormone release. Ann. Ent. Soc. Am., 92, 1–13CrossRefGoogle Scholar
Adlington, D., Randolph, S. E. and Rogers, D. J. (1996) Flying to feed or flying to mate: gender differences in the flight activity of tsetse (Glossina palpalis). Phys. Ent., 21, 85–92CrossRefGoogle Scholar
Ahmadi, A. and McClelland, G. A. H. (1985) Mosquito-mediated attraction of female mosquitoes to a host. Phys. Ent., 10, 251–5CrossRefGoogle Scholar
Ahmed, A. M., Baggott, S. L., Maingon, R. and Hurd, H. (2002) The costs of mounting an immune response are reflected in the reproductive fitness of the mosquito Anopheles gambiae. Oikos, 97, 371–7CrossRefGoogle Scholar
Ahmed, A. M., Maingon, R., Romans, P. and Hurd, H. (2001) Effects of malaria infection on vitellogenesis in Anopheles gambiae during two gonotrophic cycles. Insect Mol. Biol., 10, 347–56CrossRefGoogle ScholarPubMed
Ahmed, A. M., Maingon, R. D., Taylor, P. J. and Hurd, H. (1999) The effects of infection with Plasmodium yoelii nigeriensis on the reproductive fitness of the mosquito Anopheles gambiae. Invertebrate Reproduction and Development, 36, 217–22CrossRefGoogle Scholar
Aikawa, M., Suzuki, M. and Gutierez, Y. (1980) Pathology of malaria. In Kreier, J. P. (ed.), Malaria. New York:Academic PressGoogle Scholar
Akai, H. and Sato, S. (1973) Ultrastructure of the larval hemocytes of the silkworm, Bombyx mori L. (Lepidoptera: Bombycidae). International Journal of Insect Morphology and Embryology, 2, 207–31CrossRefGoogle Scholar
Akman, L., Yamashita, A., Watanabe, H.et al. (2002) Genome sequence of the endocellular obligate symbiont of tsetse flies, Wigglesworthia glossinidia. Nature Genetics, 32, 402–7CrossRefGoogle ScholarPubMed
Aksoy, S. (2000) Tsetse – a haven for microorganisms. Parasitology Today, 16, 114–18CrossRefGoogle ScholarPubMed
Aksoy, S., Gibson, W. C. and Lehane, M. J. (2003) Perspectives on the interactions between tsetse and trypanosomes with implications for the control of trypanosomiasis. Advances in Parasitology, 53, 1–84CrossRefGoogle Scholar
Albritton, A. (1952) Standard Values in Blood. Phiadelphia: Saunders
Alekseev, A., Rasnityn, S. and Vitlin, L. (1977) On group behaviour of females of blood-sucking mosquitoes. Communication I. Discovery of the ‘effect’ of invitation. [In Russian]Parazitologiyai Parazitarnye Bolezni, 46, 23–4Google Scholar
Alekseyev, A. N., Abdullayev, I. T., Rasnitsyn, S. P. and Martsinovskiy, M. (1984) Comparison of flight capability of Aedes aegypti that are infected and not infected with plasmodia. Med. Parazitol., 1, 11–13Google Scholar
Allan, S. A., Day, J. F. and Edman, J. D. (1987) Visual ecology of biting flies. Ann. Rev. Ent., 32, 297–316CrossRefGoogle ScholarPubMed
Allan, S. A. and Stoffolano, J. G. (1986a) Effects of background contrast on visual attraction and orientation of Tabanus nigrovittatus (Diptera: Tabanidae). Environ. Entomol., 15, 689–94CrossRefGoogle Scholar
Allan, S. A. and Stoffolano, J. G. (1986b) The effects of hue and intensity on visual attraction of adult Tabanus nigrovittatus (Diptera: Tabanidae). J. Med. Ent., 23, 83–91CrossRefGoogle Scholar
Altman, P. L. and Dittmer, D. S.(1971) Blood and other body fluids. In Respiration and Circulation. Bethesda, MDI: Federation of American Societies of Experimental Biology, p. 540
Altner, H. and Loftus, R. (1985) Ultrastructure and function of insect thermo and hygroreceptors. Ann. Rev. Ent., 30, 273–95CrossRefGoogle Scholar
Amin, O. M. and Wagner, M. E. (1983) Further notes on the function of pronotal combs in fleas (Siphonaptera). Ann. Ent. Soc. Am., 76, 232–4CrossRefGoogle Scholar
Anderson, J. R. and Ayala, S. C. (1968) Trypanosome transmitted by a Phlebotomus: first report from the Americas. Science, 161, 1023–5CrossRefGoogle ScholarPubMed
Anderson, J. R. and Hoy, J. B. (1972) Relationship between host attack rates and CO2 baited insect flight trap catches of certain Symphoromyia spp. J. Med. Ent., 9, 373–92CrossRefGoogle Scholar
Anderson, R. A. and Brust, R. A. (1996) Blood feeding success of Aedes aegypti and Culex nigripalpus (Diptera: Culicidae) in relation to defensive behavior of Japanese quail (Coturnix japonica) in the laboratory. Journal of Vector Ecology, 21, 94–104Google Scholar
Anderson, R. A., Koella, J. C. and Hurd, H. (1999) The effect of Plasmodium yoelii nigeriensis infection on the feeding persistence of Anopheles stephensi Liston throughout the sporogonic cycle. Proc. R. Soc. Lond. B Biol. Sci., 266, 1729–33CrossRefGoogle ScholarPubMed
Anderson, R. C. (1957) The life cycles of Dipetalonematid nematodes (Filarioidea, Dipetalonematidae): the problem of their evolution. J. Helminth., 31, 203–24CrossRefGoogle ScholarPubMed
Anderson, R. M. (1986) Genetic variability in resistance to parasitic invasion: population implications for invertebrate host species. In Lackie, A. M. (ed.), Immune Mechanisms in Invertebrate Vectors, Vol. 56, Oxford: Oxford University PressGoogle Scholar
Anderson, R. M. and May, R. M. (1978) Regulation and stability of host–parasite population interactions. Journal of Animal Ecology, 47, 219–47CrossRefGoogle Scholar
Ansell, J., Hamilton, K. A., Pinder, M., Walraven, G. E. L. and Lindsay, S. W. (2002) Short-range attractiveness of pregnant women to Anopheles gambiae mosquitoes. Trans. R. Soc. of Trop. Med. and Hyg., 96, 113–16CrossRefGoogle ScholarPubMed
Arlian, L. (2002) Arthropod allergens and human health. In Annual Review of Entomology. Annual Reviews Ic., Palo Alto, Vol. 47, 395–434
Arrese, E. L., Canavoso, L. E., Jouni, Z. E.et al. (2001) Lipid storage and mobilization in insects: current status and future directions. Insect Biochemistry and Molecular Biology, 31, 7–17CrossRefGoogle ScholarPubMed
Aschner, M. (1932) Experimentelle unter Suchungen über die Symbiose der Kleiderlaus. Naturwiss., 20, 501–5CrossRefGoogle Scholar
Aschner, M. (1934) Studies on the symbiosis of the body louse. I. Elimination of the symbionts by centrifugation of the eggs. Parasitol., 26, 309–14CrossRefGoogle Scholar
Aschner, M. (1946) The symbiosis of Eucampsipoda aegyptica Mcq. Bull. Soc. Ent. Egypte, 30, 1–6Google Scholar
Ashford, R. W. (2001) The Leishmania ses. In M. W. Service (ed.), Encyclopedia of Arthropod-transmitted Infections of Man and Domesticated Animals. CABI, 269–79
Ashida, M., Ochiai, M. and Niki, T. (1988) Immunolocalization of prophenoloxidase among hemocytes of the Silkworm, Bombyx mori. Tissue and Cell, 20, 599–610Google ScholarPubMed
Audy, J. R., Radovsky, F. J. and Vercammen-Grandjean, P. H. (1972) Neosomy: radical intrastadial metamorphosis associated with arthropod symbioses. J. Med. Ent., 9, 487–94CrossRefGoogle ScholarPubMed
Avila, A., Silverman, N., Diaz-Meco, M. T. and Moscat, J. (2002) The Drosophila atypical protein kinase C-ref(2)p complex constitutes a conserved module for signaling in the toll pathway. Mol. Cell. Biol. 22, 8787–95CrossRefGoogle ScholarPubMed
Ayala, S. C. and Lee, D. (1970) Saurian malaria: development of sporozoites in two species of phelebotomine sandflies. Science, 167, 891–2CrossRefGoogle Scholar
Bacot, A. M. and Martin, C. J. (1914) Observations on the mechanism of the transmission of the plague by fleas. J. Hyg., 13, 423–39Google ScholarPubMed
Bailey, L. (1952) The action of the proventriculus of the worker honeybee. J. Exp. Biol., 29, 310–27Google Scholar
Baker, J. R. (1965) The evolution of parasitic protozoa. In Taylor, A. E. R. (ed.), of the British Society for Parasitology. Oxford: Blackwell, Vol. 3Google Scholar
Baker, R. C. (1986) Pheromane-modulated movement of flying moths. In Payne, T. L., Birch, M. C. and Kennedy, C. E. J. (eds.), Mechanisms in Insect Olfaction. Oxford: Clarendon PressGoogle Scholar
Baker, T. C. (1990) Upwind flight and casting flight: complementary phasic and tonic systems used for location of sex pheromone sources by male moths. In K. B. Doving (ed.), Symposium on Olfaction and Taste, Oslo, Vol. X
Balashov, Y. (1984) Interaction between blood-sucking arthropods and their hosts, and its influence on vector potential. Ann. Rev. Ent., 29, 137–56CrossRefGoogle ScholarPubMed
Ball, G. H. (1943) Parasitism and evolution. Am. Nat., 77, 345–64CrossRefGoogle Scholar
Banziger, H. (1971) Blood-sucking moths of Malaya. Fauna, 1, 5–16Google Scholar
Baranov, N. (1935) New information on the Golubatz fly, S. columbaczense. Rev. Appl. Ent. B, 23, 275–6Google Scholar
Barrera, A. (1966) Hallazgo de Amblyopinus tiptoni Barrera, 1966 en Costa Rica, A. C. (Col.: Staph.). Acta zool. Mex., 8, 1–3Google Scholar
Barrera, A. and Machado-Allison, C. E. (1965) Coleopteros ectoparasiticos de Mamiferos. Ciencia Mexico, 23, 201–8Google Scholar
Bar-Zeev, M., Maibach, H. I. and Khan, A. A. (1977) Studies on the attraction of Aedes aegypti (Diptera: Culicidae) to man. J. Med. Ent., 14, 113–20CrossRefGoogle Scholar
Baudisch, K. (1958) Beitraäge zur Zytologie und Embryologie einiger Insektensymbiosen. Zeit. Morph. Okol. Tiere, 47, 436–88CrossRefGoogle Scholar
Baylis, M. and Mbwabi, A. L. (1995) Feeding-behavior of tsetse-flies (Glossina pallidipes Austen) on Trypanosoma-infected oxen in Kenya. Parasitology, 110, 297–305CrossRefGoogle ScholarPubMed
Beach, R., Kiilu, G. and Leeuwenburg, J. (1985) Modification of sandfly biting behaviour by Leishmania leads to increased parasite transmission. Am. J. Trop. Med. Hyg., 34, 279–83CrossRefGoogle ScholarPubMed
Beard, C. B., Mason, P. W., Aksoy, S., Tesh, R. B. and Richards, F. F. (1992) Transformation of an insect symbiont and expression of a foreign gene in the Chagas-disease vector Rhodnius prolixus. Am. J. Trop. Med. and Hyg., 46, 195–200CrossRefGoogle Scholar
Beckenbach, A. T. and Borkent, A. (2003) Molecular analysis of the biting midges (Diptera: Ceratopogonidae), based on mitochondrial cytochrome oxidase subunit 2. Mol. Phylogenet. Evol. 27, 21–35CrossRefGoogle Scholar
Beckett, E. B. and Macdonald, W. W. (1971) The development and survival of subperiodic Brugia malayi and B. pahangi larvae in a selected strain of Aedes aegypti. Trans. R. Soc. Trop. Med. Hyg., 65, 339–46CrossRefGoogle Scholar
Beenakkers, A. M. T., Horst, D. J. and Marrewijk, W. J. A. (1984) Insect flight muscle metabolism. Insect Biochemistry, 14, 243–60CrossRefGoogle Scholar
Beerntsen, B. T., James, A. A. and Christensen, B. M. (2000) Genetics of mosquito vector competence. Microbiology and Molecular Biology Reviews, 64, 115–37CrossRefGoogle ScholarPubMed
Beerntsen, B. T., Severson, D. W., Klinkhammer, J. A., Kassner, V. A. and Christensen, B. M. (1995) Aedes aegypti: A quantitative trait locus (QTL) influencing filarial worm intensity is linked to QTL for susceptibility to other mosquito-borne pathogens. Experimental Parasitology, 81, 355–62CrossRefGoogle ScholarPubMed
Beklemishev, V. N. (1957) Some general questions on the biology of bloodsucking lower flies. Meditsinskaya parazitologiya i parazitarnye bolezni, 5, 562–6Google Scholar
Belkaid, Y., Valenzuela, J. G., Kamhawi, S.et al. (2000) Delayed-type hypersensitivity to Phlebotomus papatasi sand fly bite: An adaptive response induced by the fly?Proceedings of the National Academy of Sciences of the United States of America, 97, 6704–9CrossRefGoogle Scholar
Bell, J. F., Stewart, S. J. and Nelson, W. A. (1982) Transplant of acquired resistance to Polyplax serrata (Phthiraptera: Hoplopleuridae) in skin allografts to athymic mice. J. Med. Ent., 19, 164–8CrossRefGoogle ScholarPubMed
Belzer, W. R. (1978a) Factors conducive to increased protein feeding by the blowfly Phormia regina. Phys. Ent. 3, 251–7CrossRefGoogle Scholar
Belzer, W. R. (1978b) Patterns of selective protein ingestion by the blowfly Phormia regina. Phys. Ent., 3, 169–257CrossRefGoogle Scholar
Belzer, W. R.(1978c) Recurrent nerve inhibition of protein feeding in the blowfly Phormia regina. Phys. Ent., 3, 259–63CrossRefGoogle Scholar
Belzer, W. R. (1979) Abdominal stretch receptors in the regulation of protein ingestion by the black blowfly, Phormia regina. Phys. Ent., 4, 7–14CrossRefGoogle Scholar
Bennet-Clark, H. C. (1963a) The control of meal size in the blood sucking bug, Rhodnius prolixus. J. Exp. Biol., 40, 741–50Google Scholar
Bennet-Clark, H. C. (1963b) Negative pressures produced in the pharyngeal pump of the blood-sucking bug, Rhodnius prolixus. J. Exp. Biol., 40, 223–9Google Scholar
Bennett, G. F., Fallis, A. M. and Campbell, A. G. (1972) The response of Simulium (EuSimulium) euryadminiculum Davies (Diptera: Simuliidae) to some olfactory and visual stimuli. Can J. Zool., 50, 793–800CrossRefGoogle Scholar
Bequaert, J. C. (1953) The Hippoboscidae or house-flies (Diptera) of mammals and birds. Part 1. Structure, physiology and natural history. Entomologia Americana, 32, 1–209Google Scholar
Bergman, D. K. (1996) Mouthparts and feeding mechanisms of haematophagous arthropods. In Wikel, S. K. (ed.), Immunology of Host-Ectoparasitic Arthropod Relationships. Wallingford: CABInternational, 30–61Google Scholar
Besansky, N. J. (1999) Complexities in the analysis of cryptic taxa within the genus Anopheles. Parasitologia, 41, 97–100Google ScholarPubMed
Bidlingmayer, W. L. (1994) How mosquitos see traps – role of visual responses. J. Am. Mosq. Control Assoc., 10, 272–9Google ScholarPubMed
Bidlingmayer, W. L., Day, J. F. and Evans, D. G. (1995) Effect of wind velocity on suction trap catches of some Florida mosquitoes. J. Am. Mosq. Control Assoc., 11, 295–301Google ScholarPubMed
Bidlingmayer, W. L. and Hem, D. G. (1979) Mosquito (Diptera: Culicidae) flight behaviour near conspicuous objects. Bull. Ent. Res., 69, 691–700CrossRefGoogle Scholar
Bidlingmayer, W. L. and Hem, D. G. (1980) The range of visual attraction and the effect of competitive visual attractants upon mosquito (Diptera: Culicidae) flight. Bull. Ent. Res., 70, 321–42CrossRefGoogle Scholar
Biessmann, H., Walter, M. F., Dimitratos, S. and Woods, D. (2002) Isolation of cDNA clones encoding putative odourant binding proteins from the antennae of the malaria-transmitting mosquito, Anopheles gambiae. Insect Mol. Biol., 11, 123–32CrossRefGoogle ScholarPubMed
Billingsley, P. B. (1990) The midgut ultrastructure of haematophagous arthropods. Ann. Rev. Ent., 35, 219–48CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1983) Ultrastructural changes in posterior midgut cells associated with blood-feeding in adult female Rhodnius prolixus Stal (Hemiptera: Reduviidae). Can J. Zool., 61, 1175–87CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1985) Cellular localisation of aminopeptidase in the midgut of Rhodnius prolixus Stal (Hemiptera: Reduviidae) during blood digestion. Cell and Tissue Research, 241, 421–8CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1986) The surface morphology of the midgut cells of Rhodnius prolixus Stal (Hemiptera: Reduviidae) during blood digestion. Acta Trop., 43, 355–66Google ScholarPubMed
Billingsley, P. F. and Downe, A. E. R. (1988) Ultrastructural localisation of cathepsin B in the midgut of Rhodnius prolixusStal (Hemiptera: Reduviidae) during blood digestion. International Journal of Insect Morphology and Embryology, 17, 295–302CrossRefGoogle Scholar
Billingsley, P. F. and Downe, A. E. R. (1989) Changes in the anterior midgut cells of adult female Rhodnius prolixus Stal. (Hemiptera: Reduviidae) after feeding. J. Med. Ent., 26, 104–8CrossRefGoogle Scholar
Billingsley, P. F. and Rudin, W. (1992) The role of the mosquito peritrophic membrane in bloodmeal digestion and infectivity of Plasmodium species. J. Parasit., 78, 430–40CrossRefGoogle ScholarPubMed
Bissonnette, E. Y., Rossignol, P. A. and Befus, A. D. (1993) Extracts of mosquito salivary gland inhibit tumour necrosis factor alpha. Parasite Immunology, 15, 27–33CrossRefGoogle ScholarPubMed
Bitkowska, E., Dzbenski, T. H., Szadziewska, M. and Wegner, Z. (1982) Inhibition of xenograft rejection reaction in the bug Triatoma infestans during infection with a protozoan, Tryapnosoma cruzi. J. Invert. Path., 40, 186–9CrossRefGoogle Scholar
Bize, P., Roulin, A. and Richner, H. (2003) Adoption as an offspring strategy to reduce ectoparasite exposure. Proc. R. Soc. Lond. B. Biol. Sci., 270 Suppl 1, 114–16CrossRefGoogle ScholarPubMed
Blackwell, A. (2000) Scottish biting midges: tourist attraction or deterrent?Antenna, 24, 144–50Google Scholar
Blackwell, A. and Page, S. (2003) Managing tourist health and Safety in the new millennium: global perspectives. In Managing Tourist Health and Safety in the New Millennium. Pergamon. 177–96CrossRefGoogle Scholar
Blandin, S., Moita, L. F., Kocher, T.et al. (2002) Reverse genetics in the mosquito Anopheles gambiae: targeted disruption of the Defensin gene. EMBO Report, 3, 852–6CrossRefGoogle ScholarPubMed
Boatin, B. A. (2003) The current state of the Onchocerciasis Control Programme in West Africa. Trop. Doct., 33, 209–14CrossRefGoogle ScholarPubMed
Boete, C., Paul, R. E. L. and Koella, J. C. (2002) Reduced efficacy of the immune melanization response in mosquitoes infected by malaria parasites. Parasitology, 125, 93–8CrossRefGoogle ScholarPubMed
Bos, H. J. and Laarman, J. J. (1975) Guinea pig lysine, cadaverine, and estradiol as attractants for the malaria mosquito Anopheles stephensi. Ent. Exp. Appl., 18, 161–72CrossRefGoogle Scholar
Bosch, O. J., Geier, M. and Boeckh, J. (2000) Contribution of fatty acids to olfactory host finding of female Aedes aegypti. Chemical Senses, 25, 323–30CrossRefGoogle ScholarPubMed
Bosio, C. F., Beaty, B. J. and Black, W. C. (1998) Quantitative genetics of vector competence for Dengue-2 virus in Aedes aegypti. Am. J. Trop. Med. Hyg., 59, 965–70CrossRefGoogle ScholarPubMed
Bosio, C. F., Fulton, R. E., Salasek, M. L., Beaty, B. J. and Black, W. C. T. (2000) Quantitative trait loci that control vector competence for dengue-2 virus in the mosquito Aedes aegypti. Genetics, 156, 687–98Google ScholarPubMed
Bossard, R. L. (2002) Speed and Reynolds number of jumping cat fleas (Siphonaptera: Pulicidae). Journal of the Kansas Entomological Society, 75, 52–4Google Scholar
Boulanger, N., Munks, R. J., Hamilton, J. V.et al. (2002) Epithelial innate immunity. A novel antimicrobial peptide with antiparasitic activity in the blood-sucking insect Stomoxys calcitrans. J. Biol. Chem., 277, 49921–6CrossRefGoogle ScholarPubMed
Boutros, M., Agaisse, H. and Perrimon, N. (2002) Sequential activation of signaling pathways during innate immune responses in Drosophila. Dev. Cell, 3, 711–22CrossRefGoogle ScholarPubMed
Bozza, M., Soares, M. B., Bozza, P. T.et al. (1998) The PACAP-type I receptor agonist maxadilan from sand fly saliva protects mice against lethal endotoxemia by a mechanism partially dependent on IL-10. Eur. J. Immunol., 28, 3120–73.0.CO;2-3>CrossRefGoogle ScholarPubMed
Bracken, G. K., Hanec, W. and Thorsteinson, A. J. (1962) The orientation of horseflies and deerflies (Tabanidae: Diptera). II. The role of some visual factors in the attractiveness of decoy silhouettes. Can. J. Zool., 40, 685–95CrossRefGoogle Scholar
Bracken, G. K. and Thorsteinson, A. J. (1965) The orientation behaviour of horse flies and deer flies (Tabanidae: Diptera). IV. The influence of some physical modifications of visual decoys on orientation of horse flies. Ent. Exp. Appl., 8, 314–18CrossRefGoogle Scholar
Bradbury, W. C. and Bennett, G. F. (1974) Behaviour of adult Simuliidae (Diptera). I. Response to colour and shape. Can. J. Zool., 52, 251–9CrossRefGoogle Scholar
Bradley, G. H. (1935) Notes on the southern buffalo gnat Eusimulium pecuarum (Riley) (Diptera: Simuliidae). Proc. Ent. Soc. Washington, 37, 60–4Google Scholar
Bradshaw, W. E. (1980) Blood-feeding and capacity for increase in the pitcher plant mosquito, Wyeomyia smithii. Environ. Entomol, 9, 86–9CrossRefGoogle Scholar
Bradshaw, W. E. and Holtzapfel, C. M. (1983) Life cycle strategies in Wyeomyia smithii: seasonal and geographic adaptations. In Brown, V. K. and Hodek, I. (eds.), Diapause and Life Cycle Strategies in Insects. The Hague: JunkGoogle Scholar
Brady, J. (1972) The visual responsiveness of the tsetse fly Glossina morsitans Westw. (Glossinidae) to moving objects: the effects of hunger, sex, host odour and stimulus characteristics. Bull. Ent. Res., 62, 257–79CrossRefGoogle Scholar
Brady, J. (1973) Changes in the probing responsiveness of starving tsetse flies (Glossina morsitans Westw.) (Diptera, Glossinidae). Bull. Ent. Res., 63, 247–55CrossRefGoogle Scholar
Brady, J. (1975) 'Hunger' in the tsetse fly: the nutritional correlates of behaviour. J. Insect Physiol., 21, 807–29CrossRefGoogle ScholarPubMed
Brady, J., Costantini, C., Sagnon, N., Gibson, G. and Coluzzi, M. (1997) The role of body odours in the relative attractiveness of different men to malarial vectors in Burkina Faso. Ann. Trop. Med. Parasit., 91, S121–S122CrossRefGoogle Scholar
Brady, J., Gibson, G. and Packer, M. J. (1989) Odour movement, wind direction and the problem of host finding by tsetse flies. Phys. Ent., 14, 369–380CrossRefGoogle Scholar
Brady, J., Griffiths, N. and Paynter, Q. (1995) Wind speed effects on odour source location by tsetse flies (Glossina). Phys. Ent., 20, 293–302CrossRefGoogle Scholar
Brady, J. and Shereni, A. (1988) Landing responses of the tsetse fly Glossina morsitans morsitans Weidemann and the stablefly Stomoxys calcitrans (L.). (Diptera: Glossinidae & Muscidae) to black and white patterns: a laboratory study. Bull. Ent. Res., 78, 301–11CrossRefGoogle Scholar
Braks, M. A. H., Scholte, E. J., Takken, W. and Dekker, T. (2000) Microbial growth enhances the attractiveness of human sweat for the malaria mosquito, Anopheles gambiae sensu stricto (Diptera: Culicidae). Chemoecology, 10, 129–34CrossRefGoogle Scholar
Braun, A., Hoffmann, J. A. and Meister, M. (1998) Analysis of the Drosophila host defense in domino mutant larvae, which are devoid of hemocytes. Proceedings of the National Academy of Sciences of the United States of America, 95, 14337–42CrossRefGoogle ScholarPubMed
Bray, R. S. (1963) The exo-erythrocytic phase of malaria parasites. Int. Rev. Trop. Med., 2, 41Google Scholar
Bray, R. S., McCrae, A. W. R. and Smalley, M. E. (1976) Lack of a circadian rhythm in the ability of the gametocytes of Plasmodium falciparum to infect Anopheles gambiae. Int. J. Parasit., 6, 399–401CrossRefGoogle ScholarPubMed
Brecher, G. and Wigglesworth, V. B. (1944) The transmission of Actinomyces rhodnii Erkison in Rhodnius prolixus Stal. (Hemiptera) as its influence on the growth of the host. Parasitol., 35, 220–4CrossRefGoogle Scholar
Breev, K. V. (1950) The behaviour of blood-sucking Diptera and warble flies when attacking reindeer and the responsive reactions of reindeer. [In Russian] Parasitologicheskii Sbornik, 12, 167–89
Brehelin, M. (1982) Comparative study of structure and function of blood cells from two Drosophila species. Cell and Tissue Research, 221, 607–15CrossRefGoogle ScholarPubMed
Brehelin, M. (1986) Insect haemocytes: a new classification to rule out the controversy. In Brehelin, M. (ed.), in Invertebrates. Berlin, Heidelberg, New York, Tokyo: Springer Verlag Press, 36–49Google Scholar
Brehelin, M., Zachary, D. and Hoffmann, J. A. (1978) A comparative ultrastructural study of blood cells from nine insect orders. Cell and Tissue Research, 195, 45–57CrossRefGoogle ScholarPubMed
Briegel, H., Hefti, A. and DiMarco, E. (2002) Lipid metabolism during sequential gonotrophic cycles in large and small female Aedes aegypti. J. Insect Physiol., 48, 547–54CrossRefGoogle ScholarPubMed
Briegel, H., Knusel, I. and Timmermann, S. E. (2001) Aedes aegypti: size, reserves, survival, and flight potential. Journal of Vector Ecology, 26, 21–31Google ScholarPubMed
Brook, M. L. (1985) The effect of allopreening on tick burdens of moulting eudyptid penguins. Auk, 102, 893Google Scholar
Browne, S. M. and Bennett, G. F. (1980) Colour and shape as mediators of host-seeking responses of simuliids and tabanids (Diptera) in the Tantramar marshes, New Brunswick, Canada. Can. J. Med. Entomol., 17, 58–62CrossRefGoogle Scholar
Browne, S. M. and Bennett, G. F. (1981) Response of mosquitoes (Diptera: Culicidae) to visual stimuli. J. Med. Ent., 6, 505–21CrossRefGoogle Scholar
Bruce, D. (1895) Preliminary Report on the Tsetse Fly Disease or Nagana, in Zululand. Durban: Bennett and Davis
Bruce, D. and Nabarro, D. (1903) Progress report on sleeping sickness in Uganda
Buchner, P. (1965) Endosymbiosis of Animals with Plant Microorganisms. New York: Wiley
Budd, L. T. (1999) DFID-Funded Tsetse and Trypanosomiasis Research and Development since 1980 (V. 2. Economic Analysis). London: Department for International Development
Bulet, P., Hetru, C., Dimarcq, J. L. and Hoffmann, D. (1999) Antimicrobial peptides in insects; structure and function. Developmental and Comparative Immunology, 23, 329–44CrossRefGoogle ScholarPubMed
Bungener, W. and Muller, G. (1976) Adharenz-Phänomene bei Trypanosoma congolense. Tropenmed. Parasitol., 27, 307–71Google Scholar
Burg, J. G., Knapp, F. W. and Silapanuntakul, S. (1993) Feeding Haematobia irritans (Diptera, Muscidae) adults through a nylon-reinforced silicone membrane. J. Med. Ent., 30, 462–6CrossRefGoogle ScholarPubMed
Burgess, I., Maunder, J. W. and Myint, T. T. (1983) Maintenance of the crab louse, Pthirus pubis, in the laboratory and behavioural studies using volunteers. Community Med., 5, 238–41Google Scholar
Burkett, D. A., Butler, J. F. and Kline, D. L. (1998) Field evaluation of colored light-emitting diodes as attractants for woodland mosquitoes and other diptera in north central Florida. J. Am. Mosq. Control Assoc., 14, 186–95Google ScholarPubMed
Burkhart, C. N., Stankiewicz, B. A., Pchalek, I., Kruge, M. A. and Burkhart, C. G. (1999) Molecular composition of the louse sheath. J. Parasit., 85, 559–61CrossRefGoogle ScholarPubMed
Burkot, T. R. (1988) Non-random host selection by anopheline mosquitoes. Parasitology Today, 4, 156–62CrossRefGoogle ScholarPubMed
Burkot, T. R., Narara, A., Paru, R., Graves, P. M. and Garner, P. (1989) Human host selection by anophelines: no evidence for preferential selection of malaria or microfilariae-infected individuals in a hyperendemic area. Parasitology, 98, 337–42CrossRefGoogle ScholarPubMed
Bursell, E. (1961) The behaviour of tsetse flies (Glossina swynnertoni Austen) in relation to problems of sampling. Proc. R. Ent. Soc. Lond. (A), 36, 9–20Google Scholar
Bursell, E. (1975) Substrates of oxidative metabolism in dipteran flight muscle. Comp. Biochem. Physiol., 52, 235–8Google ScholarPubMed
Bursell, E. (1977) Synthesis of proline by fat body of the tsetse fly (Glossina morsitans). Metabolic pathways. Insect Biochem., 7, 427–34CrossRefGoogle Scholar
Bursell, E. (1981) Energetics of haematophagous arthropods: influence of parasites. Parasitology, 82, 107–10Google Scholar
Bursell, E. (1984) Effects of host odour on the behaviour of tsetse. Insect Sci. Applic., 5, 345–9Google Scholar
Bursell, E. (1987) The effect of wind-borne odours on the direction of flight in tsetse flies, Glossina spp. Phys. Ent., 12, 149–56CrossRefGoogle Scholar
Bursell, E. and Taylor, P. (1980) An energy budget for Glossina (Diptera: Glossinidae). Bull. Ent. Res., 70, 187–96CrossRefGoogle Scholar
Burton, R. (1860) The Lake Regions of Central Africa. London: Longman
Butt, T. M. and Shields, K. S. (1996) The structure and behavior of gypsy moth (Lymantria dispar) hemocytes. J. Invert. Path., 68, 1–14CrossRefGoogle ScholarPubMed
Buxton, P. A. (1930) The biology of a blood-sucking bug, Rhodnius prolixus. Trans. R. Soc. Trop. Med. Hyg., 78, 227–36Google Scholar
Buxton, P. A. (1947) The Louse. London: Edward Arnold
Buxton, P. A. (1948) Experiments with lice and fleas. I. The baby mouse. Parasitol., 39, 119–24CrossRefGoogle Scholar
Canyon, D. V., Hii, J. L. K. and Muller, R. (1998) Multiple host-feeding and biting persistence of Aedes aegypti. Ann. Trop. Med. Parasit., 92, 311–16CrossRefGoogle ScholarPubMed
Cappello, M., Li, S., Chen, X. O.et al. (1998) Tsetse thrombin inhibitor: bloodmeal-induced expression of an anticoagulant in salivary glands and gut tissue of Glossina morsitans morsitans. Proceedings of the National Academy of Sciences of the United States of America, 95, 14290–5CrossRefGoogle ScholarPubMed
Carde, R. T. (1996) Odour plumes and odour-mediated flight in insects. In Bock, G. R. and G. Cardew (eds.), Olfaction in Mosquito-Host Interactions. Chichester: Wiley Ciba Foundation Symposium 200, 54–70Google Scholar
Carwardine, S. L. and Hurd, H. (1997) Effects of Plasmodium yoelii nigeriensis infection on Anopheles stephensi egg development and resorption. Med. Vet. Entomol, 11, 265–9CrossRefGoogle ScholarPubMed
Castanera, M. B., Aparicio, J. P. and Gurtler, R. E. (2003) A stage-structured stochastic model of the population dynamics of Triatoma infestans, the main vector of Chagas disease. Ecological Modelling, 162, 33–53CrossRefGoogle Scholar
Caterino, M. S., Cho, S. and Sperling, F. A. H. (2000) The current state of insect molecular systematics: a thriving Tower of Babel. Ann. Rev. Ent., 45, 1–54CrossRefGoogle Scholar
Cavanaugh, D. C. (1971) Specific effect of temperature upon transmission of the plague bacillus by the oriental rat flea, Xenopsylla cheopis. Am. J. Trop. Med. Hyg., 20, 264–73CrossRefGoogle ScholarPubMed
Chadee, D. D. and Beier, J. C. (1997) Factors influencing the duration of blood-feeding by laboratory-reared and wild Aedes aegypti (Diptera: Culicidae) from Trinidad, West Indies. Ann. of Trop. Med. Parasit., 91, 199–207CrossRefGoogle ScholarPubMed
Chadee, D. D., Beier, J. C. and Martinez, R. (1996) The effect of the cibarial armature on blood meal haemolysis of four anopheline mosquitoes. Bull. Ent. Res., 86, 351–4CrossRefGoogle Scholar
Chagas, C. (1909) Ueber eine neue Trypanosomiasis des Menschen. Memorias Institut Oswaldo Cruz, 1, 159–218CrossRef
Challier, A., Eyraud, M., Lafaye, A. and Laveissiere, C. (1977) Amelioration due rendement du piege biconique pour glossines (Diptera, Glossinidae) par l'emploi d'un cone inférieur bleu. Cah. ORSTOM, Ser. Entomol. Med. Parasitol., 15, 283–6Google Scholar
Champagne, D. E., Nussenzveig, R. H. and Ribeiro, J. M. C. (1995a) purification, partial characterization, and cloning of nitric oxide-carrying heme-proteins (nitrophorins) from salivary-glands of the bloodsucking insect Rhodnius prolixus. J. Biol. Chem., 270, 8691–5CrossRefGoogle Scholar
Champagne, D. E. and Ribeiro, J. M. C. (1994) Sialokinin-I and Sialokinin-Ii – Vasodilatory Tachykinins from the Yellow-Fever Mosquito Aedes aegypti. Proceedings of the National Academy of Sciences of the United States of America, 91, 138–42CrossRefGoogle ScholarPubMed
Champagne, D. E., Smartt, C. T., Ribeiro, J. M. C. and James, A. A. (1995b) The salivary gland-specific apyrase of the mosquito Aedes aegypti is a member of the 5ʹ-nucleotidase family. Proceedings of the National Academy of Sciences of the United States of America, 92, 694–8CrossRefGoogle Scholar
Chang, Y. H. and Judson, C. L. (1977) The role of isoleucine in differential egg production by the mosquito Aedes aegypti Linnaeus (Diptera: Culicidae) following feeding on human or guinea pig blood. Comp. Biochem. Physiol. A., 57, 23–8CrossRefGoogle Scholar
Chapman, R. F. (1961) Some experiments to determine the methods used in host finding by tsetse flies, Glossina medicorum. Bull. Ent. Res., 52, 83–97CrossRefGoogle Scholar
Chapman, R. F. (1982) Chemoreception: the significance of receptor numbers. Adv. Insect Physiol., 16, 247–356CrossRefGoogle Scholar
Charlab, R., Rowton, E. D. and Ribeiro, J. M. C. (2000) The salivary adenosine deaminase from the sand fly Lutzomyia longipalpis. Experimental Parasitology, 95, 45–53CrossRefGoogle ScholarPubMed
Charlwood, J. D., Billingsley, P. F. and Hoc, T. Q. (1995a) Mosquito-mediated attraction of female European but not African mosquitos to hosts. Ann. Trop. Med. Parasit., 89, 327–9CrossRefGoogle Scholar
Charlwood, J. D., Smith, T., Kihonda, J. (1995b) Density-independent feeding success of malaria vectors (Diptera, Culicidae) in Tanzania. Bull. Ent. Res., 85, 29–35CrossRefGoogle Scholar
Chen, C. C. and Chen, C. S. (1995) Brugia pahangi: effects of melanization on the uptake of nutrients by microfilariae in vitro. Experimental Parasitology, 81, 72–8CrossRefGoogle ScholarPubMed
Chen, C. C. and Laurence, B. R. (1985) An ultrastructural study on the encapsulation of microfilariae of Brugia pahangi in the haemocoel of Anopheles quadrimaculatus. Int. J. Parasitol, 15, 421–8CrossRefGoogle ScholarPubMed
Chikilian, M. L., Bradley, T. J., Nayar, J. K., Cashclark, C. E. and Knight, J. W. (1995) Ultrastructure of the intracellular melanization of Brugia malayi (Buckley) (Nematoda, Filarioidea) in the thoracic muscles of Anopheles quadrimaculatus (Say) (Diptera, Culicidae). International Journal of Insect Morphology and Embryology, 24, 83–92CrossRefGoogle Scholar
Chikilian, M. L., Bradley, T. J., Nayar, J. K. and Knight, J. W. (1994) Ultrastructural comparison of extracellular and intracellular encapsulation of Brugia malayi in Anopheles quadrimaculatus. Journal of Parasitology, 80, 133–40CrossRefGoogle ScholarPubMed
Choe, K. M., Werner, T., Stoven, S., Hultmark, D. and Anderson, K. V. (2002) Requirement for a peptidoglycan recognition protein (PGRP) in relish activation and antibacterial immune responses in Drosophila. Science, 296, 359–62CrossRefGoogle ScholarPubMed
Christensen, B. M. (1978) Dirofilaria immitis: effects on the longevity of Aedes trivittatus. Experimental Parasitology, 44, 116–23CrossRefGoogle Scholar
Christensen, B. M., Forton, K. F., Lafond, M. M. and Grieve, R. B. (1987) Surface changes on Brugia pahangi microfilariae and their association with immune evasion in Aedes aegypti. J. Invert. Pathol., 49, 14–18CrossRefGoogle ScholarPubMed
Christensen, B. M. and LaFond, M. M. (1986) Parasite induced suppression of the immune response in Aedes aegypti by Brugia pahangi. J. Parasit., 72, 216–19CrossRefGoogle ScholarPubMed
Christophides, G. K., Zdobnov, E., Barillas-Mury, C.et al. (2002) Immunity-related genes and gene families in Anopheles gambiae. Science, 298, 159–65CrossRefGoogle ScholarPubMed
Ciurea, I. and Dinulescu, G. (1924) Ravages causes par la mouchede Goloubatz en Roumanie; ses attaques contre les animaux et contre l'homme. Ann. Trop. Med. Parasit., 18, 323–42CrossRefGoogle Scholar
Clay, T. (1963) A new species of Haematomyzus Piaget (Phthiraptera, Insecta). Proc. Zool. Soc. Lond., 141, 153–61CrossRefGoogle Scholar
Clifford, C. M., Bell, J. F., Moore, G. J. and Raymond, C. (1967) Effects of limb disability of lousiness in mice. IV. Evidence of genetic factors in susceptibility to Polyplax serrata. Experimental Parasitology, 20, 56–67CrossRefGoogle ScholarPubMed
Coatney, G. R., Collins, W. E., McWilson, W. and Contacos, P. G. (1971) The Primate Malarias. Bethesda, MD: NIAID
Cockerell, T. D. A. (1918) New species of North American fossil beetles, cockroaches and tsetse flies. Proc. U. S. Natn. Mus., 54, 301–11CrossRefGoogle Scholar
Coetzee, M., Craig, M. and Sueur, D. (2000) Distribution of African malaria mosquitoes belonging to the Anopheles gambiae complex. Parasitology Today, 16, 74–7CrossRefGoogle ScholarPubMed
Colless, D. H. and Chellapah, W. T. (1960) Effects of body weight and size of blood meal upon egg production in Aedes aegypti (Linnaeus) (Diptera, Culicidae). Ann. Trop. Med. Parasit., 54, 475–82CrossRefGoogle Scholar
Collins, F. H., Sakai, R. K., Vernick, K. D.et al. (1986) Genetic selection of a refractory strain of the malaria vector Anopheles gambiae. Science, 234, 607–10CrossRefGoogle ScholarPubMed
Colman, R. W. (2001) Hemostasis and Thrombosis: Basic Principles and Clinical Practice. London: Lippincott, Williams and Wilkins
Coluzzi, M., Concetti, A. and Ascoli, F. (1982) Effect of cibarial armature of mosquitoes (Diptera: Culicidae) on blood-meal haemolysis. J. Insect Physiol., 28, 885–8CrossRefGoogle Scholar
Coluzzi, M., Sabatini, A., Torre, A., Di Deco, M. A. and Petrarca, V. (2002) A polygene chromosome analysis of the Anopheles gambiae species complex. Science, 298, 1415–18CrossRefGoogle Scholar
Colvin, J., Brady, J. and Gibson, G. (1989) Visually-guided, upwind turning behaviour of free-flying tsetse flies in odour-laden wind: a wind-tunnel study. Phys. Ent., 14, 31–9CrossRefGoogle Scholar
Colyer, C. N. and Hammond, C. O. (1968) Flies of the British Isles. London: Frederick Warne
Compton-Knox, P. and Hayes, K. L. (1972) Attraction of Tabanus spp. (Diptera: Tabanidae) to traps baited with carbon dioxide and other chemicals. Environ. Entomal., 1, 323–6CrossRefGoogle Scholar
Cook, S. P. and McCleskey, E. W. (2002) Cell damage excites nociceptors through release of cytosolic ATP. Pain, 95, 41–7CrossRefGoogle ScholarPubMed
Cornford, E. M., Freeman, B. J. and MacInnis, A. J. (1976) Physiological relationships and circadian periodicities in rodent trypanosomes. Trans. R. Soc. Trop. Med. Hyg., 70, 238–43CrossRefGoogle ScholarPubMed
Croft, S. L., East, J. S. and Molyneux, D. H. (1982) Antitrypanosomal factor in the haemolymph of Glossina. Acta Trop., 39, 293–302Google Scholar
Cross, M. L., Cupp, E. W. and Enriquez, F. J. (1994) Modulation of murine cellular immune-responses and cytokines by salivary-gland extract of the black fly Simulium vittatum. Tropical Medicine and Parasitology, 45, 119–124Google ScholarPubMed
Cupp, E. W., Cupp, M. S., Ribeiro, J. M. C. and Kunz, S. E. (1998a) Blood-feeding strategy of Haematobia irritans (Diptera: Muscidae). J. Med. Ent., 35, 591–5CrossRefGoogle Scholar
Cupp, E. W. and Stokes, G. M. (1976) Feeding patterns of Culex salinarius Coquillett in Jefferson parish, Louisiana. Mosq. News, 36, 332–5Google Scholar
Cupp, M. S., Ribeiro, J. M. C., Champagne, D. E. and Cupp, E. W. (1998b) Analyses of cDNA and recombinant protein for a potent vasoactive protein in saliva of a blood-feeding black fly, Simulium vittatum. J. Exp. Biol., 201, 1553–61Google Scholar
Dale, C., Young, S. A., Haydon, D. T. and Welburn, S. C. (2001) The insect endosymbiont Sodalis glossinidius utilizes a type III secretion system for cell invasion. Proceedings of the National Academy of Sciences of the United States of America, 98, 1883–8CrossRefGoogle ScholarPubMed
Dan, A., Pereira, M. H., Pesquero, J. L., Diotaiuti, L. and Beirao, P. S. L. (1999) Action of the saliva of Triatoma infestans (Heteroptera: Reduviidae) on sodium channels. J. Med. Entomol., 36, 875–9CrossRefGoogle ScholarPubMed
Daniel, T. L. and Kingsolver, J. G. (1983) Feeding strategy and the mechanics of blood sucking in insects. J. Theor. Biol., 105, 661–72CrossRefGoogle ScholarPubMed
David, C. T., Kennedy, J. S., Ludlow, A. R., Perry, J. N. and Wall, C. (1982) A reappraisal of insect flight towards a distant point source of wind-borne odor. J. Chem. Ecol., 8, 1207–15CrossRefGoogle ScholarPubMed
Davidson, G. and Draper, C. C. (1953) Field studies of some of the basic factors concerned in the transmission of malaria. Trans. R. Soc. Trop. Med. Hyg., 47, 522–35CrossRefGoogle ScholarPubMed
Davis, E. E. (1984) Regulation of sensitivity in the peripheral chemoreceptor systems for host-seeking behaviour by a haemolymph-borne factor in Aedes aegypti. J. Insect Physiol., 30, 179–83CrossRefGoogle Scholar
Davis, E. E. and Sokolove, P. G. (1975) Temperature response of the antennal receptors in the mosquito, Aedes aegypti. J. Comp. Physiol., 96, 223–36CrossRefGoogle Scholar
Day, J. F. and Edman, J. D. (1983) Malaria renders mice susceptible to mosquito feeding when gametocytes are most infective. J. Parasitol., 69, 163–70CrossRefGoogle ScholarPubMed
Day, J. F. and Edman, J. D. (1984a) The importance of disease induced changes in mammalian body temperature to mosquito blood feeding. Comp. Biochem. Physiol., 77, 447–52CrossRefGoogle Scholar
Day, J. F. and Edman, J. D. (1984b) Mosquito engorgement on normally defensive hosts depends on host activity patterns. J. Med. Ent., 21, 732–40CrossRefGoogle Scholar
Azambuja, P., Guimares, J. A. and Garcia, E. S. (1983) Haemolytic factor from the crop of Rhodnius prolixus: evidence and partial characterisation. J. Insect Physiol., 29, 833–7CrossRefGoogle Scholar
Gregorio, E., Spellman, P. T., Tzou, P., Rubin, G. M. and Lemaitre, B. (2002) The Toll and Imd pathways are the major regulators of the immune response in Drosophila. EMBOJ., 21, 2568–79CrossRefGoogle ScholarPubMed
Jong, R. and Knols, B. G. J. (1995) Selection of biting sites on man by 2 malaria mosquito species. Experientia, 51, 80–4CrossRefGoogle Scholar
DeFoliart, G. R., Grimstad, P. R. and Watts, D. M. (1987) Advances in mosquito-borne arbovirus/vector research. Ann. Rev. Ent., 32, 479–505CrossRefGoogle ScholarPubMed
Dei Cas, E., Maurois, P., Landau, I.et al. (1980) Morphologie et infectivite des gametocytes de Plasmodium inui. Annales Parasit., 55, 621–33Google Scholar
Dekker, T. and Takken, W. (1998) Differential responses of mosquito sibling species Anopheles arabiensis and An. quadriannulatus to carbon dioxide, a man or a calf. Med. Vet. Entomol., 12, 136–40CrossRefGoogle ScholarPubMed
Dekker, T., Takken, W. and Braks, M. A. H. (2001) Innate preference for host-odor blends modulates degree of anthropophagy of Anopheles gambiae sensu lato (Diptera: Culicidae). J. Med. Ent., 38, 868–71CrossRefGoogle Scholar
Dekker, T., Takken, W., Knols, B. G. J.et al. (1998) Selection of biting sites on a human host by Anopheles gambiae s. s., An. arabiensis and An. quadriannulatus. Entomologia Experimentalis et Applicata, 87, 295–300CrossRefGoogle Scholar
Denotter, C. J., Tchicaya, T. and Schutte, A. M. (1991) Effects of age, sex and hunger on the antennal olfactory sensitivity of tsetse-flies. Phys. Ent., 16, 173–82Google Scholar
Desquesnes, M. and Dia, M. L. (2003) Trypanosoma vivax: mechanical transmission in cattle by one of the most common African tabanids, Atylotus agrestis. Experimental Parasitology, 103, 35–43CrossRefGoogle ScholarPubMed
Dethier, V. G. (1954) Notes on the biting response of tsetse flies. Am. J. Trop. Med. Hyg., 3, 160–71CrossRefGoogle ScholarPubMed
Detinova, T. S. (1962) Age Grouping Methods in Diptera of Medical Importance. Geneva: World Health Organization
Dias, J. C., Silveira, A. C. and Schofield, C. J. (2002) The impact of Chagas disease control in Latin America: a review. Mem. Inst. Oswaldo Cruz, 97, 603–12CrossRefGoogle ScholarPubMed
Diatta, M., Spiegel, A., Lochouarn, L. and Fontenille, D. (1998) Similar feeding preferences of Anopheles gambiae and An. arabiensis in Senegal. Trans. R. Soc. Trop. Med. Hyg., 92, 270–2CrossRefGoogle ScholarPubMed
Dickerson, G. and Lavoipierre, M. M. J. (1959) Studies on the methods of feeding blood-sucking arthropods. II. The method of feeding adopted by the bed-bug (Cimex lectularius) when obtaining a blood meal from the mammalian host. Ann. Trop. Med. Parasit., 53, 347–57CrossRefGoogle Scholar
Dimopoulos, G. (2003) Insect immunity and its implication in mosquito-malaria interactions. Cell Microbiol., 5, 3–14CrossRefGoogle ScholarPubMed
Dimopoulos, G., Christophides, G. K., Meister, S.et al. (2002) Genome expression analysis of Anopheles gambiae: responses to injury, bacterial challenge, and malaria infection. Proceedings of the National Academy of Sciences of the United States of America, 99, 8814–19CrossRefGoogle ScholarPubMed
Dimopoulos, G., Seeley, D., Wolf, A. and Kafatos, F. C. (1998) Malaria infection of the mosquito Anopheles gambiae activates immune-responsive genes during critical transition stages of the parasite life cycle. EMBO J., 17, 6115–23CrossRefGoogle ScholarPubMed
Dobson, A. P. (1988) Parasite-induced changes in host behaviour, Q. Rev. Biol., 63(4), 140–65
Downes, J. A. (1970) The ecology of blood-sucking diptera: an evolutionary perspective. In Fallis, A. M. (ed.), and Physiology of Parasites. Toronto: University of Toronto PressGoogle Scholar
Downs, C. M., Theberge, J. B. and Smith, S. M. (1986) The influence of insects on the distribution, microhabitat choice, and behaviour of the Burwash caribou herd. Can. J. Zool., 64, 622–9CrossRefGoogle Scholar
Dryden, M. W. (1989) Host association, on-host longevity and egg production of Ctenocephalides felis felis. Vet Parasitol., 34, 117–22CrossRefGoogle ScholarPubMed
Dujardin, J. C., Banuls, A. L., Llanos-Cuentas, A.et al. (1995) Putative Leishmania hybrids in the Eastern Andean valley of Huanuco, Peru. Acta Trop., 59, 293–307CrossRefGoogle ScholarPubMed
Duncan, P. and Vigne, N. (1979) The effect of group size in horses on the rate of attacks by blood-sucking flies. Animal Behaviour, 27, 623–5CrossRefGoogle Scholar
Dunnet, G. M. (1970) Siphonaptera (Fleas). In CSIRO (ed.) The Insects of Australia. Melbourne: Melbourne University Press
Durvasula, R. V., Gumbs, A., Panackal, A.et al. (1997) Prevention of insect-borne disease: an approach using transgenic symbiotic bacteria. Proceedings of the National Academy of Sciences of the United States of America, 94, 3274–8CrossRefGoogle ScholarPubMed
Duval, J., Rajaonarivelo, E. and Rabenirainy, L. (1974) Ecologie de Styloconops spinosifrons (Carter, 1921) (Diptera, Ceratopogonidae) sur les plages de la côte Est de Madagascar. Cahiers ORSTOM, 12, 245–58Google Scholar
Dye, C. (1992) The analysis of parasite transmission by bloodsucking insects. Ann. Rev. Ent., 37, 1–19CrossRefGoogle ScholarPubMed
East, J., Molyneux, D. H., Maudlin, I. and Dukes, P. (1983) Effect of Glossina haemolymph on salivarian trypanosomes in vitro. Annals of Tropical Medicine and Parasitology, 77, 97–9CrossRefGoogle ScholarPubMed
Edman, J. D. (1974) Host-feeding patterns of Florida mosquitoes III Culex (Culex) and Culex (NeoCulex). J. Med. Ent., 11, 95–104CrossRefGoogle Scholar
Edman, J. D., Day, J. F. and Walker, E. (1985) Vector-host interplay: factors affecting disease transmission. In Lounibos, L. P., Rey, R. and Frank, J. H. (eds.), of Mosquitoes: Proceedings of a Workshop. New Beach, Florida: Florida Medical Entomology LaboratoryGoogle Scholar
Edman, J. D. and Kale, H. W. (1971) Host behaviour: its influence on the feeding success of mosquitoes. Ann. Ent. Soc. Am., 64, 513–16CrossRefGoogle Scholar
Edman, J. D. and Spielman, A. (1988) Blood feeding by vectors: physiology, ecology, behaviour and vertebrate defence. In Monath, T. P. (ed.), The Arboviruses: Epidemiology and Ecology. Baton Rouge: C.R.C. Press, Vol. 1Google Scholar
Edman, J. D. and Taylor, D. J. (1968) Culex nigripalpus: seasonal shift in the bird-mammal feeding ratio in a mosquito vector of human encephalitis. Science, 161, 67–8CrossRefGoogle Scholar
Edman, J. D., Webber, L. A. and Kale, H. W. (1972) Effect of mosquito density on the interrelationship of host behavior and mosquito feeding success. Am. J. Trop. Med. Hyg., 21, 487–91CrossRefGoogle ScholarPubMed
Eichler, D. A. (1973) Studies on Onchocerca gutterosa (Neumann, 1910) and its development in Simulium ornatum (Meigen, 1818). 3. Factors affecting the development of the parasite in its vector. J. Helm., 47, 73–88CrossRefGoogle Scholar
Eiras, A. E. and Jepson, P. C. (1994) Responses of female Aedes aegypti (Diptera, Culicidae) to host odors and convection currents using an olfactometer bioassay. Bull. Ent. Res., 84, 207–11CrossRefGoogle Scholar
Elkinton, J. S., Schal, C., Ono, T. and Carde, R. T. (1987) Pheromone puff trajectory and upwind flight of male gypsy moths in a forest. Phys. Ent., 12, 399–406CrossRefGoogle Scholar
Ellis, D. S. and Evans, D. A. (1977) Passage of Trypanosoma brucei rhodesiense through the peritrophic membrane of Glossina morsitans morsitans. Nature, 267, 834–5CrossRefGoogle ScholarPubMed
Elrod-Erickson, M., Mishra, S. and Schneider, D. (2000) Interactions between the cellular and humoral immune responses in Drosophila. Current Biology, 10, 781–4CrossRefGoogle ScholarPubMed
Elsen, P., Amoudi, M. A. and Leclercq, M. (1990) 1st record of Glossina fuscipes fuscipes Newstead, 1910 and Glossina morsitans submorsitans Newstead, 1910 in Southwestern Saudi-Arabia. Annales De La Societe Belge De Medecine Tropicale, 70, 281–7Google Scholar
Emmerson, K. C., Kim, K. C. and Price, R. D. (1973) Lice. In Flynn, R. J (ed.), Parasites of Laboratory Animals. Ames, Iowa: Iowa State University PressGoogle Scholar
Escalante, A. A. and Ayala, F. J. (1995) Evolutionary origin of Plasmodium and other apicomplexa based on ribosomal-RNAgenes. Proceedings of the National Academy of Sciences of the United States of America, 92, 5793–7CrossRefGoogle Scholar
Esseghir, S., Ready, P. D., KillickKendrick, R. and BenIsmail, R. (1997) Mitochondrial haplotypes and phylogeography of Phlebotomus vectors of Leishmania major. Insect Mol. Biol., 6, 211–25CrossRefGoogle ScholarPubMed
Evans, G. O. (1950) Studies on the bionomics of the sheep ked, Melophagus ovinus L., in West Wales. Bull. Ent. Res., 40, 459–78CrossRefGoogle Scholar
Ewert, A. (1965) Comparative migration of microfilariae and development of Brugia pahangi in various mosquitoes. Am. J. Trop. Med. Hyg., 14, 254–9CrossRefGoogle ScholarPubMed
Falleroni, D. (1927) Per la soluzione del problema malarico italiano. Riv. Malariol., 6, 344–409Google Scholar
Fallis, A. M., Bennett, G. F., Griggs, G. and Allen, T. (1967) Collecting Simulium venustum female in fan traps and on silhouettes with the aid of carbon dioxide. Can. J. Zool., 45, 1011–17CrossRefGoogle ScholarPubMed
Fallis, A. M. and Raybould, J. N. (1975) Response of two African simuliids to silhouettes and carbon dioxide. J. Med. Entomol., 12, 349–51CrossRefGoogle ScholarPubMed
Fallis, A. M. and Smith, S. M. (1964) Ether extract from birds and CO2 as attractants for some ornithophilic simuliids. Can. J. Zool., 42, 723–30CrossRefGoogle Scholar
Farkas, S. R. and Shorey, H. H. (1972) Chemical trail-following by flying insects: a mechanism for orientation to a distant odor source. Science (Washington, D.C), 178, 67–8CrossRefGoogle ScholarPubMed
Farmer, J., Maddrell, S. H. P. and Spring, J. H. (1981) Absorption of fluid by the midgut of Rhodnius. J. Exp. Biol., 94, 301–16Google Scholar
Faust, E. C., Russel, P. F. and Jung, R. C. (1977) Craig and Fausts Clinical Parasitology. Philadelphia: Lea and Febiger
Favia, G., Torre, A., Bagayoko, M.et al. (1997) Molecular identification of sympatric chromosomal forms of Anopheles gambiae and further evidence of their reproductive isolation. Insect Mol. Biol., 6, 377–83CrossRefGoogle ScholarPubMed
Feingold, B. F. and Benjamini, E. (1961) Allergy to flea bites: clinical and experimental observations. Ann. Allerg., 19, 1274–89Google ScholarPubMed
Ferdig, M. T., Beerntsen, B. T., Spray, F. J., Li, J. and Christensen, B. M. (1993) Reproductive costs associated with resistance in a mosquito-filarial worm system. Am. J. Trop. Med. Hyg., 49, 756–62CrossRefGoogle Scholar
Ferguson, H. M. and Read, A. F. (2002a) Genetic and environmental determinants of malaria parasite virulence in mosquitoes. Proc. R. Soc. Lond. B Biol. Sciences, 269, 1217–24CrossRefGoogle Scholar
Ferguson, H. M. and Read, A. F. (2002b) Why is the effect of malaria parasites on mosquito survival still unresolved?Trends in Parasitology, 18, 256–61CrossRefGoogle Scholar
Ferrari, J., Muller, C. B., Kraaijeveld, A. R. and Godfray, H. C. (2001) Clonal variation and covariation in aphid resistance to parasitoids and a pathogen. Evolution, 55, 1805–1814CrossRefGoogle Scholar
Ferris, G. R. (1931) The louse of elephants Haematomyzus elephantis Piaget (Mallophaga: Haematomyzidae). Parasitol., 23, 112–27CrossRefGoogle Scholar
Flores, G. B. and Lazzari, C. R. (1996) The role of the antennae in Triatoma infestans: Orientation towards thermal sources. J. Insect Physiol., 42, 433–40CrossRefGoogle Scholar
Foley, E. and Farrell, P. H. (2003) Nitric oxide contributes to induction of innate immune responses to gram-negative bacteria in Drosophila. Genes Dev., 17, 115–25CrossRefGoogle ScholarPubMed
Foster, W. A. (1976) Male sexual maturation of the tsetse flies Glossina morsitans Westwood and G. austeni Newstead (Diptera: Glossinidae) in relation to feeding. Bull. Ent. Res., 66, 389–99CrossRefGoogle Scholar
Fox, A. N., Pitts, R. J., Robertson, H. M., Carlson, J. R. and Zwiebel, L. J. (2001) Candidate odorant receptors from the malaria vector mosquito Anopheles gambiae and evidence of down-regulation in response to blood feeding. Proceedings of the National Academy of Sciences of the United States of America, 98, 14693–7CrossRefGoogle ScholarPubMed
Francischetti, I. M. B., Ribeiro, J. M. C., Champagne, D. and Andersen, J. (2000) Purification, cloning, expression, and mechanism of action of a novel platelet aggregation inhibitor from the salivary gland of the blood-sucking bug, Rhodnius prolixus. J. Biol. Chem., 275, 12639–50CrossRefGoogle ScholarPubMed
Francischetti, I. M. B., Valenzuela, J. G. and Ribeiro, J. M. C. (1999) Anophelin: kinetics and mechanism of thrombin inhibition. Biochemistry, 38, 16678–85CrossRefGoogle ScholarPubMed
Fredeen, F. J. H. (1961) A trap for studying the attacking behaviour of black flies Simulium articum Mall. Can. Entomol., 93, 73–8CrossRefGoogle Scholar
Freier, J. E. and Friedman, S. (1976) Effect of host infection with Plasmodium gallinaceum on the reproductive capacity of Aedes aegypti. J. Invert. Pathol., 28, 161–6CrossRefGoogle ScholarPubMed
Friend, W. G. (1978) Physical factors affecting the feeding responses of Culiseta inornata to ATP, sucrose and blood. Ann. Ent. Soc. Am., 71, 935–40CrossRefGoogle Scholar
Friend, W. G. and Smith, J. J. B. (1971) Feeding in Rhodnius prolixus: mouthpart activity and salivation and their correlation with changes of electrical resistance. J. Insect Physiol., 17, 233–43CrossRefGoogle Scholar
Friend, W. G. and Smith, J. J. B. (1975) Feeding in Rhodnius prolixus: increasing sensitivity to ATP during prolonged food deprivation. J. Insect Physiol., 21, 1081–4CrossRefGoogle ScholarPubMed
Friend, W. G. and Smith, J. J. B. (1977) Factors affecting feeding by blood-sucking insects. Ann. Rev. Ent., 22, 309–31CrossRefGoogle Scholar
Friend, W. G. and Stoffolano, J. G. (1984) Feeding responses of the horsefly, Tabanus nigrovittatus, to physical factors, ATP analogues and blood fractions. Phys. Ent., 9, 395–402CrossRefGoogle Scholar
Fu, H., Leake, C. J., Mertens, P. P. and Mellor, P. S. (1999) The barriers to bluetongue virus infection, dissemination and transmission in the vector, Culicoides variipennis (Diptera: Ceratopogonidae). Arch. Virol., 144, 747–61CrossRefGoogle Scholar
Gad, A. M., Maier, W. A. and Piekorski, G. (1979) Pathology of Anopheles stephensi after infection with Plasmodium berghei berghei. I. Mortality rate. Z. Parasit., 60, 249–61CrossRefGoogle ScholarPubMed
Gade, G. and Auerswald, L. (2002) Beetles' choice: proline for energy output: control by AKHs. Comp. Biochem. Physiol. B., 132, 117–29CrossRefGoogle ScholarPubMed
Galun, R. (1966) Feeding stimulants of the rat flea Xenopsylla cheopis Roth. Life Sci., 5, 1335–42CrossRefGoogle Scholar
Galun, R. (1986) Diversity of phagostimulants used for recognition of blood meal by haematophagous arthropods. In Borovsky, D. and Spielman, A. (eds.), Host-Regulated Development Mechanisms in Vector Arthropods. Florida: IFAS, University of FloridaGoogle Scholar
Galun, R. (1987) The evolution of purinergic receptors involved in recognition of a blood meal by haematophagous insects. Mem. Inst. Oswaldo Cruz, 82, 5–9CrossRefGoogle Scholar
Galun, R., Avidor, Y. and Bar-Zeev, M. (1963) Feeding response in Aedes aegypti: stimulation by adnosine triphosphate. Science, 124, 1674–5CrossRefGoogle Scholar
Galun, R., Friend, W. G. and Nudelman, S. (1988) Purinergic reception by culicine mosquitoes. J. Comp. Physiol. A., 163, 665–70CrossRefGoogle ScholarPubMed
Galun, R. and Kabayo, J. P. (1988) Gorging response of Glossina palpalis palpalis to ATP analogues. Phys. Ent., 13, 419–23CrossRefGoogle Scholar
Galun, R., Koontz, L. C. and Gwadz, R. W. (1985) Engorgement response of anopheline mosquitoes to blood fractions and artificial solutions. Phys. Ent., 10, 145–9CrossRefGoogle Scholar
Galun, R., Vardimonfriedman, H. and Frankenburg, S. (1993) Gorging response of culicine mosquitos (Diptera, Culicidae) to blood fractions. J. Med. Ent., 30, 513–17CrossRefGoogle Scholar
Garcia, R. and Radovsky, F. J. (1962) Haematophagy by two non-biting myscid flies and its relationship to tabanid feeding. Can. Entomol., 94, 1110–16CrossRefGoogle Scholar
Gardiner, E. M. and Strand, M. R. (1999) Monoclonal antibodies bind distinct classes of hemocytes in the moth Pseudoplusia includens. J Insect Physiol., 45, 113–26CrossRefGoogle ScholarPubMed
Garms, R., Walsh, J. F. and Davies, J. B. (1979) Studies on the reinvasion of the Onchocerciasis Control Programme in the Volta River Basin by Simulium damnosum s.l. with emphasis on the south-western areas. Tropenmed. Parasitol., 30, 345–62Google Scholar
Garrett-Jones, C. and Shidrawi, G. R. (1969) Malaria vectorial capacity of a population of Anopheles gambiae. Bulletin of the World Health Organization, 40, 531–45Google ScholarPubMed
Gaston, K. A. and Randolph, S. E. (1993) Reproductive under-performance of tsetse-flies in the laboratory, related to feeding frequency. Phys. Ent., 18, 130–6CrossRefGoogle Scholar
Gatehouse, A. G. (1970) The probing response of Stomoxys calcitrans to certain physical and olfactory stimuli. J. Insect Physiol., 16, 61–74CrossRefGoogle ScholarPubMed
Gatehouse, A. G. (1972) Some responses of tsetse flies to visual and olfactory stimuli. Nature New Biol., 236, 63–4CrossRefGoogle ScholarPubMed
Gaunt, M. W. and Miles, M. A. (2002) An insect molecular clock dates the origin of the insects and accords with palaeontological and biogeographic landmarks. Molecular Biology and Evolution, 19, 748–61CrossRefGoogle ScholarPubMed
Gaunt, M. W., Yeo, M., Frame, I. A. (2003) Mechanism of genetic exchange in American trypanosomes. Nature, 421, 936–9CrossRefGoogle ScholarPubMed
Gautret, P. (2001) Plasmodium falciparum gametocyte periodicity. Acta Trop., 78, 1–2CrossRefGoogle ScholarPubMed
Gautret, P. and Motard, A. (1999) Periodic infectivity of Plasmodium gametocytes to the vector: a review. Parasite-Journal De La Societe Francaise De Parasitologie, 6, 103–11Google ScholarPubMed
Geden, C. J. and Hogsette, J. A. (1994) Research and extension needs for integrated pest management for arthropods of veterinary importance. Proceedings of a Workshop in Lincoln, Nebraska, 12–14 April, Lincoln, NebraskaGoogle Scholar
Gee, J. C. (1975) Diuresis in the tsetse fly Glossina austeni. J. Exp. Biol., 63, 381–90Google ScholarPubMed
Geier, M., Bosch, O. J. and Boeckh, J. (1999) Ammonia as an attractive component of host odour for the yellow fever mosquito, Aedes aegypti. Chemical Senses, 24, 647–53CrossRefGoogle ScholarPubMed
Gentile, G., Della Torre, A., Maegga, B., Powell, J. R. and Caccone, A. (2002) Genetic differentiation in the African malaria vector, Anopheles gambiae s.s., and the problem of taxonomic status. Genetics, 161, 1561–78Google ScholarPubMed
Ghosh, K. N. and Mukhopadhyay, J. (1998) The effect of anti-sandfly saliva antibodies on Phlebotomus argentipes and Leishmania donovani. Int. J. Parasit., 28, 275–81CrossRefGoogle ScholarPubMed
Gibson, G. (1992) Do tsetse-flies see zebras: a field-study of the visual response of tsetse to striped targets. Phys. Ent., 17, 141–7CrossRefGoogle Scholar
Gibson, G. and Brady, J. (1985) Anemotactic flight paths of tsetse flies in relation to host odour: preliminary video study in nature. Phys. Ent., 10, 395–406CrossRefGoogle Scholar
Gibson, G. and Brady, J. (1988) Flight behaviour of tsetse flies in host odour plumes: the initial response to leaving or entering odour. Phys. Ent., 13, 29–42CrossRefGoogle Scholar
Gibson, G. and Torr, S. J. (1999) Visual and olfactory responses of haematophagous Diptera to host stimuli. Med. Vet. Entomol., 13, 2–23CrossRefGoogle ScholarPubMed
Gibson, G. and Young, S. (1991) The optics of tsetse-fly eyes in relation to their behavior and ecology. Phys. Ent., 16, 273–82CrossRefGoogle Scholar
Gikonyo, N. K., Hassanali, A., Njagi, P. G. N. and Saini, R. K. (2000) Behaviour of Glossina morsitans morsitans Westwood (Diptera: Glossinidae) on waterbuck Kobus defassa Ruppel and feeding membranes smeared with waterbuck sebum indicates the presence of allomones. Acta Trop., 77, 295–303CrossRefGoogle ScholarPubMed
Gillespie, R. D., Mbow, M. L. and Titus, R. G. (2000) The immunomodulatory factors of bloodfeeding arthropod saliva. Parasite Immunology, 22, 319–31CrossRefGoogle ScholarPubMed
Gillett, J. D. (1967) Natural selection and feeding speed in a blood sucking insect. Proc. R. Soc. London Ser. B., 167, 316–29CrossRefGoogle Scholar
Gillett, J. D. and Connor, J. (1976) Host temperature and the transmission of arboviruses by mosquitoes. Mosq. News, 36, 472–7Google Scholar
Gillies, M. T. (1980) The role of carbon dioxide in host-finding by mosquitoes (Diptera: Culicidae): a review. Bull. Ent. Res., 70, 525–32CrossRefGoogle Scholar
Gillies, M. T. and Wilkes, T. J. (1969) A comparison of the range of attraction of animal baits and carbon dioxide for some West African mosquitoes. Bull. Ent. Res., 59, 441–56CrossRefGoogle ScholarPubMed
Gillies, M. T. and Wilkes, T. J. (1970) The range of attraction of single baits for some West African mosquitoes. Bull. Ent. Res., 60, 225–35CrossRefGoogle ScholarPubMed
Gillies, M. T. and Wilkes, T. J. (1972) The range of attraction of animal baits and carbon dioxide for mosquitoes. Studies in a freshwater area of West Africa. Bull. Ent. Res., 61, 389–404CrossRefGoogle Scholar
Glasgow, J. P. (1961) The feeding habits of Glossina swynnertoni. J. An. Ecol., 30, 77–85CrossRefGoogle Scholar
Goodchild, A. J. P. (1955) Some observations on growth and egg production of the blood-sucking Reduviids, Rhodnius prolixus and Triatoma infestans. Proc. R. Ent. Soc., 30, 137–44Google Scholar
Gooding, R. H. (1968) A note on the relationship between feeding and insemination in Pediculus humanus. J. Med. Ent., 5, 265–6CrossRefGoogle ScholarPubMed
Gooding, R. H. (1972) Digestive processes of haematophagous insects. I. A literature review. Quaest. Ent., 8, 5–60Google Scholar
Gooding, R. H.(1974) Digestive processes in haematophagous insects. Control of trypsin secretion in Glossina morsitans morsitans. J. Insect Physiol., 20, 957–64CrossRefGoogle Scholar
Gooding, R. H. (1975) Inhibition of diuresis in the tsetse fly (Glossina morsitans) by ouabain or acetazolamide. Experientia, 31, 938–9CrossRefGoogle ScholarPubMed
Gooding, R. H. (1977) Digestive processes of haematophagous insects. XIV Haemolytic activity in the midgut of Glossina morsitans morsitans Westwood (Diptera: Glossinidae). Can. J. Zool., 55, 1899–1905CrossRefGoogle Scholar
Gordon, R. M., Crewe, W. and Willett, K. C. (1956) Studies on the deposition, migration and development to the blood forms of tryapnosomes belonging to the Trypanosoma brucei group. I. An account of the process of feeding adopted by the tsetse fly when obtaining a blood meal from the mammalian host, with special reference to the ejection of saliva and the relationship of the feeding process to the deposition of the metacyclic trypanosomes. Ann. Trop. Med., 50, 426–37CrossRefGoogle Scholar
Gore, T. C. and Pittman-Noblet, G. (1978) The effect of photoperiod on the deep body temperature of domestic turkeys and its relationship to the diurnal periodicity of Leucocytozoon smithi gametocytes in the peripheral blood of turkeys. Poultry Science, 57, 603–7CrossRefGoogle ScholarPubMed
Gorman, M. J., Cornel, A. J., Collins, F. H. and Paskewitz, S. M. (1996) A shared genetic mechanism for melanotic encapsulation of CM- Sephadex beads and a malaria parasite, Plasmodium cynomolgi B, in the mosquito, Anopheles gambiae. Experimental Parasitology, 84, 380–6CrossRefGoogle Scholar
Gottar, M., Gobert, V., Michel, T.et al. (2002) The Drosophila immune response against Gram-negative bacteria is mediated by a peptidoglycan recognition protein. Nature, 416, 640–4CrossRefGoogle ScholarPubMed
Gotz, P. (1986) Encapsulation in arthropods. In Brehelin, M. (ed.), Immunity in Invertebrates. Springer Verlag, Berlin Heidelberg, New York, Tokyo, 153–70CrossRefGoogle Scholar
Graf, R., Raikhel, A. S., Brown, M. R., Lea, A. O. and Briegel, H. (1986) Mosquito trypsin: immunocytochemical localisation in the midgut of blood-fed Aedes aegypti (L.). Cell, 245, 19–27Google Scholar
Graham, H. (1902) Dengue: a study of its mode of propagation and pathology. Medical Record, 61, 204–7
Grant, A. J., Wigton, B. E., Aghajanian, J. G. and O' Connell, R. J. (1995) Electrophysiological responses of receptor neurons in mosquito maxillary palp sensilla to carbon-dioxide. J. Comp. Physiol., 177, 389–96CrossRefGoogle ScholarPubMed
Grassi, B., Bignami, A. E. and Bastianelli, G. (1899) Ciclo evolutivo delle semilune nell' Anopheles claviger ed altri studi sulla malaria dall' ottobre 1898 all maggio 1899. Atti. Soc. Studi Malaria, 1, 143–27
Green, C. H. (1986) effects of colors and synthetic odors on the attraction of Glossina pallidipes and Glossina morsitans morsitans to traps and screens. Phys. Ent., 11, 411–21CrossRefGoogle Scholar
Green, C. H. (1989) The use of two-coloured screens for catching Glossina palpalis palpalis (Robineau-Desvoidy) (Diptera: Glossinidae). Bull. Ent. Res., 79, 81–93CrossRefGoogle Scholar
Green, C. H. and Cosens, D. (1983) Spectral responses of the tsetse fly, Glossina morsitans morsitans. J. Insect Physiol., 29, 795–800CrossRefGoogle Scholar
Griffiths, N. and Brady, J. (1995) Wind structure in relation to odour plumes in tsetse fly habitats. Phys. Ent., 20, 286–92CrossRefGoogle Scholar
Griffiths, N., Paynter, Q. and Brady, J. (1995) Rates of progress up odour trails by tsetse flies: a mark-release video study of the timing of odour source location by Glossina pallidipes. Phys. Ent., 20, 100–8CrossRefGoogle Scholar
Grimstad, P. R., Paulson, S. L. and Craig, G. B. Jr (1985) Vector competence of Aedes hendersoni (Diptera: Culicidae) for La Crosse virus and evidence of a salivary gland escape barrier. J. Med. Ent., 22, 447–53CrossRefGoogle ScholarPubMed
Grimstad, P. R., Ross, Q. E. and Craig, G. B. Jr (1980) Aedes triseriatus (Diptera: Culicidae) and La Crosse virus. II. Modification of mosquito feeding behaviour by virus infection. J. Med. Ent., 17, 1–7CrossRefGoogle ScholarPubMed
Grossman, G. L. and Pappas, L. G. (1991) Human skin temperature and mosquito (Diptera, Culicidae) blood feeding rate. J. Med. Ent., 28, 456–60CrossRefGoogle ScholarPubMed
Grubhoffer, L., Hypsa, V. and Volf, P. (1997) Lectins (hemagglutinins) in the gut of the important disease vectors. Parasite-Journal De La Societe Francaise De Parasitologie, 4, 203–16Google ScholarPubMed
Guarneri, A. A., Diotaiuti, L., Gontijo, N. F., Gontijo, A. F. and Pereira, M. H. (2000) Comparison of feeding behaviour of Triatoma infestans, Triatoma brasiliensis and Triatoma pseudomaculata in different hosts by electronic monitoring of the cibarial pump. J. Insect Physiol., 46, 1121–7CrossRefGoogle ScholarPubMed
Guerenstein, P. G. and Nunez, J. A. (1994) Feeding response of the hematophagous bugs Rhodnius prolixus and Triatoma infestans to saline solutions: a comparative-study. J. Insect Physiol., 40, 747–52CrossRefGoogle Scholar
Gwadz, R. W. (1969) Regulation of blood meal size in the mosquito. J. Insect Physiol., 15, 2039–44CrossRefGoogle ScholarPubMed
Haarlov, N. (1964) Life cycle and distribution pattern of Lipoptena cervi (Dipt., Hippobosc.) on Danish deer. Oikos, 15, 93–129CrossRefGoogle Scholar
Hacker, C. S. and Kilama, W. L. (1974) The relationship between Plasmodium gallinaceum density and the fecundity of Aedes aegypti. J. Invert. Pathol., 23, 101–5CrossRefGoogle ScholarPubMed
Hackett, L. W. and Missiroli, A. (1931) The natural disappearence of malaria in certain regions of Europe. Am. J. Hyg., 13, 57–78Google Scholar
Hafner, M. S., Sudman, P. D., Villablanca, F. X.et al. (1994) Disparate rates of molecular evolution in cospeciating hosts and parasites. Science, 265, 1087–90CrossRefGoogle ScholarPubMed
Hailman, J. P. (1979) Environmental light and conspicuous colours. In Burtt, E. H. (ed.), The Behavioural Significance of Colour. New York: Garland STPM PressGoogle Scholar
Hall, D. R., Beevor, P. S., Cork, A., Nesbitt, B. F. and Vale, G. A. (1984) 1-Octen-3-ol: a potent olfactory stimulant and attractant for tsetse isolated from cattle odours. Insect Sci. Applic., 5, 335–9Google Scholar
Hall, L. R. and Titus, R. G. (1995) Sand fly vector saliva selectively modulates macrophage functions that inhibit killing of Leishmania major and nitric oxide production. Journal of Immunology, 155, 3501–6Google ScholarPubMed
Halstead, S. B. (1990) Dengue. In Warren, K. S. and Mahmoud, A. A. F. (eds.), Tropical and Geographical Medicine.New York: McGraw Hill, 675–85Google Scholar
Handman, E. (2000) Cell biology of Leishmania. In Advances in Parasitology, Vol. 44, 1–39Google Scholar
Hansens, E. J., Bosler, E. M. and Robinson, J. W. (1971) Use of traps for study and control of salt marsh flies. J. Econ. Entomol., 64, 1481–6CrossRefGoogle Scholar
Hao, Z., Kasumba, I., Lehane, M. J.et al. (2001) Tsetse immune responses and trypanosome transmission: implications for the development of tsetse-based strategies to reduce trypanosomiasis. Proc. Natl. Acad. Sci. USA, 98, 12648–53CrossRefGoogle ScholarPubMed
Haque, A. and Capron, A. (1982) Transplacental transfer of rodent microfilariae induces antigen-specific tolerance in rats. Nature, 299, 361–3CrossRefGoogle ScholarPubMed
Hargrove, J. W. (1980a) The effect of ambient-temperature on the flight performance of the mature male tsetse-fly, Glossina morsitans. Phys. Ent., 5, 397–400CrossRefGoogle Scholar
Hargrove, J. W. (1980b) The importance of model size and ox odour on the alighting response of Glossina morsitans Westwood and Glossina pallidipes Austen (Diptera: Glossinidae). Bull. Ent. Res., 70, 229–34CrossRefGoogle Scholar
Hargrove, J. W., Holloway, M. T. P., Vale, G. A., Gough, A. J. E. and Hall, D. R. (1995) Catches of tsetse (Glossina spp) (Diptera, Glossinidae) from traps and targets baited with large doses of natural and synthetic host odor. Bull. Ent. Res., 85, 215–27CrossRefGoogle Scholar
Hargrove, J. W. and Williams, B. G. (1995) A cost-benefit-analysis of feeding in female tsetse. Med. Vet. Ent., 9, 109–19CrossRefGoogle ScholarPubMed
Harrington, L. C., Edman, J. D. and Scott, T. W. (2001) Why do female Aedes aegypti (Diptera: Culicidae) feed preferentially and frequently on human blood?J. Med. Ent., 38, 411–22CrossRefGoogle ScholarPubMed
Harris, J. A., Hillerton, J. E. and Morant, S. V. (1987) Effect on milk production of controlling muscid flies, and reducing fly-avoidance behaviour by the use of Fenvalerate ear tags during the dry period. Journal of Dairy Research, 54, 165–71CrossRefGoogle ScholarPubMed
Harris, P., Riordan, D. F. and Cooke, D. (1969) Mosquitoes feeding on insect larvae. Science, 164, 184–5CrossRefGoogle ScholarPubMed
Harrison, G. (1978) Mosquitoes, Malaria and Man: A History of the Hostilities since 1880. London: Murray
Hart, B. L. and Hart, L. A. (1994 ) Fly switching by Asian elephants: tool use to control parasites. Animal Behaviour, 48, 35–45CrossRefGoogle Scholar
Hawking, F. (1962) Microfilaria infestation as an instance of periodic phenomena seen in host-parasite relationships. Ann. N. Y. Acad. Sci., 98, 940–53CrossRefGoogle ScholarPubMed
Hawking, F. (1976) Circadian rhythms in Trypanosoma congolense. Trans. R. Soc. Trop. Med. Hyg., 70, 170CrossRefGoogle ScholarPubMed
Hawking, F., Gammage, K. and Worms, M. J. (1972) The asexual and sexual circadian rhythms of Plasmodium vinckei chabaudi, P. berghei and P. gallinaceum. Parasitology, 65, 189–201CrossRefGoogle ScholarPubMed
Hawking, F., Wilson, M. E. and Gammage, K. (1971) Guidance for cyclic development and short-lived maturity in in the gametocytes of Plasmodium falciparum. Parasitology, 65, 549–59Google Scholar
Hawking, F., Worms, M. J. and Gammage, K. (1968) 24- and 48-hour cycles of malaria parasites in the blood: their purpose, production and control. Trans. R. Soc. Trop. Med. Hyg., 62, 731–60CrossRefGoogle ScholarPubMed
Hawking, F., Worms, M. J., Gammage, K. and Goddard, P. A. (1966) The biological purpose of the blood-cycle of the malaria parasite Plasmodium cynomolgi. Lancet, 2, 422–4CrossRefGoogle Scholar
Hecker, H. and Rudin, W. (1981) Morphometric parameters of the midgut cells of Aedes aegypti L. (Insecta, Diptera) under various conditions. Cell, 219, 619–27Google ScholarPubMed
Helle, T. and Aspi, J. (1983) Does herd formation reduce insect harassment among reindeer? A field experiment with animal traps. Acta Zool. Fernica, 175, 129–31Google Scholar
Hendry, G. and Godwin, G. (1988) Biting midges in Scottish forestry: a costly irritant or a trivial nuisance?Scottish Forestry, 42, 113–19Google Scholar
Henry, V. G. and Conley, R. H. (1970) Some parasites of European wild hogs in the southern Appalachians. J. Wildl. Mgmt., 34, 913–17CrossRefGoogle Scholar
Hill, C. A., Fox, A. N., Pitts, R. J.et al. (2002) G protein coupled receptors in Anopheles gambiae. Science, 298, 176–8CrossRefGoogle ScholarPubMed
Hill, P., Saunders, D. S. and Campbell, J. A. (1973) The production of ‘symbiont-free’ Glossina morsitans and an associated loss of female fertility. Trans. R. Soc. Trop. Med. Hyg., 67, 727–8CrossRefGoogle ScholarPubMed
Hillyer, J. F. and Christensen, B. M. (2002) Characterization of hemocytes from the yellow fever mosquito, Aedes aegypti. Histochemistry and Cell Biology, 117, 431–40CrossRefGoogle ScholarPubMed
Hinnebusch, B. J., Fischer, E. R. and Schwan, T. G. (1998) Evaluation of the role of the Yersinia pestis plasminogen activator and other plasmid-encoded factors in temperature-dependent blockage of the flea. Journal of Infectious Diseases, 178, 1406–15CrossRefGoogle ScholarPubMed
Hinton, H. E. (1958) The phylogeny of the panorpoid orders. Ann. Rev. Ent., 3, 181–206CrossRefGoogle Scholar
Hoc, T. Q. and Schaub, G. A. (1996) Improvement of techniques for age grading hematophagous insects: ovarian oil-injection and ovariolar separation techniques. J. Med. Ent., 33, 286–9CrossRefGoogle ScholarPubMed
Hocking, B. (1953) The intrinsic range and speed of flight of insects. Trans. R. Soc. Lond., 104, 223–345Google Scholar
Hocking, B. (1957) Louse control through textile fibre size. Bull. Ent. Res., 48, 507–14CrossRefGoogle Scholar
Hocking, B. (1971) Blood-sucking behaviour of terrestrial arthropods. Ann. Rev. Ent., 16, 1–26CrossRefGoogle ScholarPubMed
Hockmeyer, W. T., Schieffer, B. A., Redington, B. C. and Eldridge, B. V. (1975) Brugia pahangi effects upon the flight capability of Aedes aegypti. Experimental Parasitology, 38, 1–5CrossRefGoogle ScholarPubMed
Hoffmann, J. A. (2003) The immune response of Drosophila. Nature, 426, 33–8CrossRefGoogle ScholarPubMed
Hoffmann, J. A. and Reichhart, J. M. (2002) Drosophila innate immunity: an evolutionary perspective. Nature Immunology, 3, 121–6CrossRefGoogle Scholar
Hogg, J. C. and Hurd, H. (1995) Plasmodium yoelli nigeriensis – the effect of high and low intensity of infection upon the egg production and bloodmeal size of Anopheles stephensi during three gonotrophic cycles. Parasitol., 111, 555–62CrossRefGoogle ScholarPubMed
Holloway, M. T. P. and Phelps, R. J. (1991) The responses of Stomoxys spp. (Diptera, Muscidae) to traps and artificial host odors in the field. Bull. Ent. Res., 81, 51–5CrossRefGoogle Scholar
Hooke, R. (1664) Micrographia.
Hopkins, F. H. E. (1949) The host associations of the lice of mammals. Proc. Zool. Soc. Lond., 119, 387–604CrossRefGoogle Scholar
Hopwood, J. A., Ahmed, A. M., Polwart, A., Williams, G. T. and Hurd, H. (2001) Malaria-induced apoptosis in mosquito ovaries: a mechanism to control vector egg production. J. Exp. Biol., 204, 2773–2780Google ScholarPubMed
Hosoi, T. (1958) Adenosine 5' phosphates as the stimulating agent in blood for inducing gorging of the mosquito. Nature, 181, 1664–5CrossRefGoogle ScholarPubMed
Hosoi, T. (1959) Identification of blood components which induce gorging in the mosquito. J. Insect Phys., 3, 191–218CrossRefGoogle Scholar
Houseman, J. G., Downe, A. E. R. and Morrison, P. E. (1985a) Similarities in digestive proteinase production in Rhodnius prolixus (Hemiptera: Reduviidae) and Stomoxys calcitrans (Diptera: Muscidae). Insect Biochem., 15, 471–4CrossRefGoogle Scholar
Houseman, J. G., Morrison, P. E. and Downe, A. E. R. (1985b) Cathepsin B and aminopeptidase in the posterior midgut of Euschistus euschistoides (Hemiptera: Phymatidae). Can. J. Zool., 63, 1288–91CrossRefGoogle Scholar
Howe, M. A. and Lehane, M. J. (1986) Post-feed buzzing in the tsetse, Glossina morsitans morsitans, is an endothermic mechanism. Phys. Ent., 11, 279–86CrossRefGoogle Scholar
Huang, C. T. (1971) Vertebrate serum inhibitors of Aedes aegypti trypsin. Insect Biochem., 1, 27–38CrossRefGoogle Scholar
Hudson, A. (1970) Notes on the piercing mouthparts of three species of mosquitoes (Diptera: Culicidae) viewed with the scanning electron miscroscope. Can. Entomal., 102, 501–9CrossRefGoogle Scholar
Hudson, B. W., Feingold, B. F. and Kartman, L. (1960) Allergy to flea bites. II. Investigations of flea bite sensitivity in humans. Experimental Parasitology, 9, 264–70CrossRefGoogle Scholar
Huff, C. (1929) Ovulation requirements of Culex pipiens Linn. Biol. Bull. (Woods Hole), 56, 347–50CrossRefGoogle Scholar
Huff, C. (1931) A proposed classification of disease transmission by arthropods. Science, 74, 456–7CrossRefGoogle Scholar
Hughes, A. L. and Piontkivska, H. (2003) Phylogeny of trypanosomatidae and bodonidae (Kinetoplastida) based on 18S rRNA: evidence for paraphyly of Trypanosoma and six other genera. Molecular Biology and Evolution, 20, 644–52CrossRefGoogle ScholarPubMed
Humphries, D. A. (1967) Function of combs in ectoparasites. Nature, 215, 319CrossRefGoogle ScholarPubMed
Hunter, D. M. and Moorhouse, D. W. (1976) The effects of Austrosimulium pestilens on the milk production of dairy cattle. Aust. Vet. J., 52, 97–9CrossRefGoogle ScholarPubMed
Huq, M. (1961) African horse sickness. Veterinary Record, 73, 123Google Scholar
Hurd, H. (1998) Parasite manipulation of insect reproduction: who benefits? Parasitology, 116, S13–21CrossRefGoogle ScholarPubMed
Hurd, H. (2003) Manipulation of medically important insect vectors by their parasites. Ann. Rev. Ent., 48, 141–61CrossRefGoogle ScholarPubMed
Hurd, H., Hogg, J. C. and Renshaw, M. (1995) Interactions between bloodfeeding, fecundity and infection in mosquitoes. Parasitology Today, 11, 411–6CrossRefGoogle Scholar
Hursey, B. S. (2001) The programme against African trypanosomiasis: aims, objectives and achievements. Trends in Parasitology, 17, 2–3CrossRefGoogle ScholarPubMed
Ibrahim, E. A. R., Ingram, G. A. and Molyneux, D. H. (1984) Haemagglutinins and parasite agglutinins in haemolymph and gut of Glossina. Tropenmed. Parasit., 35, 151–6Google ScholarPubMed
ICZN (1999) International Code of Zoological Nomenclature. London: ICZN
Irving, P., Troxler, L., Heuer, T. S.et al. (2001) A genome-wide analysis of immune responses in Drosophila. Proceedings of the National Academy of Sciences of the United States of America, 98, 15119–24CrossRefGoogle ScholarPubMed
Isawa, H., Yuda, M., Orito, Y. and Chinzei, Y. (2002) A mosquito salivary protein inhibits activation of the plasma contact system by binding to factor Ⅻ and high molecular weight kininogen. J. Biol. Chem., 277, 27651–8CrossRefGoogle ScholarPubMed
Isawa, H., Yuda, M., Yoneda, K. and Chinzei, Y. (2000) The insect salivary protein, prolixin-S, inhibits factor IXa generation and Xase complex formation in the blood coagulation pathway. J. Biol. Chem., 275, 6636–41CrossRefGoogle ScholarPubMed
Iwanaga, S. (2002) The molecular basis of innate immunity in the horseshoe crab. Current Opinion in Immunology, 14, 87–95CrossRefGoogle ScholarPubMed
James, M. T. and Harwood, R. F. (1969) Herm's Medical Entomology. London: Macmillan
Janeway, C. A., Travers, P., Walport, M. and Schlomchik, M. (2001) Immunology. Edinburgh: Churchill Livingstone
Janse, C. J., Rouwenhorst, R. J., Klooster, P. F. J., Kaay, H. J. and Overdulve, J. P. (1985) Development of Plasmodium berghei ookinetes in the midgut of Anopheles atroparvus mosquitoes and in vitro. Parasitol., 91, 219–25CrossRefGoogle ScholarPubMed
Jefferies, D. (1984) Transmission of disease by haematophagous arthropods. Unpublished Ph.D. thesis, University of Salford
Jenkins, D. W. (1964) Advances in medical entomology using radio-isotopes. Experimental Parasitology, 3, 474–90CrossRefGoogle Scholar
Jenni, L., Molyneux, D. H., Livesey, J. L. and Galun, R. (1980) Feeding behaviour of tsetse flies infected with salivarian trypanosomes. Nature, 283, 383–5CrossRefGoogle ScholarPubMed
Jobling, B. (1976) On the fascicle of blood-sucking Diptera. In addition a description of the maxillary glands in Phlebotomus papatasi, together with the musculature of the labium and pulsatory organ of both the latter species and also of some other Diptera. J. Nat. Hist., 10(4), 457–61
Johnson, K. P., Adams, R. J. and Clayton, D. H. (2002a) The phylogeny of the louse genus Brueelia does not reflect host phylogeny. Biological Journal of the Linnaean Society, 77, 233–47CrossRefGoogle Scholar
Johnson, K. P., Weckstein, J. D., Witt, C. C., Faucett, R. C. and Moyle, R. G. (2002b) The perils of using host relationships in parasite taxonomy: phylogeny of the Degeeriella complex. Mol. Phylogenet. Evol., 23, 150–7CrossRefGoogle Scholar
Jones, C. J. (1996) Immune responses to fleas, bugs and sucking lice. In Wikel, S. K. (ed.), The Immunology of Host–Ectoparasitic Arthropod Interactions. Wallingford: CAB International, 150–74Google Scholar
Jones, J. C. and Pillitt, D. R. (1973) Blood-feeding behavior of adult Aedes aegypti mosquitoes. Biol. Bull., 145, 127–39CrossRefGoogle ScholarPubMed
Jordan, A. M. (1974) Recent development in the ecology and methods of control of tsetse flies (Glossina spp.), a review. Bull. Ent. Res., 63(4), 361–99CrossRef
Jordan, A. M. (1986) Trypanosomiasis Control and African Rural Development. London: Longman
Jordan, A. M. and Curtis, C. F. (1968) The performance of Glossina austeni when fed on lop-eared rabbits and goats. Trans. R. Soc. Trop. Med. Hyg., 62, 123–4CrossRefGoogle Scholar
Jordan, A. M. and Curtis, C. F. (1972) Productivity of Glossina morsitans morsitans Westwood maintained in the laboratory, with particular reference to the sterile-insect release method. Bull. W.H.O., 46, 33–8Google Scholar
Jordan, K. (1962) Notes on the Tunga caecigena (Siphonaptera: Tungidae). Bull. Br. Mus. Nat. Hist. (Ent.), 12(4), 353–64CrossRef
Julius, D. and Basbaum, A. I. (2001) Molecular mechanisms of nociception. Nature, 413, 203–10CrossRefGoogle ScholarPubMed
Kaaya, G. P. and Ratcliffe, N. A. (1982) Comparative study of haemocytes and associated cells of some medically important dipterans. J. Morph., 173, 351–65CrossRefGoogle Scholar
Kamhawi, S. (2000) The biological and immunomodulatory properties of sand fly saliva and its role in the establishment of Leishmania infections. Microbes Infect, 2, 1765–73CrossRefGoogle ScholarPubMed
Kamhawi, S., Belkaid, Y., Modi, G., Rowton, E. and Sacks, D. (2000) Protection against cutaneous Leishmaniasis resulting from bites of uninfected sand flies. Science, 290, 1351–4CrossRefGoogle ScholarPubMed
Kangwangye, T. N. (1977) Reactions of large mammals to biting flies in Rwenzori National Park, Uganda. In C. P. F. Lima (ed.), Proceedings of the First East African Conference on Entomological Pest Control.
Kartman, L. (1953) Factors influencing infection of the mosquito with Dirofilaria immitis (Leidy, 1856). Experimental Parasitology, 2, 27–78CrossRefGoogle Scholar
Kathirithamby, J., Ross, L. D. and Johnston, J. S. (2003) Masquerading as self? Endoparasitic Strepsiptera (Insecta) enclose themselves in host-derived epidermal bag. Proc. Natl. Acad. Sci. USA, 100, 7655–9CrossRefGoogle ScholarPubMed
Katz, O., Waitumbi, J. N., Zer, R. and Warburg, A. (2000) Adenosine, AMP, and protein phosphatase activity in sandfly saliva. Am. J. Trop. Med. Hyg., 62, 145–50CrossRefGoogle ScholarPubMed
Kavaliers, M., Choleris, E. and Colwell, D. D. (2001) Learning from others to cope with biting flies: social learning of fear-induced conditioned analgesia and active avoidance. Behavioral Neuroscience, 115, 661–74CrossRefGoogle ScholarPubMed
Keiper, R. R. and Berger, J. (1982) Refuge-seeking and pest avoidance by feral horses in desert and island environments. Applied Animal Ethology, 9, 111–20CrossRefGoogle Scholar
Kellogg, F. E. (1970) Water vapour and carbon dioxide receptors in Aedes aegypti. J. Insect Physiol., 16, 99–108CrossRefGoogle ScholarPubMed
Kellogg, F. E. and Wright, R. H. (1962) The guidance of flying insects. V. Mosquito attraction. Can. Entomol., 94, 1009–16CrossRefGoogle Scholar
Kelly, D. W. (2001) Why are some people bitten more than others?Trends in Parasitology, 17, 578–81CrossRefGoogle ScholarPubMed
Kelly, D. W., Mustafa, Z. and Dye, C. (1996) Density-dependent feeding success in a field population of the sandfly, Lutzomyia longipalpis. Journal of Animal Ecology, 65, 517–27CrossRefGoogle Scholar
Kelly, D. W. and Thompson, C. E. (2000) Epidemiology and optimal foraging: modelling the ideal free distribution of insect vectors. Parasitology, 120, 319–27CrossRefGoogle ScholarPubMed
Kennedy, J. S. (1940) The visual responses of flying mosquitoes. Proc. Zool. Soc. Lond., 109, 221–42Google Scholar
Kennedy, J. S. (1983) Zigzagging and casting as a programmed response to wind-borne odour: a review. Phys. Ent., 8, 109–20CrossRefGoogle Scholar
Kettle, D. S. (1984) Medical and Veterinary Entomology. London: Croom Helm
Khan, A. A. and Maibach, H. I. (1970) A study of the probing response of Aedes aegypti. I. Effect of nutrition on probing. J. Econ. Ent., 63, 974–6CrossRefGoogle ScholarPubMed
Khan, A. A. and Maibach, H. I. (1971) A study of the probing response of Aedes aegypti. 2. Effect of desiccation and blood feeding on probing to skin and an artificial target. J. Econ. Entomol., 64, 439–42CrossRefGoogle ScholarPubMed
Kilama, W. L. and Craig, G. B. (1969) Monofactorial inheritance of susceptibility to Plasmodium gallinaceum in Aedes aegypti. Ann. Trop. Med. Parasit., 63, 419–32CrossRefGoogle ScholarPubMed
Killeen, G. F., McKenzie, F. E., Foy, B. D., Bogh, C. and Beier, J. C. (2001) The availability of potential hosts as a determinant of feeding behaviours and malaria transmission by African mosquito populations. Trans. R. Soc. Trop. Med. Hyg., 95, 469–76CrossRefGoogle ScholarPubMed
Kim, K. C. (1985) Evolution and host association of Anoplura. In Kim, K. C. (ed.), Coevolution of Parasitic Arthropods and Mammals. New York: WileyGoogle Scholar
Kim, K. D. and Adler, P. H. (1985) Evolution and host association of Anoplura. In Kim, K. C. (ed.), Coevolution of Parasitic Arthropods and Mammals.New York: WileyGoogle Scholar
Kingsolver, J. G. (1987) Mosquito host choice and the epidemiology of malaria. Am. Nat., 130, 811–27CrossRefGoogle Scholar
Kirch, H. J., Spates, G., Droleskey, R., Kloft, W. J. and Deloach, J. R. (1991a) Mechanism of hemolysis of erythrocytes by hemolytic factors from Stomoxys calcitrans (L) (Diptera, Muscidae). J. Insect Phys., 37, 851–61CrossRefGoogle Scholar
Kirch, H. J., Spates, G., Kloft, W. J. and Deloach, J. R. (1991b) The relationship of membrane-lipids to species-specific hemolysis by hemolytic factors from Stomoxys calcitrans (L) (Diptera, Muscidae). Insect Biochem., 21, 113CrossRefGoogle Scholar
Klein, T. A., Harrison, B. A., Andre, R. G., Whitmire, R. E. and Inlao, I. (1982) Detrimental effects of Plasmodium cynomolgi infections on the longevity of Anopheles dirus. Mosq. News, 42, 265–71Google Scholar
Kline, D. L. and Lemire, G. F. (1995) Field evaluation of heat as an added attractant to traps baited with carbon dioxide and octenol for Aedes taeniorhynchus. J. Am. Mosq. Control Assoc., 11, 454–6Google ScholarPubMed
Klowden, M. J. (1993) Mating and nutritional state affect the reproduction of Aedes albopictus mosquitoes. J. Am. Mosq. Control Assoc., 9, 169–73Google ScholarPubMed
Klowden, M. J., Davis, E. E. and Bowen, M. F. (1987) Role of the fat body in the regulation of host-seeking behaviour in the mosquito, Aedes aegypti. J. Insect Physiol., 33, 643–6CrossRefGoogle Scholar
Klowden, M. J., Kline, D. L., Takken, W., Wood, J. R. and Carlson, D. A. (1990) Field studies on the potential of butanone, carbon-dioxide, honey extract, 1-octen-3-ol, L-lactic acid and phenols as attractants for mosquitos. Med. Vet. Entomol., 4, 383–91Google Scholar
Klowden, M. J. and Lea, A. O. (1979) Abdominal distention terminates subsequent host-seeking behavior of Aedes aegypti following a blood meal. J. Insect Physiol., 25, 583–5CrossRefGoogle ScholarPubMed
Knols, B. G. J., Jong, R. and Takken, W. (1995) Differential attractiveness of isolated humans to mosquitoes in Tanzania. Trans. R. Soc. Trop. Med. Hyg., 89, 604–6CrossRefGoogle ScholarPubMed
Knols, B. G. J., Mboera, L. E. G. and Takken, W. (1998) Electric nets for studying odour-mediated host-seeking behaviour of mosquitoes. Med. Vet. Entomol., 12, 116–20CrossRefGoogle ScholarPubMed
Knols, B. G. J., Loon, J. J. A., Cork, A.et al. (1997) Behavioural and electrophysiological responses of the female malaria mosquito Anopheles gambiae (Diptera: Culicidae) to Limburger cheese volatiles. Bull. Ent. Res., 87, 151–9CrossRefGoogle Scholar
Koella, J. C., Agnew, P. and Michalakis, Y. (1998a) Coevolutionary interactions between host life histories and parasite life cycles. Parasitology, 116, S47–S55CrossRefGoogle Scholar
Koella, J. C. and Boete, C. (2002) A genetic correlation between age at pupation and melanization immune response of the yellow fever mosquito Aedes aegypti. Evolution Int. J. Org. Evolution, 56, 1074–9CrossRefGoogle ScholarPubMed
Koella, J. C., Sorensen, F. L. and Anderson, R. A. (1998b) The malaria parasite, Plasmodium falciparum, increases the frequency of multiple feeding of its mosquito vector, Anopheles gambiae. Proc. R. Soc. Lond. B Biol. Sci., 265, 763–8CrossRefGoogle Scholar
Komano, H., Mizuno, D. and Natori, S. (1980) Purification of a lectin induced in the haemolymph of Sarcophaga peregrina larvae on injury. J. Biol. Chem., 255, 2919–24Google Scholar
Kramer, L. D., Hardy, J. L., Presser, S. B. and Houk, E. G. (1981) Dissemination barriers for western equine encephalomyelitis virus in Culex tarsalis infected after ingestion of low viral doses. Am. J. Trop. Med. Hyg., 30, 190–7CrossRefGoogle ScholarPubMed
Krasnov, B. R., Khokhlova, I. S. and Shenbrot, G. I. (2003a) Density-dependent host selection in ectoparasites: an application of isodar theory to fleas parasitizing rodents. Oecologia, 134, 365–72CrossRefGoogle Scholar
Krasnov, B. R., Sarfati, M., Arakelyan, M. S.et al. (2003b) Host specificity and foraging efficiency in blood-sucking parasite: feeding patterns of the flea Parapulex chephrenis on two species of desert rodents. Parasitology Research, 90, 393–9CrossRefGoogle Scholar
Krynski, S., Kuchta, A. and Becla, E. (1952) Research on the nature of the noxious action of guinea-pig blood on the body louse (in Polish). Bull. Inst. Mar. Med. Gdansk, 4, 104–7Google Scholar
Ksiazkiewicz-Ilijewa, M. and Rosciszewska, E. (1979) Ultrastructure of the haemocytes of Tetrodontophora bielanensis Waga (Collembola). Cytobios, 26, 113–21Google Scholar
Kunz, S. E., Murrell, K. D., Lambert, G., James, L. F. and Terrill, C. E. (1991) Estimated losses of livestock to pests. In Pimentel, D. (ed.), CRCHandbook of Pest Management in Agriculture. Boca Raton, Florida: CRC Press, Vol. 1, pp. 69–98Google Scholar
Kurata, S. (2004) Recognition of infectious non-self and activation of immune responses by peptidoglycan recognition protein (PGRP)-family members in Drosophila. Dev. Comp. Immunol., 28, 89–95CrossRefGoogle ScholarPubMed
Kurtz, J. and Franz, K. (2003) Evidence for memory in invertebrate immunity. Nature, 425, 37–8CrossRefGoogle ScholarPubMed
Breque, G. C., Meifert, D. W. and Rye, J. (1972) Experimental control of stable flies, Stomoxys calcitrans (Diptera: Muscidae), by the release of chemosterilized adults. Can. Entomol., 104, 885–7Google Scholar
Laarman, J. J. (1958) The host-seeking behaviour of anopheline mosquitoes. Trop. Geogr. Med., 10, 293–305Google ScholarPubMed
Lackie, A. M. (1986) Evasion of insect immunity by helminth larvae. In Lackie, A. M. (ed.), Immune Mechanisms in Invertebrate Vectors, Oxford: Oxford University Press, Vol. 56, 161–78Google Scholar
Lafond, M. M., Christensen, B. M. and Lasee, B. A. (1985) Defense reactions of mosquitoes to filarial worms: potential mechanisms for avoidance of the response by Brugia pahangi microfilaria. J. Invert. Pathol., 46, 26–30CrossRefGoogle Scholar
Lainson, R. and Shaw, J. J. (1987) Evolution, classification and geographical distribution. In Peters, W. and Killick-Kendrick, R. (eds.), The Leishmaniases in Biology and Medicine. New York: Academic PressGoogle Scholar
Lall, S. B. (1969) Phagostimulants of haematophagous tabanids. Entomol. Exp. Appl., 12, 325–36CrossRefGoogle Scholar
Land, M. F., Gibson, G. and Horwood, J. (1997) Mosquito eye design: conical rhabdoms are matched to wide aperture lenses. Proc. R. Soc. Lond. B. Biol. Sci., 264, 1183–7CrossRefGoogle Scholar
Land, M. F., Gibson, G., Horwood, J. and Zeil, J. (1999) Fundamental differences in the optical structure of the eyes of nocturnal and diurnal mosquitoes. J. Comp. Physiol., 185, 91–103CrossRefGoogle Scholar
Lane, N. J. and Harrison, J. B. (1979) An unusual cell surface modification: a double plasma membrane. J. Cell Sci., 39, 355–72Google ScholarPubMed
Langley, P. A. (1970) Post-teneral development of thoracic flight musculature in the tsetse flies Glossina austeni and G. morsitans. Entomologia Exp. Appl., 13, 133–40CrossRefGoogle Scholar
Langley, P. A. and Maly, H. (1969) Membrane feeding technique for tsetse flies (Glossina spp.). Nature, 221, 855–6CrossRefGoogle Scholar
Lanzaro, G. C., Toure, Y. T., Carnahan, J.et al. (1998) Complexities in the genetic structure of Anopheles gambiae populations in west Africa as revealed by microsatellite DNA analysis. Proceedings of the National Academy of Sciences of the United States of America, 95, 14260–5CrossRefGoogle ScholarPubMed
Larrivee, D. H., Benjamini, E., Feingold, B. F. and Shimuzu, M. (1964) Histologic studies of guinea pig skin: different stages of allergic reactivity to flea bites. Exp. Parasitol., 15, 491–502CrossRefGoogle ScholarPubMed
Laurence, B. R. (1966) Intake and migration of the microfilariae of Onchocerca volvulus (Leukart) in Simulium damnosum Theobald. J. Helm., 40, 337–42CrossRefGoogle Scholar
Laurence, B. R. and Pester, F. R. N. (1961) The ability of Anopheles gambiae Giles to transmit Brugia patei (Buckley, Nelson and Heisch). J. Trop. Med. Hyg., 64, 169–71Google Scholar
Laurence, B. R. and Pester, F. R. N. (1967) Adaptation of a filarial worm, Brugia patei, to a new mosquito host, Aedes togoi. J. Helminth., 41, 365–92CrossRefGoogle Scholar
Lavine, M. D. and Strand, M. R. (2002) Insect hemocytes and their role in immunity. Insect Biochem. Mol. Biol., 32, 1295–309CrossRefGoogle ScholarPubMed
Lavoipierre, M. M. J., Dickerson, G. and Gordon, R. M. (1959) Studies on the methods of feeding of blood-sucking arthropods. I. The manner in which triatomine bugs obtain their blood meal as observed in the tissues of the living rodent, with some remarks on the effects of the bite on human volunteers. Ann. Trop. Med. Parasitol., 53, 235–50CrossRefGoogle Scholar
Lavoipierre, M. M. J. and Hamachi, M. (1961) An apparatus for observations on the feeding mechanism of the flea. Nature, 192, 998–9CrossRefGoogle Scholar
Lazzari, C. R., Reiseman, C. E. and Insausti, T. C. (1998) The role of the ocelli in the phototactic behaviour of the haematophagous bug Triatoma infestans. J. Insect Physiol., 44, 1159–62CrossRefGoogle ScholarPubMed
Lebestky, T., Chang, T., Hartenstein, V. and Banerjee, U. (2000) Specification of Drosophila hematopoietic lineage by conserved transcription factors. Science, 288, 146–9CrossRefGoogle ScholarPubMed
Lee, J. H., Rowley, W. A. and Platt, K. B. (2000) Longevity and spontaneous flight activity of Culex tarsalis (Diptera: Culicidae) infected with western equine encephalomyelitis virus. J. Med. Ent., 37, 187–93CrossRefGoogle ScholarPubMed
Lee, R. (1974) Structure and function of the fascicular stylets, and the labral and cibarial sense organs of male and female Aedes aegypti (L>). Quest. Entomol., 10, 187–215Google Scholar
Lehane, M. J. (1976a) Digestive enzyme secretion in Stomoxys calcitrans (Diptera: Muscidae). Tissue and Cell, 170, 275–87Google Scholar
Lehane, M. J. (1976b) The formation and histochemical structure of the peritrophic membrane in the stable fly, Stomoxys calcitrans. J. Insect Physiol., 22, 1551–7CrossRefGoogle Scholar
Lehane, M. J. (1977a) An hypothesis of the mechanism controlling proteolytic digestive enzyme production levels in Stomoxys calcitrans. J. Insect Physiol., 23, 713–15CrossRefGoogle Scholar
Lehane, M. J. (1977b) Transcellular absorption of lipids in the midgut of the stablefly, Stomoxys calcitrans. J. Insect Physiol., 23, 945–54CrossRefGoogle Scholar
Lehane, M. J. (1985) Determining the age of an insect. Parasitology Today, 1, 81–5CrossRefGoogle ScholarPubMed
Lehane, M. J. (1987) Quantitative evidence for merocrine secretion in an insect midgut cell. Tissue and Cell, 19, 451–561CrossRefGoogle Scholar
Lehane, M. J. (1988) Evidence for secretion by the release of cytoplasmic extrusions from midgut cells of Stomoxys calcitrans. J. Insect Physiol., 34, 949–53CrossRefGoogle Scholar
Lehane, M. J. (1989) The intracellular pathway and kinetics of digestive enzyme secretion in an insect midgut cell. Tissue and Cell, 21, 101–11CrossRefGoogle Scholar
Lehane, M. J. (1997) Peritrophic matrix structure and function. Ann. Rev. Ent., 42, 525–50CrossRefGoogle ScholarPubMed
Lehane, M. J., Aksoy, S., Gibson, W. (2003) Adult midgut EST from the tsetse fly Glossina morsitans morsitans and expression analysis of putative immune response genes. Genome Biology, 4 (10), R63CrossRefGoogle ScholarPubMed
Lehane, M. J., Aksoy, S. and Levashina, E. A. (2004) Blood-sucking insect immune responses and parasite transmission. Trends in Parasitology, in press
Lehane, M. J., Allingham, P. G. and Weglicki, P. (1996) Peritrophic matrix composition of the tsetse fly, Glossina morsitans morsitans. Cell and Tissue Research, 272, 158–62Google Scholar
Lehane, M. J. and Billingsley, P. A. (eds.) (1996) The Biology of the Insect Midgut.London: Chapman and HallCrossRefGoogle Scholar
Lehane, M. J., Crisanti, A. and Mueller, H. M. (1996) Mechanisms controlling the synthesis and secretion of digestive enzymes in insects. In Lehane, M. J. (ed.), The Insect Midgut. London: Chapman and HallCrossRefGoogle Scholar
Lehane, M. J. and Hargrove, J. (1988) Field experiments on a new method for determining age in tsetse flies (Diptera, Glossinidae). Ecol. Entomol., 13, 319–22CrossRefGoogle Scholar
Lehane, M. J. and Laurence, B. R. (1977) Flight muscle ultrastructure of susceptible and refractory mosquitoes parasitized by larval Brugia pahangi. Parasitology, 74, 87–92CrossRefGoogle ScholarPubMed
Lehane, M. J. and Mail, T. S. (1985) Determining the age of adult male and female Glossina morsitans morsitans using a new technique. Ecol. Entomol., 10, 219–24CrossRefGoogle Scholar
Lehane, M. J. and Schofield, C. J. (1981) Field experiments of dispersive flight by Triatoma infestans. Trans. R. Soc. Trop. Med. Hyg., 75, 399–400CrossRefGoogle ScholarPubMed
Lehane, M. J. and Schofield, C. J. (1982) Flight initiation in Triatoma infestans (Klug) (Hemiptera: Reduviidae). Bull. Ent. Res., 72, 497–510CrossRefGoogle Scholar
Lehane, M. J., Wu, D. and Lehane, S. M. (1997) Midgut-specific immune molecules are produced by the blood-sucking insect Stomoxys calcitrans. Proceedings of the National Academy of Sciences of the United States of America, 94, 11502–7CrossRefGoogle ScholarPubMed
Lehane, S. M., Assinder, S. J. and Lehane, M. J. (1998) Cloning, sequencing, temporal expression and tissue-specificity of two serine proteases from the midgut of the blood-feeding fly Stomoxys calcitrans. European Journal of Biochemistry, 254, 290–6CrossRefGoogle ScholarPubMed
Lemaitre, B., Reichhart, J. M. and Hoffmann, J. A. (1997) Drosophila host defense: differential induction of antimicrobial peptide genes after infection by various classes of microorganisms. Proceedings of the National Academy of Sciences of the United States of America, 94, 14614–19CrossRefGoogle ScholarPubMed
Lemos, F. J. A., Cornel, A. J. and Jacobs Lorena, M. (1996) Trypsin and aminopeptidase gene expression is affected by age and food composition in Anopheles gambiae. Insect Biochemistry and Molecular Biology, 26, 651–8CrossRefGoogle ScholarPubMed
Lester, H. M. O. and Lloyd, L. (1929) Notes on the process of digestion in tsetse flies. Bull. Ent. Res., 19, 39–60CrossRefGoogle Scholar
Leulier, F., Parquet, C., Pili-Floury, S. et al. (2003) The Drosophila immune system detects bacteria through specific peptidoglycan recognition. Nature Immunology, 4, 478–84CrossRefGoogle ScholarPubMed
Levashina, E. A., Langley, E., Green, C. et al. (1999) Constitutive activation of toll-mediated antifungal defense in serpin-deficient Drosophila. Science, 285, 1917–19CrossRefGoogle ScholarPubMed
Levashina, E. A., Moita, L. F., Blandin, S.et al. (2001) Conserved role of a complement-like protein in phagocytosis revealed by dsRNA knockout in cultured cells of the mosquito Anopheles gambiae. Cell, 104, 709–18CrossRefGoogle ScholarPubMed
Lewis, D. J. (1953) Simulium damnosum and its relation to Onchocerciasis in the Anglo-Egyptian Sudan. Bull. Ent. Res., 43, 597–644CrossRefGoogle Scholar
Lewis, L. F., Christenson, D. M. and Eddy, G. W. (1967) Rearing the long-nosed cattle louse and cattle-biting louse on host animals in Oregon. J. Econ. Entomol., 60, 755–7CrossRefGoogle Scholar
Lewis, T. and Taylor, L. R. (1965) Diurnal periodicity of flight by insects. Trans. R. Ent. Soc. Lond., 116, 393–479CrossRefGoogle Scholar
Li, X., Sina, B. and Rossignol, P. A. (1992) Probing behaviour and sporozoite delivery by Anopheles stephensi infected with Plasmodium berghei. Med. Vet. Entomol., 6, 57–61CrossRefGoogle ScholarPubMed
Ligoxygakis, P., Pelte, N., Hoffmann, J. A. and Reichart, J. M. (2002) Activation of Drosophila toll during fungal infection by a blood serine protease. Nature Reviews Immunology, 2, 545Google Scholar
Lindsay, L. B. and Galloway, T. D. (1998) Reproductive status of four species of fleas (Insecta: Siphonaptera) on Richardson's ground squirrels (Rodentia: Sciuridae) in Manitoba, Canada. J. Med. Ent., 35, 423–30CrossRefGoogle Scholar
Lindsay, S. W., Adiamah, J. H., Miller, J. E., Pleass, R. J. and Armstrong, J. R. M. (1993) Variation in attractiveness of human subjects to malaria mosquitoes (Diptera, Culicidae) in the Gambia. J. Med. Ent., 30, 368–73CrossRefGoogle Scholar
Linley, J. R. and Davies, J. B. (1971) Sandflies and tourism in Florida and the Bahamas and Caribbean area. J. Econ. Entomol., 64, 264–78CrossRefGoogle Scholar
Linsenmair, K. E. (1973) Die Windorientierurung laufender Insekten. Fortschr. Zool., 21, 59–79Google Scholar
Liu, C. T., Hou, R. F. and Chen, C. C. (1998) Formation of basement membrane-like structure terminates the cellular encapsulation of microfilariae in the haemocoel of Anopheles quadrimaculatus. Parasitology, 116 (Pt 6), 511–18CrossRefGoogle ScholarPubMed
Lochmiller, R. L. and Deerenberg, C. (2000) Trade-offs in evolutionary immunology: just what is the cost of immunity?Oikos, 88, 87–98CrossRefGoogle Scholar
Loder, P. M. J., Hargrove, J. W. and Randolph, S. E. (1998) A model for blood meal digestion and fat metabolism in male tsetse flies (Glossinidae). Phys. Ent., 23, 43–52CrossRefGoogle Scholar
Lodmell, D. L., Bell, J. F., Clifford, C. M., Moore, G. J. and Raymond, G. (1970) Effects of limb disability on lousiness in mice. V. Hierarchy disturbance on mutual grooming and reproductive capacities. Expl. Parasit., 27, 184–92CrossRefGoogle Scholar
Loke, H. and Randolph, S. E. (1995) Reciprocal regulation of fat-content and flight activity in male tsetse-flies (Glossina palpalis). Phys. Ent., 20, 243–7CrossRefGoogle Scholar
Lord, W. D., DiZinno, J. A., Wilson, M. R. et al. (1998) Isolation, amplification, and sequencing of human mitochondrial DNA obtained from human crab louse, Pthirus pubis (L.), blood meals. Journal of Forensic Sciences, 43, 1097–100CrossRefGoogle ScholarPubMed
Loudon, C. and McCulloh, K. (1999) Application of the Hagen-Poiseuille equation to fluid feeding through short tubes. Ann. Ent. Soc. Am., 92, 153–8CrossRefGoogle Scholar
Lowenberger, C. A., Ferdig, M. T., Bulet, P. et al. (1996) Aedes aegypti – induced antibacterial proteins reduce the establishment and development of Brugia malayi. Experimental Parasitology, 83, 191–201CrossRefGoogle ScholarPubMed
Lowther, J. K. and Wood, D. M. (1964) Specificity of a black fly, Simulium euryadmiculum Davies, towards its host, the common loon. Can. Entomol., 96, 911–13CrossRefGoogle Scholar
Luckhart, S. and Rosenberg, R. (1999) Gene structure and polymorphism of an invertebrate nitric oxide synthase gene. Gene, 232, 25–34CrossRefGoogle ScholarPubMed
Luckhart, S., Vodovotz, Y., Cui, L. W. and Rosenberg, R. (1998) The mosquito Anopheles stephensi limits malaria parasite development with inducible synthesis of nitric oxide. Proceedings of the National Academy of Sciences of the United States of America, 95, 5700–5CrossRefGoogle ScholarPubMed
Lyman, D. E., Monteiro, F. A., Escalante, A. A. et al. (1999) Mitochondrial DNA sequence variation among triatomine vectors of Chagas' disease. Am. J. Trop. Med. Hyg., 60, 377–86CrossRefGoogle ScholarPubMed
Lythgoe, K. A. (2000) The coevolution of parasites with host-acquired immunity and the evolution of sex. Evolution Int. J. Org. Evolution, 54, 1142–56CrossRefGoogle ScholarPubMed
Maa, T. C. and Marshall, A. G. (1981) Diptera Pupipara of the New Hebrides (South Pacific): taxonomy, zoogeography, host association and ecology. Q. Jl. Taiwan Mus., 34, 213–32Google Scholar
MacCormack, C. P. (1984) Human ecology and behaviour in malaria control in tropical Africa. Bull. WHO, 62, 81–7Google ScholarPubMed
Macdonald, W. W. (1962a) The genetic basis of susceptibility to infection with semi-periodic Brugia malayi in Aedes aegypti. Ann. Trop. Med. Parasit., 56, 373–82CrossRefGoogle Scholar
Macdonald, W. W. (1962b) The selection of a strain of Aedes aegypti susceptible to infection with semi-periodic Brugia malayi. Ann. Trop. Med. Parasit., 56, 368–72CrossRefGoogle Scholar
Macdonald, W. W. (1963) Further studies on a strain of Aedes aegypti susceptible to infection with sub-periodic Brugia malayi. Ann. Trop. Med. Parasit., 57, 452–60CrossRefGoogle ScholarPubMed
Macdonald, W. W. and Ramachandran, A. (1965) The influence of the gene fm (filarial susceptibility, Brugia malayi) on the susceptibility of Aedes aegypti to several strains of Brugia, Wuchereria and Dirofilaria. Annals of Tropical Medicine and Parasitology, 59, 64–73CrossRefGoogle ScholarPubMed
Mackie, F. P. (1907) The part played by Pediculus corporis in the transmission of relapsing fever. British Medical Journal, 2, 1706–9CrossRef
Macvicker, J. A. K., Billingsley, P. F., Djamgoz, M. B. A. and Harrow, I. D. (1994) Ouabain-sensitive Na+/K+-ATPase activity in the reservoir zone of the midgut of Stomoxys calcitrans (Diptera, Muscidae). Insect Biochemistry and Molecular Biology, 24, 151–9CrossRefGoogle Scholar
Maddrell, S. H. P. (1963) Control of ingestion in Rhodnius prolixus Stal. Nature, 198, 210CrossRefGoogle Scholar
Maddrell, S. H. P. (1980) Characteristics of epithelial transport in insect Malpighian tubules. Curr. Topics Memb. Transport, 14, 427–63CrossRefGoogle Scholar
Magesa, S. M., Mdira, Y. K., Akida, J. A., Bygbjerg, I. C. and Jakobsen, P. H. (2000) Observations on the periodicity of Plasmodium falciparum gametocytes in natural human infections. Acta Trop., 76, 239–46CrossRefGoogle ScholarPubMed
Mahmood, F. (2000) Susceptibility of geographically distinct Aedes aegypti L. from Florida to Dirofilaria immitis (Leidy) infection. J. Vector Ecol., 25, 36–47Google ScholarPubMed
Mahon, R. and Gibbs, A. (1982) Arbovirus-infected hens attract more mosquitoes. In MacKenzie, J. S. (ed.), Viral Diseases in South Esat Asia and the Western Pacific. New York: Academic PressGoogle Scholar
Maier, W. A. and Omer, O. (1973) Der einfluss von Plasmodium cathemerium auf den Aminosauregehalt und die eizahl von Culex pipiens fatigans. Z. Parasit., 42, 265–78CrossRefGoogle Scholar
Malhotra, I., Ouma, J. H., Wamachi, A. et al. (2003) Influence of maternal filariasis on childhood infection and immunity to Wuchereria bancrofti in Kenya. Infection and Immunity, 71, 5231–7CrossRefGoogle ScholarPubMed
Mallon, E. B., Loosli, R. and Schmid-Hempel, P. (2003) Specific versus nonspecific immune defense in the bumblebee, Bombus terrestris L. Evolution, 57, 1444–7Google ScholarPubMed
Mans, B. J., Louw, A. I. and Neitz, A. W. H. (2002) Evolution of hematophagy in ticks: common origins for blood coagulation and platelet aggregation inhibitors from soft ticks of the genus Ornithodoros. Molecular Biology and Evolution, 19, 1695–705CrossRefGoogle ScholarPubMed
Manson, P. (1878) On the development of Filaria sanguinis hominis, and on the mosquito considered as a nurse. J. Linn. Soc. Zool. London, 14, 304–11CrossRefGoogle Scholar
Marchoux, E. and Salinberi, A. (1903) La spirillose des poules. Annales de la Institut Pasteur, 17, 569–80
Margalit, J., Galun, R. and Rice, M. J. (1972) Mouthpart sensilla of the tsetse fly and their function. I. Feeding patterns. Ann. Trop. Med. Parasit., 66, 525–36CrossRefGoogle ScholarPubMed
Marshall, A. G. (1981) The Ecology of Ectoparasitic Insects. New York: Academic Press
Marx, R. (1955) Über die wirtsfindung und die Bedeutung de artspezifischen duftstoffes bei Cimex lectularius Linne. Z. Parasit., 17, 41–72CrossRefGoogle Scholar
Masaninga, F. and Mihok, S. (1999) Host influence on adaptation of Trypanosoma congolense metacyclics to vertebrate hosts. Med. Vet. Entomol., 13, 330–2CrossRefGoogle ScholarPubMed
Matsumoto, Y., Oda, Y., Uryu, M. and Hayakawa, Y. (2003) Insect cytokine growth-blocking peptide triggers a termination system of cellular immunity by inducing its binding protein. J. Biol. Chem., 278, 38579–85CrossRefGoogle ScholarPubMed
Matthysse, J. G. (1946) Cattle lice: their biology and control. Cornell Agr. Exp. Sta. Bull, 832, 1–67Google Scholar
Mattingley, P. F. (1965) The evolution of parasite–arthropod vector systems. In Taylor, A. E. R. (ed.), Symposium of the British Society for Parasitology. Oxford:Blackwell, Vol. 3Google Scholar
Maudlin, I. and Dukes, P. (1985) Extrachromosomal inheritance of susceptibility to trypanosome infection in tsetse flies. I. Selection of susceptible and refractory strains of Glossina morsitans morsitans. Ann. Trop. Med. Parasit., 79, 317–24CrossRefGoogle Scholar
Maudlin, I. and Ellis, D. (1985) Association between intracellular rickettsia-like infections of midgut cells and susceptibility to trypanosome infections in Glossina species. Z. Parasit., 71, 683–7CrossRefGoogle Scholar
Maudlin, I., Kabayo, J. P., Flood, M. E. T. and Evans, D. A. (1984) Serum factors and the maturation of Trypanosoma congolense infections in Glossina morsitans. Z. Parasit., 70, 11–19CrossRefGoogle ScholarPubMed
Maudlin, I. and Welburn, S. C. (1987) Lectin-mediated establishment of midgut infections of Trypanosoma congolense and Trypanosoma bruce in Glossina morsitans. Tropical Medicine and Parasitology, 38, 167–70Google Scholar
Maudlin, I., Welburn, S. C. and Milligan, P. J. M. (1998) Trypanosome infections and survival in tsetse. Parasitology, 116, S23–S28CrossRefGoogle ScholarPubMed
Mayer, M. S. and James, J. D. (1969) Attraction of Aedes aegypti (L.): responses to human arms, carbon dioxide, and air currents in a new type of olfactometer. Bull. Ent. Res., 58, 629–42CrossRefGoogle Scholar
Mayer, M. S. and James, J. D. (1970) Attraction of Aedes aegypti. II. Velocity of reaction to host with and without additional carbon dioxide. Ent. Exp. Appl., 13, 47–53CrossRefGoogle Scholar
Mbow, M. L., Bleyenberg, J. A., Hall, L. R. and Titus, R. G. (1998) Phlebotomus papatasi sand fly salivary gland lysate down-regulates a Th1, but up-regulates a Th2, response in mice infected with Leishmania major. Journal of Immunology, 161, 5571–7Google ScholarPubMed
McCabe, C. T. and Bursell, E. (1975a) Interrelationships between amino acid and lipid metabolism in the tsetse fly, Glossina morsitans. Insect Biochem., 5, 781–9CrossRefGoogle Scholar
McCabe, C. T. and Bursell, E. (1975b) Metabolism of digestive products in the tsetse fly, Glossina morsitans. Insect Biochem., 5, 769–79CrossRefGoogle Scholar
McCall, P. J. and Kelly, D. W. (2002) Learning and memory in disease vectors. Trends in Parasitology, 18, 429–33CrossRefGoogle ScholarPubMed
McCall, P. J. and Lemoh, P. A. (1997) Evidence for the ‘invitation effect’ during bloodfeeding by blackflies of the Simulium damnosum complex (Diptera, Simuliidae). Journal of Insect Behavior, 10, 299–303CrossRefGoogle Scholar
McCall, P. J., Mosha, F. W., Njunwa, K. J. and Sherlock, K. (2001) Evidence for memorized site-fidelity in Anopheles arabiensis. Trans. R. Soc. of Trop. Med. Hyg., 95, 587–90CrossRefGoogle ScholarPubMed
McDermott, M. J., Weber, E., Hunter, S. et al. (2000) Identification, cloning, and characterization of a major cat flea salivary allergen (Cte f 1). Molecular Immunology, 37, 361–75CrossRefGoogle Scholar
McGavin, G. C. (2001) Essential Entomology: An Order by Order Introduction. Oxford: Oxford University Press
McGreevy, P. B., Bryan, J. H., Oothuman, P. and Kolstrup, N. (1978) The lethal effects of the cibarial and pharyngeal armatures of mosquitoes on microfilariae. Trans. R. Soc. Trop. Med. Hyg., 74, 361–8CrossRefGoogle Scholar
McGreevy, P. B., McClelland, G. A. H. and Lavoipierre, M. M. J. (1974) Inheritance of susceptibility to Dirofilaria immitis infection in Aedes aegypti. Ann. Trop. Med. Parasit., 68, 97–109CrossRefGoogle ScholarPubMed
McKeever, S. (1977) Observations of Corethrella feeding on tree frogs (Hyla). Mosq. News, 37, 522Google Scholar
McKeever, S. and French, F. E. (1991) Corethrella (Diptera, Corethrellidae) of Eastern North-America – Laboratory Life-History and Field Responses to Anuran Calls. Ann. Ent. Soc. Am., 84, 493–7CrossRefGoogle Scholar
McKelvey, J. J. (1973) Man against Tsetse: Struggle for Africa. Ithaca: Cornell University Press
Mead-Briggs, A. R. (1964) The reproductive biology of the rabbit flea Spilopsyllus cuniculi (Dale) and the dependance of this species on the upon the breeding of its host. J. Exp. Biol., 41, 371–402Google Scholar
Medvedev, S. I. and Skylar, V. Y. (1974) Beetles (Coleoptera) from nests of small mammals in Donotsk Province (in Russian). Entomologicheskoe Obozrenie, 53, 561–71Google Scholar
Meijerink, J., Braks, M. A. H., Brack, A. A. et al. (2000) Identification of olfactory stimulants for Anopheles gambiae from human sweat samples. Journal of Chemical Ecology, 26, 1367–82CrossRefGoogle Scholar
Meister, M. and Lagueux, M. (2003) Drosophila blood cells. Cell Microbiol., 5, 573–80CrossRefGoogle ScholarPubMed
Mellink, J. J. (1981) Selections for blood-feeding efficiency in colonized Aedes aegypti. Mosq. News, 41, 119–25Google Scholar
Mellink, J. J. and Bovenkamp, W. (1981) Functional aspects of mosquito salivation in blood feeding in Aedes aegypti. Mosq. News, 41, 110–15Google Scholar
Mellor, P. S. and Boorman, J. (1980) Multiplication of the bluetongue virus in Culicoides nubeculosus (Meigen) simultaneously infected with the virus and microfilaria of Onchocerca cervicalis (Railliet and Henry). Ann. Trop. Med. Parasit., 74, 463–9CrossRefGoogle Scholar
Menezes, H. and Jared, C. (2002) Immunity in plants and animals: common ends through different means using similar tools. Comp. Biochem. Physiol. C Toxicol. Pharmacol., 132, 1–7CrossRefGoogle ScholarPubMed
Mews, A. R., Baumgartner, H., Luger, D. and Offori, E. D. (1976) Colonisation of Glossina morsitans morsitans Westw. in the laboratory using in vitro feeding techniques. Bull. Ent. Res., 65, 631–42CrossRefGoogle Scholar
Miall, R. C. (1978) The flicker fusion frequencies of six laboratory insects, and the response of the compound eye to mains fluorescent ‘ripple’. Phys. Ent., 3, 99–106CrossRefGoogle Scholar
Michel, T., Reichhart, J. M., Hoffmann, J. A. and Royet, J. (2001) Drosophila toll is activated by Gram-positive bacteria through a circulating peptidoglycan recognition protein. Nature, 414, 756–9CrossRefGoogle ScholarPubMed
Miller, N. and Lehane, M. J. (1990) In vitro perfusion studies on the peritrophic membrane of the tsetse fly Glossina morsitans morsitans (Diptera: Glossinidae). J. Insect Phys., 36, 813–18CrossRefGoogle Scholar
Minchella, D. J. (1985) Host life-history variation in response to parasitism. Parasitol., 90, 205–16CrossRefGoogle Scholar
Minchella, D. J. and Loverde, P. T. (1983) Laboratory comparison of the relative success of Biomphalaria glabrata stocks which are susceptible and insusceptible to infection with Schistosoma mansoni. Parasitology, 86, 335–44CrossRefGoogle ScholarPubMed
Mitchell, B. K. and Reinouts van Haga-Kelker, H. A. (1976) A comparison of the feeding behaviour in teneral and post-teneral Glossina morsitans (Diptera, Glossinidae) using an artificial membrane. Ent. Exp. Appl., 20, 105–12CrossRefGoogle Scholar
Mockford, E. L. (1967) Some Psocoptera from the plumage of birds. Proc. Ent. Soc. Washington, 69, 307–9Google Scholar
Mockford, E. L. (1971) Psocoptera from the dusky-footed wood rat in southern California (Psocoptera: Atropidae, Psoguillidae, Liposcelidae). Pan-Pacific Entomologist, 47, 127–40Google Scholar
Moffatt, M. R., Blakemore, D. and Lehane, M. J. (1995) Studies on the synthesis and secretion of digestive trypsin in Stomoxys calcitrans (Insecta-Diptera). Comp. Biochem. Phys. B, 110B, 291–300CrossRefGoogle Scholar
Mohr, C. O. (1943) Cattle droppings as ecological units. Ecol. Monographs, 13, 275CrossRefGoogle Scholar
Moloo, S. K. (1983) Feeding behaviour of Glossina morsitans morsitans infected with Trypanosoma vivax, T. congolense or T. brucei. Parasit., 86, 51–6CrossRefGoogle ScholarPubMed
Moloo, S. K. and Dar, F. (1985) Probing by Glossina morsitans centralis infected with pathogenic Trypanosoma species. Trans. R. Soc. of Trop. Med. Hyg., 79, 119CrossRefGoogle ScholarPubMed
Moloo, S. K. and Kutuza, S. B. (1970) Feeding and crop-emptying in Glossina brevipalpis Newstead. Acta Trop., 27, 356–77Google ScholarPubMed
Moloo, S. K., Sabwa, C. L. and Baylis, M. (2000) Feeding behaviour of Glossina pallidipes and G. morsitans centralis on Boran cattle infected with Trypanosoma congolense or T. vivax under laboratory conditions. Med. Vet. Entomol., 14., 290–9CrossRefGoogle Scholar
Molyneux, D. H. (1984) Evolution of the Trypanosomatidae: considerations of polyphyletic origins of mammalian parasites. CNRS/INSERM, 1986, 231–40Google Scholar
Molyneux, D. H., Bradley, M., Hoerauf, A., Kyelem, D. and Taylor, M. J. (2003) Mass drug treatment for lymphatic filariasis and onchocerciasis. Trends in Parasitology, 19, 516–22CrossRefGoogle ScholarPubMed
Molyneux, D. H. and Killick-Kendrick, R. (1987) Morphology, ultrastructure and life cycles. In Peters, W. and Killick-Kendrick, R. (eds.), The Leishmaniases in Biology and Medicine. New York: Academic PressGoogle Scholar
Molyneux, D. H., Killick-Kendrick, R. and Ashford, R. W. (1975) Leishmania in phlebotomid sandflies. III. The ultrastructure of Leishmania mexicana amazonensis in the midgut and pharynx of Lutzomyia longipalpis. Proc. R. Soc. B, 190, 341–57CrossRefGoogle ScholarPubMed
Montfort, W. R., Weichsel, A. and Andersen, J. F. (2000) Nitrophorins and related antihemostatic lipocalins from Rhodnius prolixus and other blood-sucking arthropods. Biochimica et Biophysica Acta – Protein Structure and Molecular Enzymology, 1482, 110–18CrossRefGoogle ScholarPubMed
Mooring, M. S., Benjamin, J. E., Harte, C. R. and Herzog, N. B. (2000) Testing the interspecific body size principle in ungulates: the smaller they come, the harder they groom. Anim. Behav., 60, 35–45CrossRefGoogle ScholarPubMed
Mooring, M. S. and Hart, B. L. (1992) Animal grouping for protection from parasites – selfish herd and encounter-dilution effects. Behaviour, 123, 173–93CrossRefGoogle Scholar
Morand, S. and Poulin, R. (1998) Density, body mass and parasite species richness of terrestrial mammals. Evolutionary Ecology, 12, 717–27CrossRefGoogle Scholar
Moret, Y. and Schmid-Hempel, P. (2000) Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 1166–8CrossRefGoogle Scholar
Moro, O. and Lerner, E. A. (1997) Maxadilan, the vasodilator from sand flies, is a specific pituitary adenylate cyclase activating peptide type I receptor agonist. J. Biol. Chem., 272, 966–70CrossRefGoogle ScholarPubMed
Morris, R. V., Shoemaker, C. B., David, J. R., Lanzaro, G. C. and Titus, R. G. (2001) Sandfly maxadilan exacerbates infection with Leishmania major and vaccinating against it protects against L. major infection. J. Immunol., 167, 5226–30CrossRefGoogle ScholarPubMed
Moskalyk, L. A. and Friend, W. G. (1994) Feeding behavior of female Aedes aegypti – effects of diet, temperature, bicarbonate and feeding technique on the response to ATP. Phys. Ent., 19, 223–9CrossRefGoogle Scholar
Moyer, B. R., Gardiner, D. W. and Clayton, D. H. (2002) Impact of feather molt on ectoparasites: looks can be deceiving. Oecologia, 131, 203–10CrossRefGoogle ScholarPubMed
Msangi, A. R., Whitaker, C. J. and Lehane, M. J. (1998) Factors influencing the prevalence of trypanosome infection of Glossina pallidipes on the Ruvu flood plain of Eastern Tanzania. Acta Trop., 70, 143–55CrossRefGoogle ScholarPubMed
Muir, L. E., Thorne, M. J. and Kay, B. H. (1992) Aedes aegypti (Diptera, Culicidae) vision – spectral sensitivity and other perceptual parameters of the female eye. J. Med. Ent., 29, 278–81CrossRefGoogle ScholarPubMed
Mukabana, W. R., Takken, W. and Knols, B. G. J. (2002a) Analysis of arthropod bloodmeals using molecular genetic markers. Trends in Parasitology, 18, 505–9CrossRefGoogle Scholar
Mukabana, W. R., Takken, W., Seda, P. et al. (2002b) Extent of digestion affects the success of amplifying human DNA from blood meals of Anopheles gambiae (Diptera: Culicidae). Bull. Ent. Res., 92, 233–9CrossRefGoogle Scholar
Mukerji, D. and Sen-Sarma, P. (1955) Anatomy and affinity of the elephant louse, Haematomyzus elephantis Piaget (Insecta: Rhyncophthiraptera). Parasitol., 45, 5–30CrossRefGoogle Scholar
Mukwaya, L. G. (1977) Genetic control of feeding preference in the mosquitoes Aedes (Stegomyia) simpsoni and aegypti. Phys. Ent., 2, 133–45CrossRefGoogle Scholar
Mullens, B. A. and Gerhardt, R. R. (1979) Feeding behaviour of some Tennessee Tabanidae. Environ. Ent., 8, 1047–51CrossRefGoogle Scholar
Muller, H. M., Catteruccia, F., Vizioli, J., DellaTorre, A. and Crisanti, A. (1995) Constitutive and blood meal-induced trypsin genes in Anopheles gambiae. Experimental Parasitology, 81, 371–85CrossRefGoogle ScholarPubMed
Muller, H. M., Crampton, J. M., Dellatorre, A., Sinden, R. and Crisanti, A. (1993) Members of a trypsin gene family in Anopheles gambiae are induced in the gut by blood meal. EMBO J., 12, 2891–900Google ScholarPubMed
Mumcuoglu, Y. and Galun, R. (1987) Engorgement response of human body lice Pediculus humanus (Insecta: Anoplura) to blood fractions and their components. Phys. Ent., 12, 171–4CrossRefGoogle Scholar
Munstermann, L. E. and Conn, J. E. (1997) Systematics of mosquito disease vectors (Diptera, Culicidae): impact of molecular biology and cladistic analysis. Ann. Rev. Ent., 42, 351–69CrossRefGoogle ScholarPubMed
Murlis, J., Willis, M. A. and Carde, R. T. (2000) Spatial and temporal structures of pheromone plumes in fields and forests. Phys. Ent., 25, 211–22CrossRefGoogle Scholar
Murray, M. D. (1957) The distribution of the eggs of mammalian lice on their hosts. II. Analysis of the oviposition behaviour of Damalinia ovis. Aust. J. Zool, 5, 19–29CrossRefGoogle Scholar
Murray, M. D. (1963) Influence of temperature on the reproduction of Damalinia equi (Denny). Aust. J. Zool., 11, 183–9CrossRefGoogle Scholar
Murray, M. D. (1987) Effects of host grooming on louse populations. Parasitology Today, 3, 276–8CrossRefGoogle ScholarPubMed
Murray, M. D. and Nicholls, D. G. (1965) Studies on the ectoparasites of seals and penguins. I. The ecology of the louse Lepidophthirus macrorhini Enderlein on the southern elephant seal, Mirounga leonina (L.). Aust. J. Zool., 13, 437–54CrossRefGoogle Scholar
Mwandawiro, C., Boots, M., Tuno, N. et al. (2000) Heterogeneity in the host preference of Japanese encephalitis vectors in Chiang Mai, northern Thailand. Trans. R. Soc. Trop. Med. Hyg., 94, 238–42CrossRefGoogle Scholar
Naksathit, A. T., Edman, J. D. and Scott, T. W. (1999) Utilization of human blood and sugar as nutrients by female Aedes aegypti (Diptera: Culicidae). J. Med. Ent, 36, 13–17CrossRefGoogle Scholar
Naksathit, A. T. and Scott, T. W. (1998) Effect of female size on fecundity and survivorship of Aedes aegypti fed only human blood versus human blood plus sugar. J. Am. Mosq. Control Assoc., 14, 148–52Google Scholar
Napier Bax, S. (1937) The senses of smell and sight in Glossina swynnertoni. Bull. Ent. Res., 28, 539–82CrossRefGoogle Scholar
Nappi, A. J., Vass, E., Frey, F. and Carton, Y. (2000) Nitric oxide involvement in Drosophila immunity. Nitric Oxide Biology and Chemistry, 4, 423–30CrossRefGoogle ScholarPubMed
Nasci, R. S. (1982) Differences in host choice between the sibling species of treehole mosquitoes Aedes triseriatus and Aedes hendersoni. Am. J. Trop. Med. Hyg., 31, 411–15CrossRefGoogle ScholarPubMed
Nelson, R. L. (1965) Carbon dioxide as an attractant for Culicoides. J. Med. Ent., 2, 56–7CrossRefGoogle ScholarPubMed
Nelson, W. A. (1987) Other blood-sucking and myiasis-producing arthropods. In Soulsby, E. J. L. (ed.), Immune Responses in Parasitic Infections: Immunology, Immunopathology and Immunoprophylaxis.Boca Raton, Florida: CRC Press, Vol. IVGoogle Scholar
Nelson, W. A., Bell, J. F., Clifford, C. M. and Keirans, A. J. (1977) Interaction of ectoparasites and their hosts. J. Med. Ent., 13, 389–428CrossRefGoogle ScholarPubMed
Nelson, W. A., Keirans, J. E., Bell, J. F. and Clifford, C. M. (1975) Host–ectoparasite relationships. J. Med. Ent., 12, 143–66CrossRefGoogle ScholarPubMed
Nelson, W. A. and Kozub, G. C. (1980) Melophagus ovinus (Diptera: Hippoboscidae): evidence of local mediation in acquired resistance of sheep to keds. J. Med. Ent., 17, 291–7CrossRefGoogle Scholar
Newson, R. M. and Holmes, R. G. (1968) Some ectoparasites of the coypu (Myocastor coypus) in eastern England. J. Anim. Ecol., 37, 471–81CrossRefGoogle Scholar
Nguu, E. K., Osir, E. O., Imbuga, M. O. and Olembo, N. K. (1996) The effect of host blood in the in vitro transformation of bloodstream trypanosomes by tsetse midgut homogenates. Med. Vet. Entomol., 10, 317–22CrossRefGoogle ScholarPubMed
Niare, O., Markianos, K., Volz, J. et al. (2002) Genetic loci affecting resistance to human malaria parasites in a west African mosquito vector population. Science, 298, 213–16CrossRefGoogle Scholar
Nieves, E. and Pimenta, P. F. P. (2002) Influence of vertebrate blood meals on the development of Leishmania (Viannia) braziliensis and Leishmania (Leishmania) amazonensis in the sand fly Lutzomyia migonei (Diptera: Psychodidae). Am. J. Trop. Med. Hyg., 67, 640–7CrossRefGoogle Scholar
Nigam, Y. and Ward, R. D. (1991) The effect of male sandfly pheromone and host factors as attractants for female Lutzomyia longipalpis (Diptera, Psychodidae). Phys. Ent., 16, 305–12CrossRefGoogle Scholar
Nogge, G. (1978) Aposymbiotic tsetse flies, Glossina morsitans morsitans obtained by feeding on rabbits immunized specifically with symbionts. J. Insect Physiol., 24, 299–304CrossRefGoogle ScholarPubMed
Nogge, G. (1981) Significance of symbionts for the maintenance of an optimal nutritional state for successful reproduction in haematophagous arthropods. Parasitology, 82, 101–4Google Scholar
Nogge, G. and Ritz, R. (1982) Number of symbionts and its regulation in tsetse flies, Glossina spp. Ent. Exp. Appl., 31, 249–54CrossRefGoogle Scholar
Noriega, F. G., Edgar, K. A., Bechet, R. and Wells, M. A. (2002) Midgut exopeptidase activities in Aedes aegypti are induced by blood feeding. J. Insect Physiol., 48, 205–12CrossRefGoogle ScholarPubMed
Noriega, F. G. and Wells, M. A. (1999) A molecular view of trypsin synthesis in the midgut of Aedes aegypti. J. Insect Physiol., 45, 613–20CrossRefGoogle ScholarPubMed
Nuttal, G. H. F. (1899) On the role of insects, arachnids, and myriapods as carriers in the spread of bacterial and parasitic disease of man and animals. A critical and historical study. Johns Hopkins Hospital Reports, 8, 1–154Google Scholar
Obiamiwe, B. A. and Macdonald, W. W. (1973) 1. The effect of heparin on the migration of Brugia pahangi microfilariae Culex pipiens. 2. The uptake of B. pahangi microfilariae in C. pipiens and the infectivity of C. pipiens in relation to microfilarial densities. 3. Evidence of a sex-linked recessive gene, sb, controlling susceptibility of C. pipiens to B. pahangi. Trans. R. Soc. Trop. Med. Hyg., 67, 32–3CrossRefGoogle Scholar
Ochiai, M., Niki, T. and Ashida, M. (1992) Immunocytochemical localization of beta-1,3-glucan recognition protein in the silkworm, Bombyx mori. Cell and Tissue Research, 268, 431–7CrossRefGoogle ScholarPubMed
Ogston, C. W. and London, W. T. (1980) Excretion of hepatitis B surface antigen by the bedbug Cimex hemipterus Fabr. Trans. R. Soc. Trop. Med. Hyg., 74, 823–5CrossRefGoogle ScholarPubMed
Olubayo, R. O., Mihok, S., Munyoki, E. and Otieno, L. H. (1994) Dynamics of host blood effects in Glossina morsitans spp. infected with Trypanosoma congolense and Trypanosoma brucei. Parasitology Research, 80, 177–81CrossRefGoogle Scholar
Meara, G. F. (1979) Variable expressions of autogeny in three mosquito species. Int. J. Invert. Reprod., 1, 253–61CrossRefGoogle Scholar
O'Meara, G. F. (1985) Ecology and autogeny in mosquitoes. In Lounibos, L. P., Rey, J. R. and Frank, J. H.,(eds.), Ecology of Mosquitoes. Florida: Florida Medical LaboratoryGoogle Scholar
O'Meara, G. F. (1987) Nutritional ecology of blood feeding diptera. In Slansky, F. and Rodriguez, J. G. (eds.),Nutritional Ecology of Insects, Mites, Spiders and Related Invertebrates. New York: WileyGoogle Scholar
Meara, G. F. and Edman, J. D. (1975) Autogenous egg production in the salt marsh mosquito, Aedes taeniorrhynchus. Biol. Bull., 149, 384–96CrossRefGoogle Scholar
Meara, G. F. and Evans, D. G. (1973) Blood-feeding requirements of the mosquito: geographical variation in Aedes taeniorhynchus. Science, 180, 1291–3CrossRefGoogle Scholar
Meara, G. F. and Evans, D. G. (1976) The influence of mating on autogenous egg development in the mosquito, Aedes taeniorrhynchus. J. Insect Physiol., 22, 613–17CrossRefGoogle Scholar
Omer, S. M. and Gillies, M. T. (1971) Loss of response to carbon dioxide in palpectomized female mosquitoes. Ent. Exp. Appl., 14, 251–2Google Scholar
Osbrink, L. A. and Rust, M. A. (1985) Cat flea (Siphonaptera: Pulicidae): factors influencing host-finding behaviour in the laboratory. Ann. Ent. Soc. Am., 78, 29–34CrossRefGoogle Scholar
Shea, B., Rebollar-Tellez, E., Ward, R. D. et al. (2002) Enhanced sandfly attraction to Leishmania-infected hosts. Trans. R. Soc. Trop. Med. Hyg., 96, 117–18CrossRefGoogle Scholar
Overal, W. L. (1980) Biology and behaviour of North American Trichobius bat flies (Diptera: Streblidae). Ph.D. thesis, University of Kansas
Overal, W. L. and Wingate, L. R. (1976) The biology of the batbug Strictimex antennatus (Hemiptera: Cimicidae) in South Africa. Ann. Natal Mus., 22, 821–8Google Scholar
Owaga, M. L. and Challier, A. (1985) Catch composition of the tsetse Glossina pallidipes Austen in revolving and stationary traps with respect to age, sex ratio and hunger stage. Insect Sci. Applic., 6, 711–18Google Scholar
Page, R. D. M., Clayton, D. H. and Paterson, A. M. (1996) Lice and cospeciation: a response to Barker. Int. J. Parasit., 26, 213–18CrossRefGoogle ScholarPubMed
Pagel, M. and Bodmer, W. (2003) A naked ape would have fewer parasites. Proc. R. Soc. Lond. B Biol. Sci., 270, Suppl 1, 117–19CrossRefGoogle ScholarPubMed
Pant, C. P., Houba, V. and Engers, H. D. (1987) Bloodmeal identification in vectors. Parasitology Today, 3, 324–6Google Scholar
Panton, L. J., Tesh, R. B., Nadeau, K. C. and Beverley, S. M. (1991) A test for genetic exchange in mixed infections of Leishmania major in the sand fly Phlebotomus papatasi. J. Protozool, 38, 224–8CrossRefGoogle ScholarPubMed
Pappas, L. G., Pappas, C. D. and Grossman, G. L. (1986) Hemodynamics of human skin during mosquito (Diptera: Culicidae) blood feeding. J. Med. Ent., 23, 581–7CrossRefGoogle ScholarPubMed
Parker, K. R. and Gooding, R. H. (1979) Effects of host anaemia, local skin factors and circulating antibodies upon biology of laboratory reared Glossina morsitans morsitans (Diptera: Glossinidae). Can. J. Zool., 57, 2393–401CrossRefGoogle Scholar
Paskewitz, S. M., Brown, M. R., Lea, A. O. and Collins, F. H. (1988) Ultrastructure of the encapsulation of Plasmodium cynomolgi (B-strain) on the midgut of a refractory strain of Anopheles gambiae. Journal of Parasitology, 74, 432–9CrossRefGoogle ScholarPubMed
Patton, W. S. and Craig, F. W. (1913) On certain haematophagous species of the genus Musca, with descriptions of two new species. Indian Journal of Medical Research, 1, 13–25Google Scholar
Peacock, A. J. (1981) Distribution of (Na+ K+)-ATPase activity in the mid-guts and hind-guts of adult Glossina morsitans and Sarcophaga nodosa and the hind-gut of Bombyx mori larvae. Comp. Biochem. Physiol. A, 69, 133–6CrossRefGoogle Scholar
Peacock, A. J. (1982) Effects of sodium transport inhibitors on diuresis and midgut (Na+ and K+) ATPase in the tsetse fly Glossina morsitans. J. Insect Physiol., 28, 553–8CrossRefGoogle Scholar
Pearman, J. V. (1960) Some African psocoptera found on rats. Entomologist, 93, 246–50Google Scholar
Pearson, T. W., Beecroft, R. P., Welburn, S. C. et al. (2000) The major cell surface glycoprotein procyclin is a receptor for induction of a novel form of cell death in African trypanosomes in vitro. Molecular and Biochemical Parasitology, 111, 333–49CrossRefGoogle ScholarPubMed
Pell, P. E. and Southern, D. I. (1976) Effect of the coccidiostat, sulphaquinoxline, on symbiosis in the tsetse fly, Glossina species. Microbios Letters, 2, 203–11Google Scholar
Pereira, H., Penido, C. M., Martins, M. S. and Diotaiuti, L. (1998) Comparative kinetics of bloodmeal intake by Triatoma infestans and Rhodnius prolixus, the two principal vectors of Chagas disease. Med. Vet. Entomol., 12, 84–8CrossRefGoogle ScholarPubMed
Pereira, M. E. A., Andrade, A. F. B. and Ribeiro, J. M. C. (1981) Lectins of distinct specificity in Rhodnius prolixus interact selectively with Trypanosoma cruzi. Science, 211, 597–9CrossRefGoogle ScholarPubMed
Perrin, N., Christe, P. and Richner, H. (1996) On host life-history response to parasitism. Oikos, 75, 317–20CrossRefGoogle Scholar
Peschken, D. P. and Thorsteinson, A. J. (1965) Visual orientation of black flies (Simuliidae: Diptera) to colour, shape and movement of targets. Ent. Exp. Appl., 8, 282–8CrossRefGoogle Scholar
Peters, W. (1968) Vorkommen, Zusammensetzung und feinstruktur peritrophischer membranen im tierreich. Zeit. Morph. Okol. Tiere., 64, 21–58CrossRefGoogle Scholar
Peters, W., Kolb, H. and Kolb-Bachofen, V. (1983) Evidence for a sugar receptor (lectin) in the peritrophic membrane of the blowfly larva, Calliphora erythrocephala Mg. (Diptera). J. Insect Physiol., 29, 275–80CrossRefGoogle Scholar
Peters, W., Zimmermann, U. and Becker, B. (1973) Investigations on the transport function and structure of peritrophic membranes. IV. Anisotropic cross bands in peritrophic membranes of Diptera. J. Insect Physiol., 19, 1067–77CrossRefGoogle Scholar
Peterson, D. G. and Brown, A. W. A. (1951) Studies of the responses of female Aedes mosquito. III. The response of Aedes aegypti (L.) to a warm body and its radiation. Bull. Ent. Res., 42, 535–41CrossRefGoogle Scholar
Phelps, R. J. and Vale, G. A. (1976) Studies on the local distribution and on the methods of host location of some Rhodesian Tabanidae (Diptera). J. Ent. Soc. S. Afr., 39, 67–81Google Scholar
Pichon, G., Awono-Ambene, H. P. and Robert, V. (2000) High heterogeneity in the number of Plasmodium falciparum gametocytes in the bloodmeal of mosquitoes fed on the same host. Parasitology, 121, 115–20CrossRefGoogle ScholarPubMed
Piot, P. and Schofield, C. J. (1986) No evidence for arthropod transmission of AIDS. Parasitology Today, 2, 294–5CrossRefGoogle Scholar
Platt, K. B., Linthicum, K. J., Myint, K. S. A. et al. (1997) Impact of dengue virus infection on feeding behavior of Aedes aegypti. Am. J. Trop. Med. Hyg., 57, 119–25CrossRefGoogle ScholarPubMed
Politzar, H. and Merot, P. (1984) Attraction of the tsetse fly Glossina morsitans submorsitans to acetone, 1 octen-3-ol, and the combination of these compounds in west Africa. Rev. Elev. Med. Vet. Pays Trop., 37, 468–73Google ScholarPubMed
Ponnudurai, T., Billingsley, P. F. and Rudin, W. (1988) Differential infectivity of Plasmodium for mosquitoes. Parasitology Today, 4, 319–21CrossRefGoogle ScholarPubMed
Port, G. R., Bateham, P. F. L. and Bryan, J. H. (1980) The relationship of host size to feeding by mosquitoes of the Anopheles gambiae complex (Diptera: Culicidae). Bull. Ent. Res., 70, 133–44CrossRefGoogle Scholar
Pospisil, J. and Zdarek, J. (1965) On the visual orientation of the stable fly (Stomoxys calcitrans L.) to colours. Acta Entomol. Bohemoslov., 62, 85–91Google Scholar
Powell, J. R., Petrarca, V., Torre, A., Caccone, A. and Coluzzi, M. (1999) Population structure, speciation, and introgression in the Anopheles gambiae complex. Parasitologia, 41, 101–13Google ScholarPubMed
Price, G. D., Smith, N. and Carlson, D. A. (1979) The attraction of female mosquitoes (Anopheles quadrimaculatus Say) to stored human emanations in conjunction with adjusted levels of relative humidity, temperature and carbon dioxide. J. Chem. Ecol., 5, 383–95CrossRefGoogle Scholar
Price, R. D. (1975) The Menacanthus eurysternus complex (Mallophaga: Menoponidae) of the Passeriformes and Piciformes (Aves). Ann. Ent. Soc. Am., 68, 617–22CrossRefGoogle Scholar
Prior, A. and Torr, S. J. (2002) Host selection by Anopheles arabiensis and An. quadriannulatus feeding on cattle in Zimbabwe. Med. Vet. Entomol., 16, 207–13CrossRefGoogle ScholarPubMed
Raikhel, A. S., Kokoza, V. A., Zhu, J. et al. (2002) Molecular biology of mosquito vitellogenesis: from basic studies to genetic engineering of antipathogen immunity. Insect Biochem. Mol. Biol., 32, 1275–86CrossRefGoogle ScholarPubMed
Ramet, M., Lanot, R., Zachary, D. and Manfruelli, P. (2002a) JNK signaling pathway is required for efficient wound healing in Drosophila. Dev. Biol., 241, 145–56CrossRefGoogle Scholar
Ramet, M., Manfruelli, P., Pearson, A., Mathey-Prevot, B. and Ezekowitz, R. A. B. (2002b) Functional genomic analysis of phagocytosis and identification of a Drosophila receptor for E-coli. Nature, 416, 644–8CrossRefGoogle Scholar
Ratcliffe, N. A. and Rowley, A. F. (1979) Role of hemocytes in defence against biological agents. In Gupta, A. P. (ed.),Insect Hemocytes. Cambridge: Cambridge University Press, 331–422CrossRefGoogle Scholar
Ratzlaff, R. E. and Wikel, S. K. (1990) Murine immune responses and immunization against Polyplax serrata (Anoplura: Polyplacidae). J. Med. Ent., 27, 1002–7CrossRefGoogle Scholar
Ready, P. D. (1978) The feeding habits of laboratory bred Lutzomyia longipalpis (Diptera: Psychodidae). J. Med. Ent., 14, 545–552CrossRefGoogle Scholar
Reddy, V. B., Kounga, K., Mariano, F. and Lerner, E. A. (2000) Chrysoptin is a potent glycoprotein IIb/IIIa fibrinogen receptor antagonist present in salivary gland extracts of the deerfly. Journal of Biological Chemistry, 275, 15861–7CrossRefGoogle ScholarPubMed
Read, W., Carrall, J., Agramonte, A. and Lazear, J. (1900) The etiology of yellow fever: a preliminary note. The Philadelphia Medical Journal, 6, 790–3
Reichardt, T. R. and Galloway, T. D. (1994) Seasonal occurrence and reproductive status of Opisocrostis bruneri (Siphonaptera, Ceratophyllidae), a flea on franklin ground-squirrel, Spermophilus franklinii (Rodentia, Sciuridae) near Birds Hill Park, Manitoba. J. Med. Ent., 31, 105–13CrossRefGoogle Scholar
Reid, G. D. F. and Lehane, M. J. (1984) Peritrophic membrane formation in three temperate simuliids, Simulium ornatum, S. equinum and S. lineatum with respect to the migration of Onchocercal microfilariae. Ann. Trop. Med. Parasit., 78, 527–39CrossRefGoogle ScholarPubMed
Reinouts van Haga, H. A. and Mitchell, B. K. (1975) Temperature receptors on tarsi of the tsetse fly Glossina morsitans West. Nature, 255, 225–6CrossRefGoogle ScholarPubMed
Reunala, T., Brummer-Korvenkontio, H., Lappalainen, P., Rasanen, L. and Palosuo, T. (1990) Immunology and treatment of mosquito bites. Clin. Exp. Allergy, 20, Suppl 4, 19–24CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. (1982) The anti-serotonin and antihistamine activities of salivary secretion of Rhodnius prolixus. J. Insect Physiol., 28, 69–75CrossRefGoogle Scholar
Ribeiro, J. M. C. (1987) Role of saliva in blood-feeding by arthropods. Ann. Rev. Ent., 32, 463–78CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. (1995) Blood-feeding arthropods – live syringes or invertebrate pharmacologists. Infectious Agents and Disease – Reviews Issues and Commentary, 4, 143–52Google ScholarPubMed
Ribeiro, J. M. C. (1998) Rhodnius prolixus salivary nitrophorins display heme-peroxidase activity. Insect Biochemistry and Molecular Biology, 28, 1051–7CrossRefGoogle Scholar
Ribeiro, J. M. C., Charlab, R., Rowton, E. D. and Cupp, E. W. (2000) Simulium vittatum (Diptera: Simuliidae) and Lutzomyia longipalpis (Diptera: Psychodidae) salivary gland hyaluronidase activity. J. Med. Ent., 37, 743–7CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C., Charlab, R. and Valenzuela, J. G. (2001) The salivary adenosine deaminase activity of the mosquitoes Culex quinquefasciatus and Aedes aegypti. J. Exp. Biol., 204, 2001–10Google ScholarPubMed
Ribeiro, J. M. C. and Francischetti, I. M. B. (2003) Role of arthropod saliva in blood feeding: sialome and post-sialome perspectives. Ann. Rev. Ent., Vol. 48, 73–88CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. and Garcia, E. S. (1981a) Platelet antiaggregating activity in the salivary secretion of the blood-sucking bug Rhodnius prolixus. Experientia, 37, 384–5CrossRefGoogle Scholar
Ribeiro, J. M. C. and Garcia, E. S. (1981b) The role of saliva in feeding in Rhodnius prolixus. J. Exp. Biol., 94, 219–30Google Scholar
Ribeiro, J. M. C., Katz, O., Pannell, L. K., Waitumbi, J. and Warburg, A. (1999) Salivary glands of the sand fly Phlebotomus papatasi contain pharmacologically active amounts of adenosine and 5-AMP. J. Exp. Biol., 202, 1551–9Google ScholarPubMed
Ribeiro, J. M. C., Rossignol, P. A. and Spielman, A. (1985) Salivary gland apyrase determines probing time in anopheline mosquitoes. J. Insect Physiol., 9, 689–92CrossRefGoogle Scholar
Ribeiro, J. M. C., Schneider, M. and Guimaraes, J. A. (1995) Purification and characterization of prolixin-S (nitrophorin-2), the salivary anticoagulant of the bloodsucking bug Rhodnius prolixus. Biochemical Journal, 308, 243–9CrossRefGoogle Scholar
Ribeiro, J. M. C. and Valenzuela, J. G. (1999) Purification and cloning of the salivary peroxidase/catechol oxidase of the mosquito Anopheles albimanus. J. Exp. Biol., 202, 809–16Google ScholarPubMed
Rice, M. J., Galun, R. and Margalit, J. (1973) Mouthpart sensilla of the tsetse fly and their function. II. Labial sensilla. Ann. Trop. Med. Parasit., 67, 101–7CrossRefGoogle ScholarPubMed
Richman, A. M., Dimopoulos, G., Seeley, D. and Kafatos, F. C. (1997) Plasmodium activates the innate immune response of Anopheles gambiae mosquitoes. EMBO J., 16, 6114–9CrossRefGoogle ScholarPubMed
Roberts, L. W. (1981) Probing by Glossina morsitans morsitans and transmission of Trypanosoma (Nannomonas) congolense. Am. J. Trop. Med. Hyg., 30, 948–51CrossRefGoogle ScholarPubMed
Roberts, R. H. (1972) Relative attractiveness of CO2 and a steer to Tabanidae, Culicidae, and Stomoxys calcitrans. Mosq. News, 32, 208–11Google Scholar
Roberts, R. H. (1977) Attractancy of two black decoys and CO2 to tabanids (Diptera: Tabanidae). Mosq. News, 37, 169–72Google Scholar
Robinson, A. (1939) The mouthparts and their function in the female mosquito, Anopheles maculipennis. Parasitol., 31, 212–42CrossRefGoogle Scholar
Rogers, K. A. and Titus, R. G. (2003) Immunomodulatory effects of Maxadilan and Phlebotomus papatasi sand fly salivary gland lysates on human primary in vitro immune responses. Parasite Immunol., 25, 127–34CrossRefGoogle ScholarPubMed
Rogers, M. E., Chance, M. L. and Bates, P. A. (2002) The role of promastigote secretory gel in the origin and transmission of the infective stage of Leishmania mexicana by the sandfly Lutzomyia longipalpis. Parasitology, 124, 495–507CrossRefGoogle ScholarPubMed
Rosenfeld, A. and Vanderberg, J. P. (1998) Identification of electrophoretically separated proteases from midgut and hemolymph of adult Anopheles stephensi mosquitoes. Journal of Parasitology, 84, 361–5CrossRefGoogle ScholarPubMed
Ross, R. (1897) On same peculier pigmented cells found in two mosquitoes fed on malaria blood. British Medical Journal, 2, 1786–8CrossRefGoogle Scholar
Ross, R. (1898) Report on the cultivation of protessoma, Labb, in grey mosquitoes. Indian Med. Gaz., 33, 401–8Google Scholar
Rossignol, P. A., Ribeiro, J. M. C., Jungery, M., Turell, M. J., Spielman, A. and Bailey, C. L. (1985) Enhanced mosquito blood-finding success on parasitaemic hosts: Evidence for vector-parasite mutualism. Proc. Nat. Acad. Sci., 82, 7725–7CrossRefGoogle Scholar
Rossignol, P. A., Ribeiro, J. M. C. and Spielman, A. (1984) Increased intradermal probing time in sporozoite-infected mosquitoes. Am. J. Trop. Med. Hyg., 33, 17–20CrossRefGoogle ScholarPubMed
Rossignol, P. A., Ribeiro, J. M. C. and Spielman, A. (1986) Increased biting rate and reduced fertility in sporozoite-infected mosquitoes. Am. J. Trop. Med. Hyg., 35, 277–9CrossRefGoogle ScholarPubMed
Rossignol, P. A. and Rossignol, A. M. (1988) Simulations of enhanced malaria transmission and host bias induced by modified vector blood location behaviour. Parasitol., 97, 363–72CrossRefGoogle ScholarPubMed
Rothschild, M. (1975) Recent advances in our knowledge of the Siphonoptera. Ann. Rev. Ent., 20, 241–59CrossRefGoogle Scholar
Rothschild, M. and Clay, T. (1952) Fleas, Flukes and Cuckoos. New York: Philosophical Library
Rothschild, M. and Ford, B. (1973) Factors influencing the breeding of the rabbit flea (Spilopsyllus cuniculi): a spring-time accelerator and a kairomone in nestling rabbit urine (with notes on Cediopsylla simplex, another ‘hormone bound’ species). J. Zool., 170, 87–137CrossRefGoogle Scholar
Rothschild, M., Schlein, Y., Parker, K. and Sternberg, S. (1972) Jump of the oriental rat flea Xenopsylla cheopis (Roths.). Nature, 239, 45–8CrossRefGoogle Scholar
Rowland, M. and Boersma, E. (1988) Changes in the spontaneous flight activity of the mosquito Anopheles stephensi by parasitization with the rodent malaria Plasmodium yoelii. Parasitology, 97, 221–7CrossRefGoogle ScholarPubMed
Rowland, M. W. and Lindsay, S. L. (1986) The circadian flight activity of Aedes aegypti parasitized with the filarial nematode Brugia pahangi. Phys. Ent., 11, 325–34CrossRefGoogle Scholar
Roy, D. N. (1936) On the role of blood in ovulation in Aedes aegypti, Linn. Bull. Ent. Res., 27, 423–9CrossRefGoogle Scholar
Royet, J., Meister, M. and Ferrandon, D. (2003) Humoral and cellular responses in Drosophilainnate immunity. In Ezekowitz, R. A. and Hoffman, J. A. (eds.), Infectious Disease: Innate Immunity. Totowa, NJ: Humana Press 137–53Google ScholarPubMed
Rubenstein, D. I. and Hohmann, M. E. (1989) Parasites and social-behavior of island feral horses. Oikos, 55, 312–20CrossRefGoogle Scholar
Rudin, W. and Hecker, H. (1979) Functional morphology of the midgut of Aedes aegypti L. (Insecta; Diptera) during blood digestion. Cell, 200, 193–203Google ScholarPubMed
Rudin, W. and Hecker, H. (1982) Functional morphology of the midgut of a sandfly as compared to other haematophagous nematocera. Tissue and Cell, 14, 751–8CrossRefGoogle Scholar
Rutberg, A. T. (1987) Horse fly harassment and the social-behavior of feral ponies. Ethology, 75, 145–54CrossRefGoogle Scholar
Sabelis, M. W. and Schippers, P. (1984) Variable wind direction and anemotactic strategies of searching for an odour plume. Oecologia, 63, 225–8CrossRefGoogle ScholarPubMed
Sacks, D. L. (1989) Metacyclogenesis in Leishmania promastigotes. Exp. Parasitol., 69, 100–3CrossRefGoogle ScholarPubMed
Sacks, D. L. and Kamhawi, S. (2001) Molecular aspects of parasite–vector and vector–host interactions in Leishmaniasis. Ann. Rev. Microbiol., 55, 453–83CrossRefGoogle ScholarPubMed
Sallum, M. A. M., Schultz, T. R., Foster, P. G. et al. (2002) Phylogeny of Anophelinae (Diptera: Culicidae) based on nuclear ribosomal and mitochondrial DNA sequences. Systematic Entomology, 27, 361–82CrossRefGoogle Scholar
Samuel, W. M. and Trainer, D. O. (1972) Lipoptena mazamae Rondani, 1878 (Diptera: Hippoboscidae) on white-tailed deer in southern Texas. J. Med. Ent., 9, 104–6CrossRefGoogle ScholarPubMed
Sandeman, R. M. (1996) Immune rsponses to mosquitoes and flies. In Wikel, S. K. (ed.), The Immunology of Host–Ectoparasitic Arthropod Interactions. Wallingford: CAB International 175–203Google Scholar
Sangiorgi, G. and Frosini, D. (1940) Di un principio emolitico (‘Cimicina’) nella saliva del Cimex lectularius. Pathologica, 32, 189–91Google Scholar
Saraiva, E. M., Pimenta, P. F., Brodin, T. N. (1995) Changes in lipophosphoglycan and gene expression associated with the development of Leishmania major in Phlebotomus papatasi. Parasitology, 111, (Pt 3), 275–87CrossRefGoogle ScholarPubMed
Sarkis, J. J. F., Guimaraes, J. A. and Ribeiro, J. M. C. (1986) Salivary apyrase of Rhodnius prolixus: kinetics and purification. Biochem. J., 233, 885–91CrossRefGoogle ScholarPubMed
Scaraffia, P. Y. and Wells, M. A. (2003) Proline can be utilized as an energy substrate during flight of Aedes aegypti females. J. Insect Physiol., 49, 591–601CrossRefGoogle ScholarPubMed
Schall, J. J. (2002) Parasite virulence. In Lewis, E. E., Campbell, J. F. and Sukdheo, M. D. K. (eds.), The Behavioural Ecology of Parasites. Wallingford: CAB InternationalCrossRefGoogle Scholar
Schiefer, B. A. et al. (1977) Plasmodium cynomolgi: effects of malaria infection on laboratory flight performance of Anopheles stephensi mosquiotoes. Exp. Parasitol., 41(2), 397–404CrossRef
Schlein, Y. (1977) Lethal effect of tetracycline on tsetse flies following damage to bacteroid symbionts. Experimentia, 33, 450–1CrossRefGoogle Scholar
Schlein, Y. and Jacobson, R. L. (1998) Resistance of Phlebotomus papatasi to infection with Leishmania donovani is modulated by components of the infective bloodmeal. Parasitology, 117, 467–73CrossRefGoogle ScholarPubMed
Schlein, Y., Warburg, A., Schnur, L. F. and Shlomai, J. (1983) Vector compatibility of Phlebotomus papatasi dependent on differentially induced digestion. Acta Trop., 40, 65–70Google ScholarPubMed
Schlein, Y., Yuval, B. and Warburg, A. (1984) Aggregation pheromone released from the palps of feeding female Phlebotomus papatasi (Psychodidae). J. Insect Physiol., 30, 153–6CrossRefGoogle Scholar
Schmid-Hempel, P. (2003) Immunology and evolution of infectious disease. Science, 300, 254CrossRefGoogle Scholar
Schmitz, H., Trenner, S., Hofmann, M. H. and Bleckmann, H. (2000) The ability of Rhodnius prolixus (Hemiptera; Reduviidae) to approach a thermal source solely by its infrared radiation. J. Insect Physiol., 46, 745–51CrossRefGoogle ScholarPubMed
Schoeler, G. B. and Wikel, S. K. (2001) Modulation of host immunity by haematophagous arthropods. Annals of Tropical Medicine and Parasitology, 95, 755–71CrossRefGoogle ScholarPubMed
Schofield, C. J. (1981) Chagas disease, triatomine bugs, and blood loss. Lancet, 1, 1316Google Scholar
Schofield, C. J. (1982) The role of blood intake in density regulation of populations of Triatoma infestans (Klug) (Hemiptera: Reduviidae). Bull. Ent. Res., 72, 617–29CrossRefGoogle Scholar
Schofield, C. J. (1985) Population dynamics and control of Triatoma infestans. Ann. Soc. Belge, Med. Trop., 65, 149–64Google ScholarPubMed
Schofield, C. J. (1988) Biosystematics of the triatominae. In Service, M. W. (ed.), Biosystematics of Haematophagous Insects. Oxford: Clarendon PressGoogle Scholar
Schofield, S. and Sutcliffe, J. F. (1996) Human individuals vary in attractiveness for host-seeking black flies (Diptera: Simuliidae) based on exhaled carbon dioxide. J. Med. Ent., 33, 102–8CrossRefGoogle ScholarPubMed
Schofield, S. and Sutcliffe, J. F. (1997) Humans vary in their ability to elicit biting responses from Simulium venustum (Diptera: Simuliidae). J. Med. Ent., 34, 64–7CrossRefGoogle Scholar
Schofield, S. and Torr, S. J. (2002) A comparison of the feeding behaviour of tsetse and stable flies. Med. Vet. Entomol., 16, 177–85CrossRefGoogle ScholarPubMed
Senghor, J. E. and Samba, E. M. (1988) Onchocerciasis control program – the human perspective. Parasitology Today, 4, 332–3CrossRefGoogle Scholar
Severson, D. W., Brown, S. E. and Knudson, D. L. (2001) Genetic and physical mapping in mosquitoes: molecular approaches. Ann. Rev. Ent., 46, 183–219CrossRefGoogle ScholarPubMed
Severson, D. W., Mori, A., Zhang, Y. and Christensen, B. M. (1994) Chromosomal mapping of two loci affecting filarial worm susceptibility in Aedes aegypti. Insect Mol. Biol., 3, 67–72CrossRefGoogle ScholarPubMed
Severson, D. W., Thathy, V., Mori, A., Zhang, Y. and Christensen, B. M. (1995) Restriction-fragment-length-polymorphism mapping of quantitative trait loci for malaria parasite susceptibility in the mosquito Aedes aegypti. Genetics, 139, 1711–17Google ScholarPubMed
Shahabuddin, M. (1998) Plasmodium ookinete development in the mosquito midgut: a case of reciprocal manipulation. Parasitology, 116, S83–S93CrossRefGoogle ScholarPubMed
Shahabuddin, M., Fields, I., Bulet, P., Hoffmann, J. A. and Miller, L. H. (1998) Plasmodium gallinaceum: differential killing of some mosquito stages of the parasite by insect defensin. Experimental Parasitology, 89, 103–12CrossRefGoogle ScholarPubMed
Shahan, M. S. and Giltner, L. T. (1945) A review of the epizootiology of equine encephalomyelitis in the United States. J. Am. Vet. Med. Assoc., 107, 279–88Google ScholarPubMed
Shin, S. W., Kokoza, V., Lobkov, I. and Raikhel, A. S. (2003) Relish-mediated immune deficiency in the transgenic mosquito Aedes aegypti. Proc. Natl. Acad. Sci. USA, 100, 2616–21CrossRefGoogle ScholarPubMed
Sieber, K. P., Huber, M., Kaslow, D. et al. (1991) The peritrophic membrane as a barrier. Its penetration by Plasmodium gallinaceum and the effect of a monoclonal antibody to ookinetes. Experimental Parasitology, 72, 145–56CrossRefGoogle ScholarPubMed
Silva, C. P., Ribeiro, A. F., Gulbenkian, S. and Terra, W. R. (1995) Organization, origin and function of the outer microvillar (perimicrovillar) membranes of Dysdercus peruvianus (Hemiptera) midgut cells. J. Insect Physiol., 41, 1093–103CrossRefGoogle Scholar
Silverman, N., Zhou, R., Stoven, S., Pandey, N., Hultmark, D. and Maniatis, T. (2000) A Drosophila IkappaB kinase complex required for Relish cleavage and antibacterial immunity. Genes Dev., 14, 2461–71CrossRefGoogle ScholarPubMed
Simond, P. L. (1898) La propagation de la peste. Annales de la Institut Pasteur, 12, 625
Sippel, W. L. and Brown, A. W. A. (1953) Studies on the responses of the female Aedes mosquito. Part V. The role of visual factors. Bull. Ent. Res., 43, 567–74CrossRefGoogle Scholar
Smit, F. G. A. M. (1972) On some adaptive structures in Siphonaptera. Folia Parasit., 19, 5–17Google ScholarPubMed
Smith, C. N., Smith, N., Gouck, H. K. et al. (1970) L-lactic acid as a factor in the attraction of Aedes aegypti to human hosts. Ann. Ent. Soc. Am., 63, 760–70CrossRefGoogle ScholarPubMed
Smith, H. V. and Titchener, R. N. (1980) Mouthparts of ectoparasites and host damage. Proc. R. Soc. Edin. B-Biol. Sci., 79, 139CrossRefGoogle Scholar
Smith, J. J. B. (1979) Effect of diet viscosity on the operation of the pharyngeal pump in the blood-feeding bug Rhodnius prolixus. J. Exp. Biol., 82, 93–104Google ScholarPubMed
Smith, J. J. B. (1984) Feeding mechanisms. In Kerkut, G. A. and Gilbert, L. I. (eds.), Comprehensive Insect Physiology, Biochemistry and Pharmacology. Oxford: PergamonGoogle Scholar
Smith, J. J. B. and Friend, W. G. (1970) Feeding in Rhodnius prolixus: responses to artificial diets as revealed by changes in electrical resistance. J. Insect Physiol., 16, 1709–20CrossRefGoogle Scholar
Smith, J. J. B. and Friend, W. G. (1982) Feeding behaviour in response to blood fractions and chemical phagostimulants in the blackfly, Simulium venustum. Phys. Ent., 7, 219–26CrossRefGoogle Scholar
Smith, K. G. V. (ed.) (1973) Insects and Other Arthropods of Medical Importance. London: British Museum (Natural History)Google Scholar
Smith, T. and Kilbourne, F. L. (1893) Investigations into the nature, causation and prevention of Texas or Southern cattle fever. U.S. Dept. Agric. Bur. Animal. Indust. Bull., Vol. 1, 301
Snodgrass, R. E. (1944) The anatomy of the Mallophaga. Occ. Pap. Calif. Acad. Sci., 6, 145–229Google Scholar
Soares, M. B., Titus, R. G., Shoemaker, C. B., David, J. R. and Bozza, M. (1998) The vasoactive peptide maxadilan from sand fly saliva inhibits TNF-alpha and induces IL-6 by mouse macrophages through interaction with the pituitary adenylate cyclase-activating polypeptide (PACAP) receptor. J. Immunol., 160, 1811–16Google ScholarPubMed
Soderhall, K. and Cerenius, L. (1998) Role of the prophenoloxidase-activating system in invertebrate immunity. Current Opinion in Immunology, 10, 23–8CrossRefGoogle ScholarPubMed
Soldatos, A. N., Metheniti, A., Mamali, I., Lambropoulou, M. and Marmaras, V. J. (2003) Distinct LPS-induced signals regulate LPS uptake and morphological changes in medfly hemocytes. Insect Biochem. Mol. Biol., 33, 1075–84CrossRefGoogle ScholarPubMed
Sorci, G., Fraipont, M. and Clobert, J. (1997) Host density and ectoparasite avoidance in the common lizard (Lacerta vivipara). Oecologia, 111, 183–8CrossRefGoogle Scholar
Soulsby, E. J. L. (1982) Helminths, Arthropods and Protozoa of Domesticated Animals. London: Bailliere Tindall
Southwood, T. R. E., Khalaf, S. and Sinden, R. E. (1975) The micro-organisms of tsetse flies. Acta Trop., 32, 259–66Google ScholarPubMed
Southworth, G. C., Mason, G. and Seed, J. R. (1968) Studies in frog trypanosomiasis. I. A 24-hour cycle in the parasitaemia level of Trypanosoma rotatorium in Rana clamitans from Louisiana. J. Parasit., 54, 255–8CrossRefGoogle Scholar
Spates, G. E. (1981) Proteolytic and haemolytic activity in the midgut of the stablefly Stomoxys calcitrans (L.): partial purification of the haemolysin. Insect Biochem., 11, 143–7CrossRefGoogle Scholar
Spates, G. E., Stipanovic, R. D., Williams, H. and Holman, G. M. (1982) Mechanisms of haemolysis in a blood-sucking dipteran, Stomoxys calcitrans. Insect Biochem., 12, 707–12CrossRefGoogle Scholar
Spindler, K. (2001) The man in the ice under special consideration of paleo-pathological evidence [in German]. Verhandlungen der Deutschen Gesellschaft für Pathologie, 85, 229–36Google Scholar
Stange, G. (1981) The ocellar component of flight equilibrium control in dragonflies. J. Comp. Physiol., 141, 335–47CrossRefGoogle Scholar
Stanko, M., Miklisova, D., Bellocq, J. G. and Morand, S. (2002) Mammal density and patterns of ectoparasite species richness and abundance. Oecologia, 131, 289–95CrossRefGoogle ScholarPubMed
Stark, K. R. and James, A. A. (1995) A factor Xa-directed anticoagulant from the salivary glands of the yellow fever mosquito Aedes aegypti. Experimental Parasitology, 81, 321–31CrossRefGoogle ScholarPubMed
Steelman, C. D. (1976) Effects of external and internal arthropod parasites on domestic livestock production. Ann. Rev. Ent, 21, 155–78CrossRefGoogle ScholarPubMed
Stevens, J. R., Noyes, H. A., Schofield, C. J. and Gibson, W. (2001) The molecular evolution of Trypanosomatidae. Advances in Parasitology, Vol. 48, 1–56CrossRefGoogle ScholarPubMed
Stierhof, Y. D., Bates, P. A., Jacobson, R. L. (1999) Filamentous proteophosphoglycan secreted by Leishmania promastigotes forms gel-like three-dimensional networks that obstruct the digestive tract of infected sandfly vectors. European Journal of Cell Biology, 78, 675–89CrossRefGoogle ScholarPubMed
Stojanovich, C. J. (1945) The head and mouthparts of the sucking lice (Insecta: Anoplura). Microentomology, 10, 1–46Google Scholar
Stoven, S., Silverman, N., Junell, A. et al. (2003) Caspase-mediated processing of the Drosophila NF-kappaB factor Relish. Proc. Natl. Acad. Sci. USA, 100, 5991–6CrossRefGoogle ScholarPubMed
Strand, M. R. and Clark, K. D. (1999) Plasmatocyte spreading peptide induces spreading of plasmatocytes but represses spreading of granulocytes. Arch. Insect Biochem. Physiol., 42, 213–233.0.CO;2-4>CrossRefGoogle ScholarPubMed
Strand, M. R. and Pech, L. L. (1995) Immunological basis for compatibility in parasitoid host relationships. Ann. Rev. Ent., 40, 31–56CrossRefGoogle ScholarPubMed
Stys, P. and Daniel, M. (1957) Lyctocoris compestris F. (Heteroptera: Anthocoridae) as a human facultative ectoparasite. Acta Societatis Entomologicae Cechoslovenicae, 54, 1–10Google Scholar
Sutcliffe, J. F. (1986) Black fly host location: a review, Can J. Zool, 64(4), 1041–53
Sutcliffe, J. F. (1987) Distance orientation of biting flies to their hosts. Insect Sci. Applic., 8, 611–16Google Scholar
Sutcliffe, J. F. and McIver, S. B. (1975) Artificial feeding of simuliids (Simulium venustum), factors associated with probing and gorging. Experientia, 31, 694–5CrossRefGoogle Scholar
Sutcliffe, J. F., Steer, D. J. and Beardsall, D. (1995) Studies of host location behavior in the black fly Simulium arcticum (Iis-10.11) (Diptera, Simuliidae) – aspects of close range trap orientation. Bull. Ent. Res., 85, 415–24CrossRefGoogle Scholar
Sutherland, D. R., Christensen, B. M. and Lasee, B. A. (1986) Midgut barrier as a possible factor in filarial worm vector competency in Aedes trivittatus. J. Invert. Path., 47, 1–7CrossRefGoogle ScholarPubMed
Sutherst, R. W., Ingram, J. S. I. and Scherm, H. (1998) Global change and vector-borne diseases. Parasitology Today, 14, 297–9CrossRefGoogle ScholarPubMed
Sutton, O. G. (1947) The problem of diffusion in the lower atmosphere. Quart. J. Roy. Meteorol. Soc., 73, 257–81CrossRefGoogle Scholar
Swellengrebel, N. H. (1929) La dissociation des fonctions sexuelles de nutritives (dissociation gonotrophique) d'Anopheles maculipennis comme cause du paludisme dans les Pays-Bas et ses rapports avec ‘l'infection domiciliare’. Ann. Inst. Pasteur, 43, 1370–89Google Scholar
Takehana, A., Katsuyama, T., Yano, T. et al. (2002) Overexpression of a pattern-recognition receptor, peptidoglycan-recognition protein-LE, activates imd/relish-mediated antibacterial defense and the prophenoloxidase cascade in Drosophila larvae. Proc. Natl. Acad. Sci. USA, 99, 13705–10CrossRefGoogle ScholarPubMed
Takken, W. (1996) Synthesis and future challenges: the response of mosquitoes to host odours. In Cardew, G. (ed.), Olfaction in Mosquito–Host Interactions. Chichester:Wiley 302–20Google Scholar
Takken, W., Kager, P. A., and Kaay, H. J., (1999) Endemische malaria terug in Nederland?Nederlands Tijdschrift voor Geneeskunde, 143, 836–8Google Scholar
Takken, W., Klowden, M. J. and Chambers, G. M. (1998) Effect of body size on host seeking and blood meal utilization in Anopheles gambiae sensu stricto (Diptera: Culicidae): the disadvantage of being small. J. Med. Ent., 35, 639–45CrossRefGoogle ScholarPubMed
Takken, W. and Knols, B. G. J. (1999) Odor-mediated behaviour of afrotropical malaria mosquitoes. Ann. Rev. Ent., 44, 131–57CrossRefGoogle ScholarPubMed
Takken, W., Loon, J. J. A. and Adam, W. (2001) Inhibition of host-seeking response and olfactory responsiveness in Anopheles gambiae following blood feeding. J. Insect Physiol., 47, 303–10CrossRefGoogle ScholarPubMed
Tashiro, H. and Schwardt, H. H. (1953) Biological studies of horse flies in New York. J. Econ. Ent., 46, 813–22CrossRefGoogle Scholar
Tawfik, M. S. (1968) Feeding mechanisms and the forces involved in some blood-sucking insects. Quaes. Ent., 4, 92–111Google Scholar
Taylor, P. J. and Hurd, H. (2001) The influence of host haematocrit on the blood feeding success of Anopheles stephensi: implications for enhanced malaria transmission. Parasitology, 122, 491–6CrossRefGoogle ScholarPubMed
Teesdale, C. (1955) Studies on the bionomics of Aedes aegypti L. in its natural habitats in a coastal region of Kenya. Bull. Ent. Res., 46, 711–42CrossRefGoogle Scholar
Tempelis, C. H. and Washino, R. K. (1967) Host feeding patterns of Culex tarsalis in the Sacramento Valley, California, with notes on other species. J. Med. Ent., 4, 315–18CrossRefGoogle ScholarPubMed
Terra, W. R. (1988a) Physiology and biochemistry of insect digestion: an evolutionary perspective. Braz. J. Med. Biol. Res., 21, 675–734Google Scholar
Terra, W. R. (2001) The origin and functions of the insect peritrophic membrane and peritrophic gel. Arch. Insect Biochem. Physiol., 47, 47–61CrossRefGoogle ScholarPubMed
Terra, W. R. and Ferreira, C. (1994) Insect digestive enzymes – properties, compartmentalization and function. Comp. Biochem. Physiol. B, 109, 1–62CrossRefGoogle Scholar
Terra, W. R., Ferreira, C . and Garcia, E. S. (1988b) Origin, distribution, properties and functions of the major Rhodnius prolixus midgut hydrolases. Insect Biochem., 18, 423–34CrossRefGoogle Scholar
Thathy, V., Severson, D. W. and Christensen, B. M. (1994) Reinterpretation of the genetics of susceptibility of Aedes aegypti to Plasmodium gallinaceum. J. Parasitol., 80, 705–12CrossRefGoogle ScholarPubMed
Theodor, O. (1967) An Illustrated Catalogue of the Rothschild Collection of Nycteribiidae (Diptera) in the British Museum (Natural History). London: British Museum
Theodos, C. M., Ribeiro, J. M. and Titus, R. G. (1991) Analysis of enhancing effect of sand fly saliva on Leishmania infection in mice. Infection and Immunity, 59, 1592–8Google ScholarPubMed
Theodos, C. M. and Titus, R. G. (1993) Salivary gland material from the sand fly Lutzomyia longipalpis has an inhibitory effect on macrophage function in vitro. Parasite Immunology, 15, 481–7CrossRefGoogle ScholarPubMed
Thompson, B. H. (1976) Studies on the attraction of Simulium damnosum s.l. (Diptera: Simuliidae) to its hosts. I. The relative importance of sight, exhaled breath and smell. Tropenmed. Parasitol., 27, 455–73Google ScholarPubMed
Thompson, W. H. and Beattey, B. J. (1977) Veneral transmission of La Crosse (California encephalitis) arbovirus in Aedes triseriatus mosquitoes. Science, 196, 530–1CrossRefGoogle ScholarPubMed
Thorsteinson, A. J. and Bracken, G. K. (1965) The orientation behavior of horse flies and deer flies (Tabanidae: Diptera). III. The use of traps in the study of orientation of tabanids in the field. Ent. Exp. Appl., 8, 189–92CrossRefGoogle Scholar
Thorsteinson, A. J., Bracken, G. K. and Tostawaryk, W. (1966) The orientation behaviour of horse flies and deer flies (Tabanidae: Diptera). VI. The influence of the number of reflecting surfaces on attractiveness to tabanids of glossy black polyhedra. Can. J. Zool., 44, 275–9CrossRefGoogle Scholar
Tillyard, R. J. (1935) The evolution of scorpion flies and their derivatives (order Mecoptera). Ann. Ent. Soc. Am., 28, 37–45CrossRefGoogle Scholar
Titus, R. G. (1998) Salivary gland lysate from the sand fly Lutzomyia longipalpis suppresses the immune response of mice to sheep red blood cells in vivo and concanavalin A in vitro. Exp. Parasitol., 89, 133–6CrossRefGoogle ScholarPubMed
Titus, R. G. and Ribeiro, J. M. C. (1990) The role of vector saliva in transmission of arthropod-borne disease. Parasitology Today, 6, 157–60CrossRefGoogle ScholarPubMed
Tobe, S. S. and Davey, K. G. (1972) Volume relationships during the pregnancy cycle of the tsetse fly Glossina austeni. Can. J. Zool., 50, 999–1010CrossRefGoogle ScholarPubMed
Torr, S. J. (1989) The host-orientated behaviour of tsetse flies (Glossina): the interaction of visual and olfactory stimuli. Phys. Ent., 14, 325–40CrossRefGoogle Scholar
Torr, S. J., Hall, D. R. and Smith, J. L. (1995) Responses of tsetse-flies (Diptera, Glossinidae) to natural and synthetic ox odors. Bull. Ent. Res., 85, 157–66CrossRefGoogle Scholar
Torr, S. J. and Mangwiro, T. N. C. (2000) Interactions between cattle and biting flies: effects on the feeding rate of tsetse. Med. Vet. Ent., 14, 400–9CrossRefGoogle ScholarPubMed
Torr, S. J., Wilson, P. J., Schofield, S. et al. (2001) Application of DNA markers to identify the individual-specific hosts of tsetse feeding on cattle. Med. Vet. Ent., 15, 78–86CrossRefGoogle ScholarPubMed
Traub, R. (1985) Coevolution of fleas and mammals. In Kim, K. C. (ed.), Coevolution of Parasitic Arthropods and Mammals. New York: WileyGoogle Scholar
Trpis, M., Duhrkopf, R. E. and Parker, K. L. (1981) Non-Mendelian inheritance of mosquito susceptibility to infection with Brugia malayi and Brugia pahangi. Science, 211, 1435–7CrossRefGoogle ScholarPubMed
Trudeau, W. L., Fernandez-Caldas, E., Fox, R. W. (1993) Allergenicity of the cat flea (Ctenocephalides felis felis). Clinical and Experimental Allergy: Journal of the British Society for Allergy and Clinical Immunology, 23, 377–83CrossRefGoogle Scholar
Turell, M. J., Bailey, C. L. and Rossi, C. A. (1984a) Increased mosquito feeding on Rift Valley fever virus-infected lambs. Am. J. Trop. Med. Hyg., 33, 1232–8CrossRefGoogle Scholar
Turell, M. J., Rossignol, P. A., Spielman, A., Rossi, C. A. and Bailey, C. L. (1984b) Enhanced arboviral transmission by mosquitoes that concurrently ingested microfilaria. Science, 225, 1039–41CrossRefGoogle Scholar
Turner, D. A. and Invest, J. F. (1973) Laboratory analyses of vision in tsetse flies (Dipt., Glossinidae). Bull. Ent. Res., 62, 343–57CrossRefGoogle Scholar
Tzou, P., Gregorio, E. and Lemaitre, B. (2002) How Drosophila combats microbial infection: a model to study innate immunity and host-pathogen interactions. Curr. Opin. Microbiol., 5, 102–10CrossRefGoogle ScholarPubMed
Tzou, P., Ohresser, S., Ferrandon, D. et al. (2000) Tissue-specific inducible expression of antimicrobial peptide genes in Drosophila surface epithelia. Immunity, 13, 737–48CrossRefGoogle ScholarPubMed
Underhill, G. W. (1940) Some factors influencing feeding activity of simuliids in the field. J. Econ. Entomol., 33, 915–17CrossRefGoogle Scholar
Vale, G. A. (1974a) New field methods for studying the response of tsetse flies (Diptera, Glossinidae) to hosts. Bull. Ent. Res., 64, 199–208CrossRefGoogle Scholar
Vale, G. A. (1974b) The response of tsetse flies (Diptera, Glossinidae) to mobile and stationary baits. Bull. Ent. Res., 64, 545–88CrossRefGoogle Scholar
Vale, G. A. (1977) Feeding responses of tsetse flies (Diptera: Glossinidae) to stationary hosts. Bull. Ent. Res., 67, 635–49CrossRefGoogle Scholar
Vale, G. A. (1980) Flight as a factor in host-finding behaviour of tsetse flies (Diptera: Glossinidae). Bull. Ent. Res., 70, 299–307CrossRefGoogle Scholar
Vale, G. A. (1982) The trap-orientated behaviour of tsetse flies (Glossinidae) and other Diptera. Bull. Ent. Res., 72, 71–93CrossRefGoogle Scholar
Vale, G. A. (1983) The effects of odours, wind direction and wind speeds on the distribution of Glossina (Diptera: Glossinidae) and other insects near stationary targets. Bull. Ent. Res., 73, 53–64CrossRefGoogle Scholar
Vale, G. A. and Hall, D. R. (1985a) The role of 1-octen-3-ol, acetone and carbon dioxide in the attraction of tsetse flies, Glossina spp. (Diptera: Glossinidae), to ox odour. Bull. Ent. Res., 75, 209–17CrossRefGoogle Scholar
Vale, G. A. and Hall, D. R. (1985b) The use of 1-octen-3-ol, acetone and carbon dioxide to improve baits for tsetse flies, Glossina spp. (Diptera: Glossinidae). Bull. Ent. Res., 75, 219–31CrossRefGoogle Scholar
Vale, G. A., Hall, D. R. and Gough, A. J. E. (1988) The olfactory responses of tsetse flies, Glossina spp. (Diptera: Glossinidae), to phenols and urine in the field. Bull. Ent. Res., 78, 293–300CrossRefGoogle Scholar
Valenzuela, J. G., Belkaid, Y., Rowton, E. and Ribeiro, J. M. C. (2001) The salivary apyrase of the blood-sucking sand fly Phlebotomus papatasi belongs to the novel Cimex family of apyrases. J. Exp. Biol., 204, 229–37Google ScholarPubMed
Valenzuela, J. G., Francischetti, I. M. B. and Ribeiro, J. M. C. (1999) Purification, cloning, and synthesis of a novel salivary anti-thrombin from the mosquito Anopheles albimanus. Biochemistry, 38, 11209–15CrossRefGoogle ScholarPubMed
Valenzuela, J. G., Pham, V. M., Garfield, M. K., Francischetti, I. M. B. and Ribeiro, J. M. C. (2002) Toward a description of the sialome of the adult female mosquito Aedes aegypti. Insect Biochemistry and Molecular Biology, 32, 1101–22CrossRefGoogle Scholar
Valenzuela, J. G. and Ribeiro, J. M. C. (1998) Purification and cloning of the salivary nitrophorin from the hemipteran Cimex lectularius. J. Exp. Biol., 201, 2659–64Google ScholarPubMed
Handel, E. (1984) Metabolism of nutrients in the adult mosquito. Mosq. News, 44, 573–9Google Scholar
Naters, W. M. V., Otter, C. J. and Cuisance, D. (1998) The interaction of taste and heat on the biting response of the tsetse fly Glossina fuscipes fuscipes. Phys. Ent., 23, 285–8Google Scholar
Vargaftig, B. B., Chignard, M. and Benveniste, J. (1981) Present concepts on the mechanism of platelet aggregation. Biochem. Pharmacol., 30, 263–71
Vaughan, J. A. and Turell, M. J. (1996) Facilitation of Rift Valley fever virus transmission by Plasmodium berghei sporozoites in Anopheles stephensi mosquitoes. Am. J. Trop. Med. Hyg., 55, 407–9CrossRefGoogle ScholarPubMed
Venkatesh, K. and Morrison, P. E. (1982) Blood meal as a regulator of triacylglycerol synthesis in the haematophagous stable fly, Stomoxys calcitrans. J. Comp. Physiol., 147, 49–52CrossRefGoogle Scholar
Venkatesh, K., Morrison, P. E. and Kallapur, V. L. (1981) Influence of blood meals on the conversion of D-(U-14C)-glucose to lipid in the fat body of the haematophagous stablefly, Stomoxys calcitrans. Comp. Biochem. Physiol., 68, 425–9Google Scholar
Vernick, K. D., Fujioka, H., Seeley, D. C. et al. (1995) Plasmodium gallinaceum – a refractory mechanism of ookinete killing in the mosquito, Anopheles gambiae. Experimental Parasitology, 80, 583–95CrossRefGoogle ScholarPubMed
Victoir, K. and Dujardin, J. C. (2002) How to succeed in parasitic life without sex? Asking Leishmania. Trends Parasitol., 18, 81–5CrossRefGoogle ScholarPubMed
Voskamp, K. E., Otter, C. J. and Noorman, N. (1998) Electroantennogram responses of tsetse flies (Glossina pallidipes) to host odours in an open field and riverine woodland. Phys. Ent., 23, 176–83CrossRefGoogle Scholar
Waage, J. K. (1979) The evolution of insect/vertebrate associations. Biological Journal of the Linnaeon Society, 12, 187–224CrossRefGoogle Scholar
Waage, J. K. (1981) How the zebra got its stripes – biting flies as selective agents in the evolution of zebra coloration. J. Ent. Soc. Sth. Afr., 44, 351–8Google Scholar
Waage, J. K. and Davies, C. R. (1986) Host-mediated competition in a bloodsucking insect community. Journal of Animal Ecology, 55, 171–80CrossRefGoogle Scholar
Waage, J. K. and Nondo, J. (1982) Host behaviour and mosquito feeding success: an experimental study. Trans. R. Soc. Trop. Med. Hyg., 76, 119–22CrossRefGoogle Scholar
Wahid, I., Sunahara, T. and Mogi, M. (2003) Maxillae and mandibles of male mosquitoes and female autogenous mosquitoes (Diptera: Culicidae). J. Med. Ent., 40, 150–8CrossRefGoogle Scholar
Ward, R. A. (1963) Genetic aspects of the susceptibility of mosquitoes to malaria infections. Exp. Parasit., 13, 328–41CrossRefGoogle Scholar
Warnes, M. L. (1995) Field studies on the effect of cattle skin secretion on the behavior of tsetse. Med. Vet. Ent., 9, 284–8CrossRefGoogle ScholarPubMed
Warnes, M. L. and Finlayson, L. H. (1985) Responses of the stable fly, Stomoxys calcitrans (L.) (Diptera: Muscidae), to carbon dioxide and host odours. I. Activation. Bull. Ent. Res., 75, 519–27CrossRefGoogle Scholar
Warnes, M. L. and Finlayson, L. H. (1986) Electroantennogram responses of the stable fly, Stomoxys calcitrans, to carbon dioxide and other odours. Phys. Ent., 11, 469–73CrossRefGoogle Scholar
Warnes, M. L. and Finlayson, L. H. (1987) Effect of host behaviour on host preference in Stomoxys calcitrans. Med. Vet. Ent., 1, 53–7CrossRefGoogle ScholarPubMed
Waterhouse, D. F. (1953) The occurrence and significance of the peritrophic membrane, with special reference to adult Lepidoptera and Diptera. Aust. J. Zool., 1, 299–318CrossRefGoogle Scholar
Watts, D. M., Pantuwatana, S., Defoliart, G. S., Yuill, T. M. and Thompson, W. H. (1973) Transovarial transmission of La Crosse virus (California encephalitis group) in the mosquito, Aedes triseriatus. Science, 182, 1140–1CrossRefGoogle Scholar
Webb, P. A., Happ, C. M., Maupin, G. O. et al. (1989) Potential for insect transmission of HIV: experimental exposure of Cimex hemipterus and Toxorhynchites amboinensis to human immunodeficiency virus. J. Infect Dis., 160, 970–7CrossRefGoogle ScholarPubMed
Webber, L. A. and Edman, J. D. (1972) Anti-mosquito behaviour of ciconiiform birds. Animal Behaviour, 20, 228–32CrossRefGoogle Scholar
Webster, J. P. and Woolhouse, M. E. J. (1999) Cost of resistance: relationship between reduced fertility and increased resistance in a snail-schistosome host-parasite system. Proc. R. Soc. Lond. B Sci., 266, 391–6CrossRefGoogle Scholar
Wee, W. L. and Anderson, J. R. (1995) Tethered flight capabilities and survival of Lambornella clarki-infected, blood-fed, and gravid Aedes sierrensis (Diptera, Culicidae). J. Med. Ent., 32, 153–60Google Scholar
Weitz, B. (1963) The feeding habits of Glossina. Bull. WHO, 28, 711–29Google ScholarPubMed
Wekesa, J. W., Copeland, R. S. and Mwangi, R. W. (1992) Effect of Plasmodium falciparum on blood feeding behavior of naturally infected Anopheles mosquitoes in western Kenya. Am. J. Trop. Med. Hyg., 47, 484–8CrossRefGoogle ScholarPubMed
Welburn, S. C., et al. (1987) In vitro cultivation of rickettsia-like organisms from Glossina spp. Ann. Trop. Med. Parasit., 81(4), 331–5CrossRef
Welburn, S. C., Arnold, K., Maudlin, I. and Gooday, G. W. (1993) Rickettsia-like organisms and chitinase production in relation to transmission of trypanosomes by tsetse-flies. Parasitology, 107, 141–5CrossRefGoogle ScholarPubMed
Welburn, S. C. and Murphy, N. B. (1998) Prohibitin and RACK homologues are up-regulated in trypanosomes induced to undergo apoptosis and in naturally occurring terminally differentiated forms. Cell Death and Differentiation, 5 (7), 615–22CrossRef
Wells, E. A. (1982) Trypanosomiasis in the absence of tsetse. In Baker, J. R. (ed.), Perspectives in Trypansosomiasis Research. Chichester:Research Studies PressGoogle Scholar
Wen, Y., Muir, L. E. and Kay, B. H. (1997) Response of Culex quinquefasciatus to visual stimuli. J. Am. Mosq. Control Assoc., 13, 150–2Google ScholarPubMed
Wenk, P. (1962) Anatomie des Kopfes von Wilhelmia equina (Simuliidae syn. Melusinidae, Diptera). Zool. Jahrb. Abt. Ontog. Tiere, 80, 81–134Google Scholar
Wenk, P. and Schlorer, G. (1963) Wirtsorientierung und Kopulation bei blutsaugenden Simuliiden (Diptera). Z. Tropenmed. Parasitol., 14, 177–91Google Scholar
Werner, T., Liu, G., Kang, D. et al. (2000) A family of peptidoglycan recognition proteins in the fruit fly Drosophila melanogaster. Proc. Natl. Acad. Sci. USA, 97, 13772–7CrossRefGoogle ScholarPubMed
Werner-Reiss, U., Galun, R., Crnjar, R. and Liscia, A. (1999) Factors modulating the blood feeding behavior and the electrophysiological responses of labral apical chemoreceptors to adenine nucleotides in the mosquito Aedes aegypti (Culicidae). J. Insect Physiol., 45, 801–8CrossRefGoogle Scholar
Weyer, F. (1960) Biological relationships between lice (Anoplura) and microbial agents. Ann. Rev. Ent., 5, 405–20CrossRefGoogle Scholar
Wharton, R. H. (1957) Studies on filariasis in Malaya: observations on the development of Wuchereria malayi in Mansonia (Mansonioides) longipalpis. Ann. Trop. Med. Parasit., 51, 278–96CrossRefGoogle ScholarPubMed
White, G. B. (1974) Anopheles gambiae complex and disease transmission in Africa. Trans. R. Soc. Trop. Med. Hyg., 4, 278–98CrossRefGoogle Scholar
White, G. B., Magayuka, S. A. and Boreham, P. F. L. (1972) Comparative studies on sibling species of the Anopheles gambiae Giles complex (Dipt. Culicidae): bionomics and vectorial capacity of species A and species B at Segera, Tanzania. Bull. Ent. Res., 62, 295–317CrossRefGoogle Scholar
White, G. B. and Rosen, B. (1973) Comparative studies on sibling species of the Anopheles gambiae Giles complex (Dipt. Culicidae). II. Ecology of species A and B in savanna around Kaduna, Nigeria, during transition from wet to dry season. Bull. Ent. Res., 62, 613–25CrossRefGoogle Scholar
Whiting, M. F. (2002) Mecoptera is paraphyletic: multiple genes and phylogeny of Mecoptera and Siphonaptera. Zoologica Scripta, 31, 93–104CrossRefGoogle Scholar
Whitten, M. M. and Ratcliffe, N. A. (1999) In vitro superoxide activity in the haemolymph of the West Indian leaf cockroach, Blaberus discoidalis. J. Insect Physiol., 45, 667–75CrossRefGoogle ScholarPubMed
Wigglesworth, V. B. (1941) The sensory physiology of the human louse Pediculus humanus corporis de Greer (Anoplura). Parasitology, 33, 67–109CrossRefGoogle Scholar
Wigglesworth, V. B. (1979) Secretory activities of plasmatocytes and oenocytoids during the moulting cycle in an insect (Rhodnius). Tissue and Cell, 11, 69–78CrossRefGoogle Scholar
Wigglesworth, V. B. and Gillett, J. D. (1934) The function of antennae in Rhodnius prolixus and the mechanism of orientation of the host. J. Exp Biol., 11, 120–39Google Scholar
Williams, B. (1994) Models of trap seeking by tsetse-flies – anemotaxis, klinokinesis and edge-detection. Journal of Theoretical Biology, 168, 105–15CrossRefGoogle Scholar
Williams, P. D. and Day, T. (2001) Interactions between sources of mortality and the evolution of parasite virulence. Proc. R. Soc. Lond. B Sci., 268, 2331–7CrossRefGoogle ScholarPubMed
Wilson, J. J., Neame, P. V. and Kelton, J. G. (1982) Infection induced thrombocytopaenia. Seminars in Thrombosis and Haemostasis, 8, 217–33CrossRefGoogle Scholar
Wilson, M. (1978) The functional organisation of locust ocelli. J. Comp. Physiol., 124, 297–316CrossRefGoogle Scholar
Wilson, R., Chen, C. W. and Ratcliffe, N. A. (1999) Innate immunity in insects: the role of multiple, endogenous serum lectins in the recognition of foreign invaders in the cockroach, Blaberus discoidalis. Journal of Immunology, 162, 1590–6Google ScholarPubMed
Woke, P. A. (1937) Comparative effects of the blood of man and of canary on egg production of Culex pipiens Linn. J. Parasit., 23, 311–13CrossRefGoogle Scholar
Wood, D. M. (1964) Studies on the beetles Leptinillus validus (Horn) and Platypsyllus castoris Rissema (Coleoptera: Leptinidae) from beaver. Proceedings of the Entomological Society of Ontario, 95, 33–63Google Scholar
Wood, S. F. (1942) Observations on vectors of Chagas' disease in the United States. I. California. Bull. Calif. Acad. Sci., 41, 61–9Google Scholar
Worms, M. J. (1972) Circadian and seasonal rhythms in blood parasites. In Canning, E. U. and Wright, C. A. (eds.), Behavioural Aspects of Parasite Transmission. London: Linnean SocietyGoogle Scholar
Wright, R. H. (1958) The olfactory guidance of flying insects. Can. Entomol., 90, 81–9CrossRefGoogle Scholar
Wright, R. H. (1968) Tunes to which mosquitoes dance. New Sci., 37, 694–7Google Scholar
Wright, R. H. and Kellogg, F. E. (1962) Response of Aedes aegypti to moist convection currents. Nature, 194, 402–3CrossRefGoogle ScholarPubMed
Xie, H., Bain, O. and Williams, S. A. (1994) Molecular phylogenetic studies on filarial parasites based on 5s ribosomal spacer sequences. Parasite-Journal de la Societe Française de Parasitologie, 1, 141–51Google ScholarPubMed
Xu, P. X., Zwiebel, L. J. and Smith, D. P. (2003) Identification of a distinct family of genes encoding atypical odorant-binding proteins in the malaria vector mosquito, Anopheles gambiae. Insect Mol. Biol., 12, 549–60CrossRefGoogle ScholarPubMed
Yajima, M., Takada, M., Takahashi, N. et al. (2003) A newly established in vitro culture using transgenic Drosophila reveals functional coupling between the phospholipase A2-generated fatty acid cascade and lipopolysaccharide-dependent activation of the immune deficiency (imd) pathway in insect immunity. Biochem. J., 371, 205–10CrossRefGoogle ScholarPubMed
Yu, X. Q. and Kanost, M. R. (2000) Immulectin-2, a lipopolysaccharide specific lectin from an insect, Manduca sexta, is induced in response to gram-negative bacteria. J. Biol. Chem., 275, 37373–81CrossRefGoogle ScholarPubMed
Yuill, T. M. (1983) The role of mammals in the maintainence and dissemination of La Crosse virus. In Calisher, C. H. and Thompson, W. H. (eds.), California Serogroup Viruses. New York: Alan R. LissGoogle Scholar
Zahedi, M. (1994) The fate of Brugia pahangi microfilariae in Armigeres subalbatus during the first 48 hours post ingestion. Tropical Medicine and Parasitology, 45, 33–5Google ScholarPubMed
Zhang, D., Cupp, M. S. and Cupp, E. W. (2002) Thrombostasin: purification, molecular cloning and expression of a novel anti-thrombin protein from horn fly saliva. Insect Biochemistry and Molecular Biology, 32, 321–30CrossRefGoogle ScholarPubMed
Zhang, Y., Ribeiro, J. M. C., Guimaraes, J. A. and Walsh, P. N. (1998) Nitrophorin-2: a novel mixed-type reversible specific inhibitor of the intrinsic factor-X activating complex. Biochemistry, 37, 10681–90CrossRefGoogle ScholarPubMed
Zheng, L. (1999) Genetic basis of encapsulation response in Anopheles gambiae. Parasitologia, 41, 181–4Google ScholarPubMed
Zheng, L., Cornel, A. J., Wang, R. (1997) Quantitative trait loci for refractoriness of Anopheles gambiae to Plasmodium cynomolgi B. Science, 276, 425–8CrossRefGoogle ScholarPubMed
Zheng, L., Wang, S., Romans, P., et al. (2003) Quantitative trait loci in Anopheles gambiae controlling the encapsulation response against Plasmodium cynomolgi Ceylon. BMC Genet, 4, 16CrossRefGoogle ScholarPubMed
Zieler, H., Garon, C. F., Fischer, E. R. and Shahabuddin, M. (2000) A tubular network associated with the brush-border surface of the Aedes aegypti midgut: implications for pathogen transmission by mosquitoes. J. Exp. Biol., 203, 1599–611Google ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • M. J. Lehane, Liverpool School of Tropical Medicine
  • Book: The Biology of Blood-Sucking in Insects
  • Online publication: 08 January 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511610493.011
Available formats
×

Save book to Dropbox

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

  • References
  • M. J. Lehane, Liverpool School of Tropical Medicine
  • Book: The Biology of Blood-Sucking in Insects
  • Online publication: 08 January 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511610493.011
Available formats
×

Save book to Google Drive

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

  • References
  • M. J. Lehane, Liverpool School of Tropical Medicine
  • Book: The Biology of Blood-Sucking in Insects
  • Online publication: 08 January 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511610493.011
Available formats
×