Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-7tdvq Total loading time: 0 Render date: 2024-08-17T05:01:30.516Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  14 August 2009

Boris R. Krasnov
Affiliation:
Ben-Gurion University of the Negev, Israel
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
Functional and Evolutionary Ecology of Fleas
A Model for Ecological Parasitology
, pp. 466 - 582
Publisher: Cambridge University Press
Print publication year: 2008

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

Abouheif, E. & Fairbairn, D. J. (1997). A comparative analysis of allometry for sexual size dimorphism: assessing Rensch's rule. American Naturalist, 149, 540–562.CrossRefGoogle Scholar
Abrahams, M. (2004). Mark but this flea. Guardian, 30 November 2004.Google Scholar
Abramsky, Z., Bowers, M. A. & Rosenzweig, M. L. (1986). Detecting interspecific competition in the field: testing the regression method. Oikos, 47, 199–204.CrossRefGoogle Scholar
Abramsky, Z., Rosenzweig, M. L., Pinshow, B., et al. (1990). Habitat selection: an experimental field-test with two gerbil species. Ecology, 71, 2358–2369.CrossRefGoogle Scholar
Abu-Madi, M. A., Behnke, J. M., Mikhail, M., Lewis, J. W. & Al-Kaabi, M. L. (2005). Parasite populations in the brown rat Rattus norvegicus from Doha, Qatar between years: the effect of host age, sex and density. Journal of Helminthology, 79, 105–111.CrossRefGoogle ScholarPubMed
Acosta, R. (2005). Relación huésped-parásito en pulgas (Insecta: Siphonaptera) y roedores (Mammalia: Rodentia) del estado de Querétaro, México. Folia Entomologica Mexicana, 44, 37–47.Google Scholar
Acosta, R. & Morrone, J. J. (2005). A new species of Hystrichopsylla Taschenberg (Siphonaptera: Hystrichopsyllidae) from the Mexican transition zone. Zootaxa, 1027, 213–238.CrossRefGoogle Scholar
Adjemian, J. C. Z., Girvetz, E., Beckett, L. & Foley, J. E. (2006). Analysis of genetic algorithm for rule-set production (GARP) modeling approach for predicting distributions of fleas implicated as vectors of plague, Yersinia pestis, in California. Journal of Medical Entomology, 43, 93–103.Google ScholarPubMed
Ageyev, V. S. & Sludsky, A. A. (1985). Materials on fleas of small mammals in the eastern Pamir and perspective of the epizootologic monitoring of this region. In Important Questions of Epidemiological Monitoring in the Natural Foci of Plague: Natural Focality of Plague in High Mountains, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 10–12 (in Russian).Google Scholar
Ageyev, V. S., Arzhannikova, A. S., Tlegenov, T. T., Samarin, E. G. & Serzhanov, O. S. (1983). Interspecific interactions among five species of fleas co-occurring on the midday jirds. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 209–210 (in Russian).Google Scholar
Ageyev, V. S., Serzhanov, O. S., Arzhannikova, A. S. & Tlegenov, T. T. (1984). Interspecific interactions among imagoes of five species of fleas (Siphonaptera) parasitic on gerbils. Parazitologiya, 18, 185–190 (in Russian).Google Scholar
Akin, D. M. (1984). Relationship between feeding and reproduction in the cat flea, Ctenocephalides felis (Bouché). Unpublished M.Sc. thesis, University of Florida, Gainesville, FL.
Alania, I. I., Rostigaev, B. A., Shiranovich, P. I. & Dzneladze, M. T. (1964). Data on the flea fauna of Adzharia. Proceedings of the Armenian Anti-Plague Station, 3, 407–435 (in Russian).Google Scholar
Alekseev, A. N. (1961). On the bionomics of fleas Ceratophyllus (Nosopsyllus) consimilis Wagn., 1898 (Ceratophyllidae, Aphaniptera). Zoologicheskyi Zhurnal, 40, 1840–1847 (in Russian).Google Scholar
Alekseev, A. N., Grebenyuk, R. V., Tchirov, P. A. & Kadysheva, A. M. (1971). On the relationships between the listeriosis pathogen (Listeria monocytogenes) and fleas. Parazitologiya, 5, 113–118 (in Russian).Google Scholar
Alexander, J. O. (1986). The physiology of itch. Parasitology Today, 2, 345–351.CrossRefGoogle ScholarPubMed
Allan, R. M. (1956). A study of the populations of the rabbit flea Spilopsyllus cuniculi (Dale) on the wild rabbit, Oryctolagus cuniculus, in north-east Scotland. Proceedings of the Royal Entomological Society of London A, 31, 145–152.CrossRefGoogle Scholar
Allander, K. (1998). The effects of an ectoparasite on reproductive success in the great tit: a 3-year experimental study. Canadian Journal of Zoology, 76, 19–25.CrossRefGoogle Scholar
Allred, D. M. (1968). Fleas of the National Reactor Testing Station. Great Basin Naturalist, 28, 73–87.Google Scholar
Allsopp, P. G. (1997). Probability of describing an Australian scarab beetle: influence of body size and distribution. Journal of Biogeography, 24, 717–724.CrossRefGoogle Scholar
Almeida-Neto, M., Guimarães, P. R. & Lewinsohn, T. M. (2007). On nestedness analyses: rethinking matrix temperature and anti-nestedness. Oikos, 116, 716–722.CrossRefGoogle Scholar
Altizer, S., Nunn, C. L., Thrall, P. H., et al. (2004). Social organization and parasite risk in mammals: integrating theory and empirical studies. Annual Review of Ecology and Systematics, 34, 517–547.CrossRefGoogle Scholar
Amin, O. M. (1974). Comb variation in the rabbit flea Cediopsylla simplex (Baker). Journal of Medical Entomology, 11, 227–230.CrossRefGoogle Scholar
Amin, O. M. (1976). Host associations and seasonal occurrence of fleas from southeastern Wisconsin mammals, with observations on morphologic variations. Journal of Medical Entomology, 13, 179–192.CrossRefGoogle ScholarPubMed
Amin, O. M. (1982). The significance of pronotal comb patterns in flea–host lodging adaptations. Wiadomosci Parazytologiczne, 28, 93–94.Google Scholar
Amin, O. M. & Sewell, R. G. (1977). Comb variations in the squirrel and chipmunk fleas, Orchopeas h. howardii (Baker) and Megabothris acerbus (Jordan) (Siphonaptera), with notes on the significance of pronotal comb patterns. American Midland Naturalist, 98, 207–212.CrossRefGoogle Scholar
Amin, O. M. & Wagner, M. E. (1983). Further notes on the function of pronotal combs in fleas (Siphonaptera). Annals of the Entomological Society of America, 76, 232–234.CrossRefGoogle Scholar
Amin, O. M., Wells, T. R. & Gately, H. L. (1974). Comb variation in the cat flea, Ctenocephalides f. felis (Bouché). Annals of the Entomological Society of America, 67, 831–834.CrossRefGoogle Scholar
Amin, O. M., Liu, J., Li, S.-J., Zhang, Y.-M. & Sun, L.-Z. (1993). Development and longevity of Nosopsyllus laeviceps kuzenkovi (Siphonaptera) from Inner Mongolia under laboratory conditions. Journal of Parasitology, 79, 193–197.CrossRefGoogle ScholarPubMed
Amrine, J. W. & Lewis, R. E. (1978). The topography of the exoskeleton of Cediopsylla simplex (Baker 1895) (Siphonaptera: Pulicidae). I. The head and its appendages. Journal of Parasitology, 64, 343–358.CrossRefGoogle ScholarPubMed
Anderson, P. C. & Kok, O. B. (2003). Ectoparasites of springhares in the Northern Cape Province, South Africa. South African Journal of Wildlife Research, 33, 23–32.Google Scholar
Anderson, R. M. & Gordon, D. M. (1982). Processes influencing the distribution of parasite numbers within host populations with special emphasis on parasite-induced host mortality. Parasitology, 85, 373–398.CrossRefGoogle Scholar
Anderson, R. M. & May, R. M. (1978). Regulation and stability of host–parasite population interactions. I. Regulatory processes. Journal of Animal Ecology, 47, 219–247.CrossRefGoogle Scholar
Anderson, R. M., Gordon, D. M., Crawley, M. J. & Hassell, M. P. (1982). Variability in the abundance of animal and plant species. Nature, 296, 245–248.CrossRefGoogle Scholar
Andreotti, R., Gomes, A., Malavazi-Piza, K. C., et al. (2002). BmTI antigens induce a bovine protective immune response against Boophilus microplus tick. International Immunopharmacology, 2, 557–563.CrossRefGoogle ScholarPubMed
Anholt, B. R. & Werner, E. E. (1998). Predictable changes in predation mortality as a consequence of changes in food availability and predation risk. Evolutionary Ecology, 12, 729–738.CrossRefGoogle Scholar
Anonymous. (2004). Human plague in 2002 and 2003. Weekly Epidemiological Record, 79, 301–306.
Apanius, V. (1998). Stress and immune defence. In Stress and Behavior, Advances in the Study of Behavior, vol. 27, ed. M⊘ller, A. P., Milinski, M. & Slater, P. J. B.. New York: Academic Press, pp. 133–153.Google Scholar
Araújo, F. R., Silva, M. P., Lopes, A. A., et al. (1988). Severe cat flea infestation of dairy calves in Brazil. Veterinary Parasitology, 80, 83–86.CrossRefGoogle Scholar
Arneberg, P. (2002). Host population density and body mass as determinants of species richness in parasite communities: comparative analyses of directly transmitted nematodes of mammals. Ecography, 25, 88–94.CrossRefGoogle Scholar
Arneberg, P., Skorping, A. & Read, A. F. (1997). Is population density a species character? Comparative analyses of the nematode parasites of mammals. Oikos, 80, 289–300.CrossRefGoogle Scholar
Arneberg, P., Skorping, A., Grenfell, B. & Read, A. F. (1998). Host densities as determinants of abundance in parasite communities. Proceedings of the Royal Society of London B, 265, 1283–1289.CrossRefGoogle Scholar
Arzamasov, I. T. (1969). Ectoparasite assemblages of insectivores in Belorussia. In Fauna and Ecology of Animals in Belorussia, ed. Anonymous. Minsk, USSR: Academy of Sciences of the Belorussian SSR, pp. 212–221 (in Russian).Google Scholar
Atkinson, W. D. & Shorrocks, B. (1981). Competition on a divided and ephemeral recourse: a simulation model. Journal of Animal Ecology, 50, 461–471.CrossRefGoogle Scholar
Atmar, W. & Patterson, B. D. (1993). The measure of order and disorder in the distribution of species in fragmented habitat. Oecologia, 96, 373–382.CrossRefGoogle ScholarPubMed
Audy, J. R., Radovsky, F. J. & Vercammen-Grandjean, P. H. (1972). Neosomy: radical intrastadial metamorphosis associated with arthropod symbioses. Journal of Medical Entomology, 9, 487–494.CrossRefGoogle ScholarPubMed
Autino, G. A. & Lareschi, M. (1998). Siphonaptera. In Biodiversidad de Artópodos Argentinos, ed. Morrone, J. J. & Ciscaron, S.. La Plata, Argentina: Ediciones SUR, pp. 279–290.Google Scholar
Bacot, A. W. (1914). A study of bionomics of the common rat fleas and other species associated with human habitations, with special reference to the influence of temperature and humidity at various periods of the life history of the insect. Journal of Hygiene, 13, 447–654.Google ScholarPubMed
Bacot, A. W. & Martin, C. J. (1914). Observations on the mechanism of the transmission of plague by fleas. Journal of Hygiene, 13, 423–439.Google ScholarPubMed
Bahmanyar, M. & Cavanaugh, D. C. (1976). Plague Manual. Geneva, Switzerland: World Health Organization.Google Scholar
Baker, K. P. & Elharam, S. (1992). The biology of Ctenocephalides canis in Ireland. Veterinary Parasitology, 45, 141–146.CrossRefGoogle ScholarPubMed
Balashov, Y. S., Bibikova, V. A., Murzakhmetova, K. & Polunina, O. A. (1961). A flea as an environment of the plague pathogen. I. Feeding and digestion in uninfected fleas. In Proceedings of the Interdisciplinary Conference Dedicated to the 40th Anniversary of the Kazakh Soviet Socialist Republic, ed. Anonymous. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 27–30 (in Russian).
Balashov, Y. S., Bibikova, V. A., Murzakhmetova, K. & Polunina, O. A. (1965). Feeding and failure of the foregut valve function in fleas. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 35, 471–476 (in Russian).Google Scholar
Ball, S. L. & Baker, R. L. (1996). Predator-induced life history changes: antipredator behavior costs or facultative life history shifts?Ecology, 77, 1116–1124.CrossRefGoogle Scholar
Ballabeni, P. & Ward, P. I. (1993). Local adaptation of the trematode Diplostomum phoxini to the European minnow Phoxinus phoxinus, its second intermediate host. Functional Ecology, 7, 84–90.CrossRefGoogle Scholar
Banbura, J., Blondel, J., Wilde-Lambrechts, H. & Perret, P. (1995). Why do female blue tits (Parus caeruleus) bring fresh plants to their nests?Journal of Ornithology, 136, 217–221.CrossRefGoogle Scholar
Banet, M. (1986). Fever in mammals: is it beneficial?Yale Journal of Biology and Medicine, 59, 117–124.Google ScholarPubMed
Banfield, A. W. F. (1974). The Mammals of Canada. Toronto, ON: University of Toronto Press.Google Scholar
Bansemir, A. D. & Sukhdeo, M. V. K. (1996). Habitat selection of a gastrointestinal parasite: proximal cues involved in decision making. Parasitology, 113, 311–316.CrossRefGoogle Scholar
Barger, M. A. & Esch, G. W. (2002). Host specificity and the distribution–abundance relationship in a community of parasites infecting fishes in streams of North Carolina. Journal of Parasitology, 88, 446–453.Google Scholar
Barker, S. C. (1991). Evolution of host–parasite associations among species of lice and rock-wallabies: coevolution?International Journal for Parasitology, 21, 497–501.CrossRefGoogle ScholarPubMed
Barnard, C. J., Behnke, J. M., Gage, A. R, Brown, H. & Smithurst, P. R. (1998). The role of parasite-induced immunodepression, rank and social environment in the modulation of behaviour and hormone concentration in male laboratory mice (Mus musculus). Proceedings of the Royal Society of London B, 265, 693–701.CrossRefGoogle Scholar
Barnes, A. M. (1965). Three new species of the genus Anomiopsyllus. Pan-Pacific Entomologist, 41, 272–280.Google Scholar
Barnes, A. M. & Radovsky, F. J. (1969). A new Tunga (Siphonaptera) from the Nearctic region with description of all stages. Journal of Medical Entomology, 6, 19–36.CrossRefGoogle ScholarPubMed
Barnes, A. M., Tipton, V. J. & Wildie, J. A. (1977). The subfamily Anomiopsyllinae (Hystrichopsyllidae: Siphonaptera). I. A revision of the genus Anomiopsyllus Baker. Great Basin Naturalist, 37, 138–206.CrossRefGoogle Scholar
Barrera, A. (1968). The altitudinal distribution of the Siphonaptera of Mount Popocatepetl (Mexico) and a biogeographical interpretation of it. Anales del Instituto de Biologia, Universidad Nacional Autónoma de Mexico, Serie Zoologia, 39, 35–100.Google Scholar
Barriere, P., Beaucournu, J. C., Menier, K. & Colyn, M. (2002). A new flea species of the genus Allopsylla Beaucournu et Fain, 1982 (Siphonaptera: Ischnopsyllidae) from Central African Republic, on a poorly known molossid bat. Parasite, 9, 233–237.Google Scholar
Bartholomew, G. A. & Casey, T. M. (1978). Oxygen consumption of moths during rest, pre-flight warm-up, and flight in relation to body size and wing morphology. Journal of Experimental Biology, 76, 11–25.Google Scholar
Bartkowska, K. (1973). Siphonaptera Tatr Polskich. Fragmenta Faunistica (Warszawa), 19, 227–281.CrossRefGoogle Scholar
Bartolomucci, A., Palanza, P., Gaspani, L., et al. (2001). Social status in mice: behavioral, endocrine and immune changes are context dependent. Physiology and Behavior, 73, 401–410.CrossRefGoogle ScholarPubMed
Bar-Zeev, M. & Sternberg, S. (1962). Factors affecting the feeding of fleas (Xenopsylla cheopis Rothsch.) through a membrane. Entomologia Experimentalis et Applicata, 5, 60–68.CrossRefGoogle Scholar
Bashenina, N. V. (1962). The Ecology of the Common Vole. Moscow, USSR: Moscow University Press (in Russian).Google Scholar
Bashenina, N. V. (1977). Pathways of Adaptations in the Myomorph Rodents. Moscow, USSR: Nauka (in Russian).Google Scholar
Bashenina, N. V. (ed.) (1981). The Bank Vole. Moscow, USSR: Nauka (in Russian).Google Scholar
Bates, J. K. (1962). Field studies on the behaviour of bird fleas. I. Behaviour of the adults of three species of bird fleas in the field. Parasitology, 52, 113–132.CrossRefGoogle Scholar
Baumgartner, D. L. & Kane, A. (1986). Wood ducks as accidental hosts of the squirrel flea, Orchopeas howardi (Siphonaptera, Ceratophyllidae). Great Lakes Entomologist, 19, 249–250.Google Scholar
Bayreuther, K. & Brauning, S. (1971). Die Cytogenetik der Flohe (Aphaniptera). II. Xenopsylla cheopis Rothschild, 1093, und Leptopsylla segnis Schonherr, 1811. Chromosoma, 33, 19–29.CrossRefGoogle Scholar
Bazanova, L. P. & Khabarov, A. V. (2000). Block formation in fleas Citellophilus tesquorum altaicus Ioff, 1936 in relation to the gender of an insect. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 69, 42–44 (in Russian).Google Scholar
Bazanova, L. P. & Mayevsky, M. P. (1996). The duration of persistence of the plaque microbe in the organism of a flea Citellophilus tesquorum altaicus.Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 65, 45–48 (in Russian).Google Scholar
Bazanova, L. P., Voronova, G. A. & Tokmakova, E. G. (2000). Differences in the proventriculus blockage between males and females of Xenopsylla cheopis (Siphonaptera: Pulicidae). Parazitologiya, 34, 56–59 (in Russian).Google Scholar
Bazanova, L. P., Nikitin, A. Y. & Mayevsky, M. P. (2004). Seasonal changes in the aggregated state of the causative agent of plague in the organism of Citellophilus tesquorum altaicus (Siphonaptera). Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 73, 3–6 (in Russian).Google Scholar
Beaucournu, J. C. & Menier, K. (1998). Le genre Ctenocephalides Stiles et Collins, 1930 (Siphonaptera, Pulicidae). Parasite, 5, 3–16.CrossRefGoogle Scholar
Beaucournu, J. C. & Pascal, M. (1998). Origine biogéographique de Nosopsyllus fasciatus (Bosc, 1800) (Siphonaptera – Ceratophyllidae) et observations sur son hôte primitif. Biogeographica, 74, 135–132.Google Scholar
Beaucournu, J. C. & Wells, K. (2004). Trois espèces nouvelles du genre Medwayella Traub, 1972 (Insecta: Siphonaptera: Pygiopsyllidae) de Sabah (Malaisie Orientale, Ile de Bornêo). Parasite, 11, 373–377.CrossRefGoogle Scholar
Beaucournu, J. C., Piver, M. & Guiguen, C. (1993). Actualité de la conquête de l'Afrique intertropical par Pulex irritans Linné, 1758. Bulletin de la Société de Pathologie Exotique, 86, 290–294.Google Scholar
Beaucournu, J. C., Kock, D. & Menier, K. (1997). La souris Mus musculus L., 1758 est-elle l'hôte primitif de la puce Leptopsylla segnis (Schonherr, 1811) (Insecta: Siphonaptera)?Biogeographica, 73, 1–12.Google Scholar
Beaucournu, J. C., Degeilh, B. & Guiguen, C. (2005). Les puces (Insecta: Siphonaptera) parasites d'oiseaux: diversité taxonomique et dispersion biogéographique. Parasite, 12, 111–121.CrossRefGoogle Scholar
Beaucournu, J. C., Vergara, P., Balboa, L. & Gonzalez-Acuna, D. A. (2006). Description d'une nouvelle puce d'oiseau provenant du Chili (Siphonaptera: Ceratophyllidae). Parasite, 13, 227–230.CrossRefGoogle Scholar
Beck, W. (1999). Landwirtschaftliche Nutztiere als Vektoren von parasitären Epizoonoseerregern und zoophilen Dermatophyten. Hautarzt, 50, 621–628.CrossRefGoogle Scholar
Beck, W. & Clark, H. H. (1997). Differentialdiagnose medizinisch relevanter Flohspezies und ihre Bedeutung in der Dermatologie. Hautarzt, 48, 714–719.CrossRefGoogle Scholar
Behnke, J. M., Harris, P. D., Bajer, A., et al. (2004). Variation in the helminth community structure in spiny mice (Acomys dimidiatus) from four montane wadis in the St Katherine region of the Sinai Peninsula in Egypt. Parasitology, 129, 379–398.CrossRefGoogle ScholarPubMed
Beissinger, S. R. & Westphal, M. I. (1998). On the use of demographic models of population viability in endangered species management. Journal of Wildlife Management, 62, 821–841.CrossRefGoogle Scholar
Bekenov, Z. E., Serzhan, O. S., Alashbayev, M. A., et al. (2000). Fluctuations in X. skrjabini population inhabiting the Northern Cis-Aral autonomous plague focus and their causes. Quarantinable and Zoonotic Infections in Kazakhstan, 2, 52–56 (in Russian).Google Scholar
Bell, G. & Burt, A. (1991). The comparative biology of parasite species diversity: intestinal helminths of freshwater fishes. Journal of Animal Ecology, 60, 1046–1063.CrossRefGoogle Scholar
Bell, P. J., Burton, H. R., & Vanfraneker, J. A. (1988). Aspects of the biology of Glaciopsyllus antarcticus (Siphonaptera, Ceratophyllidae) during the breeding-season of a host (Fulmarus glacialoides). Polar Biology, 8, 403–410.CrossRefGoogle Scholar
Belokopytova, A. M., Tchumakova, I. V. & Kozlov, M. P. (1983). Feeding of Xenopsylla conformis, Nosopsyllus laeviceps and Citellophilus tesquorum on reptiles. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 214–215 (in Russian).Google Scholar
Belyavtseva, L. I. (2002). Phenology of fleas Citellophilus tesquorum inhabiting different areas of North Caucasus. Quarantinable and Zoonotic Infections in Kazakhstan, 6, 30–34 (in Russian).Google Scholar
Bengtson, S. A., Brinck-Lindroth, G., Lundqvist, L., Nilsson, A. & Rundgren, S. (1986). Ectoparasites on small mammals in Iceland: origin and population characteristics of a species-poor insular community. Holarctic Ecology, 9, 143–148.Google Scholar
Benjamini, E., Feingold, B. F., Young, J. D., Kartman, L. & Shimizu, M. (1963). Allergy to flea bites. IV. In vitro collection and antigenic properties of the oral secretion of the cat flea. Experimental Parasitology, 13, 143–154.CrossRefGoogle ScholarPubMed
Bennet-Clark, H. C. & Lucey, E. C. (1967). The jump of the flea: a study of the energetics and a model of the mechanism. Journal of Experimental Biology, 47, 59–67.Google Scholar
Bennett, G. F., Caines, J. R. & Bishop, M. A. (1988). Influence of blood parasites on the body mass of passeriform birds. Journal of Wildlife Diseases, 24, 339–343.CrossRefGoogle ScholarPubMed
Benton, A. H., Cerwonka, R. & Hill, J. (1959). Observations on host perception in fleas. Journal of Parasitology, 45, 614.CrossRefGoogle Scholar
Benton, A. H., Surman, M. & Krinsky, W. L. (1979). Observations on the feeding habits of some larval fleas (Siphonaptera). Journal of Parasitology, 65, 671–672.CrossRefGoogle Scholar
Berendyaeva, E. L. & Kudryavtseva, K. F. (1969). Density of fleas on Marmota sibirica in the Upper-Arym autonomous plague focus. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 55–58 (in Russian).Google Scholar
Bernotat-Danielowski, S. & Knülle, W. (1986). Ultrastructure of the rectal sac, the site of water vapor uptake from the atmosphere in larvae of the oriental rat flea Xenopsylla cheopis. Tissue and Cell, 18, 437–455.CrossRefGoogle Scholar
Berseth, W. C. & Zubac, P. (1987). Habitat associations of fleas, Siphonaptera, parasitizing the short-tailed shrew, Blarina brevicauda. Canadian Field Naturalist, 101, 594–596.Google Scholar
Betz, O. & Fuhrmann, S. (2001). Life history traits in different life forms of predaceous Stenus beetles (Coleoptera, Staphylinidae), living in waterside environments. Netherlands Journal of Zoology, 51, 371–393.CrossRefGoogle Scholar
Beutel, R. G. & Gorb, S. N. (2001). Ultrastructure of attachment specializations of hexapods (Arthropoda): evolutionary patterns inferred from a revised ordinal phylogeny. Journal of Zoological Systematics and Evolutionary Research, 39, 177–207.CrossRefGoogle Scholar
Beutel, R. G. & Pohl, H. (2006). Endopterygote systematics: where do we stand and what is the goal (Hexapoda, Arthropoda)?Systematic Entomology, 31, 202–219.CrossRefGoogle Scholar
Beveridge, I. & Chilton, N. B. (2001). Co-evolutionary relationships between the nematode subfamily Cloacininae and its macropodid marsupial hosts. International Journal for Parasitology, 21, 976–996.CrossRefGoogle Scholar
Bgytova, C. I. (1963). Data on the ecology of fleas. IV. Copulation and egg maturation in Ctenophthalmus dolichus under different environmental conditions. In Proceedings of the Scientific Conference of the Natural Focality and Prophylaxis of Plague, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 15–16 (in Russian).Google Scholar
Bibikov, D. I. & Bibikova, V. A. (1955). Studying of the Isabelline wheatear and its ectoparasites. Zoologicheskyi Zhurnal, 34, 399–407 (in Russian).Google Scholar
Bibikova, V. A. (1956). On the biology of fleas parasitic on marmots. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 2, 49–51 (in Russian).Google Scholar
Bibikova, V. A. (1963). Use of microscopy and spectral analysis when studying the hosts' blood in the fleas' stomachs. In Proceedings of the 4th Scientific Conference on Natural Focality and Prophylaxis of Plague, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 21–22 (in Russian).
Bibikova, V. A. (1965). Duration of metamorphosis in Ceratophyllus trispinus balkhaschensis Mik. 1958 under experimental conditions. In Proceedings of the 4th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute and Kainar, pp. 31–32 (in Russian).Google Scholar
Bibikova, V. A. (1977). Contemporary views on the interrelationships between fleas and the pathogens of human and animal diseases. Annual Review of Entomology, 22, 23–32.CrossRefGoogle ScholarPubMed
Bibikova, V. A. & Gerasimova, N. G. (1967). On the biology of Xenopsylla skrjabini Ioff, 1928. II. Flea feeding under experimental conditions. Zoologicheskyi Zhurnal, 46, 730–736 (in Russian).Google Scholar
Bibikova, V. A. & Gerasimova, N. G. (1973). Feeding of fleas Xenopsylla nuttalli Ioff, 1930 in the experiments. Problems of Particularly Dangerous Diseases, 29, 122–125 (in Russian).Google Scholar
Bibikova, V. A. & Klassovsky, L. N. (1974). Transmission of Plague by Fleas. Moscow, USSR: Meditsina (in Russian).Google Scholar
Bibikova, V. A. & Zhovty, I. F. (1980). Review of certain studies of fleas in the USSR, 1967–1976. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 257–272.Google Scholar
Bibikova, V. A., Ilyinskaya, V. L., Kaluzhenova, Z. P., Morozova, I. V. & Shmuter, M. F. (1963). On the biology of fleas of the genus Xenopsylla in the Sary-Ishik-Otrau Desert. Zoologicheskyi Zhurnal, 42, 1045–1050 (in Russian).Google Scholar
Bibikova, V. A., Zolotova, S. I., Murzakhmetova, K. & Leonova, T. N. (1971). Fecundity and its dynamics in two fleas parasitic on the great gerbil under experimental conditions. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 367–369 (in Russian).Google Scholar
Bidashko, F. G., Grazhdanov, A. K., Medzykhovsky, G. A., et al. (2001). Results of the studies of flea abundance in human houses in the plague focus of the West-Kazakhstan Region. Quarantinable and Zoonotic Infections in Kazakhstan, 4, 89–93 (in Russian).Google Scholar
Bidashko, F. G., Grazhdanov, A. K., Tanitovsky, B. A., et al. (2004). Attack rate of fleas Pulex irritans on humans in various biotopes of human houses. Quarantinable and Zoonotic Infections in Kazakhstan, 9, 58–61 (in Russian).Google Scholar
Biliński, S. M., Büning, J. & Simiczyjew, B. (1998). The ovaries of Mecoptera: basic similarities. Folia Histochemica et Cytobiologica, 36, 189–196.Google ScholarPubMed
Bilyalov, Z. A., Ershova, L. S., Sviridov, G. G., Sokolov, P. N. & Arakelyants, V. S. (1989). A case of experimental transmission of the causative agent of plague from Ornithodoros tartakovskyi ticks to fleas of the great gerbil. Parazitologiya, 23, 362–364 (in Russian).Google Scholar
Birtles, R. J., Harrison, T. G. & Molyneux, D. H. (1994). Grahamella in small woodland mammals in the UK: isolation, prevalence and host-specificity. Annals of Tropical Medicine and Parasitology, 88, 317–327.
Bitam, I., Parola, P., Cruz, Dittmar K., et al. (2006). First molecular detection of Rickettsia felis in fleas from Algeria. American Journal of Tropical Medicine and Hygiene, 74, 532–535.Google ScholarPubMed
Black, C. C. & Krishtalka, L. (1986). Rodents, bats, and insectivores from the Plio-Pleistocene sediments to the east of Lake Turkana, Kenya. Contributions in Science, Natural History Museum of Los Angeles County, 372, 1–15.Google Scholar
Blackburn, T. M. & Gaston, K. J. (2001). Linking patterns in macroecology. Journal of Animal Ecology, 70, 338–352.CrossRefGoogle Scholar
Blanc, G. & Baltazard, M. (1941). Transmission du bacille de Withmore par la puce du rat Xenopsylla cheopis. Comptes Rendus de l'Académie des Sciences, 213, 541–543.Google Scholar
Blanc, G. & Baltazard, M. (1942). Sur le mécanisme de la transmission de la peste par Xenopsylla cheopis. Comptes Rendus de l'Académie des Sciences, 136, 645–647.Google Scholar
Blanc, G. & Baltazard, M. (1944a). Contribution à l'étude du comportement des microbes pathogènes chez les insectes hématophages. Premier mémoire. Archives de l'Institut Pasteur du Maroc, 3, 21–49.Google Scholar
Blanc, G. & Balatazard, M. (1944b). Revue chronologique des recherches expérimentales sur la transmission et la conservation naturelle des typhus. Archives de l'Institut Pasteur du Maroc, 2, 535–715.Google Scholar
Blank, S. M., Kutzscher, C., Masello, J. F., Pilgrim, R. L. C. & Quillfeldt, P. (2007). Stick-tight fleas in the nostrils and below the tongue: evolution of an extraordinary infestation site in Hectopsylla (Siphonaptera: Pulicidae). Zoological Journal of the Linnean Society, 149, 117–137.CrossRefGoogle Scholar
Blaski, M. (1989). Fleas occurring on rodents living in the area of coal-mine Boleslaw Smialy in Laziska (Poland). Prace Naukowe Uniwersytetu Slaskiego w Katowicach, 1035, 92–96 (in Polish).Google Scholar
Blaski, M. (1991). Fleas (Siphonaptera) collected on the rodents from two industrial plants in Silesia. Prace Naukowe Uniwersytetu Slaskiego w Katowicach, 1184, 87–88 (in Polish).Google Scholar
Blaski, M. (2004). Seasonal dynamics of Ceratophyllus hirundinis (Curtis, 1826) (Siphonaptera, Insecta) in the nests of Delichon urbica (L.). Wiadomoῦci Parazytologiczne, 50, 31–34 (in Polish).Google Scholar
Blount, J. D., Houston, D. C., M⊘ller, A. P. & Wright, J. (2003). Do individual branches of immune defence correlate? A comparative case study of scavenging and non-scavenging birds. Oikos, 102, 340–350.CrossRefGoogle Scholar
Bock, C. E., Cruz, A., Grant, M. C., Aid, C. S. & Strong, T. R. (1992). Field experimental evidence for diffuse competition among southwestern riparian birds. American Naturalist, 140, 515–528.CrossRefGoogle ScholarPubMed
Bodrova, T. V. & Zhovty, I. F. (1961). Fleas of the Daurian ground squirrel in the area of the Zun-Torei Lake (S.-E. Trans-Baikalia). Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 1, 82–85 (in Russian).Google Scholar
Boeken, B. & Shachak, M. (1998). The dynamics of abundance and incidence of annual plant species richness during colonization in a desert. Ecography, 21, 63–73.CrossRefGoogle Scholar
Bogdanov, I. I., Malkova, M. G., Yakimenko, V. V. & Tantsev, A. K. (2001). Parasite–host associations between fleas and small mammals in the Omsk Region. Parazitologiya, 35, 184–191 (in Russian).Google Scholar
Bohn, T. & Amundsen, P. A. (2001). The competitive edge of an invading specialist. Ecology, 82, 2150–2163.CrossRefGoogle Scholar
Bolger, D. T., Alberts, A. C. & Soulé, M. E. (1991). Occurrence patterns of bird species in habitat fragments: sampling, extinction, and nested species subsets. American Naturalist, 137, 155–166.CrossRefGoogle Scholar
Bossard, R. L. (2002). Speed and Reynolds number of jumping cat fleas (Siphonaptera: Pulicidae). Journal of the Kansas Entomological Society, 75, 52–54.Google Scholar
Bossard, R. L. (2006). Mammal and flea relationships in the Great Basin Desert: from H. J. Egoscue's collections. Journal of Parasitology, 92, 260–266.CrossRefGoogle Scholar
Bossard, R. L., Broce, A. B. & Dryden, M. W. (2000). Effects of circadian rhythms and other bioassay factors on cat flea (Pulicidae: Siphonaptera) susceptibility to insecticides. Journal of the Kansas Entomological Society, 73, 21–29.Google Scholar
Botelho, J. R. & Linardi, P. M. (1996). Interrelações entre ectoparasitos e roedores em ambientes silvestre e urbano de Belo Horizonte, Minas Gerais, Brasil. Revista Brasileira de Entomologia, 40, 425–430.Google Scholar
Bouchet, F., Guidon, N., Dittmar, K., et al. (2003). Parasite remains in archeological sites. Memórias do Instituto Oswaldo Cruz, 98, 39–47.CrossRefGoogle Scholar
Boudreaux, H. B. (1979). Arthropod Phylogeny with Special Reference to Insects. New York: John Wiley.Google Scholar
Boughton, R. K., Atwell, J. W. & Schoech, S. J. (2006). An introduced generalist parasite, the sticktight flea (Echidnophaga gallinacea), and its pathology in the threatened Florida scrub-jay (Aphelocoma coerulescens). Journal of Parasitology, 92, 941–948.CrossRefGoogle Scholar
Boulouis, H.-J., Chang, C.-C., Henn, J. B., Kasten, R. W. & Chomel, B. B. (2005). Factors associated with the rapid emergence of zoonotic Bartonella infections. Veterinary Research, 36, 383–410.CrossRefGoogle ScholarPubMed
Bouslama, Z., Chabi, Y. & Lambrechts, M. M. (2001). Chicks resist high parasite intensities in an Algerian population of blue tits. Ecoscience, 8, 320–324.CrossRefGoogle Scholar
Bouslama, Z., Lambrechts, M. M., Ziane, N., Djenidi, R. D. & Chabi, Y. (2002). The effect of nest ectoparasites on parental provisioning in a north-African population of the blue tit Parus caeruleus. Ibis, 144, E73–E78.CrossRefGoogle Scholar
Boycott, A. E. (1913). The reaction of flea bites. Journal of Pathology and Bacteriology, 17, 110.Google Scholar
Bozeman, F. M., Sonenshine, D. E., Williams, M. S., et al. (1981). Experimental infection of ectoparasitic arthropods with Rickettsia prowazekii (GvF-16 strain) and transmission to flying squirrels. American Journal of Tropical Medicine and Hygiene, 30, 253–263.CrossRefGoogle ScholarPubMed
Bradley, T. J. & Hetz, S. (2001). Specific dynamic action in the insect Rhodnius prolixus. American Zoologist, 41, 1397.Google Scholar
Braks, M. A. H., Honorio, N. A., Lounibos, L. P., Oliveira, R. L. & Juliano, S. A. (2004). Interspecific competition between two invasive species of container mosquitoes, Aedes aegypti and Aedes albopictus (Diptera: Culicidae), in Brazil. Annals of the Entomological Society of America, 97, 130–139.CrossRefGoogle Scholar
Brändle, M., Amarell, U., Auge, H., Klotz, S. & Brandl, R. (2001). Plant and insect diversity along a pollution gradient: understanding species richness across trophic levels. Biodiversity and Conservation, 10, 1497–1511.CrossRefGoogle Scholar
Brandt, C. A. (1992). Social factors in immigration and emigration. In Animal Dispersal: Small Mammals as a Model, ed. Stenseth, N. C. & Lidicker, W. Z.. London: Chapman & Hall, pp. 96–141.CrossRefGoogle Scholar
Bray, J. R. & Curtis, J. T. (1957). An ordination of the upland forest communities of Southern Wisconsin. Ecological Monographs, 27, 325–349.
Breitschwerdt, E. B. & Kordick, D. L. (2000). Bartonella infection in animals: carriership, reservoir potential, pathogenicity, and zoonotic potential for human infection. Clinical Microbiology Reviews, 13, 428–438.CrossRefGoogle ScholarPubMed
Brinck, G. (1966). Siphonaptera from small mammals in natural foci of tick-borne encephalitis virus in Sweden. Opuscula Entomologica, 31, 156–170.Google Scholar
Brinck, G. & Löfqvist, J. (1973). The hedgehog Erinaceus europeus and its flea Archaeopsylla erinacei. Zoon (Supplement), 1, 97–103.Google Scholar
Brinck-Lindroth, G. (1968). Host spectra and distribution of fleas of small mammals in Swedish Lapland. Opuscula Entomologica, 33, 327–358.Google Scholar
Brinkerhoff, R. J., Markeson, A. B., Knouft, J. A., Gage, K. L. & Montenieri, J. A. (2006). Abundance patterns of two Oropsylla (Ceratophyllidae: Siphonaptera) species on black-tailed prairie dog (Cynomys ludovicianus) hosts. Journal of Vector Ecology, 31, 355–363.CrossRefGoogle ScholarPubMed
Brinkhof, M. W. G., Heeb, P., Kölliker, M. & Richner, H. (1999). Immunocompetence of nestling great tits in relation to rearing environment and parentage. Proceedings of the Royal Society of London B, 266, 2315–2322.CrossRefGoogle Scholar
Brooks, D. R. (1988). Macroevolutionary comparisons of host and parasite phylogenies. Annual Review of Ecology and Systematics, 19, 235–259.CrossRefGoogle Scholar
Brooks, D. R. & McLennan, D. A. (1991). Phylogeny, Ecology, and Behaviour: A Research Program in Comparative Biology. Chicago, IL: University of Chicago Press.Google Scholar
Brouwer, L. & Komdeur, J. (2004). Green nesting material has a function in mate attraction in the European starling. Animal Behaviour, 67, 539–548.CrossRefGoogle Scholar
Brown, C. R. & Brown, M. B. (1986). Ectoparasitism as a cost of coloniality in cliff swallows (Hirundo pyrrhonota). Ecology, 67, 1206–1218.CrossRefGoogle Scholar
Brown, C. R. & Brown, M. B. (1992). Ectoparasitism as a cause of natal dispersal in cliff swallows. Ecology, 73, 1718–1723.CrossRefGoogle Scholar
Brown, C. R., Brown, M. B. & Rannala, B. (1995). Ectoparasites reduce long-term survival of their avian host. Proceedings of the Royal Society of London B, 262, 313–319.CrossRefGoogle Scholar
Brown, J. H. (1975) Geographical ecology of desert rodents. In Ecology and Evolution of Communities, ed. Cody, M. L. & Diamond, J. M.. Cambridge, MA: Harvard University Press, pp. 315–341.Google Scholar
Brown, J. H. (1984). On the relationship between abundance and distribution of species. American Naturalist, 124, 255–279.CrossRefGoogle Scholar
Brown, J. H. (1995). Macroecology. Chicago, IL: University of Chicago Press.Google Scholar
Brown, J. H. & Lomolino, M. V. (1998). Biogeography. Sunderland, MA: Sinauer Associates.Google Scholar
Brown, J. H., Stevens, G. C. & Kaufman, D. M. (1996). The geographic range: size, shape, boundaries, and internal structure. Annual Review of Ecology and Systematics, 27, 597–623.CrossRefGoogle Scholar
Brown, J. S., Kotler, B. P. & Mitchell, W. M. (1994). Foraging theory, patch use, and the structure of a Negev Desert granivore community. Ecology, 75, 2286–2300.CrossRefGoogle Scholar
Brown, S. J. (1989). Pathological consequences of feeding by hematophagous arthropods: comparison of feeding strategies. In Proceedings of a Symposium ‘Physiological Interactions between Hematophagous Arthropods and their Vertebrate Hosts’, ed. Johnston, C. J. & Williams, R. E.. Lanham, MD: Entomological Society of America, pp. 4–14.Google Scholar
Bruce, W. N. (1948). Studies on the biological requirements of the cat flea. Annals of the Entomological Society of America, 41, 346–352.CrossRefGoogle Scholar
Bryukhanova, L. V. (1966). Reproduction and feeding of fleas parasitic on ground squirrels. In Particularly Dangerous Diseases in Caucasus, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 37–40 (in Russian).Google Scholar
Bryukhanova, L. V. & Myalkovskaya, C. A. (1974). On the duration of metamorphosis of fleas in the nests of the pygmy ground squirrel. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 121–124 (in Russian).Google Scholar
Bryukhanova, L. V. & Surkova, L. A. (1970). On the biology of Frontopsylla semura Wagn. et Ioff, 1926. In Vectors of Particularly Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 247–252 (in Russian).Google Scholar
Bryukhanova, L. V. & Surkova, L. A. (1983). Digestion of the blood of ground squirrels by fleas. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 220–222 (in Russian).Google Scholar
Bryukhanova, L. V., Sardar, E. A. & Levi, M. I. (1961). On the method of measurement the amount of blood taken by fleas. Proceedings of Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, 5, 28–32 (in Russian).Google Scholar
Bryukhanova, L. V., Darskaya, N. F. & Surkova, L. A. (1978). Blood digestion by fleas Leptopsylla segnis Schöncher. Parazitologiya, 12, 383–386 (in Russian).Google Scholar
Bryukhanova, L. V., Darskaya, N. F., Surkova, L. A. & Karandina, R. S. (1983). Digestion of the blood of gerbils by fleas. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 219–220 (in Russian).Google Scholar
Buckland, P. C. & Sadler, J. P. (1989). A biogeography of the human flea, Pulex irritans L. (Siphonaptera, Pulicidae). Journal of Biogeography, 16, 115–120.CrossRefGoogle Scholar
Buckle, A. P. (1978). The mark, release and recapture of fleas in a wild population of woodmice, Apodemus sylvaticus. Journal of Zoology, 186, 563–567.Google Scholar
Buechler, K., Fitze, P. S., Gottstein, B., Jacot, A. & Richner, H. (2002). Parasite-induced maternal response in a natural bird population. Journal of Animal Ecology, 71, 247–252.CrossRefGoogle Scholar
Buchman, K. (1991). Relationship between host size of Anguilla anguilla and the infection level of the monogeneans Pseudodactylogyrus spp. Journal of Fish Biology, 35, 599–601.CrossRefGoogle Scholar
Burdelov, A. S., Sabilayev, A. S., Mu, T. srepov, et al. (1999). Spatial distribution of fleas Xenopsylla skrjabini and X. gerbilli minax in colonies of the great gerbil in the north-western Pri-Balkhash area. Parazitologiya, 33, 493–496 (in Russian).Google Scholar
Burdelov, L. A., Zhubanazarov, I. Z. & Rudenchik, N. F. (1989). The size of small mammals and the number of fleas parasitizing them. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 58, 42–45 (in Russian).Google Scholar
Burdelov, S. A., Leiderman, M., Krasnov, B. R., Khokhlova, I. S. & Degen, A. A. (2007). Locomotor response to light and surface angle in three species of desert fleas. Parasitology Research, 100, 973–982.CrossRefGoogle ScholarPubMed
Burdelova, N. V. (1996). Flea fauna of some small mammals in the Dzhungarskyi Ala-Tau Ridge. In Proceedings of the Conference ‘Ecological Aspects of Epidemiology and Epizootology of Plague and Other Dangerous Diseases’, ed. Burdelov, L. A.. Almaty, Kazakhstan: Kazakh Scientific Center for Quarantine and Zoonotic Diseases, pp. 119–120 (in Russian).Google Scholar
Burdelova, N. V. & Burdelov, V. A. (1983). Dynamics of flea species composition at high and low density of the great gerbil in the Akdala Valley. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 222–223 (in Russian).Google Scholar
Burroughs, A. L. (1947). Sylvatic plague studies: the vector efficiency of nine species of fleas compared with Xenopsylla cheopis. Journal of Hygiene, 43, 371–396.CrossRefGoogle Scholar
Burrows, M. & Wolf, H. (2002). Jumping and kicking in the false stick insect Prosarthria teretrirostris: kinematics and motor control. Journal of Experimental Biology, 205, 1519–1530.Google ScholarPubMed
Bursell, E. (1974). Environmental aspects: humidity. In The Physiology of Insects, vol. 2, ed. Rockstein, M.. New York: Academic Press, pp. 44–84.Google Scholar
Bursten, S. N., Kimsey, R. B. & Owings, D. H. (1997). Ranging of male Oropsylla montana fleas via male California ground squirrel (Spermophilus beecheyi) juveniles. Journal of Parasitology, 83, 804–809.CrossRefGoogle ScholarPubMed
Busalaeva, N. N. & Fedosenko, A. K. (1964). Fleas parasitic on small mammals in the high mountain areas of the Trans-Ili Ala-Tau Ridge. Proceedings of the Institute of Zoology of Academy of Sciences of the Kazakh SSR, 22, 177–183 (in Russian).Google Scholar
Bush, A. O. & Holmes, J. C. (1986). Intestinal helminths of lesser scaup ducks: patterns of association. Canadian Journal of Zoology, 64, 132–141.CrossRefGoogle Scholar
Bush, A. O., Lafferty, K. D., Lotz, J. M. & Shostak, A. W. (1997). Parasitology meets ecology on its own terms: Margolis et al. revisited. Journal of Parasitology, 83, 575–583.CrossRefGoogle Scholar
Butler, J. M. & Roper, T. J. (1996). Ectoparasites and sett use in European badgers. Animal Behaviour, 52, 621–629.CrossRefGoogle Scholar
Buxton, P. A. (1948). Experiments with mice and fleas. I. The baby mouse. Parasitology, 39, 119–124.CrossRefGoogle ScholarPubMed
Byers, G. W. (1996). More on the origin of Siphonaptera. Journal of the Kansas Entomological Society, 69, 274–277.Google Scholar
Cabrero-Sanudo, F. J. & Lobo, J. M. (2003). Estimating the number of species not yet described and their characteristics: the case of the Western Palaearctic dung beetle species (Coleoptera, Scarabaeoidea). Biodiversity and Conservation, 12, 147–166.CrossRefGoogle Scholar
Cadiergues, M. C., Joubert, C. & Franc, M. (2000). A comparison of jump performances of the dog flea, Ctenocephalides canis (Curtis, 1826) and the cat flea, Ctenocephalides felis felis (Bouché, 1835). Veterinary Parasitology, 92, 239–241.CrossRefGoogle Scholar
Cadiergues, M. C., Santamarta, D., Mallet, X. & Franc, M. (2001). First blood meal of Ctenocephalides canis (Siphonaptera: Pulicidae) on dogs: time to initiation of feeding and duration. Journal of Parasitology, 87, 214–215.CrossRefGoogle ScholarPubMed
Cai, W.-F., Fang, Z. & Yang, G.-R. (2000). Effects of temperature, humidity on the population emergence rate of Xenopsylla cheopis in laboratory. Chinese Journal of Vector Biology and Control, 11, 358–361 (in Chinese).Google Scholar
Caley, M. J. & Schluter, D. (1997). The relationship between local and regional diversity. Ecology, 78, 70–80.CrossRefGoogle Scholar
Callaway, R. M. & King, L. (1996). Temperature-driven variation in substrate oxygenation and the balance of competition and facilitation. Ecology, 77, 1189–1195.CrossRefGoogle Scholar
Callaway, R. M. & Walker, L. R. (1997). Competition and facilitation: a synthetic approach to interactions in plant communities. Ecology, 78, 1958–1965.CrossRefGoogle Scholar
Calvete, C., Blanco-Aguiar, J. A., Virgós, E., Cabezas-Díaz, S. & Villafuerte, R. (2004). Spatial variation in helminth community structure in the red-legged partridge (Alectoris rufa L.): effects of definitive host density. Parasitology, 129, 101–113.CrossRefGoogle ScholarPubMed
Campos, E. G., Maupin, G. O., Barnes, A. M. & Eads, R. B. (1985). Seasonal occurrence of fleas (Siphonaptera) on rodents in a foothills habitat in Larimer County, Colorado, USA. Journal of Medical Entomology, 22, 266–270.CrossRefGoogle Scholar
Carlier, Y. & Truyens, C. (1995). Influence of maternal infection on offspring resistance towards parasites. Parasitology Today, 11, 94–99.CrossRefGoogle ScholarPubMed
Carney, J. P. & Dick, T. A. (2000). Helminth communities of yellow perch (Perca flavescens (Mitchill)): determinants of pattern. Canadian Journal of Zoology, 78, 538–555.CrossRefGoogle Scholar
Caro, A., Combes, C. & Euzet, L. (1997). What makes a fish a suitable host for Monogenea in the Mediterranean?Journal of Helminthology, 71, 203–210.CrossRefGoogle Scholar
Carson, H. L. (1959). Genetic conditions that promote or retard the formation of species. Cold Spring Harbor Symposia on Quantitative Biology, 24, 87–103.CrossRefGoogle ScholarPubMed
Case, T. J. & Taper, M. L. (2000). Interspecific competition, environmental gradients, gene flow, and the coevolution of species' borders. American Naturalist, 155, 583–605.CrossRefGoogle ScholarPubMed
Case, T. J., Holt, R. D., McPeek, M. A. & Keitt, T. H. (2005). The community context of species' borders: ecological and evolutionary perspectives. Oikos, 108, 28–46.CrossRefGoogle Scholar
Castleberry, S. B., Castleberry, N. L., Wood, P. B., Ford, W. M. & Mengak, M. T. (2003). Fleas (Siphonaptera) of the Allegheny woodrat (Neotoma magister) in West Virginia with comments on host specificity. American Midland Naturalist, 149, 233–236.CrossRefGoogle Scholar
Castro, J. M., Nolan, V. & Ketterson, E. D. (2001). Steroid hormones and immune function: experimental studies in wild and captive dark-eyed juncos (Junco hyemalis). American Naturalist, 157, 408–420.Google Scholar
Cavitt, J. F., Pearse, A. T. & Miller, T. A. (1999). Brown thrasher nest reuse: a time saving resource, protection from search-strategy predators, or cues for nest-site selection?Condor, 101, 859–862.CrossRefGoogle Scholar
Chadee, D. D. (1994). Distribution patterns of Tunga penetrans within a community in Trinidad, West Indies. Journal of Tropical Medicine and Hygiene, 97, 167–170.Google ScholarPubMed
Chandy, L. & Prasad, R. (1987). Behavioral resistance of hosts of flea infestation. Proceedings of the Indian National Science Academy B, 53, 27–30.Google Scholar
Charleston, M. A. (1998). Jungles: a new solution to the host/parasite phylogeny reconciliation problem. Mathematical Biosciences, 149, 191–223.CrossRefGoogle ScholarPubMed
Chase, J. M. (1996 ). Differential competitive interactions and the included niche: an experimental analysis with grasshoppers. Oikos, 76, 103–112.CrossRefGoogle Scholar
Chastel, O. & Beaucournu, J. C. (1992). Specificity and ecoethology of bird fleas in Kerguelen Islands. Annales de Parasitologie Humaine et Comparée, 67, 213–220.CrossRefGoogle Scholar
Chekchak, T., Chapuis, J. L., Pisanu, B. & Bousses, P. (2000). Introduction of the rabbit flea, Spilopsyllus cuniculi (Dale), to a subantarctic island (Kerguelen Archipelago) and its assessment as a vector of myxomatosis. Wildlife Research, 27, 91–101.CrossRefGoogle Scholar
Chesson, P. & Huntly, N. (1997). The roles of harsh and fluctuating conditions in the dynamics of ecological communities. American Naturalist, 150, 519–553.CrossRefGoogle ScholarPubMed
Christe, P., Oppliger, A. & Richner, H. (1994). Ectoparasite affects choice and use of roost sites in the great tit (Parus major). Animal Behaviour, 47, 895–898.CrossRefGoogle Scholar
Christe, P., Richner, H. & Oppliger, A. (1996a). Begging, food provisioning, and nestling competition in great tit broods infested with ectoparasites. Behavioral Ecology, 7, 127–131.CrossRefGoogle Scholar
Christe, P., Richner, H. & Oppliger, A. (1996b). Of great tits and fleas: sleep baby sleep …. Animal Behaviour, 52, 1087–1092.CrossRefGoogle Scholar
Christe, P., M⊘ller, A. P. & Lope, F. (1998). Immunocompetence and nestling survival in the house martin: the tasty chick hypothesis. Oikos, 83, 175–179.CrossRefGoogle Scholar
Christe, P., Arlettaz, R. & Vogel, P. (2000). Variation in intensity of a parasitic mite (Spinturnix myoti) in relation to the reproductive cycle and immunocompetence of its bat host (Myotis myotis). Ecology Letters, 3, 207–212.CrossRefGoogle Scholar
Christe, P., Giorgi, M. S., Vogel, P. & Arlettaz, R. (2003). Differential species-specific ectoparasitic mite intensities in two intimately coexisting sibling bat species: resource-mediated host attractiveness or parasite specialization?Journal of Animal Ecology, 72, 866–872.CrossRefGoogle Scholar
Christe, P., Michaux, J. & Morand, S. (2006). Biological conservation and parasitism. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 593–613.CrossRefGoogle Scholar
Christian, K. A. & Bedford, G. S. (1995). Physiological consequences of filarial parasites in the frillneck lizard, Chlamydosaurus kingii, in Northern Australia. Canadian Journal of Zoology, 73, 2302–2306.CrossRefGoogle Scholar
Christodoulopoulos, G. & Theodoropoulos, G. (2003). Infestation of dairy goats with the human flea, Pulex irritans, in central Greece. Veterinary Record, 152, 371–372.CrossRefGoogle ScholarPubMed
Christodoulopoulos, G., Theodoropoulos, G., Kominakis, A. & Theis, J. H. (2006). Biological, seasonal and environmental factors associated with Pulex irritans infestation of dairy goats in Greece. Veterinary Parasitology, 137, 137–143.CrossRefGoogle Scholar
Clark, F. & McNeil, D. A. C. (1981). The variation in population densities of fleas in house martin nests in Leicestershire. Ecological Entomology, 6, 379–386.CrossRefGoogle Scholar
Clark, F. & McNeil, D. A. C. (1991). Temporal variation in the population densities of fleas in house martin nests (Delichon u. urbica (Linnaeus)) in Leicestershire, U.K. Entomologist's Gazette, 42, 281–288.Google Scholar
Clark, F. & McNeil, D. A. C. (1993). A study of some factors affecting mortality overwinter in three congeneric species of bird fleas. Entomologist, 112, 55–66.Google Scholar
Clark, F., Greenwood, M. T. & Smith, J. S. (1993a). The use of an insect activity monitor in behavioral studies of the flea, Xenopsylla cheopis (Rothschild). Bulletin of the Society of Vector Ecology, 18, 26–32.Google Scholar
Clark, F., McNeil, D. A. C. & Hill, L. A. (1993b). Studies on the dispersal of three congeneric species of flea monoxenous to the house martin (Delichon urbica (L.)). Entomologist, 112, 85–94.Google Scholar
Clark, F., Deadman, D., Greenwood, M. T. & Larsen, K. S. (1997). A circadian rhythm of locomotor activity in newly emerged Ceratophyllus sciurorum. Medical and Veterinary Entomology, 11, 213–216.CrossRefGoogle ScholarPubMed
Clark, F., Deadman, D., Greenwood, M. T., Larsen, K. S. & Pudney, S. (1999). Effects of feeding on the circadian rhythm of Ceratophyllus s. sciurorum (Siphonaptera: Ceratophyllidae). Journal of Vector Ecology, 24, 78–82.Google Scholar
Clark, L. & Mason, J. R. (1988). Effect of biologically active plants used as nest material and the derived benefit to starling nestlings. Oecologia, 77, 174–180.CrossRefGoogle ScholarPubMed
Clarke, K. R. & Warwick, R. M. (1998). A taxonomic distinctness index and its statistical properties. Journal of Applied Ecology, 35, 523–531.CrossRefGoogle Scholar
Clarke, K. R. & Warwick, R. M. (1999). The taxonomic distinctness measure of biodiversity: weighting of step lengths between hierarchical levels. Marine Ecology Progress Series, 184, 21–29.CrossRefGoogle Scholar
Clarke, K. R. & Warwick, R. M. (2001). A further biodiversity index applicable to species lists: variation in taxonomic distinctness. Marine Ecology Progress Series, 216, 265–278.CrossRefGoogle Scholar
Clayton, D. H. & Cotgreave, P. (1994). Relationship of bill morphology to grooming behaviour in birds. Animal Behaviour, 47, 195–201.CrossRefGoogle Scholar
Clayton, D. H. & Moore, J. (1997). Introduction. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 1–6.Google Scholar
Clayton, D. H. & Tompkins, D. M. (1994). Ectoparasite virulence is linked to mode of transmission. Proceedings of the Royal Society of London B, 256, 211–217.CrossRefGoogle ScholarPubMed
Clayton, D. H. & Tompkins, D. M. (1995). Comparative effects of mites and lice on the reproductive success of rock doves (Columba livia). Parasitology, 110, 195–206.
Clayton, D. H., Al-Tamimi, S. & Johnston, K. P. (2003). The ecological basis of coevolutionary history. In Tangled Trees: Phylogeny, Cospeciation and Coevolution, ed. Page, R. D. M.. Chicago, IL: University of Chicago Press, pp. 310–342.Google Scholar
Clements, A. N. (1992). The Biology of Mosquitoes: Development, Nutrition and Reproduction, vol. 1. Wallingford, UK: CAB International.Google Scholar
Clements, J. F. (2007). The Clements Checklist of Birds of the World, 6th edn. London: Christopher Helm.
Cole, L. C. & Koepke, J. A. (1946). A study of rodent ectoparasites in Mobile, Alabama. Public Health Reports, 61, 1469–1487.CrossRefGoogle Scholar
Cole, L. C. & Koepke, J. A. (1947a). Problems of interpretation of the data of rodent–ectoparasite surveys. Public Health Reports(Supplement), 202, 1–24.Google Scholar
Cole, L. C. & Koepke, J. A. (1947b). A study of rodent ectoparasites in Honolulu. Public Health Reports (Supplement), 202, 25–41.Google Scholar
Cole, L. C. & Koepke, J. A. (1947c). A study of rodent ectoparasites in Savannah, Georgia. Public Health Reports (Supplement), 202, 42–59.Google Scholar
Cole, L. C. & Koepke, J. A. (1947d). A study of rodent ectoparasites in Dothan, Alabama. Public Health Reports (Supplement), 202, 61–71.Google Scholar
Collen, B., Purvis, A. & Gittleman, J. L. (2004). Biological correlates of description date in carnivores and primates. Global Ecology and Biogeography, 13, 459–467.CrossRefGoogle Scholar
Collinge, S. K., Johnson, W. C., Ray, C., et al. (2005). Landscape structure and plague occurrence in black-tailed prairie dogs on grasslands of the western USA. Landscape Ecology, 20, 941–955.CrossRefGoogle Scholar
Colwell, R. K. (2000). Rensch's rule crosses the line: convergent allometry of sexual size dimorphism in hummingbirds and flower mites. American Naturalist, 156, 495–510.Google ScholarPubMed
Combes, C. (1997). Fitness of parasites: pathology and selection. International Journal for Parasitology, 27, 1–10.CrossRefGoogle Scholar
Combes, C. (2001). Parasitism: The Ecology and Evolution of Intimate Interactions. Chicago, IL: University of Chicago Press.Google Scholar
Combes, C. (2005). The Art of Being a Parasite. Chicago, IL: University of Chicago Press.Google Scholar
Cooke, B. D. (1990a). Rabbit burrows as environments for the European rabbit fleas, Spilopsyllus cuniculi (Dale), in arid South Australia. Australian Journal of Zoology, 38, 317–325.CrossRefGoogle Scholar
Cooke, B. D. (1990b). Notes on the comparative reproductive biology and the laboratory breeding of the rabbit flea Xenopsylla cunicularis Smit (Siphonaptera, Pulicidae). Australian Journal of Zoology, 38, 527–534.CrossRefGoogle Scholar
Cooke, B. D. (1999). Notes on the life history of the rabbit flea Caenopsylla laptevi ibera Beaucornu & Marquez, 1987 (Siphonaptera: Ceratophyllidae) in eastern Spain. Parasite, 6, 347–354.CrossRefGoogle Scholar
Cooke, B. D. & Skewes, M. A. (1988). The effect of temperature and humidity on the survival and development of the European rabbit flea Spilopsyllus cuniculi (Dale). Australian Journal of Zoology, 36, 449–459.CrossRefGoogle Scholar
Cooper, J. E. (1976). Tunga penetrans infestation in pigs. Veterinary Record, 98, 472.CrossRefGoogle ScholarPubMed
Cornell, H. V. (1993). Unsaturated patterns in species assemblages: the role of regional processes in setting local species richness. In Species Diversity in Ecological Communities: Historical and Geographical Perspectives, ed. Ricklefs, R. F. & Schluter, D.. Chicago, IL: University of Chicago Press, pp. 243–252.Google Scholar
Cornell, H. V. & Lawton, J. H. (1992). Species interactions, local and regional processes, and limits to richness of ecological communities: a theoretical perspective. Journal of Animal Ecology, 61, 1–12.CrossRefGoogle Scholar
Correia, T. R., Souza, C. P., Fernandes, J. I., et al. (2003). Life cycle of Ctenocephalides felis felis (Bouché, 1835) (Siphonaptera, Pulicidae) from different artificial diets. Revista Brasiliera de Zoociências, Juiz de Fora, 5, 153–160.Google Scholar
Côté, I. M. & Poulin, R. (1995). Parasitism and group size in social animals: a meta-analysis. Behavioral Ecology, 6, 159–165.CrossRefGoogle Scholar
Cotgreave, P. & Clayton, D. H. (1994). Comparative analysis of time spent grooming by birds in relation to parasite load. Behaviour, 131, 171–187.CrossRefGoogle Scholar
Cotton, M. J. (1970a). The reproductive biology of Ctenophthalmus nobilis (Rothschild) (Siphonaptera). Proceedings of the Royal Entomological Societyof London A, 45, 141–148.CrossRefGoogle Scholar
Cotton, M. J. (1970b). The life history of the hen flea, Ceratophyllus gallinae (Schrank) (Siphonaptera, Ceratophyllidae). Entomologist, 103, 45–48.Google Scholar
Cox, F. E. G. (2001). Concomitant infections, parasites and immune responses. Parasitology, 122, S23–S38.CrossRefGoogle ScholarPubMed
Cox, R., Stewart, P. D. & Macdonald, D. W. (1999). The ectoparasites of the European badger, Meles meles, and the behavior of the host-specific flea, Paraceras melis. Journal of Insect Behavior, 12, 245–265.CrossRefGoogle Scholar
Craw, R. C., Grehan, J. R. & Heads, M. J. (1999). Panbiogeography: Tracking the History of Life.New York: Oxford University Press.Google Scholar
Cresswell, J. E., Vidal-Martinez, V. M. & Crichton, N. J. (1995). The investigation of saturation in the species richness of communities: some comments on methodology. Oikos, 72, 301–304.CrossRefGoogle Scholar
Croll, N. A., Anderson, R. M., Gyerkos, T. W. & Ghardian, E. (1982). The population biology and control of Ascaris lumbricoides in a rural community in Iran. Transactions of the Royal Society of Tropical Medicine and Hygiene, 76, 187–197.CrossRefGoogle Scholar
Crooks, K. R., Scott, C. A., Angeloni, L., et al. (2001). Ectoparasites of the island fox on Santa Cruz Island. Journal of Wildlife Diseases, 37, 189–193.CrossRefGoogle Scholar
Crooks, K. R., Garcelon, D. K., Scott, C. A., et al. (2004). Ectoparasites of a threatened insular endemic mammalian carnivore: the island spotted skunk. American Midland Naturalist, 151, 35–41.CrossRefGoogle Scholar
Crowell, K. L. & Pimm, S. L. (1976). Competition and niche shifts introduced onto small islands. Oikos, 27, 251–258.CrossRefGoogle Scholar
Crum, G. E., Knapp, F. W. & White, G. M. (1974). Response of the cat flea, Ctenocephalides felis (Bouche), and the Oriental rat flea, Xenopsylla cheopis (Rothschild), to electromagnetic radiation in the 300–700 nanometer range. Journal of Medical Entomology, 11, 88–94.CrossRefGoogle Scholar
Cumming, G. S. & Bernard, R. T. (1997). Rainfall, food abundance and timing of parturition in African bats. Oecologia, 111, 309–317.CrossRefGoogle ScholarPubMed
Cyprich, D., Krumpál, M. & Rolníková, T. (1999). Occurrence and distribution of Ceratophyllus gallinae (Schrank, 1803) (Siphonaptera) in Slovakia. Folia Faunistica Slovaca, 4, 79–88 (in Slovak).Google Scholar
Cyprich, D., Pinowski, J. & Krumpál, M. (2002). Seasonal changes in numbers of fleas (Siphonaptera) in nests of the house sparrow (Passer domesticus) and tree sparrow (P. montanus) in Warsaw surroundings (Poland). Acta Parasitologica, 47, 58–65.Google Scholar
Cyprich, D., Štiavnická, L. & Kiefer, M. S. (2003). Information about occurrence and distribution specific flea species (Siphonaptera) of the ground squirrel (Spermophilus citellus L., 1758) in Slovakia: the genus Citellophilus and Neopsylla. Folia Faunistica Slovaca, 8, 47–51 (in Slovak).Google Scholar
Cyprich, D., Krumpál, M. & Duda, M. (2006). Characteristic of flea fauna (Siphonaptera) of ecological group of birds (Aves) nesting on the ground. Entomofauna Carpathica, 18, 1–4 (in Slovak).Google Scholar
Dallai, R., Lupetti, P., Afzelius, B. A. & Frati, F. (2003). Sperm structure of Mecoptera and Siphonaptera (Insecta) and the phylogenetic position of Boreus hyemalis. Zoomorphology, 122, 211–220.CrossRefGoogle Scholar
Dampf, A. (1911). Palaeopsylla klebsiana n. sp., eine fossiler Floh aus dem baltischen Bernstein. Schriften der Physikalisch–Ökonomischen Gesellschaft zu Königsberg, 51, 248–259.Google Scholar
Darby, C., Ananth, S. L., Tan, L. & Hinnebusch, B. J. (2005). Identification of gmhA, a Yersinia pestis gene required for flea blockage, by using a Caenorhabditis elegans biofilm system. Infection and Immunity, 73, 7236–7242.CrossRefGoogle ScholarPubMed
Darskaya, N. F. (1954). Fleas of the Daurian ground squirrel. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 12, 245–257 (in Russian).Google Scholar
Darskaya, N. F. (1955). Ecological characteristics of Xenopsylla gerbilli caspica. I. Fleas parasitic on the great gerbil, associated with the ecological features of their host. In Natural Focality of Human Diseases and Regional Epidemiology, ed. ,Anonymous. Moscow, USSR: Medgiz, pp. 400–407 (in Russian).Google Scholar
Darskaya, N. F. (1964). On the comparative ecology of fleas belonging to the genus Ceratophyllus. Ectoparasites, 4, 31–180 (in Russian).Google Scholar
Darskaya, N. F. (1970). Ecological comparisons of some fleas of the USSR fauna. Zoologicheskyi Zhurnal, 49, 729–745 (in Russian).Google Scholar
Darskaya, N. F. & Besedina, K. P. (1961). On the possibility of fleas feeding on reptiles. Proceeding of Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, 5, 33–39 (in Russian).Google Scholar
Darskaya, N. F. & Karandina, P. C. (1974). Observations on pre-imaginal development of Neopsylla setosa Wagn., 1898: fleas parasitic on ground squirrels. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Disieases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 134–137 (in Russian).Google Scholar
Darskaya, N. F., Brukhanova, L. V. & Kunitskaya, N. T. (1965). On the method of investigation of flea oviposition. In Problems of General Zoology and Medical Entomology, ed. ,Anonymous. Moscow, USSR: Moscow University Press, pp. 6–9.Google Scholar
Darskaya, N. F., Bragina, Z. S. & Petrov, V. G. (1970). On fleas of the common vole and shrews in dependence of sharp density fluctuations of these mammals. In Vectors of Particularly Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 132–152 (in Russian).Google Scholar
Davidson, D. W. (1980). Some consequences of diffuse competition in a desert ant community. American Naturalist, 116, 92–105.CrossRefGoogle Scholar
Davidson, D. W. (1985). An experimental study of diffuse competition in harvester ants. American Naturalist, 125, 500–506.CrossRefGoogle Scholar
Davis, D. E. (1951). Observations on rat ectoparasites and typhus fever in San Antonio, Texas. Public Health Reports, 66, 1717–1726.CrossRefGoogle ScholarPubMed
Davis, R. M., Smith, R. T., Madon, M. B. & Sitko-Cleugh, E. (2002). Flea, rodent and plague ecology at Chichupate Campground, Ventura County, California. Journal of Vector Ecology, 27, 107–127.Google ScholarPubMed
Dawson, L. H. J., Renaud, F., Guégan, J. F. & MeeÛs, T. (2000). Experimental evidence of asymmetrical competition between two species of parasitic copepods. Proceedings of the Royal Society of London B, 267, 1973–1978.CrossRefGoogle ScholarPubMed
Dawson, R. D. (2004). Does fresh vegetation protect avian nests from ectoparasites? An experiment with tree swallows. Canadian Journal of Zoology, 82, 1005–1010.
Dawson, R. D. & Bortolotti, G. R. (1997). Ecology of parasitism of nestling American kestrels by Carnus hemapterus (Diptera, Carnidae). Canadian Journal of Zoology, 75, 2021–2026.CrossRefGoogle Scholar
Day, J. F. & Benton, A. H. (1980). Population dynamics and coevolution of adult siphonapteran parasites of the southern flying squirrel (Glaucomys volans volans). American Midland Naturalist, 103, 333–338.CrossRefGoogle Scholar
Cardoso, Albuquerque V. & Linardi, P. M. (2006). Scanning electron microscopy studies of sensilla and other structures of the head of Polygenis (Polygenis) tripus (Siphonapera: Rhopalopsyllidae). Micron, 37, 557–565.CrossRefGoogle Scholar
Dean, S. R. & Meola, R. W. (1997). Effect of juvenile hormone and juvenile hormone mimics on sperm transfer from the testes of the male cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 34, 485–488.CrossRefGoogle Scholar
Dean, S. R. & Meola, R. W. (2002a). Factors influencing sperm transfer and insemination in cat fleas (Siphonaptera: Pulicidae) fed on an artificial membrane system. Journal of Medical Entomology, 39, 475–479.CrossRefGoogle Scholar
Dean, S. R. & Meola, R. W. (2002b). Effect of diet composition on weight gain, sperm transfer, and insemination in the cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 39, 370–375.CrossRefGoogle Scholar
Carvalho, R. W., Almeida, A. B., Barbosa-Silva, S. C., et al. (2003). The patterns of tungiasis in Araruama township, state of Rio de Janeiro, Brazil. Memórias do Instituto Oswaldo Cruz, 98, 31–36.CrossRefGoogle ScholarPubMed
Degen, A. A. (1997). Ecophysiology of Small Desert Mammals. New York: Springer-Verlag.CrossRefGoogle Scholar
Degen, A. A. (2006). Effect of macroparasites on the energy budget of small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 371–399.CrossRefGoogle Scholar
Degen, A. A., Kam, M., Khokhlova, I. S., Krasnov, B. R. & Barraclough, T. (1998). Average daily metabolic rate of rodents: habitat and dietary comparisons. Functional Ecology, 12, 63–73.CrossRefGoogle Scholar
Delahay, R. J., Speakman, J. R. & Moss, R. (1995). The energetic consequences of parasitism: effects of a developing infection of Trichostongylus tenius (Nematoda) of red grouse (Lagopus lagopus scoticus) energy balance, body weight and condition. Parasitology, 110, 473–482.CrossRefGoogle Scholar
Lope, F., M⊘ller, A. P. & Cruz, C. (1998). Parasitism, immune response and reproductive success in the house martin Delichon urbica. Oecologia, 114, 188–193.CrossRefGoogle Scholar
Demas, G. E. & Nelson, R. J. (1998). Photoperiod, ambient temperature, and food availability interact to affect reproductive and immune function in adult male deer mice (Peromyscus maniculatus). Journal of Biological Rhythms, 13, 253–262.CrossRefGoogle Scholar
Demin, E. P., Zagniborodova, E. N., Sageev, M. T., et al. (1970). Some features of ecology of fleas parasitic on Rhombomys opimus in western Turkmenistan in relation to their importance in the epizootology of plague. Problems of Particularly Dangerous Diseases, 11, 49–56 (in Russian).Google Scholar
Moraes, L. B., Bossi, D. E. P. & Linhares, A. X. (2003). Siphonaptera parasites of wild rodents and marsupials trapped in three mountain ranges of the Atlantic Forest in Southeastern Brazil. Memórias do Instituto Oswaldo Cruz, 98, 1071–1076.CrossRefGoogle ScholarPubMed
Hollander, N. & Allen, J. R. (1986). Cross-reactive antigens between a tick Dermacentor variabilis (Acari: Ixodidae) and a mite Prosoptes cuniculi (Acari: Psoroptidae). Journal of Medical Entomology, 23, 44–50.CrossRefGoogle Scholar
Deoras, P. J. & Prasad, R. S. (1967). Feeding mechanisms of Indian fleas X. cheopis (Roths.) and X. astia (Roths.). Indian Journal of Medical Research, 55, 1041–1050.Google Scholar
Pedro, N., Delgado, M. J., Gancedo, B. & Alonso-Bedate, M. (2003). Changes in glucose, glycogen, thyroid activity and hypothalamic catecholamines in tench by starvation and refeeding. Journal of Comparative Physiology B, 173, 475–481.CrossRefGoogle ScholarPubMed
Desdevises, Y., Jovelin, R., Jousson, O. & Morand, S. (2000). Comparison of ribosomal DNA sequences of Lamellodiscus spp. (Monogenea, Diplectanidae) parasitising Pagellus (Sparidae, Teleostei) in the north Mediterranean Sea: species divergence and coevolutionary interactions. International Journal for Parasitology, 30, 741–746.CrossRefGoogle ScholarPubMed
Desdevises, Y., Morand, S., Jousson, O. & Legendre, P. (2002a). Coevolution between Lamellodiscus (Monogenea: Diplectanidae) and Sparidae (Teleostei): the study of a complex host–parasite system. Evolution, 56, 2459–2471.CrossRefGoogle Scholar
Desdevises, I., Morand, S. & Legendre, P. (2002b). Evolution and determinants of host specificity in the genus Lamellodiscus (Monogenea). Biological Journal of the Linnean Society, 77, 431–443.CrossRefGoogle Scholar
Devi, S. G. & Prasad, R. S. (1985). Phagostimulants in artificial feeding systems of rat fleas Xenopsylla cheopis and Xenopsylla astia. Proceedings of the Indian National Science Academy B, 51, 566–573.Google Scholar
Dezfuli, B. S., Giari, L., Biaggi, S. & Poulin, R. (2001). Associations and interactions among intestinal helminths of the brown trout, Salmo trutta, in northern Italy. Journal of Helminthology, 75, 331–336.Google ScholarPubMed
Dick, C. W. & Patterson, B. D. (2006). Bat flies: obligate parasites of bats. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 179–194.CrossRefGoogle Scholar
Dick, C. W., Gannon, M. R., Little, W. E. & Patrick, M. J. (2003). Ectoparasite associations of bats from central Pennsylvania. Journal of Medical Entomology, 40, 813–819.CrossRefGoogle ScholarPubMed
Diamond, J. M. (1975). Assembly of species communities: chance or competition? In Ecology and Evolution of Communities, ed. Cody, M. L. & Diamond, J. M.. Cambridge, MA: Harvard University Press, pp. 342–444.Google Scholar
Cruz, Dittmar K. & Whiting, M. (2003). Genetic and phylogeographic structure of populations of Pulex simulans (Siphonaptera) in Peru inferred from two genes (CytB and CoII). Parasitology Research, 91, 55–59.CrossRefGoogle Scholar
Cruz, Dittmar K., Mamat, U., Whiting, M., et al. (2003). Techniques of DNA-studies on prehispanic ectoparasites (Pulex sp., Pulicidae, Siphonaptera) from animal mummies of the Chiribaya Culture, Southern Peru. Memórias do Instituto Oswaldo Cruz, 98, 53–58.Google Scholar
Dobson, A. P. (1985). The population dynamics of competition between parasites. Parasitology, 91, 317–347.
Dobson, A. P. (1990). Models for multi-species parasite–host communities. In Parasite Communities: Patterns and Processes, ed. Esch, G., Bush, A. O. & Aho, J. M.. London: Chapman & Hall, pp. 261–288.Google Scholar
Dobson, A. P. & Roberts, M. (1994). The population dynamics of parasitic helminth communities. Parasitology, 109, S97–S108.
Dogiel, V. A., Petrushevski, G. K. & Polyanski, Y. I. (1961). Parasitology of Fishes. Edinburgh, UK: Oliver & Boyd.Google Scholar
Dong, B. (1991). Experimental studies on the transmission of hemorrhagic fever with renal syndrome virus by gamasid mites and fleas. Chinese Medical Journal, 71, 502–504 (in Chinese).Google Scholar
Dowling, A. P. G. (2006). Mesostigmatid mites as parasites of small mammals: systematics, ecology, and the evolution of parasitic associations. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 103–117.CrossRefGoogle Scholar
Downing, J. A. (1986). Spatial heterogeneity: evolved behaviour or mathematical artefact. Nature, 323, 255–257.CrossRefGoogle Scholar
Drobney, R. D., Train, C. T. & Gredrickson, L. H. (1983). Dynamics of the platyhelminth fauna of wood ducks in relation to food habits and reproductive state. Journal of Parasitology, 69, 375–380.CrossRefGoogle ScholarPubMed
Dryden, M. W. (1989). Host association, on-host longevity and egg production of Ctenocephalides felis felis. Veterinary Parasitology, 34, 117–122.CrossRefGoogle ScholarPubMed
Dryden, M. W. & Blakemore, J. (1989). A review of flea allergy dermatitis in the dog and cat. Companion Animal Practice, 19, 10–16.Google Scholar
Dryden, M. W. & Broce, A. B. (1993). Development of a trap for collecting newly emerged Ctenocephalides felis (Siphonaptera: Pulicidae) in homes. Journal of Medical Entomology, 30, 901–906.CrossRefGoogle Scholar
Dubinina, V. B. & Dubinin, M. N. (1951). Parasite fauna of mammals of the Dauric Steppe. Fauna and Ecology of Rodents, 4, 98–156 (in Russian).Google Scholar
Dubyansky, M. A., Dubyanskaya. L. D., Zhubanazarov, I. Z. & Filipchenko, V. E. (1975). Alternative method for prognosis of abundance of Xenopsylla fleas. In Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 93–96 (in Russian).Google Scholar
Duffy, J. E., Richardson, J. P. & France, K. E. (2005). Ecosystem consequences of diversity depend on food chain length in estuarine vegetation. Ecology Letters, 8, 301–309.CrossRefGoogle Scholar
Dufva, R, & Allander, K. (1996). Variable effects of the hen flea Ceratophyllus gallinae on the breeding success of the great tit Parus major in relation to weather conditions. Ibis, 138, 772–777.CrossRefGoogle Scholar
Dunnet, G. M. & Mardon, D. K. (1974). A monograph of Australian fleas (Siphonaptera). Australian Journal of Zoology (Supplemental Series), 30, 1–273.Google Scholar
Durden, L. A. (1995). Fleas (Siphonaptera) of cotton mice on a Georgia Barrier Island: a depauperate fauna. Journal of Parasitology, 81, 526–529.CrossRefGoogle ScholarPubMed
Durden, L. A. & Beaucournu, J. C. (2002). Gymnomeropsylla n. gen. (Siphonaptera: Pygiopsyllidae) from Sulawesi, Indonesia, with the description of two new species. Parasite, 9, 225–232.CrossRefGoogle ScholarPubMed
Durden, L. A. & Kollars, T. M. (1997). The fleas (Siphonaptera) of Tennessee. Journal of Vector Ecology, 22, 13–22.Google ScholarPubMed
Durden, L. A., Ellis, B. A., Banks, C. W., Crowe, J. D. & Oliver, J. H. (2004). Ectoparasites of gray squirrels in two different habitats and screening of selected ectoparasites for bartonellae. Journal of Parasitology, 90, 485–489.CrossRefGoogle ScholarPubMed
Durden, L. A., Judy, T. N., Martin, J. E. & Spedding, L. S. (2005). Fleas parasitizing domestic dogs in Georgia, USA: species composition and seasonal abundance. Veterinary Parasitology, 130, 157–162.CrossRefGoogle ScholarPubMed
Durden, L. A., Cunningham, M. W., McBride, R. & Ferree, B. (2006). Ectoparasites of free-ranging pumas and jaguars in the Paraguayan Chaco. Veterinary Parasitology, 137, 189–193.CrossRefGoogle ScholarPubMed
Dynesius, M. & Jansson, R. (2000). Evolutionary consequences of changes in species' geographical distributions driven by Milankovitch climate oscillations. Proceedings of the National Academy of Sciences of the USA, 97, 9115–9120.CrossRefGoogle ScholarPubMed
Ebert, D. (1994). Virulence and local adaptation of a horizontally transmitted parasite. Science, 265, 1084–1086.CrossRefGoogle ScholarPubMed
Eckstein, R. A. & Hart, B. L. (2000a). The organization and control of grooming in cats. Applied Animal Behaviour Science, 68, 131–140.CrossRefGoogle Scholar
Eckstein, R. A. & Hart, B. L. (2000b). Grooming and control of fleas in cats. Applied Animal Behaviour Science, 68, 141–150.CrossRefGoogle Scholar
Edney, E. B. (1945). Laboratory studies on the bionomics of the rat fleas, Xenopsylla brasiliensis Baker and X. cheopis Roths. I. Certain effects of light, temperature and humidity on the rate of development and on adult longevity. Bulletin of Entomological Research, 35, 399–416.Google Scholar
Edney, E. B. (1947a). Laboratory studies on the bionomics of the rat fleas, Xenopsylla brasiliensis Baker and X. cheopis Rothsch. II. Water relations during the cocoon period. Bulletin of Entomological Research, 38, 263–280.CrossRefGoogle Scholar
Edney, E. B. (1947b). Laboratory studies on the bionomics of the rat fleas, Xenopsylla brasiliensis Baker and X. cheopis Roths. III. Further factors affecting adult longevity. Bulletin of Entomological Research, 38, 389–404.CrossRefGoogle Scholar
Eeley, H. A. C. & Foley, R. A. (1999). Species richness, species range size and ecological specialization among African primates: geographical patterns and conservation implications. Biodiversity and Conservation, 8, 1033–1056.CrossRefGoogle Scholar
Eeva, T., Lehikoinen, E. & Nurmi, J. (1994). Effects of ectoparasites on breeding success of great tits (Parus major) and pied flycatchers (Ficedula hypoleuca) in an air-pollution gradient. Canadian Journal of Zoology, 72, 624–635.CrossRefGoogle Scholar
Egoscue, H. J. (1976). Flea exchange between deer mice and some associated small mammals in western Utah. Great Basin Naturalist, 36, 475–480.Google Scholar
Eisele, M., Heukelbach, J., Marck, E., et al. (2003). Investigations on the biology, epidemiology, pathology and control of Tunga penetrans in Brazil. I. Natural history of tungiasis in man. Parasitology Research, 90, 87–99.Google Scholar
Eiseman, C. H. & Binnengton, K. C. (1994). The peritrophic membrane: its formation, structure, chemical composition and permeability in relation to vaccination against ectoparasitic arthropods. International Journal for Parasitology, 24, 15–26.CrossRefGoogle Scholar
Elliot, J. M. (1977). Some Methods for Statistical Analysis of Samples of Benthic Invertebrates, 2nd edn, Freshwater Biological Association Sceintific Publications No. 25. Ambleside, UK: Titus Wilson & Son.Google Scholar
Ellis, B. A., Regnery, R. L., Beati, L., et al. (1999). Rats of the genus Rattus are reservoir hosts for pathogenic Bartonella species: an Old World origin for a New World disease. Journal of Infectious Diseases, 180, 220–224.CrossRefGoogle ScholarPubMed
Elmes, G. W., Barr, B., Thomas, J. A. & Clarke, R. T. (1999). Extreme host specificity by Microdon mutabilis (Diptera: Syrphidae), a social parasite of ants. Proceedings of the Royal Society of London B, 266, 447–453.CrossRefGoogle Scholar
Elshanskaya, N. I. & Popov, M. N. (1972). Zoologico-parasitological characteristics of the River Kenkeme valley (Central Yakutia). In Theriology, vol. 1, ed. Kolosova, L. D. & Lukyanova, I. V.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 368–372 (in Russian).Google Scholar
Emelianova, N. D. & Shtilmark, F. R. (1967). Fleas of insectivores, rodents and lagomorphs of the central part of Western Sayan. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 27, 241–253 (in Russian).Google Scholar
Engelthaler, D. M., Hinnebusch, B. J., Rittner, C. M. & Gage, K. L. (2000). Quantitative competitive PCR as a technique for exploring flea–Yersina pestis dynamics. American Journal of Tropical Medicine and Hygiene, 62, 552–560.CrossRefGoogle ScholarPubMed
Enquist, B. J., Haskell, J. P. & Tiffney, B. H. (2002). General patterns of taxonomic and biomass partitioning in extant and fossil plant communities. Nature, 419, 610–613.CrossRefGoogle ScholarPubMed
Esbérard, C. (2001). Infestation of Rhynchopsyllus pulex (Siphonaptera: Tungidae) on Molossus molossus (Chiroptera) in Southeastern Brazil. Memórias do Instituto Oswaldo Cruz, 96, 1169–1170.CrossRefGoogle Scholar
Euzet, L. & Combes, C. (1980). Les problèmes de l'espèce chez les animaux parasites. Bulletin de la Société Zoologique de France, 40, 239–285.Google Scholar
Evans, F. G. & Freeman, R. B. (1950). On the relationship of some mammal fleas to their hosts. Annals of the Entomological Society of America, 43, 320–333.CrossRefGoogle Scholar
Ezenwa, V. O. (2004). Host social behavior and parasitic infection: a multifactorial approach. Behavioral Ecology, 15, 446–454.CrossRefGoogle Scholar
Fain, A. & Hyland, K. W. (1985). Evolution of astigmatid mites on mammals. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 641–658.Google Scholar
Fairbairn, D. J. (1997). Allometry for sexual size dimorphism: pattern and process in the coevolution of body size in males and females. Annual Review of Ecology and Systematics, 28, 659–687.CrossRefGoogle Scholar
Fairbairn, D. J. (2005). Allometry for sexual size dimorphism: testing two hypotheses for Rensch's rule in the water strider Aquarius remigis. American Naturalist, 166, S69–S84.Google ScholarPubMed
Farhang-Azad, A., Traub, R. & Wisseman, C. L. (1983). Rickettsia mooseri infection in the fleas Leptopsylla segnis and Xenopsylla cheopis. American Journal of Tropical Medicine and Hygiene, 32, 1392–1400.CrossRefGoogle ScholarPubMed
Farhang-Azad, A., Traub, R., Sofi, M. & Wisseman, C. L. (1984). Experimental murine typhus infection in the cat fleas, Ctenocephalides felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 21, 675–680.CrossRefGoogle Scholar
Farhang-Azad, A., Traub, R. & Boqar, S. (1985). Transovarial transmission of murine typhus rickettsia in Xenopsylla cheopis fleas. Science, 227, 543–545.CrossRefGoogle ScholarPubMed
Farhang-Azad, A., Radulovic, S., Higgins, J. A., Noden, B. H. & Troyer, J. M. (1997). Flea-borne rickettsioses: ecologic considerations. Emerging Infectious Diseases, 3, 319–327.CrossRefGoogle Scholar
Faulkenberry, G. D. & Robbins, R. G. (1980). Statistical measures of interspecific association between the fleas of the gray-tailed vole, Microtus canicaudus Miller. Entomological News, 91, 93–101.Google Scholar
Faust, E. C. & Maxwell, T. A. (1930). The finding of the larva of the chigo, Tunga penetrans, in scrapings from human skin. Archives of Dermatology and Syphilology, 22, 94–97.CrossRefGoogle Scholar
Fauth, P. T., Krementz, D. G. & Hines, J. E. (1991). Ectoparasitism and the role of green nesting material in the European starling. Oecologia, 88, 22–29.CrossRefGoogle ScholarPubMed
Fedorov, Y. V., Igolkin, N. I. & Tyushnikova, M. K. (1959). Some data on virus-carrying fleas in areas of tick-borne encephalitis and lymphocytic choriomenengitis. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 28, 149–152 (in Russian).Google Scholar
Feingold, B. F., Benjamini, E. & Michaeli, D. (1968). The allergic responses to insect bites. Annual Review of Entomology, 13, 138–158.CrossRefGoogle Scholar
Feliu, C., Renaud, F., Catzeflis, F., et al. (1997). Comparative analysis of parasite species richness of Iberian rodents. Parasitology, 115, 453–466.CrossRefGoogle ScholarPubMed
Fellis, K. J., Negovetich, N. J., Esch, G. W., Horak, I. G. & Boomker, J. (2003). Patterns of association, nestedness, and species co-occurrence of helminth parasites in the greater kudu, Tragelaphus strepsiceros, in the Kruger National Park, South Africa, and the Etosha National Park, Namibia. Journal of Parasitology, 89, 899–907.CrossRefGoogle ScholarPubMed
Felsenstein, J. (1985). Phylogenies and the comparative method. American Naturalist, 125, 1–15.CrossRefGoogle Scholar
Feng, X.-Y., Ni, E.-J., Cao, P.-G., Zheng, C.-J. & Yang, G.-H. (2004). Investigation of the distributive features of flea population during epidemic outbreak of plague in Longlin County, Guangxi. Acta Parasitologica et Medica Entomologica Sinica, 11, 235–237 (in Chinese).Google Scholar
Feng, Y.-M., Li, W. & Wang, Z.-Y. (2003). Observation on the life cycle of Ischnopsyllus octactenus at laboratory. Chinese Journal of Vector Biology and Control, 14, 202–203 (in Chinese).Google Scholar
Fenner, F. & Ratcliff, F. N. (1965). Myxomatosis. Cambridge, UK: Cambridge University Press.Google Scholar
Feoktistov, A. Z., Vasiliev, G. I. & Kraminsky, N. N. (1968). Passage of the virus of tick-borne encephalitis via fleas Xenopsylla cheopis Roths. In Problems of Epidemiology and Epizootology of Paricularly Dangerous Diseases, ed. Anonymous. Kyzyl, USSR: The Tuva Anti-Plague Station, pp. 317–320.Google Scholar
Ferkin, M. H., Sorokin, E. S. & Johnston, R. E. (1996). Self-grooming as a sexually dimorphic communicative behaviour in meadow voles, Microtus pennsylvanicus. Animal Behaviour, 51, 801–810.CrossRefGoogle Scholar
Ferrari, N., Cattadori, I. M., Nespereira, J., Rizzoli, A. & Hudson, P. J. (2004). The role of host sex in parasite dynamics: field experiments on the yellow-necked mouse Apodemus flavicollis. Ecology Letters, 7, 88–94.CrossRefGoogle Scholar
Ferrari, N., Rosá, R., Pugliese, A. & Hudson, P. J. (2007). The role of sex in parasite dynamics: model simulations on transmission of Heligmosomoides polygyrus in populations of yellow-necked mice, Apodemus flavicollis. International Journal for Parasitology, 37, 341–349.CrossRefGoogle ScholarPubMed
Fichet-Calvet, E., Jomaa, I., Ismail, Ben R. & Ashford, R. W. (2000). Pattern of infection of haemoparasites in the fat sand rat, Psammomys obesus, in Tunisia and effect on the host. Annals of Tropical Medicine and Parasitology, 94, 55–68.CrossRefGoogle Scholar
Fieberg, J. & Ellner, S. P. (2000). When is it meaningful to estimate an extinction probability?Ecology, 81, 2040–2047.CrossRefGoogle Scholar
Fielden, L. J., Rechav, Y. & Bryson, N. R. (1992). Acquired immunity to larvae of Amblyomma marmoreum and A. hebraeum by tortoises, guinea-pigs and guinea-fowl. Medical and Veterinary Entomology, 6, 251–254.CrossRefGoogle Scholar
Fielden, L. J., Jones, R. M., Goldberg, M. & Rechav, Y. (1999). Feeding and respiratory gas exchange in the American dog tick, Dermacentor variabilis. Journal of Insect Physiology, 45, 297–304.CrossRefGoogle ScholarPubMed
Fielden, L. J., Krasnov, B. R. & Khokhlova, I. S. (2001). Respiratory gas exchange in the flea Xenopsylla conformis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 38, 735–739.CrossRefGoogle Scholar
Fielden, L. J., Krasnov, B. R., Still, K. & Khokhlova, I. S. (2002). Water balance in two species of desert fleas, Xenopsylla ramesis and X. conformis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 39, 875–881.CrossRefGoogle Scholar
Fielden, L. J., Krasnov, B. R., Khokhlova, I. S. & Arakelyan, M. S. (2004). Respiratory gas exchange in the desert flea Xenopsylla ramesis (Siphonaptera: Pulicidae): response to temperature and blood-feeding. Comparative Biochemistry and Physiology A, 137, 557–565.CrossRefGoogle ScholarPubMed
Filimonova, S. A. (1986). Changes in the ultra-structure of the intestinal epithelium of Xenopsylla cheopis (Siphonaptera) after emerging from cocoons and beginning of feeding. Parazitologiya, 20, 99–105 (in Russian).Google Scholar
Filimonova, S. A. (1989). A morphologic analysis of digestion in Leptopsylla segnis (Siphonaptera: Leptopsyllidae) fleas. Parazitologiya, 23, 480–488 (in Russian).Google ScholarPubMed
Fiorello, C. V., Robbins, R. G., Maffei, L. & Wade, S. E. (2006). Parasites of free-ranging small canids and felids in the Bolivian Chaco. Journal of Zoo and Wildlife Medicine, 37, 130–134.CrossRefGoogle ScholarPubMed
Fisher, R. A. (1930). The Genetical Theory of Natural Selection. Oxford, UK: Oxford University Press.CrossRefGoogle Scholar
Fitze, P. S. & Richner, H. (2002). Differential effects of a parasite on ornamental structures based on melanins and carotenoids. Behavioral Ecology, 13, 401–407.CrossRefGoogle Scholar
Fitze, P. S., Clobert, J. & Richner, H. (2004a). Long-term life-history consequences of ectoparasite-modulated growth and development. Ecology, 85, 2018–2026.CrossRefGoogle Scholar
Fitze, P. S., Tschirren, B. & Richner, H. (2004b). Life history and fitness consequences of ectoparasites. Journal of Animal Ecology, 73, 216–226.CrossRefGoogle Scholar
Fleming, T. H., Breitwisch, R. L. & Whitesides, G. W. (1987). Patterns of tropical vertebrate frugivore diversity. Annual Review of Ecology and Systematics, 18, 91–109.CrossRefGoogle Scholar
Flux, J. E. C. (1972). Seasonal and regional abundance of fleas on hares in Kenya. Journal of East African Natural History Society, 29, 1–8.Google Scholar
Folstad, I. & Karter, A. J. (1992). Parasites, bright males, and the immunocompetence handicap. American Naturalist, 139, 603–622.CrossRefGoogle Scholar
Forbes, M. R., Alisauskas, R. T., McLaughlin, J. D. & Cuddington, K. M. (1999). Explaining co-occurrence among helminth species of lesser snow geese (Chen caerulescens) during their winter and spring migration. Oecologia, 120, 613–620.CrossRefGoogle Scholar
Foster, W. A. & Olkowski, W. (1968). Natural invasion of artificial cliff swallow nests by Oeciacus vicarius (Hemiptera: Cimicidae) and Ceratophyllus petrochelidoni (Siphonaptera: Ceratophyllidae). Journal of Medical Entomology, 5, 488–491.CrossRefGoogle Scholar
Fowler, J. A., Cohen, S. & Greenwood, M. T. (1983). Seasonal variation in the infestation of blackbirds by fleas. Bird Study, 30, 240–242.CrossRefGoogle Scholar
Fox, B. J. & Brown, J. H. (1993). Assembly rules for the functional groups in North American desert rodent communities. Oikos, 67, 358–370.CrossRefGoogle Scholar
Fox, B. J. & Luo, J. (1996). Estimating competition coefficients from census data: a re-examination of the regression technique. Oikos, 77, 291–300.CrossRefGoogle Scholar
Fox, I., Fox, R. I. & Bayona, I. G. (1966). Fleas feed on the lizards in the laboratory in Puerto Rico. Journal of Medical Entomology, 2, 395–396.CrossRefGoogle Scholar
Fox, J. W., McGrady-Steed, J. & Petchey, O. L. (2000). Testing for local species saturation with nonindependent regional species pools. Ecology Letters, 3, 198–206.CrossRefGoogle Scholar
Fox, L. R. (1975). Cannibalism in natural populations. Annual Review of Ecology and Systematics, 6, 87–106.CrossRefGoogle Scholar
Fox, L. R. & Morrow, P. A. (1981). Specialization: species property or local phenomenon?Science, 211, 887–893.CrossRefGoogle ScholarPubMed
Franc, M., Choquart, P. & Cadiergues, M. C. (1998). Species of fleas found on dogs in France. Revue de Médecine Vétérinaire, 149, 135–140.Google Scholar
Freeman, R. B. & Madsen, H. (1949). A parasitic flea larva. Nature, 164, 187–188.CrossRefGoogle ScholarPubMed
Fretwell, S. D. & Lucas, H. L. (1970). On territorial behavior and other factors influencing habitat distribution in birds. I. Theoretical development. Acta Biotheoretica, 19, 16–36.CrossRefGoogle Scholar
Frigessi, A., Holden, M., Marshall, C., et al. (2005). A Bayesian model for the population dynamics of two interacting species, with application to great gerbils and fleas in south-eastern Kazakhstan. Biometrics, 61, 231–239.CrossRefGoogle Scholar
Fry, J. D. (1996). The evolution of host specialization: are trade-offs overrated? American Naturalist, 148, S84–S107.
Fuller, G. K. (1974). Observations on flea attachment at low hair densities on man. Journal of Natural History, 8, 207–213.CrossRefGoogle Scholar
Futuyma, D. J. & Moreno, G. (1988). The evolution of ecological specialization. Annual Review of Ecology and Systematics, 19, 207–233.CrossRefGoogle Scholar
Futuyma, D. J. & Slatkin, M. (1983). Coevolution. Sunderland, MA: Sinauer Associates.Google Scholar
Gabbutt, P. D. (1961). The distribution of some small mammals and their associated fleas from central Labrador. Ecology, 42, 518–525.CrossRefGoogle Scholar
Gäde, G. (2002). Sexual dimorphism in the pyrgomorphid grasshopper Phymateus morbillosus: from wing morphometry and flight behaviour to flight physiology and endocrinology. Physiological Entomology, 27, 51–57.CrossRefGoogle Scholar
Gage, K. L. & Kosoy, M. Y. (2005). Natural history of plague: perspectives from more than a century of research. Annual Review of Entomology, 50, 505–528.CrossRefGoogle Scholar
Gage, K. L., Ostfeld, R. S. & Olson, J. G. (1995). Nonviral vector-borne zoonoses associated with mammals in the United States. Journal of Mammalogy, 76, 695–715.CrossRefGoogle Scholar
Galaktionov, K. V. (1996). Life cycles and distribution of seabird helminths in Arctic and subArctic regions. Bulletin of the Scandinavian Society for Parasitology, 6, 31–49.Google Scholar
Galbe, J. & Oliver, J. H. (1992). Immune response of lizards and rodents to larval Ixodes scapularis (Acari, Ixodidae). Journal of Medical Entomology, 29, 774–783.CrossRefGoogle Scholar
Gallivan, G. J. & Horak, I. G. (1997). Body size and habitat as determinants of tick infestations of wild ungulates in South Africa. South African Journal of Wildlife Research, 27, 63–70.Google Scholar
Galun, R. (1975). Research into alternative arthropod control measures against livestock pests (part 1). In Workshop on the Ecology and Control of External Parasites of Economic Importance on Bovines in Latin America, ed. Thompson, K. C.. Cali, Colombia: Centro Internacional de Agricultura Tropical (CIAT), pp. 155–161.Google Scholar
Garland, T., Harvey, P. H. & Ives, A. R. (1992). Procedures for the analysis of comparative data using phylogenetically independent contrasts. American Naturalist, 41, 18–32.Google Scholar
Garland, T., Dickerman, A. W. C., Janis, M. & Jones, J. A. (1993). Phylogenetic analysis of covariance by computer simulation. Systematic Biology, 42, 265–292.CrossRefGoogle Scholar
Gaston, K. J. (1996). Spatial covariance in the species richness of higher taxa. In Aspects of the Genesis and Maintenance of Biological Diversity, ed. Hochberg, M. E., Clobert, J. & Barbault, R.. Oxford, UK: Oxford University Press, pp. 221–242.Google Scholar
Gaston, K. J. (2003). The Structure and Dynamics of Geographic Ranges. Oxford, UK: Oxford University Press.Google Scholar
Gaston, K. J. & Blackburn, T. M. (2000). Pattern and Process in Macroecology. Oxford, UK: Blackwell Science.CrossRefGoogle Scholar
Gaston, K. J. & Blackburn, T. M. (2003). Dispersal and the interspecific abundance–occupancy relationship in British birds. Global Ecology and Biogeography, 12, 373–379.CrossRefGoogle Scholar
Gaston, K. J., Blackburn, T. M. & Loder, N. (1995). Which species are described first? The case of North American butterflies. Biodiversity and Conservation, 4, 119–127.CrossRefGoogle Scholar
Gaston, K. J., Blackburn, T. M. & Lawton, J. H. (1997). Interspecific abundance–range size relationships: an appraisal of mechanisms. Journal of Animal Ecology, 66, 579–601.CrossRefGoogle Scholar
Gauzshtein, D. M., Kunitsky, V. N., Kunitskaya, N. T. & Filimonov, V. I. (1965). On the time spent on the host body in fleas parasitic on the great gerbil. In Proceedings of the 4th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute and Kainar, pp. 66–68 (in Russian).Google Scholar
Gauzshtein, D. M., Kunitsky, V. N., Gubaidullina, V. S., et al. (1967). On the phenology of reproduction in some fleas parasitic on the great gerbil in the southern Balkhash Region. In Proceedings of the 5th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 160–163 (in Russian).Google Scholar
Geigy, R. & Herbig, A. (1949). Die Hypertrophie der Organe beim Weibchen von Tunga penetrans. Acta Tropica, 6, 246–262.Google Scholar
Gerasimova, N. G. (1970). Metamorphosis of fleas Xenopsylla nuttalli Ioff, 1930. In Vectors of Dangerous Diseases and Their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 316–322 (in Russian).Google Scholar
Gerasimova, N. G. (1973). Some reproductive parameters in Xenopsylla skrjabini and X. nuttalli. Problems of Particularly Dangerous Diseases, 29, 117–121 (in Russian).Google Scholar
Gerasimova, N. G., Denisova, N. G., Denisov, P. S., Knyazeva, T. V. & Lavrovsky, A. A. (1977). Species composition and population dynamics of fleas on the pygmy ground squirrel in the stable foci of plague in the Ergeni Upland. Parazitologiya, 11, 446–452 (in Russian).Google Scholar
Gillespie, R. D., Mbow, M. L. & Titus, R. G. (2000). The immunomodulatory factors of bloodfeeding arthropod saliva. Parasite Immunology, 22, 319–331.CrossRefGoogle ScholarPubMed
Gillespie, S. H., Smith, G. L. & Osbourn, A. (2004). Microbe–Vector Interactions in Vector-Borne Diseases. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Giorgi, M. S., Arlettaz, R., Christe, P. & Vogel, P. (2001). The energetic grooming costs imposed by a parasitic mite (Spinturnix myoti) upon its bat host (Myotis myotis). Proceedings of the Royal Society of London B, 268, 2071–2075.CrossRefGoogle Scholar
Gliwicz, J. (1992). Patterns of dispersal in non-cyclic populations of small rodents. In Animal Dispersal: Small Mammals as a Model, ed. Stenseth, N. C. & Lidicker, W. Z.. London: Chapman & Hall, pp. 147–159.CrossRefGoogle Scholar
Goater, C. P. & Ward, P. I. (1992). Negative effects of Rhabdias bufonis (Nematoda) on the growth and survival of toads (Bufo bufo). Oecologia, 89, 161–165.CrossRefGoogle Scholar
Gobel, E. & Krampitz, H. E. (1982). Histologische Untersuchungen zur Gamogonie und Sporogonie von Hepatozoon erhardovae in experimentell infizierten Rattenflöhen (Xenopsylla cheopis). Zeitschrift für Parasitenkunde, 67, 261–271.CrossRefGoogle Scholar
Gong, Y.-L., Li, Z.-L. & Ma, L.-M. (2004). Further research of the bloodsucking activities of the flea Citellophilus tesquorum sungaris. Acta Parasitologica et Medica Entomologica Sinica, 11, 47–49 (in Chinese).Google Scholar
Gong, Z.-D., Xie, B.-Q. & Ling, J.-B. (1996). Ecology and fauna of fleas on Mt. Gaoligong of Yunnan. Zoological Research, 17, 59–67 (in Chinese).Google Scholar
Gong, Z.-D., Duan, X.-D., Feng, X.-G., Wu, X.-Y. & Liu, Q. (1999). The fauna and ecology of fleas in Cangshan Mountain and Erhai Lake Nature Reserve, Dali. Zoological Research, 20, 451–456 (in Chinese).Google Scholar
Gong, Z.-D., Wu, H.-Y., Duan, X.-D., Feng, X.-G. & Yang, G.-R. (2000). Fauna and community ecology of fleas in Lincang region, Yunnan province. Acta Parasitologica et Medica Entomologica Sinica, 7, 160–169 (in Chinese).Google Scholar
Gong, Z.-D., Zheng, D., Wu, H.-Y, et al. (2001). The relationship between the geographical distribution trends of flea species diversity and the important environmental factor in the Hengduan Mountains, Yunnan. Biodiversity Science, 9, 319–328 (in Chinese).Google Scholar
Gong, Z.-D., Wu, H.-Y., Duan, X.-D., et al. (2004). Vertical distribution pattern and fauna characteristics of flea communities in the Mt. Wuliang Nature Reserve, Jingdong, Yunnan. Chinese Journal of Vector Biology and Control, 15, 344–348 (in Chinese).Google Scholar
Gong, Z.-D., Wu, H.-Y., Duan, X.-D., et al. (2005). Species richness and vertical distribution pattern of flea fauna in Hengduan Mountains of western Yunnan, China. Biodiversity Science, 13, 279–289 (in Chinese).CrossRefGoogle Scholar
Gong, Z.-D., Zhang, L.-Y., Duan, X.-D., et al. (2007). Species richness and fauna of fleas along a latitudinal gradient in the Three Parallel Rivers landscape, China. Biodiversity Science, 15, 61–69 (in Chinese).Google Scholar
González, M. T. & Poulin, R. (2005). Spatial and temporal predictability of the parasite community structure of a benthic marine fish along its distributional range. International Journal for Parasitology, 35, 1369–1377.CrossRefGoogle ScholarPubMed
Gotelli, N. J. (2000). Null model analysis of species co-occurrence patterns. Ecology, 81, 2606–2621.CrossRefGoogle Scholar
Gotelli, N. J. & Arnett, A. E. (2000). Biogeographic effects of red fire ant invasion. Ecology Letters, 3, 257–261.CrossRefGoogle Scholar
Gotelli, N. J. & Entsminger, G. L. (2001). Swap and fill algorithms in null model analysis: rethinking the Knight's Tour. Oecologia, 129, 281–291.CrossRefGoogle ScholarPubMed
Gotelli, N. J. & Entsminger, G. L. (2006). EcoSim: Null Models Software for Ecology. Version 7. Jericho, VT: Acquired Intelligence Inc. & Kesey-Bear. Available online at http://garyentsminger.com/ecosim.htm.Google Scholar
Gotelli, N. J. & Graves, G. R. (1996). Null Models in Ecology. Washington, DC: Smithsonian Institution Press.Google Scholar
Gotelli, N. J. & McCabe, D. J. (2002). Species co-occurrence: a meta-analysis of J. M. Diamond's assembly rules model. Ecology, 83, 2091–2096.CrossRefGoogle Scholar
Gotelli, N. J. & Rohde, K. (2002). Co-occurrence of ectoparasites of marine fishes: a null model analysis. Ecology Letters, 5, 86–94.CrossRefGoogle Scholar
Goüy de Bellocq, J., Sarà, M., Casanova, J. C., Feliu, C. & Morand, S. (2003). A comparison of the structure of helminth communities in the woodmouse, Apodemus sylvaticus, on islands of the western Mediterranean and continental Europe. Parasitology Research, 90, 64–70.Google Scholar
Goüy de Bellocq, J., Krasnov, B. R., Khokhlova, I. S., Ghazaryan, L. & Pinshow, B. (2006a). Immunocompetence and flea parasitism in a desert rodent. Functional Ecology, 20, 637–646.CrossRefGoogle Scholar
Goüy de Bellocq, J., Krasnov, B. R., Khokhlova, I. S. & Pinshow, B. (2006b). Temporal dynamics of a T-cell mediated immune response in desert rodents. Comparative Biochemistry and Physiology A, 145, 554–559.CrossRefGoogle Scholar
Gracia, M. J., Lucientes, J., Castillo, J. A., et al. (2000). Pulex irritans infestation in dogs. Veterinary Record, 147, 748–749.Google ScholarPubMed
Grafen, A. (1989). The phylogenetic regression. Philosophical Transactions of the Royal Society of London B, 326, 119–157.CrossRefGoogle ScholarPubMed
Gray, C. A., Gray, P. N. & Pence, D. B. (1989). Influence of social status on the helminth community of late-winter mallards. Canadian Journal of Zoology, 67, 1937–1944.CrossRefGoogle Scholar
Grazhdanov, A. K., Bidashko, F. G., Tanitovky, V. A., et al. (2002). Comparative fecundity of fleas parasitic on Meriones gerbils in the laboratory. Quarantinable and Zoonotic Infections in Kazakhstan, 6, 39–43 (in Russian).Google Scholar
Grebenyuk, R. V. (1951). Sheep Vermipsylleses and their Control. Frunze, USSR: Ylym (in Russian).Google Scholar
Greene, W. K., Carnegie, R. L., Shaw, S. E., Thompson, R. C. A. & Penhale, W. J. (1993). Characterization of allergens of the cat flea, Ctenocephalides felis: detection and frequency of IgE antibodies in canine sera. Parasite Immunology, 15, 69–74.CrossRefGoogle ScholarPubMed
Greenwood, M. T., Clark, F. & Smith, J. S. (1991). Automatic recording of flea activity. Medical and Veterinary Entomology, 5, 93–100.CrossRefGoogle ScholarPubMed
Gregory, R. D., Montgomery, S. S. J. & Montgomery, W. I. (1992). Population biology of Heligmosomoides polygyrus (Nematoda) in the wood mouse. Journal of Animal Ecology, 61, 749–757.CrossRefGoogle Scholar
Gregory, R. D., Keymer, A. E. & Harvey, P. H. (1996). Helminth parasite richness among vertebrates. Biodiversity and Conservation, 5, 985–997.CrossRefGoogle Scholar
Greives, T. J., McGlothlin, J. W., Jawor, J. M., Demas, G. E. & Ketterson, E. D. (2006). Testosterone and innate immune function inversely covary in a wild population of breeding dark-eyed juncos (Junco hyemalis). Functional Ecology, 20, 812–818.CrossRefGoogle Scholar
Grenfell, B. T. (1992). Parasitism and the dynamics of ungulate grazing systems. American Naturalist, 139, 907–929.CrossRefGoogle Scholar
Grenfell, B. T. & Dobson, A. P. (eds.) (1995). Ecology of Infectious Diseases in Natural Populations. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Grenfell, B. T., Dietz, K. & Roberts, M. G. (1995). Modelling the immuno-epidemiology of macroparasites in naturally-fluctuated host populations. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 362–383.CrossRefGoogle Scholar
Griffiths, D. (1999). On investigation local–regional species richness relationships. Journal of Animal Ecology, 68, 1051–1055.CrossRefGoogle Scholar
Grodzinski, W. & Wunder, B. A. (1975). Ecological energetics of small mammals. In Small Mammals: Their Productivity and Population Dynamics, ed. Golley, F. B., Petrusewitz, K. & Ryszkowski, L.. Cambridge, UK: Cambridge University Press, pp. 173–204.Google Scholar
Gromov, V. S., Krasnov, B. R. & Shenbrot, G. I. (2000). Space use in Wagner's gerbil Gerbillus dasyurus in the Negev Highlands, Israel. Acta Theriologica, 45, 175–182.CrossRefGoogle Scholar
Gruen, J. R. & Weissman, S. M. (1997). Evolving views of the major histocompatibility complex. Blood, 90, 4252–4265.Google ScholarPubMed
Gubareva, N. P., Akiev, A. K., Zemelman, B. M. & Abdurakhmanov, G. A. (1976). Effect of some factors on formation of the plague blockage in Ceratophyllus tesquorum and Neopsylla setosa setosa. Parazitologiya, 10, 315–319 (in Russian).Google Scholar
Guégan, J.-F. & Hugueny, B. A. (1994). A nested parasite species subset pattern in tropical fish host as major determinant of parasite infracommunity structure. Oecologia, 100, 184–189.CrossRefGoogle ScholarPubMed
Guégan, J.-F. & Kennedy, C. R. (1996). Parasite richness/sampling effort/host range: the fancy three-piece jigsaw puzzle. Parasitology Today, 12, 367–369.CrossRefGoogle ScholarPubMed
Guégan, J.-F. & Morand, S. (1996). Polyploid hosts: strange attractors for parasites! Oikos, 7, 366–370.CrossRefGoogle Scholar
Guégan, J.-F., Lambert, A., Leveque, C. & Euzet, L. (1992). Can host body size explain the parasite species richness in tropical freshwater fishes?Oecologia, 90, 197–204.CrossRefGoogle ScholarPubMed
Guégan, J.-F., Morand, S. & Poulin, R. (2005). Are there general laws in parasite community ecology? The emergence of spatial parasitology and epidemiology. In Parasitism and Ecosystems, ed. Thomas, F., Guégan, J.-F. & Renaud, F.. Oxford, UK: Oxford University Press, pp. 22–42.CrossRefGoogle Scholar
Guerrero, O. M., Chinchilla, M. & Abrahams, E. (1997). Increasing of Toxoplasma gondii (Coccidia: Sarcocystidae) infections by Trypanosoma lewisi (Kinetoplastida, Trypanosomatidae) in white rats. Revista de Biologia Tropical, 45, 877–882.Google Scholar
Gulland, F. M. D. (1995). The impact of infectious diseases on wild animal populations: a review. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 20–51.CrossRefGoogle Scholar
Guo, T. & Xu, R. (1999). Trophic niche of flea in the southern slope of the Himalaya Mountains. Chinese Journal of Applied Ecology, 10, 67–70 (in Chinese).Google Scholar
Guo, X.-G., Gong, Z.-D., Qian, T.-J., et al. (2000). Flea fauna investigation in some foci of human plague in Yunnan, China. Acta Zootaxonomica Sinica, 25, 291–297 (in Chinese).Google Scholar
Gurevitch, J., Morrow, L. L., Wallace, A. & Walsh, J. S. (1992). A meta-analysis of competition in field experiments. American Naturalist, 140, 539–572.CrossRefGoogle Scholar
Gurtler, R. E., Cohen, J. E., Cecere, M. C. & Chuit, R. (1997). Shifting host choices of the vector of Chagas disease, Triatoma infestans, in relation to the availability of hosts in houses in north-west Argentina. Journal of Applied Ecology, 34, 699–715.CrossRefGoogle Scholar
Gusev, V. M., Petrosyan, E. A., Guseva, A. A., Eigelis, Y. K. & Tchernyavsky, A. M. (1962). Wild birds: carriers of ectoparasites in the Trans-Caucasus. Proceedings of the Azerbaijanian Anti-Plague Station, 3, 177–184 (in Russian).Google Scholar
Guseva, A. A. & Kosminsky, R. B. (1974). Feeding and reproduction of Frontopsylla elata caspica Ioff et Arg., 1934 (Ceratophyllidae, Siphonaptera) in experiments. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 132–134 (in Russian).Google Scholar
Gustafson, C. R., Bickford, A. A., Cooper, G. L. & Charlton, B. R. (1997). Sticktight fleas associated with fowl pox in a backyard chicken flock in California. Avian Diseases, 41, 1006–1009.CrossRefGoogle Scholar
Gwinner, H. & Berger, S. (2005). European starlings: nestling condition, parasites and green nest material during the breeding season. Journal of Ornithology, 146, 365–371.CrossRefGoogle Scholar
Haag-Wackernagel, D. & Spiewak, R. (2004). Human infestation by pigeon fleas (Ceratophyllus columbae) from feral pigeons. Annals of Agricultural and Environmental Medicine, 11, 343–346.Google ScholarPubMed
Haas, G. E. (1965). Comparative suitability of the four murine rodents of Hawaii as hosts for Xenopsylla vexabilis and X. cheopis (Siphonaptera). Journal of Medical Entomology, 2, 75–83.CrossRefGoogle Scholar
Haas, G. E. (1966). A technique for estimating the total number of rodent fleas in cane fields in Hawaii. Journal of Medical Entomology, 2, 392–394.CrossRefGoogle ScholarPubMed
Haas, G. E. (1969). Quantitative relationships between fleas and rodents in a Hawaiian cane field. Pacific Science, 23, 70–82.Google Scholar
Haas, G. E. & Kucera, J. R. (2004). Fleas (Siphonaptera) in nests of voles (Microtus spp.) in montane habitats of three regions of Utah. Western North American Naturalist, 64, 346–352.Google Scholar
Haas, G. E. & Wilson, N. (1985). Rodent fleas (Siphonaptera) in tree cavities of woodpeckers in Alaska, USA. Canadian Field Naturalist, 100, 554–556.Google Scholar
Haas, G. E. & Wilson, N. (1998). Polygenis martinezbaezi (Siphonaptera: Rhopalopsyllidae) reared from a rodent nest found in the Peloncillo Mountains of southwestern New Mexico. Journal of Medical Entomology, 35, 431–432.CrossRefGoogle ScholarPubMed
Haas, G. E., Johnson, L. & Wilson, N. (1980). Siphonaptera from mammals in Alaska, USA. Supplement 2. Southeastern Alaska. Journal of the Entomological Society of British Columbia, 77, 43–46.Google Scholar
Haas, G. E., Rumfelt, T. & Wilson, N. (1981). Fleas (Siphonaptera) from nests of the tree swallow Iridoprocne bicolor and the violet-green swallow Tachycineta thalassina in Alaska, USA. Wasmann Journal of Biology, 39, 37–41.Google Scholar
Haas, G. E., Kucera, J. R., Runck, A. M., Macdonald, S. O. & Cook, J. A. (2005). Mammal fleas (Siphonaptera: Ceratophyllidae) new for Alaska and the southeastern mainland collected during seven years of a field survey of small mammals. Journal of the Entomological Society of British Columbia, 102, 65–75.Google Scholar
Haeselbarth, E., Segerman, J. & Zumpt, F. (1966). The arthropod parasites of vertebrates in Africa south of the Sahara (Ethiopian region). III. Insecta excl. Phthiraptera. Publications of the South African Institute for Medical Research, 13, 1–283.Google Scholar
Hafner, M. S. & Nadler, S. A. (1988). Phylogenetic trees support the coevolution of parasites and their hosts. Nature, 332, 258–259.CrossRefGoogle ScholarPubMed
Hafner, M. S. & Nadler, S. A. (1990). Cospeciation in host–parasite assemblages: comparative analysis of rates of evolution and timing of cospeciation events. Systematic Zoology, 39, 192–204.CrossRefGoogle Scholar
Hafner, M. S. & Page, R. D. M. (1995). Molecular phylogenies and host–parasite cospeciation: gophers and lice as a model system. Philosophical Transactions of the Royal Society of London B, 349, 77–83.CrossRefGoogle ScholarPubMed
Haftorn, S. (1994). The act of tremble-thrusting in tit nests, performance and possible function. Fauna Norvegica C, 17, 55–74.Google Scholar
Haitlinger, R. (1970). Die Flöhe (Siphonaptera) der Kleinsäuger aus den West- und Mittelsudeten. Polskie Pismo Entomologiczne, 40, 749–762.Google Scholar
Haitlinger, R. (1971). Aphanipterofauna drobnych gryzoni i owadozernych Wrocławia. Zootechnika, 30, 9–22.Google Scholar
Haitlinger, R. (1973). The parasitological investigation of small mammals of the Góry Sowie (Middle Sudetes). I. Siphonaptera (Insecta). Polskie Pismo Entomologiczne, 43, 499–519.Google Scholar
Haitlinger, R. (1974). Fleas (Siphonaptera) of small mammals of the Pieniny, Poland. Polskie Pismo Entomologiczne, 44, 765–788.Google Scholar
Haitlinger, R. (1975). The parasitological investigation of small mammals of the Góry Sowie (Middle Sudetes). II. Siphonaptera (Insecta). Polskie Pismo Entomologiczne, 45, 373–396.Google Scholar
Haitlinger, R. (1977). The parasitological investigation of small mammals of the Góry Sowie (Middle Sudetes). VI. Siphonaptera, Anoplura, Acarina. Polskie Pismo Entomologiczne, 47, 429–492.Google Scholar
Haitlinger, R. (1981). Structure of Arthropod communities occurring on Microtus arvalis (Pall.) in various habitats. I. Faunistic differentiation, dominance structure, arthropod infestation intensiveness in relation to habitats and host population dynamics. Polish Ecological Studies, 7, 271–292.Google Scholar
Haitlinger, R. (1989). Arthropods (Acari, Anoplura, Siphonaptera, Coleoptera) of small mammals of the Babia Góra Mts. Acta Zoologica Cracoviensia, 32, 15–56.Google Scholar
Hallas, T. & Bang, P. (1976). Fleas caught on small mammals at seven locations in eastern Denmark. Flora og Fauna, 82, 11–18.Google Scholar
Halliwell, R. E. W. & Longino, S. J. (1985). IgE and IgG antibodies to flea antigen in differing dog populations. Veterinary Immunology and Immunopathology, 8, 215–223.CrossRefGoogle ScholarPubMed
Halliwell, R. E. W. & Schemmer, K. R. (1987). The role of basophils in the immunopathogenesis of hypersensitivity to fleas (Ctenocephalides felis) in dogs. Veterinary Immunology and Immunopathology, 15, 203–213.CrossRefGoogle Scholar
Halvorsen, O. (1985). On the relationship between social status of a host and risk of parasitic infection. Oikos, 47, 71–74.CrossRefGoogle Scholar
Hamilton, W. D. & Zuk, M. (1982). Heritable true fitness and bright birds: a role for parasites?Science, 218, 384–387.CrossRefGoogle Scholar
Hamilton, P. B., Stevens, J. R., Holz, P., et al. (2005). The inadvertent introduction into Australia of Trypanosoma nabiasi, the trypanosome of the European rabbit (Oryctolagus cuniculus), and its potential for biocontrol. Molecular Ecology, 14, 3167–3175.CrossRefGoogle Scholar
Hanski, I. (1982). Communities of bumblebees: testing the core–satellite hypothesis. Annales Zoologici Fennici, 19, 65–73.Google Scholar
Hanski, I. (1998). Metapopulation dynamics. Nature, 396, 41–49.CrossRefGoogle Scholar
Hanski, I., Kouki, J. & Halkka, A. (1993). Three explanations of the positive relationship between distribution and abundance of species. In Species Diversity in Ecological Communities: Historical and Geographical Perspectives, ed. Ricklefs, R. E. & Schluter, D.. Chicago, IL: University of Chicago Press, pp. 108–116.Google Scholar
Hanski, I., Moilanen, A. & Gyllenberg, M. (1996). Minimum viable metapopulation size. American Naturalist, 147, 527–541.CrossRefGoogle Scholar
Harper, G. H., Marchant, A. & Boddington, D. G. (1992). The ecology of the hen flea Ceratophyllus gallinae and the moorhen flea Dasypsyllus gallinulae in nestboxes. Journal of Animal Ecology, 61, 317–327.CrossRefGoogle Scholar
Harrison, J. F. & Roberts, S. P. (2000). Flight respiration and energetics. Annual Review of Physiology, 62, 179–205.CrossRefGoogle ScholarPubMed
Hart, B. L. (1990). Behavioral adaptations to pathogens and parasites: five strategies. Neuroscience and Biobehavioural Reviews, 14, 273–294.CrossRefGoogle ScholarPubMed
Hart, B. L. & Pryor, P. A. (2004). Developmental and hair-coat determinants of grooming behaviour in goats and sheep. Animal Behaviour, 67, 11–19.CrossRefGoogle Scholar
Hart, B. L., Hart, L. A., Mooring, M. S. & Olubayo, R. (1992). Biological basis of grooming behaviour in antelope: the body size, vigilance and habitat principles. Animal Behaviour, 44, 615–631.CrossRefGoogle Scholar
Hartley, S. & Shorrocks, B. (2002). A general framework for the aggregation model of coexistence. Journal of Animal Ecology, 71, 651–662.CrossRefGoogle Scholar
Hartwell, W. V., Quan, S. F., Scott, K. G. & Kartman, L. (1958). Observations on flea transfer between hosts: a mechanism in the spread of bubonic plague. Science, 127, 814.CrossRefGoogle ScholarPubMed
Harvey, P. H. & Pagel, M. D. (1991). The Comparative Method in Evolutionary Biology. Oxford, UK: Oxford University Press.Google Scholar
Hasibender, G. & Dye, C. (1988). Population dynamics of mosquito-borne disease: persistence in a completely heterogenous environment. Theoretical Population Biology, 33, 31–53.CrossRefGoogle Scholar
Hastriter, M. W. (1997). Establishment of the tungid flea, Tunga monositus (Siphonaptera: Pulicidae), in the United States. Great Basin Naturalist, 57, 281–282.Google Scholar
Hastriter, M. W. (2000a). Jordanopsylla becki (Siphonaptera: Ctenophthalmidae), a new species of flea from the Nevada Test Site. Proceedings of the Entomological Society of Washington, 102, 135–141.Google Scholar
Hastriter, M. W. (2000b). Echidnophaga suricatta (Siphonaptera: Pulicidae), a new species of flea from the Northern Cape Province, South Africa. African Zoology, 35, 77–83.CrossRefGoogle Scholar
Hastriter, M. W. (2001a). Five new species and new subgenus of fleas (Siphonaptera: Chimaeropsyllidae, Ctenophthalmidae) from Africa. Proceedings of the Entomological Society of Washington, 103, 832–848.Google Scholar
Hastriter, M. W. (2001b). Fleas (Siphonaptera: Ctenophthalmidae and Rhopalopsyllidae) from Argentina and Chile with two new species from the rock rat, Aconaemys fuscus, in Chile. Annals of Carnegie Museum, 70, 169–178.Google Scholar
Hastriter, M. W. (2004). Revision of the flea genus Jellisonia Traub, 1944 (Siphonaptera: Ceratophyllidae). Annals of Carnegie Museum, 73, 213–238.Google Scholar
Hastriter, M. W. & Eckerlin, R. P. (2003). Jellisonia painteri (Siphonaptera: Ceratophyllidae), a new species of flea from Guatemala. Annals of Carnegie Museum, 72, 215–221.Google Scholar
Hastriter, M. W. & Haas, G. E. (2005). Bionomics and distribution of species of Hystrichopsylla in Arizona and New Mexico, with a description of Hystrichopsylla dippiei oblique, n. ssp. (Siphonaptera: Hystrichopsyllidae). Journal of Vector Ecology, 30, 251–262.Google Scholar
Hastriter, M. W. & Tipton, V. J. (1975). Fleas (Siphonaptera) associated with small mammals of Morocco. Journal of the Egyptian Public Health Association, 50, 79–169.Google Scholar
Hastriter, M. W. & Whiting, M. E. (2002). Macropsylla novaehollandiae (Siphonaptera: Hystrichopsyllidae), a new species of flea from Tasmania. Proceedings of the Entomological Society of Washington, 104, 663–671.Google Scholar
Hastriter, M. W. & Whiting, M. F. (2003). Siphonaptera (fleas). In Encyclopedia of Insects, ed. Resh, V. H. & Carde, R.. Orlando, FL: Elsevier Science, pp. 1039–1045.Google Scholar
Hastriter, M. W., Egoscue, H. J. & Traub, R. (1998). A description of the male of Jordanopsylla allredi Traub and Tipton, 1951, and characterization of the tribes within Anomiopsyllinae (Siphonaptera: Ctenophthalmidae). Proceedings of the Entomological Society of Washington, 100, 141–146.Google Scholar
Hastriter, M. W., Zyzak, M. D., Soto, R., et al. (2002). Fleas (Siphonaptera) from Ancash Department, Peru with the description of a new species, Ectinorus alejoi (Rhopalopsyllidae), and the description of the male of Plocopsylla pallas (Rothschild, 1914) (Sephanocircidae). Annals of Carnegie Museum, 71, 87–106.Google Scholar
Hastriter, M. W., Haas, G. E. & Wilson, N. (2006). New distribution records for Stenoponia americana (Baker) and Stenoponia ponera Traub and Johnson (Siphonaptera: Ctenophthalmidae) with a review of records from the southwestern United States. Zootaxa, 1253, 51–59.Google Scholar
Haukisalmi, V. & Henttonen, H. (1990). The impact of climatic factors and host density on the long-term population dynamics of vole helminths. Oecologia, 83, 309–315.CrossRefGoogle ScholarPubMed
Haukisalmi, V. & Henttonen, H. (1993). Coexistence in helminths of the bank vole Clethrionomys glareolus. I. Patterns of co-occurrence. Journal of Animal Ecology, 62, 221–229.CrossRefGoogle Scholar
Hawlena, H., Abramsky, Z. & Krasnov, B. R. (2005). Age-biased parasitism and density-dependent distribution of fleas (Siphonaptera) on a desert rodent. Oecologia, 146, 200–208.CrossRefGoogle Scholar
Hawlena, H., Krasnov, B. R., Abramsky, Z., et al. (2006a). Flea infestation and energy requirements of rodent hosts: are there general rules?Functional Ecology, 20, 1028–1036.CrossRefGoogle Scholar
Hawlena, H., Khokhlova, I. S., Abramsky, Z. & Krasnov, B. R. (2006b). Age, intensity of infestation by flea parasites and body mass loss in a rodent host. Parasitology, 133, 187–193.CrossRefGoogle Scholar
Hawlena, H., Abramsky, Z. & Krasnov, B. R. (2006c). Ectoparasites and age-dependent survival in a desert rodent. Oecologia, 148, 30–39.CrossRefGoogle Scholar
Hawlena, H., Abramsky, Z., Krasnov, B. R. & Saltz, D. (2007a). Host defence versus intraspecific competition in the regulation of infrapopulations of the flea Xenopsylla conformis on its rodent host Meriones crassus. International Journal for Parasitology, 37, 919–925.CrossRefGoogle Scholar
Hawlena, H., Bashary, D., Abramsky, Z. & Krasnov, B. R. (2007b). Benefits, costs and constraints of anti-parasitic grooming in adult and juvenile rodents. Ethology, 113, 394–402.CrossRefGoogle Scholar
He, J.-H., Liang, Y. & Zhang, H.-Y. (1997). A study on the transmission of plague though seven kinds of fleas in rat type and wild rodent type plague foci in Yunnan. Chinese Journal of Epidemiology, 18, 236–240 (in Chinese).Google ScholarPubMed
Heath, A. W., Arfsten, A., Yamanaka, M., Dryden, M. W. & Dale, B. (1994). Vaccination against the cat flea Ctenocephalides felis felis. Parasite Immunology, 16, 187–191.CrossRefGoogle ScholarPubMed
Hecht, O. (1943). La reacciones da la piel contra las picaduras de insectos como fenómenas alergicos. Revista de Sanidad y Asistencia Social, 8, 945–959.Google Scholar
Heckmann, R., Gansen, B. & Hom, M. (1967). Maternal transfer of immunity to rat coccidiosis: Eimeria nieschulzi (Dieben – 1924). Journal of Protozoology, S14, 35.Google Scholar
Hecnar, S. J. & Closkey, M' R. T. (1997). Patterns of nestedness and species association in a pond-dwelling amphibian fauna. Oikos, 80, 371–381.CrossRefGoogle Scholar
Heeb, P., Werner, I., Richner, H. & Kölliker, M. (1996). Horizontal transmission and reproductive rates of hen fleas in great tit nests. Journal of Animal Ecology, 65, 474–484.CrossRefGoogle Scholar
Heeb, P., Werner, I., Kölliker, M. & Richner, H. (1998). Benefits of induced host responses against an ectoparasite. Proceedings of the Royal Society of London B, 265, 51–56.CrossRefGoogle Scholar
Heeb, P., Werner, I., Mateman, A. C., et al. (1999). Ectoparasite infestation and sex-biased local recruitment of hosts. Nature, 400, 63–65.CrossRefGoogle ScholarPubMed
Heeb, P., Kölliker, M. & Richner, H. (2000). Bird–ectoparasite interactions, nest humidity, and ectoparasite community structure. Ecology, 81, 958–968.Google Scholar
Heino, J., Muotka, T. & Paavola, R. (2003). Determinants of macroinvertebrate diversity in headwater streams: regional and local influences. Journal of Animal Ecology, 72, 425–434.CrossRefGoogle Scholar
Heitman, T. L., Koski, K. G. & Scott, M. E. (2003). Energy deficiency alters behaviours involved in transmission of Heligmosomoides polygyrus (Nematoda) in mice. Canadian Journal of Zoology, 81, 1767–1773.CrossRefGoogle Scholar
Heller-Haupt, A., Kagaruki, L. K. & Varma, M. G. R. (1996). Resistance and cross-resistance in rabbits to adults of three species of African ticks (Acari: Ixodidae). Experimental and Applied Acarology, 20, 155–165.CrossRefGoogle Scholar
Hemmes, R. B., Alvarado, A. & Hart, B. L. (2002). Use of California bay foliage by wood rats for possible fumigation of nest-borne ectoparasites. Behavioral Ecology, 13, 381–385.CrossRefGoogle Scholar
Henry, P. Y., Poulin, B., Rousset, F., Renaud, F. & Thomas, F. (2004). Infestation by the mite Harpirhynchus nidulans in the bearded tit Panurus biarmicus. Bird Study, 51, 34–40.CrossRefGoogle Scholar
Heukelbach, J., Oliveira, F. A. S., Hesse, G. & Feldmeier, H. (2001). Tungiasis: a neglected health problem of poor communities. Tropical Medicine and International Health, 6, 267–272.CrossRefGoogle ScholarPubMed
Heukelbach, J., Wilcke, T. & Feldmeier, H. (2004). Cutaneous larva migrans (creeping eruption) in an urban slum in Brazil. International Journal of Dermatology, 43, 511–515.CrossRefGoogle Scholar
Hillebrand, H. (2005). Regressions of local on regional diversity do not reflect the importance of local interactions or saturation of local diversity. Oikos, 110, 195–198.CrossRefGoogle Scholar
Hillebrand, H. & Blenckner, T. (2002). Regional and local impact on species diversity: from pattern to processes. Oecologia, 132, 479–491.CrossRefGoogle ScholarPubMed
Hinaidy, H. K. (1991). The biology of Dipylidium caninum. Zentralblatt für Veterinärmedizin B, 38, 329–336.Google ScholarPubMed
Hinkle, N. C., Koehler, P. G. & Kern, W. H. (1991). Hematophagous strategies of the cat flea (Siphonaptera: Pulicidae). Florida Entomologist, 74, 377–385.CrossRefGoogle Scholar
Hinkle, N. C., Koehler, P. G. & Patterson, R. S. (1998). Host grooming efficiency for regulation of cat flea (Siphonaptera: Pulicidae) populations. Journal of Medical Entomology, 35, 266–269.CrossRefGoogle ScholarPubMed
Hinnebusch, B. J., Gage, K. L. & Schwan, T. G. (1998). Estimation of vector infectivity rates for plague by means of a standard curve-based competitive polymerase chain reaction method to quantify Yersinia pestis in fleas. American Journal of Tropical Medicine and Hygiene, 58, 562–569.CrossRefGoogle ScholarPubMed
Hinton, H. E. (1958). The phylogeny of the Panorpoid orders. Annual Review of Entomology, 3, 181–206.CrossRefGoogle Scholar
Hoberg, E. P., Brooks, D. R. & Siegel-Causey, D. (1997). Host–parasite co-speciation: history, principles, and prospects. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 212–235.Google Scholar
Holland, C. (1984). Interactions between Moniliformis (Acanthocephala) and Nippostrongylus (Nematoda) in the small intestine of laboratory rats. Parasitology, 88, 303–315.Google ScholarPubMed
Holland, G. P. (1955). Primary and secondary sexual characteristics of some Ceratophyllinae, with notes on the mechanism of copulation (Siphonaptera). Transactions of the Royal Entomological Society of London, 107, 233–248.CrossRefGoogle Scholar
Holland, G. P. (1964). Evolution, classification and host relationships of Siphonaptera. Annual Review of Entomology, 9, 123–146.CrossRefGoogle Scholar
Holland, G. P. (1985). The fleas of Canada, Alaska and Greenland (Siphonaptera). Memoirs of the Entomological Society of Canada, 130, 1–631.Google Scholar
Holmes, J. C. & Price P. W. (1986). Communities of parasites. In Community Ecology: Patterns and Processes, ed. Kikkawa, J. & Anderson, D. J.. Oxford, UK: Blackwell Science, pp. 187–213.Google Scholar
Holmstad, P. R., Hudson, P. J. & Skorping, A. (2005). The influence of a parasite community on the dynamics of a host population: a longitudinal study on willow ptarmigan and their parasites. Oikos, 111, 377–391.CrossRefGoogle Scholar
Holt, R. D., Dobson, A. P., Begon, M., Bowers, R. G. & Schauber, E. M. (2003). Parasite establishment in host communities. Ecology Letters, 6, 837–842.CrossRefGoogle Scholar
Honnay, O., Hermy, M. & Coppin, P. (1999). Nested plant communities in deciduous forest fragments: species relaxation or nested habitats?Oikos, 84, 119–129.CrossRefGoogle Scholar
Hoogland, J. L. & Sherman, P. W. (1976). Advantages and disadvantages of bank swallow (Riparia riparia) coloniality. Ecological Monographs, 46, 33–58.CrossRefGoogle Scholar
Hopkins, G. H. E. (1957). Host associations of Siphonaptera. In 1st Symposium on Host Specificity amongst Parasites of Vertebrates, ed. Baer, J. G.. Neuchâtel, Switzerland: Institut de Zoologie, Université de Neuchâtel, pp. 64–87.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1953). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 1, Tungidae and Pulicidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1956). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 2, Coptopsyllidae, Vermipsyllidae, Stephanocircidae, Ischnopsyllidae, Hypsophthalmidae and Xiphiopsyllidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1962). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 3, Hystrichopsyllidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1966). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 4, Hystrichopsyllidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1971). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 5, Leptopsyllidae and Ancistropsyllidae. London: Trustees of the British Museum.Google Scholar
Hopla, C. E. (1980). Fleas as vectors of tularemia in Alaska. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 287–300.Google Scholar
Hörak, P., Ots, I., Vellau, H., Spottiswoode, C. & M⊘ller, A. P. (2001). Carotenoid-based plumage coloration reflects hemoparasite infection and local survival in breeding great tits. Oecologia, 126, 166–173.CrossRefGoogle ScholarPubMed
Hsu, M.-H. & Wu, W.-J. (2000). Effects of multiple mating on female reproductive output in the cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 37, 828–834.CrossRefGoogle Scholar
Hsu, M.-H. & Wu, W.-J. (2001). Off-host observations of mating and postmating behaviors in the cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 38, 352–360.CrossRefGoogle Scholar
Hsu, M.-H., Hsu, T.-C. & Wu, W.-J. (2002). Distribution of cat fleas (Siphonaptera: Pulicidae) on the cat. Journal of Medical Entomology, 39, 685–688.CrossRefGoogle ScholarPubMed
Hu, X.-L., He, J.-H., Zhang, H.-Y., Zhao, W.-H. & Liang, Y. (1996). Evaluation on qualities of house rat fleas from laboratory breeding. Endemic Diseases Bulletin, 11, 14–17 (in Chinese).Google Scholar
Hu, X.-L., He, J.-H., Zhang, H.-Y., et al. (1998). Experimental breeding and life history of the flea Ctenophthalmus quadratus. Endemic Diseases Bulletin, 13, 26–28 (in Chinese).Google Scholar
Hu, X.-L., He, J.-H. & Zhang, H.-Y. (2001). Body weight and quantity of blood-sucking of six species of fleas in Yunnan province, China. Chinese Journal of Pest Control, 17, 393–395 (in Chinese).Google Scholar
Hudson, B. W. & Prince, F. M. (1958a). A method for large-scale rearing of the cat flea, Ctenocephalides felis felis (Bouché). Bulletin of World Health Organization, 19, 1126–1129.Google Scholar
Hudson, B. W. & Prince, F. M. (1958b). Culture methods for fleas Pulex irritans (L.) and Pulex simulans Baker. Bulletin of World Health Organization, 19, 1129–1133.Google Scholar
Hudson, B. W., Feingold, B. F. & Kartman, L. (1960a). Allergy to flea bites. I. Experimental induction of flea-bite sensitivity in guinea pigs. Experimental Parasitology, 9, 18–24.CrossRefGoogle Scholar
Hudson, B. W., Feingold, B. F. & Kartman, L. (1960b). Allergy to flea bites. II. Investigations of flea bite sensitivity in humans. Experimental Parasitology, 9, 264–270.CrossRefGoogle Scholar
Hudson, P. J. & Dobson, A. P. (1995). Macroparasites: observed patterns. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 144–176.CrossRefGoogle Scholar
Hudson, P. J. & Dobson, A. P. (1997). Host–parasite processes and demographic consequences. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J. M.. Oxford, UK: Oxford University Press, pp. 128–154.Google Scholar
Hudson, P. J. & Greenman, J. (1998). Competition mediated by parasites: biological and theoretical progress. Trends in Ecology and Evolution, 13, 387–390.CrossRefGoogle ScholarPubMed
Hughes, J. B. (2000). The scale of resource specialization and the distribution and abundance of lycaenid butterflies. Oecologia, 123, 375–383.CrossRefGoogle ScholarPubMed
Hughes, T. P., Baird, A. H., Dinsdale, E. A., et al. (2000). Supply-side ecology works both ways: the link between benthic adults, fecundity, and larval recruits. Ecology, 81, 2241–2249.CrossRefGoogle Scholar
Hughes, V. L. & Randolph, S. E. (2001). Testosterone depresses innate and acquired resistance to ticks in natural rodent hosts: a force for aggregated distributions of parasites. Journal of Parasitology, 87, 49–54.CrossRefGoogle ScholarPubMed
Hugueny, B. A. & Guégan, J.-F. (1997). Community nestedness and the proper way to assess statistical significance by Monte-Carlo tests: some comments on Worthen and Rohde's (1996) paper. Oikos, 80, 572–574.CrossRefGoogle Scholar
Humphreys, N. E. & Grencis, R. K. (2002). Effects of ageing on the immunoregulation of parasitic infection. Infection and Immunity, 70, 5148–5157.CrossRefGoogle ScholarPubMed
Humphries, D. A. (1966). The function of combs in fleas. Entomological Monthly Magazine, 102, 232–236.Google Scholar
Humphries, D. A. (1967a). The mating behaviour of the hen flea Ceratophyllus gallinae (Schrank) (Siphonaptera: Insecta). Animal Behaviour, 15, 82–90.CrossRefGoogle Scholar
Humphries, D. A. (1967b). The action of the male genitalia during the copulation of the hen flea, Ceratophyllus galinnae (Schrank). Proceedings of the Royal Entomological Society of London A, 42, 101–106.CrossRefGoogle Scholar
Humphries, D. A. (1967c). Function of combs in ectoparasites. Nature, 215, 319.CrossRefGoogle Scholar
Humphries, D. A. (1968). The host-finding behavior of the hen flea, Ceratophyllus gallinae (Schrank) (Siphonaptera). Parasitology, 59, 403–414.CrossRefGoogle Scholar
Humphries, D. A. (1969). Behavioral aspects of the ecology of the sand-martin flea Ceratophyllus styx jordani Smit (Siphonaptera). Parasitology, 59, 311–334.CrossRefGoogle Scholar
Hunter, M. D. & Price, P. W. (1992). Playing chutes and ladders: heterogeneity and the relative roles of bottom–up and top–down forces in natural communities. Ecology, 73, 724–732.Google Scholar
Hürka, K. (1963a). Bat fleas (Aphaniptera, Ischnopsyllidae) of Czechoslovakia: contribution to the distribution, morphology, bionomy, ecology and systematics. I. Subgenus Ischnopsyllus Westw. Acta Faunistica Entomologica Musei Nationalis Pragae, 9, 57–120.Google Scholar
Hürka, K. (1963b). Bat fleas (Aphaniptera, Ischnopsyllidae) of Czechoslovakia: contribution to the distribution, morphology, bionomy, ecology and systematics. II. Subgenus Hexactenopsylla Oud., genus Rhinolophopsylla Oud., subgenus Nycteridopsylla Oud., subgenus Dinycteropsylla Ioff. Acta Universitatis Carolinae, Biologica, 3, 1–73.Google Scholar
Ioff, I. G. (1941). Ecology of Fleas in Relevance to their Medical Importance. Pyatygorsk, USSR: Pyatygorsk Publishers (in Russian).Google Scholar
Ioff, I. G. (1949). Aphaniptera of Kyrgyzstan. Ectoparasites, 1, 5–212 (in Russian).Google Scholar
Ioff, I. G. (1950). The alakurt. Materials to Knowledge of Fauna and Flora of the USSR [Materialy k Poznaniju Fauny i Flory SSSR], 2, 4–29 (in Russian).Google Scholar
Ioff, I. G. & Tiflov, V. E. (1954). The Key to Identification of Aphaniptera of the South-East of the USSR. Stavropol, USSR: Stavropol Publishers (in Russian).Google Scholar
Ioff, I. G., Tiflov, V. E., Argyropulo, A. I., et al. (1946). News species of fleas (Aphaniptera). Medical Parasitology [Meditsinskaya Parazitologiya], 15, 85–94 (in Russian).Google Scholar
Ioff, I. G., Mikulin, M. A. & Scalon, O. N. (1965). The Key to Identification of Fleas of the Middle Asia and Kazakhstan. Moscow, USSR: Meditsina (in Russian).Google Scholar
Iqbal, Q. L. (1973). On the presence of mating pheromone in the rat flea Nosopsyllus fasciatus (Bosc.). Pakistan Journal of Zoology, 5, 123–125.Google Scholar
Iqbal, Q. L. (1974). Host-finding behaviour of the rat flea Nosopsyllus fasciatus (Bosc.). Biologia, Lahore, 20, 147–150.Google Scholar
Iqbal, Q. J. & Humphries, D. A. (1970). Temperature as a critical factor in the mating behavior of the rat flea, Nosopsyllus fasciatus (Bosc.). Parasitology, 61, 375–380.CrossRefGoogle Scholar
Iqbal, Q. J. & Humphries, D. A. (1974). The mating behavior of the rat flea Nosopsyllus fasciatus Bosc. Pakistan Journal of Zoology, 6, 163–174.Google Scholar
Iqbal, Q. J. & Humphries, D. A. (1976). Remating in the rat flea Nosopsyllus fasciatus (Bosc.). Pakistan Journal of Zoology, 8, 39–41.Google Scholar
Iqbal, Q. J. & Humphries, D. A. (1983). Feeding behavior of the rat flea Nosopsyllus fasciatus. Pakistan Journal of Zoology, 14, 71–74.Google Scholar
Ives, A. R. (1988a). Aggregation and the coexistence of competitors. Annales Zoologici Fennici, 25, 75–88.Google Scholar
Ives, A. R. (1988b). Covariance, coexistence and the population dynamics of two competitors using a patchy resource. Journal of Theoretical Biology, 133, 345–361.CrossRefGoogle Scholar
Ives, A. R. (1991). Aggregation and coexistence in a carrion fly community. Ecological Monographs, 61, 75–94.CrossRefGoogle Scholar
Iwao, K. & Ohsaki, N. (1996). Inter- and intraspecific interactions among larvae of specialist and generalist parasitoids. Research on Population Ecology, 38, 265–273.CrossRefGoogle Scholar
Izsák, J. & Papp, L. (1995). Application of the quadratic entropy index for diversity studies on drosophilid species assemblages. Environmental and Ecological Statistics, 2, 213–224.CrossRefGoogle Scholar
Jackson, T. P. (2000). Adaptation to living in an open arid environment: lessons from the burrow structure of two South African whistling rats, Parotomys brantsii and P. littledalei. Journal of Arid Environments, 46, 345–355.CrossRefGoogle Scholar
Jaenike, J. (1990). Host specialization in phytophagous insects. Annual Review of Ecology and Systematics, 21, 243–273.CrossRefGoogle Scholar
Jaenike, J. & James, A. C. (1991). Aggregation and the coexistence of mycophagous Drosophila. Journal of Animal Ecology, 60, 913–928.CrossRefGoogle Scholar
Jameson, E. W. (1985). Pleioxenous host-restriction in fleas. Journal of Natural History, 19, 861–876.CrossRefGoogle Scholar
Jameson, E. W. (1999). Host–ectoparasite relationships among North American chipmunks. Acta Theriologica, 44, 225–231.CrossRefGoogle Scholar
Jamieson, B. G. M. (1987). The Ultrastructure and Phylogeny of Insect Spermatozoa. Cambridge, UK: Cambridge University Press.Google Scholar
Janeway, C., Travers, P., Walport, M. & Capra, J. (1999). Immunobiology: The Immune System in Health and Disease. New York: Garland.Google Scholar
Janion, S. M. (1962). Flea infestation of three rodent species: Apodemus agrarius, Apodemus flavicollis and Clethrionomys glareolus at the period of Apodemus agrarius mass occurrence. Bulletin of the Polish Academy of Sciences, Series of Biological Sciences, 10, 361–366.Google Scholar
Janion, S. M. (1968). Certain host–parasite relationships between rodents (Muridae) and fleas (Aphaniptera). Ekologia Polska, 16, 561–606.Google Scholar
Jarrett, W. F. (1975). Cat leukemia and its viruses. Advances in Veterinary Science and Comparative Medicine, 19, 165–193.Google ScholarPubMed
Jell, P. A. & Duncan, P. M. (1986). Invertebrates, mainly insects, from the freshwater, Lower Cretaceous, Koonwarra Fossil Bed (Korumburra Group), South Gippsland, Victoria. Memoir of the Association of Australasian Palaeontologists, 3, 111–205.Google Scholar
Jellison, W. L. (1959). Fleas and disease. Annual Review of Entomology, 4, 389–414.CrossRefGoogle Scholar
Johnsen, T. S. & Zuk, M. (1999). Parasites and tradeoffs in the immune response of female red jungle fowl. Oikos, 86, 487–492.CrossRefGoogle Scholar
Johnson, K. P., Adams, R. J. & Clayton, D. H. (2002). The phylogeny of the louse genus Brueelia does not reflect host phylogeny. Biological Journal of the Linnean Society, 77, 233–247.CrossRefGoogle Scholar
Johnson, K. P., Bush, S. E. & Clayton, D. H. (2005). Correlated evolution of host and parasite body size: tests of Harrison's rule using birds and lice. Evolution, 59, 1744–1753.Google ScholarPubMed
Johnston, C. M. & Brown, S. J. (1985). Xenopsylla cheopis: cellular expression of hypersensitivity to guinea pigs. Experimental Parasitology, 59, 81–89.CrossRefGoogle ScholarPubMed
Johnston, M., Johnston, D. & Richardson, A. (2005). Digestive capabilities reflect the major food sources in three species of talitrid amphipods. Comparative Biochemistry and Physiology B, 140, 251–257.CrossRefGoogle ScholarPubMed
Jokela, J., Schmid-Hempel, P. & Rigby, M. C. (2000). Dr Pangloss restrained by the Red Queen: steps towards a unified defence theory. Oikos, 89, 267–274.CrossRef
Jones, C. J. (1996). Immune responses to fleas, bugs and sucking lice. In The Immunology of Host–Ectoparasitic Arthropod Relationships, ed. Wikel, S. K.. Wallingford, UK: CAB International, pp. 150–174.Google Scholar
Jordan, K. (1945). On the deciduous frontal tubercle of some genera of Siphonaptera. Proceedings of the Royal Entomological Society of London B, 14, 113–116.Google Scholar
Jordan, K. (1962). Notes on the Tunga caecigena (Siphonaptera: Tungidae). Bulletin of the British Museum of Natural History, Entomology, 12, 353–364.Google Scholar
Jordan, K. & Rothschild, N. C. (1915a). On some Siphonaptera collected by W. Ruckbeil in East Turkestan. Ectoparasites, 1, 1–24.Google Scholar
Jordan, K. & Rothschild, N. C. (1915b). Contribution to our knowledge of American Siphonaptera. Ectoparasites, 1, 45–60.Google Scholar
Joseph, S. A. (1974). Incidence of the flea Ancistropsylla nepalensis Lewis, 1968 on the barking deer (Muntiacus muntjak aureus Smith, 1827) in India. Indian Veterinary Journal, 51, 356–358.Google Scholar
Joseph, S. A. & Mani, K. R. (1980). Cervus unicolor niger: the Indian sambar, a new host for Ancistropsylla nepalensis Lewis, 1968. Cheiron, 9, 200–202.Google Scholar
Joy, J. E. & Briscoe, N. J. (1994). Parasitic arthropods of white-footed mice at McClintock Wildlife Station, West Virginia. Journal of the American Mosquito Control Association, 10, 108–111.Google Scholar
Juricova, Z., Halouzka, J. & Hubalek, Z. (2002). Serologic survey for antibodies to Borrelia burgdorferi in rodents and detection of spirochaetes in ticks and fleas in South Moravia (Czech Republic). Biologia, Bratislava, 57, 383–387.Google Scholar
Jurík, M. (1974). Bionomics of fleas in birds' nests in the territory of Czechoslovakia. Acta Scientiarum Naturalium Brno, 8, 1–54.Google Scholar
Jurík, M. (1983a). To the knowledge of ecological conditions affecting the occurrence of specific and non-specific flea species on their hosts (Talpa europaea – Siphonaptera). Biologia, Bratislava, 38, 949–957.Google Scholar
Jurík, M. (1983b). Ceratophyllus vagabundus insularis Rothschild, 1906 and Ceratophyllus rossittensis Dampf, 1913 in Czechoslovakia (Siphonaptera). Folia Parasitologica, 30, 169–173.Google Scholar
Kaal, J. F., Baker, K. & Torgerson, P. R. (2006). Epidemiology of flea infestation of ruminants in Libya. Veterinary Parasitology, 141, 313–318.CrossRefGoogle ScholarPubMed
Kadatskaya, K. P. (1983). Facultative imaginal diapause in fleas Xenopsylla conformis (Siphonaptera). Parazitologiya, 17, 370–374 (in Russian).Google Scholar
Kadatskaya, K. P. & Kadatsky, N. G. (1983). Comparative data on abundance of fleas parasitic on Meriones erythrourus in the western and eastern parts of the Apsheron Peninsula in relation to the plague epizootic in 1976–1978. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 237–238 (in Russian).Google Scholar
Kadatskaya, K. P. & Shirova, L. F. (1983). Seasonal changes of reproduction in fleas Xenopsylla conformis in Azerbaijan. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 238–240 (in Russian).Google Scholar
Kam, M. & Degen, A. A. (1993). Energetics of lactation and growth in the fat sand rat, Psammomys obesus: new perspectives of resource partitioning and the effect of litter size. Journal of Theoretical Biology, 162, 353–369.CrossRefGoogle Scholar
Kam, M. & Degen, A. A. (1997). Energy requirements and the efficiency of utilization of metabolizable energy in free-living animals: evaluation of existing theories and generation of a new model. Journal of Theoretical Biology, 184, 101–104.CrossRefGoogle Scholar
Kam, M., Khokhlova, I. S. & Degen, A. A. (1997). Granivory and plant selection by desert gerbils of different body size. Ecology, 78, 2218–2229.CrossRefGoogle Scholar
Bai, Kamala M. & Prasad, R. S. (1979). Influence of nutrition on maturation of male rat fleas, Xenopsylla cheopis and X. astia. Journal of Medical Entomology, 16, 164–165.CrossRefGoogle Scholar
Kaňuch, P., Krištín, A. & Krištofik, J. (2005). Phenology, diet, and ectoparasites of Leisler's bat (Nyctalus leisleri) in the western Carpathians (Slovakia). Acta Chiropterologica, 7, 249–257.CrossRefGoogle Scholar
Karandina, P. C. & Darskaya, N. F (1974). Observations on the pre-imaginal development in fleas parasitic on ground squirrels: Ceratophyllus (Citellophilus) tesquorum Wagn., 1898. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 143–144 (in Russian).Google Scholar
Kareiva, P. & Wennergren, U. (1995). Connecting landscape patterns to ecosystem and population processes. Nature, 373, 299–302.CrossRefGoogle Scholar
Karlson, R. H., Cornell, H. V. & Hughes, T. P. (2004). Coral communities are regionally enriched along an oceanic biodiversity gradient. Nature, 429, 867–870.CrossRefGoogle ScholarPubMed
Kartman, L., Prince, F. M., Quan, S. F. & Stark, H. E. (1958). New knowledge of the ecology of sylvatic plague. Annals of the New York Academy of Sciences, 70, 668–711.CrossRefGoogle ScholarPubMed
Kavaliers, M. & Colwell, D. (1994). Parasite infection attenuates nonopioid mediated predator-induced analgesia in mice. Physiology and Behavior, 55, 505–510.CrossRefGoogle ScholarPubMed
Kavaliers, M. & Colwell, D. D. (1995). Reduced spatial learning in mice infected with the nematodeHeligmosomoides polygyrus. Parasitology, 110, 591–597.CrossRefGoogle ScholarPubMed
Kavaliers, M., Colwell, D. D. & Choleris, E. (1998). Parasitized female mice display reduced aversive responses to the odours of infected males. Proceedings of the Royal Society of London B, 265, 1111–1118.CrossRefGoogle ScholarPubMed
K͡dra, A. H., Kruszewicz, A. G., Mazgajski, T. D. & Modlińska, E. (1996). The effects of the presence of fleas in nestboxes on fledglings of pied flycatchers and great tits. Acta Parasitologica, 41, 211–213.Google Scholar
Keeling, M. G. & Gilligan, C. A. (2000a). Metapopulation dynamics of bubonic plague. Nature, 407, 903–906.CrossRefGoogle Scholar
Keeling, M. G. & Gilligan, C. A. (2000b). Bubonic plague: a metapopulation model of a zoonosis. Proceedings of the Royal Society of London B, 267, 2219–2230.CrossRefGoogle Scholar
Kehr, J. D., Heukelbach, J., Mehlhorn, H. & Feldmeier, H. (2007). Morbidity assessment in sand flea disease (tungiasis). Parasitology Research, 100, 413–421.CrossRefGoogle Scholar
Kelly, D. W. & Thompson, C. E. (2000). Epidemiology and optimal foraging: modeling the ideal free distribution of insect vectors. Parasitology, 120, 319–327.CrossRefGoogle Scholar
Kelly, D. W., Mustafa, Z. & Dye, C. (1996). Density-dependent feeding success in a field population of the sandfly Lutzomyia longipalpis. Journal of Animal Ecology, 65, 517–527.CrossRefGoogle Scholar
Kennedy, C. R. & Bush, A. O. (1994). The relationship between pattern and scale in parasite communities: a stranger in a strange land. Parasitology, 109, 187–196.CrossRefGoogle Scholar
Kern, W. H. (1993). The autecology of the cat flea (Ctenocephalides felis felis Bouché) and the synecology of the cat flea and its domestic hosts (Felis catus). Unpublished Ph.D. thesis, University of Florida, Gainesville, FL.
Kern, W. H., Koehler, P. G. & Patterson, R. S. (1992). Diel patterns of cat flea (Siphonaptera, Pulicidae) egg and fecal deposition. Journal of Medical Entomology, 29, 203–206.CrossRefGoogle ScholarPubMed
Kern, W. H., Richman, D. L., Koehler, P. G. & Brenner, R. J. (1999). Outdoor survival and development of immature cat fleas (Siphonaptera: Pulicidae) in Florida. Journal of Medical Entomology, 36, 207–211.Google Scholar
Kettle, D. S. (1995). Medical and Veterinary Entomology. Wallingford, UK: CAB International.Google Scholar
Key, B. H. & Kemp, D. H. (1994). Vaccines against arthropods. American Journal of Tropical Medicine and Hygiene, 50, 87–96.CrossRefGoogle Scholar
Khalid, M. L., Morsy, T. A., Shennawy, El S. F., et al. (1992). Studies on flea fauna in El Fayoum Governorate, Egypt. Journal of the Egyptian Society of Parasitology, 22, 783–799.Google ScholarPubMed
Kharlamov, V. P. (1965). Changes in feeding activity and mobility of the flea Xenopsylla cheopis marked with radioactive 32P. Zoologicheskyi Zhurnal, 44, 547–550 (in Russian).Google Scholar
Khokhlova, I. S. & Knyazeva, T. V. (1983). The effect of spatial and social structure of the house mouse populations on flea assemblages. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 165–167 (in Russian).Google Scholar
Khokhlova, I. S., Krasnov, B. R., Shenbrot, G. I. & Degen, A. A. (1994). Seasonal body mass changes and habitat distribution in several rodent species from the Ramon erosion cirque, Negev Highlands, Israel. Zoologicheskyi Zhurnal, 73, 115–121 (in Russian).Google Scholar
Khokhlova, I. S., Krasnov, B. R., Shenbrot, G. I. & Degen, A. A. (2001). Body mass and environment: a study in Negev rodents. Israel Journal of Zoology, 47, 1–14.CrossRefGoogle Scholar
Khokhlova, I. S., Krasnov, B. R., Kam, M., Burdelova, N. V. & Degen, A. A. (2002). Energy cost of ectoparasitism: the flea Xenopsylla ramesis on the desert gerbil Gerbillus dasyurus. Journal of Zoology, 258, 349–354.CrossRefGoogle Scholar
Khokhlova, I. S., Spinu, M., Krasnov, B. R. & Degen, A. A. (2004a). Immune response to fleas in a wild desert rodent: effect of parasite species, parasite burden, sex of host and host parasitological experience. Journal of Experimental Biology, 207, 2725–2733.CrossRefGoogle Scholar
Khokhlova, I. S., Spinu, M., Krasnov, B. R. & Degen, A. A. (2004b). Immune responses to fleas in two rodent species differing in natural prevalence of infestation and diversity of flea assemblages. Parasitology Research, 94, 304–311.CrossRefGoogle Scholar
Khokhlova, I. S., Hovhanyan, A., Krasnov, B. R. & Degen, A. A. (2007). Reproductive success in two species of desert fleas: density-dependence and host effect. Journal of Experimental Biology, 210, 2121–2127.CrossRefGoogle ScholarPubMed
Khrustselevsky, V. P., Sokolova, A. A. & Balabas, N. G. (1971). Materials on the reproduction of Xenopsylla gerbilli in the Moyynkum Desert. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 436–439 (in Russian).Google Scholar
Khudyakov, I. S. (1965). Fleas (Aphaniptera) of the coastal zone of southern Primorie Region. Entomological Review, 44, 117–122 (in Russian).Google Scholar
Kiefer, M., Klimaszewski, S. M. & Krumpál, M. (1982). Zoogeographical regionalization of Mongolia on the basis of flea fauna (Siphonaptera). Polskie Pismo Entomologiczne, 52, 13–29.Google Scholar
Kiefer, M., Krumpál, M., Cendsuren, N., Lobachev, V. S. & Chotolchu, N. (1984). Checklist, distribution and bibliography of Mongolian Siphonaptera. In Erforschung biologischer Ressourcen der mongolischen Volksrepublik, vol. 4, ed. Stubbe, M., Hilbig, W. & Dawaa, N.. Halle, Germany: Wissenschaftliche Beiträge Universität Halle-Wittenberg, pp. 91–123.Google Scholar
Kilpatrick, A. M. & Ives, A. R. (2003). Species interactions can explain Taylor's power law for ecological time series. Nature, 422, 65–68.CrossRefGoogle ScholarPubMed
Kim, K. C. (1985a). Evolution and host associations of Anoplura. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 197–232.Google Scholar
Kim, K. C. (1985b). Evolutionary relationships of parasitic arthropods and mammals. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 3–82.Google Scholar
King, C. M. (1976). The fleas of a population of weasels in Wytham Wood, Oxford. Journal of Zoology, 180, 525–535.CrossRefGoogle Scholar
King, C. M. & Moody, J. E. (1982). The biology of the stoat (Mustela erminea) in the National Parks of New Zealand. VII. Fleas. New Zealand Journal of Zoology, 9, 141–144.CrossRefGoogle Scholar
Kings, R. C. & Teasly, M. (1980). Insect oogenesis: some generalities and their bearing on the ovarian development of fleas. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 337–340.Google Scholar
Kingsolver, J. G. (1987). Mosquito host choice and the epidemiology of malaria. American Naturalist, 130, 811–827.CrossRefGoogle Scholar
Kiriakova, A. N., Koptzev, L. A. & Koptzeva, Z. G. (1970). Annual number of generations of Xenopsylla fleas in the northern Kyzylkum Desert. Parazitologiya, 6, 528–536 (in Russian).Google Scholar
Kirillova, N. Y., Kirillov, A. A. & Ivashkina, V. A. (2006). Ectoparasites of the edible dormouse Glis glis L. of the Samarskaya Luka Peninsula (Russia). Polish Journal of Ecology, 54, 387–390.Google Scholar
Kirk, W. D. J. (1991). The size relationship between insects and their hosts. Ecological Entomology, 16, 351–359.CrossRefGoogle Scholar
Kisielewska, K. (1970). Ecological organization of intestinal helminth groupings in Clethrionomys glareolus (Schreb.) (Rodentia). III. Structure of helminth groupings in C. glareolus populations of various forest biocoenoses in Poland. Acta Parasitologica, 18, 163–176.Google Scholar
Klasing, K. C. (1998). Nutritional modulation of resistance to infectious diseases. Poultry Science, 77, 1119–1125.CrossRefGoogle ScholarPubMed
Klassen, G. J. (1992). Coevolution: a history of the macroevolutionary approach to studying host–parasite associations. Journal of Parasitology, 78, 573–587.CrossRefGoogle ScholarPubMed
Kleiber, M. (1961). The Fire of Life: An Introduction to Animal Energetics. New York: John Wiley.Google Scholar
Klein, J. (1990). Immunology. Oxford, UK: Blackwell Science.
Klein, J. M. (1966). Données écologiques et biologiques sur Synopsyllus fonquerniei Wagner et Roubaud, 1932 (Siphonaptera) puce du rat péridomestique, dans la région de Tananarive. Cahiers ORSTOM, Série Entomologie Médicale et Parasitologie, 4, 3–29.Google Scholar
Klein, J. M., Simonkovich, E., Alonso, J. M. & Baranton, G. (1975). Observations écologiques dans une zone epizootique de peste en Mauritanie. II. Les puces de rongeurs (Insecta, Siphonaptera). Cahiers ORSTOM, Série Entomologie Médicale et Parasitologie, 13, 29–39.Google Scholar
Klein, S. L. & Nelson, R. J. (1998a). Sex and species differences in cell-mediated immune responses in voles. Canadian Journal of Zoology, 76, 1394–1398.CrossRefGoogle Scholar
Klein, S. L. & Nelson, R. J. (1998b). Adaptive immune responses are linked to the mating system of arvicoline rodents. American Naturalist, 151, 59–67.CrossRefGoogle Scholar
Klompen, J. S. H., Black, W. C., Keirans, J. E. & Oliver, J. H. (1996). Evolution of ticks. Annual Review of Entomology, 41, 141–161.CrossRefGoogle ScholarPubMed
Knopf, P. M. & Coghlan, R. L. (1989). Maternal transfer of resistance to Schistosoma mansoni. Journal of Parasitology, 75, 398–404.CrossRefGoogle ScholarPubMed
Knülle, W. (1967). Physiological properties and biological implications of the water vapour sorption mechanism in larvae of the oriental rat flea, Xenopsylla cheopis (Roths). Journal of Insect Physiology, 13, 333–357.CrossRefGoogle Scholar
Koehler, P. G., Leppla, N. C. & Patterson, S. (1990). Circadian rhythm in the cat flea Ctenocephalides felis (Siphonaptera: Pulicidae). In Chronobiology: Its Role in Clinical Medicine, General Biology, and Agriculture, part B, ed. Hayes, D. K., Pauly, J. E. & Reiter, R. J.. New York: Wiley-Liss, pp. 661–665.Google Scholar
Kolpakova, S. A. (1950). Migration of fleas from the burrows of the great gerbil. Ectoparasites, 2, 115–128 (in Russian).Google Scholar
Kondrashkina, K. I. & Dudnikova, A. F. (1962). Oxygen consumption in fleas parasitic on the ground squirrels as a physiological test of their viability. In Particularly Dangerous and Natural Diseases, ed. Anonymous. Moscow, USSR: Medgiz, pp. 63–69 (in Russian).Google Scholar
Kondrashkina, K. I. & Dudnikova, A. F. (1968). Oxygen consumption in fleas parasitic on the Norway rats. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 87–91 (in Russian).Google Scholar
Kondrashkina, K. I. & Gerasimova, N. G. (1971). Dependence of the metabolic rate of two common fleas parasitic on the great gerbil Rhombomys opimus on their physiological conditions. Problems of Particularly Dangerous Diseases, 17, 32–37 (in Russian).Google Scholar
Konkova, K. V. & Timofeeva, A. A. (1970). Studies of fleas (Siphonaptera) of the Sakhalin and Kuril Islands. In Vectors of Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 371–390 (in Russian).Google Scholar
Konnov, N. P., Demchenko, T. A., Anisimov, P. I., Kondrashkina, K. I. & Lukyanova, A. D. (1986). Pathogenic effect of the plague microbe on the flea Xenopsylla cheopis and ultrastructure of the agent at various periods of its stay in the vector. Parazitologiya, 20, 19–22 (in Russian).Google Scholar
Korallo, N. P., Vinarski, M. V., Krasnov, B. R., et al. (2007). Are there general rules governing parasite diversity? Small mammalian hosts and gamasid mite assemblages. Diversity and Distributions, 13, 353–360.CrossRefGoogle Scholar
Korneeva, L. A. & Sadovenko, E. V. (1990). Patterns of development of the fleas of the great gerbils in the laboratory. In Advantages of Medical Entomology and Acarology in the USSR, ed. Medvedev, G. S.. Leningrad, USSR: All-Union Entomological Society and Zoological Institute, Academy of Sciences of the USSR, pp. 12–14 (in Russian).Google Scholar
Korovin, E. P. (1961). Vegetation of Middle Asia and Southern Kazakhstan. Tashkent, USSR: Academy of Sciences of the Uzbek SSR Press (in Russian).Google Scholar
Korpimaki, E., Tolonen, P. & Bennett, G. F. (1995). Blood parasites, sexual selection and reproductive success of European kestrels. Ecoscience, 2, 335–343.CrossRefGoogle Scholar
Koshkin, S. M. (1966). Materials on flea fauna in Sovetskaya Gavan. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 242–248 (in Russian).Google Scholar
Kosminsky, R. B. (1960). Reproduction of fleas parasitic on mice in the field and experimental conditions. In Parasitological Problems: Proceedings of the 3rd Parasitological Conference of the Ukrainian SSR, ed. Anonymous. Kiev, USSR: Insitute of Zoology of the Academy of Sciences of the Ukrainian SSR, pp. 326–328 (in Russian).Google Scholar
Kosminsky, R. B. (1961). Study of bionomics of the fleas parasitic on the house mice. Unpublished Ph.D. thesis, Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, Stavropol, USSR (in Russian).
Kosminsky, R. B. (1962). Characteristics of the life cycles of fleas parasitic on the house mice in the buildings of a steppe settlement in the Stavropol Region. In Problems of Ecology, vol. 4, ed. Anonymous. Kiev, USSR: Kiev University Press, pp. 65–66 (in Russian).Google Scholar
Kosminsky, R. B. (1965). Feeding and reproduction in fleas parasitic on the house mice in the field and under experiment. Zoologicheskyi Zhurnal, 44, 1372–1375 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1974a). On gonotrophic activity of some fleas belonging to the genus Ctenophthalmus Kolenati (Siphonaptera). In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 154–156 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1974b). On biology of Amphipsylla rossica Wagn., 1912 (Ceratophyllidae, Siphonaptera): a flea with all-year-round activity. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 156–158 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1975a). Age-related changes in the imago of Ctenophthalmus wagneri Tifl, 1928 (Ctenophthalmidae, Siphonaptera). Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 44, 96–100 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1975b). Feeding behavior and reproduction of Ctenophthalmus wagneri Tifl., 1928 (Ctenophthalmidae, Siphonaptera) under experimental conditions. Parazitologiya, 9, 265–270 (in Russian).Google Scholar
Kosminsky, R. B. & Udovitskaya, E. Y. (1975). Upper temperature boundary of life of imago of some species of fleas (Siphonaptera). Proceedings of the Academy of Sciences of the USSR [Doklady Akademii Nauk SSSR], 222, 500–503 (in Russian).Google Scholar
Kosminsky, R. B., Avetisyan, G. A. & Talybov, A. N. (1970). Annual cycle of Ctenophthalmus wladimiri Isajeva-Gurvich, 1948: common flea species of the common vole in the south-east of Trans-Caucasian Upland. In Vectors of Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 79–107 (in Russian).Google Scholar
Kosminsky, R. B., Bryukhanova, L. V., Darskaya, N. F., et al. (1974). Annual cycle of Ceratophyllus (Nosopsyllus) consimilis Wagn., 1898 (Ceratophyllidae, Siphonaptera) in the Stavropol Upland. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 152–154 (in Russian).Google Scholar
Kosoy, M. Y., Regnery, R., Tzianabos, T., et al. (1997). Distribution, diversity, and host specificity of Bartonella in rodents from the southern United States. American Journal of Tropical Medicine and Hygiene, 57, 578–588.CrossRefGoogle Scholar
Kosoy, M. Y., Mandel, E., Green, D., Marston, E. L. & Childs, J. E. (2004). Prospective studies of Bartonella of rodents. I. Demographic and temporal patterns in population dynamics. Vector-Borne and Zoonotic Diseases, 4, 285–95.CrossRefGoogle ScholarPubMed
Kotler, B. P., Brown, J. S. & Subach, A. (1993). Mechanisms of species coexistence of optimal foragers: temporal partitioning by two species of sand dune gerbils. Oikos, 67, 548–556.CrossRefGoogle Scholar
Kotti, B. K. & Kovalevsky, I. (1995). Fleas parasitic on small mammals in the area between the Amur and Bureya Rivers. Zoologicheskyi Zhurnal, 74, 70–76 (in Russian).Google Scholar
Kouki, J. & Hanski, I. (1995). Population aggregation facilitates coexistence of many competing carrion fly species. Oikos, 72, 223–227.CrossRefGoogle Scholar
Kozlovskaya, O. L. (1958). Flea (Aphaniptera) fauna of rodents of the valley of the Ussury River in the Khabarovsk Region. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 109–116 (in Russian).Google Scholar
Kozlovskaya, O. L. & Demidova, A. A. (1958). Data on the ecology of fleas parasitic on the field mouse in the Khabarovsk Region. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 59–64 (in Russian).Google Scholar
Krämer, F. & Mencke, N. (2001). Flea Biology and Control: The Biology of the Cat Flea – Control and Prevention with Imidacloprid in Small Animals. New York: Springer-Verlag.CrossRefGoogle Scholar
Krampitz, H. E. (1964). Über Vorkomen und Verhalten von Haemococcidien der Gattung Hepatozoon Miller, 1908 (Protozoa, Adeloidea) in mittel- und südeuropäischen Säugern. Acta Tropica, 21, 114–154.Google Scholar
Krampitz, H. E. (1981). Development of Hepatozoon erhardovae Krampitz, 1964, (Protozoa: Haemogregarinidae) in experimental mammalian and arthropod hosts. II. Sexual development in fleas and sporozoite indices in xenodiagnosis. Transactions of the Royal Society of Tropical Medicine and Hygiene, 75, 155–157.CrossRefGoogle ScholarPubMed
Krasnov, B. R. & Khokhlova, I. S. (2001). The effect of behavioural interactions on the exchange of flea (Siphonaptera) between two rodent species. Journal of Vector Ecology, 26, 181–190.Google ScholarPubMed
Krasnov, B. R. & Knyazeva, T. V. (1983). Ectoparasite exchange between the midday jird, Meriones meridianus, and the house mouse, Mus musculus, under experiment. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 243–244 (in Russian).Google Scholar
Krasnov, B. R. & Shenbrot, G. I. (2002). Coevolutionary events in history of association of jerboas (Rodentia: Dipodidae) and their flea parasites. Israel Journal of Zoology, 48, 331–350.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S., Degen, A. A. & Rogovin, K. V. (1996a). On the biology of Sundevall's jird (Meriones crassus Sundevall) in Negev Highlands, Israel. Mammalia, 60, 375–391.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Ivanitskaya, E. Y. (1996b). Spatial structure of rodent community in the Ramon erosion cirque, Negev Highlands (Israel). Journal of Arid Environments, 32, 319–327.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Medvedev, S. G., Vatschenok, V. S. & Khokhlova, I. S. (1997). Host–habitat relations as an important determinant of spatial distribution of flea assemblages (Siphonaptera) on rodents in the Negev Desert. Parasitology, 114, 159–173.CrossRefGoogle ScholarPubMed
Krasnov, B. R., Shenbrot, G. I., Medvedev, S. G., Khokhlova, I. S. & Vatschenok, V. S. (1998). Habitat-dependence of a parasite–host relationship: flea assemblages in two gerbil species of the Negev Desert. Journal of Medical Entomology, 35, 303–313.CrossRefGoogle ScholarPubMed
Krasnov, B. R., Hastriter, M., Medvedev, S. G., et al. (1999). Additional records of fleas (Siphonaptera) on wild rodents in the southern part of Israel. Israel Journal of Zoology, 45, 333–340.Google Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2001a). The effect of temperature and humidity on the survival of pre-imaginal stages of two flea species (Siphonaptera: Pulicidae). Journal of Medical Entomology, 38, 629–637.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2001b). Development rates of two Xenopsylla flea species in relation to air temperature and humidity. Medical and Veterinary Entomology, 15, 249–258.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Oguzoglu, I. & Burdelova, N. V. (2002a). Host discrimination by two desert fleas using an odour cue. Animal Behaviour, 64, 33–40.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2002b). The effect of substrate on survival and development of two species of desert fleas (Siphonaptera: Pulicidae). Parasite, 9, 135–142.CrossRefGoogle Scholar
Krasnov, B. R., Burdelova, N. V., Shenbrot, G. I. & Khokhlova, I. S. (2002c). Annual cycles of four flea species (Siphonaptera) in the central Negev Desert. Medical and Veterinary Entomology, 16, 266–276.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2002d). Time to survival under starvation in two flea species (Siphonaptera: Pulicidae) at different air temperatures and relative humidities. Journal of Vector Ecology, 27, 70–81.Google Scholar
Krasnov, B. R., Khokhlova, I. S. & Shenbrot, G. I. (2002e). The effect of host density on ectoparasite distribution: an example with a desert rodent parasitized by fleas. Ecology, 83, 164–175.CrossRefGoogle Scholar
Krasnov, B. R., Burdelov, S. A., Khokhlova, I. S. & Burdelova, N. V. (2003a). Sexual size dimorphism, morphological traits and jump performance in seven species of desert fleas (Siphonaptera). Journal of Zoology, 261, 181–189.CrossRefGoogle Scholar
Krasnov, B. R., Sarfati, M., Arakelyan, M. S., et al. (2003b). Host-specificity and foraging efficiency in blood-sucking parasite: feeding patterns of a flea Parapulex chephrenis on two species of desert rodents. Parasitology Research, 90, 393–399.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S. & Shenbrot, G. I. (2003c). Density-dependent host selection in ectoparasites: an application of isodar theory to fleas parasitizing rodents. Oecologia, 134, 365–373.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Burdelova, N. V., Mirzoyan, N. S. & Degen, A. A. (2004a). Fitness consequences of density-dependent host selection in ectoparasites: testing reproductive patterns predicted by isodar theory in fleas parasitizing rodents. Journal of Animal Ecology, 73, 815–820.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Burdelov, S. A. & Fielden, L. J. (2004b). Metabolic rate and jumping performance in seven species of desert fleas. Journal of Insect Physiology, 50, 149–156.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2004c). Relationships between parasite abundance and the taxonomic distance among a parasite's host species: an example with fleas parasitic on small mammals. International Journal for Parasitology, 34, 1289–1297.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I. & Khokhlova, I. S. (2004d). Sampling fleas: the reliability of host infestation data. Medical and Veterinary Entomology, 18, 232–240.CrossRefGoogle Scholar
Krasnov, B. R., Mouillot, D., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2004e). Geographical variation in host specificity of fleas (Siphonaptera) parasitic on small mammals: the influence of phylogeny and local environmental conditions. Ecography, 27, 787–797.CrossRefGoogle Scholar
Krasnov, B. R., Poulin, R., Shenbrot, G. I., Mouillot, D. & Khokhlova, I. S. (2004f). Ectoparasitic ‘jacks-of-all-trades’: relationship between abundance and host specificity in fleas (Siphonaptera) parasitic on small mammals. American Naturalist, 164, 506–515.Google Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Degen, A. A. (2004g). Flea species richness and parameters of host body, host geography and host ‘milieu’. Journal of Animal Ecology, 73, 1121–1128.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Degen, A. A. (2004h). Relationship between host diversity and parasite diversity: flea assemblages on small mammals. Journal of Biogeography, 31, 1857–1866.CrossRefGoogle Scholar
Krasnov, B. R., Poulin, R., Shenbrot, G. I., Mouillot, D. & Khokhlova, I. S. (2005a). Host specificity and geographic range in haematophagous ectoparasites. Oikos, 108, 449–456.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2005b). Diversification of ectoparasite assemblages and climate: an example with fleas parasitic on small mammals. Global Ecology and Biogeography, 14, 167–175.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S., Hawlena, H., Khokhlova, I. S. & Shenbrot, G. I. (2005c). Sex-biased parasitism, seasonality and sexual size dimorphism in desert rodents. Oecologia, 146, 209–217.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Arakelyan, M. S. & Degen, A. A. (2005d). Is a starving host tastier? Reproduction in fleas parasitizing food limited rodents. Functional Ecology, 19, 625–631.CrossRefGoogle Scholar
Krasnov, B. R., Burdelova, N. V., Khokhlova, I. S., Shenbrot, G. I. & Degen, A. A. (2005e). Pre-imaginal interspecific competition in two flea species parasitic on the same rodent host. Ecological Entomology, 30, 146–155.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Mouillot, D., Khokhlova, I. S. & Poulin, R. (2005f). What are the factors determining the probability of discovering a flea species (Siphonaptera)?Parasitology Research, 97, 228–237.CrossRefGoogle Scholar
Krasnov, B. R., Mouillot, D., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2005g). Abundance patterns and coexistence processes in communities of fleas parasitic on small mammals. Ecography, 28, 453–464.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S., Khokhlova, I. S., Shenbrot, G. I. & Hawlena, H. (2005h). Abundance and distribution of fleas on desert rodents: linking Taylor's power law to ecological specialization and epidemiology. Parasitology, 131, 825–837.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Miklisova, D. & Morand, S. (2005i). Distribution of fleas (Siphonaptera) among small mammals: mean abundance predicts prevalence via simple epidemiological model. International Journal for Parasitology, 35, 1097–1101.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2005j). Nested pattern in flea assemblages across the host's geographic range. Ecography, 28, 475–484.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Mouillot, D., Khokhlova, I. S. & Poulin, R. (2005k). Spatial variation in species diversity and composition of flea assemblages in small mammalian hosts: geographic distance or faunal similarity?Journal of Biogeography, 32, 633–644.CrossRefGoogle Scholar
Krasnov, B. R., Mouillot, D., Khokhlova, I. S., Shenbrot, G. I. & Poulin, R. (2005l). Covariance in species diversity and facilitation among non-interactive parasite taxa: all against the host. Parasitology, 131, 557–568.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M. & Morand, S. (2006a). Age-dependent flea (Siphonaptera) parasitism in rodents: a host's life history matters. Journal of Parasitology, 92, 242–248.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S., Mouillot, D., et al. (2006b). Resource predictability and host specificity in fleas: the effect of host body mass. Parasitology, 133, 81–88.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Mouillot, D., Khokhlova, I. S. & Poulin, R. (2006c). Ecological characteristics of flea species relate to their suitability as plague vectors. Oecologia, 149, 474–481.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2006d). Is abundance a species attribute? An example with haematophagous ectoparasites. Oecologia, 150, 132–140.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Miklisova, D. & Morand, S. (2006e). Host specificity, parasite community size and the relation between abundance and its variance. Evolutionary Ecology, 20, 75–91.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Khokhlova, I. S., et al. (2006f). Aggregation and species coexistence in fleas parasitic on small mammals. Ecography, 29, 159–168.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S., Hawlena, H. & Degen, A. A. (2006g). Temporal variation in parasite infestation of a host individual: does a parasite-free host remain uninfested permanently?Parasitology Research, 99, 541–545.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M. & Morand, S. (2006h). Are ectoparasite communities structured? Species co-occurrence, temporal variation and null models. Journal of Animal Ecology, 75, 1330–1339.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Khokhlova, I. S., et al. (2006i). Relationships between local and regional species richness in flea communities of small mammalian hosts: saturation and spatial scale. Parasitology Research, 98, 403–413.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S. & Poulin, R. (2006j). Patterns of macroparasite diversity in small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 197–232.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Miklisova, D. & Morand, S. (2006k). Habitat variation in species composition of flea assemblages on small mammals in central Europe. Ecological Research, 21, 460–469.CrossRefGoogle Scholar
Krasnov, B. R., Korine, C., Burdelova, N. V., Khokhlova, I. S. & Pinshow, B. (2007a). Between-host phylogenetic distance and feeding efficiency in haematophagous ectoparasites: rodent fleas and a bat host. Parasitology Research, 101, 365–371.CrossRefGoogle Scholar
Krasnov, B. R., Hovhanyan, A., Khokhlova, I. S. & Degen, A. A. (2007b). Density-dependence and feeding success in haematophagous ectoparasites. Parasitology, 134, 1379–1386.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2007c). Geographic variation in the ‘bottom-up’ control of diversity: fleas and their small mammalian hosts. Global Ecology and Biogeography, 16, 179–186.CrossRefGoogle Scholar
Krebs, C. J. (1994). Ecology: The Experimental Analysis of Distribution and Abundance, 4th edn. New York: HarperCollins.Google Scholar
Krebs, C. J. (1996). Population cycles revisited. Journal of Mammalogy, 77, 8–24.CrossRefGoogle Scholar
Kristan, D. M. (2002). Maternal and direct effects of the intestinal nematode Heligmosomoides polygyrus on offspring growth and susceptibility to infection. Journal of Experimental Biology, 205, 3967–3977.Google ScholarPubMed
Kristensen, N. P. (1975). The phylogeny of hexapod ‘orders’: a critical review of recent accounts. Zeitschrift für zoologische Systematik und Evolutionsforschung, 13, 1–44.CrossRefGoogle Scholar
Kristensen, N. P. (1981). Phylogeny of insect orders. Annual Review of Entomology, 26, 135–157.CrossRefGoogle Scholar
Krivokhatsky, V. A. (1984). Seasonal changes in the distribution of fleas in burrow passages as indicator of their migration activity. Parazitologiya, 18, 150–153 (in Russian).Google Scholar
Krylov, D. G. (1986). On fauna and ecology of fleas parasitic on small mammals from the Moscow Region, Russian Federation, USSR. Parazitologiya, 20, 356–363 (in Russian).Google Scholar
Kucheruk, V. V. (1983). Mammal burrows: their structure, topology and use. Fauna and Ecology of Rodents, 15, 5–54 (in Russian).Google Scholar
Kucheruk, V. V., Kulik, I. L. & Dubrovsky, Y. A. (1972). The great gerbil as a desert life form. Fauna and Ecology of Rodents, 11, 5–70 (in Russian).Google Scholar
Kulakova, Z. G. (1962). On the role of fleas in transmission of the tick-borne encephalitis (experimental results). Bulletin of the Moscow Naturalist Society, Series Biology, 67, 144–145 (in Russian).Google Scholar
Kulakova, Z. G. (1964). Feeding of Xenopsylla gerbilli caspica and some other flea species. Ectoparasites, 4, 205–219 (in Russian).Google Scholar
Kumar, R. & Kumar, R. (1996). Cross-resistance to Hyalomma anatolicum anatolicum ticks in rabbits immunized with midgut antigens of Hyalomma dromedarii. Indian Journal of Animal Sciences, 66, 657–661.Google Scholar
Kunitskaya, N. T. (1960). Study of the reproductive organs of female fleas and identification of their physiological age. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 29, 688–701 (in Russian).Google Scholar
Kunitskaya, N. T. (1970). Structure of the ovaries in fleas. Parazitologiya, 4, 444–450 (in Russian).Google Scholar
Kunitskaya, N. T., Gauzshtein, D. M., Kunitsky, V. N., Rodionov, I. A. & Filimonov, V. I. (1965a). Feeding activity of fleas parasitic on the great gerbil in experiments. In Proceedings of the 4th Scientific Conference on Natural Focality and Prophylaxis of Plague, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 135–137 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1965b). On the reproduction of fleas parasitic on the great gerbil. In Proceedings of the 4th Scientific Conference on Natural Focality and Prophylaxis of Plague, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 137–138 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1969). Reproduction and age structure of populations of fleas from genera Coptopsylla and Paradoxopsyllus in the southern Balkhash region. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 74–76 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N., Gauzshtein, D. M., Morozova, I. V. & Savelova, N. M. (1971). Age composition of imago in populations of Xenopsylla gerbilli and Xenopsylla hirtipes in the southern Pri-Balkhashie. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 387–389 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1974). On fecundity of fleas Xenopsylla gerbilli in experiments. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 323–325 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1977). Phenological age of fleas and an attempt to analyze age composition of natural populations of Xenopsylla gerbilli Wagn. Parazitologiya, 11, 202–210 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N., Gauzshtein, D. M. & Savelova, N. M. (1979). Spatial distribution of the larvae of fleas parasitic on the great gerbil in the host's burrow. Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 107–110 (in Russian).Google Scholar
Kunitsky, V. N. (1961). On the environmental conditions of fleas parasitic on gerbils in the southeastern Azerbaijan SSR. Zoologicheskyi Zhurnal, 40, 848–858 (in Russian).Google Scholar
Kunitsky, V. N. (1970). Essay on the comparative ecology of fleas parasitic on gerbils in southwestern Azerbaijan. In Vectors of Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 153–227 (in Russian).Google Scholar
Kunitsky, V. N. & Kunitskaya, N. T. (1962). Fleas of southwestern Azerbaijan. Proceedings of the Azerbaijanian Anti-Plague Station, 3, 156–169 (in Russian).Google Scholar
Kunitsky, V. N., Kunitskaya, N. T., Gauzshtein, D. M. & Podkovyrova, T. S. (1963). Reproduction and development of Ceratophyllus laeviceps Wagn. in natural and experimental conditions. In Proceedings of Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Anonymous. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 126–128 (in Russian).Google Scholar
Kunitsky, V. N., Gauzshtein, D. M. & Kunitskaya, N. T. (1971a). On the effect of the soil moisture of the longevity of fleas under low ambient temperatures. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 395–397 (in Russian).Google Scholar
Kunitsky, V. N., Volkov, V. M., Lelikova, Z. F., et al. (1971b). Changes in the numbers of fleas in the artificially decreased populations of the great gerbil in the northeastern part of the Caspian Lowland. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 392–395 (in Russian).Google Scholar
Kunitsky, V. N., Volkov, V. M., Lelikova, Z. F. & Agunkova, O. S. (1974). On annual number of generations of Xenopsylla skrjabini in the Caspian Lowlands. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 328–3330 (in Russian).Google Scholar
Kuris, A. M., Blaustein, A. R. & Aho, J. J. (1980). Hosts as islands. American Naturalist, 116, 570–586.CrossRefGoogle Scholar
Kusiluka, L. J. M., Kambarage, D. M., Matthewman, R. W., Daborn, C. J. & Harrison, L. J. S. (1995). Prevalence of ectoparasites of goats in Tanzania. Journal of Applied Animal Research, 7, 69–74.CrossRefGoogle Scholar
Kuznetsov, A. A. & Matrosov, A. N. (2003). The use of individual marking of fleas (Siphonaptera) for studies of their dispersal by hosts. Zoologicheskyi Zhurnal, 82, 964–972 (in Russian).Google Scholar
Kuznetsov, A. A., Matrosov, A. N., Nikitin, P. N. & Eigelis, S. Y. (1993). Method of individual marking of fleas and the results of the application of this method for the study of the dispersal of ectoparasites of gerbils in the Volga-Ural Sands. Problems of Particularly Dangerous Diseases, 73, 58–64 (in Russian).Google Scholar
Kuznetsov, A. A., Matrosov, A. N., Chyong, L. T. V. & Dat, D. T. (1999). Movements of the synantropous rats and their fleas in the settlements of the southern Vietnam. Problems of Particularly Dangerous Diseases, 79, 59–65 (in Russian).Google Scholar
Kyriazakis, I., Tolkamp, B. J. & Hutchings, M. R. (1998). Towards a functional explanation for the occurrence of anorexia during parasitic infections. Animal Behaviour, 56, 265–274.CrossRefGoogle ScholarPubMed
Labandeira, C. C. (1997). Insect mouthparts: ascertaining the paleobiology of insect feeding strategies. Annual Review of Ecology and Systematics, 28, 153–193.CrossRefGoogle Scholar
Labunets, N. F. (1967). Zoogeographic characteristics of the western Khangay. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 27, 231–240 (in Russian).Google Scholar
Lafferty, K. D. & Kuris, A. M. (2000). Parasite–host modelling meets reality: adaptive peaks and their ecological attributes. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R., Morand, S. & Skorping, A.. Amsterdam, the Netherlands: Elsevier Science, pp. 9–26.Google Scholar
Lafferty, K. D. & Kuris, A. M. (2002). Trophic strategies, animal diversity and body size. Trends in Ecology and Evolution, 17, 507–513.CrossRefGoogle Scholar
Lambrechts, M. M. & Santos, Dos A. (2000). Aromatic herbs in Corsican blue tit nests: the ‘Potpourri’ hypothesis. Acta Oecologica, 21, 175–178.CrossRefGoogle Scholar
Lang, J. D. (1996). Factors affecting the seasonal abundance of ground squirrel and wood rat fleas (Siphonaptera) in San Diego County, California. Journal of Medical Entomology, 33, 790–804.CrossRefGoogle Scholar
Lapchin, L. & Guillemaud, T. (2005). Asymmetry in host and parasitoid diffuse coevolution: when the Red Queen has to keep a finger in more than one pie. Frontiers in Zoology, 2, 4. doi:10.1186/1742-9994-2-4.CrossRefGoogle Scholar
Lareschi, M. (2006). The relationship of sex and ectoparasite infestation in the water rat Scapteromys aquaticus (Rodentia: Cricetidae) in La Plata, Argentina. Revista de Biologia Tropical, 54, 673–679.CrossRefGoogle Scholar
Larrivee, D. H., Benjamini, E., Feingold, B. F. & Shimizu, M. (1964). Histologic studies of guinea pig skin: different stages of allergic reactivity to flea bites. Experimental Parasitology, 15, 491–502.CrossRefGoogle ScholarPubMed
Larsen, K. S. (1995). Laboratory rearing of the squirrel flea Ceratophyllus sciurorum sciurorum with notes on its biology. Entomologia Experimentalis et Applicata, 76, 241–245.CrossRefGoogle Scholar
Larson, O. R. (1973). North Dakota fleas. IV. Cold tolerance in the bird flea Ceratophillus idius (Jordan and Rothschild). Proceedings of the North Dakota Academy of Sciences, 26, 51–55.Google Scholar
Lauer, D. M. & Sonenshine, D. E. (1978). Bionomics of the squirrel flea, Orchopeas howardi (Siphonaptera: Ceratophyllidae), in laboratory and field colonies of the southern flying squirrel, Glaucomys volans, using radiolabelling technique. Journal of Medical Entomology, 15, 1–10.CrossRefGoogle Scholar
Launay, H. (1989). Facteurs écologiques influençant la répartition et la dynamique des populations de Xenopsylla cunicularis Smit, 1957 (Insecta: Siphonaptera) puce inféodée au lapin de garenne, Oryctolagus cuniculus (L.). Vie et Milieu, 39, 111–120.Google Scholar
Lavoipierre, M. M. J., Radovsky, F. J. & Budwiser, P. D. (1979). The feeding process of a tungid flea, Tunga monositus (Siphonaptera: Tungidae), and its relationship to the host inflammatory and repair response. Journal of Medical Entomology, 15, 187–217.CrossRefGoogle Scholar
Lawrence, W. & Foil, L. D. (2002). The effect of diet upon pupal development and cocoon formation by the cat flea (Siphonaptera: Pulicidae). Journal of Vector Ecology, 27, 39–43.Google Scholar
Lawton, J. H. & Hassell, M. P. (1981). Asymmetrical competition in insects. Nature, 289, 793–795.CrossRefGoogle Scholar
Layene, J. N. (1954). The biology of the red squirrel, Tamiasciurus hudsonicus loquax (Bangs), in central New York. Ecological Monographs, 24, 227–267.CrossRefGoogle Scholar
Layene, J. N. (1963). A study of the parasites of the Florida mouse, Peromyscus floridanus, in relation to host and environmental factors. Tulane Studies in Zoology and Botany, 11, 1–27.Google Scholar
Lee, C. Y., Alexander, P. S., Yang, V. V. C. & Yu, J. Y. L. (2001). Seasonal reproductive activity of male formosan wood mice (Apodemus semotus): relationships to androgen levels. Journal of Mammalogy, 82, 700–708.2.0.CO;2>CrossRefGoogle Scholar
Lee, P. L. M. & Clayton, D. H. (1995). Population biology of swift (Apus apus) ectoparasites in relation to host reproductive success. Ecological Entomology, 20, 43–50.CrossRefGoogle Scholar
Lee, S. E., Johnstone, I. P., Lee, R. P. & Opdebeeck, J. P. (1999). Putative salivary allergens of the cat flea, Ctenocephalides felis. Veterinary Immunology and Immunopathology, 69, 229–237.CrossRefGoogle ScholarPubMed
Lee, W. B. & Houston, D. C. (1993). The effect of diet quality on gut anatomy in British voles (Microtinae). Journal of Comparative Physiology B, 163, 337–339.CrossRefGoogle Scholar
Leeson, H. S. (1936). Further experiments upon the longevity of Xenopsylla cheopis Roths. (Siphonaptera). Parasitology, 28, 403–409.CrossRefGoogle Scholar
Legendre, P., Galzin, R. & Harmelin-Vivien, M. L. (1997). Relating behavior to habitat: solutions to the fourth-corner problem. Ecology, 78, 547–562.Google Scholar
Legendre, P., Desdevises, Y. & Bazin, E. (2002). A statistical test for host–parasite coevolution. Systematic Biology, 51, 217–234.CrossRefGoogle ScholarPubMed
Lehane, M. (2005). The Biology of Blood-Sucking in Insects, 2nd edn. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Lehmann, T. (1992). Reproductive activity of Synosternus cleopatrae (Siphonaptera: Pulicidae) in relation to host factors. Journal of Medical Entomology, 29, 946–952.CrossRefGoogle ScholarPubMed
Lehmann, T. (1993). Ectoparasites: direct impact on host fitness. Parasitology Today, 9, 8–13.CrossRefGoogle ScholarPubMed
Lehmann, T. (1994). Reinfestation analysis to estimate ectoparasite population size, emergence and mortality. Journal of Medical Entomology, 31, 257–264.CrossRefGoogle ScholarPubMed
Leonov, Y. A. (1958). Fleas parasitic on rodents in the southern part of Primorie (Far East). Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 147–152 (in Russian).Google Scholar
Lepage, D. (2006). Avibase: The World Bird Database. Available online at www.bsc-eoc.org/avibase/avibase.jsp?pg=home&lang=EN
Letov, G. S., Emelianova, N. D., Letova, G. I. & Sulimov, A. D. (1966). Rodents and their ectoparasites in the settlements of Tuva. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 270–276 (in Russian).Google Scholar
Letova, G. I., Letov, G. S. & Mamontova, E. V. (1969). Parasitological description of the Mungun-Taigin part of the Altai plague focus. Problems of Particularly Dangerous Infections, 10, 55–60 (in Russian).Google Scholar
Levenbook, L. (1985). Storage proteins. In Comprehensive Insect Biochemistry, Physiology and Pharmacology, ed. Gilbert, L. I. & Kerkut, G.. Oxford, UK: Pergamon Press, pp. 307–346.Google Scholar
Levine, J. M. (1999). Indirect facilitation: evidence and predictions from a riparian community. Ecology, 80, 1762–1769.CrossRefGoogle Scholar
Levine, J. M. & Rees, M. (2002). Coexistence and relative abundance in annual plant assemblages: the roles of competition and colonization. American Naturalist, 160, 452–467.CrossRefGoogle ScholarPubMed
Levins, R. (1968). Evolution in Changing Environments. Princeton, NJ: Princeton University Press.Google Scholar
Lewis, R. E. (1998). Résumé of the Siphonaptera (Insecta) of the world. Journal of Medical Entomology, 35, 377–389.CrossRefGoogle Scholar
Lewis, R. E. & Eckerlin, R. P. (2004). A new species of Hystrichopsylla Taschenberg, 1880 (Siphonaptera: Hystrichopsyllidae) from Guatemala. Proceedings of the Entomological Society of Washington, 106, 757–760.Google Scholar
Lewis, R. E. & Grimaldi, D. (1997). A pulicid flea in Miocene amber from the Dominican Republic (Insecta: Siphonaptera: Pulicidae). American Museum Novitates, 3205, 1–9.Google Scholar
Lewis, R. E. & Haas, G. E. (2001). A review of the North American Catallagia Rothschild, 1915, with the description of a new species (Siphonaptera: Ctenophthalmidae: Neopsyllinae: Phalacropsyllini). Journal of Vector Ecology, 26, 51–69.Google Scholar
Lewis, R. E. & Lewis, J. H. (1990). An annotated checklist of the fleas (Siphonaptera) of the Middle East. Fauna of Saudi Arabia, 11, 251–277.Google Scholar
Lewis, R. E. & Stone, E. (2001). Psittopsylla mexicana, a new genus and species of bird flea from Chihuahua, Mexico (Siphonaptera: Ceratophyllidae: Ceratophyllinae). Journal of the New York Entomological Society, 109, 360–366.CrossRefGoogle Scholar
Lewis, R. E., Lewis, J. H. & Maser, C. (1988). The Fleas of the Pacific Northwest. Corvallis, OR: Oregon State University Press.Google Scholar
Li, B.-G., Zhang, P., Watanabe, K. & Tan, C.-L. (2002). Does allogrooming serve a hygienic function in the Sichuan snub-nosed monkey (Rhinopithecus roxellana)?Acta Zoologica Sinica, 48, 705–715 (in Chinese).Google Scholar
Li, W., Wang, Z.-Y., Wang, C., Zhang, Z.-J. & Ye, R.-Y. (2004). Observations on the life cycle of Neopsylla teratura Rothschild, 1913 in the laboratory. Endemic Diseases Bulletin, 19, 8–9 (in Chinese).Google Scholar
Li, Z.-L. & Ma, L.-M. (1999). Aggregation degree of distribution of two fleas, Citellophilus tesquorum sungaris (Jordan) and Neopsylla bidentatiformis Wagner, on host body. Entomological Knowledge, 36, 89–91 (in Chinese).Google Scholar
Li, Z.-L. & Zhang, Y.-X. (1997). Analysis on the yearly dynamics relation between body flea index and population of Citellus dauricus. Acta Entomologica Sinica, 40, 166–170 (in Chinese).Google Scholar
Li, Z.-L. & Zhang, Y.-X. (1998). The yearly dynamics relationship between burrow nest flea index and population of Citellus dauricus. Acta Entomologica Sinica, 41, 77–81 (in Chinese).Google Scholar
Li, Z.-L., Zhang, W.-R. & Ma, L.-M. (1995). Analysis of the relations among flea index, populations of Meriones unguiculatus and meteorological factors. Acta Entomologica Sinica, 38, 442–447 (in Chinese).Google Scholar
Li, Z.-L., Zhang, W.-R. & Yan, W.-L. (2000). Studies on dynamics of body and burrow nest fleas of Meriones unguiculatus. Acta Entomologica Sinica, 43, 58–63 (in Chinese).Google Scholar
Li, Z.-L., Yang, Y. & Chen, S.-G. (2001a). An analysis on the population dynamics of fleas in the man-made plague focuses of Harbin suburbs. Acta Entomologica Sinica, 44, 507–511 (in Chinese).Google Scholar
Li, Z.-L., Liu, T.-C. & Niu, Y. (2001b). Studies on dynamics of body and burrow fleas of Microtus brandti and succession of their community. Acta Entomologica Sinica, 44, 327–331 (in Chinese).Google Scholar
Li, Z.-L., Yu, G.-J. & Chen, D. (2002). Statespace model between Nosopsyllus laeviceps kuzenkovi and population of Meriones unguiculatus. Acta Entomologica Sinica, 45, 132–133 (in Chinese).Google Scholar
Liao, H.-R. & Lin, D.-H. (1993). Laboratorial observation on some biological characters of two rat fleas in south China. Endemic Diseases Bulletin, 8, 61–64 (in Chinese).Google Scholar
Lighton, J. R. B. (1991). Measurements in insects. In Concise Encyclopedia on Biological and Biomedical Measurement Systems, ed. Payne, C. A.. Oxford, UK: Pergamon Press, pp. 201–208.Google Scholar
Lighton, J. R. B., Fielden, L. J. & Rechav, Y. (1993). Discontinuous ventilation in a non-insect, the tick Amblyomma marmoreum (Acari, Ixodidae): characterization and metabolic modulation. Journal of Experimental Biology, 180, 229–245.Google Scholar
Liker, A., Markus, M., Vazar, A., Zemankovics, E. & Rózsa, L. (2001). Distribution of Carnus hemapterus in a starling colony. Canadian Journal of Zoology, 79, 574–580.CrossRefGoogle Scholar
Linardi, P. M. & Guimarães, L. R. (2000). Sifonápteros do Brasil. São Paulo, Brazil: Museu de Zoologia da Universidade de São Paulo.Google Scholar
Linardi, P. M., Botelho, J. R. & Cunha, H. C. (1985). Ectoparasites of rodents of the urban region of Belo Horizonte, MG. II. Variations of the infestation indices in Rattus norvegicus norvegicus. Memórias do Instituto Oswaldo Cruz, 80, 227–232.CrossRefGoogle Scholar
Linardi, P. M., Gomes, A. F., Botelho, J. R. & Lopes, C. M. L. (1994). Some ectoparasites of commensal rodents from Huambo, Angola. Journal of Medical Parasitology, 31, 754–756.Google ScholarPubMed
Linardi, P. M., DeMaria, M. & Botelho, J. R. (1997). Effects of larval nutrition on the postembryonic development of Ctenocephalides felis felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 34, 494–497.CrossRefGoogle Scholar
Lindén, M. (1991). Divorce in great tits: chance or choice? An experimental approach. American Naturalist, 138, 1039–1048.CrossRefGoogle Scholar
Lindsay, L. R. & Galloway, T. D. (1997). Seasonal activity and temporal separation of four species of fleas (Insecta: Siphonaptera) infesting Richardson's ground squirrels, Spermophilus richardsoni (Rodentia: Sciuridae), in Manitoba, Canada. Canadian Journal of Zoology, 75, 1310–1322.CrossRefGoogle Scholar
Lindsay, L. R. & Galloway, T. D. (1998). Reproductive status of four species of fleas (Insecta: Siphonaptera) on Richardson's ground squirrel (Rodentia: Sciuridae) in Manitoba, Canada. Journal of Medical Entomology, 35, 423–430.CrossRefGoogle Scholar
Linley, J. R., Benton, A. H. & Day, J. F. (1994). Ultrastructure of the eggs of seven flea species (Siphonaptera). Journal of Medical Entomology, 31, 813–827.CrossRefGoogle Scholar
Linsdale, J. M. (1947). The California Ground Squirrel. Berkeley, CA: University of California Press.Google Scholar
Linsdale, J. M. & Davis, B. S. (1956). Taxonomic appraisal and occurrence of fleas at the Hastings Reservation in Central California. University of California Publications in Zoology, 54, 293–370.Google Scholar
Linsdale, J. M. & Tevis, L. P. (1951). The Dusky-Footed Wood Rat. Berkeley, CA: University of California Press.Google Scholar
Litvinova, E. A. (2004). On biology and ecology of Ctenophthalmus congeneroides Wagner, 1929 (Siphonaptera: Hystrichopsyllidae): one of the most abundant fleas of rodents in the Primorie Region. A. I. Kurentsov's Annual Memorial Meetings, 15, 104–107 (in Russian).Google Scholar
Liu, J., Li, S.-J., Amin, O. M. & Zhang, Y.-M. (1993). Blood-feeding of the gerbil flea Nosopsyllus laeviceps kuzenkovi (Yagubyants), vector of plague in Inner Mongolia, China. Medical and Veterinary Entomology, 7, 54–58.Google Scholar
Liu, Z.-Y., Wu, H.-Y., Li, G.-Z., et al. (1986). Fauna Sinica. Insecta. Siphonaptera. Beijing: Science Press (in Chinese).Google Scholar
Lively, C. M. (1989). Adaptation by a parasitic nematode to local populations of its snail host. Evolution, 50, 1663–1671.CrossRefGoogle Scholar
Lloyd, M. (1967). Mean crowding. Journal of Animal Ecology, 36, 1–30.CrossRefGoogle Scholar
Lo, C. M., Morand, S. & Galzon, R. (1998). Parasite diversity/host age and size relationship in three coral-reef fish from French Polynesia. International Journal for Parasitology, 28, 1695–1708.CrossRefGoogle ScholarPubMed
Lochmiller, R. L. & Dabbert, C. B. (1993). Immunocompetence, environmental stress, and the regulation of animal populations. Trends in Comparative Biochemistry and Physiology, 1, 823–855.Google Scholar
Lochmiller, R. L. & Deerenberg, C. (2000). Trade-offs in evolutionary immunology: just what is the cost of immunity?Oikos, 88, 87–98.CrossRefGoogle Scholar
Lochmiller, R. L., Vestey, M. R. & Boren, J. C. (1993). Relationship between protein nutritional status and immunocompetence in northern bobwhite chicks. Auk, 110, 503–510.CrossRefGoogle Scholar
Lombardero, M. J., Ayres, M. P., Hofstetter, R. W., Moser, J. C. & Lepzig, K. D. (2003). Strong indirect interactions of Tarsonemus mites (Acarina: Tarsonemidae) and Dendroctonous frontalis. Oikos, 102, 243–252.CrossRefGoogle Scholar
Lomnicki, A. (1988). Population Ecology of Individuals. Princeton, NJ: Princeton University Press.Google ScholarPubMed
López-Sepulcre, A. & Kokko, H. (2005). Territorial defense, territory size, and population regulation. American Naturalist, 166, 317–329.CrossRefGoogle ScholarPubMed
Lorange, E. A., Race, B. L., Sebbane, F. & Hinnebusch, B. J. (2005). Poor vector competence of fleas and the evolution of hypervirulence in Yersinia pestis. Journal of Infectious Diseases, 191, 1907–1912.CrossRefGoogle ScholarPubMed
Loreau, M. (2000). Are communities saturated? On the relationship of S, J and T diversity. Ecology Letters, 3, 73–76.CrossRefGoogle Scholar
Losos, J. B., Leal, M., Glor, R. E., et al. (2003). Niche lability in the evolution of a Caribbean lizard community. Nature, 424, 542–545.CrossRefGoogle ScholarPubMed
Lott, D. F. (1991). Intraspecific Variation in the Social Systems of Wild Vertebrates. Cambridge, UK: Cambridge University Press.Google Scholar
Louw, J. P., Horak, I. G. & Braack, L. E. O. (1993). Fleas and lice on scrub hares (Lepus saxatilis). Onderstepoort Journal of Veterinary Research, 60, 95–101.Google Scholar
Louw, J. P., Horak, I. G., Horak, M. L. & Braack, L. E. O. (1995). Fleas, lice and mites on scrub hares (Lepus saxatilis) in Northern and Eastern Transvaal and in KwaZulu-Natal, South Africa. Onderstepoort Journal of Veterinary Research, 62, 133–137.Google ScholarPubMed
Lu, L. & Wu, H. (2001). The molecular phylogeny of some species of the bidentatiformis group of the genus Neopsylla based on 16s rRNA gene. Acta Entomologica Sinica, 44, 548–554 (in Chinese).Google Scholar
Lu, L. & Wu, H. (2002). The variation of rDNA ITS2 sequences in nine species of Neopsylla (Siphonaptera: Ctenophthalmidae). Acta Parasitologica et Medica Entomologica Sinica, 9, 106–113 (in Chinese).Google Scholar
Lu, L. & Wu, H. (2005). Morphological phylogeny of Geusibia Jordan, 1932 (Siphonaptera: Leptopsyllidae) and the host–parasite relationships with pikas. Systematic Parasitology, 61, 65–78.Google Scholar
Luchetti, A., Mantovani, B., Pampiglione, S. & Trentini, M. (2005). Molecular characterization of Tunga trimamillata and T. penetrans (Insecta, Siphonaptera, Tungidae): taxonomy and genetic variability. Parasite, 12, 123–129.CrossRefGoogle Scholar
Ludwig, D. (1999). Is it meaningful to estimate a probability of extinction?Ecology, 80, 298–310.CrossRefGoogle Scholar
Lukashevich, E. D. & Mostovsky, M. B. (2003). Haematophagous insects in the fossil record. Paleontologicheskyi Zhurnal, 37, 48–56 (in Russian).Google Scholar
Lundqvist, L. (1988). Reproductive strategies of ectoparasites on small mammals. Canadian Journal of Zoology, 66, 774–781.CrossRefGoogle Scholar
Lundqvist, L. & Brinck-Lindroth, G. (1990). Patterns of coexistence: ectoparasites on small mammals in northern Fennoscandia. Holarctic Ecology, 13, 39–49.Google Scholar
Ma, L.-M. (1983). Distribution of fleas in the hair coat of the host. Acta Entomologica Sinica, 26, 409–412 (in Chinese).Google Scholar
Ma, L.-M. (1988). Abundance of fleas in relation to population fluctuations of their hosts. Acta Entomologica Sinica, 31, 50–54 (in Chinese).Google Scholar
Ma, L.-M. (1989). The distribution of fleas on the host body in relation to temperature and the number of fleas. Acta Entomologica Sinica, 32, 68–73 (in Chinese).Google Scholar
Ma, L.-M. (1990). Observations on longevity of adult Neopsylla bidentatiformis Wagner and Citellophilus tesquorum sungaris (Jordan) under different conditions. Entomological Knowledge, 27, 358–359 (in Chinese).Google Scholar
Ma, L.-M. (1993a). Resistance of fleas Neopsylla bidentatiformis and Citellophilus (Citellophilus) tesquorum sungaris to low temperature. Endemic Diseases Bulletin, 8, 71–73 (in Chinese).Google Scholar
Ma, L.-M. (1993b). The sex ratios of some fleas in north China. Acta Entomologica Sinica, 36, 63–66 (in Chinese).Google Scholar
Ma, L.-M. (1994a). Observation on survival of fleas and their hosts under high temperature in field of northern China. Acta Zoologica Sinica, 40, 100–104 (in Chinese).Google Scholar
Ma, L.-M. (1994b). Laboratory studies on fleas Neopsylla bidentatiformis and Citellophilus tesquorum sungaris attacking and leaving from hosts. Acta Entomologica Sinica, 36, 63–66 (in Chinese).Google Scholar
Ma, L.-M. (1995). Activity of Neopsylla bidentatiformis Wagner and Citellophilus tesquorum sungaris Jordan. Entomological Knowledge, 32, 225–227 (in Chinese).Google Scholar
Ma, L.-M. (1997). The monstrosities of fleas reared in laboratory. Entomological Journal of East China, 6, 107–109 (in Chinese).Google Scholar
Ma, L.-M. (2000). Body length of fleas in relation to some factors, and influence of host nutrition on fleas. Acta Parasitologica et Medica Entomologica Sinica, 7, 235–240 (in Chinese).Google Scholar
Ma, L.-M. (2002). Relationship of hunger tolerance to some environmental and physiologic factors in fleas Neopsylla bidentatiformis and Citellophilus tesquorum sungaris. Acta Parasitologica et Medica Entomologica Sinica, 9, 246–248 (in Chinese).Google Scholar
MacArthur, R. H. (1955). Fluctuations in animal populations, and a measure of community stability. Ecology, 36, 533–536.CrossRefGoogle Scholar
MacArthur, R. H. (1972). Geographical Ecology. New York: Harper & Row.Google Scholar
MacArthur, R. H. & Levins, R. (1964). Competition, habitat selection and character displacement in a patchy environment. Proceedings of the National Academy of Sciences of the USA, 51, 1207–1210.CrossRefGoogle Scholar
MacArthur, R. H. & Levins, R. (1967). The limiting similarity of convergence and divergence of coexisting species. American Naturalist, 101, 377–385.CrossRefGoogle Scholar
MacArthur, R. H. & Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacInnis, A. J. (1976). How parasites find hosts: some thoughts on the inception of host–parasite integration. In Ecological Aspects of Parasitology, ed. Kennedy, C. R.. Amsterdam, the Netherlands: North Holland, pp. 3–20.Google Scholar
Madhavi, R. & Anderson, R. M. (1985). Variability in the susceptibility of the fish host, Poecilia reticulata, to infection with Gyrodactylus bullatarudis (Monogenea). Parasitology, 91, 531–544.CrossRefGoogle Scholar
Main, A. J. (1983). Fleas (Siphonaptera) on small mammals in Connecticut, USA. Journal of Medical Entomology, 20, 33–39.CrossRefGoogle ScholarPubMed
Maizels, R. M., Balic, A., Gomez-Escobar, N., et al. (2004). Helminth parasites: masters of regulation. Immunological Reviews, 201, 89–116.CrossRefGoogle Scholar
Makundi, R. H. & Kilonzo, B. S. (1994). Seasonal dynamics of rodent fleas and its implication on control strategies in Lushoto district, north-eastern Tanzania. Journal of Applied Entomology, 118, 165–171.CrossRefGoogle Scholar
Manhert, V. (1972). Zum Auftreten von Kleinsäuger-Flöhen auf ihren Wirten in Abhängigkeit von Jahreszeit und Höhenstufen. Oecologia, 8, 400–418.CrossRefGoogle Scholar
Mans, B. J., Louw, A. I. & 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, 1695–1705.CrossRefGoogle ScholarPubMed
Mappes, T., Mappes, J. & Kotiaho, J. (1994). Ectoparasites, nest-site choice and breeding success in the pied flycatcher. Oecologia, 98, 147–149.CrossRefGoogle ScholarPubMed
Margalit, Y. & Shulov, A. S. (1972). Effect of temperature on development of prepupa and pupa of the rat flea, Xenopsylla cheopis Rothschild. Journal of Medical Entomology, 9, 117–125.CrossRefGoogle ScholarPubMed
Margolis, L., Esch, G. W., Holmes, J. C., Kuris, A. M. & Schad, G. A. (1982). The use of ecological terms in parasitology (report of an ad hoc committee of the American Society of Parasitologists). Journal of Parasitology, 68, 131–133.CrossRefGoogle Scholar
Marie, J.-L., Fournier, P.-E., Rolain, J.-M., et al. (2006). Molecular detection of Bartonella quintana, B. elizabethae, B. koehlerae, B. doshiae, B. taylorii, and Rickettsia felis in rodent fleas collected in Kabul, Afghanistan. American Journal of Tropical Medicine and Hygiene, 74, 436–439.Google ScholarPubMed
Marshall, A. G. (1981a). The Ecology of Ectoparasitic Insects. London: Academic Press.Google Scholar
Marshall, A. G. (1981b). Sex ratio in ectoparasitic insects. Ecological Entomology, 6, 155–174.CrossRefGoogle Scholar
Martin, T. E., M⊘ller, A. P., Merino, S. & Clobert, J. (2001). Does clutch size evolve in response to parasites and immunocompetence?Proceedings of the National Academy of Sciences of the USA, 98, 2071–2076.CrossRefGoogle ScholarPubMed
Mashek, H., Licznerski, B. & Pincus, S. (1997). Tungiasis in New York. International Journal of Dermatology, 36, 276–278.Google ScholarPubMed
Mashtakov, V. I. (1969). Density dynamics of gerbils and their fleas in different landscape-ecological regions. Problems of Particularly Dangerous Infections, 8, 100–105 (in Russian).Google Scholar
Maslennikova, Z. P., Bibikova, V. A. & Morozova, I. V. (1967). Abundance of Xenopsylla fleas in winter micropopulations in the association with the abundance of the great gerbil. In Proceedings of the 5th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 176–178 (in Russian).Google Scholar
Matejusová, I., Morand, S. & Gelnar, M. (2000). Nestedness in assemblages of gyrodactylids (Monogenea: Gyrodactylidea) parasitising two species of cyprinid: with reference to generalists and specialists. International Journal for Parasitology, 30, 1153–1158.CrossRefGoogle ScholarPubMed
Matthews, J. W. (2004). Effects of site and species characteristics on nested patterns of species composition in sedge meadows. Plant Ecology, 174, 271–278.CrossRefGoogle Scholar
May, R. M. & Anderson, R. M. (1978). Regulation and stability of host–parasite population interactions. II. Destabilizing processes. Journal of Animal Ecology, 47, 455–461.CrossRefGoogle Scholar
May, R. M. & Anderson, R. M. (1979). Population biology of infectious diseases. II. Nature, 280, 455–461.CrossRefGoogle Scholar
Mayevsky, M. P., Bazanova, L. P. & Popkov, A. F. (1999). Winter survival of the causative agent of plaque in the long-tailed ground squirrel in the Tuva natural focus. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 68, 55–58 (in Russian).Google Scholar
Mazgajski, T. D., Kedra, A. H., Modlińska, E. & Samborski, J. (1997). Siphonaptera influence the condition of starling Sturnus vulgaris nestlings. Acta Ornithologica, 32, 185–190.Google Scholar
McCallum, H. & Dobson, A. P. (1995). Detecting disease and parasite threats to endangered species and ecosystems. Trends in Ecology and Evolution, 10, 190–194.CrossRefGoogle ScholarPubMed
McCoy, G. W. & Mitzmain, M. B. (1909). The regional distribution of fleas on rodents. Parasitology, 2, 297–304.CrossRefGoogle Scholar
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, 361–375.CrossRefGoogle Scholar
McKenzie, A. A. (1990). The ruminant dental grooming apparatus. Zoological Journal of the Linnean Society, 99, 117–128.CrossRefGoogle Scholar
McKinney, P. & McDonald, L. C. (2001). Tungiasis. Medical Journal, 2, 3–10.Google Scholar
McTier, T. L., George, J. E. & Bennet, S. N. (1981). Resistance and cross-resistance of guinea pigs to Dermacentor andersoni Stiles, D. variabilis (Say), Amblyomma americanum (Linnaeus) and Ixodes scapularis Say. Journal of Parasitogy, 67, 813–822.CrossRefGoogle ScholarPubMed
Mead-Briggs, A. R. (1964). The reproductive biology of the rabbit flea Spilopsyllus cuniculi (Dale) and the dependence of this species upon the breeding of its host. Journal of Experimental Biology, 41, 371–402.Google Scholar
Mead-Briggs, A. R. & Rudge, A. J. B. (1960). Breeding of the rabbit flea, Spilopsyllus cuniculi (Dale): requirement of a ‘factor’ from a pregnant rabbit for ovarian maturation. Nature, 187, 1136–1137.CrossRefGoogle Scholar
Mead-Briggs, A. R. & Vaughan, J. A. (1969). Some requirements for mating in the rabbit flea, Spilopsyllus cuniculi (Dale). Journal of Experimental Biology, 51, 495–511.Google Scholar
Mead-Briggs, A. R., Vaughan, J. A. & Rennison, B. D. (1975). Seasonal variation in numbers of the rabbit flea on the wild rabbit. Parasitology, 70, 103–118.CrossRefGoogle ScholarPubMed
Mears, S., Clark, F., Greenwood, M. & Larsen, K. S. (2002). Host location, survival and fecundity of the Oriental rat flea Xenopsylla cheopis (Siphonaptera: Pulicidae) in relation to black rat Rattus rattus (Rodentia: Muridae) host sex and age. Bulletin of Entomological Research, 92, 375–384.CrossRefGoogle Scholar
Medvedev, S. G. (1989a). Structure of the head capsule in fleas (Siphonaptera). I. Entomological Review, 68, 1–18.Google Scholar
Medvedev, S. G. (1989b). Ecological characteristics and distribution of fleas of the family Ischnopsyllidae (Siphonaptera). Parasitological Collection, 36, 21–43 (in Russian).Google Scholar
Medvedev, S. G. (1990). Evolution of fleas parasitic on Chiroptera. Parazitologiya, 24, 457–465 (in Russian).Google Scholar
Medvedev, S. G. (1994). Morphological basis of classification of the order Siphonaptera. Entomological Review, 73, 22–43 (in Russian).Google Scholar
Medvedev, S. G. (1996). Geographical distribution of families of fleas (Siphonaptera). Entomological Review, 76, 978–992.Google Scholar
Medvedev, S. G. (1997a). Host–parasite relations in fleas (Siphonaptera). I. Entomological Review, 77, 318–337.Google Scholar
Medvedev, S. G. (1997b). Host–parasite relations in fleas (Siphonaptera). II. Entomological Review, 77, 511–521.Google Scholar
Medvedev, S. G. (1998). Fauna and host–parasite relations of fleas (Siphonaptera) in the Palaearctic. Entomological Review, 78, 292–308.Google Scholar
Medvedev, S. G. (2000a). Fauna and host–parasite associations of fleas (Siphonaptera) in different zoogeographical regions of the world. I. Entomological Review, 80, 409–435.Google Scholar
Medvedev, S. G. (2000b). Fauna and host–parasite associations of fleas (Siphonaptera) in different zoogeographical regions of the world. II. Entomological Review, 80, 640–655.Google Scholar
Medvedev, S. G. (2001). Morphological structure of thoracic and abdominal ctenidia of fleas (Siphonaptera). Parazitologiya, 35, 291–306 (in Russian).Google Scholar
Medvedev, S. G. (2002). Specific features of the distribution and host associations of fleas (Siphonaptera). Entomological Review, 82, 1165–1177.Google Scholar
Medvedev, S. G. (2003a). Morphological adaptations of fleas (Siphonaptera) to parasitism. I. Entomological Review, 82, 40–62 (in Russian).Google Scholar
Medvedev, S. G. (2003b). Morphological adaptations of fleas (Siphonaptera) to parasitism. II. Entomological Review, 82, 820–835 (in Russian).Google Scholar
Medvedev, S. G. (2004). Morphological adaptations of fleas (Siphonaptera) to parasitism. III. Entomological Review, 83, 313–333 (in Russian).Google Scholar
Medvedev, S. G. (2005). An Attempted System Analysis of the Evolution of the Order of Fleas (Siphonaptera), Lectures in Memoriam N. A. Kholodkovsky, No. 57. St Petersburg, Russia: Russian Entomological Society and Zoological Institute of Russian Academy of Sciences (in Russian).Google Scholar
Medvedev, S. G. & Lobanov, A. L. (1999). Information-analytical system of the World fauna of fleas (Siphonaptera): results and prospects. Entomological Review, 79, 654–665.Google Scholar
Medvedev, S. G. & Krasnov, B. R. (2006). Fleas: permanent satellites of small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 161–177.CrossRefGoogle Scholar
Medvedev, S. G., Khabilov, T. K. & Rybin, S. N. (1984). On the biology of the bat fleas (Ischnopsyllidae: Siphonaptera) in the Middle Asia and Kazakhstan. Parazitologiya, 18, 140–149 (in Russian).Google Scholar
Medvedev, S. G., Lobanov, A. L. & Lyangouzov, I. A. (2005). World Database of Fleas (Nov 2004 Version). In Species 2000 and ITIS Catalogue of Life: 2005 Annual Checklist, ed. Bisby, F. A., Ruggiero, M. A., Wilson, K. L., et al.Reading, MA: Species 2000 (CD-ROM).Google Scholar
Medzykhovsky, G. A. (1971). On the method of measurement of the duration of uninterrupted stay of fleas on a host. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 404–406 (in Russian).Google Scholar
Mellanby, K. (1933). The influence of the temperature and humidity on the pupation of Xenopsylla cheopis. Bulletin of Entomological Research, 24, 197–202.CrossRefGoogle Scholar
Méndez, E. (1977). Mammalian–siphonapteran associations, the environment, and biogeography of mammals of southwestern Colombia. Quaestiones Entomologicae, 13, 91–182.Google Scholar
Meng, F.-X., Feng, Y.-L., Chen, J.-Q., Song, X.-P. & Liu, Q.-Y. (2006). The sex ratio and adult eclosion of Xenopsylla cheopis in laboratory. Chinese Journal of Vector Biology and Control, 17, 15–16 (in Chinese).Google Scholar
Menier, K. (2003). Infestation of dairy goats with Pulex irritans. Veterinary Record, 153, 128–128.Google ScholarPubMed
Merila, J. & Allander, K. (1995). Do great tits (Parus major) prefer ectoparasite-free roost sites? – An experiment. Ethology, 99, 53–60.CrossRefGoogle Scholar
Merino, S. & Potti, J. (1996). Weather dependent effects of nest ectoparasites on their bird hosts. Ecography, 19, 107–113.CrossRefGoogle Scholar
Merino, S., Minguez, E. & Belliure, B. (1999). Ectoparasite effects on nestling European storm-petrels. Waterbirds, 22, 297–301.CrossRefGoogle Scholar
Metzger, M. E. & Rust, M. K. (1992). Egg production and emergence of adult cat fleas (Siphonaptera: Pulicidae) exposed to different photoperiods. Journal of Medical Entomology, 33, 651–655.CrossRefGoogle Scholar
Metzger, M. E. & Rust, M. K. (1997). Effect of temperature on cat flea (Siphonaptera: Pulicidae) development and overwintering. Journal of Medical Entomology, 34, 173–178.CrossRefGoogle ScholarPubMed
Meyer, K. F. (1947). The prevention of plague in the light of newer knowledge. Annals of the New York Academy of Sciences, 48, 429–467.CrossRefGoogle Scholar
Michelsen, V. (1997). A revised interpretation of the mouthparts in adult fleas (Insecta, Siphonaptera). Zoologischer Anzeiger, 235, 217–223.Google Scholar
Miklisova, D. & Stanko, M. (1992). Negative binomial distribution as a model for fleas on small rodents. Biologia, Bratislava, 52, 647–652.Google Scholar
Mikulin, M. A. (1956). Data on fleas of the Middle Asia. II. Fauna and some characteristic of geographic distribution of fleas parasitic on the great gerbil in deserts of the southern Trans-Balkhash Desert. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 2, 95–107 (in Russian).Google Scholar
Mikulin, M. A. (1958). Data on fleas of the Middle Asia and Kazakhstan. V. Fleas of the Tarbagatai Mountains. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 4, 227–240 (in Russian).Google Scholar
Mikulin, M. A. (1959a). Data on fleas of the Middle Asia and Kazakhstan. VIII. Fleas of the Akmolinsk Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 5, 237–245 (in Russian).Google Scholar
Mikulin, M. A. (1959b). Data on fleas of the Middle Asia and Kazakhstan. X. Fleas of the eastern Balkhash Desert, Trans-Alakul Desert and Sungorian Gates. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 6, 205–220 (in Russian).Google Scholar
Milazzo, C., Goüy de Bellocq, J., Cagnin, M., et al. (2003). Helminths and ectoparasites of Rattus rattus and Mus musculus from Sicily, Italy. Comparative Parasitology, 70, 199–204.CrossRefGoogle Scholar
Milinski, M. (1990). Parasites and host decision-making. In Parasitism and Host Behaviour, ed. Barnard, C. J. & Behnke, J. M.. London: Taylor & Francis, pp. 95–116.Google Scholar
Milinski, M. & Bakker, T. C. M. (1990). Female sticklebacks use male coloration in mate choice and hence avoid parasitized males. Nature, 344, 330–333.CrossRefGoogle Scholar
Miller, D. H. & Benton, A. H. (1970). Cold tolerance of some adult fleas (Ceratophyllidae: Siphonaptera). Canadian Field Naturalist, 84, 396–397.Google Scholar
Miller, R. A. (1996). The aging immune system: primer and prospectus. Science, 273, 70–74.CrossRefGoogle ScholarPubMed
Mineur, Y. S., Prasol, D. J., Belzung, C. & Crusio, W. E. (2003). Agonistic behavior and unpredictable chronic mild stress in mice. Behavior Genetics, 33, 513–519.CrossRefGoogle Scholar
Mironov, A. N. & Pasyukov, V. V. (1987). Observations on the construction of cocoons by fleas Nosopsyllus fasciatus. Parazitologiya, 21, 10–15 (in Russian).Google Scholar
Mitchell-Jones, A. J., Amori, G., Boganowicz, W., et al. (1999). The Atlas of European Mammals. London: T. & A.D. Poyser.Google Scholar
Miyamoto, K. & Hashimoto, Y. (2000). Outbreak of the sparrow flea bite cases in Hokkaido, Japan. Medical Entomology and Zoology, 51, 111 (in Japanese).CrossRefGoogle Scholar
Moeller, D. A. (2004). Facilitative interactions among plants via shared pollinators. Ecology, 85, 3289–3301.CrossRefGoogle Scholar
Mohr, C. O. (1958). Relation to mean number of fleas to prevalence of infestation on rats. American Journal of Tropical Medicine and Hygiene, 7, 519–522.CrossRefGoogle ScholarPubMed
Moll, A. A. & Leary, O' S. B. (1945). Plague in the Americas: Historical and Quasi-Epidemiological Survey. Washington, DC: The Pan American Sanitary Bureau.Google Scholar
M⊘ller, A. P. (1989). Parasites, predators and nest boxes: facts and artefacts in the nest boxes studies of birds?Oikos, 56, 421–423.CrossRefGoogle Scholar
M⊘ller, A. P. (1991). Ectoparasite loads affect optimal clutch size in swallows. Functional Ecology, 5, 351–359.CrossRefGoogle Scholar
M⊘ller, A. P. (1993). Parasites differentially increase the degree of fluctuating asymmetry in secondary sexual characters. Journal of Evolutionary Biology, 5, 691–699.CrossRefGoogle Scholar
M⊘ller, A. P. (1997). Parasitism and the evolution of host life history. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 105–127.Google Scholar
M⊘ller, A. P. & Lope, F. (1999). Senescence in a short-lived migratory bird: age-dependent morphology, migration, reproduction and parasitism. Journal of Animal Ecology, 68, 163–171.CrossRefGoogle Scholar
M⊘ller, A. P., Christe, P. & Garamszegi, L. Z. (2005). Coevolutionary arms races: increased host immune defense promotes specialization by avian fleas. Journal of Evolutionary Biology, 18, 46–59.CrossRefGoogle Scholar
Moore, J. (2002). Parasites and the Behavior of Animals. New York: Oxford University Press.Google Scholar
Moore, S. L. & Wilson, K. (2002). Parasites as a viability cost of sexual selection in natural populations of mammals. Science, 297, 2015–2018.CrossRefGoogle ScholarPubMed
Mooring, M. S. (1995). The effect of tick challenge on grooming rate by impala. Animal Behaviour, 50, 377–392.CrossRefGoogle Scholar
Mooring, M. S. & Hart, B. L. (1995). Costs of allogrooming in impala: distraction from vigilance. Animal Behaviour, 49, 1414–1416.CrossRefGoogle Scholar
Mooring, M. S. & Hart, B. L. (1997). Self grooming in impala mothers and lambs: testing the body size and tick challenge principles. Animal Behaviour, 53, 925–934.CrossRefGoogle Scholar
Mooring, M. S. & Samuel, W. M. (1999). Premature loss of winter hair in free-ranging moose (Alces alces) infested with winter ticks (Dermacentor albipictus) is correlated with grooming rate. Canadian Journal of Zoology, 77, 148–156.CrossRefGoogle Scholar
Mooring, M. S., Benjamin, J. E., Harte, C. R. & Herzog, N. B. (2000). Testing the interspecific body size principle in ungulates: the smaller they come, the harder they groom. Animal Behaviour, 60, 35–45.CrossRefGoogle ScholarPubMed
Mooring, M. S., Reisig, D. D., Niemeyer, J. M. & Osborne, E. R. (2002). Sexually and developmentally dimorphic grooming: a comparative survey of the Ungulata. Ethology, 108, 911–934.CrossRefGoogle Scholar
Mooring, M. S., Blumstein, D. T. & Stoner, C. J. (2004). The evolution of parasite-defence grooming in ungulates. Biological Journal of the Linnean Society, 81, 17–37.CrossRefGoogle Scholar
Morand, S. (1996). Life-history traits in parasitic nematodes: a comparative approach for the search of invariants. Functional Ecology, 10, 210–218.CrossRefGoogle Scholar
Morand, S. (2000). Wormy world: comparative tests of theoretical hypotheses on parasite species richness. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R., Morand, S. & Skorping, A.. Amsterdam, the Netherlands: Elsevier Science, pp. 63–79.Google Scholar
Morand, S. & Guégan, J.-F. (2000). Distribution and abundance of parasite nematodes: ecological specialization, phylogenetic constraints or simply epidemiology?Oikos, 88, 563–573.CrossRefGoogle Scholar
Morand, S. & Harvey, P. H. (2000). Mammalian metabolism, longevity and parasite species richness. Proceedings of the Royal Society of London B, 267, 1999–2003.CrossRefGoogle ScholarPubMed
Morand, S. & Poulin, R. (1998). Density, body mass and parasite species richness of terrestrial mammals. Evolutionary Ecology, 12, 717–727.CrossRefGoogle Scholar
Morand, S. & Poulin, R. (2000). Nematode parasite species richness and the evolution of spleen size in birds. Canadian Journal of Zoology, 78, 1356–1360.CrossRefGoogle Scholar
Morand, S. & Poulin, R. (2003). Phylogenies, the comparative method and parasite evolutionary ecology. Advances in Parasitology, 54, 281–302.CrossRefGoogle ScholarPubMed
Morand, S., Pointier, J.-P., Borel, G. & Theron, A. (1993). Pairing probability of schistosomes related to their distribution among the host population. Ecology, 74, 2444–2449.CrossRefGoogle Scholar
Morand, S., Poulin, R., Rohde, K. & Hayward, C. (1999). Aggregation and species coexistence of ectoparasites of marine fishes. International Journal for Parasitology, 29, 663–672.CrossRefGoogle ScholarPubMed
Morand, S., Hafner, M. S., Page, R. D. M. & Reed, D. L. (2000). Comparative body size relationships in pocket gophers and their chewing lice. Biological Journal of the Linnean Society, 70, 239–249.CrossRefGoogle Scholar
Morand, S., Goüy de Bellocq, J., Stanko, M. & Miklisova, D. (2004). Is sex-biased ectoparasitism related to sexual size dimorphism in small mammals of central Europe?Parasitology, 129, 505–510.CrossRefGoogle ScholarPubMed
Moret, Y. & Schmid-Hempel, P. (2000). Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 1166–1168.CrossRefGoogle Scholar
Morlan, H. B. (1955). Mammal fleas of Santa Fe County, New Mexico. Texas Reports on Biology and Medicine, 13, 93–125.Google ScholarPubMed
Morozkina, E. A., Lysenko, L. S. & Kafarskaya, D. G. (1970). Materials on the ecology of fleas of the red marmot in the eastern Pamir Mountains. In Vectors of Particularly Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 337–341 (in Russian).Google Scholar
Morozkina, E. A., Lysenko, L. S. & Kafarskaya, D. G. (1971). Fleas of the red marmot (Marmota caudata) and other animals inhabiting the Gissar Ridge. Problems of Particularly Dangerous Infections, 17, 38–44 (in Russian).Google Scholar
Morozov, Y. A. (1974). The level of flea infestation in the great gerbils of different age. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 337–338 (in Russian).Google Scholar
Morozov, Y. A., Rapoport, L. P. & Kovtun, I. P. (1972). Parasitological contacts between the great gerbils (Rhombomys opimus Licht.) and the yellow ground squirrels (Citellus fulvus Licht.) in the Tchu-Moyynkum Desert. Parazitologiya, 6, 334–337 (in Russian).Google Scholar
Morris, D. W. (1987a). Ecological scale and habitat use. Ecology, 68, 362–369.CrossRefGoogle Scholar
Morris, D. W. (1987b). Spatial scale and the cost of density-dependent habitat selection. Evolutionary Ecology, 1, 379–388.CrossRefGoogle Scholar
Morris, D. W. (1988). Habitat-dependent population regulation and community structure. Evolutionary Ecology, 2, 253–269.CrossRefGoogle Scholar
Morris, D. W. (1990). Temporal variation, habitat selection and community structure. Oikos, 59, 303–312.CrossRefGoogle Scholar
Morris, D. W. (2003). Toward an ecological synthesis: a case for habitat selection. Oecologia, 136, 1–13.CrossRefGoogle ScholarPubMed
Morrison, M. L., Marcot, B. G. & Mannan, R. W. (1992). Wildlife–Habitat Relationships: Concepts and Applications. Madison, WI: University of Wisconsin Press.Google Scholar
Morrone, J. J. & Acosta, R. (2006). A synopsis of the fleas (Insecta: Siphonaptera) parasitizing New World species of Soricidae (Mammalia: Insectivora). Zootaxa, 1354, 1–30.Google Scholar
Morrone, J. J. & Gutiérrez, A. (2004). Do fleas (Insecta: Siphonaptera) parallel their mammal host diversification in the Mexican transition zone?Journal of Biogeography, 32, 1315–1325.CrossRefGoogle Scholar
Morrone, J. J., Acosta, R. & Gutiérrez, A. (2000). Cladistics, biogeography, and host relationships of the flea subgenus Ctenophthalmus (Alloctenus), with the description of a new Mexican species (Siphonaptera: Ctenophthalmidae). Journal of the New York Entomological Society, 108, 1–12.CrossRefGoogle Scholar
Morsy, T. A., Kady, El G. A., Salama, M. M. & Sabry, A. H. (1993). The seasonal abundance of Gerbillus pyramidum and their flea ectoparasites in Al Arish, North Sinai Governorate, Egypt. Journal of the Egyptian Society of Parasitology, 23, 269–276.Google ScholarPubMed
Moser, B. A., Koehler, P. G. & Patterson, R. S. (1991). Effect of larval diet on cat flea (Siphonaptera: Pulicidae) developmental times and adult emergence. Journal of Economic Entomology, 84, 1257–1261.CrossRefGoogle ScholarPubMed
Moskalenko, V. V. (1958). On the effect of temperature on the behaviour of fleas after deaths of their host. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 181–184 (in Russian).Google Scholar
Moskalenko, V. V. (1963a). On the longevity of several flea species parasitic on rodents in the Primorie Region. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 6, 166–169 (in Russian).Google Scholar
Moskalenko, V. V. (1963b). On the effect of temperature on reproduction of some flea species from the Primorie Region under laboratory conditions. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 6, 162–165 (in Russian).Google Scholar
Moskalenko, V. V. (1966). On the frequency of feeding in fleas from the Primorie Region. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 349–354 (in Russian).
Mouillot, D. & Poulin, R. (2004). Taxonomic partitioning shedding light on the diversification of parasite communities. Oikos, 104, 205–207.CrossRefGoogle Scholar
Mouillot, D., Krasnov, B. R., Gaston, K., Shenbrot, G. I. & Poulin, R. (2006). Conservatism of host specificity in parasites. Ecography, 29, 596–602.CrossRefGoogle Scholar
Moura, M. O., Bordignon, M. & Graciolli, G. (2003). Host characteristics do not affect community structure of ectoparasites on the fishing bat Noctilio leporinus (L., 1758) (Mammalia: Chiroptera). Memórias do Instituto Oswaldo Cruz, 98, 811–815.CrossRefGoogle Scholar
Moynahan, E. J. (1987). Kawasaki disease: a novel feline virus transmitted by fleas. Lancet, i (8526), 195.CrossRefGoogle Scholar
Muehlen, M., Heukelbach, J., Wilcke, T., et al. (2003). Investigations on the biology, epidemiology, pathology and control of Tunga penetrans in Brazil. II. Prevalence, parasite load and topographic distribution of lesions in the population of a traditional fishing village. Parasitology Research, 90, 449–455.CrossRefGoogle ScholarPubMed
Muirhead-Thomson, R. C. (1968). Ecology of Insect Vector Populations. New York: Academic Press.Google Scholar
Mulenga, A., Sugimoto, C. & Onuma, M. (2000). Issues in tick vaccine development: identification and characterization of potential candidate vaccine antigens. Microbes and Infection, 2, 1353–1361.CrossRefGoogle ScholarPubMed
Mules, M. W. (1940). Notes on the life history and artificial breeding of the Australian ‘stickfast’ flea Echidnophaga mirmecobii Rothschild. Australian Journal of Experimental Biological and Medical Science, 18, 385–390.CrossRefGoogle Scholar
Muller, G. H., Kirk, R. W., Scott, D. W., et al. (2001). Muller and Kirk's Small Animal Dermatology, 6th edn. Philadelphia, PA: Saunders College Publishing.Google Scholar
Muñoz, L., Milenko, A. & Casanueva, M. E. (2003). Prevalencia e intensidad de ectoparasitos asociados a Tadarida brasiliensis (Geoffroy Saint-Hilaire, 1824) (Chiroptera: Molossidae) en Concepción. Gayana, 67, 1–8.Google Scholar
Murzakhmetova, K. (1958). On studying physiological activity of fleas parasitic on gerbils. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 4, 223–226 (in Russian).Google Scholar
Myalkovskaya, S. A. (1983). Some ecological characteristics of fleas parasitic on the pygmy ground squirrels in Dagestan. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 254–255 (in Russian).Google Scholar
Nakazawa, K., Oisi, I. & Kumi, S. (1957). Seasonal prevalence of rodent fleas. Medical Entomology and Zoology, 8, 11–13 (in Japanese).CrossRefGoogle Scholar
Naumov, R. L. & Gutova, V. P. (1984). Experimental study of the participation of gamasid mites and fleas in circulation of the tick-borne encephalitis virus (a review). Parazitologiya, 18, 106–115 (in Russian).Google Scholar
Navarro, C., Marzal, A., Lope, F. & M⊘ller, A. P. (2003). Dynamics of an immune response in house sparrows Passer domesticus in relation to time of day, body condition and blood parasite infection. Oikos, 101, 291–298.CrossRefGoogle Scholar
Nayak, N. C. & Bhowmik, M. K. (1990). Goat flea (order Siphonaptera) as a possible vector for the transmission of caprine mycoplasmal polyarthritis with septicaemia. Preventive Veterinary Medicine, 9, 259–266.CrossRefGoogle Scholar
Nazarova, I. V. (1981). Fleas of the Volga–Kama Region. Moscow, USSR: Nauka (in Russian).Google Scholar
Neal, E. G. & Roper, T. J. (1991). The environmental impact of badgers (Meles meles) and their setts. Symposia of the Zoological Society of London, 63, 89–106.Google Scholar
Nee, S. (1994). How populations persist. Nature, 367, 123–124.CrossRefGoogle Scholar
Nee, S., Hassell, M. P. & May, R. M. (1997). Two-species metapopulation models. In Metapopulation Biology: Ecology, Genetics, and Evolution, ed. Hanski, I. & Gilpin, M. E.. San Diego, CA: Academic Press, pp. 123–147.Google Scholar
Nekola, J. C. & White, P. S. (1999). The distance decay of similarity in biogeography and ecology. Journal of Biogeography, 26, 867–878.CrossRefGoogle Scholar
Nelson, R. J. & Demas, G. E. (1996). Seasonal changes in immune function. Quarterly Review of Biology, 71, 511–548.CrossRefGoogle ScholarPubMed
Nelzina, E. N., Danilova, G. M. & Tchernova, G. I. (1963). Abundance and spatial distribution of micropopulations of hematophagous arthropods in microhabitats of Citellus pygmaeus. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 32, 45–54 (in Russian).Google Scholar
Němec, F. (1993). Flea community inhabiting nests of the common vole (Microtus arvalis Pallas, 1779) in west Bohemian farmland. Folia Musei Rerum Naturalium Bohemiae Occidentalis Zoologica, 38, 1–37.Google Scholar
Neter, J., Wasserman, W. & Kutner, M. H. (1990). Applied Linear Statistical Models: Regression, Analysis of Variance, and Experimental Design. Homewood, IL: Richard D. Irwin.
Neuhaus, P. (2003). Parasite removal and its impact on litter size and body condition in Columbian ground squirrels (Spermophilus columbianus). Proceedings of the Royal Society of London B, 270, S213–S215.CrossRefGoogle Scholar
Newton, I. (1998). Population Limitation in Birds. London: Academic Press.Google Scholar
Nijssen, A. (1985). Body temperature regulation by care of the fur in rats Rattus norvegicus albinos. Netherlands Journal of Zoology, 35, 423–437.CrossRefGoogle Scholar
Nikitina, N. A. (1961). Results of marking and recapturing of small mammals in the Komi Soviet Republic. Bulletin of the Moscow Naturalist Society, Series Biology, 66, 15–25 (in Russian).Google Scholar
Nikitina, N. A. & Nikolaeva, G. (1979). Study of the ability of some rodents to get rid of fleas. Zoologicheskyi Zhurnal, 58, 931–933 (in Russian).Google Scholar
Nikitina, N. A. & Nikolaeva, G. (1981). Ability of rodents to clean themselves of specific and non-specific fleas. Zoologicheskyi Zhurnal, 60, 165–167 (in Russian).Google Scholar
Nikulshin, S. V. (1980). Descriptions of the annual cycles of fleas (Aphaniptera) parasitic on Citellus musicus from the Baksan Valley. Parazitologiya, 14, 134–141 (in Russian).Google Scholar
Nikulshin, S. V. & Shinkareva, V. N. (1983). Seasonal patterns of ecology of fleas Citellophilus tesquorum in the Enikol Tract. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 260–261 (in Russian).Google Scholar
Nilsson, J. A. (2003). Ectoparasitism in marsh tits: costs and functional explanations. Behavioral Ecology, 14, 175–181.CrossRefGoogle Scholar
Njau, B. C. & Nyindo, M. (1987). Detection of immune response in rabbits infested with Rhipicephalus appendiculatus andRhipicephalus evertsi evertsi. Research in Veterinary Science, 43, 217–221.Google Scholar
Njunwa, K. J., Mwaiko, G. L., Kilonzo, B. S. & Mhina, J. I. (1989). Seasonal patterns of rodents, fleas and plague status in the Western Usambara Mountains, Tanzania. Medical and Veterinary Entomology, 3, 17–22.CrossRefGoogle ScholarPubMed
Nordling, D., Andersson, M., Zohari, S. & Gustafsson, L. (1998). Reproductive effort reduces specific immune response and parasite resistance. Proceedings of the Royal Society of London B, 265, 1291–1298.CrossRefGoogle Scholar
Norman, R., Bowers, R. G., Begon, M. & Hudson, P. J. (1999). Persistence of tick-borne virus in the presence of multiple host species: tick reservoirs and parasite mediated competition. Journal of Theoretical Biology, 200, 111–118.CrossRefGoogle ScholarPubMed
Nosil, P. (2002). Transition rates between specialization and generalization in phytophagous insects. Evolution, 56, 1701–1706.CrossRefGoogle ScholarPubMed
Novokreshchenova, N. S. (1960). Materials on ecology of fleas of Citellus pygmaeus in relation to their epizootological importance. Proceedings of the Anti-Plague ‘Microb’ Institute, 4, 444–456 (in Russian).Google Scholar
Novokreshchenova, N. S. (1962). Materials on comparative ecology of three species of fleas parasitic on the great gerbil. In Particularly Dangerous and Natural Diseases, ed. ,Anonymous. Moscow, USSR: Medgiz, pp. 53–63 (in Russian).Google Scholar
Novokreshchenova, N. S. & Kuznetsova, G. S. (1964). Ecological characteristic of fleas of the great gerbil in regions with permanent plague epizootics. Zoologicheskyi Zhurnal, 43, 1638–1647 (in Russian).Google Scholar
Novokreshchenova, N. S., Soldatkin, I. S. & Levoshina, A. I. (1968). Comparative frequency of blood meals in various flea species, measured in the laboratory conditions using radioactive indicators. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 49–54 (in Russian).Google Scholar
Novokreshchenova, N. S., Zagniborodova, E. N., Zabegalova, M. N., et al. (1975). Multiannual dynamics of density in fleas parasitic on the great gerbils in Turkmenistan. Problems of Particularly Dangerous Diseases, 43/44, 84–90 (in Russian).Google Scholar
Novozhilova, E. N. (1977). Ectoparasites of small mammals and inhabitants of their burrows in the Pre-Polar Ural. Proceedings of the Komi Branch of the Academy of Sciences of the USSR, 34, 125–139 (in Russian).Google Scholar
Nuriev, K. K., Rapoport, L. P., Shokputov, T. M., Orlova, L. M. & Duisenbiev, D. M. (2004). Materials on rodent fleas from mountain regions of southern Kazakhstan. Quarantinable and Zoonotic Infections in Kazakhstan, 9, 66–70 (in Russian).Google Scholar
Oberdorff, T., Hugueny, B., Compin, A. & Belkessam, D. (1998). Non-interactive fish communities in the coastal streams of north-western France. Journal of Animal Ecology, 67, 472–484.CrossRefGoogle Scholar
Brien, O' E. L. & Dawson, R. D. (2005). Perceived risk of ectoparasitism reduces primary reproductive investment in tree swallows Tachycineta bicolor. Journal of Avian Biology, 36, 269–275.CrossRefGoogle Scholar
Donnell, O' M. J. & Machin, J. (1988). Water vapor absorption by terrestrial organisms. Advances in Comparative and Environmental Physiology, 2, 47–90.CrossRefGoogle Scholar
Oguge, N., Rarieya, M. & Ondiaka, P. (1997). A preliminary survey of macroparasite community of rodents of Kahawa, central Kenya. Belgian Journal of Zoology, 127, S113–S118.Google Scholar
Hare, O' N. (2005). Asking the right questions. New Zealand Listener, 197 (Jan. 29–Feb. 4), No. 3377.Google Scholar
Olsen, N. J. & Kovacs, W. J. (1996). Gonadal steroids and immunity. Endocrine Review, 17, 369–384.Google ScholarPubMed
Olsson, K. & Allander, K. (1995). Do fleas, and/or old nest material, influence nest-site preference in hole-nesting passerines? Ethology, 101, 160–170.CrossRefGoogle Scholar
Olsufiev, N. G. (1975). Taxonomy, Microbiology and Laboratory Diagnostics of the Tularaemia Pathogen. Moscow, USSR: Meditsina (in Russian).Google Scholar
Olsufiev, N. G. & Dunaeva, T. N. (1960). Epizootology (natural focality) of tularaemia. In Tularemia, ed. Olsufiev, N. G. & Rudnev, G. P.. Moscow, USSR: Meditsina, pp. 136–206 (in Russian).Google Scholar
Oparina, O., Starozhitskaya, G. S. & Samurov, M. A. (1989). Human blood-sucking ability of fleas parasitic on rodents and small gerbils. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 58, 23–25 (in Russian).Google Scholar
Opdebeeck, J. P. & Slacek, B. (1993). An attempt to protect cats against infestation with Ctenocephalides felis felis using gut membrane antigens as a vaccine. International Journal for Parasitology, 23, 1063–1067.CrossRefGoogle ScholarPubMed
Oppliger, A., Richner, H. & Christe, P. (1994). Effect of an ectoparasite on lay date, nest-site choice, desertion, and hatching success in the great tit (Parus major). Behavioral Ecology, 5, 130–134.CrossRefGoogle Scholar
Orell, M., Rytkönen, S. & Ilomäki, K. (1993). Do pied flycatchers prefer nest boxes with old nest material?Annales Zoologici Fennici, 30, 313–316.Google Scholar
Osacar-Jimenez, J. J., Lucientes-Curdi, J. & Calvete-Margolles, C. (2001). Abiotic factors influencing the ecology of wild rabbit fleas in north-eastern Spain. Medical and Veterinary Entomology, 15, 157–166.CrossRefGoogle ScholarPubMed
Osbrink, W. L. & Rust, M. (1984). Fecundity and longevity of the adult cat flea Ctenocephalides felis felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 21, 727–731.CrossRefGoogle Scholar
Osbrink, W. L. & Rust, M. (1985). Cat flea (Siphonaptera: Pulicidae): factors influencing host finding behaviour in the laboratory. Annals of the Entomological Society of America, 78, 29–34.CrossRefGoogle Scholar
Ostfeld, R. & Keesing, F. (2000). The function of biodiversity in the ecology of vector-borne zoonotic diseases. Canadian Journal of Zoology, 78, 2061–2078.CrossRefGoogle Scholar
Overal, W. L. (1980). Host-relations of the batfly Megistopoda aranea (Diptera: Streblidae) in Panama. University of Kansas Scientific Bulletin, 52, 1–20.Google Scholar
Öztürk, L., Pelin, Z., Karadeniz, D., et al. (1999). Effects of 48 hours' sleep deprivation on human immune profile. Sleep Research Online, 2, 107–111.Google Scholar
Pacala, S. W. & Dobson, A. P. (1988). The relation between the number of parasites/host and host age: population dynamic causes and maximum likelihood estimation. Parasitology, 96, 197–210.CrossRefGoogle ScholarPubMed
Pacejka, A. J., Gratton, C. M. & Thompson, C. F. (1998). Do potentially virulent mites affect house wren (Troglodytes aedon) reproductive success?Ecology, 79, 1797–1806.CrossRefGoogle Scholar
Page, R. M. D. (1990). Component analysis: a valiant failure?Cladistics, 6, 119–136.CrossRefGoogle Scholar
Page, R. M. D. (1993a). Parasites, phylogeny and cospeciation. International Journal for Parasitology, 23, 499–506.CrossRefGoogle Scholar
Page, R. M. D. (1993b). Genes, organisms, and areas: the problem of multiple lineages. Systematic Biology, 42, 77–84.CrossRefGoogle Scholar
Page, R. M. D. (1994a). Parallel phylogenies: reconstructing the history of host–parasite assemblages. Cladistics, 10, 155–173.CrossRefGoogle Scholar
Page, R. M. D. (1994b). Maps between trees and cladistic analysis of historical associations among genes, organisms, and areas. Systematic Biology, 43, 58–77.Google Scholar
Page, R. D. M. (ed.) (2003). Tangled Trees: Phylogeny, Cospeciation, and Coevolution. Chicago, IL: University of Chicago Press.Google Scholar
Page, R. D. M. & Charleston, M. D. (1998). Trees within trees: phylogeny and historical associations. Trends in Ecology and Evolution, 13, 356–359.CrossRefGoogle ScholarPubMed
Palombit, R. A., Cheney, D. L. & Seyfarth, R. M. (2001). Female–female competition for male ‘friends’ in wild chacma baboons, Papio cynocephalus ursinus. Animal Behaviour, 61, 1159–1171.CrossRefGoogle Scholar
Pampiglione, S., Trentini, M., Fioravanti, M. L., Onore, G. & Rivasi, F. (2003). Additional description of a new species of Tunga (Siphonaptera) from Ecuador. Parasite, 10, 9–15.CrossRefGoogle ScholarPubMed
Panchenko, G. M. (1971). Ecological study of larvae of the rat fleas of Siberia and Far East. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 229–231 (in Russian).Google Scholar
Paramonov, B. B., Emelianova, N. D., Zarubina, V. N. & Kontrimavitchus, V. L. (1966). Materials for the study of ectoparasites of rodents and shrews of the Kamchatka Peninsula. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 333–341 (in Russian).Google Scholar
Park, T. (1948). Experimental studies of interspecific competition. I. Competition between populations of the flour beetles, Tribolium confusum Duval and Tribolium castaneum Herbst. Ecological Monographs, 18, 265–308.CrossRefGoogle Scholar
Parman, D. C. (1923). Biological notes on the hen flea, Echidnophaga gallinacea. Journal of Agricultural Research, 23, 1007–1009.Google Scholar
Parola, R. & Raoult, D. (2006). Tropical rickettsioses. Clinics in Dermatology, 24, 191–200.CrossRefGoogle ScholarPubMed
Parola, P., Davoust, B. & Raoult, D. (2005). Tick- and flea-borne rickettsial emerging zoonoses. Veterinary Research, 36, 469–492.CrossRefGoogle ScholarPubMed
Parsons, P. A. (1990). Fluctuating asymmetry: an epigenetic measure of stress. Biological Reviews, 17, 391–421.Google Scholar
Pascal, M., Beaucournu, J. C. & Lorvelec, O. (2004). An enigma: the lack of Siphonaptera on wild rats and mice on densely populated tropical islands. Acta Parasitologica, 49, 168–172.Google Scholar
Paterson, A. M. & Banks, J. (2001). Analytical approaches to measuring cospeciation of host and parasites: through a glass, darkly. International Journal for Parasitology, 31, 1012–1022.CrossRefGoogle ScholarPubMed
Paterson, A. M. & Gray, R. D. (1997). Host–parasite co-speciation, host switching, and missing the boat. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 236–250.Google Scholar
Paterson, A. M., Gray, R. D. & Wallis, G. P. (1993). Parasites, petrels and penguins: does louse presence reflect seabird phylogeny?International Journal for Parasitology, 23, 515–526.CrossRefGoogle Scholar
Paterson, A. M., Wallis, G. P., Wallis, L. J. & Gray, R. D. (2000). Seabird and louse coevolution: Complex histories revealed by 12S rRNA sequences and reconciliation analyses. Systematic Biology, 38, 144–153.Google Scholar
Patrick, M. J. (1991). Distribution of enteric helminthes in Glaucomys volans L. (Sciuridae): a test for competition. Ecology, 72, 755–758.CrossRefGoogle Scholar
Patterson, B. D. (1990). On the temporal development of nested subset patterns of species composition. Oikos, 59, 330–342.CrossRefGoogle Scholar
Patterson, B. D. & Atmar, W. (1986). Nested subsets and the structure of insular mammalian faunas and archipelagos. Biological Journal of the Linnean Society, 28, 65–82.CrossRefGoogle Scholar
Patterson, B. D. & Brown, J. H. (1991). Regionally nested patterns of species composition in granivorous rodent assemblages. Journal of Biogeography, 18, 395–402.CrossRefGoogle Scholar
Pauller, O. F. & Tchipizubova, P. A. (1958). Ecology of fleas of the Daurian ground squirrel in the Trans-Baikalia. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 161–179 (in Russian).Google Scholar
Pauller, O. F., Elshanskaya, N. I. & Shvetsova, I. V. (1966). Ecological and faunistical review of mammalian and bird ectoparasites in the tularemia focus of the Selenga River delta. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 322–332 (in Russian).Google Scholar
Peach, W. J., Fowler, J. A. & Greenwood, M. T. (1987). Seasonal variation in the infestation of starlings Sturnus vulgaris by fleas (Siphonaptera). Bird Study, 34, 251–252.CrossRefGoogle Scholar
Pennings, S. C. & Silliman, B. R. (2005). Linking biogeography and community ecology: latitudinal variation in plant–herbivore interaction strength. Ecology, 86, 2310–2319.CrossRefGoogle Scholar
Perlman, S. J. & Jaenike, J. (2001). Competitive interactions and persistence of two nematode species that parasitize Drosophila recens. Ecology Letters, 4, 577–584.CrossRefGoogle Scholar
Perry, J. N. (1988). Some models for spatial variability of animal species. Oikos, 51, 124–130.CrossRefGoogle Scholar
Perry, J. N. & Taylor, L. R. (1986). Stability of real interacting populations in space and time: implications, alternatives and negative binomial kc. Journal of Animal Ecology, 55, 1053–1068.CrossRefGoogle Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Peterson, A. T., Soberon, J. & Sanchez-Cordero, V. (1999). Conservatism of ecological niches in evolutionary time. Science, 285, 1265–1267.CrossRefGoogle ScholarPubMed
Petit, C., Hossaert-McKey, M., Perret, P., Blondel, J. & Lambrechts, M. M. (2002). Blue tits use selected plants and olfaction to maintain an aromatic environment for nestlings. Ecology Letters, 5, 585–589.CrossRefGoogle Scholar
Peus, F. (1968). Über die beiden Bernstein-Flöhe (Insecta, Siphonaptera). Paläontologische Zeitschrift, 42, 62–72.CrossRefGoogle Scholar
Peus, F. (1970). Zur Kenntnis der Flöhe Deutschlands (Insecta, Siphonaptera). IV. Faunistik und Ökologie der Säugeteierflöhe. Zoologische Jahrbücher, Abteilung für Systematic Ökologie und Geographie der Tiere, 99, 400–418.Google Scholar
Piersma, T. & Drent, J. (2003). Phenotypic flexibility and the evolution of organismal design. Trends in Ecology and Evolution, 18, 228–233.CrossRefGoogle Scholar
Pigage, H. K. & Larson, O. R. (1983). The detection of glycerol in overwintering purple martin fleas, Ceratophyllus idius. Comparative Biochemistry and Physiology A, 75, 593–595.CrossRefGoogle Scholar
Pigage, H. K., Pigage, J. C. & Tillman, J. F. (2005). Fleas associated with the northern pocket gopher (Thomomys talpoides) in Elbert County, Colorado. Western North American Naturalist, 65, 210–214.Google Scholar
Pilgrim, R. L. C. & Galloway, T. D. (2004). Descriptions of flea larvae (Siphonaptera: Ceratophyllidae, Leptopsyllidae) found in nests of the house martin, Delichon urbica (Aves: Hirundinidae), in Great Britain. Journal of Natural History, 38, 473–502.CrossRefGoogle Scholar
Pimm, S. L. (1979). The structure of food webs. Theoretical Population Biology, 16, 144–158.CrossRefGoogle ScholarPubMed
Pimm, S. L. & Rosenzweig, M. L. (1981). Competitors and habitat use. Oikos, 37, 1–6.CrossRefGoogle Scholar
Pollitzer, R. (1960). A review of recent literature on plague. Bulletin of the World Health Organization, 23, 313–400.Google ScholarPubMed
Pollitzer, R. & Meyer, K. F. (1961). The ecology of plague. In Studies of Disease Ecology, ed. May, J. F.. New York: Hafner, pp. 433–590.Google Scholar
Ponomarenko, A. G. (1976). The new insect from the Cretaceous of the Trans-Baikalia was a probable parasite of pterosaurs. Paleontologicheskyi Zhurnal, 3, 102–106 (in Russian).Google Scholar
Ponomarenko, A. G. (1986). Scarabaeiformis incertae sedis. Transactions of the Joint Soviet–Mongolian Paleontological Expedition, 28, 110–112 (in Russian).Google Scholar
Popov, V. N. & Verzhutsky, D. B. (1988). Characteristics of intrapopulation aggregations of the long-tailed ground squirrel (Citellus undulatus Pall.) under density decline. Bulletin of the Moscow Naturalist Society, Series Biology, 93, 47–50 (in Russian).Google Scholar
Popova, A. S. (1968). Flea fauna of the Moyynkum Desert. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 402–406 (in Russian).Google Scholar
Poulin, R. (1993). The disparity between observed and uniform distibutions: a new look at parasite aggregation. International Journal for Parasitology, 23, 937–944.CrossRefGoogle Scholar
Poulin, R. (1995a). Clutch size and egg size in free-living and parasitic copepods: a comparative analysis. Evolution, 49, 325–336.CrossRefGoogle Scholar
Poulin, R. (1995b). Phylogeny, ecology, and the richness of parasite communities in vertebrates. Ecological Monographs, 65, 283–302.CrossRefGoogle Scholar
Poulin, R. (1996a) Sexual inequalities in helminth infections: a cost of being male?American Naturalist, 147, 289–295.CrossRefGoogle Scholar
Poulin, R. (1996b). Richness, nestedness and randomness in parasite infracommunity structure. Oecologia, 105, 545–551.CrossRefGoogle Scholar
Poulin, R. (1997). Parasite faunas of freshwater fish: the relationship between richness and the specificity of parasites. International Journal for Parasitology, 27, 1091–1098.CrossRefGoogle ScholarPubMed
Poulin, R. (1998). Large-scale patterns of host use by parasites of freshwater fishes. Ecology Letters, 1, 118–128.CrossRefGoogle Scholar
Poulin, R. (1999). Body size vs. abundance among parasite species: positive relationships?Ecography, 22, 246–250.CrossRefGoogle Scholar
Poulin, R. (2003). The decay of similarity with geographical distance in parasite communities of vertebrate hosts. Journal of Biogeography, 30, 1609–1615.CrossRefGoogle Scholar
Poulin, R. (2005). Relative infection levels and taxonomic distances among the host species used by a parasite: insights into parasite specialization. Parasitology, 130, 109–115.CrossRefGoogle ScholarPubMed
Poulin, R. (2006). Variation in infection parameters among populations within parasite species: intrinsic properties versus local factors. International Journal for Parasitology, 36, 877–885.CrossRefGoogle ScholarPubMed
Poulin, R. (2007a). Evolutionary Ecology of Parasites: From Individuals to Communities, 2nd edn. Princeton, NJ: Princeton University Press.Google Scholar
Poulin, R. (2007b). Are there general laws in parasite ecology?Parasitology, 134, 763–776.CrossRefGoogle Scholar
Poulin, R. & Guégan, J.-F. (2000). Nestedness, antinestedness, and relationship between prevalence and intensity in ectoparasite assemblages of marine fish: a spatial model of species co-existence. International Journal for Parasitology, 30, 1147–1152.CrossRefGoogle Scholar
Poulin, R. & Hamilton, W. J. (1997). Ecological correlates of body size and egg size in parasitic Ascothoracida and Rhizocephala (Crustacea). Acta Oecologica, 18, 621–635.CrossRefGoogle Scholar
Poulin, R. & Morand, S. (2004). Parasite Biodiversity. Washington, DC: Smithsonian Institution Press.Google Scholar
Poulin, R. & Mouillot, D. (2003). Parasite specialization from a phylogenetic perspective: a new index of host specificity. Parasitology, 126, 473–480.CrossRefGoogle ScholarPubMed
Poulin, R. & Mouillot, D. (2004a). The relationship between specialization and local abundance: the case of helminth parasites of birds. Oecologia, 140, 372–378.CrossRefGoogle Scholar
Poulin, R. & Mouillot, D. (2004b). The evolution of taxonomic diversity in helminth assemblages of mammalian hosts. Evolutionary Ecology, 18, 231–247.CrossRefGoogle Scholar
Poulin, R. & Mouillot, D. (2005a). Combining phylogenetic and ecological information into a new index of host specificity. Journal of Parasitology, 91, 511–514.CrossRefGoogle Scholar
Poulin, R. & Mouillot, D. (2005b). Host specificity and the probability of discovering species of helminth parasites. Parasitology, 130, 709–715.CrossRefGoogle Scholar
Poulin, R. & Valtonen, E. T. (2001). Nested assemblages resulting from host-size variation: the case of endoparasite communities in fish hosts. International Journal for Parasitology, 31, 194–1204.CrossRefGoogle ScholarPubMed
Poulin, R. & Valtonen, E. T. (2002). The predictability of helminth community structure in space: a comparison of fish populations from adjacent lakes. International Journal for Parasitology, 30, 1235–1243.CrossRefGoogle Scholar
Poulin, P., Brodeur, J. & Moore, J. (1994). Parasite manipulation of host behaviour: should hosts always lose?Oikos, 70, 479–484.CrossRefGoogle Scholar
Poulin, R., Mouillot, D. & George-Nascimento, M. (2003). The relationship between species richness and productivity in metazoan parasite communities. Oecologia, 137, 277–285.CrossRefGoogle ScholarPubMed
Poulin, R., Krasnov, B. R. & Morand, S. (2006a). Patterns of host specificity in parasites exploiting small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 233–256.CrossRefGoogle Scholar
Poulin, R., Krasnov, B. R., Shenbrot, G. I., Mouillot, D. & Khokhlova, I. S. (2006b). Evolution of host specificity in fleas: is it directional and irreversible?International Journal for Parasitology, 36, 185–191.CrossRefGoogle Scholar
Prasad, R. S. (1969). Influence of host on fecundity of the Indian rat flea, Xenopsylla cheopis (Roths.). Journal of Medical Entomology, 6, 443–447.CrossRefGoogle Scholar
Prasad, R. S. (1972). Different site selections by the rat fleas Xenopsyllsa cheopis and Xenopsyllsa astia (Siphonaptera, Pulicidae). Entomologist's Gazette, 108, 63–64.Google Scholar
Prasad, R. S. (1973). Studies on host–flea relationships. II. Sex hormones of the host and fecundity of rat fleas Xenopsylla astia (Rothschild) and Xenopsylla cheopis (Rothschild) (Siphonaptera). Indian Journal of Medical Research, 61, 38–44.Google Scholar
Prasad, R. S. (1976). Studies on host–flea relationships. IV. Progesterone and cortisone do not influence reproductive potentials of rat fleas Xenopsylla astia (Rothschild) and Xenopsylla cheopis (Rothschild) (Siphonaptera). Parasitology Research, 50, 81–46.Google Scholar
Prasad, R. S. (1987). Host dependency among haematophagous insects: a case study on flea–host association. Proceedings of the Indian Academy of Sciences, 96, 349–360.CrossRefGoogle Scholar
Prasad, R. S. & Kamala Bai, M. (1976). Studies on host specificity in fleas. In Insect and Host Specificity: Proceedings of the Symposium on Problems of Host Specificity, ed. Ananthakrishnan, T. N.. Madras, India: Loyola College, pp. 111–115.Google Scholar
Price, T., Lovette, I. J., Bermingham, E., Gibbs, H. L. & Richman, A. D. (2000). The imprint of history on communities of North American and Asian warblers. American Naturalist, 156, 354–367.CrossRefGoogle ScholarPubMed
Prokopiev, V. N. (1969). Morphological types of cocoons and mechanism of emergence of imago fleas. In Proceedings of the 4th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute and Kainar, pp. 182–184 (in Russian).Google Scholar
Promislow, D. E. L. (1992). Costs of sexual selection in natural populations of mammals. Proceedings of the Royal Society of London B, 247, 203–210.CrossRefGoogle Scholar
Puchala, P. (2004). Detrimental effects of larval blow files (Protocalliphora azurea) on nestlings and breeding success of tree sparrows (Passer montanus). Canadian Journal of Zoology, 82, 1285–1290.CrossRefGoogle Scholar
Pullen, S. R. & Meola, R. W. (1995). Survival and reproduction of the cat flea (Siphonaptera: Pulicidae) fed human blood on an artificial membrane system. Journal of Medical Entomology, 32, 467–4670.CrossRefGoogle ScholarPubMed
Pulliam, H. R. (2000). On the relationship between niche and distribution. Ecology Letters, 3, 349–361.CrossRefGoogle Scholar
Punsky, E. E. & Zagniborodova, E. N. (1964). Role of fleas Xenopsylla gerbilli gerbilli in maintenance and transmission of the pathogen of erysepeloid. Proceedings of the Turkmenian Scientific Institute of Epidemiology and Hygiene, 6, 345–347 (in Russian).Google Scholar
Pushnitsa, F. A., Shevchenko, S. F., Mironov, A. N., et al. (1978). Present status of abundance and habitat distribution of the ground squirrels and their fleas in the eastern parts of the Rostov Region. Problems of Particularly Dangerous Diseases, 60, 48–52 (in Russian).Google Scholar
Qi, Y.-M. (1990a). Fine structures of the reproductive system of three flea species: development of female genitalia. Acta Entomologica Sinica, 33, 182–188 (in Chinese).Google Scholar
Qi, Y.-M. (1990b). Fine structures of the reproductive system of three flea species: development of male genitalia. Acta Entomologica Sinica, 33, 403–411 (in Chinese).Google Scholar
Qian, T.-J., Gong, Z.-D. & Guo, X.-G. (2000). Sex ratio analysis on dominant flea species of the flea community in the foci of human plague in Yunnan. Medical Journal of Dali College, 9, 1–3 (in Chinese).Google Scholar
Radovsky, F. J. (1985). Evolution of mammalian mesostigmatid mites. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 441–504.Google Scholar
Randolph, S. E. (1977). Changing spatial relationships in a population of Apodemus sylvaticus with the onset of breeding. Journal of Animal Ecology, 46, 653–676.CrossRefGoogle Scholar
Randolph, S. E. (1994). Density-dependent acquired resistance to ticks in natural hosts, independent of concurrent infection with Babesia microti. Parasitology, 108, 413–419.CrossRefGoogle ScholarPubMed
Rapoport, E. H. (1982). Areography: Geographic Strategies of Species. London: Pergamon Press.Google Scholar
Rapoport, L. P., Morozov, Y. A. & Korneyev, G. A. (1976). Intensity of parasitological contacts in the colonies of the great gerbils under different population densities of animals and their fleas. Parazitologiya, 10, 392–396 (in Russian).Google Scholar
Rapoport, L. P., Kondratenko, L. P., Orlova, L. M., et al. (2007). Fleas in the Moyynkum Desert and their epozootological role. Zoologicheskyi Zhurnal, 86, 44–51 (in Russian).Google Scholar
Rasnitsyn, A. P. (1992). Strashila incredibilis, a new enigmatic mecopteroid insect with possible siphonapteran affinities from upper Jurassic of Siberia. Psyche, 99, 323–333.CrossRefGoogle Scholar
Rasnitsyn, A. P. (2002a). Infraclass Gryllones Laicharting, 1781: the grylloneans. In History of Insects, ed. Rasnitsyn, A. P. & Quicke, D. L. J.. Dordrecht, the Netherlands: Kluwer, pp. 254–324.CrossRefGoogle Scholar
Rasnitsyn, A. P. (2002b). Order Pulicida Billbergh, 1820: the fleas. In History of Insects, ed. Rasnitsyn, A. P. & Quicke, D. L. J.. Dordrecht, the Netherlands: Kluwer, pp. 239–242.CrossRefGoogle Scholar
Rassokhina, O. S., Starozhitskaya, G. S. & Knyazeva, T. V. (1985). Fecundity of four flea species in the laboratory colonies. Parazitologiya, 19, 488–490 (in Russian).Google Scholar
Raszl, S. M., Cabral, D. D. & Linardi, P. M. (1998). Xenopsylla cheopis on dogs from Brazil: first report. Arquivo Brasileiro de Medicina Veterinária e Zootecnia, 50, 211–212 (in Portugese).Google Scholar
Raszl, S. M., Cabral, D. D. & Linardi, P. M. (1999). Notas sobre Sifonápteros (Pulicidae, Tungidae e Rhopalopsyllidae) de carnívoros domésticos brasileiros. Revista Brasiliera de Entomologia, 43, 95–97.Google Scholar
Ratovonjato, J., Duchemin, J. B. & Chanteau, S. (2000). Optimized method for rearing fleas (Xenopsylla cheopis and Synopsyllus fonquerniei). Archives de l'Institut Pasteur de Madagascar, 66, 75–77.Google Scholar
Rausher, M. D. (1993). The evolution of habitat preference: avoidance and adaptation. In Evolution of Insect Pests: Patterns of Variation, ed. Kim, K. C. & McPherson, B. A.. New York: John Wiley, pp. 259–283.Google Scholar
Ravkin, Y. S. & Sapegina, V. F. (1990). Fleas of rodents of the southern taiga of the Pri-Angarie. Bulletin of the Siberian Branch of the Academy of Sciences of the USSR, Series Biological Sciences, 3, 63–68 (in Russian).Google Scholar
Rea, J. G. & Irwin, S. W. B. (1994). The ecology of host-finding behaviour and parasite transmission: past and future perspectives. Parasitology, 109, S31–S39.CrossRefGoogle ScholarPubMed
Rechav, Y. (1992). Naturally acquired resistance to ticks: a global view. Insect Science and its Application, 13, 495–504.Google Scholar
Rechav, Y. & Dauth, J. (1987). Development of resistance in rabbits to immature stages of the ixodid tick Rhipicephalus appendiculatus. Medical and Veterinary Entomology, 1, 177–183.CrossRefGoogle ScholarPubMed
Rechav, Y. & Fielden, L. J. (1995). The effect of host resistance on the metabolic rate of engorged females of Rhipicephalus evertsi evertsi. Medical and Veterinary Entomology, 9, 289–292.CrossRefGoogle ScholarPubMed
Rechav, Y & Fielden, L. J. (1997). The effect of various host species on the feeding performance of immature stages of the tick Hyalomma truncatum (Acari: Ixodidae). Experimental and Applied Acarology, 21, 551–559.CrossRefGoogle Scholar
Rechav, Y., Heller-Haupt, A. & Varma, M. G. R. (1989). Resistance and cross-resistance in guinea-pigs and rabbits to immature stages of ixodid ticks. Medical and Veterinary Entomology, 3, 333–336.CrossRefGoogle ScholarPubMed
Redford, K. H. & Eisenberg, J. F. (1992). Mammals of the Neotropics: The Southern Cone, vol. 2, Chile, Argentine, Uruguay, Paraguay. Chicago, IL: University of Chicago Press.Google Scholar
Reichardt, T. R. & 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. Journal of Medical Entomology, 31, 105–113.CrossRefGoogle Scholar
Reichman, O. J. & Smith, S. C. (1990). Burrows and burrowing behavior by mammals. In Current Mammology, vol. 2, ed. Genoways, H. H.. New York: Plenum Press, pp. 197–244.Google Scholar
Reiczigel, J. & Rózsa, L. (1998). Host-mediated site-segregation of ectoparasites: an individual-based simulation study. Journal of Parasitology, 84, 491–498.CrossRefGoogle Scholar
Reinhold, K. (1999). Energetically costly behaviour and the evolution of resting metabolic rate in insects. Functional Ecology, 13, 217–224.CrossRefGoogle Scholar
Reiss, M. J. (1986). Sexual dimorphism in body size: are larger species more dimorphic? Journal of Theoretical Biology, 121, 163–172.
Reiss, M. J. (1989). The Allometry of Growth and Reproduction. Cambridge, UK: Cambridge University Press.
Reitblat, A. G. & Belokopytova, A. M. (1974). On cannibalism and predatory habits of flea larvae. Zoologicheskyi Zhurnal, 53, 135–137 (in Russian).Google Scholar
Řehaček, J. (1961). Transmission of tick-borne encephalitis virus by fleas. Journal of Hygiene, Epidemiology and Immunology (Prague), 5, 282–285 (in Russian).Google ScholarPubMed
Rementsova, M. M. (1962). Brucellosis in Wildlife. Alma-Ata, USSR: Kainar (in Russian).Google Scholar
Rendell, W. B. & Verbeek, N. A. M. (1996). Old nest material in nest boxes of tree swallows: effects on nest-site choice and nest building. Auk, 113, 319–328.CrossRefGoogle Scholar
Rensch, B. (1960). Evolution above the Species Level. New York: Columbia University Press.Google Scholar
Reshetnikova, P. I. (1959). Flea fauna of the Kustanai Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 6, 261–265 (in Russian).Google Scholar
Ribeiro, J. M. C. (1987). Role of saliva in blood feeding in arthropods. Annual Review of Entomology, 32, 463–478.CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. (1995). Blood-feeding arthropods: live syringes or invertebrate pharmacologists?Infectious Agents and Disease – Reviews Issues and Commentary, 4, 143–152.Google ScholarPubMed
Ribeiro, J. M. C. (1996). Common problems of arthropod vectors of disease. In The Biology of Disease Vectors, ed. Beaty, B. J. & Marquardt, W. C.. Niwot, CO: University of Colorado Press, pp. 25–33.Google Scholar
Ribeiro, J. M. C., Vaughan, J. A. & Farhang-Azad, A. (1990). Characterization of the salivary apyrase activity of three rodent flea species. Comparative Biochemistry and Physiology B, 95, 215–218.CrossRefGoogle ScholarPubMed
Richards, P. A. & Richards, A. G. (1969). Acanthae: a new type of cuticular process in the proventriculus of Mecoptera and Siphonaptera. Zoologische Jahrbücher, Abteilung für Anatomie und Ontogenie der Tiere, 86, 158–176.Google Scholar
Richner, H. (1996). Flohzirkus im Vogelnest: Wirt–Parasiten-Interaktionen in der Brutzeit. Ornithologische Beobachter, 93, 103–110.Google Scholar
Richner, H. (1998). Host–parasite interactions and life-history evolution. Zoology, 101, 333–344.Google Scholar
Richner, H. & Heeb, P. (1995). Are clutch and brood size patterns in birds shaped by ectoparasites?Oikos, 73, 435–441.CrossRefGoogle Scholar
Richner, H., Oppliger, A. & Christe, P. (1993). Effect of an ectoparasite on reproduction in great tits. Journal of Animal Ecology, 62, 703–710.CrossRefGoogle Scholar
Ricotta, C. (2004). A parametric diversity measure combining the relative abundances and taxonomic distinctiveness of species. Diversity and Distributions, 10, 143–146.CrossRefGoogle Scholar
Riddoch, B. J., Greenwood, M. T. & Ward, R. D. (1984). Aspects of the population structure of the sand martin flea, Ceratophyllus styx, in Britain. Journal of Natural History, 18, 475–484.CrossRefGoogle Scholar
Riek, E. F. (1970). Lower Cretaceous fleas. Nature, 227, 746–747.CrossRefGoogle ScholarPubMed
Ritter, R. C. & Epstein, A. N. (1974). Saliva lost by grooming: major item in rats' water economy. Behavioral Biology, 11, 581–585.CrossRefGoogle Scholar
Robbins, R. G. & Faulkenberry, G. D. (1982). A population model for fleas of the gray-tailed vole, Microtus canicaudus Miller. Entomological News, 93, 70–74.Google Scholar
Roberts, M. G., Smith, G. & Grenfell, B. T. (1995). Matematical models for macroparasites of wildlife. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 177–208.CrossRefGoogle Scholar
Robson, D. S. (1972). Appendix: Statistical tests of significance. Journal of Theoretical Biology, 34, 350–352.Google Scholar
Rödl, P. (1979). Investigation of the transfer of fleas among small mammals using radioactive phosphorus. Folia Parasitologica, 26, 265–274.Google ScholarPubMed
Rodrigues, A. F. S. F. & Daemon, E. (2004). Ixodideos e sifonapteros em Cerdocyon thous L. (Carnivora, Canidae) procedentes da zona da Mata Mineira, Brasil.Arquivos do Instituto Biologico São Paulo, 71, 371–372.Google Scholar
Rodríguez, Z., Moreira, E. C., Linardi, P. M. & Santos, H. A. (1999). Notes on the bat flea Hormopsylla fosteri (Siphonaptera: Ischnopsyllidae) infesting Molossops abrasus (Chiroptera). Memórias do Instituto Oswaldo Cruz, 94, 727–728.CrossRefGoogle Scholar
Rodríguez-Gironés, M. A. & Santamaría, L. (2006). A new algorithm to calculate the nestedness temperature of presence–absence matrices. Journal of Biogeography, 33, 924–935.CrossRefGoogle Scholar
Roehrig, J. T., Piesman, J., Hunt, A. R., et al. (1992). The hamster immune-response to tick-transmitted Borrelia burgdorferi differs from the response to needle-inoculated, cultured organisms. Journal of Immunology, 149, 3648–3653.Google ScholarPubMed
Rogovin, K. A., Randall, J., Kolosova, I. & Moshkin, M. (2003). Social correlates of stress in adult males of the great gerbil, Rhombomys opimus, in years of high and low population densities. Hormones and Behavior, 43, 132–139.CrossRefGoogle ScholarPubMed
Rogowitz, G. L. & Chappell, M. A. (2000). Energy metabolism of eucalyptus-boring beetles at rest and during locomotion: gender makes a difference. Journal of Experimental Biology, 203, 1131–1139.Google ScholarPubMed
Rohde, K. (1979). A critical evaluation of intrinsic and extrinsic factors responsible for niche restriction in parasites. American Naturalist, 114, 648–671.CrossRefGoogle Scholar
Rohde, K. (1985). Increased viviparity of marine parasites at high latitudes. Hydrobiologia, 137, 197–201.CrossRefGoogle Scholar
Rohde, K. (1992). Latitudinal gradients in species diversity: the search for the primary cause. Oikos, 65, 514–527.CrossRefGoogle Scholar
Rohde, K. (1994). Niche restriction in parasites: proximate and ultimate causes. Parasitology, 109, S69–S84.CrossRefGoogle ScholarPubMed
Rohde, K. (1996). Rapoport's rule is a local phenomenon and cannot explain latitudinal gradients in species diversity. Biodiversity Letters, 3, 10–13.CrossRefGoogle Scholar
Rohde, K. (1998). Is there a fixed number of niches for endoparasites of fish?International Journal for Parasitology, 28, 1861–1865.CrossRefGoogle Scholar
Rohde, K. (1999). Latitudinal gradients in species diversity and Rapoport's rule re-visited: a review of recent work and what can parasites teach us about the causes of the gradients?Ecography, 22, 593–613.CrossRefGoogle Scholar
Rohde, K., Hayward, C. & Heap, M. (1995). Aspects of the ecology of metazoan ectoparasites of marine fishes. International Journal for Parasitology, 25, 945–970.CrossRefGoogle ScholarPubMed
Rohde, K., Worthen, W., Heap, M., Hugueny, B. A. & Guégan, J.-F. (1998). Nestedness in assemblages of metazoan ecto- and endoparasites of marine fish. International Journal for Parasitology, 28, 543–549.CrossRefGoogle ScholarPubMed
Rolff, J. (2002). Bateman's principle and immunity. Proceedings of the Royal Society of London B, 269, 867–872.CrossRefGoogle ScholarPubMed
Roman, E. & Pichot, J. (1975). Fleas of mammals in bird nests during winter. Bulletin Mensuel de la Société Linnéenne de Lyon, 44, 53–57.CrossRefGoogle Scholar
Ronquist, F. (1995). Reconstructing the history of host–parasite associations using generalized parsimony. Cladistics, 11, 73–89.CrossRefGoogle Scholar
Ronquist, F. (2001). TreeFitter, Version 1.0. Available online at www.ebc.uu.se/systzoo/research/treefitter/treefitter.html.Google Scholar
Ronquist, F. & Liljeblad, J. (2001). Evolution of the gall wasp–host plant association. Evolution, 55, 2503–2522.Google ScholarPubMed
Roper, T. J., Ostler, J. R., Schmid, T. K. & Christian, S. F. (2001). Sett use in European badgers Meles meles. Behaviour, 138, 173–187.CrossRefGoogle Scholar
Roper, T. J., Jackson, T. P., Conradt, L. & Bennett, N. C. (2002). Burrow use and the influence of ectoparasites in Brants' whistling rat Parotomys brantsii. Ethology, 108, 557–564.CrossRefGoogle Scholar
Rosenfeld, J. S. (2002). Functional redundancy in ecology and conservation. Oikos, 98, 156–162.CrossRefGoogle Scholar
Rosenzweig, M. L. (1981). A theory of habitat selection. Ecology, 62, 327–335.CrossRefGoogle Scholar
Rosenzweig, M. L. (1989). Habitat selection as a source of biological diversity. Evolutionary Ecology, 1, 315–330.CrossRefGoogle Scholar
Rosenzweig, M. L. (1991). Habitat selection and population interactions: the search for mechanism. American Naturalist, 137, 5–28.CrossRefGoogle Scholar
Rosenzweig, M. L. (1992). Species diversity gradients: we know more and less than we thought. Journal of Mammalogy, 73, 715–730.CrossRefGoogle Scholar
Rosenzweig, M. L. (1995). Species Diversity in Space and Time. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Rosenzweig, M. L. & Abramsky, Z. (1985). Detecting density-dependent habitat selection. American Naturalist, 126, 405–417.CrossRefGoogle Scholar
Rosenzweig, M. L. & Ziv, Y. (1999). The echo pattern of species diversity: pattern and processes. Ecography, 22, 614–628.CrossRefGoogle Scholar
Rosenzweig, M. L., Abramsky, Z., Kotler, B. P. & Mitchell, W. A. (1985). Can interaction coefficients be determined from census data?Oecologia, 66, 194–198.CrossRefGoogle Scholar
Rossin, A. & Malizia, A. I. (2002). Relationship between helminth parasites and demographic attributes of a population of the subterranean rodent Ctenomys talarum (Rodentia: Octodontidae). Journal of Parasitology, 88, 1268–1270.CrossRefGoogle Scholar
Rothschild, N. C. (1915a). A synopsis of the British Siphonaptera. Entomological Monthly Magazine, 51, 49–112.Google Scholar
Rothschild, N. C. (1915b). On Neopsylla and some allied genera of Siphonaptera. Ectoparasites, 1, 30–44.Google Scholar
Rothschild, M. (1965a). Fleas. Scientific American, 213, 44–53.CrossRefGoogle Scholar
Rothschild, M. (1965b). The rabbit flea and hormones. Endeavour, 24, 162–168.Google Scholar
Rothschild, M. (1969). Notes of fleas: with the first record of the mermithid nematode from the order. Proceedings of the British Entomological and Natural History Society, 1, 1–8.Google Scholar
Rothschild, M. (1973). Note given at meeting on Pulex irritans found in Viking pit excavations. Proceedings of the Royal Entomological Society of London A, 38, 29.Google Scholar
Rothschild, M. (1975). Recent advances in our knowledge of the order Siphonaptera. Annual Review of Entomology, 20, 241–259.CrossRefGoogle ScholarPubMed
Rothschild, M. (1992). Neosomy in fleas, and the sessile life-style. Journal of Zoology, 226, 613–629.CrossRefGoogle Scholar
Rothschild, M. & Clay, T. (1952). Fleas, Flukes and Cuckoos: A Study of Bird Parasites, 3rd edn. London: Collins.Google Scholar
Rothschild, M. & Ford, R. (1966). Hormones of the vertebrate host controlling ovarian regression and copulation of the rabbit flea. Nature, 211, 261–266.CrossRefGoogle ScholarPubMed
Rothschild, M. & Ford, R. (1969). Does a pheromone-like factor from the nestling rabbit stimulate impregnation and maturation in the rabbit flea?Nature, 221, 1169–1170.CrossRefGoogle ScholarPubMed
Rothschild, M. & Ford, R. (1972). Breeding cycle of the flea Cediopsylla simplex is controlled by breeding cycle of host. Science, 178, 625–626.CrossRefGoogle ScholarPubMed
Rothschild, M. & Ford, R. (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). Journal of Zoology, 170, 87–137.CrossRefGoogle Scholar
Rothschild, M. & Hinton, H. E. (1968). Holding organs on the antennae of male fleas. Proceedings of the Royal Entomological Society of London A, 43, 105–107.CrossRefGoogle Scholar
Rothschild, M. & Neville, C. (1967). Fleas: insects which fly with their legs. Proceedings of the Royal Entomological Society of London C, 32, 9–10.Google Scholar
Rothschild, M. & Schlein, J. (1975). The jumping mechanism of Xenospylla cheopis. I. Exoskeletal structures and musculature. Philosophical Transactions of the Royal Society of London B, 271, 457–489.CrossRefGoogle Scholar
Rothschild, M., Ford, B. & Hughes, M. (1970). Maturation of the rabbit flea (Spilopsyllus cuniculi) and the oriental flea (Xenopsylla cheopis): some effects of mammalian hormones on development and impregnation. Transactions of the Zoological Society of London, 32, 105–188.Google Scholar
Rothschild, M., Schlein, J., Parker, K. & Sternberg, S. (1972). Jump of the oriental rat flea Xenopsylla cheopis (Roths.). Nature, 239, 45–48.CrossRefGoogle Scholar
Rothschild, M., Schlein, J., Parker, K., Neville, C. & Sternberg, S. (1973). The flying leap of the flea. Scientific American, 229, 92–100.CrossRefGoogle Scholar
Rothschild, M., Schlein, J., Parker, K., Neville, C. & Sternberg, S. (1975). The jumping mechanism of Xenospylla cheopis. III. Execution of the jump and activity. Philosophical Transactions of the Royal Society of London B, 271, 499–515.CrossRefGoogle Scholar
Roulin, A., Jeanmonod, J. & Blanc, T. (1997). Green plant material on common buzzard's (Buteo buteo) nests during the rearing of chicks. Alauda, 65, 251–257.Google Scholar
Roulin, A., Riols, C., Dijkstra, C. & Ducrest, A. (2001). Female plumage spottiness and parasite resistance in the barn owl (Tyto alba). Behavioral Ecology, 12, 103–110.CrossRefGoogle Scholar
Rousset, F., Thomas, F., MeeÛs, T. & Renaud, F. (1996). Inference of parasite-induced host mortality from distribution of parasite loads. Ecology, 77, 2203–2211.CrossRefGoogle Scholar
Roy, B. A. (2001). Patterns of association between crucifers and their flower-mimic pathogens: host jumps are more common than coevolution or cospeciation. Evolution, 55, 41–53.CrossRefGoogle ScholarPubMed
Rózsa, L., Rekasi, J. & Reiczigel, J. (1996). Relationship of host coloniality to the population ecology of avian lice (Insecta: Phthiraptera). Journal of Animal Ecology, 65, 242–248.CrossRefGoogle Scholar
Rudenchik, Y. V., Soldatkin, I. S., Klimova, Z. I. & Severova, Z. A. (1967). On the relationships between abundance of fleas and abundance of the great gerbils. In Proceedings of the 5th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 181–183 (in Russian).Google Scholar
Rudolph, D. & Knülle, W. (1982). Novel uptake systems for atmospheric water vapor among insects. Journal of Experimental Zoology, 222, 321–333.CrossRefGoogle Scholar
Ruffer, D. G. (1965). Burrows and burrowing behavior of Onychomys leucogaster. Journal of Mammalogy, 46, 241–247.CrossRefGoogle Scholar
Rust, M. K. (1992). Influence of photoperiod on egg production of cat fleas (Siphonaptera: Pulicidae) infesting cats. Journal of Medical Entomology, 29, 242–245.CrossRefGoogle ScholarPubMed
Rust, M. K. (1994). Interhost movement of adult cat fleas (Siphonaptera: Pulicidae). Journal of Medical Entomology, 31, 486–489.CrossRefGoogle Scholar
Ryba, J., Rodl, P., Bartos, L., Daniel, M. & Cerny, V. (1986). Some features of the ecology of fleas inhabiting the nests of the suslik (Citellus citellus (L.). I. Population dynamics, sex ratio, feeding, reproduction. Folia Parasitologica, 33, 265–275.Google ScholarPubMed
Ryba, J., Rodl, P., Bartos, L., Daniel, M. & Cerny, V. (1987). Some features of the ecology of fleas inhabiting the nests of the suslik (Citellus citellus (L.). II. The influence of mesostigmatid mites on fleas. Folia Parasitologica, 34, 61–68.Google Scholar
Ryckman, R. E. (1971). Plague vector studies. I. The rate of transfer of fleas among Citellus, Rattus and Sylvilagus under field conditions in southern California. Journal of Medical Entomology, 8, 535–540.CrossRefGoogle ScholarPubMed
Rytkönen, S., Lehtonen, R. & Orell, M. (1998). Breeding great tits Parus major avoid nestboxes infested with fleas. Ibis, 140, 687–690.CrossRefGoogle Scholar
Rzhevskaya, A. E., Rapoport, L. P., Orlova, L. M., Nuriev, K. K. & Suslova, L. P. (1991). Fleas in the eastern Kyzylkum Desert and their epizootic importance. Parazitologiya, 25, 504–511 (in Russian).Google Scholar
Sabilaev, A. S., Davydova, V. N. & Pole, D. S. (2003). About rodents' flea fauna on the left bank of the Ili River. Quarantinable and Zoonotic Infections in Kazakhstan, 7, 148–149 (in Russian).Google Scholar
Saino, N., Calza, S. & M⊘ller, A. P. (1998). Effects of a dipteran ectoparasite on immune response and growth trade-offs in barn swallow, Hirundo rustica, nestlings. Oikos, 81, 217–228.CrossRefGoogle Scholar
Sakaguti, K. & Jameson, E. W. (1962). The Siphonaptera of Japan. Pacific Insects Monographs, 3, 1–169.Google Scholar
Salas, V. & Herrera, E. A. (2004). Intestinal helminths of capybaras, Hydrochoerus hydrochaeris, from Venezuela. Memórias do Instituto Oswaldo Cruz, 99, 563–566.CrossRefGoogle ScholarPubMed
Salkeld, D. J. & Stapp, P. (2006). Seroprevalence rates and transmission of plague (Yersinia pestis) in mammalian carnivores. Vector-Borne and Zoonotic Diseases, 6, 231–239.CrossRefGoogle Scholar
Samarina, G. P., Alekseev, A. N. & Shiranovich, P. I. (1968). Study of fecundity of the rat fleas (Xenopsylla cheopis Rothschild and Ceratophyllus fasciatus Bosc.) when fed on different host species. Zoologicheskyi Zhurnal, 47, 261–268 (in Russian).Google Scholar
Samurov, M. A. (1985). Life history and the prognosis of abundance of fleas parasitizing gerbils of the genus Meriones in the Volga–Ural Sands. Unpublished Ph.D. thesis, Institute of Zoology of Academy of Science of the Kazakh SSR, Alma-Ata, USSR (in Russian).Google Scholar
Samurov, M. A. & Ageyev, V. S. (1983). Annual number of generations of fleas Xenopsylla conformis Wagn. (Siphonaptera, Pulicidae) in the Volga–Ural Sands. Entomological Review, 62, 226–269 (in Russian).Google Scholar
Samurov, M. A. & Yakunin, B. M. (1979). On the annual number of generations of fleas Ceratophyllus laeviceps in the Volga–Ural Sands. In Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 1, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 128–130 (in Russian).Google Scholar
Samurov, M. A. & Yakunin, B. M. (1980). Age composition and cycle of reproductive activity in female fleas Ceratophyllus laeviceps Wagn. (Siphonaptera) in the Volga–Ural Sands. Entomological Review, 59, 510–512 (in Russian).Google Scholar
Sapegina, V. F. (1976). Distribution of fleas parasitic on small mammals and birds in the southern taiga of the Pri–Irtyshie Region. Parazitologiya, 10, 397–400 (in Russian).Google Scholar
Sapegina, V. F. (1988). Fleas of small mammals and birds in the forest-park area of the city of Novosibirsk. Parazitologiya, 22, 132–136 (in Russian).Google ScholarPubMed
Sapegina, V. F. & Kharitonova, N. N. (1969). On the ability of bird fleas to transmit virus of the Omsk haemorhaggic fever in the experiment. In Migrating Birds and their Role in the Circulation of Arboviruses, ed. Tcherepanov, A. I.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 263–267 (in Russian).Google Scholar
Sapegina, V. F., Yudin, B. S. & Dudareva, G. V. (1980a). Materials on the biology of fleas of the Taimyr and Gydanskyi Penunsulae. In Parasitic Insects and Ticks of Siberia, ed. Davydova, M. S.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 225–231 (in Russian).Google Scholar
Sapegina, V. F., Ravkin, Y. S., Lukianova, I. V. & Sebeleva, G. G. (1980b). Fleas of the forest zone of western Siberia. In Problems of Zoogeography and Faunal History, ed. Belyshev, B. F. & Ravkin, Y. S.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 94–166 (in Russian).Google Scholar
Sapegina, V. F., Lukianova, I. V. & Fomin, B. N. (1981a). Fleas of small mammals in northern foothills of the Altai Mountains and Upper Ob River Region. In Biological Problems of Natural Foci, ed. Maximov, A. A.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 167–176 (in Russian).Google Scholar
Sapegina, V. F., Yudin, B. S. & Yudina, S. A. (1981b). Fleas of small mammals in the northern taiga of the southern Taimyr Peninsula. Bulletin of the Siberian Branch of the Academy of Sciences of the USSR, Series Biological Sciences, 1, 96–104 (in Russian).Google Scholar
Sapegina, V. F., Vartapetov, L. G. & Pokrovskaya, I. V. (1990). The fleas of small mammals in the northern taiga of western Siberia. Parazitologiya, 24, 56–62 (in Russian).Google ScholarPubMed
Sarfati, M., Krasnov, B. R., Ghazaryan, L., et al. (2005). Energy costs of blood digestion in a host-specific haematophagous parasite. Journal of Experimental Biology, 208, 2489–2496.CrossRefGoogle Scholar
Sasal, P., Trouvé, S., Müller-Graf, C. & Morand, S. (1999). Specificity and host predictability: a comparative analysis among monogenean parasites of fish. Journal of Animal Ecology, 68, 437–444.CrossRefGoogle Scholar
Saxena, V. K. (1987). Rodent–ectoparasite association in selected biotopes of Mirzapur and Varanasi districts of Uttar Pradesh. Journal of Communicable Diseases, 19, 310–316.Google ScholarPubMed
Saxena, V. K. (1999). Mesostigmatid mite infestations of rodents in diverse biotopes of central and southern India. Journal of Parasitology, 85, 147–149.CrossRefGoogle ScholarPubMed
Scalon, O. I. (1981). On fleas from eastern Mongolia with description of male and female of Echidnophaga tiscadaea Smit, 1967 (Siphonaptera). Parazitologiya, 15, 280–287 (in Russian).Google Scholar
Schall, J. J. (1990). The ecology of lizard malaria. Parasitology Today, 6, 264–269.CrossRefGoogle ScholarPubMed
Scharf, W. C. (1991). Geographic distribution of Siphonaptera collected from small mammals on Lake Michigan islands. Great Lakes Entomologist, 24, 39–43.Google Scholar
Scharf, W. C. (1998). Fleas (Siphonaptera) from migrating owls: passengers on the journey. Michigan Birds and Natural History, 5, 167–171.Google Scholar
Scheidt, V. J. (1988). Flea allergy dermatitis. In The Veterinary Clinics of North America: Small Animal Practice, ed. White, S. D.. Philadelphia, PA: Saunders College Publishing, pp. 1023–1042.Google Scholar
Schelhaas, D. P. & Larson, O. R. (1989). Cold hardiness and winter survival in the bird flea, Ceratophyllus idius. Journal of Insect Physiology, 35, 149–153.CrossRefGoogle Scholar
Schemmer, K. R. & Halliwell, R. E. (1987). Efficacy of alum-precipitated flea antigen for hyposensitization of flea-allergic dogs. Seminars in Veterinary Medicine and Surgery (Small Animal), 2, 195–198.Google ScholarPubMed
Schlein, Y. (1980). Morphological similarities between the skeletal structures of Siphonaptera and Mecoptera. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 359–367.Google Scholar
Schluter, D. (1984). A variance test for detecting species associations, with some example applications. Ecology, 65, 998–1005.CrossRefGoogle Scholar
Schmid-Hempel, P. (2003). Variation in immune defence as a question of evolutionary ecology. Proceeding of the Royal Society of London B, 270, 357–366.CrossRefGoogle ScholarPubMed
Schmid-Hempel, P. & Ebert, D. (2003). On the evolutionary ecology of specific immune defence. Trends in Ecology and Evolution, 18, 27–32.CrossRefGoogle Scholar
Schmidt-Nielsen, K. (1990). Animal Physiology: Adaptation and Environment, 4th edn. Cambridge, UK: Cambridge University Press.Google Scholar
Schoener, T. W. (1974). Competition and the form of the habitat shift. Theoretical Population Biology, 6, 265–307.CrossRefGoogle ScholarPubMed
Schofield, S. & Torr, S. J. (2002). A comparison of feeding behaviour of tsetse and stable flies. Medical and Veterinary Entomology, 16, 177–185.CrossRefGoogle ScholarPubMed
Schönrogge, K., Gardner, M. G., Elmes, G. W., et al. (2006). Host propagation permits extreme local adaptation in a social parasite of ants. Ecological Letters, 9, 1032–1040.CrossRefGoogle Scholar
Schradin, C. & Pillay, N. (2005). Demography of the striped mouse (Rhabdomys pumilio) in the succulent karoo. Mammalian Biology, 70, 84–92.CrossRefGoogle Scholar
Schwan, T. G. (1975). Flea reinfestation on the California meadow vole (Microtus californicus). Journal of Medical Entomology, 21, 760.CrossRefGoogle Scholar
Schwan, T. G. (1986). Seasonal abundance of fleas (Siphonaptera) on grassland rodents in Lake Nakuru National Park, Kenya, and potential for plague transmission. Bulletin of Entomological Research, 76, 633–648.CrossRefGoogle Scholar
Schwan, T. G. (1993). Sex ratio and phoretic mites of fleas (Siphonaptera, Pulicidae and Hystrichopsyllidae) on the Nile grass rat (Arvicanthis niloticus) in Kenya. Journal of Medical Entomology, 30, 122–135.CrossRefGoogle Scholar
Schwan, T. G. & Schwan, V. R. (1980). Observations on the fleas Xenopsylla debilis and Xenopsylla difficilis (Siphonaptera) infesting the gerbil Tatera nigricauda in Southern Kenya. African Journal of Ecology, 18, 267–272.CrossRefGoogle Scholar
Scofield, A., Riera, M. D. F., Eliseri, C., Massard, C. L. & Linardi, P. M. (2005). Ocorrência de Rhopalopsyllus lutzi lutzi (Baker) (Siphonaptera, Rhopalopsyllidae) em Canis familiaris (Linnaeus) de zona rural do município de Piraí, Rio de Janeiro, Brasil. Revista Brasileira de Entomologia, 49, 159–161.CrossRefGoogle Scholar
Seal, S. C. & Bhattacharji, L. M. (1961). Epidemiological studies of plague in Calcutta. I. Bionomics of two species of rat fleas and distribution, densities and resistance of rodents in relation to the epidemiology of plague in Calcutta. Indian Journal of Medical Research, 49, 974–1007.Google Scholar
Segerman, J. (1995). Siphonaptera of Southern Africa: handbook for the identification of fleas. Publications of the South African Institute for Medical Research, 57, 1–264.Google Scholar
Serzhan, O. S. (2002). Pathways of evolution of the faunistic complexes of rodent fleas in Kazakhstan, Middle Asia and adjacent regions and their role in the endemism of plague. Quarantinable and Zoonotic Diseases in Kazakhstan, 6, 83–90 (in Kazakh).Google Scholar
Serzhan, O. S. & Ageyev, V. S. (2000). Geographical distribution and host complexes of plague-infected fleas in relation to some problems of paleogenesis of plague enzootics. Quarantinable and Zoonotic Diseases in Kazakhstan, 2, 183–192 (in Russian).Google Scholar
Sevenster, J. G. (1996). Aggregation and coexistence. I. Theory and analysis. Journal of Animal Ecology, 65, 297–307.CrossRefGoogle Scholar
Sgonina, K. (1935). Die Reizphysiologie des Igelflöhs (Archaeopsylla erinacei Bouché) und seiner Larve. Zeitschrift für Parasitenkunde, 7, 539–571.CrossRefGoogle Scholar
Shafi, M. M., Ali, R., Ghazi, R. R. & Noor, U. N. (1988). Flea index studies of synanthropic rats in Karachi, Pakistan. Acta Parasitologica Polonica, 33, 185–194.Google Scholar
Shaftesbury, A. D. (1934). The Siphonaptera (fleas) of North Carolina, with special reference to sex ratios. Journal of the Elisha Mitchell Scientific Society, 49, 247–263.Google Scholar
Shargal, E., Kronfeld-Schor, N. & Dayan, T. (2000). Population biology and spatial relationships of coexisting spiny mice (Acomys) in Israel. Journal of Mammalogy, 81, 1046–1052.2.0.CO;2>CrossRefGoogle Scholar
Sharif, M. (1949). Effects of constant temperature and humidity on the development of the larvae and the pupae of the three Indian species of Xenopsylla (Insecta: Siphonaptera). Philosophical Transactions of the Royal Society of London B, 233, 581–633.CrossRefGoogle Scholar
Shaw, D. J. & Dobson, A. P. (1995). Patterns of macroparasite abundance and aggregation in wildlife populations: a quantitative review. Parasitology, 111, S111–S127.CrossRefGoogle ScholarPubMed
Shaw, D. J., Grenfell, B. T. & Dobson, A. P. (1998). Patterns of macroparasite aggregation in wildlife host populations. Parasitology, 117, 597–610.CrossRefGoogle ScholarPubMed
Shaw, J. L. & Moss, R. (1990). Effect of the caecal nematode Trichostrongylus tenius on egg-laying by captive red grouse. Research in Veterinary Science, 48, 253–258.Google Scholar
Shaw, S. E., Kenny, M. J., Tasker, S. & Birtles, R. J. (2004). Pathogen carriage by the cat flea Ctenocephalides felis (Bouché) in the United Kingdom. Veterinary Microbiology, 102, 183–188.CrossRefGoogle ScholarPubMed
Shchedrin, V. I. (1974). Morphological and histochemical data on the blood digestion in some flea species: vectors of plague. Unpublished Ph.D. thesis, All-Union Scientific Anti-Plague Institute ‘Microb’, Saratov, USSR (in Russian).Google Scholar
Shchedrin, V. I., Loktev, N. A. & Lunina, E. A. (1974). Morphological and histochemical data on digestion in fleas P. irritans. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 281–283 (in Russian).Google Scholar
Sheldon, B. C. & Verhulst, S. (1996). Ecological immunology: costly parasite defenses and trade-offs in evolutionary ecology. Trends in Ecology and Evolution, 11, 317–321.CrossRefGoogle Scholar
Shenbrot, G. I. & Krasnov, B. R. (2002). Can interaction coefficients be determined from census data? Testing two estimation methods with Negev Desert rodents. Oikos, 99, 47–58.CrossRefGoogle Scholar
Shenbrot, G. I., Krasnov, B. R. & Khokhlova, I. S. (1994). On the biology of Gerbillus henleyi (Rodentia: Gerbillidae) in the Negev Highlands, Israel. Mammalia, 58, 581–589.CrossRefGoogle Scholar
Shenbrot, G. I., Sokolov, V. E., Geptner, V. G. & Kovalskaya, Y. M. (1995). The Mammals of Russia and Adjacent Regions: Dipodoidea.Moscow, Russia: Nauka (in Russian).Google Scholar
Shenbrot, G. I., Krasnov, B. R. & Khokhlova, I. S. (1997). On the biology of Wagner's gerbil Gerbillus dasyurus (Wagner, 1842) (Rodentia: Gerbillidae) in the Negev Highlands, Israel. Mammalia, 61, 467–486.Google Scholar
Shenbrot, G. I., Krasnov, B. R. & Rogovin, K. A. (1999a). Spatial Ecology of Desert Rodent Communities. New York: Springer-Verlag.CrossRefGoogle Scholar
Shenbrot, G. I., Krasnov, B. R. & Khokhlova, I. S. (1999b). Notes on the biology of the bushy-tailed jird, Sekeetamys calurus, in the central Negev, Israel. Mammalia, 63, 374–377.Google Scholar
Shenbrot, G. I., Krasnov, B. R., Khokhlova, I. S., Demidova, T. & Fielden, L. J. (2002). Habitat-dependent differences in architecture and microclimate of the Sundevall's jird (Meriones crassus) burrows in the Negev Desert, Israel. Journal of Arid Environments, 51, 265–279.CrossRefGoogle Scholar
Shenbrot, G. I., Krasnov, B. R. & Lu, L. (2007). Geographic range size and host specificity in ectoparasites: A case study with Amphipsylla fleas and rodent hosts. Journal of Biogeography, 34, 1679–1690.CrossRefGoogle Scholar
Shepherd, R. C. H. & Edmonds, J. W. (1979). The distribution of the stickfast fleas, Echidnophaga myrmecobii Rothschild and E. perilis Jordan, on the wild rabbit, Oryctolagus cuniculus (L.). Australian Journal of Zoology, 27, 261–271.CrossRefGoogle Scholar
Shevchenko, V. L., Samurov, M. A., Kaimashnikov, V. I. & Polyakov, V. K. (1971). Some patterns of variation in the abundance of the midday jirds and fleas Xenopsylla conformis in the Volga–Ural Sands. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 449–450 (in Russian).Google Scholar
Shevchenko, V. L., Grazhdanov, A. K., Zharinova, L. K. & Andreeva, T. A. (1976). Abilities of avian fleas Frontopsylla frontalis alatau Fed., 1946 to infect rodents with plague. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 45, 49–52 (in Russian).Google ScholarPubMed
Shields, W. M. & Crook, J. R. (1987). Barn swallow coloniality: a net cost for group breeding in the Adirondacks. Ecology, 68, 1373–1386.CrossRefGoogle Scholar
Shinozaki, Y., Sh, T. iibashi, Yoshizawa, K., et al. (2004). Ectoparasites of the Pallas squirrel, Callosciurus erythraeus, introduced to Japan. Medical and Veterinary Entomology, 18, 61–63.CrossRefGoogle ScholarPubMed
Shorrocks, B. (1996). Local diversity: a problem with too many solutions. In The Genesis and Maintenance of Biological Diversity, ed. Hochberg, M., Clobert, J. & Barbault, R.. Oxford, UK: Oxford University Press, pp. 104–122.Google Scholar
Shorrocks, B. & Rosewell, J. (1986). Guild size in drosophilids: a simulation model. Journal of Animal Ecology, 55, 527–541.CrossRefGoogle Scholar
Shryock, J. A. & Houseman, R. M. (2005). A comparison of fecal protein content in male and female cat fleas, Ctenocephalides felis (Bouché) (Siphonaptera: Pulicidae). Florida Entomologist, 88, 335–337.CrossRefGoogle Scholar
Shubber, A. H., Lloyd, S. & Soulsby, E. J. L. (1981). Infection with gastrointestinal helminths: effect of lactation and maternal transfer of immunity. Parasitology Research, 65, 181–189.Google ScholarPubMed
Shudo, E. & Iwasa, Y. (2001). Inducible defense against pathogens and parasites: optimal choice among multiple options. Journal of Theoretical Biology, 209, 233–247.CrossRefGoogle ScholarPubMed
Shulov, A. & Naor, D. (1964). Experiments on the olfactory responses and host specificity of the Oriental rat flea (Xenopsylla cheopis). Parasitology, 54, 225–231.CrossRefGoogle Scholar
Shurin, J. B. & Allen, E. G. (2001). Effects of competition, predation, and dispersal on species richness at local and regional scales. American Naturalist, 158, 624–637.CrossRefGoogle ScholarPubMed
Shurin, J. B., Havel, J. E., Leibold, M. A. & Pinel-Alloul, B. (2000). Local and regional zooplankton species richness: a scale-independent test for saturation. Ecology, 81, 3062–3073.CrossRef
Shutler, D. & Campbell, A. A. (2007). Experimental addition of greenery reduces flea loads in nests of a non-greenery using species, the tree swallow Tachycineta bicolor. Journal of Avian Biology, 38, 7–12.CrossRefGoogle Scholar
Shutler, D., Petersen, S. D., Dawson, R. D. & Campbell, A. (2003). Sex ratios of fleas (Siphonaptera: Ceratophyllidae) in nests of tree swallows (Passeriformes: Hirundinidae) exposed to different chemicals. Environmental Entomology, 32, 1045–1048.CrossRefGoogle Scholar
Shutler, D., Mullie, A. & Clark, R. G. (2004). Tree swallow reproductive investment, stress, and parasites. Canadian Journal of Zoology, 82, 442–448.CrossRefGoogle Scholar
Shwartz, E. A., Berendyaeva, E. L. & Grebenyuk, R. V. (1958). Fleas parasitic on rodents in the Frunze Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 4, 255–261 (in Russian).Google Scholar
Siemann, E. (1998). Experimental tests of effects of plant productivity and diversity on grassland arthropod diversity. Ecology, 79, 2057–2070.CrossRefGoogle Scholar
Silverman, J. & Appel, A. G. (1994). Adult cat flea (Siphonaptera, Pulicidae) excretion of host blood proteins in relation to larval nutrition. Journal of Medical Entomology, 31, 265–271.CrossRefGoogle ScholarPubMed
Silverman, J. & Rust, M. K. (1983). Some abiotic factors affecting the survival of the cat flea, Ctenocephalides felis (Siphonaptera: Pulicidae). Environmental Entomology, 12, 490–495.CrossRefGoogle Scholar
Silverman, J. & Rust, M. K. (1985). Extended longevity of the pre-emerged adult of the cat flea (Siphonaptera: Pulicidae) and factors stimulating emergence from the pupal cocoon. Annals of the Entomological Society of America, 78, 763–768.CrossRefGoogle Scholar
Silverman, J., Rust, M. K. & Reierson, D. A. (1981). Influence of temperature and humidity on survival and development of the cat flea, Ctenocephalides felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 18, 78–83.CrossRefGoogle Scholar
Silvertown, J., Franco, M. & Harper, J. L. (eds.) (1997). Plant Life Histories: Ecology, Phylogeny and Evolution. New York: Cambridge University Press.Google Scholar
Simiczyjew, B. & Margas, W. (2001). Ovary structure in the bat flea Ischnopsyllus spp. (Siphonaptera: Ischnopsyllidae): phylogenetic implications. Zoologica Poloniae, 46, 5–14.Google Scholar
Šimková, A., Desdevises, Y., Gelnar, M. & Morand, S. (2000). Coexistence of nine gill ectoparasites (Dactylogyrus: Monogenea) parasitizing the roach (Rutilus rutilus L.): history and present ecology. International Journal for Parasitology, 30, 1077–1088.CrossRefGoogle Scholar
Šimková, A., Desdevises, Y., Gelnar, M. & Morand, S. (2001). Morphometric correlates of host specificity in Dactylogyrus species (Monogenea) parasites of European cyprinid fish. Parasitology, 123, 169–177.CrossRefGoogle ScholarPubMed
Šimková, A., Kadlec, D., Gelnar, M. & Morand, S. (2002). Abundance–prevalence relationship of gill congeneric ectoparasites: testing the core–satellite hypothesis and ecological specialization. Parasitology Research, 88, 682–686.Google Scholar
Šimková, A., Sitko, J., Okulewicz, J. & Morand, S. (2003). Occurrence of intermediate hosts and structure of digenean communities of the black-headed gull, Larus ridibundus (L.). Parasitology, 126, 69–78.CrossRefGoogle Scholar
Šimková, A., Verneau, O., Gelnar, M. & Morand, S. (2006). Specificity and specialization of congeneric monogeneans parasitizing cyprinid fish. Evolution, 60, 1023–1037.CrossRefGoogle ScholarPubMed
Sinelshchikov, V. A. (1956). Study of flea fauna of the Pavlodar Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 2, 147–153 (in Russian).Google Scholar
Singh, S. K. & Girschick, H. J. (2003). Tick–host interactions and their immunological implications in tick-borne diseases. Current Science, 85, 1284–1298.Google Scholar
Skinner, J. D. & Smithers, H. N. (1990). The Mammals of the Southern African Subregion, 2nd edn. Pretoria, South Africa: University of Pretoria Press.Google Scholar
Skorping, A., Read, A. F. & Keymer, A. E. (1991). Life history covariation in intestinal nematodes of mammals. Oikos, 60, 365–372.CrossRefGoogle Scholar
Skuratowicz, W. (1960). Pchly (Aphaniptera) ptaków i ssaków Bialowieskiego Parku Narodowego. Annales Zoologici (Warszawa), 19, 1–32.Google Scholar
Skuratowicz, W. (1967). Pchly – Siphonaptera (Aphaniptera): Klucze do Oznacznia Owadow Polski 29. Warsaw: Panstwowe Wydawnictwo Naukowe.Google Scholar
Slomczyński, R., Kaliński, A., Wawrzyńiak, J., et al. (2006). Effects of experimental reduction in nest micro-parasite and macro-parasite loads on nestling hemoglobin level in blue tits Parus caeruleus. Acta Oecologica, 30, 223–227.CrossRefGoogle Scholar
Slonov, M. N. (1965). On the biology of a flea Ceratophyllus tamias Wagn., 1927. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 34, 485–487 (in Russian).Google ScholarPubMed
Smetana, A. (1965). On the transmission of tick-borne encephalitis by fleas. Acta Virologica, 9, 375–379.Google ScholarPubMed
Smit, F. G. A. M. (1954). Lopper: Danmarks Fauna, 60. Copenhagen, Denmark: G. E. C. Gads Forlag.Google Scholar
Smit, F. G. A. M. (1962a). Neotunga euloidea gen. n., sp. n. (Siphonaptera, Pulicidae). Bulletin of the British Museum of Natural History, Entomology, 12, 365–378.CrossRefGoogle Scholar
Smit, F. G. A. M. (1962b). Siphonaptera collected from moles and their nests at Wilp, Netherlands, by Jhr. W. C. Van Heurn. Tijdschrift voor Entomologie, 105, 29–44.Google Scholar
Smit, F. G. A. M. (1972). On some adaptive structures in Siphonaptera. Folia Parasitologica, 19, 5–17.Google ScholarPubMed
Smit, F. G. A. M. (1974). Siphonaptera collected by Dr J. Martens in Nepal. Senckenbergiana Biologica, 55, 357–398.Google Scholar
Smit, F. G. A. M. (1977). An unusual form of the stick-tight flea Echidnophaga gallinacea. Revue Zoologique Africaine, 91, 198–199.Google Scholar
Smit, F. G. A. M. (1978). Fossil ‘fleas’. Flea News, 14, 1–2.Google Scholar
Smit, F. G. A. M. (1982). Classification of the Siphonaptera. In Synopsis and Classification of Living Organisms, vol. 2, ed. Parker, S.. New York: McGraw-Hill, pp. 557–563.Google Scholar
Smit, F. G. A. M. (1987). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptern) in the British Museum (Natural History), vol. 7, Malacopsylloidea (Malacopsyllidae and Rhopalopsyllidae). London: Oxford University Press.
Smith, A. (1951). The effect of relative humidity on the activity of the tropical rat flea Xenopsylla cheopis (Roths.). Bulletin of Entomological Research, 42, 585–600.CrossRefGoogle Scholar
Smith, A. (1980). Lack of interspecific interactions of Everglades rodents on two spatial scales. Acta Theriologica, 25, 61–70.CrossRefGoogle Scholar
Smith, A., Telfer, S., Burthe, S., Bennett, M. & Begon, M. (2005). Trypanosomes, fleas and field voles: ecological dynamics of a host–vector–parasite interaction. Parasitology, 131, 355–365.CrossRefGoogle ScholarPubMed
Smith, F. A., Brown, J. H., Haskell, J. P., et al. (2004). Similarity of mammalian body size across the taxonomic hierarchy and across space and time. American Naturalist, 163, 672–691.CrossRefGoogle ScholarPubMed
Smith, S. A. & Clay, M. E. (1985). Morphology of the antennae of the bat flea Myodopsylla insignis (Siphonaptera: Ischnopsyllidae). Journal of Medical Entomology, 22, 64–71.CrossRefGoogle Scholar
Smits, J. E., Bortolotti, G. R. & Tella, J. L. (1999). Simplifying the phytohaemagglutinin skin-testing technique in studies of avian immunocompetence. Functional Ecology, 13, 567–572.CrossRefGoogle Scholar
Snodgrass, R. E. (1944). The feeding apparatus of biting and sucking insects affecting man and animals. Smithsonian Miscellaneous Collections, 104, 1–107.Google Scholar
Sobey, W. R., Menzies, W. & Conolly, D. (1974). Myxomatosis: some observations on breeding the European rabbit flea Spilopsyllus cuniculi (Dale) in an animal house. Journal of Hygiene, 71, 453–465.CrossRefGoogle Scholar
Sokolova, A. A. & Popova, A. S. (1969). On the biology of fleas Coptopsylla lamellifer. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 90–92 (in Russian).Google Scholar
Sokolova, A. A., Balabas, N. G. & Trofimenko, I. P. (1971). Data on reproduction of Coptopsylla lamellifer in the Moyynkum Desert. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 416–417 (in Russian).Google Scholar
Soliman, S., Marzouk, A. S., Main, A. J. & Montasser, A. A. (2001). Effect of sex, size, and age of commensal rat hosts on the infestation parameters of their ectoparasites in a rural area of Egypt. Journal of Parasitology, 87, 1307–1316.CrossRefGoogle Scholar
Soloshenko, I. Z. (1958). Role of haematophagous arthropods in the maintenance of the epizootics of leptospiroses in the foci of these diseases. In Proceedings of the 10th Conference on Parasitological Problems and Diseases with Natural Focality, ed. Anonymous. Moscow–Leningrad, USSR, pp. 139–140 (in Russian).Google Scholar
Soloshenko, I. Z. (1962). Role of haematophagous arthropods in transmission and maintanence of pathogenous leptospires. II. Relationships between haematophagous arthropods and leptospiroses. Journal of Microbiology [Zhurnal Mikrobiologii], 4, 31–34 (in Russian).Google Scholar
Solovieva, A. V., Alania, I. I. & Kosminsky, R. B. (1976). On the ecology of fleas Ctenophthalmus (Euctenophthalmus) strigosus Rostigaev et Zolotova, 1964 (Ctenophthalmidae, Siphonaptera) in southern Trans-Caucasus. Problems of Particularly Dangerous Diseases, 51, 46–49 (in Russian).Google Scholar
Sorci, G. (1996). Patterns of haemogregarine load, aggregation and prevalence as a function of host age in the lizard Lacerta vivipara. Journal of Parasitology, 82, 676–678.CrossRefGoogle ScholarPubMed
Sorci, G., Defraipont, M. & Clobert, J. (1997). Host density and ectoparasite avoidance in the common lizard (Lacerta vivipara). Oecologia, 11, 183–188.CrossRefGoogle Scholar
Soshina, Y. F. (1973). The rate of flea infestation in common myomorph rodents in the forest zone of the Krym Mountains. Parazitologiya, 7, 31–35 (in Russian).Google Scholar
Sosnina, E. F. (1967a). The dependence of the infestation and species composition of rodent ectoparasites on the host's habitat (on the example of Rattus turkestanicus). Wiadomosci Parazytologiczne, 13, 637–641.Google Scholar
Sosnina, E. F. (1967b). An attempt of biocoenotical analysis of the assemblages of arthropods collected from rodents. Parasitological Collection, 23, 61–69 (in Russian).Google Scholar
Sotnikova, A. N. & Soldatov, G. M. (1969). Extraction of the virus of the tick-borne encephalitis from fleas Ceratophyllus tamias Wagn. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 33, 622–624 (in Russian).Google Scholar
Southwood, T. R. E. (1966). Ecological Methods. London: Chapman & Hall.Google Scholar
Souza, W. P. (1994). Patterns and processes in communities of helminth parasites. Trends in Ecology and Evolution, 9, 52–57.CrossRefGoogle Scholar
Sreter-Lancz, Z., Tornyai, K., Szell, Z., Sreter, T. & Marialigeti, K. (2006). Bartonella infections in fleas (Siphonaptera: Pulicidae) and lack of Bartonellae in ticks (Acari: Ixodidae) from Hungary. Folia Parasitologica, 53, 313–316.CrossRefGoogle ScholarPubMed
Srivastava, D. (1999). Using local–regional richness plots to test for species saturation: pitfalls and potentials. Journal of Animal Ecology, 68, 1–16.CrossRefGoogle Scholar
Stanko, M. (1987). Siphonaptera of small mammals in the northern part of the Krupina Plain. Stredné Slovensko, Zborník Stredoslovenského Múzea, Banská Bystrica, 6, 108–117 (in Slovak).Google Scholar
Stanko, M. (1988). Fleas (Siphonaptera) of small mammals in eastern part of the Volovské Vrchy Mountains. Acta Rerum Naturalium Musei Nationalis Slovaci Bratislava, 34, 29–40 (in Slovak).Google Scholar
Stanko, M. (1994). Fleas synusy (Siphonaptera) of small mammals from the central part of the East-Slovakian lowlands. Biologia, Bratislava, 49, 239–246.Google Scholar
Stanko, M., Miklisova, D., Goüy de Bellocq, J. & Morand, S. (2002). Mammal density and patterns of ectoparasite species richness and abundance. Oecologia, 131, 289–295.CrossRefGoogle ScholarPubMed
Stanko, M., Krasnov, B. R. & Morand, S. (2006). Relationship between host density and parasite distribution: inferring regulating mechanisms from census data. Journal of Animal Ecology, 75, 575–583.CrossRefGoogle ScholarPubMed
Stanko, M., Krasnov, B. R., Miklisova, D. & Morand, S. (2007). Simple epidemiological model predicts the relationships between prevalence and abundance in ixodid ticks. Parasitology, 134, 59–68.CrossRefGoogle ScholarPubMed
Starikov, V. P. & Sapegina, V. F. (1987). Ectoparasites of small mammals in the forest-steppe Trans-Ural Region. In Ecology and Geography of Arthropods in Siberia, ed. Tcherepanov, A. I.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 76–83 (in Russian).Google Scholar
Stark, H. E. (2002). Population dynamics of adult fleas (Siphonaptera) on hosts and in nests of the California vole. Journal of Medical Entomology, 39, 818–824.CrossRefGoogle ScholarPubMed
Stark, H. E. & Miles, V. I. (1962). Ecological studies of wild rodent plague in the San Francisco Bay area of California. VI. The relative abundance of the certain flea species and their host relationships on coexisting wild and domestic rodents. American Journal of Tropical Medicine and Hygiene, 11, 525–534.CrossRefGoogle ScholarPubMed
Starozhitskaya, G. S. (1968). Effect of gonotrophic cycle of fleas on the duration of their uninterrupted stay on a host. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 59–64 (in Russian).Google Scholar
Starozhitskaya, G. S. (1970). Diapause in fleas of the genus Xenopsylla and its epizootological importance. Problems of Particularly Dangerous Diseases, 13, 148–155 (in Russian).Google Scholar
Statzner, B., Dolédec, S. & Hugueny, B. (2004). Biological trait composition of European stream invertebrate communities: assessing the effects of various trait filter types. Ecography, 27, 470–488.CrossRefGoogle Scholar
Stearns, S. C. (1992). The Evolution of Life Histories. New York: Oxford University Press.Google Scholar
Steele, W. K., Pilgrim, R. L. C. & Palma, R. L. (1997). Occurrence of the flea Glaciopsyllus antarcticus and avian lice in central Dronning Maud Land. Polar Biology, 18, 292–294.CrossRefGoogle Scholar
Stenseth, N. C, Samia, N. I., Viljugrein, H., et al. (2006). Plague dynamics are driven by climate variation. Proceedings of the National Academy of Sciences of the USA, 103, 13110–13115.CrossRefGoogle ScholarPubMed
Stepanova, N. A. & Mitropolsky, O. V. (1971). Relationships between co-occurring fleas Xenopsylla hirtipes and Xenopsylla gerbilli. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 418–420 (in Russian).Google Scholar
Stepanova, N. A. & Mitropolsky, O. V. (1977). Spatial distribution of two sympathric species of fleas parasitic on the great gerbil in the Kyzyl-Kum Desert. Parazitologiya, 11, 147–152 (in Russian).Google Scholar
Stevens, G. C. (1989). The latitudinal gradient in geographical range: how so many species coexist in the tropics. American Naturalist, 133, 240–256.CrossRefGoogle Scholar
Stevenson, H. L., Bai, Y., Kosoy, M. Y., et al. (2003). Detection of novel Bartonella strains and Yersinia pestis in pairie dogs and their fleas (Siphonaptera: Ceratophyllidae and Pulicidae) using multiplex polymerase chain reaction. Journal of Medical Entomology, 40, 329–337.CrossRefGoogle ScholarPubMed
Stevenson, H. L., Labruna, M. B., Montenieri, J. A., et al. (2005). Detection of Rickettsia felis in a New World flea species, Anomiopsyllus nudata (Siphonaptera: Ctenophthalmidae). Journal of Medical Entomology, 42, 163–167.CrossRefGoogle Scholar
Stewart, M. A. & Evans, F. C. (1941). A comparative study of rodent and burrow flea populations. Proceedings of the Society for Experimental Biology and Medicine, 47, 140–142.CrossRefGoogle Scholar
Stewart, P. D. & MacDonald, D. W. (2003). Badgers and badger fleas: strategies and counter-strategies. Ethology, 109, 751–764.CrossRefGoogle Scholar
Stireman, J. O. (2005). The evolution of generalization? Parasitoid flies and the perils of inferring host range evolution from phylogenies. Journal of Evolutionary Biology, 18, 325–336.CrossRefGoogle ScholarPubMed
Stone, L. & Roberts, A. (1991). Conditions for a species to gain advantage from the presence of competitors. Ecology, 72, 1964–1972.CrossRefGoogle Scholar
Strahan, R. (ed.) (1983). The Complete Book of Australian Mammals. North Ryde, Australia: Collins, Angus & Robertson.Google Scholar
Streilein, J. W. (1990). Skin associated lymphoid tissues (SALT): the next generation. In Skin Immune System (SIS), ed. Bos, J. D.. Boca Raton, FL: CRC Press, pp. 25–48.Google Scholar
Studdert, V. P. & Arundel, J. H. (1988). Dermatitis of the pinnae of cats in Australia associated with the European rabbit flea (Spilopsyllus cuniculi). Veterinary Record, 123, 624–625.Google Scholar
Štys, P. & Bilinski, S. M. (1990). Ovariole types and the phylogeny of hexapods. Biology Review, 65, 401–429.CrossRefGoogle Scholar
Sukhanova, V. I., Tchernikina, M. A., Sosnovtseva, V. P., et al. (1978). Multiannual dynamics of abundance in fleas parasitic on the great gerbil in northern Turkmenistan. Problems of Particularly Dangerous Diseases, 60, 53–57 (in Russian).Google Scholar
Sukhdeo, M. V. K. (1997). Earth's third environment: the worm's eye view. BioScience, 47, 141–149.Google Scholar
Sukhdeo, M. V. K. (2000). Inside the vertebrate host: ecological strategies by parasites living in the third environment. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R., Morand, S. & Skorping, A.. Amsterdam, the Netherlands: Elsevier Science, pp. 43–62.Google Scholar
Sukhdeo, M. V. K. & Bansemir, A. D. (1996). Critical resources that influence habitat selection decisions by gastrointestinal helminth parasites. International Journal for Parasitology, 26, 483–498.CrossRefGoogle ScholarPubMed
Sukhdeo, M. V. K., Sukhdeo, S. C. & Bansemir, A. D. (2002). Interactions between intestinal nematodes and vertebrate hosts. In The Behavioural Ecology of Parasites, ed. Lewis, E. E., Campbell, J. F. & Sukhdeo, M. V. K.. Wallingford, UK: CAB International, pp. 223–242.CrossRefGoogle Scholar
Suleimenov, B. M. (2004). Mechanism of Plague Enzootic. Almaty, Kazakhstan: Almaty (in Russian).Google Scholar
Suntsov, V. V. & Suntsova, N. I. (2003). Origin and genesis of natural and anthropogenic plague foci: ecological, geographical and social aspects. In Ecological and Epizootological Aspects of Plague in Vietnam, ed. Korzun, L. P. & Suntsov, V. V.. Moscow (Russia), Ho Chi Minh, Buonmathuot (Vietnam): GEOS, pp. 109–149 (in Russian).Google Scholar
Suntsov, V. V. & Suntsova, N. I. (2006). Plague: Origin and Evolution of Epizootic System (Ecological, Geographical and Social Aspects). Moscow, Russia: KMK Scientific Press (in Russian).Google Scholar
Suntsov, V. V., Vi, Li Thi K. & Suntsova, N. I. (1992a). Some features of the flea (Insecta, Siphonaptera) fauna of small mammals in Vietnam. Zoologicheskyi Zhurnal, 71, 88–94 (in Russian).Google Scholar
Suntsov, V. V., Huong, L. T. & Suntsova, N. I. (1992b). Notes on fleas (Siphonaptera) in the plague foci on the Tay Nguyen Plateau (Vietnam). Parazitologiya, 26, 516–520 (in Russian).Google Scholar
Suter, P. R. (1964). Biologie von Echidnophaga gallinacea (Westw.) und Vergleich mit andern Verhaltenstypen bei Flöhen. Acta Tropica, 21, 193–238.Google Scholar
Sutherland, W. J. (1983). Aggregation and the ‘ideal free’ distribution. Journal of Animal Ecology, 52, 821–828.CrossRefGoogle Scholar
Sutherland, W. J. (1996). From Individual Behaviour to Population Ecology. Oxford, UK: Oxford University Press.Google Scholar
Sviridov, G. G. (1963). Application of radioactive isotopes for the study of some problems of flea ecology. II. The contact of animals and intensity of the exchange of ectoparasites in the population of Rhombomis opimus. Zoologicheskyi Zhurnal, 42, 546–550 (in Russian).Google Scholar
Syrvatcheva, N. G. (1964). Data on flea fauna of the Kabardino-Balkarian ASSR. Proceedings of the Armenian Anti-Plague Station, 3, 389–405 (in Russian).Google Scholar
Szidat, L. (1940). Beiträge zum Aubfau eines natürlichen Systems der Trematoden. I. Die Entwicklung von Echinocercaria choanophila U. Szidat zu Cathaemasia hians und die Ableitung der Fasciolidae von den Echinostomidae. Zeitschrift für Parasitenkunde, 11, 239–283.CrossRefGoogle Scholar
Tabor, S. P., Williams, D. F., Germano, D. J. & Thomas, R. E. (1993). Fleas (Siphonaptera) infesting giant kangaroo rats (Dipodomys ingens) on the Elkhorn and Carrizo plains, San Luis Obispo County, California. Journal of Medical Entomology, 30, 291–294.CrossRefGoogle Scholar
Takahashi, K., Tuno, N. & Kagaya, T. (2005). The relative importance of spatial aggregation and resource partitioning on the coexistence of mycophagous insects. Oikos, 109, 125–134.CrossRefGoogle Scholar
Talybov, A. N. (1974). Some data on the lifespan of fleas parasitic on the common vole in the Trans-Caucasus mountains. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 183–185 (in Russian).Google Scholar
Talybov, A. N. (1975). Life expectancy of Ctenophthalmus wladimiri Is.-Gurv., 1948 (Siphonaptera, Ctenophthalmidae) under laboratory conditions. Parazitologiya, 9, 354–358 (in Russian).Google ScholarPubMed
Talybov, A. N. (1976). Development of the pre-imaginal phases of flea Ctenophthalmus wladimiri Is.-Gurv., 1948. Parazitologiya, 10, 320–324 (in Russian).Google Scholar
Tanitovsky, V. A., Bidashko, F. G., Grazhdanov, A. K. & Dauletova, S. B. (2004). Species structure and number of fleas parasitizing small mammals in the middle part of the Ural River valley. Quarantinable and Zoonotic Infections in Kazakhstan, 9, 76–80 (in Russian).Google Scholar
Tarasevich, L. N., Tagiltsev, A. A. & Malkov, G. B. (1969). Results of virological examination of ixodid ticks and fleas in the southern Omsk Region. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 38, 705–707 (in Russian).Google Scholar
Tarshis, I. B. (1956). Feeding techniques for blood-sucking arthropods. Proceedings of the 10th International Congress of Entomology, 3, 767–784.Google Scholar
Taylor, J. & Purvis, A. (2003). Have mammals and their chewing lice diversified in parallel? In Tangled Trees: Phylogeny, Cospeciation, and Coevolution, ed. Page, R. D. M.. Chicago, IL: University of Chicago Press, pp. 240–261.Google Scholar
Taylor, L. H., Mackinnon, M. J. & Read, A. F. (1998). Virulence of mixed-clone and single-clone infections of the rodent malaria Plasmodium chabaudi. Evolution, 52, 583–591.CrossRefGoogle ScholarPubMed
Taylor, L. R. (1961). Aggregation, variance and the mean. Nature, 189, 732–735.CrossRefGoogle Scholar
Taylor, L. R. & Taylor, R. A. J. (1977). Aggregation, migration and population dynamics. Nature, 265, 415–421.CrossRefGoogle Scholar
Taylor, L. R. & Woiwod, I. P. (1980). Temporal stability as a density-dependent species characteristic. Journal of Animal Ecology, 49, 209–224.CrossRefGoogle Scholar
Taylor, L. R., Woiwod, I. P. & Perry, J. N. (1979). The negative binomial as a dynamic ecological model and density-dependence of k. Journal of Animal Ecology, 48, 289–304.CrossRefGoogle Scholar
Taylor, R. A. J., Lindquist, R. K. & Shipp, J. L. (1998). Variation and consistency in spatial distribution as measured by Taylor's power law. Environmental Entomology, 27, 191–201.CrossRefGoogle Scholar
Taylor, S. D., Cruz, Dittmar K., Porter, M. L. & Whiting, M. F. (2005). Characterization of the long-wavelength opsin from Mecoptera and Siphonaptera: does a flea see?Molecular Biology and Evolution, 22, 1165–1174.CrossRefGoogle ScholarPubMed
Tchernova, N. A. (1971). Reproduction of Xenopsylla skrjabini and their preferences for different elements of the great gerbil burrow in the Mangyshlak Peninsula. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 443–444 (in Russian).Google Scholar
Tchimanina, B. M. & Kozlovskaya, O. L. (1971a). Experimental study of the role of fleas Ctenophthalmus congeneroides Wagn. and Neopsylla bidentatiformis Wagn. in the circulation of the tick-borne encephalitis virus. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 235–236 (in Russian).Google Scholar
Tchimanina, B. M. & Kozlovskaya, O. L. (1971b). Experimental study of the circulation of the tick-borne encephalitis virus in the nests of the forest voles via fleas Frontopsylla elata botis, Jord., 1929. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 237–238 (in Russian).Google Scholar
Tchumakova, I. V. & Kozlov, M. P. (1983). Quantitative parameters of mortality and survival in fleas Nosopsyllus consimilis at different stages of the metamorphosis. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 280–281 (in Russian).Google Scholar
Tchumakova, I. V., Tokanev, F. I. & Kozlov, M. P. (1978). Dependence of the reproduction capacity of fleas (Aphaniptera) on the recurrence of mating. Parazitologiya, 12, 292–296 (in Russian).Google Scholar
Tchumakova, I. V., Kozlov, M. P. & Belokopytova, A. (1981). Estimation of the dependence of the reproduction of rodent fleas (Siphonaptera) on feeding by experimental breeding of fleas on different hosts. Entomological Review, 60, 562–569 (in Russian).Google Scholar
Tchumakova, I. V., Ermolova, N. V. & Shaposhnikova, L. I. (2002). Principles for prediction of population densities of fleas parasitic on rodents. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 72, 45–48 (in Russian).Google Scholar
Telfer, S., Bown, K. J., Sekules, R., et al. (2005). Disruption of a host–parasite system following the introduction of an exotic host species. Parasitology, 130, 661–668.CrossRefGoogle ScholarPubMed
Tella, J. L. (2002). The evolutionary transition to coloniality promotes higher blood parasitism in birds. Journal of Evolutionary Biology, 15, 32–41.CrossRefGoogle Scholar
Tella, J. L., Scheuerlein, A. & Ricklefs, R. E. (2002). Is cell-mediated immunity related to the evolution of life-history strategies in birds?Proceedings of the Royal Society of London B, 269, 1059–1066.CrossRefGoogle ScholarPubMed
Tellam, R. L., Smith, D., Kemp, D. H. & Willadsen, P. (1992). Vaccination against ticks. In Animal Parasite Control Utilizing Biotechnology, ed. Yong, W. K.. Boca Raton, FL: CRC Press, pp. 303–331.Google Scholar
Tenquist, J. D. & Charleston, W. A. G. (2001). A revision of the annotated checklist of ectoparasites of terrestrial mammals in New Zealand. Journal of the Royal Society of New Zealand, 31, 481–542.CrossRefGoogle Scholar
Teplinskaya, T. A., Labunetz, N. F. & Kuliev, M. T. (1983). Seasonal dynamics of age structure and reproduction of Xenopsylla conformis in the Caspian natural plague focus. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 277–278 (in Russian).Google Scholar
Terborgh, J. W. & Faaborg, J. (1980). Saturation of bird communities in the West Indies. American Naturalist, 116, 178–195.CrossRefGoogle Scholar
Hofstede, H. M. & Fenton, M. B. (2005). Relationships between roost preferences, ectoparasite density and grooming behaviour of neotropical bats. Journal of Zoology, 266, 333–340.CrossRefGoogle Scholar
Theodor, O. & Costa, M. (1967). A Survey of the Parasites of Wild Mammals and Birds in Israel. vol. 1, wEctoparasites. Jerusalem: Israel Academy of Science and Humanities.Google Scholar
Théron, A. & Combes, C. (1995). Asynchrony of infection timing, habitat preference, and sympatric speciation of schistosome parasites. Evolution, 49, 372–375.CrossRefGoogle ScholarPubMed
Thomas, C. D. & Hanski, I. (1997). Butterfly metapopulations. In Metapopulation Biology: Ecology, Genetics, and Evolution, ed. Hanski, I. & Gilpin, M. E.. San Diego, CA: Academic Press, pp. 359–386.Google Scholar
Thomas, K. & Shutler, D. (2001). Ectoparasites, nestling growth, parental feeding rates, and begging intensity of tree swallows. Canadian Journal of Zoology, 79, 346–353.CrossRefGoogle Scholar
Thomas, R. (1988). A review of flea collection records from Onychomys leucogaster with observations on the role of grasshopper mice in the epizootiology of wild rodent plague. Great Basin Naturalist, 48, 83–95.Google Scholar
Thomas, R. (1996). Fleas and the agents they transmit. In The Biology of Disease Vectors, ed. Beaty, B. J. & Marquardt, W. C.. Niwot, CO: University of Colorado Press, pp. 146–159.Google Scholar
Thompson, C. F. & Neill, A. J. (1991). House wrens do not prefer clean nestboxes. Animal Behaviour, 42, 1022–1024.CrossRefGoogle Scholar
Thompson, C. W., Hillgarth, N., Leu, M. & McClure, H. E. (1997). High parasite load in house finches (Carpodacus mexicanus) is correlated with reduced expression of a sexually selected trait. American Naturalist, 149, 270–294.CrossRefGoogle Scholar
Thompson, J. N. (1994). The Coevolutionary Process. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Thompson, J. N. (2005). The Geographic Mosaic of Coevolution. Chicago, IL: University of Chicago Press.Google Scholar
Tian, J. (1995). Niches of 27 flea species in the natural focus of plague in Jianchuan, Yunnan Province. Endemic Diseases Bulletin, 10, 27–32 (in Chinese).Google Scholar
Tiflov, V. E. (1959). Role of fleas in epizootology of tularemia. Proceedings of the Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, 2, 363–392 (in Russian).Google Scholar
Tiflov, V. E. (1964). Destiny of the bacterial cultures in the organism of a flea. Ectoparasites, 4, 181–198 (in Russian).Google Scholar
Tiflov, V. E. & Ioff, I. G. (1932). Observations on the biology of fleas. Herald of Microbiology and Epidemiology [Vestnik Mikrobiologii i Epidemiologii], 11, 95–117 (in Russian).Google Scholar
Tillyard, R. J. (1926). The Insects of Australia and New Zealand. Sydney Australia: Angus and Robertson.Google Scholar
Tillyard, R. J. (1935). The evolution of the scorpion-flies and their derivatives (order Mecoptera). Annals of the Entomological Society of America, 28, 1–45.CrossRefGoogle Scholar
Timi, J. T. & Poulin, R. (2003). Parasite community structure within and across host populations of a marine pelagic fish: how repeatable is it?International Journal for Parasitology, 33, 1353–1362.CrossRefGoogle Scholar
Tipton, V. J. & Machado-Allison, C. E. (1972). Fleas of Venezuela. Brigham Young University Scientific Bulletin, Biological Series, 17, 1–115.Google Scholar
Tipton, V. J. & Méndez, E. (1966). The fleas (Siphonaptera) of Panama. In Ectoparasites of Panama, ed. Wenzel, R. L. & Tipton, V. J.. Chicago, IL: Field Museum of Natural History, pp. 289–385.Google Scholar
Tipton, V. J. & Méndez, E. (1968). New species of fleas (Siphonaptera) from Cerro Potosi, Mexico, with notes on ecology and host–parasite relationships. Pacific Insects, 10, 177–214.Google Scholar
Toft, C. A. & Karter, A. J. (1990). Parasite–host coevolution. Trends in Ecology and Evolution, 5, 326–329.CrossRefGoogle ScholarPubMed
Tofts, R. & Silvertown, J. (2000). A phylogenetic approach to community assembly from a local species pool. Proceedings of the Royal Society of London B, 267, 363–369.CrossRefGoogle ScholarPubMed
Tokeshi, M. (1999). Species Coexistence: Ecological and Evolutionary Perspectives. Oxford, UK: Blackwell Science.Google Scholar
Tokmakova, E. G., Verzhutsky, D. B. & Bazanova, L. P. (2006). Formation of the proventriculus blockage, alimentary activity and mortality in fleas Amphipsylla primaris primaris infected with Yersinia pestis. Parazitologiya, 40, 215–224 (in Russian).Google Scholar
Trager, W. (1939). Acquired immunity to ticks. Journal of Parasitology, 25, 57–81.CrossRefGoogle Scholar
Tränkle, S. B. (1989). Wirtspecifizität und Wanderaktivität des Katzenflohes Ctenocephalides felis (Bouché). Unpublished M.Sc. thesis, Albert Ludwigs Universität, Freiburg im Beisgau, Germany.Google Scholar
Traub, R. (1972a). The Gunong Benom Expedition 1967. Ⅻ. Notes on zoogeography, convergent evolution and taxonomy of fleas (Siphonaptera), based on collection from Gunong Benom and elsewhere in South-East Asia. 2. Convergent evolution. Bulletin of the British Museum Natural History, Zoology, 23, 309–387.Google Scholar
Traub, R. (1972b). The relationship between the spines, combs and other skeletal features of fleas (Siphonaptera) and the vestiture, affinities and habits of their hosts. Journal of Medical Entomology, 9, 601.Google Scholar
Traub, R. (1980). The zoogeography and evolution of some fleas, lice and mammals. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 93–172.Google Scholar
Traub, R. (1985). Coevolution of fleas and mammals. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 295–437.Google Scholar
Traub, R., Wisseman, C. L. & Farhang-Azad, A. (1978). The ecology of murine typhus: a critical review. Tropical Diseases Bulletin, 75, 237–317.Google ScholarPubMed
Traub, R., Rothschild, M. & Haddow, J. F. (1983). The Ceratophyllidae: Key to the Genera and Host Relationships.Cambridge, UK: Cambridge University Press.Google Scholar
Tripet, F. & Richner, H. (1997a). Host responses to ectoparasites: food compensation by parent blue tits. Oikos, 78, 557–561.CrossRefGoogle Scholar
Tripet, F. & Richner, H. (1997b). The coevolutionary potential of a ‘generalist’ parasite, the hen flea Ceratophyllus gallinae. Parasitology, 115, 419–427.CrossRefGoogle Scholar
Tripet, F. & Richner, H. (1999a). Density-dependent processes in the population dynamics of a bird ectoparasite Ceratophyllus gallinae. Ecology, 80, 1267–1277.CrossRefGoogle Scholar
Tripet, F. & Richner, H. (1999b). Demography of the hen flea Ceratophyllus gallinae in blue tit Parus caeruleus nests. Journal of Insect Behavior, 12, 159–174.CrossRefGoogle Scholar
Tripet, F., Christe, P. & M⊘ller, A. P. (2002a). The importance of host spatial distribution for parasite specialization and speciaton: a comparative study of bird fleas (Siphonaptera: Ceratophyllidae). Journal of Animal Ecology, 71, 735–748.CrossRefGoogle Scholar
Tripet, F., Glaser, M. & Richner, H. (2002b). Behavioural responses to ectoparasites: time-budget adjustments and what matters to blue tits Parus caeruleus infested by fleas. Ibis, 144, 461–469.CrossRefGoogle Scholar
Tripet, F., Jacot, A. & Richner, H. (2002c). Larval competition affects the life histories and dispersal behaviour of an avian ectoparasite. Ecology, 83, 935–945.CrossRefGoogle Scholar
Trivers, R. L. & Willard, D. E. (1973). Natural selection of parental ability to vary sex ratio of offspring. Science, 179, 90–92.CrossRefGoogle ScholarPubMed
Trudeau, W. L., Fernandez-Caldas, E. & Fox, R. W. (1993). Allergenicity of the cat flea (Ctenocephalides felis felis). Clinical and Experimental Allergy, 23, 377–383.CrossRefGoogle Scholar
Trukhachev, N. N. (1971). Effect of host on the offspring of fleas Xenopsylla cheopis. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 423–425 (in Russian).Google Scholar
Tschirren, B., Fitze, P. S. & Richner, H. (2003). Sexual dimorphism in susceptibility to parasites and cell-mediated immunity in great tit nestlings. Journal of Animal Ecology, 72, 839–845.CrossRefGoogle Scholar
Tschirren, B., Richner, H. & Schwabl, H. (2004). Ectoparasite-modulated deposition of maternal androgens in great tit eggs. Proceedings of the Royal Society of London B, 271, 1371–1375.CrossRefGoogle ScholarPubMed
Tschirren, B., Saladin, V., Fitze, P. S., Schwabl, H. & Richner, H. (2005). Maternal yolk testosterone does not modulate parasite susceptibility or immune function in great tit nestlings. Journal of Animal Ecology, 74, 675–682.CrossRefGoogle Scholar
Tschirren, B., Bischoff, L. L., Saladin, V. & Richner, H. (2007a). Host condition and host immunity affect parasite fitness in a bird–ectoparasite system. Functional Ecology, 21, 372–378.CrossRefGoogle Scholar
Tschirren, B., Fitze, P. S. & Richner, H. (2007b). Maternal modulation of natal dispersal in a passerine bird: an adaptive strategy to cope with parasitism?American Naturalist, 169, 87–93.CrossRefGoogle Scholar
Uchikawa, K., Sato, A. & Kugimoto, M. (1967). A report on the flea fauna on the Oki Islands. Medical Entomology and Zoology, 18, 14–17 (in Japanese).CrossRefGoogle Scholar
Ulmanen, I. & Myllymäki, A. (1971). Species composition and numbers of fleas (Siphonaptera) in a local population of the field vole, Microtus agrestis (L.). Annales Zoologici Fennici, 8, 374–384.Google Scholar
Uvarov, B. P. (1931). Insects and climate. Transactions of the Royal Entomological Society of London, 79, 1–247.CrossRefGoogle Scholar
Vainikka, A., Jokinen, E. I., Kortet, R. & Taskinen, J. (2004). Gender- and season-dependent relationships between testosterone, oestradiol and immune functions in wild roach. Journal of Fish Biology, 64, 227–240.CrossRefGoogle Scholar
Valone, T. J. & Hoffman, C. D. (2002). Effects of regional pool size on local diversity in small-scale annual plant communities. Ecology Letters, 5, 477–480.CrossRefGoogle Scholar
Valtonen, E. T., Pulkkinen, K., Poulin, R. & Julkunen, M. (2001). The structure of parasite component communities in brackish water fishes of the northeastern Baltic Sea. Parasitology, 122, 471–481.CrossRefGoogle ScholarPubMed
Vandermeer, J. (1990). Indirect and diffuse interactions: complicated cycles in a population embedded in a large community. Journal of Theoretical Biology, 142, 429–442.CrossRefGoogle Scholar
Vansulin, S. A. (1961). Ecology of fleas parasitic on the great gerbils. In Proceedings of the Interdisciplinary Conference Dedicated to the 40th Anniversary of the Kazakh Soviet Socialist Republic, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 47–49 (in Russian).Google Scholar
Vansulin, S. A. (1965). On the ecology of fleas (Aphaniptera) of the great gerbil. Entomological Review, 44, 307–314 (in Russian).Google Scholar
Vansulin, S. A. & Volkova, L. A. (1962). Fur structure in Rhombomys opimus Licht. and its effect on the abundance of fleas parasitizing these rodents in different seasons. Zoologicheskyi Zhurnal, 41, 147–150 (in Russian).Google Scholar
Vuren, D. (1996). Ectoparasites, fitness, and social behaviour of yellow-bellied marmots. Ethology, 102, 686–694.CrossRefGoogle Scholar
Varma, M. G. R. & Page, R. J. C. (1966). The epidemiology of louping ill in Ayshire, Scotland: ectoparasites of small mammals. I. (Siphonaptera). Journal of Medical Entomology, 3, 331–335.CrossRefGoogle Scholar
Varma, M. G. R., Hellerhaupt, A., Trinder, P. K. E. & Langi, A. O. (1990). Immunization of guinea-pigs against Rhipicephalus appendiculatus adult ticks using homogenates from unfed immature ticks. Immunology, 71, 133–138.Google ScholarPubMed
Vashchenok, V. S. (1966a). Histological description of the oogenesis in fleas Echidnophaga oschanini Wagn. (Pulicidae, Aphaniptera). Zoologicheskyi Zhurnal, 45, 1821–1831 (in Russian).Google Scholar
Vashchenok, V. S. (1966b). Morphophysiological changes in the organism of fleas Echidnophaga oschanini Wagn. (Aphaniptera, Pulicidae) during feeding and reproduction. Entomological Review, 45, 715–727 (in Russian).Google Scholar
Vashchenok, V. S. (1967a). Gonotrophic relationships in fleas Ceratophyllus consimilis Wagn. (Aphaniptera, Ceratophyllidae). Parasitological Collection, 13, 222–235 (in Russian).Google Scholar
Vashchenok, V. S. (1967b). On the ecology of fleas Echidnophaga oshanini Wagn. (Pulicidae, Aphaniptera) in the Tuva ASSR. Parazitologiya, 1, 27–35 (in Russian).Google Scholar
Vashchenok, V. S. (1974). Activity of blood digestion in fleas. In Proceedings of the 7th Meeting of the All-Union Entomological Society, ed. Anonymous. Leningrad, USSR: Zoological Institute of the Academy of Sciences of the USSR, p. 209 (in Russian).Google Scholar
Vashchenok, V. S. (1979). Maintanence of the causative agent of the yersiniosis in fleas Xenopsylla cheopis. Parazitologiya, 13, 19–25 (in Russian).Google Scholar
Vashchenok, V. S. (1984). Fleas and agents of bacterial diseases of humans and animals. Parasitological Collection, 32, 79–123 (in Russian).Google Scholar
Vashchenok, V. S. (1988). Fleas: Vectors of Pathogens Causing Diseases in Humans and Animals. Leningrad, USSR: Nauka (in Russian).Google Scholar
Vashchenok, V. S. (1993). Factors regulating egg production in fleas Leptopsylla segnis (Leptopsyllidae: Siphonaptera). Parazitologiya, 27, 382–388 (in Russian).Google Scholar
Vashchenok, V. S. (1995). The dependence of the egg-laying activity in the fleas Leptopsylla segnis (Siphonaptera: Leptopsyllidae) on their abundance on a host. Parazitologiya, 29, 267–271.Google Scholar
Vashchenok, V. S. (2001). Age changes of fecundity in fleas Leptopsylla segnis (Siphonaptera: Leptopsyllidae). Parazitologiya, 35, 460–464 (in Russian).Google Scholar
Vashchenok, V. S. (2006). Species composition, host preferences and niche differentiation in fleas (Siphonaptera) parasitic on small mammals in the Ilmen–Volkhov Lowland. Parazitologiya, 40, 425–437 (in Russian).Google Scholar
Vashchenok, V. S. & Solina, L. T. (1969). On blood digestion in fleas Xenopsylla cheopis Roths. (Aphaniptera, Pulicidae). Parazitologiya, 3, 451–460 (in Russian).Google Scholar
Vashchenok, V. S. & Solina, L. T. (1972). Age-related changes in the fat tissue of female fleas Xenopsylla cheopis. Zoologicheskyi Zhurnal, 60, 79–85 (in Russian).Google Scholar
Vashchenok, V. S. & Tchirov, P. A. (1976). Histological study of fleas Ceratophyllus consimilis Wagn. infected with the causative agent of listeriosis (Listeria monocytogenes). Parazitologiya, 10, 61–66 (in Russian).Google Scholar
Vashchenok, V. S. & Tretiakov, K. A. (2003). Seasonal dynamic of flea (Siphonaptera) abundance on Clethrionomys glareolus in the northern part of the Novgorod Region. Parazitologiya, 37, 177–189 (in Russian).Google Scholar
Vashchenok, V. S. & Tretiakov, K. A. (2004). Seasonal dynamic of flea (Siphonaptera) abundance on the common shrew Sorex araneus in the northern part of the Novgorod Region. Parazitologiya, 38, 503–514 (in Russian).Google Scholar
Vashchenok, V. S. & Tretiakov, K. A. (2005). Seasonal dynamic of flea (Siphonaptera) abundance on the pygmy woodmouse Apodemus uralensis in the northern part of the Novgorod Region. Parazitologiya, 39, 270–277 (in Russian).Google Scholar
Vashchenok, V. S., SolinaL, T. L, T. & Zhirnov, A. E. (1976). Digestion of blood of different animals by fleas Xenopsylla cheopis. Parazitologiya, 10, 544–549 (in Russian).Google Scholar
Vashchenok, V. S., Bryukhanova, L. V. & Shchedrin, V. I. (1985). Characteristics of feeding and digestion in fleas. Parasitological Collection, 33, 134–148 (in Russian).Google Scholar
Vashchenok, V. S., Karandina, R. S. & Bryukhanova, L. V. (1988). Amount of blood consumed by different flea species in the experiments. Parazitologiya, 22, 312–320 (in Russian).Google Scholar
Vashchenok, V. S., Sheikin, A. O. & Serzhanov, O. S. (1992). Morphophysiological characteristics of fleas Xenopsylla gerbilli during the autumn–winter diapause. Parasitological Collection, 37, 5–15 (In Russian).Google Scholar
Vasiliev, G. I. (1961). Observations on flea breeding in the laboratory. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 2, 97–99 (in Russian).Google Scholar
Vasiliev, G. I. (1966). On ectoparasites and their hosts in relation to the plague epizootic in the Bajan-Khongor Aimak (Mongolian People's Republic). Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 277–281 (in Russian).Google Scholar
Vasiliev, G. I. (1971). Fleas of the long-tailed ground squirrel (species composition, ecology, epizootological importance for plague). Unpublished Ph.D. thesis, Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, Irkutsk, USSR (in Russian).
Vasiliev, G. I. & Zhovty, I. F. (1961). An attempt to investigate the rules of the distribution of micropopulations of fleas in a microhabitat. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 1, 88–91 (in Russian).Google Scholar
Vasiliev, G. I. & Zhovty, I. F. (1971). On the annual cycle of Oropsylla asiatica Wagn., 1929 (Siphonaptera) parasitic on Spermophilus ungulatus in the Siberian Cis-Baikalia. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 227–229 (in Russian).Google Scholar
Vaughan, J. A. & Coombs, M. E. (1979). Laboratory breeding of the European rabbit flea, Spilopsyllus cuniculi (Dale). Journal of Hygiene, 83, 521–530.CrossRefGoogle Scholar
Vaughan, J. A. & Mead-Briggs, A. R. (1970). Host-finding behaviour of the rabbit flea, Spilopsyllus cuniculi with special reference to the significance of urine as an attractant. Parasitology, 61, 397–409.CrossRefGoogle Scholar
Vaughan, J. A., Jerse, A. E. & Farhang-Azad, A. (1989). Rat leucocyte's response to the bites of rat fleas (Siphonaptera: Pulicidae). Journal of Medical Entomology, 26, 449–453.CrossRefGoogle Scholar
Vázquez, D. P. & Aizen, M. A. (2003). Null model analyses of specialization in plant–pollinator interactions. Ecology, 84, 2493–2501.CrossRefGoogle Scholar
Vázquez, D. P. & Stevens, R. D. (2004). The latitudinal gradient in niche breadth: concepts and evidence. American Naturalist, 164, E1–E19.CrossRefGoogle ScholarPubMed
Vázquez, D. P., Poulin, R., Krasnov, B. R. & Shenbrot, G. I. (2005). Species abundance patterns and the distribution of specialization in host–parasite interaction networks. Journal of Animal Ecology, 74, 946–955.CrossRefGoogle Scholar
Verts, B. J. (1961). Observations on the fleas (Siphonaptera) of some small mammals in northwestern Illinois. American Midland Naturalist, 66, 471–476.CrossRefGoogle Scholar
Verzhutsky, D. B., Zonov, G. B. & Popov, V. V. (1990). Epizootological importance of flea accumulation in the aggregations of female long-tailed ground squirrels in the Tuva plague focus. Parazitologiya, 24, 186–92 (in Russian).Google Scholar
Via, S. (2001). Sympatric speciation in animals: the ugly duckling grows up. Trends in Ecology and Evolution, 16, 381–390.CrossRefGoogle ScholarPubMed
Vidal-Martinez, V. M. & Poulin, R. (2003). Spatial and temporal repeatability in parasite community structure of tropical fish hosts. Parasitology, 127, 387–398.CrossRefGoogle ScholarPubMed
Viitala, J., Hakkarainen, H. & Ylönen, H. (1994). Different dispersal in Clethrionomys and Microtus. Annales Zoologici Fennici, 31, 411–415.Google Scholar
Violovich, N. A. (1969). Landscape and geographic distribution of fleas. In Biological Regionalization of the Novosibirsk Region, ed. Maximov, A. A.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 211–221 (in Russian).Google Scholar
Visser, M., Rehbein, S. & Wiedemann, C. (2001). Species of flea (Siphonaptera) infesting pets and hedgehogs in Germany. Journal of Veterinary Medicine B, 48, 197–202.CrossRefGoogle ScholarPubMed
Vobis, M., D'Haese, J., Mehlhorn, H., Heukelbach, J., Mencke, N. & Feldmeier, H. (2005). Molecular biological investigations of Brazilian Tunga sp. isolates from man, dogs, cats, pigs and rats. Parasitological Research, 96, 107–112.CrossRefGoogle ScholarPubMed
Volfertz, A. A. & Kolpakova, S. A. (1946). On epizootology of tularemia: role of fleas Ctenophthalmus orientalis Wagn. in epizootology of tularemia. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 16, 83–84 (in Russian).Google Scholar
Volis, S., Mendlinger, S., Olswig-Whittaker, L., Safriel, U. N. & Orlovsky, N. (1998). Phenotypic variation and stress resistance in core and peripheral populations of Hordeum spontaneum. Biodiversity and Conservation, 7, 799–813.CrossRefGoogle Scholar
Vysotskaya, S. O. (1967). Biocoenotical relationships between ectoparasites of rodents and other inhabitants of their nests. Parasitological Collection, 23, 19–60 (in Russian).Google Scholar
Waage, J. K. (1979). The evolution of insect/vertebrate associations. Biological Journal of the Linnean Society, 12, 187–224.CrossRefGoogle Scholar
Wade, S. E. & Georgi, J. R. (1988). Survival and reproduction of artificially fed cat fleas, Ctenocephalides felis Bouché (Siphonaptera: Pulicidae). Journal of Medical Entomology, 25, 186–190.CrossRefGoogle Scholar
Waeber, P. O. & Hemelrijk, C. K. (2003). Female dominance and social structure in Alaotran gentle lemurs. Behaviour, 140, 1235–1246.CrossRefGoogle Scholar
Wagner, J. (1929). About new species of Palaearctic fleas (Aphaniptera). II. AnnualReports of the Zoological Museum of the Academy of Sciences of the USSR, 30, 531–547 (in Russian).Google Scholar
Wakelin, D. (1996). Immunity to Parasites: How Parasitic Infections Are Controlled, 2nd edn. Cambridge, UK: Cambridge University Press.Google Scholar
Walker, M., Steiner, S., Brinkhof, M. W. G. & Richner, H. (2003). Induced responses of nestling great tits reduce hen flea reproduction. Oikos, 102, 67–74.CrossRefGoogle Scholar
Wall, R. & Shearer, D. (2001). Veterinary Ectoparasites: Biology, Pathology and Control, 2nd edn. Oxford: Blackwell Science.CrossRefGoogle Scholar
Walshe, B. M. (1948). The oxygen requirements and thermal resistance of chironomid larvae from flowing and from still waters. Journal of Experimental Biology, 25, 35–44.Google Scholar
Walton, D. W. & Hong, H.-K. (1976). Fleas of small mammals from the endemic haemorrhagic fever zones of Kyonggi and Kangwon Provinces of the Repubic of Korea. Korean Journal of Parasitology, 14, 17–24.CrossRefGoogle ScholarPubMed
Walton, D. W. & Tun, U. M. (1978). Fleas of small mammals from Rangoon, Burma. Southeast Asian Journal of Tropical Medicine and Public Health, 9, 369–377.Google ScholarPubMed
Wang, G.-L., Xi, N., Adily, S., et al. (2004a). Studies on some characteristics of bioecology and morphology of Vermipsylla alakurt. Endemic Diseases Bulletin, 19, 25–27 (in Chinese).Google Scholar
Wang, G.-L., Xi, N., Dang, X.-S., et al. (2004b). Pathogen identification of vermipsyllosis of domestic animal in Bazhou, Xinjiang. Chinese Journal of Veterinary Parasitology, 12, 2–3 (in Chinese).Google Scholar
Warburg, A., Saraiva, E., Lanzaro, G. C., Titus, R. G. & Neva, F. (1994). Saliva of Lutzomyia longipalpis sibling species differs in its composition and capacity to enhance leishmaniasis. Philosophical Transactions of the Royal Society of London B, 345, 223–230.CrossRefGoogle ScholarPubMed
Ward, S. A. (1992). Assessing functional explanations of host specificity. American Naturalist, 139, 883–891.CrossRefGoogle Scholar
Warwick, R. M. & Clarke, K. R. (2001). Practical measures of marine biodiversity based on relatedness of species. Oceanography and Marine Biology, 39, 207–231.Google Scholar
Watkins, R. A., Moshier, S. E. & Pinter, A. J. (2006). The flea Megabothris abantis: an invertebrate host of Hepatozoon sp. and a likely definitive host in Hepatozoon infections of the montane vole, Microtus montanus. Journal of Wildlife Diseases, 42, 386–390.CrossRefGoogle Scholar
Watt, C., Dobson, A. P. & Grenfell, B. T. (1995). Glossary. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 510–521.CrossRefGoogle Scholar
Watts, M. M., Pascoe, D. & Carroll, K. (2002). Population responses of the freshwater amphipod Gammarus pulex (L.) to an environmental estrogen, 17 alpha-ethinylestradiol. Environmental Toxicology and Chemistry, 21, 445–450.CrossRefGoogle Scholar
Webb, C. T., Brooks, C. P., Gage, K. L. & Antolin, M. F. (2006). Classic flea-borne transmission does not drive plague epizootics in prairie dogs. Proceedings of the National Academy of Sciences of the USA, 103, 6236–6241.CrossRefGoogle Scholar
Webb, D. R., Porter, W. P. & Mcclure, P. A. (1990). Development of insulation in juvenile rodents: functional compromise in insulation. Functional Ecology, 4, 251–256.CrossRefGoogle Scholar
Webber, L. A. & Edman, J. D. (1972). Anti-mosquito behaviour of ciconiiform birds. Animal Behaviour, 20, 228–232.CrossRefGoogle Scholar
Wedekind, C. (1992). Detailed information about parasites revealed by sexual ornamentation. Proceedings of the Royal Society of London B, 247, 169–174.CrossRefGoogle Scholar
Wegner, Z. (1970). Lice (Anoplura) of small mammals caught in Dobrogea (Roumania). SocietƷʝii de Þiinʝe Biologia din Republica SocialistƷ România, Zoologia, 20, 305–314.Google Scholar
Wenk, P. (1953). Der Kopf von Ctenocephalus canis (Curt.) (Aphaniptera). Zoologische Jahrbücher, Abteilung für Anatomie und Ontogenie der Tiere, 73, 103–164.Google Scholar
Wenzel, R. L. & Tipton, V. J. (1966). Some relationships between mammal hosts and their ectoparasites. In Ectoparasites of Panama, ed. Wenzel, R. L. & Tipton, V. J.. Chicago, IL: Field Museum of Natural History, pp. 677–723.Google Scholar
Wesołowski, T. & Stańska, M. (2001). High ectoparasite loads in hole-nesting birds: a nestbox bias?Journal of Avian Biology, 32, 281–285.CrossRefGoogle Scholar
Wessels, W. (1998). Gerbillidae from the Miocene and Pliocene of Europe. Mitteilungen bayerische Staatssammlung für Paläontologie und historische Geologie, 38, 187–207.Google Scholar
Whitehead, M. D., Burton, H. R., Bell, P. J., Arnould, J. P. Y. & Rounsevell, D. E. (1991). A further contribution on the biology of the Antarctic flea, Glaciopsyllus antarcticus (Siphonaptera: Ceratophyllidae). Polar Biology, 11, 379–383.CrossRefGoogle Scholar
Whiteman, N. K. & Parker, P. G. (2004). Body condition and parasite load predict territory ownership in the Galapagos hawk. Condor, 106, 915–921.CrossRefGoogle Scholar
Whiteman, N. K. & Parker, P. G. (2005). Using parasites to infer host population history: a new rationale for parasite conservation. Animal Conservation, 8, 175–181.CrossRefGoogle Scholar
Whiting, M. F., Carpenter, J. C., Wheeler, Q. D. & Wheeler, W. C. (1997). The Strepsiptera problem: phylogeny of the holometabolous insect orders inferred from 18S and 28S ribosomal DNA sequences and morphology. Systematic Biology, 46, 1–68.Google ScholarPubMed
Whiting, M. F. (2002a). Phylogeny of the holometabolous insect orders: molecular evidence. Zoologica Scripta, 31, 3–15.CrossRefGoogle Scholar
Whiting, M. F. (2002b). Mecoptera is paraphyletic: multiple genes and phylogeny of Mecoptera and Siphonaptera. Zoologica Scripta, 31, 93–104.CrossRefGoogle Scholar
Whiting, M. F., Whiting, A. S. & Hastriter, M. W. (2003). A comprehensive phylogeny of Mecoptera and Siphonaptera. Entomologische Abhandlungen, 61, 169.Google Scholar
Widmann, O. (1922). Extracts from the diary of Otto Widmann. Transactions of the Academy of Science of St Louis, 24, 1–77.Google Scholar
Wikel, S. K. (1984). Immunomodulation of host responses to ectoparasite infestation: an overview. Veterinary Parasitology, 14, 321–339.CrossRefGoogle ScholarPubMed
Wikel, S. K. (ed.) (1996). The Immunology of Host–Ectoparasitic Arthropod Relationships. Wallingford, UK: CAB International.Google Scholar
Wikel, S. K. & Alarcon-Chaidez, F. J. (2001). Progress toward molecular characterization of ectoparasite modulation of host immunity. Veterinary Parasitology, 101, 275–287.CrossRefGoogle ScholarPubMed
Willadsen, P. (1980). Immunity to ticks. Advances in Parasitology, 18, 293–313.CrossRefGoogle ScholarPubMed
Willadsen, P. (1987). Immunological approaches to the control of ticks. International Journal for Parasitology, 17, 671–677.CrossRefGoogle ScholarPubMed
Willadsen, P. (2001). The molecular revolution in the development of vaccines against ectoparasites. Veterinary Parasitology, 101, 353–367.CrossRefGoogle ScholarPubMed
Willadsen, P. (2006). Vaccination against ectoparasites. Parasitology, 133, S9–S25.CrossRefGoogle ScholarPubMed
Willadsen, P., Bird, P. E., Cobon, G. & Hungerford, J. (1995). Commercialization of a recombinant vaccine against Boophilus microplus. Parasitology, 110, 543–550.CrossRefGoogle Scholar
Williams, B. (1991). Adaptations to endoparasitism in the larval integument and respiratory system of the flea Uropsylla tasmanica Rothschild (Siphonaptera, Pygiopsyllidae). Australian Journal of Zoology, 39, 77–90.CrossRefGoogle Scholar
Williams, B. (1993). Reproductive success of cat fleas, Ctenocephalides felis, on calves as unusual hosts. Medical and Veterinary Entomology, 7, 94–98.CrossRefGoogle ScholarPubMed
Williams, R. T. (1971). Observations on the behaviour of the European rabbit flea, Spilopsyllus cuniculi (Dale), on a natural population of wild rabbits, Oryctolagus cuniculus (L.), in Australia. Australian Journal of Zoology, 19, 41–51.CrossRefGoogle Scholar
Williams, R. T. & Paper, I. (1971). Observations on the dispersal of the European rabbit flea, Spilopsyllus cuniculi (Dale), through a natural population of wild rabbits, Oryctolagus cuniculus (L.). Australian Journal of Zoology, 19, 129–140.CrossRefGoogle Scholar
Willmann, R. (1981a). Das Exoskelett der männlichen Genitalien der Mecoptera (Insecta). I. Morphologie. Zeitschrift für zoologische Systematik und Evolutionsforschung, 19, 96–150.CrossRefGoogle Scholar
Willmann, R. (1981b). Das Exoskelett der männlichen Genitalien der Mecoptera (Insecta). II. Die phylogenetischen Beziehungen der Schnabelfliegen-Familien. Zeitschrift für zoologische Systematik und Evolutionsforschung, 19, 153–174.CrossRefGoogle Scholar
Wilson, D. E. & Reeder, D. M. (eds.) (2005). Mammal Species of the World: A Taxonomic and Geographic Reference, 3rd edn. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Wilson, K., Bj⊘rnstad, O. N., Dobson, A. P., et al. (2001). Heterogeneities in macroparasite infections: patterns and processes. In The Ecology of Wildlife Diseases, ed. Hudson, P. J., Rizzoli, A., Grenfell, B. T., Heesterbeek, H. & Dobson, A. P.. Oxford, UK: Oxford University Press, pp. 6–44.Google Scholar
Wilson, N. A. (1961). The ectoparasites (Ixodides, Anoplura and Siphonaptera) of Indiana Mammals. Unpublished Ph.D. thesis, Purdue University, West Lafayette, IN.
Wilson, N. A. & Durden, L. A. (2003). Ectoparasites of terrestrial vertebrates inhabiting the Georgia Barrier Islands, USA: an inventory and preliminary biogeographical analysis. Journal of Biogeography, 30, 1207–1220.CrossRefGoogle Scholar
Wilson, N. A., Telford, S. R. & Forrester, D. J. (1991). Ectoparasites of a population of urban gray squirrels in northern Florida. Journal of Medical Entomology, 28, 461–464.CrossRefGoogle ScholarPubMed
Windsor, D. A. (1990). Heavenly hosts. Nature, 348, 104.CrossRefGoogle Scholar
Windsor, D. A. (1995). Equal rights for parasites. Conservation Biology, 9, 1–2.CrossRefGoogle Scholar
Winkel, W. (1975a). Vergleichend-brutbiologische Untersuchungen an fünf Meisenarten (Parus spp.) in einem niedersächsischen Aufforstungsgebeit mit japanischer Lärche Larix leptolepis. Die Vogelwelt, 96, 41–63.Google Scholar
Winkel, W. (1975b). Vergleichend-brutbiologische Untersuchungen an fünf Meisenarten (Parus spp.) in einem niedersächsischen Aufforstungsgebeit mit japanischer Lärche Larix leptolepis. Die Vogelwelt, 96, 104–114.Google Scholar
Withers, P. C. (1992). Comparative Animal Physiology. Fort Worth, TX: Saunders College Publishing.Google Scholar
Witt, L. H., Linardi, P. M., Meckes, O., et al. (2004). Blood-feeding of Tunga penetrans males. Medical and Veterinary Entomology, 18, 439–441.CrossRefGoogle ScholarPubMed
Worthen, W. B. (1996). Community composition and nested-subsets analyses: basic descriptors for community ecology. Oikos, 76, 417–426.CrossRefGoogle Scholar
Worthen, W. B. & Rohde, K. (1996). Nested subsets analyses of colonization-dominated communities: metazoan ectoparasites of marine fishes. Oikos, 75, 471–478.CrossRefGoogle Scholar
Wright, D. H., Patterson, B. D., Mikkelson, G. M., Cutler, A. & Atmar, W. (1998). A comparative analysis of nested subset patterns of species composition. Oecologia, 113, 1–20.CrossRefGoogle Scholar
Xun, H. & Qi, Y.-M. (2004). Histochemistry of three enzymes in newly emerged and engorged adults of rat fleas Monopsyllus anisus (Rothschild) and Leptopsylla segnis (Schönherr). Acta Entomologica Sinica, 47, 444–448 (in Chinese).Google Scholar
Xun, H. & Qi, Y.-M. (2005). Histochemistry of fat and nonspecific esterase in newly emerged and sucked adults of rat fleas Monopsyllus anisus (Rothschild) and Leptopsylla segnis (Schoenherr). Acta Entomologica Sinica, 48, 829–832 (in Chinese).Google Scholar
Yadav, A., Khajuria, J. K. & Devi, J. (2006). Cat flea infestation in goats. Indian Veterinary Journal, 83, 439–440.Google Scholar
Yakunin, B. M. & Kunitskaya, N. T. (1980). On the effect of high relative humidity on reproduction and longevity of fleas Xenopsylla skrjabini. In Problems of Natural Focality of Plague: Proceedings of the 4th Soviet–Mongol Conference of Specialists from the Anti-Plague Establishments, vol. 1, ed. Golubinsky, E. P.. Irkutsk, USSR: Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, pp. 88–89 (in Russian).Google Scholar
Yakunin, B. M., Zolotova, S. I., Serzhanov, O. S., et al. (1971). On density dynamics and reproduction of Pulex irritans. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 450–452 (in Russian).Google Scholar
Yakunin, B. M., Tchernova, N. A. & Kunitskaya, N. T. (1979). Annual number of generations of fleas Xenopsylla skrjabini in the Mangyshlak Peninsula (Aphaniptera). Parazitologiya, 13, 510–515 (in Russian).Google Scholar
Yamauchi, T. (2005). Human dermatitis caused by the house-martin flea, Ceratophyllus farreni chaoi (Siphonaptera: Ceratophyllidae) in Shimane Prefecture, Japan. Medical Entomology and Zoology, 56, 49–52 (in Japanese).CrossRefGoogle Scholar
Yensen, E., Baird, C. R. & Sherman, P. W. (1996). Larger ectoparasites of the Idaho ground squirrel (Spermophilus brunneus). Great Basin Naturalist, 56, 237–246.CrossRefGoogle Scholar
Yeruham, I. & Koren, O. (2003). Severe infestation of a she-ass with the cat flea Ctenocephalides felis felis (Bouché, 1835). Veterinary Parasitology, 115, 365–367.CrossRefGoogle Scholar
Yeruham, I., Rosen, S. & Hadani, A. (1989). Mortality in calves, lambs and kids caused by severe infestation with the cat flea Ctenocephalides felis felis (Bouché, 1835) in Israel. Veterinary Parasitology, 30, 351–356.CrossRefGoogle Scholar
Ying, B., Kosoy, M. Y, Maupin, G. O., Tsuchiya, K. R. & Gage, K. L. (2002). Genetic and ecological characteristics of Bartonella communities in rodents in southern China. American Journal of Tropical Medicine and Hygiene, 66, 622–627.CrossRefGoogle Scholar
Yinon, U., Shulov, A. & Margalit, Y. (1967). The hygroreaction of the larvae of the Oriental rat flea, Xenopsylla cheopis Rothsch. (Siphonaptera: Pulicidae). Parasitology, 57, 315–319.CrossRefGoogle Scholar
Yudin, B. S., Krivosheev, V. G. & Belyaev, V. G. (1976). Small Mammals of the Northern Far East. Novosibirsk, USSR: Nauka, Siberian Branch (in Russian).Google Scholar
Yue, B.-S., Zou, F.-D., Sun, Q.-Z. & Li, J. (2002). Mating behavior of the cat flea, Ctenocephalides felis Bouché (Siphonaptera: Pulicidae) and male response to female extract on an artificial feeding system. Acta Entomologica Sinica, 9, 29–34 (in Chinese).Google Scholar
Yurgenson, I. A. & Maksimov, V. N. (1981). Effect of air temperature and relative humidity on the pre-imaginal development of fleas Ctenophthalmus teres (Siphonaptera). Parazitologiya, 15, 38–46 (in Russian).Google Scholar
Yushchenko, G. V. (1965). On the problem of the pseudotuberculosis research in the USSR. Proceedings of the Scientific Institute of Vaccines and Sera and Tomsk Medicine Institute, 16, 167–173 (in Russian).Google Scholar
Zagniborodova, E. N. (1960). Fauna and ecology of fleas in western Turmenistan. In Problems of Natural Foci and Epizootology of Plague in Turkmenistan, ed. Fenyuk, B. K.. Saratov, USSR: Turkmenian Anti-Plague Station and All-Union Scientific Anti-Plague Institute ‘Microb’, pp. 320–334 (in Russian).Google Scholar
Zagniborodova, E. N. (1965). Epizootological importance of migrating fleas of the great gerbil in Turkmenistan. Proceedings of the Academy of Sciences of the Turkmenian SSR, Biology, 5, 65–70 (in Russian).Google Scholar
Zagniborodova, E. N. (1968). Long-term study of the ecology of fleas of the great gerbil in the southern part of the central Kara-Kum Desert. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 78–86 (in Russian).Google Scholar
Zahavi, A. (1977). The cost of honesty (further remarks on the handicap principle). Journal of Theoretical Biology, 67, 603–605.CrossRefGoogle Scholar
Zahn, A. & Rupp, D. (2004). Ectoparasite load in European vespertilionid bats. Journal of Zoology, 262, 383–391.CrossRefGoogle Scholar
Zakson-Aiken, M., Gregory, L. M. & Shoop, W. L. (1996). Reproductive strategies of the cat flea (Siphonaptera: Pulicidae): parthenogenesis and autogeny?Journal of Medical Entomology, 33, 395–397.CrossRefGoogle ScholarPubMed
Zatsarinina, G. V. (1972). Salvic alakurt (Dorcadia dorcadia Roth., Aphaniptera, Vermipsillidae): the pest of the deer breeding. Proceedings of the Entomological Sector [Trudy Entomologisheskogo Sektora], 3, 1–108 (in Russian).Google Scholar
Zavala-Velazquez, J. E., Ruiz-Sosa, J. A., Sanchez-Elias, R. A., Becerra-Carmona, G. & Walker, D. H. (2000). Rickettsia felis rickettsiosis in Yucatan. Lancet, 356, 1079–1080.CrossRefGoogle ScholarPubMed
Zeigler, R. & Ibrahim, M. M. (2001). Formation of lipid reserves in fat body and eggs of the yellow fever mosquito, Aedes aegypti. Journal of Insect Physiology, 47, 623–627.CrossRefGoogle Scholar
Zenuto, R. R., Antinuchi, C. D. & Busch, C. (2002). Bioenergetics of reproduction and pup development in a subterranean rodent (Ctenomys talarum). Physiological and Biochemical Zoology, 75, 469–478.CrossRefGoogle Scholar
Zhang, T., Yu, X.-M. & Zhang, S.-Y. (2005). Investigation on the community composition of the small mammals and the parastic fleas in Xingning, Guangdong. Chinese Journal of Vector Biology and Control, 16, 446–447 (in Chinese).Google Scholar
Zhang, Y., Jin, S., Quan, G., et al. (1997). Distribution of Mammalian Species in China. Beijing: China Forestry Publishing House.Google Scholar
Zhang, Y.-Z., Gong, Z.-D., Feng, X.-G., et al. (2002). Study on the relationship between fleas and hosts in Mt. Baicaoling, Yunnan Province, China. Endemic Diseases Bulletin, 17, 22–23 (in Chinese).Google Scholar
Zhao, L., Jin, H.-L., She, R.-P., et al. (2006). A rodent model for allergic dermatitis induced by flea antigens. Veterinary Immunology and Immunopathology, 114, 285–296.CrossRefGoogle ScholarPubMed
Zhovty, I. F. (1963). Some contradictory questions of ecology of the rodent fleas in association with their epidemiological importance. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 6, 96–104 (in Russian).Google Scholar
Zhovty, I. F. (1967). Effect of rodents' life history on ecological conditions of their fleas. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 27, 195–210 (in Russian).Google Scholar
Zhovty, I. F. (1970). Essays on the ecology of fleas parasitic on rodents in Siberia and Far East. II. Fleas of marmots. In Vectors of Particularly Dangerous Diseases and Their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 253–283 (in Russian).Google Scholar
Zhovty, I. F. & Kopylova, O. A. (1957). Fleas of the Daurian pika in the period of the increase of its density. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 15, 293–298 (in Russian).Google Scholar
Zhovty, I. F. & Leonov, Y. A. (1958). Abundance of fleas on the Norway rats in the settlements of the southern part of the Primorie Region (Far East) and some patterns of its variation. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 75–89 (in Russian).Google Scholar
Zhovty, I. F. & Peshkov, B. I. (1958). Observations on the overwintering of fleas parasitic on the grey marmots in the Trans-Baikalia. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 27–32 (in Russian).Google Scholar
Zhovty, I. F. & Vasiliev, G. I. (1962a). Temperature conditions of rodents' fur as an environment for fleas. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 4, 152–156 (in Russian).Google Scholar
Zhovty, I. F. & Vasiliev, G. I. (1962b). On the self-cleaning from fleas in rodents. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 4, 156–160 (in Russian).Google Scholar
Zhovty, I. F., Netchaeva, L. K., Koshkin, S. M., et al. (1983). Patterns of seasonal density fluctuations in populations of the rat fleas in the Primorie Region (Far East) and a search for the criteria for their prognosis. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 234–235 (in Russian).Google Scholar
Zolotova, S. I. (1968). Comparative ecological descriptions of fleas of the great gerbil, Xenopsylla gerbilli minax Jord., 1926 and Ctenophthalmus dolichus Ioff, 1953. Unpublished Ph.D. thesis, The Middle Asian Scientific Anti-Plague Institute, Alma-Ata, USSR (in Russian).Google Scholar
Zolotova, S. I. & Afanasieva, O. V. (1969). Materials on the ecology of fleas of the great gerbil. IV. Duration of pre-imaginal development in Ctenophthalmus dolichus. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 66–68 (in Russian).Google Scholar
Zolotova, S. I. & Iskhanova, Z. A. (1979). Relationships between fleas Xenopsylla gerbilli minax and X. skrjabini in the area of overlapping of their geographic ranges. In Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 99–101 (in Russian).Google Scholar
Zolotova, S. I. & Varshavskaya, P. N. (1974). Age composition of imago in the population of Xenopsylla skrjabini in the northern Cis-Aral Region. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 316–318 (in Russian).Google Scholar
Zolotova, S. I., Pavlova, A. E. & Yakunin, B. M. (1971). Longevity of Pullex irritans after feeding on a non-specific host. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 377–378 (in Russian).Google Scholar
Zolotova, S. I., Bibikova, V. I. & Murzakhmetova, K. (1979). On the fecundity of fleas Xenopsylla gerbilli minax parasitic on the great gerbil (Aphaniptera). Parazitologiya, 13, 497–502.Google Scholar
Zuk, M. & McKean, K. A. (1996). Sex differences in parasite infections: patterns and processes. International Journal for Parasitology, 26, 1009–1024.CrossRefGoogle ScholarPubMed
Abouheif, E. & Fairbairn, D. J. (1997). A comparative analysis of allometry for sexual size dimorphism: assessing Rensch's rule. American Naturalist, 149, 540–562.CrossRefGoogle Scholar
Abrahams, M. (2004). Mark but this flea. Guardian, 30 November 2004.Google Scholar
Abramsky, Z., Bowers, M. A. & Rosenzweig, M. L. (1986). Detecting interspecific competition in the field: testing the regression method. Oikos, 47, 199–204.CrossRefGoogle Scholar
Abramsky, Z., Rosenzweig, M. L., Pinshow, B., et al. (1990). Habitat selection: an experimental field-test with two gerbil species. Ecology, 71, 2358–2369.CrossRefGoogle Scholar
Abu-Madi, M. A., Behnke, J. M., Mikhail, M., Lewis, J. W. & Al-Kaabi, M. L. (2005). Parasite populations in the brown rat Rattus norvegicus from Doha, Qatar between years: the effect of host age, sex and density. Journal of Helminthology, 79, 105–111.CrossRefGoogle ScholarPubMed
Acosta, R. (2005). Relación huésped-parásito en pulgas (Insecta: Siphonaptera) y roedores (Mammalia: Rodentia) del estado de Querétaro, México. Folia Entomologica Mexicana, 44, 37–47.Google Scholar
Acosta, R. & Morrone, J. J. (2005). A new species of Hystrichopsylla Taschenberg (Siphonaptera: Hystrichopsyllidae) from the Mexican transition zone. Zootaxa, 1027, 213–238.CrossRefGoogle Scholar
Adjemian, J. C. Z., Girvetz, E., Beckett, L. & Foley, J. E. (2006). Analysis of genetic algorithm for rule-set production (GARP) modeling approach for predicting distributions of fleas implicated as vectors of plague, Yersinia pestis, in California. Journal of Medical Entomology, 43, 93–103.Google ScholarPubMed
Ageyev, V. S. & Sludsky, A. A. (1985). Materials on fleas of small mammals in the eastern Pamir and perspective of the epizootologic monitoring of this region. In Important Questions of Epidemiological Monitoring in the Natural Foci of Plague: Natural Focality of Plague in High Mountains, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 10–12 (in Russian).Google Scholar
Ageyev, V. S., Arzhannikova, A. S., Tlegenov, T. T., Samarin, E. G. & Serzhanov, O. S. (1983). Interspecific interactions among five species of fleas co-occurring on the midday jirds. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 209–210 (in Russian).Google Scholar
Ageyev, V. S., Serzhanov, O. S., Arzhannikova, A. S. & Tlegenov, T. T. (1984). Interspecific interactions among imagoes of five species of fleas (Siphonaptera) parasitic on gerbils. Parazitologiya, 18, 185–190 (in Russian).Google Scholar
Akin, D. M. (1984). Relationship between feeding and reproduction in the cat flea, Ctenocephalides felis (Bouché). Unpublished M.Sc. thesis, University of Florida, Gainesville, FL.
Alania, I. I., Rostigaev, B. A., Shiranovich, P. I. & Dzneladze, M. T. (1964). Data on the flea fauna of Adzharia. Proceedings of the Armenian Anti-Plague Station, 3, 407–435 (in Russian).Google Scholar
Alekseev, A. N. (1961). On the bionomics of fleas Ceratophyllus (Nosopsyllus) consimilis Wagn., 1898 (Ceratophyllidae, Aphaniptera). Zoologicheskyi Zhurnal, 40, 1840–1847 (in Russian).Google Scholar
Alekseev, A. N., Grebenyuk, R. V., Tchirov, P. A. & Kadysheva, A. M. (1971). On the relationships between the listeriosis pathogen (Listeria monocytogenes) and fleas. Parazitologiya, 5, 113–118 (in Russian).Google Scholar
Alexander, J. O. (1986). The physiology of itch. Parasitology Today, 2, 345–351.CrossRefGoogle ScholarPubMed
Allan, R. M. (1956). A study of the populations of the rabbit flea Spilopsyllus cuniculi (Dale) on the wild rabbit, Oryctolagus cuniculus, in north-east Scotland. Proceedings of the Royal Entomological Society of London A, 31, 145–152.CrossRefGoogle Scholar
Allander, K. (1998). The effects of an ectoparasite on reproductive success in the great tit: a 3-year experimental study. Canadian Journal of Zoology, 76, 19–25.CrossRefGoogle Scholar
Allred, D. M. (1968). Fleas of the National Reactor Testing Station. Great Basin Naturalist, 28, 73–87.Google Scholar
Allsopp, P. G. (1997). Probability of describing an Australian scarab beetle: influence of body size and distribution. Journal of Biogeography, 24, 717–724.CrossRefGoogle Scholar
Almeida-Neto, M., Guimarães, P. R. & Lewinsohn, T. M. (2007). On nestedness analyses: rethinking matrix temperature and anti-nestedness. Oikos, 116, 716–722.CrossRefGoogle Scholar
Altizer, S., Nunn, C. L., Thrall, P. H., et al. (2004). Social organization and parasite risk in mammals: integrating theory and empirical studies. Annual Review of Ecology and Systematics, 34, 517–547.CrossRefGoogle Scholar
Amin, O. M. (1974). Comb variation in the rabbit flea Cediopsylla simplex (Baker). Journal of Medical Entomology, 11, 227–230.CrossRefGoogle Scholar
Amin, O. M. (1976). Host associations and seasonal occurrence of fleas from southeastern Wisconsin mammals, with observations on morphologic variations. Journal of Medical Entomology, 13, 179–192.CrossRefGoogle ScholarPubMed
Amin, O. M. (1982). The significance of pronotal comb patterns in flea–host lodging adaptations. Wiadomosci Parazytologiczne, 28, 93–94.Google Scholar
Amin, O. M. & Sewell, R. G. (1977). Comb variations in the squirrel and chipmunk fleas, Orchopeas h. howardii (Baker) and Megabothris acerbus (Jordan) (Siphonaptera), with notes on the significance of pronotal comb patterns. American Midland Naturalist, 98, 207–212.CrossRefGoogle Scholar
Amin, O. M. & Wagner, M. E. (1983). Further notes on the function of pronotal combs in fleas (Siphonaptera). Annals of the Entomological Society of America, 76, 232–234.CrossRefGoogle Scholar
Amin, O. M., Wells, T. R. & Gately, H. L. (1974). Comb variation in the cat flea, Ctenocephalides f. felis (Bouché). Annals of the Entomological Society of America, 67, 831–834.CrossRefGoogle Scholar
Amin, O. M., Liu, J., Li, S.-J., Zhang, Y.-M. & Sun, L.-Z. (1993). Development and longevity of Nosopsyllus laeviceps kuzenkovi (Siphonaptera) from Inner Mongolia under laboratory conditions. Journal of Parasitology, 79, 193–197.CrossRefGoogle ScholarPubMed
Amrine, J. W. & Lewis, R. E. (1978). The topography of the exoskeleton of Cediopsylla simplex (Baker 1895) (Siphonaptera: Pulicidae). I. The head and its appendages. Journal of Parasitology, 64, 343–358.CrossRefGoogle ScholarPubMed
Anderson, P. C. & Kok, O. B. (2003). Ectoparasites of springhares in the Northern Cape Province, South Africa. South African Journal of Wildlife Research, 33, 23–32.Google Scholar
Anderson, R. M. & Gordon, D. M. (1982). Processes influencing the distribution of parasite numbers within host populations with special emphasis on parasite-induced host mortality. Parasitology, 85, 373–398.CrossRefGoogle Scholar
Anderson, R. M. & May, R. M. (1978). Regulation and stability of host–parasite population interactions. I. Regulatory processes. Journal of Animal Ecology, 47, 219–247.CrossRefGoogle Scholar
Anderson, R. M., Gordon, D. M., Crawley, M. J. & Hassell, M. P. (1982). Variability in the abundance of animal and plant species. Nature, 296, 245–248.CrossRefGoogle Scholar
Andreotti, R., Gomes, A., Malavazi-Piza, K. C., et al. (2002). BmTI antigens induce a bovine protective immune response against Boophilus microplus tick. International Immunopharmacology, 2, 557–563.CrossRefGoogle ScholarPubMed
Anholt, B. R. & Werner, E. E. (1998). Predictable changes in predation mortality as a consequence of changes in food availability and predation risk. Evolutionary Ecology, 12, 729–738.CrossRefGoogle Scholar
Anonymous. (2004). Human plague in 2002 and 2003. Weekly Epidemiological Record, 79, 301–306.
Apanius, V. (1998). Stress and immune defence. In Stress and Behavior, Advances in the Study of Behavior, vol. 27, ed. M⊘ller, A. P., Milinski, M. & Slater, P. J. B.. New York: Academic Press, pp. 133–153.Google Scholar
Araújo, F. R., Silva, M. P., Lopes, A. A., et al. (1988). Severe cat flea infestation of dairy calves in Brazil. Veterinary Parasitology, 80, 83–86.CrossRefGoogle Scholar
Arneberg, P. (2002). Host population density and body mass as determinants of species richness in parasite communities: comparative analyses of directly transmitted nematodes of mammals. Ecography, 25, 88–94.CrossRefGoogle Scholar
Arneberg, P., Skorping, A. & Read, A. F. (1997). Is population density a species character? Comparative analyses of the nematode parasites of mammals. Oikos, 80, 289–300.CrossRefGoogle Scholar
Arneberg, P., Skorping, A., Grenfell, B. & Read, A. F. (1998). Host densities as determinants of abundance in parasite communities. Proceedings of the Royal Society of London B, 265, 1283–1289.CrossRefGoogle Scholar
Arzamasov, I. T. (1969). Ectoparasite assemblages of insectivores in Belorussia. In Fauna and Ecology of Animals in Belorussia, ed. Anonymous. Minsk, USSR: Academy of Sciences of the Belorussian SSR, pp. 212–221 (in Russian).Google Scholar
Atkinson, W. D. & Shorrocks, B. (1981). Competition on a divided and ephemeral recourse: a simulation model. Journal of Animal Ecology, 50, 461–471.CrossRefGoogle Scholar
Atmar, W. & Patterson, B. D. (1993). The measure of order and disorder in the distribution of species in fragmented habitat. Oecologia, 96, 373–382.CrossRefGoogle ScholarPubMed
Audy, J. R., Radovsky, F. J. & Vercammen-Grandjean, P. H. (1972). Neosomy: radical intrastadial metamorphosis associated with arthropod symbioses. Journal of Medical Entomology, 9, 487–494.CrossRefGoogle ScholarPubMed
Autino, G. A. & Lareschi, M. (1998). Siphonaptera. In Biodiversidad de Artópodos Argentinos, ed. Morrone, J. J. & Ciscaron, S.. La Plata, Argentina: Ediciones SUR, pp. 279–290.Google Scholar
Bacot, A. W. (1914). A study of bionomics of the common rat fleas and other species associated with human habitations, with special reference to the influence of temperature and humidity at various periods of the life history of the insect. Journal of Hygiene, 13, 447–654.Google ScholarPubMed
Bacot, A. W. & Martin, C. J. (1914). Observations on the mechanism of the transmission of plague by fleas. Journal of Hygiene, 13, 423–439.Google ScholarPubMed
Bahmanyar, M. & Cavanaugh, D. C. (1976). Plague Manual. Geneva, Switzerland: World Health Organization.Google Scholar
Baker, K. P. & Elharam, S. (1992). The biology of Ctenocephalides canis in Ireland. Veterinary Parasitology, 45, 141–146.CrossRefGoogle ScholarPubMed
Balashov, Y. S., Bibikova, V. A., Murzakhmetova, K. & Polunina, O. A. (1961). A flea as an environment of the plague pathogen. I. Feeding and digestion in uninfected fleas. In Proceedings of the Interdisciplinary Conference Dedicated to the 40th Anniversary of the Kazakh Soviet Socialist Republic, ed. Anonymous. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 27–30 (in Russian).
Balashov, Y. S., Bibikova, V. A., Murzakhmetova, K. & Polunina, O. A. (1965). Feeding and failure of the foregut valve function in fleas. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 35, 471–476 (in Russian).Google Scholar
Ball, S. L. & Baker, R. L. (1996). Predator-induced life history changes: antipredator behavior costs or facultative life history shifts?Ecology, 77, 1116–1124.CrossRefGoogle Scholar
Ballabeni, P. & Ward, P. I. (1993). Local adaptation of the trematode Diplostomum phoxini to the European minnow Phoxinus phoxinus, its second intermediate host. Functional Ecology, 7, 84–90.CrossRefGoogle Scholar
Banbura, J., Blondel, J., Wilde-Lambrechts, H. & Perret, P. (1995). Why do female blue tits (Parus caeruleus) bring fresh plants to their nests?Journal of Ornithology, 136, 217–221.CrossRefGoogle Scholar
Banet, M. (1986). Fever in mammals: is it beneficial?Yale Journal of Biology and Medicine, 59, 117–124.Google ScholarPubMed
Banfield, A. W. F. (1974). The Mammals of Canada. Toronto, ON: University of Toronto Press.Google Scholar
Bansemir, A. D. & Sukhdeo, M. V. K. (1996). Habitat selection of a gastrointestinal parasite: proximal cues involved in decision making. Parasitology, 113, 311–316.CrossRefGoogle Scholar
Barger, M. A. & Esch, G. W. (2002). Host specificity and the distribution–abundance relationship in a community of parasites infecting fishes in streams of North Carolina. Journal of Parasitology, 88, 446–453.Google Scholar
Barker, S. C. (1991). Evolution of host–parasite associations among species of lice and rock-wallabies: coevolution?International Journal for Parasitology, 21, 497–501.CrossRefGoogle ScholarPubMed
Barnard, C. J., Behnke, J. M., Gage, A. R, Brown, H. & Smithurst, P. R. (1998). The role of parasite-induced immunodepression, rank and social environment in the modulation of behaviour and hormone concentration in male laboratory mice (Mus musculus). Proceedings of the Royal Society of London B, 265, 693–701.CrossRefGoogle Scholar
Barnes, A. M. (1965). Three new species of the genus Anomiopsyllus. Pan-Pacific Entomologist, 41, 272–280.Google Scholar
Barnes, A. M. & Radovsky, F. J. (1969). A new Tunga (Siphonaptera) from the Nearctic region with description of all stages. Journal of Medical Entomology, 6, 19–36.CrossRefGoogle ScholarPubMed
Barnes, A. M., Tipton, V. J. & Wildie, J. A. (1977). The subfamily Anomiopsyllinae (Hystrichopsyllidae: Siphonaptera). I. A revision of the genus Anomiopsyllus Baker. Great Basin Naturalist, 37, 138–206.CrossRefGoogle Scholar
Barrera, A. (1968). The altitudinal distribution of the Siphonaptera of Mount Popocatepetl (Mexico) and a biogeographical interpretation of it. Anales del Instituto de Biologia, Universidad Nacional Autónoma de Mexico, Serie Zoologia, 39, 35–100.Google Scholar
Barriere, P., Beaucournu, J. C., Menier, K. & Colyn, M. (2002). A new flea species of the genus Allopsylla Beaucournu et Fain, 1982 (Siphonaptera: Ischnopsyllidae) from Central African Republic, on a poorly known molossid bat. Parasite, 9, 233–237.Google Scholar
Bartholomew, G. A. & Casey, T. M. (1978). Oxygen consumption of moths during rest, pre-flight warm-up, and flight in relation to body size and wing morphology. Journal of Experimental Biology, 76, 11–25.Google Scholar
Bartkowska, K. (1973). Siphonaptera Tatr Polskich. Fragmenta Faunistica (Warszawa), 19, 227–281.CrossRefGoogle Scholar
Bartolomucci, A., Palanza, P., Gaspani, L., et al. (2001). Social status in mice: behavioral, endocrine and immune changes are context dependent. Physiology and Behavior, 73, 401–410.CrossRefGoogle ScholarPubMed
Bar-Zeev, M. & Sternberg, S. (1962). Factors affecting the feeding of fleas (Xenopsylla cheopis Rothsch.) through a membrane. Entomologia Experimentalis et Applicata, 5, 60–68.CrossRefGoogle Scholar
Bashenina, N. V. (1962). The Ecology of the Common Vole. Moscow, USSR: Moscow University Press (in Russian).Google Scholar
Bashenina, N. V. (1977). Pathways of Adaptations in the Myomorph Rodents. Moscow, USSR: Nauka (in Russian).Google Scholar
Bashenina, N. V. (ed.) (1981). The Bank Vole. Moscow, USSR: Nauka (in Russian).Google Scholar
Bates, J. K. (1962). Field studies on the behaviour of bird fleas. I. Behaviour of the adults of three species of bird fleas in the field. Parasitology, 52, 113–132.CrossRefGoogle Scholar
Baumgartner, D. L. & Kane, A. (1986). Wood ducks as accidental hosts of the squirrel flea, Orchopeas howardi (Siphonaptera, Ceratophyllidae). Great Lakes Entomologist, 19, 249–250.Google Scholar
Bayreuther, K. & Brauning, S. (1971). Die Cytogenetik der Flohe (Aphaniptera). II. Xenopsylla cheopis Rothschild, 1093, und Leptopsylla segnis Schonherr, 1811. Chromosoma, 33, 19–29.CrossRefGoogle Scholar
Bazanova, L. P. & Khabarov, A. V. (2000). Block formation in fleas Citellophilus tesquorum altaicus Ioff, 1936 in relation to the gender of an insect. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 69, 42–44 (in Russian).Google Scholar
Bazanova, L. P. & Mayevsky, M. P. (1996). The duration of persistence of the plaque microbe in the organism of a flea Citellophilus tesquorum altaicus.Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 65, 45–48 (in Russian).Google Scholar
Bazanova, L. P., Voronova, G. A. & Tokmakova, E. G. (2000). Differences in the proventriculus blockage between males and females of Xenopsylla cheopis (Siphonaptera: Pulicidae). Parazitologiya, 34, 56–59 (in Russian).Google Scholar
Bazanova, L. P., Nikitin, A. Y. & Mayevsky, M. P. (2004). Seasonal changes in the aggregated state of the causative agent of plague in the organism of Citellophilus tesquorum altaicus (Siphonaptera). Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 73, 3–6 (in Russian).Google Scholar
Beaucournu, J. C. & Menier, K. (1998). Le genre Ctenocephalides Stiles et Collins, 1930 (Siphonaptera, Pulicidae). Parasite, 5, 3–16.CrossRefGoogle Scholar
Beaucournu, J. C. & Pascal, M. (1998). Origine biogéographique de Nosopsyllus fasciatus (Bosc, 1800) (Siphonaptera – Ceratophyllidae) et observations sur son hôte primitif. Biogeographica, 74, 135–132.Google Scholar
Beaucournu, J. C. & Wells, K. (2004). Trois espèces nouvelles du genre Medwayella Traub, 1972 (Insecta: Siphonaptera: Pygiopsyllidae) de Sabah (Malaisie Orientale, Ile de Bornêo). Parasite, 11, 373–377.CrossRefGoogle Scholar
Beaucournu, J. C., Piver, M. & Guiguen, C. (1993). Actualité de la conquête de l'Afrique intertropical par Pulex irritans Linné, 1758. Bulletin de la Société de Pathologie Exotique, 86, 290–294.Google Scholar
Beaucournu, J. C., Kock, D. & Menier, K. (1997). La souris Mus musculus L., 1758 est-elle l'hôte primitif de la puce Leptopsylla segnis (Schonherr, 1811) (Insecta: Siphonaptera)?Biogeographica, 73, 1–12.Google Scholar
Beaucournu, J. C., Degeilh, B. & Guiguen, C. (2005). Les puces (Insecta: Siphonaptera) parasites d'oiseaux: diversité taxonomique et dispersion biogéographique. Parasite, 12, 111–121.CrossRefGoogle Scholar
Beaucournu, J. C., Vergara, P., Balboa, L. & Gonzalez-Acuna, D. A. (2006). Description d'une nouvelle puce d'oiseau provenant du Chili (Siphonaptera: Ceratophyllidae). Parasite, 13, 227–230.CrossRefGoogle Scholar
Beck, W. (1999). Landwirtschaftliche Nutztiere als Vektoren von parasitären Epizoonoseerregern und zoophilen Dermatophyten. Hautarzt, 50, 621–628.CrossRefGoogle Scholar
Beck, W. & Clark, H. H. (1997). Differentialdiagnose medizinisch relevanter Flohspezies und ihre Bedeutung in der Dermatologie. Hautarzt, 48, 714–719.CrossRefGoogle Scholar
Behnke, J. M., Harris, P. D., Bajer, A., et al. (2004). Variation in the helminth community structure in spiny mice (Acomys dimidiatus) from four montane wadis in the St Katherine region of the Sinai Peninsula in Egypt. Parasitology, 129, 379–398.CrossRefGoogle ScholarPubMed
Beissinger, S. R. & Westphal, M. I. (1998). On the use of demographic models of population viability in endangered species management. Journal of Wildlife Management, 62, 821–841.CrossRefGoogle Scholar
Bekenov, Z. E., Serzhan, O. S., Alashbayev, M. A., et al. (2000). Fluctuations in X. skrjabini population inhabiting the Northern Cis-Aral autonomous plague focus and their causes. Quarantinable and Zoonotic Infections in Kazakhstan, 2, 52–56 (in Russian).Google Scholar
Bell, G. & Burt, A. (1991). The comparative biology of parasite species diversity: intestinal helminths of freshwater fishes. Journal of Animal Ecology, 60, 1046–1063.CrossRefGoogle Scholar
Bell, P. J., Burton, H. R., & Vanfraneker, J. A. (1988). Aspects of the biology of Glaciopsyllus antarcticus (Siphonaptera, Ceratophyllidae) during the breeding-season of a host (Fulmarus glacialoides). Polar Biology, 8, 403–410.CrossRefGoogle Scholar
Belokopytova, A. M., Tchumakova, I. V. & Kozlov, M. P. (1983). Feeding of Xenopsylla conformis, Nosopsyllus laeviceps and Citellophilus tesquorum on reptiles. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 214–215 (in Russian).Google Scholar
Belyavtseva, L. I. (2002). Phenology of fleas Citellophilus tesquorum inhabiting different areas of North Caucasus. Quarantinable and Zoonotic Infections in Kazakhstan, 6, 30–34 (in Russian).Google Scholar
Bengtson, S. A., Brinck-Lindroth, G., Lundqvist, L., Nilsson, A. & Rundgren, S. (1986). Ectoparasites on small mammals in Iceland: origin and population characteristics of a species-poor insular community. Holarctic Ecology, 9, 143–148.Google Scholar
Benjamini, E., Feingold, B. F., Young, J. D., Kartman, L. & Shimizu, M. (1963). Allergy to flea bites. IV. In vitro collection and antigenic properties of the oral secretion of the cat flea. Experimental Parasitology, 13, 143–154.CrossRefGoogle ScholarPubMed
Bennet-Clark, H. C. & Lucey, E. C. (1967). The jump of the flea: a study of the energetics and a model of the mechanism. Journal of Experimental Biology, 47, 59–67.Google Scholar
Bennett, G. F., Caines, J. R. & Bishop, M. A. (1988). Influence of blood parasites on the body mass of passeriform birds. Journal of Wildlife Diseases, 24, 339–343.CrossRefGoogle ScholarPubMed
Benton, A. H., Cerwonka, R. & Hill, J. (1959). Observations on host perception in fleas. Journal of Parasitology, 45, 614.CrossRefGoogle Scholar
Benton, A. H., Surman, M. & Krinsky, W. L. (1979). Observations on the feeding habits of some larval fleas (Siphonaptera). Journal of Parasitology, 65, 671–672.CrossRefGoogle Scholar
Berendyaeva, E. L. & Kudryavtseva, K. F. (1969). Density of fleas on Marmota sibirica in the Upper-Arym autonomous plague focus. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 55–58 (in Russian).Google Scholar
Bernotat-Danielowski, S. & Knülle, W. (1986). Ultrastructure of the rectal sac, the site of water vapor uptake from the atmosphere in larvae of the oriental rat flea Xenopsylla cheopis. Tissue and Cell, 18, 437–455.CrossRefGoogle Scholar
Berseth, W. C. & Zubac, P. (1987). Habitat associations of fleas, Siphonaptera, parasitizing the short-tailed shrew, Blarina brevicauda. Canadian Field Naturalist, 101, 594–596.Google Scholar
Betz, O. & Fuhrmann, S. (2001). Life history traits in different life forms of predaceous Stenus beetles (Coleoptera, Staphylinidae), living in waterside environments. Netherlands Journal of Zoology, 51, 371–393.CrossRefGoogle Scholar
Beutel, R. G. & Gorb, S. N. (2001). Ultrastructure of attachment specializations of hexapods (Arthropoda): evolutionary patterns inferred from a revised ordinal phylogeny. Journal of Zoological Systematics and Evolutionary Research, 39, 177–207.CrossRefGoogle Scholar
Beutel, R. G. & Pohl, H. (2006). Endopterygote systematics: where do we stand and what is the goal (Hexapoda, Arthropoda)?Systematic Entomology, 31, 202–219.CrossRefGoogle Scholar
Beveridge, I. & Chilton, N. B. (2001). Co-evolutionary relationships between the nematode subfamily Cloacininae and its macropodid marsupial hosts. International Journal for Parasitology, 21, 976–996.CrossRefGoogle Scholar
Bgytova, C. I. (1963). Data on the ecology of fleas. IV. Copulation and egg maturation in Ctenophthalmus dolichus under different environmental conditions. In Proceedings of the Scientific Conference of the Natural Focality and Prophylaxis of Plague, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 15–16 (in Russian).Google Scholar
Bibikov, D. I. & Bibikova, V. A. (1955). Studying of the Isabelline wheatear and its ectoparasites. Zoologicheskyi Zhurnal, 34, 399–407 (in Russian).Google Scholar
Bibikova, V. A. (1956). On the biology of fleas parasitic on marmots. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 2, 49–51 (in Russian).Google Scholar
Bibikova, V. A. (1963). Use of microscopy and spectral analysis when studying the hosts' blood in the fleas' stomachs. In Proceedings of the 4th Scientific Conference on Natural Focality and Prophylaxis of Plague, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 21–22 (in Russian).
Bibikova, V. A. (1965). Duration of metamorphosis in Ceratophyllus trispinus balkhaschensis Mik. 1958 under experimental conditions. In Proceedings of the 4th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute and Kainar, pp. 31–32 (in Russian).Google Scholar
Bibikova, V. A. (1977). Contemporary views on the interrelationships between fleas and the pathogens of human and animal diseases. Annual Review of Entomology, 22, 23–32.CrossRefGoogle ScholarPubMed
Bibikova, V. A. & Gerasimova, N. G. (1967). On the biology of Xenopsylla skrjabini Ioff, 1928. II. Flea feeding under experimental conditions. Zoologicheskyi Zhurnal, 46, 730–736 (in Russian).Google Scholar
Bibikova, V. A. & Gerasimova, N. G. (1973). Feeding of fleas Xenopsylla nuttalli Ioff, 1930 in the experiments. Problems of Particularly Dangerous Diseases, 29, 122–125 (in Russian).Google Scholar
Bibikova, V. A. & Klassovsky, L. N. (1974). Transmission of Plague by Fleas. Moscow, USSR: Meditsina (in Russian).Google Scholar
Bibikova, V. A. & Zhovty, I. F. (1980). Review of certain studies of fleas in the USSR, 1967–1976. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 257–272.Google Scholar
Bibikova, V. A., Ilyinskaya, V. L., Kaluzhenova, Z. P., Morozova, I. V. & Shmuter, M. F. (1963). On the biology of fleas of the genus Xenopsylla in the Sary-Ishik-Otrau Desert. Zoologicheskyi Zhurnal, 42, 1045–1050 (in Russian).Google Scholar
Bibikova, V. A., Zolotova, S. I., Murzakhmetova, K. & Leonova, T. N. (1971). Fecundity and its dynamics in two fleas parasitic on the great gerbil under experimental conditions. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 367–369 (in Russian).Google Scholar
Bidashko, F. G., Grazhdanov, A. K., Medzykhovsky, G. A., et al. (2001). Results of the studies of flea abundance in human houses in the plague focus of the West-Kazakhstan Region. Quarantinable and Zoonotic Infections in Kazakhstan, 4, 89–93 (in Russian).Google Scholar
Bidashko, F. G., Grazhdanov, A. K., Tanitovsky, B. A., et al. (2004). Attack rate of fleas Pulex irritans on humans in various biotopes of human houses. Quarantinable and Zoonotic Infections in Kazakhstan, 9, 58–61 (in Russian).Google Scholar
Biliński, S. M., Büning, J. & Simiczyjew, B. (1998). The ovaries of Mecoptera: basic similarities. Folia Histochemica et Cytobiologica, 36, 189–196.Google ScholarPubMed
Bilyalov, Z. A., Ershova, L. S., Sviridov, G. G., Sokolov, P. N. & Arakelyants, V. S. (1989). A case of experimental transmission of the causative agent of plague from Ornithodoros tartakovskyi ticks to fleas of the great gerbil. Parazitologiya, 23, 362–364 (in Russian).Google Scholar
Birtles, R. J., Harrison, T. G. & Molyneux, D. H. (1994). Grahamella in small woodland mammals in the UK: isolation, prevalence and host-specificity. Annals of Tropical Medicine and Parasitology, 88, 317–327.
Bitam, I., Parola, P., Cruz, Dittmar K., et al. (2006). First molecular detection of Rickettsia felis in fleas from Algeria. American Journal of Tropical Medicine and Hygiene, 74, 532–535.Google ScholarPubMed
Black, C. C. & Krishtalka, L. (1986). Rodents, bats, and insectivores from the Plio-Pleistocene sediments to the east of Lake Turkana, Kenya. Contributions in Science, Natural History Museum of Los Angeles County, 372, 1–15.Google Scholar
Blackburn, T. M. & Gaston, K. J. (2001). Linking patterns in macroecology. Journal of Animal Ecology, 70, 338–352.CrossRefGoogle Scholar
Blanc, G. & Baltazard, M. (1941). Transmission du bacille de Withmore par la puce du rat Xenopsylla cheopis. Comptes Rendus de l'Académie des Sciences, 213, 541–543.Google Scholar
Blanc, G. & Baltazard, M. (1942). Sur le mécanisme de la transmission de la peste par Xenopsylla cheopis. Comptes Rendus de l'Académie des Sciences, 136, 645–647.Google Scholar
Blanc, G. & Baltazard, M. (1944a). Contribution à l'étude du comportement des microbes pathogènes chez les insectes hématophages. Premier mémoire. Archives de l'Institut Pasteur du Maroc, 3, 21–49.Google Scholar
Blanc, G. & Balatazard, M. (1944b). Revue chronologique des recherches expérimentales sur la transmission et la conservation naturelle des typhus. Archives de l'Institut Pasteur du Maroc, 2, 535–715.Google Scholar
Blank, S. M., Kutzscher, C., Masello, J. F., Pilgrim, R. L. C. & Quillfeldt, P. (2007). Stick-tight fleas in the nostrils and below the tongue: evolution of an extraordinary infestation site in Hectopsylla (Siphonaptera: Pulicidae). Zoological Journal of the Linnean Society, 149, 117–137.CrossRefGoogle Scholar
Blaski, M. (1989). Fleas occurring on rodents living in the area of coal-mine Boleslaw Smialy in Laziska (Poland). Prace Naukowe Uniwersytetu Slaskiego w Katowicach, 1035, 92–96 (in Polish).Google Scholar
Blaski, M. (1991). Fleas (Siphonaptera) collected on the rodents from two industrial plants in Silesia. Prace Naukowe Uniwersytetu Slaskiego w Katowicach, 1184, 87–88 (in Polish).Google Scholar
Blaski, M. (2004). Seasonal dynamics of Ceratophyllus hirundinis (Curtis, 1826) (Siphonaptera, Insecta) in the nests of Delichon urbica (L.). Wiadomoῦci Parazytologiczne, 50, 31–34 (in Polish).Google Scholar
Blount, J. D., Houston, D. C., M⊘ller, A. P. & Wright, J. (2003). Do individual branches of immune defence correlate? A comparative case study of scavenging and non-scavenging birds. Oikos, 102, 340–350.CrossRefGoogle Scholar
Bock, C. E., Cruz, A., Grant, M. C., Aid, C. S. & Strong, T. R. (1992). Field experimental evidence for diffuse competition among southwestern riparian birds. American Naturalist, 140, 515–528.CrossRefGoogle ScholarPubMed
Bodrova, T. V. & Zhovty, I. F. (1961). Fleas of the Daurian ground squirrel in the area of the Zun-Torei Lake (S.-E. Trans-Baikalia). Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 1, 82–85 (in Russian).Google Scholar
Boeken, B. & Shachak, M. (1998). The dynamics of abundance and incidence of annual plant species richness during colonization in a desert. Ecography, 21, 63–73.CrossRefGoogle Scholar
Bogdanov, I. I., Malkova, M. G., Yakimenko, V. V. & Tantsev, A. K. (2001). Parasite–host associations between fleas and small mammals in the Omsk Region. Parazitologiya, 35, 184–191 (in Russian).Google Scholar
Bohn, T. & Amundsen, P. A. (2001). The competitive edge of an invading specialist. Ecology, 82, 2150–2163.CrossRefGoogle Scholar
Bolger, D. T., Alberts, A. C. & Soulé, M. E. (1991). Occurrence patterns of bird species in habitat fragments: sampling, extinction, and nested species subsets. American Naturalist, 137, 155–166.CrossRefGoogle Scholar
Bossard, R. L. (2002). Speed and Reynolds number of jumping cat fleas (Siphonaptera: Pulicidae). Journal of the Kansas Entomological Society, 75, 52–54.Google Scholar
Bossard, R. L. (2006). Mammal and flea relationships in the Great Basin Desert: from H. J. Egoscue's collections. Journal of Parasitology, 92, 260–266.CrossRefGoogle Scholar
Bossard, R. L., Broce, A. B. & Dryden, M. W. (2000). Effects of circadian rhythms and other bioassay factors on cat flea (Pulicidae: Siphonaptera) susceptibility to insecticides. Journal of the Kansas Entomological Society, 73, 21–29.Google Scholar
Botelho, J. R. & Linardi, P. M. (1996). Interrelações entre ectoparasitos e roedores em ambientes silvestre e urbano de Belo Horizonte, Minas Gerais, Brasil. Revista Brasileira de Entomologia, 40, 425–430.Google Scholar
Bouchet, F., Guidon, N., Dittmar, K., et al. (2003). Parasite remains in archeological sites. Memórias do Instituto Oswaldo Cruz, 98, 39–47.CrossRefGoogle Scholar
Boudreaux, H. B. (1979). Arthropod Phylogeny with Special Reference to Insects. New York: John Wiley.Google Scholar
Boughton, R. K., Atwell, J. W. & Schoech, S. J. (2006). An introduced generalist parasite, the sticktight flea (Echidnophaga gallinacea), and its pathology in the threatened Florida scrub-jay (Aphelocoma coerulescens). Journal of Parasitology, 92, 941–948.CrossRefGoogle Scholar
Boulouis, H.-J., Chang, C.-C., Henn, J. B., Kasten, R. W. & Chomel, B. B. (2005). Factors associated with the rapid emergence of zoonotic Bartonella infections. Veterinary Research, 36, 383–410.CrossRefGoogle ScholarPubMed
Bouslama, Z., Chabi, Y. & Lambrechts, M. M. (2001). Chicks resist high parasite intensities in an Algerian population of blue tits. Ecoscience, 8, 320–324.CrossRefGoogle Scholar
Bouslama, Z., Lambrechts, M. M., Ziane, N., Djenidi, R. D. & Chabi, Y. (2002). The effect of nest ectoparasites on parental provisioning in a north-African population of the blue tit Parus caeruleus. Ibis, 144, E73–E78.CrossRefGoogle Scholar
Boycott, A. E. (1913). The reaction of flea bites. Journal of Pathology and Bacteriology, 17, 110.Google Scholar
Bozeman, F. M., Sonenshine, D. E., Williams, M. S., et al. (1981). Experimental infection of ectoparasitic arthropods with Rickettsia prowazekii (GvF-16 strain) and transmission to flying squirrels. American Journal of Tropical Medicine and Hygiene, 30, 253–263.CrossRefGoogle ScholarPubMed
Bradley, T. J. & Hetz, S. (2001). Specific dynamic action in the insect Rhodnius prolixus. American Zoologist, 41, 1397.Google Scholar
Braks, M. A. H., Honorio, N. A., Lounibos, L. P., Oliveira, R. L. & Juliano, S. A. (2004). Interspecific competition between two invasive species of container mosquitoes, Aedes aegypti and Aedes albopictus (Diptera: Culicidae), in Brazil. Annals of the Entomological Society of America, 97, 130–139.CrossRefGoogle Scholar
Brändle, M., Amarell, U., Auge, H., Klotz, S. & Brandl, R. (2001). Plant and insect diversity along a pollution gradient: understanding species richness across trophic levels. Biodiversity and Conservation, 10, 1497–1511.CrossRefGoogle Scholar
Brandt, C. A. (1992). Social factors in immigration and emigration. In Animal Dispersal: Small Mammals as a Model, ed. Stenseth, N. C. & Lidicker, W. Z.. London: Chapman & Hall, pp. 96–141.CrossRefGoogle Scholar
Bray, J. R. & Curtis, J. T. (1957). An ordination of the upland forest communities of Southern Wisconsin. Ecological Monographs, 27, 325–349.
Breitschwerdt, E. B. & Kordick, D. L. (2000). Bartonella infection in animals: carriership, reservoir potential, pathogenicity, and zoonotic potential for human infection. Clinical Microbiology Reviews, 13, 428–438.CrossRefGoogle ScholarPubMed
Brinck, G. (1966). Siphonaptera from small mammals in natural foci of tick-borne encephalitis virus in Sweden. Opuscula Entomologica, 31, 156–170.Google Scholar
Brinck, G. & Löfqvist, J. (1973). The hedgehog Erinaceus europeus and its flea Archaeopsylla erinacei. Zoon (Supplement), 1, 97–103.Google Scholar
Brinck-Lindroth, G. (1968). Host spectra and distribution of fleas of small mammals in Swedish Lapland. Opuscula Entomologica, 33, 327–358.Google Scholar
Brinkerhoff, R. J., Markeson, A. B., Knouft, J. A., Gage, K. L. & Montenieri, J. A. (2006). Abundance patterns of two Oropsylla (Ceratophyllidae: Siphonaptera) species on black-tailed prairie dog (Cynomys ludovicianus) hosts. Journal of Vector Ecology, 31, 355–363.CrossRefGoogle ScholarPubMed
Brinkhof, M. W. G., Heeb, P., Kölliker, M. & Richner, H. (1999). Immunocompetence of nestling great tits in relation to rearing environment and parentage. Proceedings of the Royal Society of London B, 266, 2315–2322.CrossRefGoogle Scholar
Brooks, D. R. (1988). Macroevolutionary comparisons of host and parasite phylogenies. Annual Review of Ecology and Systematics, 19, 235–259.CrossRefGoogle Scholar
Brooks, D. R. & McLennan, D. A. (1991). Phylogeny, Ecology, and Behaviour: A Research Program in Comparative Biology. Chicago, IL: University of Chicago Press.Google Scholar
Brouwer, L. & Komdeur, J. (2004). Green nesting material has a function in mate attraction in the European starling. Animal Behaviour, 67, 539–548.CrossRefGoogle Scholar
Brown, C. R. & Brown, M. B. (1986). Ectoparasitism as a cost of coloniality in cliff swallows (Hirundo pyrrhonota). Ecology, 67, 1206–1218.CrossRefGoogle Scholar
Brown, C. R. & Brown, M. B. (1992). Ectoparasitism as a cause of natal dispersal in cliff swallows. Ecology, 73, 1718–1723.CrossRefGoogle Scholar
Brown, C. R., Brown, M. B. & Rannala, B. (1995). Ectoparasites reduce long-term survival of their avian host. Proceedings of the Royal Society of London B, 262, 313–319.CrossRefGoogle Scholar
Brown, J. H. (1975) Geographical ecology of desert rodents. In Ecology and Evolution of Communities, ed. Cody, M. L. & Diamond, J. M.. Cambridge, MA: Harvard University Press, pp. 315–341.Google Scholar
Brown, J. H. (1984). On the relationship between abundance and distribution of species. American Naturalist, 124, 255–279.CrossRefGoogle Scholar
Brown, J. H. (1995). Macroecology. Chicago, IL: University of Chicago Press.Google Scholar
Brown, J. H. & Lomolino, M. V. (1998). Biogeography. Sunderland, MA: Sinauer Associates.Google Scholar
Brown, J. H., Stevens, G. C. & Kaufman, D. M. (1996). The geographic range: size, shape, boundaries, and internal structure. Annual Review of Ecology and Systematics, 27, 597–623.CrossRefGoogle Scholar
Brown, J. S., Kotler, B. P. & Mitchell, W. M. (1994). Foraging theory, patch use, and the structure of a Negev Desert granivore community. Ecology, 75, 2286–2300.CrossRefGoogle Scholar
Brown, S. J. (1989). Pathological consequences of feeding by hematophagous arthropods: comparison of feeding strategies. In Proceedings of a Symposium ‘Physiological Interactions between Hematophagous Arthropods and their Vertebrate Hosts’, ed. Johnston, C. J. & Williams, R. E.. Lanham, MD: Entomological Society of America, pp. 4–14.Google Scholar
Bruce, W. N. (1948). Studies on the biological requirements of the cat flea. Annals of the Entomological Society of America, 41, 346–352.CrossRefGoogle Scholar
Bryukhanova, L. V. (1966). Reproduction and feeding of fleas parasitic on ground squirrels. In Particularly Dangerous Diseases in Caucasus, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 37–40 (in Russian).Google Scholar
Bryukhanova, L. V. & Myalkovskaya, C. A. (1974). On the duration of metamorphosis of fleas in the nests of the pygmy ground squirrel. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 121–124 (in Russian).Google Scholar
Bryukhanova, L. V. & Surkova, L. A. (1970). On the biology of Frontopsylla semura Wagn. et Ioff, 1926. In Vectors of Particularly Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 247–252 (in Russian).Google Scholar
Bryukhanova, L. V. & Surkova, L. A. (1983). Digestion of the blood of ground squirrels by fleas. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 220–222 (in Russian).Google Scholar
Bryukhanova, L. V., Sardar, E. A. & Levi, M. I. (1961). On the method of measurement the amount of blood taken by fleas. Proceedings of Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, 5, 28–32 (in Russian).Google Scholar
Bryukhanova, L. V., Darskaya, N. F. & Surkova, L. A. (1978). Blood digestion by fleas Leptopsylla segnis Schöncher. Parazitologiya, 12, 383–386 (in Russian).Google Scholar
Bryukhanova, L. V., Darskaya, N. F., Surkova, L. A. & Karandina, R. S. (1983). Digestion of the blood of gerbils by fleas. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 219–220 (in Russian).Google Scholar
Buckland, P. C. & Sadler, J. P. (1989). A biogeography of the human flea, Pulex irritans L. (Siphonaptera, Pulicidae). Journal of Biogeography, 16, 115–120.CrossRefGoogle Scholar
Buckle, A. P. (1978). The mark, release and recapture of fleas in a wild population of woodmice, Apodemus sylvaticus. Journal of Zoology, 186, 563–567.Google Scholar
Buechler, K., Fitze, P. S., Gottstein, B., Jacot, A. & Richner, H. (2002). Parasite-induced maternal response in a natural bird population. Journal of Animal Ecology, 71, 247–252.CrossRefGoogle Scholar
Buchman, K. (1991). Relationship between host size of Anguilla anguilla and the infection level of the monogeneans Pseudodactylogyrus spp. Journal of Fish Biology, 35, 599–601.CrossRefGoogle Scholar
Burdelov, A. S., Sabilayev, A. S., Mu, T. srepov, et al. (1999). Spatial distribution of fleas Xenopsylla skrjabini and X. gerbilli minax in colonies of the great gerbil in the north-western Pri-Balkhash area. Parazitologiya, 33, 493–496 (in Russian).Google Scholar
Burdelov, L. A., Zhubanazarov, I. Z. & Rudenchik, N. F. (1989). The size of small mammals and the number of fleas parasitizing them. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 58, 42–45 (in Russian).Google Scholar
Burdelov, S. A., Leiderman, M., Krasnov, B. R., Khokhlova, I. S. & Degen, A. A. (2007). Locomotor response to light and surface angle in three species of desert fleas. Parasitology Research, 100, 973–982.CrossRefGoogle ScholarPubMed
Burdelova, N. V. (1996). Flea fauna of some small mammals in the Dzhungarskyi Ala-Tau Ridge. In Proceedings of the Conference ‘Ecological Aspects of Epidemiology and Epizootology of Plague and Other Dangerous Diseases’, ed. Burdelov, L. A.. Almaty, Kazakhstan: Kazakh Scientific Center for Quarantine and Zoonotic Diseases, pp. 119–120 (in Russian).Google Scholar
Burdelova, N. V. & Burdelov, V. A. (1983). Dynamics of flea species composition at high and low density of the great gerbil in the Akdala Valley. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 222–223 (in Russian).Google Scholar
Burroughs, A. L. (1947). Sylvatic plague studies: the vector efficiency of nine species of fleas compared with Xenopsylla cheopis. Journal of Hygiene, 43, 371–396.CrossRefGoogle Scholar
Burrows, M. & Wolf, H. (2002). Jumping and kicking in the false stick insect Prosarthria teretrirostris: kinematics and motor control. Journal of Experimental Biology, 205, 1519–1530.Google ScholarPubMed
Bursell, E. (1974). Environmental aspects: humidity. In The Physiology of Insects, vol. 2, ed. Rockstein, M.. New York: Academic Press, pp. 44–84.Google Scholar
Bursten, S. N., Kimsey, R. B. & Owings, D. H. (1997). Ranging of male Oropsylla montana fleas via male California ground squirrel (Spermophilus beecheyi) juveniles. Journal of Parasitology, 83, 804–809.CrossRefGoogle ScholarPubMed
Busalaeva, N. N. & Fedosenko, A. K. (1964). Fleas parasitic on small mammals in the high mountain areas of the Trans-Ili Ala-Tau Ridge. Proceedings of the Institute of Zoology of Academy of Sciences of the Kazakh SSR, 22, 177–183 (in Russian).Google Scholar
Bush, A. O. & Holmes, J. C. (1986). Intestinal helminths of lesser scaup ducks: patterns of association. Canadian Journal of Zoology, 64, 132–141.CrossRefGoogle Scholar
Bush, A. O., Lafferty, K. D., Lotz, J. M. & Shostak, A. W. (1997). Parasitology meets ecology on its own terms: Margolis et al. revisited. Journal of Parasitology, 83, 575–583.CrossRefGoogle Scholar
Butler, J. M. & Roper, T. J. (1996). Ectoparasites and sett use in European badgers. Animal Behaviour, 52, 621–629.CrossRefGoogle Scholar
Buxton, P. A. (1948). Experiments with mice and fleas. I. The baby mouse. Parasitology, 39, 119–124.CrossRefGoogle ScholarPubMed
Byers, G. W. (1996). More on the origin of Siphonaptera. Journal of the Kansas Entomological Society, 69, 274–277.Google Scholar
Cabrero-Sanudo, F. J. & Lobo, J. M. (2003). Estimating the number of species not yet described and their characteristics: the case of the Western Palaearctic dung beetle species (Coleoptera, Scarabaeoidea). Biodiversity and Conservation, 12, 147–166.CrossRefGoogle Scholar
Cadiergues, M. C., Joubert, C. & Franc, M. (2000). A comparison of jump performances of the dog flea, Ctenocephalides canis (Curtis, 1826) and the cat flea, Ctenocephalides felis felis (Bouché, 1835). Veterinary Parasitology, 92, 239–241.CrossRefGoogle Scholar
Cadiergues, M. C., Santamarta, D., Mallet, X. & Franc, M. (2001). First blood meal of Ctenocephalides canis (Siphonaptera: Pulicidae) on dogs: time to initiation of feeding and duration. Journal of Parasitology, 87, 214–215.CrossRefGoogle ScholarPubMed
Cai, W.-F., Fang, Z. & Yang, G.-R. (2000). Effects of temperature, humidity on the population emergence rate of Xenopsylla cheopis in laboratory. Chinese Journal of Vector Biology and Control, 11, 358–361 (in Chinese).Google Scholar
Caley, M. J. & Schluter, D. (1997). The relationship between local and regional diversity. Ecology, 78, 70–80.CrossRefGoogle Scholar
Callaway, R. M. & King, L. (1996). Temperature-driven variation in substrate oxygenation and the balance of competition and facilitation. Ecology, 77, 1189–1195.CrossRefGoogle Scholar
Callaway, R. M. & Walker, L. R. (1997). Competition and facilitation: a synthetic approach to interactions in plant communities. Ecology, 78, 1958–1965.CrossRefGoogle Scholar
Calvete, C., Blanco-Aguiar, J. A., Virgós, E., Cabezas-Díaz, S. & Villafuerte, R. (2004). Spatial variation in helminth community structure in the red-legged partridge (Alectoris rufa L.): effects of definitive host density. Parasitology, 129, 101–113.CrossRefGoogle ScholarPubMed
Campos, E. G., Maupin, G. O., Barnes, A. M. & Eads, R. B. (1985). Seasonal occurrence of fleas (Siphonaptera) on rodents in a foothills habitat in Larimer County, Colorado, USA. Journal of Medical Entomology, 22, 266–270.CrossRefGoogle Scholar
Carlier, Y. & Truyens, C. (1995). Influence of maternal infection on offspring resistance towards parasites. Parasitology Today, 11, 94–99.CrossRefGoogle ScholarPubMed
Carney, J. P. & Dick, T. A. (2000). Helminth communities of yellow perch (Perca flavescens (Mitchill)): determinants of pattern. Canadian Journal of Zoology, 78, 538–555.CrossRefGoogle Scholar
Caro, A., Combes, C. & Euzet, L. (1997). What makes a fish a suitable host for Monogenea in the Mediterranean?Journal of Helminthology, 71, 203–210.CrossRefGoogle Scholar
Carson, H. L. (1959). Genetic conditions that promote or retard the formation of species. Cold Spring Harbor Symposia on Quantitative Biology, 24, 87–103.CrossRefGoogle ScholarPubMed
Case, T. J. & Taper, M. L. (2000). Interspecific competition, environmental gradients, gene flow, and the coevolution of species' borders. American Naturalist, 155, 583–605.CrossRefGoogle ScholarPubMed
Case, T. J., Holt, R. D., McPeek, M. A. & Keitt, T. H. (2005). The community context of species' borders: ecological and evolutionary perspectives. Oikos, 108, 28–46.CrossRefGoogle Scholar
Castleberry, S. B., Castleberry, N. L., Wood, P. B., Ford, W. M. & Mengak, M. T. (2003). Fleas (Siphonaptera) of the Allegheny woodrat (Neotoma magister) in West Virginia with comments on host specificity. American Midland Naturalist, 149, 233–236.CrossRefGoogle Scholar
Castro, J. M., Nolan, V. & Ketterson, E. D. (2001). Steroid hormones and immune function: experimental studies in wild and captive dark-eyed juncos (Junco hyemalis). American Naturalist, 157, 408–420.Google Scholar
Cavitt, J. F., Pearse, A. T. & Miller, T. A. (1999). Brown thrasher nest reuse: a time saving resource, protection from search-strategy predators, or cues for nest-site selection?Condor, 101, 859–862.CrossRefGoogle Scholar
Chadee, D. D. (1994). Distribution patterns of Tunga penetrans within a community in Trinidad, West Indies. Journal of Tropical Medicine and Hygiene, 97, 167–170.Google ScholarPubMed
Chandy, L. & Prasad, R. (1987). Behavioral resistance of hosts of flea infestation. Proceedings of the Indian National Science Academy B, 53, 27–30.Google Scholar
Charleston, M. A. (1998). Jungles: a new solution to the host/parasite phylogeny reconciliation problem. Mathematical Biosciences, 149, 191–223.CrossRefGoogle ScholarPubMed
Chase, J. M. (1996 ). Differential competitive interactions and the included niche: an experimental analysis with grasshoppers. Oikos, 76, 103–112.CrossRefGoogle Scholar
Chastel, O. & Beaucournu, J. C. (1992). Specificity and ecoethology of bird fleas in Kerguelen Islands. Annales de Parasitologie Humaine et Comparée, 67, 213–220.CrossRefGoogle Scholar
Chekchak, T., Chapuis, J. L., Pisanu, B. & Bousses, P. (2000). Introduction of the rabbit flea, Spilopsyllus cuniculi (Dale), to a subantarctic island (Kerguelen Archipelago) and its assessment as a vector of myxomatosis. Wildlife Research, 27, 91–101.CrossRefGoogle Scholar
Chesson, P. & Huntly, N. (1997). The roles of harsh and fluctuating conditions in the dynamics of ecological communities. American Naturalist, 150, 519–553.CrossRefGoogle ScholarPubMed
Christe, P., Oppliger, A. & Richner, H. (1994). Ectoparasite affects choice and use of roost sites in the great tit (Parus major). Animal Behaviour, 47, 895–898.CrossRefGoogle Scholar
Christe, P., Richner, H. & Oppliger, A. (1996a). Begging, food provisioning, and nestling competition in great tit broods infested with ectoparasites. Behavioral Ecology, 7, 127–131.CrossRefGoogle Scholar
Christe, P., Richner, H. & Oppliger, A. (1996b). Of great tits and fleas: sleep baby sleep …. Animal Behaviour, 52, 1087–1092.CrossRefGoogle Scholar
Christe, P., M⊘ller, A. P. & Lope, F. (1998). Immunocompetence and nestling survival in the house martin: the tasty chick hypothesis. Oikos, 83, 175–179.CrossRefGoogle Scholar
Christe, P., Arlettaz, R. & Vogel, P. (2000). Variation in intensity of a parasitic mite (Spinturnix myoti) in relation to the reproductive cycle and immunocompetence of its bat host (Myotis myotis). Ecology Letters, 3, 207–212.CrossRefGoogle Scholar
Christe, P., Giorgi, M. S., Vogel, P. & Arlettaz, R. (2003). Differential species-specific ectoparasitic mite intensities in two intimately coexisting sibling bat species: resource-mediated host attractiveness or parasite specialization?Journal of Animal Ecology, 72, 866–872.CrossRefGoogle Scholar
Christe, P., Michaux, J. & Morand, S. (2006). Biological conservation and parasitism. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 593–613.CrossRefGoogle Scholar
Christian, K. A. & Bedford, G. S. (1995). Physiological consequences of filarial parasites in the frillneck lizard, Chlamydosaurus kingii, in Northern Australia. Canadian Journal of Zoology, 73, 2302–2306.CrossRefGoogle Scholar
Christodoulopoulos, G. & Theodoropoulos, G. (2003). Infestation of dairy goats with the human flea, Pulex irritans, in central Greece. Veterinary Record, 152, 371–372.CrossRefGoogle ScholarPubMed
Christodoulopoulos, G., Theodoropoulos, G., Kominakis, A. & Theis, J. H. (2006). Biological, seasonal and environmental factors associated with Pulex irritans infestation of dairy goats in Greece. Veterinary Parasitology, 137, 137–143.CrossRefGoogle Scholar
Clark, F. & McNeil, D. A. C. (1981). The variation in population densities of fleas in house martin nests in Leicestershire. Ecological Entomology, 6, 379–386.CrossRefGoogle Scholar
Clark, F. & McNeil, D. A. C. (1991). Temporal variation in the population densities of fleas in house martin nests (Delichon u. urbica (Linnaeus)) in Leicestershire, U.K. Entomologist's Gazette, 42, 281–288.Google Scholar
Clark, F. & McNeil, D. A. C. (1993). A study of some factors affecting mortality overwinter in three congeneric species of bird fleas. Entomologist, 112, 55–66.Google Scholar
Clark, F., Greenwood, M. T. & Smith, J. S. (1993a). The use of an insect activity monitor in behavioral studies of the flea, Xenopsylla cheopis (Rothschild). Bulletin of the Society of Vector Ecology, 18, 26–32.Google Scholar
Clark, F., McNeil, D. A. C. & Hill, L. A. (1993b). Studies on the dispersal of three congeneric species of flea monoxenous to the house martin (Delichon urbica (L.)). Entomologist, 112, 85–94.Google Scholar
Clark, F., Deadman, D., Greenwood, M. T. & Larsen, K. S. (1997). A circadian rhythm of locomotor activity in newly emerged Ceratophyllus sciurorum. Medical and Veterinary Entomology, 11, 213–216.CrossRefGoogle ScholarPubMed
Clark, F., Deadman, D., Greenwood, M. T., Larsen, K. S. & Pudney, S. (1999). Effects of feeding on the circadian rhythm of Ceratophyllus s. sciurorum (Siphonaptera: Ceratophyllidae). Journal of Vector Ecology, 24, 78–82.Google Scholar
Clark, L. & Mason, J. R. (1988). Effect of biologically active plants used as nest material and the derived benefit to starling nestlings. Oecologia, 77, 174–180.CrossRefGoogle ScholarPubMed
Clarke, K. R. & Warwick, R. M. (1998). A taxonomic distinctness index and its statistical properties. Journal of Applied Ecology, 35, 523–531.CrossRefGoogle Scholar
Clarke, K. R. & Warwick, R. M. (1999). The taxonomic distinctness measure of biodiversity: weighting of step lengths between hierarchical levels. Marine Ecology Progress Series, 184, 21–29.CrossRefGoogle Scholar
Clarke, K. R. & Warwick, R. M. (2001). A further biodiversity index applicable to species lists: variation in taxonomic distinctness. Marine Ecology Progress Series, 216, 265–278.CrossRefGoogle Scholar
Clayton, D. H. & Cotgreave, P. (1994). Relationship of bill morphology to grooming behaviour in birds. Animal Behaviour, 47, 195–201.CrossRefGoogle Scholar
Clayton, D. H. & Moore, J. (1997). Introduction. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 1–6.Google Scholar
Clayton, D. H. & Tompkins, D. M. (1994). Ectoparasite virulence is linked to mode of transmission. Proceedings of the Royal Society of London B, 256, 211–217.CrossRefGoogle ScholarPubMed
Clayton, D. H. & Tompkins, D. M. (1995). Comparative effects of mites and lice on the reproductive success of rock doves (Columba livia). Parasitology, 110, 195–206.
Clayton, D. H., Al-Tamimi, S. & Johnston, K. P. (2003). The ecological basis of coevolutionary history. In Tangled Trees: Phylogeny, Cospeciation and Coevolution, ed. Page, R. D. M.. Chicago, IL: University of Chicago Press, pp. 310–342.Google Scholar
Clements, A. N. (1992). The Biology of Mosquitoes: Development, Nutrition and Reproduction, vol. 1. Wallingford, UK: CAB International.Google Scholar
Clements, J. F. (2007). The Clements Checklist of Birds of the World, 6th edn. London: Christopher Helm.
Cole, L. C. & Koepke, J. A. (1946). A study of rodent ectoparasites in Mobile, Alabama. Public Health Reports, 61, 1469–1487.CrossRefGoogle Scholar
Cole, L. C. & Koepke, J. A. (1947a). Problems of interpretation of the data of rodent–ectoparasite surveys. Public Health Reports(Supplement), 202, 1–24.Google Scholar
Cole, L. C. & Koepke, J. A. (1947b). A study of rodent ectoparasites in Honolulu. Public Health Reports (Supplement), 202, 25–41.Google Scholar
Cole, L. C. & Koepke, J. A. (1947c). A study of rodent ectoparasites in Savannah, Georgia. Public Health Reports (Supplement), 202, 42–59.Google Scholar
Cole, L. C. & Koepke, J. A. (1947d). A study of rodent ectoparasites in Dothan, Alabama. Public Health Reports (Supplement), 202, 61–71.Google Scholar
Collen, B., Purvis, A. & Gittleman, J. L. (2004). Biological correlates of description date in carnivores and primates. Global Ecology and Biogeography, 13, 459–467.CrossRefGoogle Scholar
Collinge, S. K., Johnson, W. C., Ray, C., et al. (2005). Landscape structure and plague occurrence in black-tailed prairie dogs on grasslands of the western USA. Landscape Ecology, 20, 941–955.CrossRefGoogle Scholar
Colwell, R. K. (2000). Rensch's rule crosses the line: convergent allometry of sexual size dimorphism in hummingbirds and flower mites. American Naturalist, 156, 495–510.Google ScholarPubMed
Combes, C. (1997). Fitness of parasites: pathology and selection. International Journal for Parasitology, 27, 1–10.CrossRefGoogle Scholar
Combes, C. (2001). Parasitism: The Ecology and Evolution of Intimate Interactions. Chicago, IL: University of Chicago Press.Google Scholar
Combes, C. (2005). The Art of Being a Parasite. Chicago, IL: University of Chicago Press.Google Scholar
Cooke, B. D. (1990a). Rabbit burrows as environments for the European rabbit fleas, Spilopsyllus cuniculi (Dale), in arid South Australia. Australian Journal of Zoology, 38, 317–325.CrossRefGoogle Scholar
Cooke, B. D. (1990b). Notes on the comparative reproductive biology and the laboratory breeding of the rabbit flea Xenopsylla cunicularis Smit (Siphonaptera, Pulicidae). Australian Journal of Zoology, 38, 527–534.CrossRefGoogle Scholar
Cooke, B. D. (1999). Notes on the life history of the rabbit flea Caenopsylla laptevi ibera Beaucornu & Marquez, 1987 (Siphonaptera: Ceratophyllidae) in eastern Spain. Parasite, 6, 347–354.CrossRefGoogle Scholar
Cooke, B. D. & Skewes, M. A. (1988). The effect of temperature and humidity on the survival and development of the European rabbit flea Spilopsyllus cuniculi (Dale). Australian Journal of Zoology, 36, 449–459.CrossRefGoogle Scholar
Cooper, J. E. (1976). Tunga penetrans infestation in pigs. Veterinary Record, 98, 472.CrossRefGoogle ScholarPubMed
Cornell, H. V. (1993). Unsaturated patterns in species assemblages: the role of regional processes in setting local species richness. In Species Diversity in Ecological Communities: Historical and Geographical Perspectives, ed. Ricklefs, R. F. & Schluter, D.. Chicago, IL: University of Chicago Press, pp. 243–252.Google Scholar
Cornell, H. V. & Lawton, J. H. (1992). Species interactions, local and regional processes, and limits to richness of ecological communities: a theoretical perspective. Journal of Animal Ecology, 61, 1–12.CrossRefGoogle Scholar
Correia, T. R., Souza, C. P., Fernandes, J. I., et al. (2003). Life cycle of Ctenocephalides felis felis (Bouché, 1835) (Siphonaptera, Pulicidae) from different artificial diets. Revista Brasiliera de Zoociências, Juiz de Fora, 5, 153–160.Google Scholar
Côté, I. M. & Poulin, R. (1995). Parasitism and group size in social animals: a meta-analysis. Behavioral Ecology, 6, 159–165.CrossRefGoogle Scholar
Cotgreave, P. & Clayton, D. H. (1994). Comparative analysis of time spent grooming by birds in relation to parasite load. Behaviour, 131, 171–187.CrossRefGoogle Scholar
Cotton, M. J. (1970a). The reproductive biology of Ctenophthalmus nobilis (Rothschild) (Siphonaptera). Proceedings of the Royal Entomological Societyof London A, 45, 141–148.CrossRefGoogle Scholar
Cotton, M. J. (1970b). The life history of the hen flea, Ceratophyllus gallinae (Schrank) (Siphonaptera, Ceratophyllidae). Entomologist, 103, 45–48.Google Scholar
Cox, F. E. G. (2001). Concomitant infections, parasites and immune responses. Parasitology, 122, S23–S38.CrossRefGoogle ScholarPubMed
Cox, R., Stewart, P. D. & Macdonald, D. W. (1999). The ectoparasites of the European badger, Meles meles, and the behavior of the host-specific flea, Paraceras melis. Journal of Insect Behavior, 12, 245–265.CrossRefGoogle Scholar
Craw, R. C., Grehan, J. R. & Heads, M. J. (1999). Panbiogeography: Tracking the History of Life.New York: Oxford University Press.Google Scholar
Cresswell, J. E., Vidal-Martinez, V. M. & Crichton, N. J. (1995). The investigation of saturation in the species richness of communities: some comments on methodology. Oikos, 72, 301–304.CrossRefGoogle Scholar
Croll, N. A., Anderson, R. M., Gyerkos, T. W. & Ghardian, E. (1982). The population biology and control of Ascaris lumbricoides in a rural community in Iran. Transactions of the Royal Society of Tropical Medicine and Hygiene, 76, 187–197.CrossRefGoogle Scholar
Crooks, K. R., Scott, C. A., Angeloni, L., et al. (2001). Ectoparasites of the island fox on Santa Cruz Island. Journal of Wildlife Diseases, 37, 189–193.CrossRefGoogle Scholar
Crooks, K. R., Garcelon, D. K., Scott, C. A., et al. (2004). Ectoparasites of a threatened insular endemic mammalian carnivore: the island spotted skunk. American Midland Naturalist, 151, 35–41.CrossRefGoogle Scholar
Crowell, K. L. & Pimm, S. L. (1976). Competition and niche shifts introduced onto small islands. Oikos, 27, 251–258.CrossRefGoogle Scholar
Crum, G. E., Knapp, F. W. & White, G. M. (1974). Response of the cat flea, Ctenocephalides felis (Bouche), and the Oriental rat flea, Xenopsylla cheopis (Rothschild), to electromagnetic radiation in the 300–700 nanometer range. Journal of Medical Entomology, 11, 88–94.CrossRefGoogle Scholar
Cumming, G. S. & Bernard, R. T. (1997). Rainfall, food abundance and timing of parturition in African bats. Oecologia, 111, 309–317.CrossRefGoogle ScholarPubMed
Cyprich, D., Krumpál, M. & Rolníková, T. (1999). Occurrence and distribution of Ceratophyllus gallinae (Schrank, 1803) (Siphonaptera) in Slovakia. Folia Faunistica Slovaca, 4, 79–88 (in Slovak).Google Scholar
Cyprich, D., Pinowski, J. & Krumpál, M. (2002). Seasonal changes in numbers of fleas (Siphonaptera) in nests of the house sparrow (Passer domesticus) and tree sparrow (P. montanus) in Warsaw surroundings (Poland). Acta Parasitologica, 47, 58–65.Google Scholar
Cyprich, D., Štiavnická, L. & Kiefer, M. S. (2003). Information about occurrence and distribution specific flea species (Siphonaptera) of the ground squirrel (Spermophilus citellus L., 1758) in Slovakia: the genus Citellophilus and Neopsylla. Folia Faunistica Slovaca, 8, 47–51 (in Slovak).Google Scholar
Cyprich, D., Krumpál, M. & Duda, M. (2006). Characteristic of flea fauna (Siphonaptera) of ecological group of birds (Aves) nesting on the ground. Entomofauna Carpathica, 18, 1–4 (in Slovak).Google Scholar
Dallai, R., Lupetti, P., Afzelius, B. A. & Frati, F. (2003). Sperm structure of Mecoptera and Siphonaptera (Insecta) and the phylogenetic position of Boreus hyemalis. Zoomorphology, 122, 211–220.CrossRefGoogle Scholar
Dampf, A. (1911). Palaeopsylla klebsiana n. sp., eine fossiler Floh aus dem baltischen Bernstein. Schriften der Physikalisch–Ökonomischen Gesellschaft zu Königsberg, 51, 248–259.Google Scholar
Darby, C., Ananth, S. L., Tan, L. & Hinnebusch, B. J. (2005). Identification of gmhA, a Yersinia pestis gene required for flea blockage, by using a Caenorhabditis elegans biofilm system. Infection and Immunity, 73, 7236–7242.CrossRefGoogle ScholarPubMed
Darskaya, N. F. (1954). Fleas of the Daurian ground squirrel. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 12, 245–257 (in Russian).Google Scholar
Darskaya, N. F. (1955). Ecological characteristics of Xenopsylla gerbilli caspica. I. Fleas parasitic on the great gerbil, associated with the ecological features of their host. In Natural Focality of Human Diseases and Regional Epidemiology, ed. ,Anonymous. Moscow, USSR: Medgiz, pp. 400–407 (in Russian).Google Scholar
Darskaya, N. F. (1964). On the comparative ecology of fleas belonging to the genus Ceratophyllus. Ectoparasites, 4, 31–180 (in Russian).Google Scholar
Darskaya, N. F. (1970). Ecological comparisons of some fleas of the USSR fauna. Zoologicheskyi Zhurnal, 49, 729–745 (in Russian).Google Scholar
Darskaya, N. F. & Besedina, K. P. (1961). On the possibility of fleas feeding on reptiles. Proceeding of Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, 5, 33–39 (in Russian).Google Scholar
Darskaya, N. F. & Karandina, P. C. (1974). Observations on pre-imaginal development of Neopsylla setosa Wagn., 1898: fleas parasitic on ground squirrels. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Disieases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 134–137 (in Russian).Google Scholar
Darskaya, N. F., Brukhanova, L. V. & Kunitskaya, N. T. (1965). On the method of investigation of flea oviposition. In Problems of General Zoology and Medical Entomology, ed. ,Anonymous. Moscow, USSR: Moscow University Press, pp. 6–9.Google Scholar
Darskaya, N. F., Bragina, Z. S. & Petrov, V. G. (1970). On fleas of the common vole and shrews in dependence of sharp density fluctuations of these mammals. In Vectors of Particularly Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 132–152 (in Russian).Google Scholar
Davidson, D. W. (1980). Some consequences of diffuse competition in a desert ant community. American Naturalist, 116, 92–105.CrossRefGoogle Scholar
Davidson, D. W. (1985). An experimental study of diffuse competition in harvester ants. American Naturalist, 125, 500–506.CrossRefGoogle Scholar
Davis, D. E. (1951). Observations on rat ectoparasites and typhus fever in San Antonio, Texas. Public Health Reports, 66, 1717–1726.CrossRefGoogle ScholarPubMed
Davis, R. M., Smith, R. T., Madon, M. B. & Sitko-Cleugh, E. (2002). Flea, rodent and plague ecology at Chichupate Campground, Ventura County, California. Journal of Vector Ecology, 27, 107–127.Google ScholarPubMed
Dawson, L. H. J., Renaud, F., Guégan, J. F. & MeeÛs, T. (2000). Experimental evidence of asymmetrical competition between two species of parasitic copepods. Proceedings of the Royal Society of London B, 267, 1973–1978.CrossRefGoogle ScholarPubMed
Dawson, R. D. (2004). Does fresh vegetation protect avian nests from ectoparasites? An experiment with tree swallows. Canadian Journal of Zoology, 82, 1005–1010.
Dawson, R. D. & Bortolotti, G. R. (1997). Ecology of parasitism of nestling American kestrels by Carnus hemapterus (Diptera, Carnidae). Canadian Journal of Zoology, 75, 2021–2026.CrossRefGoogle Scholar
Day, J. F. & Benton, A. H. (1980). Population dynamics and coevolution of adult siphonapteran parasites of the southern flying squirrel (Glaucomys volans volans). American Midland Naturalist, 103, 333–338.CrossRefGoogle Scholar
Cardoso, Albuquerque V. & Linardi, P. M. (2006). Scanning electron microscopy studies of sensilla and other structures of the head of Polygenis (Polygenis) tripus (Siphonapera: Rhopalopsyllidae). Micron, 37, 557–565.CrossRefGoogle Scholar
Dean, S. R. & Meola, R. W. (1997). Effect of juvenile hormone and juvenile hormone mimics on sperm transfer from the testes of the male cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 34, 485–488.CrossRefGoogle Scholar
Dean, S. R. & Meola, R. W. (2002a). Factors influencing sperm transfer and insemination in cat fleas (Siphonaptera: Pulicidae) fed on an artificial membrane system. Journal of Medical Entomology, 39, 475–479.CrossRefGoogle Scholar
Dean, S. R. & Meola, R. W. (2002b). Effect of diet composition on weight gain, sperm transfer, and insemination in the cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 39, 370–375.CrossRefGoogle Scholar
Carvalho, R. W., Almeida, A. B., Barbosa-Silva, S. C., et al. (2003). The patterns of tungiasis in Araruama township, state of Rio de Janeiro, Brazil. Memórias do Instituto Oswaldo Cruz, 98, 31–36.CrossRefGoogle ScholarPubMed
Degen, A. A. (1997). Ecophysiology of Small Desert Mammals. New York: Springer-Verlag.CrossRefGoogle Scholar
Degen, A. A. (2006). Effect of macroparasites on the energy budget of small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 371–399.CrossRefGoogle Scholar
Degen, A. A., Kam, M., Khokhlova, I. S., Krasnov, B. R. & Barraclough, T. (1998). Average daily metabolic rate of rodents: habitat and dietary comparisons. Functional Ecology, 12, 63–73.CrossRefGoogle Scholar
Delahay, R. J., Speakman, J. R. & Moss, R. (1995). The energetic consequences of parasitism: effects of a developing infection of Trichostongylus tenius (Nematoda) of red grouse (Lagopus lagopus scoticus) energy balance, body weight and condition. Parasitology, 110, 473–482.CrossRefGoogle Scholar
Lope, F., M⊘ller, A. P. & Cruz, C. (1998). Parasitism, immune response and reproductive success in the house martin Delichon urbica. Oecologia, 114, 188–193.CrossRefGoogle Scholar
Demas, G. E. & Nelson, R. J. (1998). Photoperiod, ambient temperature, and food availability interact to affect reproductive and immune function in adult male deer mice (Peromyscus maniculatus). Journal of Biological Rhythms, 13, 253–262.CrossRefGoogle Scholar
Demin, E. P., Zagniborodova, E. N., Sageev, M. T., et al. (1970). Some features of ecology of fleas parasitic on Rhombomys opimus in western Turkmenistan in relation to their importance in the epizootology of plague. Problems of Particularly Dangerous Diseases, 11, 49–56 (in Russian).Google Scholar
Moraes, L. B., Bossi, D. E. P. & Linhares, A. X. (2003). Siphonaptera parasites of wild rodents and marsupials trapped in three mountain ranges of the Atlantic Forest in Southeastern Brazil. Memórias do Instituto Oswaldo Cruz, 98, 1071–1076.CrossRefGoogle ScholarPubMed
Hollander, N. & Allen, J. R. (1986). Cross-reactive antigens between a tick Dermacentor variabilis (Acari: Ixodidae) and a mite Prosoptes cuniculi (Acari: Psoroptidae). Journal of Medical Entomology, 23, 44–50.CrossRefGoogle Scholar
Deoras, P. J. & Prasad, R. S. (1967). Feeding mechanisms of Indian fleas X. cheopis (Roths.) and X. astia (Roths.). Indian Journal of Medical Research, 55, 1041–1050.Google Scholar
Pedro, N., Delgado, M. J., Gancedo, B. & Alonso-Bedate, M. (2003). Changes in glucose, glycogen, thyroid activity and hypothalamic catecholamines in tench by starvation and refeeding. Journal of Comparative Physiology B, 173, 475–481.CrossRefGoogle ScholarPubMed
Desdevises, Y., Jovelin, R., Jousson, O. & Morand, S. (2000). Comparison of ribosomal DNA sequences of Lamellodiscus spp. (Monogenea, Diplectanidae) parasitising Pagellus (Sparidae, Teleostei) in the north Mediterranean Sea: species divergence and coevolutionary interactions. International Journal for Parasitology, 30, 741–746.CrossRefGoogle ScholarPubMed
Desdevises, Y., Morand, S., Jousson, O. & Legendre, P. (2002a). Coevolution between Lamellodiscus (Monogenea: Diplectanidae) and Sparidae (Teleostei): the study of a complex host–parasite system. Evolution, 56, 2459–2471.CrossRefGoogle Scholar
Desdevises, I., Morand, S. & Legendre, P. (2002b). Evolution and determinants of host specificity in the genus Lamellodiscus (Monogenea). Biological Journal of the Linnean Society, 77, 431–443.CrossRefGoogle Scholar
Devi, S. G. & Prasad, R. S. (1985). Phagostimulants in artificial feeding systems of rat fleas Xenopsylla cheopis and Xenopsylla astia. Proceedings of the Indian National Science Academy B, 51, 566–573.Google Scholar
Dezfuli, B. S., Giari, L., Biaggi, S. & Poulin, R. (2001). Associations and interactions among intestinal helminths of the brown trout, Salmo trutta, in northern Italy. Journal of Helminthology, 75, 331–336.Google ScholarPubMed
Dick, C. W. & Patterson, B. D. (2006). Bat flies: obligate parasites of bats. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 179–194.CrossRefGoogle Scholar
Dick, C. W., Gannon, M. R., Little, W. E. & Patrick, M. J. (2003). Ectoparasite associations of bats from central Pennsylvania. Journal of Medical Entomology, 40, 813–819.CrossRefGoogle ScholarPubMed
Diamond, J. M. (1975). Assembly of species communities: chance or competition? In Ecology and Evolution of Communities, ed. Cody, M. L. & Diamond, J. M.. Cambridge, MA: Harvard University Press, pp. 342–444.Google Scholar
Cruz, Dittmar K. & Whiting, M. (2003). Genetic and phylogeographic structure of populations of Pulex simulans (Siphonaptera) in Peru inferred from two genes (CytB and CoII). Parasitology Research, 91, 55–59.CrossRefGoogle Scholar
Cruz, Dittmar K., Mamat, U., Whiting, M., et al. (2003). Techniques of DNA-studies on prehispanic ectoparasites (Pulex sp., Pulicidae, Siphonaptera) from animal mummies of the Chiribaya Culture, Southern Peru. Memórias do Instituto Oswaldo Cruz, 98, 53–58.Google Scholar
Dobson, A. P. (1985). The population dynamics of competition between parasites. Parasitology, 91, 317–347.
Dobson, A. P. (1990). Models for multi-species parasite–host communities. In Parasite Communities: Patterns and Processes, ed. Esch, G., Bush, A. O. & Aho, J. M.. London: Chapman & Hall, pp. 261–288.Google Scholar
Dobson, A. P. & Roberts, M. (1994). The population dynamics of parasitic helminth communities. Parasitology, 109, S97–S108.
Dogiel, V. A., Petrushevski, G. K. & Polyanski, Y. I. (1961). Parasitology of Fishes. Edinburgh, UK: Oliver & Boyd.Google Scholar
Dong, B. (1991). Experimental studies on the transmission of hemorrhagic fever with renal syndrome virus by gamasid mites and fleas. Chinese Medical Journal, 71, 502–504 (in Chinese).Google Scholar
Dowling, A. P. G. (2006). Mesostigmatid mites as parasites of small mammals: systematics, ecology, and the evolution of parasitic associations. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 103–117.CrossRefGoogle Scholar
Downing, J. A. (1986). Spatial heterogeneity: evolved behaviour or mathematical artefact. Nature, 323, 255–257.CrossRefGoogle Scholar
Drobney, R. D., Train, C. T. & Gredrickson, L. H. (1983). Dynamics of the platyhelminth fauna of wood ducks in relation to food habits and reproductive state. Journal of Parasitology, 69, 375–380.CrossRefGoogle ScholarPubMed
Dryden, M. W. (1989). Host association, on-host longevity and egg production of Ctenocephalides felis felis. Veterinary Parasitology, 34, 117–122.CrossRefGoogle ScholarPubMed
Dryden, M. W. & Blakemore, J. (1989). A review of flea allergy dermatitis in the dog and cat. Companion Animal Practice, 19, 10–16.Google Scholar
Dryden, M. W. & Broce, A. B. (1993). Development of a trap for collecting newly emerged Ctenocephalides felis (Siphonaptera: Pulicidae) in homes. Journal of Medical Entomology, 30, 901–906.CrossRefGoogle Scholar
Dubinina, V. B. & Dubinin, M. N. (1951). Parasite fauna of mammals of the Dauric Steppe. Fauna and Ecology of Rodents, 4, 98–156 (in Russian).Google Scholar
Dubyansky, M. A., Dubyanskaya. L. D., Zhubanazarov, I. Z. & Filipchenko, V. E. (1975). Alternative method for prognosis of abundance of Xenopsylla fleas. In Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 93–96 (in Russian).Google Scholar
Duffy, J. E., Richardson, J. P. & France, K. E. (2005). Ecosystem consequences of diversity depend on food chain length in estuarine vegetation. Ecology Letters, 8, 301–309.CrossRefGoogle Scholar
Dufva, R, & Allander, K. (1996). Variable effects of the hen flea Ceratophyllus gallinae on the breeding success of the great tit Parus major in relation to weather conditions. Ibis, 138, 772–777.CrossRefGoogle Scholar
Dunnet, G. M. & Mardon, D. K. (1974). A monograph of Australian fleas (Siphonaptera). Australian Journal of Zoology (Supplemental Series), 30, 1–273.Google Scholar
Durden, L. A. (1995). Fleas (Siphonaptera) of cotton mice on a Georgia Barrier Island: a depauperate fauna. Journal of Parasitology, 81, 526–529.CrossRefGoogle ScholarPubMed
Durden, L. A. & Beaucournu, J. C. (2002). Gymnomeropsylla n. gen. (Siphonaptera: Pygiopsyllidae) from Sulawesi, Indonesia, with the description of two new species. Parasite, 9, 225–232.CrossRefGoogle ScholarPubMed
Durden, L. A. & Kollars, T. M. (1997). The fleas (Siphonaptera) of Tennessee. Journal of Vector Ecology, 22, 13–22.Google ScholarPubMed
Durden, L. A., Ellis, B. A., Banks, C. W., Crowe, J. D. & Oliver, J. H. (2004). Ectoparasites of gray squirrels in two different habitats and screening of selected ectoparasites for bartonellae. Journal of Parasitology, 90, 485–489.CrossRefGoogle ScholarPubMed
Durden, L. A., Judy, T. N., Martin, J. E. & Spedding, L. S. (2005). Fleas parasitizing domestic dogs in Georgia, USA: species composition and seasonal abundance. Veterinary Parasitology, 130, 157–162.CrossRefGoogle ScholarPubMed
Durden, L. A., Cunningham, M. W., McBride, R. & Ferree, B. (2006). Ectoparasites of free-ranging pumas and jaguars in the Paraguayan Chaco. Veterinary Parasitology, 137, 189–193.CrossRefGoogle ScholarPubMed
Dynesius, M. & Jansson, R. (2000). Evolutionary consequences of changes in species' geographical distributions driven by Milankovitch climate oscillations. Proceedings of the National Academy of Sciences of the USA, 97, 9115–9120.CrossRefGoogle ScholarPubMed
Ebert, D. (1994). Virulence and local adaptation of a horizontally transmitted parasite. Science, 265, 1084–1086.CrossRefGoogle ScholarPubMed
Eckstein, R. A. & Hart, B. L. (2000a). The organization and control of grooming in cats. Applied Animal Behaviour Science, 68, 131–140.CrossRefGoogle Scholar
Eckstein, R. A. & Hart, B. L. (2000b). Grooming and control of fleas in cats. Applied Animal Behaviour Science, 68, 141–150.CrossRefGoogle Scholar
Edney, E. B. (1945). Laboratory studies on the bionomics of the rat fleas, Xenopsylla brasiliensis Baker and X. cheopis Roths. I. Certain effects of light, temperature and humidity on the rate of development and on adult longevity. Bulletin of Entomological Research, 35, 399–416.Google Scholar
Edney, E. B. (1947a). Laboratory studies on the bionomics of the rat fleas, Xenopsylla brasiliensis Baker and X. cheopis Rothsch. II. Water relations during the cocoon period. Bulletin of Entomological Research, 38, 263–280.CrossRefGoogle Scholar
Edney, E. B. (1947b). Laboratory studies on the bionomics of the rat fleas, Xenopsylla brasiliensis Baker and X. cheopis Roths. III. Further factors affecting adult longevity. Bulletin of Entomological Research, 38, 389–404.CrossRefGoogle Scholar
Eeley, H. A. C. & Foley, R. A. (1999). Species richness, species range size and ecological specialization among African primates: geographical patterns and conservation implications. Biodiversity and Conservation, 8, 1033–1056.CrossRefGoogle Scholar
Eeva, T., Lehikoinen, E. & Nurmi, J. (1994). Effects of ectoparasites on breeding success of great tits (Parus major) and pied flycatchers (Ficedula hypoleuca) in an air-pollution gradient. Canadian Journal of Zoology, 72, 624–635.CrossRefGoogle Scholar
Egoscue, H. J. (1976). Flea exchange between deer mice and some associated small mammals in western Utah. Great Basin Naturalist, 36, 475–480.Google Scholar
Eisele, M., Heukelbach, J., Marck, E., et al. (2003). Investigations on the biology, epidemiology, pathology and control of Tunga penetrans in Brazil. I. Natural history of tungiasis in man. Parasitology Research, 90, 87–99.Google Scholar
Eiseman, C. H. & Binnengton, K. C. (1994). The peritrophic membrane: its formation, structure, chemical composition and permeability in relation to vaccination against ectoparasitic arthropods. International Journal for Parasitology, 24, 15–26.CrossRefGoogle Scholar
Elliot, J. M. (1977). Some Methods for Statistical Analysis of Samples of Benthic Invertebrates, 2nd edn, Freshwater Biological Association Sceintific Publications No. 25. Ambleside, UK: Titus Wilson & Son.Google Scholar
Ellis, B. A., Regnery, R. L., Beati, L., et al. (1999). Rats of the genus Rattus are reservoir hosts for pathogenic Bartonella species: an Old World origin for a New World disease. Journal of Infectious Diseases, 180, 220–224.CrossRefGoogle ScholarPubMed
Elmes, G. W., Barr, B., Thomas, J. A. & Clarke, R. T. (1999). Extreme host specificity by Microdon mutabilis (Diptera: Syrphidae), a social parasite of ants. Proceedings of the Royal Society of London B, 266, 447–453.CrossRefGoogle Scholar
Elshanskaya, N. I. & Popov, M. N. (1972). Zoologico-parasitological characteristics of the River Kenkeme valley (Central Yakutia). In Theriology, vol. 1, ed. Kolosova, L. D. & Lukyanova, I. V.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 368–372 (in Russian).Google Scholar
Emelianova, N. D. & Shtilmark, F. R. (1967). Fleas of insectivores, rodents and lagomorphs of the central part of Western Sayan. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 27, 241–253 (in Russian).Google Scholar
Engelthaler, D. M., Hinnebusch, B. J., Rittner, C. M. & Gage, K. L. (2000). Quantitative competitive PCR as a technique for exploring flea–Yersina pestis dynamics. American Journal of Tropical Medicine and Hygiene, 62, 552–560.CrossRefGoogle ScholarPubMed
Enquist, B. J., Haskell, J. P. & Tiffney, B. H. (2002). General patterns of taxonomic and biomass partitioning in extant and fossil plant communities. Nature, 419, 610–613.CrossRefGoogle ScholarPubMed
Esbérard, C. (2001). Infestation of Rhynchopsyllus pulex (Siphonaptera: Tungidae) on Molossus molossus (Chiroptera) in Southeastern Brazil. Memórias do Instituto Oswaldo Cruz, 96, 1169–1170.CrossRefGoogle Scholar
Euzet, L. & Combes, C. (1980). Les problèmes de l'espèce chez les animaux parasites. Bulletin de la Société Zoologique de France, 40, 239–285.Google Scholar
Evans, F. G. & Freeman, R. B. (1950). On the relationship of some mammal fleas to their hosts. Annals of the Entomological Society of America, 43, 320–333.CrossRefGoogle Scholar
Ezenwa, V. O. (2004). Host social behavior and parasitic infection: a multifactorial approach. Behavioral Ecology, 15, 446–454.CrossRefGoogle Scholar
Fain, A. & Hyland, K. W. (1985). Evolution of astigmatid mites on mammals. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 641–658.Google Scholar
Fairbairn, D. J. (1997). Allometry for sexual size dimorphism: pattern and process in the coevolution of body size in males and females. Annual Review of Ecology and Systematics, 28, 659–687.CrossRefGoogle Scholar
Fairbairn, D. J. (2005). Allometry for sexual size dimorphism: testing two hypotheses for Rensch's rule in the water strider Aquarius remigis. American Naturalist, 166, S69–S84.Google ScholarPubMed
Farhang-Azad, A., Traub, R. & Wisseman, C. L. (1983). Rickettsia mooseri infection in the fleas Leptopsylla segnis and Xenopsylla cheopis. American Journal of Tropical Medicine and Hygiene, 32, 1392–1400.CrossRefGoogle ScholarPubMed
Farhang-Azad, A., Traub, R., Sofi, M. & Wisseman, C. L. (1984). Experimental murine typhus infection in the cat fleas, Ctenocephalides felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 21, 675–680.CrossRefGoogle Scholar
Farhang-Azad, A., Traub, R. & Boqar, S. (1985). Transovarial transmission of murine typhus rickettsia in Xenopsylla cheopis fleas. Science, 227, 543–545.CrossRefGoogle ScholarPubMed
Farhang-Azad, A., Radulovic, S., Higgins, J. A., Noden, B. H. & Troyer, J. M. (1997). Flea-borne rickettsioses: ecologic considerations. Emerging Infectious Diseases, 3, 319–327.CrossRefGoogle Scholar
Faulkenberry, G. D. & Robbins, R. G. (1980). Statistical measures of interspecific association between the fleas of the gray-tailed vole, Microtus canicaudus Miller. Entomological News, 91, 93–101.Google Scholar
Faust, E. C. & Maxwell, T. A. (1930). The finding of the larva of the chigo, Tunga penetrans, in scrapings from human skin. Archives of Dermatology and Syphilology, 22, 94–97.CrossRefGoogle Scholar
Fauth, P. T., Krementz, D. G. & Hines, J. E. (1991). Ectoparasitism and the role of green nesting material in the European starling. Oecologia, 88, 22–29.CrossRefGoogle ScholarPubMed
Fedorov, Y. V., Igolkin, N. I. & Tyushnikova, M. K. (1959). Some data on virus-carrying fleas in areas of tick-borne encephalitis and lymphocytic choriomenengitis. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 28, 149–152 (in Russian).Google Scholar
Feingold, B. F., Benjamini, E. & Michaeli, D. (1968). The allergic responses to insect bites. Annual Review of Entomology, 13, 138–158.CrossRefGoogle Scholar
Feliu, C., Renaud, F., Catzeflis, F., et al. (1997). Comparative analysis of parasite species richness of Iberian rodents. Parasitology, 115, 453–466.CrossRefGoogle ScholarPubMed
Fellis, K. J., Negovetich, N. J., Esch, G. W., Horak, I. G. & Boomker, J. (2003). Patterns of association, nestedness, and species co-occurrence of helminth parasites in the greater kudu, Tragelaphus strepsiceros, in the Kruger National Park, South Africa, and the Etosha National Park, Namibia. Journal of Parasitology, 89, 899–907.CrossRefGoogle ScholarPubMed
Felsenstein, J. (1985). Phylogenies and the comparative method. American Naturalist, 125, 1–15.CrossRefGoogle Scholar
Feng, X.-Y., Ni, E.-J., Cao, P.-G., Zheng, C.-J. & Yang, G.-H. (2004). Investigation of the distributive features of flea population during epidemic outbreak of plague in Longlin County, Guangxi. Acta Parasitologica et Medica Entomologica Sinica, 11, 235–237 (in Chinese).Google Scholar
Feng, Y.-M., Li, W. & Wang, Z.-Y. (2003). Observation on the life cycle of Ischnopsyllus octactenus at laboratory. Chinese Journal of Vector Biology and Control, 14, 202–203 (in Chinese).Google Scholar
Fenner, F. & Ratcliff, F. N. (1965). Myxomatosis. Cambridge, UK: Cambridge University Press.Google Scholar
Feoktistov, A. Z., Vasiliev, G. I. & Kraminsky, N. N. (1968). Passage of the virus of tick-borne encephalitis via fleas Xenopsylla cheopis Roths. In Problems of Epidemiology and Epizootology of Paricularly Dangerous Diseases, ed. Anonymous. Kyzyl, USSR: The Tuva Anti-Plague Station, pp. 317–320.Google Scholar
Ferkin, M. H., Sorokin, E. S. & Johnston, R. E. (1996). Self-grooming as a sexually dimorphic communicative behaviour in meadow voles, Microtus pennsylvanicus. Animal Behaviour, 51, 801–810.CrossRefGoogle Scholar
Ferrari, N., Cattadori, I. M., Nespereira, J., Rizzoli, A. & Hudson, P. J. (2004). The role of host sex in parasite dynamics: field experiments on the yellow-necked mouse Apodemus flavicollis. Ecology Letters, 7, 88–94.CrossRefGoogle Scholar
Ferrari, N., Rosá, R., Pugliese, A. & Hudson, P. J. (2007). The role of sex in parasite dynamics: model simulations on transmission of Heligmosomoides polygyrus in populations of yellow-necked mice, Apodemus flavicollis. International Journal for Parasitology, 37, 341–349.CrossRefGoogle ScholarPubMed
Fichet-Calvet, E., Jomaa, I., Ismail, Ben R. & Ashford, R. W. (2000). Pattern of infection of haemoparasites in the fat sand rat, Psammomys obesus, in Tunisia and effect on the host. Annals of Tropical Medicine and Parasitology, 94, 55–68.CrossRefGoogle Scholar
Fieberg, J. & Ellner, S. P. (2000). When is it meaningful to estimate an extinction probability?Ecology, 81, 2040–2047.CrossRefGoogle Scholar
Fielden, L. J., Rechav, Y. & Bryson, N. R. (1992). Acquired immunity to larvae of Amblyomma marmoreum and A. hebraeum by tortoises, guinea-pigs and guinea-fowl. Medical and Veterinary Entomology, 6, 251–254.CrossRefGoogle Scholar
Fielden, L. J., Jones, R. M., Goldberg, M. & Rechav, Y. (1999). Feeding and respiratory gas exchange in the American dog tick, Dermacentor variabilis. Journal of Insect Physiology, 45, 297–304.CrossRefGoogle ScholarPubMed
Fielden, L. J., Krasnov, B. R. & Khokhlova, I. S. (2001). Respiratory gas exchange in the flea Xenopsylla conformis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 38, 735–739.CrossRefGoogle Scholar
Fielden, L. J., Krasnov, B. R., Still, K. & Khokhlova, I. S. (2002). Water balance in two species of desert fleas, Xenopsylla ramesis and X. conformis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 39, 875–881.CrossRefGoogle Scholar
Fielden, L. J., Krasnov, B. R., Khokhlova, I. S. & Arakelyan, M. S. (2004). Respiratory gas exchange in the desert flea Xenopsylla ramesis (Siphonaptera: Pulicidae): response to temperature and blood-feeding. Comparative Biochemistry and Physiology A, 137, 557–565.CrossRefGoogle ScholarPubMed
Filimonova, S. A. (1986). Changes in the ultra-structure of the intestinal epithelium of Xenopsylla cheopis (Siphonaptera) after emerging from cocoons and beginning of feeding. Parazitologiya, 20, 99–105 (in Russian).Google Scholar
Filimonova, S. A. (1989). A morphologic analysis of digestion in Leptopsylla segnis (Siphonaptera: Leptopsyllidae) fleas. Parazitologiya, 23, 480–488 (in Russian).Google ScholarPubMed
Fiorello, C. V., Robbins, R. G., Maffei, L. & Wade, S. E. (2006). Parasites of free-ranging small canids and felids in the Bolivian Chaco. Journal of Zoo and Wildlife Medicine, 37, 130–134.CrossRefGoogle ScholarPubMed
Fisher, R. A. (1930). The Genetical Theory of Natural Selection. Oxford, UK: Oxford University Press.CrossRefGoogle Scholar
Fitze, P. S. & Richner, H. (2002). Differential effects of a parasite on ornamental structures based on melanins and carotenoids. Behavioral Ecology, 13, 401–407.CrossRefGoogle Scholar
Fitze, P. S., Clobert, J. & Richner, H. (2004a). Long-term life-history consequences of ectoparasite-modulated growth and development. Ecology, 85, 2018–2026.CrossRefGoogle Scholar
Fitze, P. S., Tschirren, B. & Richner, H. (2004b). Life history and fitness consequences of ectoparasites. Journal of Animal Ecology, 73, 216–226.CrossRefGoogle Scholar
Fleming, T. H., Breitwisch, R. L. & Whitesides, G. W. (1987). Patterns of tropical vertebrate frugivore diversity. Annual Review of Ecology and Systematics, 18, 91–109.CrossRefGoogle Scholar
Flux, J. E. C. (1972). Seasonal and regional abundance of fleas on hares in Kenya. Journal of East African Natural History Society, 29, 1–8.Google Scholar
Folstad, I. & Karter, A. J. (1992). Parasites, bright males, and the immunocompetence handicap. American Naturalist, 139, 603–622.CrossRefGoogle Scholar
Forbes, M. R., Alisauskas, R. T., McLaughlin, J. D. & Cuddington, K. M. (1999). Explaining co-occurrence among helminth species of lesser snow geese (Chen caerulescens) during their winter and spring migration. Oecologia, 120, 613–620.CrossRefGoogle Scholar
Foster, W. A. & Olkowski, W. (1968). Natural invasion of artificial cliff swallow nests by Oeciacus vicarius (Hemiptera: Cimicidae) and Ceratophyllus petrochelidoni (Siphonaptera: Ceratophyllidae). Journal of Medical Entomology, 5, 488–491.CrossRefGoogle Scholar
Fowler, J. A., Cohen, S. & Greenwood, M. T. (1983). Seasonal variation in the infestation of blackbirds by fleas. Bird Study, 30, 240–242.CrossRefGoogle Scholar
Fox, B. J. & Brown, J. H. (1993). Assembly rules for the functional groups in North American desert rodent communities. Oikos, 67, 358–370.CrossRefGoogle Scholar
Fox, B. J. & Luo, J. (1996). Estimating competition coefficients from census data: a re-examination of the regression technique. Oikos, 77, 291–300.CrossRefGoogle Scholar
Fox, I., Fox, R. I. & Bayona, I. G. (1966). Fleas feed on the lizards in the laboratory in Puerto Rico. Journal of Medical Entomology, 2, 395–396.CrossRefGoogle Scholar
Fox, J. W., McGrady-Steed, J. & Petchey, O. L. (2000). Testing for local species saturation with nonindependent regional species pools. Ecology Letters, 3, 198–206.CrossRefGoogle Scholar
Fox, L. R. (1975). Cannibalism in natural populations. Annual Review of Ecology and Systematics, 6, 87–106.CrossRefGoogle Scholar
Fox, L. R. & Morrow, P. A. (1981). Specialization: species property or local phenomenon?Science, 211, 887–893.CrossRefGoogle ScholarPubMed
Franc, M., Choquart, P. & Cadiergues, M. C. (1998). Species of fleas found on dogs in France. Revue de Médecine Vétérinaire, 149, 135–140.Google Scholar
Freeman, R. B. & Madsen, H. (1949). A parasitic flea larva. Nature, 164, 187–188.CrossRefGoogle ScholarPubMed
Fretwell, S. D. & Lucas, H. L. (1970). On territorial behavior and other factors influencing habitat distribution in birds. I. Theoretical development. Acta Biotheoretica, 19, 16–36.CrossRefGoogle Scholar
Frigessi, A., Holden, M., Marshall, C., et al. (2005). A Bayesian model for the population dynamics of two interacting species, with application to great gerbils and fleas in south-eastern Kazakhstan. Biometrics, 61, 231–239.CrossRefGoogle Scholar
Fry, J. D. (1996). The evolution of host specialization: are trade-offs overrated? American Naturalist, 148, S84–S107.
Fuller, G. K. (1974). Observations on flea attachment at low hair densities on man. Journal of Natural History, 8, 207–213.CrossRefGoogle Scholar
Futuyma, D. J. & Moreno, G. (1988). The evolution of ecological specialization. Annual Review of Ecology and Systematics, 19, 207–233.CrossRefGoogle Scholar
Futuyma, D. J. & Slatkin, M. (1983). Coevolution. Sunderland, MA: Sinauer Associates.Google Scholar
Gabbutt, P. D. (1961). The distribution of some small mammals and their associated fleas from central Labrador. Ecology, 42, 518–525.CrossRefGoogle Scholar
Gäde, G. (2002). Sexual dimorphism in the pyrgomorphid grasshopper Phymateus morbillosus: from wing morphometry and flight behaviour to flight physiology and endocrinology. Physiological Entomology, 27, 51–57.CrossRefGoogle Scholar
Gage, K. L. & Kosoy, M. Y. (2005). Natural history of plague: perspectives from more than a century of research. Annual Review of Entomology, 50, 505–528.CrossRefGoogle Scholar
Gage, K. L., Ostfeld, R. S. & Olson, J. G. (1995). Nonviral vector-borne zoonoses associated with mammals in the United States. Journal of Mammalogy, 76, 695–715.CrossRefGoogle Scholar
Galaktionov, K. V. (1996). Life cycles and distribution of seabird helminths in Arctic and subArctic regions. Bulletin of the Scandinavian Society for Parasitology, 6, 31–49.Google Scholar
Galbe, J. & Oliver, J. H. (1992). Immune response of lizards and rodents to larval Ixodes scapularis (Acari, Ixodidae). Journal of Medical Entomology, 29, 774–783.CrossRefGoogle Scholar
Gallivan, G. J. & Horak, I. G. (1997). Body size and habitat as determinants of tick infestations of wild ungulates in South Africa. South African Journal of Wildlife Research, 27, 63–70.Google Scholar
Galun, R. (1975). Research into alternative arthropod control measures against livestock pests (part 1). In Workshop on the Ecology and Control of External Parasites of Economic Importance on Bovines in Latin America, ed. Thompson, K. C.. Cali, Colombia: Centro Internacional de Agricultura Tropical (CIAT), pp. 155–161.Google Scholar
Garland, T., Harvey, P. H. & Ives, A. R. (1992). Procedures for the analysis of comparative data using phylogenetically independent contrasts. American Naturalist, 41, 18–32.Google Scholar
Garland, T., Dickerman, A. W. C., Janis, M. & Jones, J. A. (1993). Phylogenetic analysis of covariance by computer simulation. Systematic Biology, 42, 265–292.CrossRefGoogle Scholar
Gaston, K. J. (1996). Spatial covariance in the species richness of higher taxa. In Aspects of the Genesis and Maintenance of Biological Diversity, ed. Hochberg, M. E., Clobert, J. & Barbault, R.. Oxford, UK: Oxford University Press, pp. 221–242.Google Scholar
Gaston, K. J. (2003). The Structure and Dynamics of Geographic Ranges. Oxford, UK: Oxford University Press.Google Scholar
Gaston, K. J. & Blackburn, T. M. (2000). Pattern and Process in Macroecology. Oxford, UK: Blackwell Science.CrossRefGoogle Scholar
Gaston, K. J. & Blackburn, T. M. (2003). Dispersal and the interspecific abundance–occupancy relationship in British birds. Global Ecology and Biogeography, 12, 373–379.CrossRefGoogle Scholar
Gaston, K. J., Blackburn, T. M. & Loder, N. (1995). Which species are described first? The case of North American butterflies. Biodiversity and Conservation, 4, 119–127.CrossRefGoogle Scholar
Gaston, K. J., Blackburn, T. M. & Lawton, J. H. (1997). Interspecific abundance–range size relationships: an appraisal of mechanisms. Journal of Animal Ecology, 66, 579–601.CrossRefGoogle Scholar
Gauzshtein, D. M., Kunitsky, V. N., Kunitskaya, N. T. & Filimonov, V. I. (1965). On the time spent on the host body in fleas parasitic on the great gerbil. In Proceedings of the 4th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute and Kainar, pp. 66–68 (in Russian).Google Scholar
Gauzshtein, D. M., Kunitsky, V. N., Gubaidullina, V. S., et al. (1967). On the phenology of reproduction in some fleas parasitic on the great gerbil in the southern Balkhash Region. In Proceedings of the 5th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 160–163 (in Russian).Google Scholar
Geigy, R. & Herbig, A. (1949). Die Hypertrophie der Organe beim Weibchen von Tunga penetrans. Acta Tropica, 6, 246–262.Google Scholar
Gerasimova, N. G. (1970). Metamorphosis of fleas Xenopsylla nuttalli Ioff, 1930. In Vectors of Dangerous Diseases and Their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 316–322 (in Russian).Google Scholar
Gerasimova, N. G. (1973). Some reproductive parameters in Xenopsylla skrjabini and X. nuttalli. Problems of Particularly Dangerous Diseases, 29, 117–121 (in Russian).Google Scholar
Gerasimova, N. G., Denisova, N. G., Denisov, P. S., Knyazeva, T. V. & Lavrovsky, A. A. (1977). Species composition and population dynamics of fleas on the pygmy ground squirrel in the stable foci of plague in the Ergeni Upland. Parazitologiya, 11, 446–452 (in Russian).Google Scholar
Gillespie, R. D., Mbow, M. L. & Titus, R. G. (2000). The immunomodulatory factors of bloodfeeding arthropod saliva. Parasite Immunology, 22, 319–331.CrossRefGoogle ScholarPubMed
Gillespie, S. H., Smith, G. L. & Osbourn, A. (2004). Microbe–Vector Interactions in Vector-Borne Diseases. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Giorgi, M. S., Arlettaz, R., Christe, P. & Vogel, P. (2001). The energetic grooming costs imposed by a parasitic mite (Spinturnix myoti) upon its bat host (Myotis myotis). Proceedings of the Royal Society of London B, 268, 2071–2075.CrossRefGoogle Scholar
Gliwicz, J. (1992). Patterns of dispersal in non-cyclic populations of small rodents. In Animal Dispersal: Small Mammals as a Model, ed. Stenseth, N. C. & Lidicker, W. Z.. London: Chapman & Hall, pp. 147–159.CrossRefGoogle Scholar
Goater, C. P. & Ward, P. I. (1992). Negative effects of Rhabdias bufonis (Nematoda) on the growth and survival of toads (Bufo bufo). Oecologia, 89, 161–165.CrossRefGoogle Scholar
Gobel, E. & Krampitz, H. E. (1982). Histologische Untersuchungen zur Gamogonie und Sporogonie von Hepatozoon erhardovae in experimentell infizierten Rattenflöhen (Xenopsylla cheopis). Zeitschrift für Parasitenkunde, 67, 261–271.CrossRefGoogle Scholar
Gong, Y.-L., Li, Z.-L. & Ma, L.-M. (2004). Further research of the bloodsucking activities of the flea Citellophilus tesquorum sungaris. Acta Parasitologica et Medica Entomologica Sinica, 11, 47–49 (in Chinese).Google Scholar
Gong, Z.-D., Xie, B.-Q. & Ling, J.-B. (1996). Ecology and fauna of fleas on Mt. Gaoligong of Yunnan. Zoological Research, 17, 59–67 (in Chinese).Google Scholar
Gong, Z.-D., Duan, X.-D., Feng, X.-G., Wu, X.-Y. & Liu, Q. (1999). The fauna and ecology of fleas in Cangshan Mountain and Erhai Lake Nature Reserve, Dali. Zoological Research, 20, 451–456 (in Chinese).Google Scholar
Gong, Z.-D., Wu, H.-Y., Duan, X.-D., Feng, X.-G. & Yang, G.-R. (2000). Fauna and community ecology of fleas in Lincang region, Yunnan province. Acta Parasitologica et Medica Entomologica Sinica, 7, 160–169 (in Chinese).Google Scholar
Gong, Z.-D., Zheng, D., Wu, H.-Y, et al. (2001). The relationship between the geographical distribution trends of flea species diversity and the important environmental factor in the Hengduan Mountains, Yunnan. Biodiversity Science, 9, 319–328 (in Chinese).Google Scholar
Gong, Z.-D., Wu, H.-Y., Duan, X.-D., et al. (2004). Vertical distribution pattern and fauna characteristics of flea communities in the Mt. Wuliang Nature Reserve, Jingdong, Yunnan. Chinese Journal of Vector Biology and Control, 15, 344–348 (in Chinese).Google Scholar
Gong, Z.-D., Wu, H.-Y., Duan, X.-D., et al. (2005). Species richness and vertical distribution pattern of flea fauna in Hengduan Mountains of western Yunnan, China. Biodiversity Science, 13, 279–289 (in Chinese).CrossRefGoogle Scholar
Gong, Z.-D., Zhang, L.-Y., Duan, X.-D., et al. (2007). Species richness and fauna of fleas along a latitudinal gradient in the Three Parallel Rivers landscape, China. Biodiversity Science, 15, 61–69 (in Chinese).Google Scholar
González, M. T. & Poulin, R. (2005). Spatial and temporal predictability of the parasite community structure of a benthic marine fish along its distributional range. International Journal for Parasitology, 35, 1369–1377.CrossRefGoogle ScholarPubMed
Gotelli, N. J. (2000). Null model analysis of species co-occurrence patterns. Ecology, 81, 2606–2621.CrossRefGoogle Scholar
Gotelli, N. J. & Arnett, A. E. (2000). Biogeographic effects of red fire ant invasion. Ecology Letters, 3, 257–261.CrossRefGoogle Scholar
Gotelli, N. J. & Entsminger, G. L. (2001). Swap and fill algorithms in null model analysis: rethinking the Knight's Tour. Oecologia, 129, 281–291.CrossRefGoogle ScholarPubMed
Gotelli, N. J. & Entsminger, G. L. (2006). EcoSim: Null Models Software for Ecology. Version 7. Jericho, VT: Acquired Intelligence Inc. & Kesey-Bear. Available online at http://garyentsminger.com/ecosim.htm.Google Scholar
Gotelli, N. J. & Graves, G. R. (1996). Null Models in Ecology. Washington, DC: Smithsonian Institution Press.Google Scholar
Gotelli, N. J. & McCabe, D. J. (2002). Species co-occurrence: a meta-analysis of J. M. Diamond's assembly rules model. Ecology, 83, 2091–2096.CrossRefGoogle Scholar
Gotelli, N. J. & Rohde, K. (2002). Co-occurrence of ectoparasites of marine fishes: a null model analysis. Ecology Letters, 5, 86–94.CrossRefGoogle Scholar
Goüy de Bellocq, J., Sarà, M., Casanova, J. C., Feliu, C. & Morand, S. (2003). A comparison of the structure of helminth communities in the woodmouse, Apodemus sylvaticus, on islands of the western Mediterranean and continental Europe. Parasitology Research, 90, 64–70.Google Scholar
Goüy de Bellocq, J., Krasnov, B. R., Khokhlova, I. S., Ghazaryan, L. & Pinshow, B. (2006a). Immunocompetence and flea parasitism in a desert rodent. Functional Ecology, 20, 637–646.CrossRefGoogle Scholar
Goüy de Bellocq, J., Krasnov, B. R., Khokhlova, I. S. & Pinshow, B. (2006b). Temporal dynamics of a T-cell mediated immune response in desert rodents. Comparative Biochemistry and Physiology A, 145, 554–559.CrossRefGoogle Scholar
Gracia, M. J., Lucientes, J., Castillo, J. A., et al. (2000). Pulex irritans infestation in dogs. Veterinary Record, 147, 748–749.Google ScholarPubMed
Grafen, A. (1989). The phylogenetic regression. Philosophical Transactions of the Royal Society of London B, 326, 119–157.CrossRefGoogle ScholarPubMed
Gray, C. A., Gray, P. N. & Pence, D. B. (1989). Influence of social status on the helminth community of late-winter mallards. Canadian Journal of Zoology, 67, 1937–1944.CrossRefGoogle Scholar
Grazhdanov, A. K., Bidashko, F. G., Tanitovky, V. A., et al. (2002). Comparative fecundity of fleas parasitic on Meriones gerbils in the laboratory. Quarantinable and Zoonotic Infections in Kazakhstan, 6, 39–43 (in Russian).Google Scholar
Grebenyuk, R. V. (1951). Sheep Vermipsylleses and their Control. Frunze, USSR: Ylym (in Russian).Google Scholar
Greene, W. K., Carnegie, R. L., Shaw, S. E., Thompson, R. C. A. & Penhale, W. J. (1993). Characterization of allergens of the cat flea, Ctenocephalides felis: detection and frequency of IgE antibodies in canine sera. Parasite Immunology, 15, 69–74.CrossRefGoogle ScholarPubMed
Greenwood, M. T., Clark, F. & Smith, J. S. (1991). Automatic recording of flea activity. Medical and Veterinary Entomology, 5, 93–100.CrossRefGoogle ScholarPubMed
Gregory, R. D., Montgomery, S. S. J. & Montgomery, W. I. (1992). Population biology of Heligmosomoides polygyrus (Nematoda) in the wood mouse. Journal of Animal Ecology, 61, 749–757.CrossRefGoogle Scholar
Gregory, R. D., Keymer, A. E. & Harvey, P. H. (1996). Helminth parasite richness among vertebrates. Biodiversity and Conservation, 5, 985–997.CrossRefGoogle Scholar
Greives, T. J., McGlothlin, J. W., Jawor, J. M., Demas, G. E. & Ketterson, E. D. (2006). Testosterone and innate immune function inversely covary in a wild population of breeding dark-eyed juncos (Junco hyemalis). Functional Ecology, 20, 812–818.CrossRefGoogle Scholar
Grenfell, B. T. (1992). Parasitism and the dynamics of ungulate grazing systems. American Naturalist, 139, 907–929.CrossRefGoogle Scholar
Grenfell, B. T. & Dobson, A. P. (eds.) (1995). Ecology of Infectious Diseases in Natural Populations. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Grenfell, B. T., Dietz, K. & Roberts, M. G. (1995). Modelling the immuno-epidemiology of macroparasites in naturally-fluctuated host populations. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 362–383.CrossRefGoogle Scholar
Griffiths, D. (1999). On investigation local–regional species richness relationships. Journal of Animal Ecology, 68, 1051–1055.CrossRefGoogle Scholar
Grodzinski, W. & Wunder, B. A. (1975). Ecological energetics of small mammals. In Small Mammals: Their Productivity and Population Dynamics, ed. Golley, F. B., Petrusewitz, K. & Ryszkowski, L.. Cambridge, UK: Cambridge University Press, pp. 173–204.Google Scholar
Gromov, V. S., Krasnov, B. R. & Shenbrot, G. I. (2000). Space use in Wagner's gerbil Gerbillus dasyurus in the Negev Highlands, Israel. Acta Theriologica, 45, 175–182.CrossRefGoogle Scholar
Gruen, J. R. & Weissman, S. M. (1997). Evolving views of the major histocompatibility complex. Blood, 90, 4252–4265.Google ScholarPubMed
Gubareva, N. P., Akiev, A. K., Zemelman, B. M. & Abdurakhmanov, G. A. (1976). Effect of some factors on formation of the plague blockage in Ceratophyllus tesquorum and Neopsylla setosa setosa. Parazitologiya, 10, 315–319 (in Russian).Google Scholar
Guégan, J.-F. & Hugueny, B. A. (1994). A nested parasite species subset pattern in tropical fish host as major determinant of parasite infracommunity structure. Oecologia, 100, 184–189.CrossRefGoogle ScholarPubMed
Guégan, J.-F. & Kennedy, C. R. (1996). Parasite richness/sampling effort/host range: the fancy three-piece jigsaw puzzle. Parasitology Today, 12, 367–369.CrossRefGoogle ScholarPubMed
Guégan, J.-F. & Morand, S. (1996). Polyploid hosts: strange attractors for parasites! Oikos, 7, 366–370.CrossRefGoogle Scholar
Guégan, J.-F., Lambert, A., Leveque, C. & Euzet, L. (1992). Can host body size explain the parasite species richness in tropical freshwater fishes?Oecologia, 90, 197–204.CrossRefGoogle ScholarPubMed
Guégan, J.-F., Morand, S. & Poulin, R. (2005). Are there general laws in parasite community ecology? The emergence of spatial parasitology and epidemiology. In Parasitism and Ecosystems, ed. Thomas, F., Guégan, J.-F. & Renaud, F.. Oxford, UK: Oxford University Press, pp. 22–42.CrossRefGoogle Scholar
Guerrero, O. M., Chinchilla, M. & Abrahams, E. (1997). Increasing of Toxoplasma gondii (Coccidia: Sarcocystidae) infections by Trypanosoma lewisi (Kinetoplastida, Trypanosomatidae) in white rats. Revista de Biologia Tropical, 45, 877–882.Google Scholar
Gulland, F. M. D. (1995). The impact of infectious diseases on wild animal populations: a review. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 20–51.CrossRefGoogle Scholar
Guo, T. & Xu, R. (1999). Trophic niche of flea in the southern slope of the Himalaya Mountains. Chinese Journal of Applied Ecology, 10, 67–70 (in Chinese).Google Scholar
Guo, X.-G., Gong, Z.-D., Qian, T.-J., et al. (2000). Flea fauna investigation in some foci of human plague in Yunnan, China. Acta Zootaxonomica Sinica, 25, 291–297 (in Chinese).Google Scholar
Gurevitch, J., Morrow, L. L., Wallace, A. & Walsh, J. S. (1992). A meta-analysis of competition in field experiments. American Naturalist, 140, 539–572.CrossRefGoogle Scholar
Gurtler, R. E., Cohen, J. E., Cecere, M. C. & Chuit, R. (1997). Shifting host choices of the vector of Chagas disease, Triatoma infestans, in relation to the availability of hosts in houses in north-west Argentina. Journal of Applied Ecology, 34, 699–715.CrossRefGoogle Scholar
Gusev, V. M., Petrosyan, E. A., Guseva, A. A., Eigelis, Y. K. & Tchernyavsky, A. M. (1962). Wild birds: carriers of ectoparasites in the Trans-Caucasus. Proceedings of the Azerbaijanian Anti-Plague Station, 3, 177–184 (in Russian).Google Scholar
Guseva, A. A. & Kosminsky, R. B. (1974). Feeding and reproduction of Frontopsylla elata caspica Ioff et Arg., 1934 (Ceratophyllidae, Siphonaptera) in experiments. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 132–134 (in Russian).Google Scholar
Gustafson, C. R., Bickford, A. A., Cooper, G. L. & Charlton, B. R. (1997). Sticktight fleas associated with fowl pox in a backyard chicken flock in California. Avian Diseases, 41, 1006–1009.CrossRefGoogle Scholar
Gwinner, H. & Berger, S. (2005). European starlings: nestling condition, parasites and green nest material during the breeding season. Journal of Ornithology, 146, 365–371.CrossRefGoogle Scholar
Haag-Wackernagel, D. & Spiewak, R. (2004). Human infestation by pigeon fleas (Ceratophyllus columbae) from feral pigeons. Annals of Agricultural and Environmental Medicine, 11, 343–346.Google ScholarPubMed
Haas, G. E. (1965). Comparative suitability of the four murine rodents of Hawaii as hosts for Xenopsylla vexabilis and X. cheopis (Siphonaptera). Journal of Medical Entomology, 2, 75–83.CrossRefGoogle Scholar
Haas, G. E. (1966). A technique for estimating the total number of rodent fleas in cane fields in Hawaii. Journal of Medical Entomology, 2, 392–394.CrossRefGoogle ScholarPubMed
Haas, G. E. (1969). Quantitative relationships between fleas and rodents in a Hawaiian cane field. Pacific Science, 23, 70–82.Google Scholar
Haas, G. E. & Kucera, J. R. (2004). Fleas (Siphonaptera) in nests of voles (Microtus spp.) in montane habitats of three regions of Utah. Western North American Naturalist, 64, 346–352.Google Scholar
Haas, G. E. & Wilson, N. (1985). Rodent fleas (Siphonaptera) in tree cavities of woodpeckers in Alaska, USA. Canadian Field Naturalist, 100, 554–556.Google Scholar
Haas, G. E. & Wilson, N. (1998). Polygenis martinezbaezi (Siphonaptera: Rhopalopsyllidae) reared from a rodent nest found in the Peloncillo Mountains of southwestern New Mexico. Journal of Medical Entomology, 35, 431–432.CrossRefGoogle ScholarPubMed
Haas, G. E., Johnson, L. & Wilson, N. (1980). Siphonaptera from mammals in Alaska, USA. Supplement 2. Southeastern Alaska. Journal of the Entomological Society of British Columbia, 77, 43–46.Google Scholar
Haas, G. E., Rumfelt, T. & Wilson, N. (1981). Fleas (Siphonaptera) from nests of the tree swallow Iridoprocne bicolor and the violet-green swallow Tachycineta thalassina in Alaska, USA. Wasmann Journal of Biology, 39, 37–41.Google Scholar
Haas, G. E., Kucera, J. R., Runck, A. M., Macdonald, S. O. & Cook, J. A. (2005). Mammal fleas (Siphonaptera: Ceratophyllidae) new for Alaska and the southeastern mainland collected during seven years of a field survey of small mammals. Journal of the Entomological Society of British Columbia, 102, 65–75.Google Scholar
Haeselbarth, E., Segerman, J. & Zumpt, F. (1966). The arthropod parasites of vertebrates in Africa south of the Sahara (Ethiopian region). III. Insecta excl. Phthiraptera. Publications of the South African Institute for Medical Research, 13, 1–283.Google Scholar
Hafner, M. S. & Nadler, S. A. (1988). Phylogenetic trees support the coevolution of parasites and their hosts. Nature, 332, 258–259.CrossRefGoogle ScholarPubMed
Hafner, M. S. & Nadler, S. A. (1990). Cospeciation in host–parasite assemblages: comparative analysis of rates of evolution and timing of cospeciation events. Systematic Zoology, 39, 192–204.CrossRefGoogle Scholar
Hafner, M. S. & Page, R. D. M. (1995). Molecular phylogenies and host–parasite cospeciation: gophers and lice as a model system. Philosophical Transactions of the Royal Society of London B, 349, 77–83.CrossRefGoogle ScholarPubMed
Haftorn, S. (1994). The act of tremble-thrusting in tit nests, performance and possible function. Fauna Norvegica C, 17, 55–74.Google Scholar
Haitlinger, R. (1970). Die Flöhe (Siphonaptera) der Kleinsäuger aus den West- und Mittelsudeten. Polskie Pismo Entomologiczne, 40, 749–762.Google Scholar
Haitlinger, R. (1971). Aphanipterofauna drobnych gryzoni i owadozernych Wrocławia. Zootechnika, 30, 9–22.Google Scholar
Haitlinger, R. (1973). The parasitological investigation of small mammals of the Góry Sowie (Middle Sudetes). I. Siphonaptera (Insecta). Polskie Pismo Entomologiczne, 43, 499–519.Google Scholar
Haitlinger, R. (1974). Fleas (Siphonaptera) of small mammals of the Pieniny, Poland. Polskie Pismo Entomologiczne, 44, 765–788.Google Scholar
Haitlinger, R. (1975). The parasitological investigation of small mammals of the Góry Sowie (Middle Sudetes). II. Siphonaptera (Insecta). Polskie Pismo Entomologiczne, 45, 373–396.Google Scholar
Haitlinger, R. (1977). The parasitological investigation of small mammals of the Góry Sowie (Middle Sudetes). VI. Siphonaptera, Anoplura, Acarina. Polskie Pismo Entomologiczne, 47, 429–492.Google Scholar
Haitlinger, R. (1981). Structure of Arthropod communities occurring on Microtus arvalis (Pall.) in various habitats. I. Faunistic differentiation, dominance structure, arthropod infestation intensiveness in relation to habitats and host population dynamics. Polish Ecological Studies, 7, 271–292.Google Scholar
Haitlinger, R. (1989). Arthropods (Acari, Anoplura, Siphonaptera, Coleoptera) of small mammals of the Babia Góra Mts. Acta Zoologica Cracoviensia, 32, 15–56.Google Scholar
Hallas, T. & Bang, P. (1976). Fleas caught on small mammals at seven locations in eastern Denmark. Flora og Fauna, 82, 11–18.Google Scholar
Halliwell, R. E. W. & Longino, S. J. (1985). IgE and IgG antibodies to flea antigen in differing dog populations. Veterinary Immunology and Immunopathology, 8, 215–223.CrossRefGoogle ScholarPubMed
Halliwell, R. E. W. & Schemmer, K. R. (1987). The role of basophils in the immunopathogenesis of hypersensitivity to fleas (Ctenocephalides felis) in dogs. Veterinary Immunology and Immunopathology, 15, 203–213.CrossRefGoogle Scholar
Halvorsen, O. (1985). On the relationship between social status of a host and risk of parasitic infection. Oikos, 47, 71–74.CrossRefGoogle Scholar
Hamilton, W. D. & Zuk, M. (1982). Heritable true fitness and bright birds: a role for parasites?Science, 218, 384–387.CrossRefGoogle Scholar
Hamilton, P. B., Stevens, J. R., Holz, P., et al. (2005). The inadvertent introduction into Australia of Trypanosoma nabiasi, the trypanosome of the European rabbit (Oryctolagus cuniculus), and its potential for biocontrol. Molecular Ecology, 14, 3167–3175.CrossRefGoogle Scholar
Hanski, I. (1982). Communities of bumblebees: testing the core–satellite hypothesis. Annales Zoologici Fennici, 19, 65–73.Google Scholar
Hanski, I. (1998). Metapopulation dynamics. Nature, 396, 41–49.CrossRefGoogle Scholar
Hanski, I., Kouki, J. & Halkka, A. (1993). Three explanations of the positive relationship between distribution and abundance of species. In Species Diversity in Ecological Communities: Historical and Geographical Perspectives, ed. Ricklefs, R. E. & Schluter, D.. Chicago, IL: University of Chicago Press, pp. 108–116.Google Scholar
Hanski, I., Moilanen, A. & Gyllenberg, M. (1996). Minimum viable metapopulation size. American Naturalist, 147, 527–541.CrossRefGoogle Scholar
Harper, G. H., Marchant, A. & Boddington, D. G. (1992). The ecology of the hen flea Ceratophyllus gallinae and the moorhen flea Dasypsyllus gallinulae in nestboxes. Journal of Animal Ecology, 61, 317–327.CrossRefGoogle Scholar
Harrison, J. F. & Roberts, S. P. (2000). Flight respiration and energetics. Annual Review of Physiology, 62, 179–205.CrossRefGoogle ScholarPubMed
Hart, B. L. (1990). Behavioral adaptations to pathogens and parasites: five strategies. Neuroscience and Biobehavioural Reviews, 14, 273–294.CrossRefGoogle ScholarPubMed
Hart, B. L. & Pryor, P. A. (2004). Developmental and hair-coat determinants of grooming behaviour in goats and sheep. Animal Behaviour, 67, 11–19.CrossRefGoogle Scholar
Hart, B. L., Hart, L. A., Mooring, M. S. & Olubayo, R. (1992). Biological basis of grooming behaviour in antelope: the body size, vigilance and habitat principles. Animal Behaviour, 44, 615–631.CrossRefGoogle Scholar
Hartley, S. & Shorrocks, B. (2002). A general framework for the aggregation model of coexistence. Journal of Animal Ecology, 71, 651–662.CrossRefGoogle Scholar
Hartwell, W. V., Quan, S. F., Scott, K. G. & Kartman, L. (1958). Observations on flea transfer between hosts: a mechanism in the spread of bubonic plague. Science, 127, 814.CrossRefGoogle ScholarPubMed
Harvey, P. H. & Pagel, M. D. (1991). The Comparative Method in Evolutionary Biology. Oxford, UK: Oxford University Press.Google Scholar
Hasibender, G. & Dye, C. (1988). Population dynamics of mosquito-borne disease: persistence in a completely heterogenous environment. Theoretical Population Biology, 33, 31–53.CrossRefGoogle Scholar
Hastriter, M. W. (1997). Establishment of the tungid flea, Tunga monositus (Siphonaptera: Pulicidae), in the United States. Great Basin Naturalist, 57, 281–282.Google Scholar
Hastriter, M. W. (2000a). Jordanopsylla becki (Siphonaptera: Ctenophthalmidae), a new species of flea from the Nevada Test Site. Proceedings of the Entomological Society of Washington, 102, 135–141.Google Scholar
Hastriter, M. W. (2000b). Echidnophaga suricatta (Siphonaptera: Pulicidae), a new species of flea from the Northern Cape Province, South Africa. African Zoology, 35, 77–83.CrossRefGoogle Scholar
Hastriter, M. W. (2001a). Five new species and new subgenus of fleas (Siphonaptera: Chimaeropsyllidae, Ctenophthalmidae) from Africa. Proceedings of the Entomological Society of Washington, 103, 832–848.Google Scholar
Hastriter, M. W. (2001b). Fleas (Siphonaptera: Ctenophthalmidae and Rhopalopsyllidae) from Argentina and Chile with two new species from the rock rat, Aconaemys fuscus, in Chile. Annals of Carnegie Museum, 70, 169–178.Google Scholar
Hastriter, M. W. (2004). Revision of the flea genus Jellisonia Traub, 1944 (Siphonaptera: Ceratophyllidae). Annals of Carnegie Museum, 73, 213–238.Google Scholar
Hastriter, M. W. & Eckerlin, R. P. (2003). Jellisonia painteri (Siphonaptera: Ceratophyllidae), a new species of flea from Guatemala. Annals of Carnegie Museum, 72, 215–221.Google Scholar
Hastriter, M. W. & Haas, G. E. (2005). Bionomics and distribution of species of Hystrichopsylla in Arizona and New Mexico, with a description of Hystrichopsylla dippiei oblique, n. ssp. (Siphonaptera: Hystrichopsyllidae). Journal of Vector Ecology, 30, 251–262.Google Scholar
Hastriter, M. W. & Tipton, V. J. (1975). Fleas (Siphonaptera) associated with small mammals of Morocco. Journal of the Egyptian Public Health Association, 50, 79–169.Google Scholar
Hastriter, M. W. & Whiting, M. E. (2002). Macropsylla novaehollandiae (Siphonaptera: Hystrichopsyllidae), a new species of flea from Tasmania. Proceedings of the Entomological Society of Washington, 104, 663–671.Google Scholar
Hastriter, M. W. & Whiting, M. F. (2003). Siphonaptera (fleas). In Encyclopedia of Insects, ed. Resh, V. H. & Carde, R.. Orlando, FL: Elsevier Science, pp. 1039–1045.Google Scholar
Hastriter, M. W., Egoscue, H. J. & Traub, R. (1998). A description of the male of Jordanopsylla allredi Traub and Tipton, 1951, and characterization of the tribes within Anomiopsyllinae (Siphonaptera: Ctenophthalmidae). Proceedings of the Entomological Society of Washington, 100, 141–146.Google Scholar
Hastriter, M. W., Zyzak, M. D., Soto, R., et al. (2002). Fleas (Siphonaptera) from Ancash Department, Peru with the description of a new species, Ectinorus alejoi (Rhopalopsyllidae), and the description of the male of Plocopsylla pallas (Rothschild, 1914) (Sephanocircidae). Annals of Carnegie Museum, 71, 87–106.Google Scholar
Hastriter, M. W., Haas, G. E. & Wilson, N. (2006). New distribution records for Stenoponia americana (Baker) and Stenoponia ponera Traub and Johnson (Siphonaptera: Ctenophthalmidae) with a review of records from the southwestern United States. Zootaxa, 1253, 51–59.Google Scholar
Haukisalmi, V. & Henttonen, H. (1990). The impact of climatic factors and host density on the long-term population dynamics of vole helminths. Oecologia, 83, 309–315.CrossRefGoogle ScholarPubMed
Haukisalmi, V. & Henttonen, H. (1993). Coexistence in helminths of the bank vole Clethrionomys glareolus. I. Patterns of co-occurrence. Journal of Animal Ecology, 62, 221–229.CrossRefGoogle Scholar
Hawlena, H., Abramsky, Z. & Krasnov, B. R. (2005). Age-biased parasitism and density-dependent distribution of fleas (Siphonaptera) on a desert rodent. Oecologia, 146, 200–208.CrossRefGoogle Scholar
Hawlena, H., Krasnov, B. R., Abramsky, Z., et al. (2006a). Flea infestation and energy requirements of rodent hosts: are there general rules?Functional Ecology, 20, 1028–1036.CrossRefGoogle Scholar
Hawlena, H., Khokhlova, I. S., Abramsky, Z. & Krasnov, B. R. (2006b). Age, intensity of infestation by flea parasites and body mass loss in a rodent host. Parasitology, 133, 187–193.CrossRefGoogle Scholar
Hawlena, H., Abramsky, Z. & Krasnov, B. R. (2006c). Ectoparasites and age-dependent survival in a desert rodent. Oecologia, 148, 30–39.CrossRefGoogle Scholar
Hawlena, H., Abramsky, Z., Krasnov, B. R. & Saltz, D. (2007a). Host defence versus intraspecific competition in the regulation of infrapopulations of the flea Xenopsylla conformis on its rodent host Meriones crassus. International Journal for Parasitology, 37, 919–925.CrossRefGoogle Scholar
Hawlena, H., Bashary, D., Abramsky, Z. & Krasnov, B. R. (2007b). Benefits, costs and constraints of anti-parasitic grooming in adult and juvenile rodents. Ethology, 113, 394–402.CrossRefGoogle Scholar
He, J.-H., Liang, Y. & Zhang, H.-Y. (1997). A study on the transmission of plague though seven kinds of fleas in rat type and wild rodent type plague foci in Yunnan. Chinese Journal of Epidemiology, 18, 236–240 (in Chinese).Google ScholarPubMed
Heath, A. W., Arfsten, A., Yamanaka, M., Dryden, M. W. & Dale, B. (1994). Vaccination against the cat flea Ctenocephalides felis felis. Parasite Immunology, 16, 187–191.CrossRefGoogle ScholarPubMed
Hecht, O. (1943). La reacciones da la piel contra las picaduras de insectos como fenómenas alergicos. Revista de Sanidad y Asistencia Social, 8, 945–959.Google Scholar
Heckmann, R., Gansen, B. & Hom, M. (1967). Maternal transfer of immunity to rat coccidiosis: Eimeria nieschulzi (Dieben – 1924). Journal of Protozoology, S14, 35.Google Scholar
Hecnar, S. J. & Closkey, M' R. T. (1997). Patterns of nestedness and species association in a pond-dwelling amphibian fauna. Oikos, 80, 371–381.CrossRefGoogle Scholar
Heeb, P., Werner, I., Richner, H. & Kölliker, M. (1996). Horizontal transmission and reproductive rates of hen fleas in great tit nests. Journal of Animal Ecology, 65, 474–484.CrossRefGoogle Scholar
Heeb, P., Werner, I., Kölliker, M. & Richner, H. (1998). Benefits of induced host responses against an ectoparasite. Proceedings of the Royal Society of London B, 265, 51–56.CrossRefGoogle Scholar
Heeb, P., Werner, I., Mateman, A. C., et al. (1999). Ectoparasite infestation and sex-biased local recruitment of hosts. Nature, 400, 63–65.CrossRefGoogle ScholarPubMed
Heeb, P., Kölliker, M. & Richner, H. (2000). Bird–ectoparasite interactions, nest humidity, and ectoparasite community structure. Ecology, 81, 958–968.Google Scholar
Heino, J., Muotka, T. & Paavola, R. (2003). Determinants of macroinvertebrate diversity in headwater streams: regional and local influences. Journal of Animal Ecology, 72, 425–434.CrossRefGoogle Scholar
Heitman, T. L., Koski, K. G. & Scott, M. E. (2003). Energy deficiency alters behaviours involved in transmission of Heligmosomoides polygyrus (Nematoda) in mice. Canadian Journal of Zoology, 81, 1767–1773.CrossRefGoogle Scholar
Heller-Haupt, A., Kagaruki, L. K. & Varma, M. G. R. (1996). Resistance and cross-resistance in rabbits to adults of three species of African ticks (Acari: Ixodidae). Experimental and Applied Acarology, 20, 155–165.CrossRefGoogle Scholar
Hemmes, R. B., Alvarado, A. & Hart, B. L. (2002). Use of California bay foliage by wood rats for possible fumigation of nest-borne ectoparasites. Behavioral Ecology, 13, 381–385.CrossRefGoogle Scholar
Henry, P. Y., Poulin, B., Rousset, F., Renaud, F. & Thomas, F. (2004). Infestation by the mite Harpirhynchus nidulans in the bearded tit Panurus biarmicus. Bird Study, 51, 34–40.CrossRefGoogle Scholar
Heukelbach, J., Oliveira, F. A. S., Hesse, G. & Feldmeier, H. (2001). Tungiasis: a neglected health problem of poor communities. Tropical Medicine and International Health, 6, 267–272.CrossRefGoogle ScholarPubMed
Heukelbach, J., Wilcke, T. & Feldmeier, H. (2004). Cutaneous larva migrans (creeping eruption) in an urban slum in Brazil. International Journal of Dermatology, 43, 511–515.CrossRefGoogle Scholar
Hillebrand, H. (2005). Regressions of local on regional diversity do not reflect the importance of local interactions or saturation of local diversity. Oikos, 110, 195–198.CrossRefGoogle Scholar
Hillebrand, H. & Blenckner, T. (2002). Regional and local impact on species diversity: from pattern to processes. Oecologia, 132, 479–491.CrossRefGoogle ScholarPubMed
Hinaidy, H. K. (1991). The biology of Dipylidium caninum. Zentralblatt für Veterinärmedizin B, 38, 329–336.Google ScholarPubMed
Hinkle, N. C., Koehler, P. G. & Kern, W. H. (1991). Hematophagous strategies of the cat flea (Siphonaptera: Pulicidae). Florida Entomologist, 74, 377–385.CrossRefGoogle Scholar
Hinkle, N. C., Koehler, P. G. & Patterson, R. S. (1998). Host grooming efficiency for regulation of cat flea (Siphonaptera: Pulicidae) populations. Journal of Medical Entomology, 35, 266–269.CrossRefGoogle ScholarPubMed
Hinnebusch, B. J., Gage, K. L. & Schwan, T. G. (1998). Estimation of vector infectivity rates for plague by means of a standard curve-based competitive polymerase chain reaction method to quantify Yersinia pestis in fleas. American Journal of Tropical Medicine and Hygiene, 58, 562–569.CrossRefGoogle ScholarPubMed
Hinton, H. E. (1958). The phylogeny of the Panorpoid orders. Annual Review of Entomology, 3, 181–206.CrossRefGoogle Scholar
Hoberg, E. P., Brooks, D. R. & Siegel-Causey, D. (1997). Host–parasite co-speciation: history, principles, and prospects. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 212–235.Google Scholar
Holland, C. (1984). Interactions between Moniliformis (Acanthocephala) and Nippostrongylus (Nematoda) in the small intestine of laboratory rats. Parasitology, 88, 303–315.Google ScholarPubMed
Holland, G. P. (1955). Primary and secondary sexual characteristics of some Ceratophyllinae, with notes on the mechanism of copulation (Siphonaptera). Transactions of the Royal Entomological Society of London, 107, 233–248.CrossRefGoogle Scholar
Holland, G. P. (1964). Evolution, classification and host relationships of Siphonaptera. Annual Review of Entomology, 9, 123–146.CrossRefGoogle Scholar
Holland, G. P. (1985). The fleas of Canada, Alaska and Greenland (Siphonaptera). Memoirs of the Entomological Society of Canada, 130, 1–631.Google Scholar
Holmes, J. C. & Price P. W. (1986). Communities of parasites. In Community Ecology: Patterns and Processes, ed. Kikkawa, J. & Anderson, D. J.. Oxford, UK: Blackwell Science, pp. 187–213.Google Scholar
Holmstad, P. R., Hudson, P. J. & Skorping, A. (2005). The influence of a parasite community on the dynamics of a host population: a longitudinal study on willow ptarmigan and their parasites. Oikos, 111, 377–391.CrossRefGoogle Scholar
Holt, R. D., Dobson, A. P., Begon, M., Bowers, R. G. & Schauber, E. M. (2003). Parasite establishment in host communities. Ecology Letters, 6, 837–842.CrossRefGoogle Scholar
Honnay, O., Hermy, M. & Coppin, P. (1999). Nested plant communities in deciduous forest fragments: species relaxation or nested habitats?Oikos, 84, 119–129.CrossRefGoogle Scholar
Hoogland, J. L. & Sherman, P. W. (1976). Advantages and disadvantages of bank swallow (Riparia riparia) coloniality. Ecological Monographs, 46, 33–58.CrossRefGoogle Scholar
Hopkins, G. H. E. (1957). Host associations of Siphonaptera. In 1st Symposium on Host Specificity amongst Parasites of Vertebrates, ed. Baer, J. G.. Neuchâtel, Switzerland: Institut de Zoologie, Université de Neuchâtel, pp. 64–87.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1953). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 1, Tungidae and Pulicidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1956). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 2, Coptopsyllidae, Vermipsyllidae, Stephanocircidae, Ischnopsyllidae, Hypsophthalmidae and Xiphiopsyllidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1962). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 3, Hystrichopsyllidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1966). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 4, Hystrichopsyllidae. London: Trustees of the British Museum.Google Scholar
Hopkins, G. H. E. & Rothschild, M. (1971). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptera) in the British Museum (Natural History), vol. 5, Leptopsyllidae and Ancistropsyllidae. London: Trustees of the British Museum.Google Scholar
Hopla, C. E. (1980). Fleas as vectors of tularemia in Alaska. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 287–300.Google Scholar
Hörak, P., Ots, I., Vellau, H., Spottiswoode, C. & M⊘ller, A. P. (2001). Carotenoid-based plumage coloration reflects hemoparasite infection and local survival in breeding great tits. Oecologia, 126, 166–173.CrossRefGoogle ScholarPubMed
Hsu, M.-H. & Wu, W.-J. (2000). Effects of multiple mating on female reproductive output in the cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 37, 828–834.CrossRefGoogle Scholar
Hsu, M.-H. & Wu, W.-J. (2001). Off-host observations of mating and postmating behaviors in the cat flea (Siphonaptera: Pulicidae). Journal of Medical Entomology, 38, 352–360.CrossRefGoogle Scholar
Hsu, M.-H., Hsu, T.-C. & Wu, W.-J. (2002). Distribution of cat fleas (Siphonaptera: Pulicidae) on the cat. Journal of Medical Entomology, 39, 685–688.CrossRefGoogle ScholarPubMed
Hu, X.-L., He, J.-H., Zhang, H.-Y., Zhao, W.-H. & Liang, Y. (1996). Evaluation on qualities of house rat fleas from laboratory breeding. Endemic Diseases Bulletin, 11, 14–17 (in Chinese).Google Scholar
Hu, X.-L., He, J.-H., Zhang, H.-Y., et al. (1998). Experimental breeding and life history of the flea Ctenophthalmus quadratus. Endemic Diseases Bulletin, 13, 26–28 (in Chinese).Google Scholar
Hu, X.-L., He, J.-H. & Zhang, H.-Y. (2001). Body weight and quantity of blood-sucking of six species of fleas in Yunnan province, China. Chinese Journal of Pest Control, 17, 393–395 (in Chinese).Google Scholar
Hudson, B. W. & Prince, F. M. (1958a). A method for large-scale rearing of the cat flea, Ctenocephalides felis felis (Bouché). Bulletin of World Health Organization, 19, 1126–1129.Google Scholar
Hudson, B. W. & Prince, F. M. (1958b). Culture methods for fleas Pulex irritans (L.) and Pulex simulans Baker. Bulletin of World Health Organization, 19, 1129–1133.Google Scholar
Hudson, B. W., Feingold, B. F. & Kartman, L. (1960a). Allergy to flea bites. I. Experimental induction of flea-bite sensitivity in guinea pigs. Experimental Parasitology, 9, 18–24.CrossRefGoogle Scholar
Hudson, B. W., Feingold, B. F. & Kartman, L. (1960b). Allergy to flea bites. II. Investigations of flea bite sensitivity in humans. Experimental Parasitology, 9, 264–270.CrossRefGoogle Scholar
Hudson, P. J. & Dobson, A. P. (1995). Macroparasites: observed patterns. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 144–176.CrossRefGoogle Scholar
Hudson, P. J. & Dobson, A. P. (1997). Host–parasite processes and demographic consequences. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J. M.. Oxford, UK: Oxford University Press, pp. 128–154.Google Scholar
Hudson, P. J. & Greenman, J. (1998). Competition mediated by parasites: biological and theoretical progress. Trends in Ecology and Evolution, 13, 387–390.CrossRefGoogle ScholarPubMed
Hughes, J. B. (2000). The scale of resource specialization and the distribution and abundance of lycaenid butterflies. Oecologia, 123, 375–383.CrossRefGoogle ScholarPubMed
Hughes, T. P., Baird, A. H., Dinsdale, E. A., et al. (2000). Supply-side ecology works both ways: the link between benthic adults, fecundity, and larval recruits. Ecology, 81, 2241–2249.CrossRefGoogle Scholar
Hughes, V. L. & Randolph, S. E. (2001). Testosterone depresses innate and acquired resistance to ticks in natural rodent hosts: a force for aggregated distributions of parasites. Journal of Parasitology, 87, 49–54.CrossRefGoogle ScholarPubMed
Hugueny, B. A. & Guégan, J.-F. (1997). Community nestedness and the proper way to assess statistical significance by Monte-Carlo tests: some comments on Worthen and Rohde's (1996) paper. Oikos, 80, 572–574.CrossRefGoogle Scholar
Humphreys, N. E. & Grencis, R. K. (2002). Effects of ageing on the immunoregulation of parasitic infection. Infection and Immunity, 70, 5148–5157.CrossRefGoogle ScholarPubMed
Humphries, D. A. (1966). The function of combs in fleas. Entomological Monthly Magazine, 102, 232–236.Google Scholar
Humphries, D. A. (1967a). The mating behaviour of the hen flea Ceratophyllus gallinae (Schrank) (Siphonaptera: Insecta). Animal Behaviour, 15, 82–90.CrossRefGoogle Scholar
Humphries, D. A. (1967b). The action of the male genitalia during the copulation of the hen flea, Ceratophyllus galinnae (Schrank). Proceedings of the Royal Entomological Society of London A, 42, 101–106.CrossRefGoogle Scholar
Humphries, D. A. (1967c). Function of combs in ectoparasites. Nature, 215, 319.CrossRefGoogle Scholar
Humphries, D. A. (1968). The host-finding behavior of the hen flea, Ceratophyllus gallinae (Schrank) (Siphonaptera). Parasitology, 59, 403–414.CrossRefGoogle Scholar
Humphries, D. A. (1969). Behavioral aspects of the ecology of the sand-martin flea Ceratophyllus styx jordani Smit (Siphonaptera). Parasitology, 59, 311–334.CrossRefGoogle Scholar
Hunter, M. D. & Price, P. W. (1992). Playing chutes and ladders: heterogeneity and the relative roles of bottom–up and top–down forces in natural communities. Ecology, 73, 724–732.Google Scholar
Hürka, K. (1963a). Bat fleas (Aphaniptera, Ischnopsyllidae) of Czechoslovakia: contribution to the distribution, morphology, bionomy, ecology and systematics. I. Subgenus Ischnopsyllus Westw. Acta Faunistica Entomologica Musei Nationalis Pragae, 9, 57–120.Google Scholar
Hürka, K. (1963b). Bat fleas (Aphaniptera, Ischnopsyllidae) of Czechoslovakia: contribution to the distribution, morphology, bionomy, ecology and systematics. II. Subgenus Hexactenopsylla Oud., genus Rhinolophopsylla Oud., subgenus Nycteridopsylla Oud., subgenus Dinycteropsylla Ioff. Acta Universitatis Carolinae, Biologica, 3, 1–73.Google Scholar
Ioff, I. G. (1941). Ecology of Fleas in Relevance to their Medical Importance. Pyatygorsk, USSR: Pyatygorsk Publishers (in Russian).Google Scholar
Ioff, I. G. (1949). Aphaniptera of Kyrgyzstan. Ectoparasites, 1, 5–212 (in Russian).Google Scholar
Ioff, I. G. (1950). The alakurt. Materials to Knowledge of Fauna and Flora of the USSR [Materialy k Poznaniju Fauny i Flory SSSR], 2, 4–29 (in Russian).Google Scholar
Ioff, I. G. & Tiflov, V. E. (1954). The Key to Identification of Aphaniptera of the South-East of the USSR. Stavropol, USSR: Stavropol Publishers (in Russian).Google Scholar
Ioff, I. G., Tiflov, V. E., Argyropulo, A. I., et al. (1946). News species of fleas (Aphaniptera). Medical Parasitology [Meditsinskaya Parazitologiya], 15, 85–94 (in Russian).Google Scholar
Ioff, I. G., Mikulin, M. A. & Scalon, O. N. (1965). The Key to Identification of Fleas of the Middle Asia and Kazakhstan. Moscow, USSR: Meditsina (in Russian).Google Scholar
Iqbal, Q. L. (1973). On the presence of mating pheromone in the rat flea Nosopsyllus fasciatus (Bosc.). Pakistan Journal of Zoology, 5, 123–125.Google Scholar
Iqbal, Q. L. (1974). Host-finding behaviour of the rat flea Nosopsyllus fasciatus (Bosc.). Biologia, Lahore, 20, 147–150.Google Scholar
Iqbal, Q. J. & Humphries, D. A. (1970). Temperature as a critical factor in the mating behavior of the rat flea, Nosopsyllus fasciatus (Bosc.). Parasitology, 61, 375–380.CrossRefGoogle Scholar
Iqbal, Q. J. & Humphries, D. A. (1974). The mating behavior of the rat flea Nosopsyllus fasciatus Bosc. Pakistan Journal of Zoology, 6, 163–174.Google Scholar
Iqbal, Q. J. & Humphries, D. A. (1976). Remating in the rat flea Nosopsyllus fasciatus (Bosc.). Pakistan Journal of Zoology, 8, 39–41.Google Scholar
Iqbal, Q. J. & Humphries, D. A. (1983). Feeding behavior of the rat flea Nosopsyllus fasciatus. Pakistan Journal of Zoology, 14, 71–74.Google Scholar
Ives, A. R. (1988a). Aggregation and the coexistence of competitors. Annales Zoologici Fennici, 25, 75–88.Google Scholar
Ives, A. R. (1988b). Covariance, coexistence and the population dynamics of two competitors using a patchy resource. Journal of Theoretical Biology, 133, 345–361.CrossRefGoogle Scholar
Ives, A. R. (1991). Aggregation and coexistence in a carrion fly community. Ecological Monographs, 61, 75–94.CrossRefGoogle Scholar
Iwao, K. & Ohsaki, N. (1996). Inter- and intraspecific interactions among larvae of specialist and generalist parasitoids. Research on Population Ecology, 38, 265–273.CrossRefGoogle Scholar
Izsák, J. & Papp, L. (1995). Application of the quadratic entropy index for diversity studies on drosophilid species assemblages. Environmental and Ecological Statistics, 2, 213–224.CrossRefGoogle Scholar
Jackson, T. P. (2000). Adaptation to living in an open arid environment: lessons from the burrow structure of two South African whistling rats, Parotomys brantsii and P. littledalei. Journal of Arid Environments, 46, 345–355.CrossRefGoogle Scholar
Jaenike, J. (1990). Host specialization in phytophagous insects. Annual Review of Ecology and Systematics, 21, 243–273.CrossRefGoogle Scholar
Jaenike, J. & James, A. C. (1991). Aggregation and the coexistence of mycophagous Drosophila. Journal of Animal Ecology, 60, 913–928.CrossRefGoogle Scholar
Jameson, E. W. (1985). Pleioxenous host-restriction in fleas. Journal of Natural History, 19, 861–876.CrossRefGoogle Scholar
Jameson, E. W. (1999). Host–ectoparasite relationships among North American chipmunks. Acta Theriologica, 44, 225–231.CrossRefGoogle Scholar
Jamieson, B. G. M. (1987). The Ultrastructure and Phylogeny of Insect Spermatozoa. Cambridge, UK: Cambridge University Press.Google Scholar
Janeway, C., Travers, P., Walport, M. & Capra, J. (1999). Immunobiology: The Immune System in Health and Disease. New York: Garland.Google Scholar
Janion, S. M. (1962). Flea infestation of three rodent species: Apodemus agrarius, Apodemus flavicollis and Clethrionomys glareolus at the period of Apodemus agrarius mass occurrence. Bulletin of the Polish Academy of Sciences, Series of Biological Sciences, 10, 361–366.Google Scholar
Janion, S. M. (1968). Certain host–parasite relationships between rodents (Muridae) and fleas (Aphaniptera). Ekologia Polska, 16, 561–606.Google Scholar
Jarrett, W. F. (1975). Cat leukemia and its viruses. Advances in Veterinary Science and Comparative Medicine, 19, 165–193.Google ScholarPubMed
Jell, P. A. & Duncan, P. M. (1986). Invertebrates, mainly insects, from the freshwater, Lower Cretaceous, Koonwarra Fossil Bed (Korumburra Group), South Gippsland, Victoria. Memoir of the Association of Australasian Palaeontologists, 3, 111–205.Google Scholar
Jellison, W. L. (1959). Fleas and disease. Annual Review of Entomology, 4, 389–414.CrossRefGoogle Scholar
Johnsen, T. S. & Zuk, M. (1999). Parasites and tradeoffs in the immune response of female red jungle fowl. Oikos, 86, 487–492.CrossRefGoogle Scholar
Johnson, K. P., Adams, R. J. & Clayton, D. H. (2002). The phylogeny of the louse genus Brueelia does not reflect host phylogeny. Biological Journal of the Linnean Society, 77, 233–247.CrossRefGoogle Scholar
Johnson, K. P., Bush, S. E. & Clayton, D. H. (2005). Correlated evolution of host and parasite body size: tests of Harrison's rule using birds and lice. Evolution, 59, 1744–1753.Google ScholarPubMed
Johnston, C. M. & Brown, S. J. (1985). Xenopsylla cheopis: cellular expression of hypersensitivity to guinea pigs. Experimental Parasitology, 59, 81–89.CrossRefGoogle ScholarPubMed
Johnston, M., Johnston, D. & Richardson, A. (2005). Digestive capabilities reflect the major food sources in three species of talitrid amphipods. Comparative Biochemistry and Physiology B, 140, 251–257.CrossRefGoogle ScholarPubMed
Jokela, J., Schmid-Hempel, P. & Rigby, M. C. (2000). Dr Pangloss restrained by the Red Queen: steps towards a unified defence theory. Oikos, 89, 267–274.CrossRef
Jones, C. J. (1996). Immune responses to fleas, bugs and sucking lice. In The Immunology of Host–Ectoparasitic Arthropod Relationships, ed. Wikel, S. K.. Wallingford, UK: CAB International, pp. 150–174.Google Scholar
Jordan, K. (1945). On the deciduous frontal tubercle of some genera of Siphonaptera. Proceedings of the Royal Entomological Society of London B, 14, 113–116.Google Scholar
Jordan, K. (1962). Notes on the Tunga caecigena (Siphonaptera: Tungidae). Bulletin of the British Museum of Natural History, Entomology, 12, 353–364.Google Scholar
Jordan, K. & Rothschild, N. C. (1915a). On some Siphonaptera collected by W. Ruckbeil in East Turkestan. Ectoparasites, 1, 1–24.Google Scholar
Jordan, K. & Rothschild, N. C. (1915b). Contribution to our knowledge of American Siphonaptera. Ectoparasites, 1, 45–60.Google Scholar
Joseph, S. A. (1974). Incidence of the flea Ancistropsylla nepalensis Lewis, 1968 on the barking deer (Muntiacus muntjak aureus Smith, 1827) in India. Indian Veterinary Journal, 51, 356–358.Google Scholar
Joseph, S. A. & Mani, K. R. (1980). Cervus unicolor niger: the Indian sambar, a new host for Ancistropsylla nepalensis Lewis, 1968. Cheiron, 9, 200–202.Google Scholar
Joy, J. E. & Briscoe, N. J. (1994). Parasitic arthropods of white-footed mice at McClintock Wildlife Station, West Virginia. Journal of the American Mosquito Control Association, 10, 108–111.Google Scholar
Juricova, Z., Halouzka, J. & Hubalek, Z. (2002). Serologic survey for antibodies to Borrelia burgdorferi in rodents and detection of spirochaetes in ticks and fleas in South Moravia (Czech Republic). Biologia, Bratislava, 57, 383–387.Google Scholar
Jurík, M. (1974). Bionomics of fleas in birds' nests in the territory of Czechoslovakia. Acta Scientiarum Naturalium Brno, 8, 1–54.Google Scholar
Jurík, M. (1983a). To the knowledge of ecological conditions affecting the occurrence of specific and non-specific flea species on their hosts (Talpa europaea – Siphonaptera). Biologia, Bratislava, 38, 949–957.Google Scholar
Jurík, M. (1983b). Ceratophyllus vagabundus insularis Rothschild, 1906 and Ceratophyllus rossittensis Dampf, 1913 in Czechoslovakia (Siphonaptera). Folia Parasitologica, 30, 169–173.Google Scholar
Kaal, J. F., Baker, K. & Torgerson, P. R. (2006). Epidemiology of flea infestation of ruminants in Libya. Veterinary Parasitology, 141, 313–318.CrossRefGoogle ScholarPubMed
Kadatskaya, K. P. (1983). Facultative imaginal diapause in fleas Xenopsylla conformis (Siphonaptera). Parazitologiya, 17, 370–374 (in Russian).Google Scholar
Kadatskaya, K. P. & Kadatsky, N. G. (1983). Comparative data on abundance of fleas parasitic on Meriones erythrourus in the western and eastern parts of the Apsheron Peninsula in relation to the plague epizootic in 1976–1978. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 237–238 (in Russian).Google Scholar
Kadatskaya, K. P. & Shirova, L. F. (1983). Seasonal changes of reproduction in fleas Xenopsylla conformis in Azerbaijan. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 238–240 (in Russian).Google Scholar
Kam, M. & Degen, A. A. (1993). Energetics of lactation and growth in the fat sand rat, Psammomys obesus: new perspectives of resource partitioning and the effect of litter size. Journal of Theoretical Biology, 162, 353–369.CrossRefGoogle Scholar
Kam, M. & Degen, A. A. (1997). Energy requirements and the efficiency of utilization of metabolizable energy in free-living animals: evaluation of existing theories and generation of a new model. Journal of Theoretical Biology, 184, 101–104.CrossRefGoogle Scholar
Kam, M., Khokhlova, I. S. & Degen, A. A. (1997). Granivory and plant selection by desert gerbils of different body size. Ecology, 78, 2218–2229.CrossRefGoogle Scholar
Bai, Kamala M. & Prasad, R. S. (1979). Influence of nutrition on maturation of male rat fleas, Xenopsylla cheopis and X. astia. Journal of Medical Entomology, 16, 164–165.CrossRefGoogle Scholar
Kaňuch, P., Krištín, A. & Krištofik, J. (2005). Phenology, diet, and ectoparasites of Leisler's bat (Nyctalus leisleri) in the western Carpathians (Slovakia). Acta Chiropterologica, 7, 249–257.CrossRefGoogle Scholar
Karandina, P. C. & Darskaya, N. F (1974). Observations on the pre-imaginal development in fleas parasitic on ground squirrels: Ceratophyllus (Citellophilus) tesquorum Wagn., 1898. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 143–144 (in Russian).Google Scholar
Kareiva, P. & Wennergren, U. (1995). Connecting landscape patterns to ecosystem and population processes. Nature, 373, 299–302.CrossRefGoogle Scholar
Karlson, R. H., Cornell, H. V. & Hughes, T. P. (2004). Coral communities are regionally enriched along an oceanic biodiversity gradient. Nature, 429, 867–870.CrossRefGoogle ScholarPubMed
Kartman, L., Prince, F. M., Quan, S. F. & Stark, H. E. (1958). New knowledge of the ecology of sylvatic plague. Annals of the New York Academy of Sciences, 70, 668–711.CrossRefGoogle ScholarPubMed
Kavaliers, M. & Colwell, D. (1994). Parasite infection attenuates nonopioid mediated predator-induced analgesia in mice. Physiology and Behavior, 55, 505–510.CrossRefGoogle ScholarPubMed
Kavaliers, M. & Colwell, D. D. (1995). Reduced spatial learning in mice infected with the nematodeHeligmosomoides polygyrus. Parasitology, 110, 591–597.CrossRefGoogle ScholarPubMed
Kavaliers, M., Colwell, D. D. & Choleris, E. (1998). Parasitized female mice display reduced aversive responses to the odours of infected males. Proceedings of the Royal Society of London B, 265, 1111–1118.CrossRefGoogle ScholarPubMed
K͡dra, A. H., Kruszewicz, A. G., Mazgajski, T. D. & Modlińska, E. (1996). The effects of the presence of fleas in nestboxes on fledglings of pied flycatchers and great tits. Acta Parasitologica, 41, 211–213.Google Scholar
Keeling, M. G. & Gilligan, C. A. (2000a). Metapopulation dynamics of bubonic plague. Nature, 407, 903–906.CrossRefGoogle Scholar
Keeling, M. G. & Gilligan, C. A. (2000b). Bubonic plague: a metapopulation model of a zoonosis. Proceedings of the Royal Society of London B, 267, 2219–2230.CrossRefGoogle Scholar
Kehr, J. D., Heukelbach, J., Mehlhorn, H. & Feldmeier, H. (2007). Morbidity assessment in sand flea disease (tungiasis). Parasitology Research, 100, 413–421.CrossRefGoogle Scholar
Kelly, D. W. & Thompson, C. E. (2000). Epidemiology and optimal foraging: modeling the ideal free distribution of insect vectors. Parasitology, 120, 319–327.CrossRefGoogle Scholar
Kelly, D. W., Mustafa, Z. & Dye, C. (1996). Density-dependent feeding success in a field population of the sandfly Lutzomyia longipalpis. Journal of Animal Ecology, 65, 517–527.CrossRefGoogle Scholar
Kennedy, C. R. & Bush, A. O. (1994). The relationship between pattern and scale in parasite communities: a stranger in a strange land. Parasitology, 109, 187–196.CrossRefGoogle Scholar
Kern, W. H. (1993). The autecology of the cat flea (Ctenocephalides felis felis Bouché) and the synecology of the cat flea and its domestic hosts (Felis catus). Unpublished Ph.D. thesis, University of Florida, Gainesville, FL.
Kern, W. H., Koehler, P. G. & Patterson, R. S. (1992). Diel patterns of cat flea (Siphonaptera, Pulicidae) egg and fecal deposition. Journal of Medical Entomology, 29, 203–206.CrossRefGoogle ScholarPubMed
Kern, W. H., Richman, D. L., Koehler, P. G. & Brenner, R. J. (1999). Outdoor survival and development of immature cat fleas (Siphonaptera: Pulicidae) in Florida. Journal of Medical Entomology, 36, 207–211.Google Scholar
Kettle, D. S. (1995). Medical and Veterinary Entomology. Wallingford, UK: CAB International.Google Scholar
Key, B. H. & Kemp, D. H. (1994). Vaccines against arthropods. American Journal of Tropical Medicine and Hygiene, 50, 87–96.CrossRefGoogle Scholar
Khalid, M. L., Morsy, T. A., Shennawy, El S. F., et al. (1992). Studies on flea fauna in El Fayoum Governorate, Egypt. Journal of the Egyptian Society of Parasitology, 22, 783–799.Google ScholarPubMed
Kharlamov, V. P. (1965). Changes in feeding activity and mobility of the flea Xenopsylla cheopis marked with radioactive 32P. Zoologicheskyi Zhurnal, 44, 547–550 (in Russian).Google Scholar
Khokhlova, I. S. & Knyazeva, T. V. (1983). The effect of spatial and social structure of the house mouse populations on flea assemblages. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 165–167 (in Russian).Google Scholar
Khokhlova, I. S., Krasnov, B. R., Shenbrot, G. I. & Degen, A. A. (1994). Seasonal body mass changes and habitat distribution in several rodent species from the Ramon erosion cirque, Negev Highlands, Israel. Zoologicheskyi Zhurnal, 73, 115–121 (in Russian).Google Scholar
Khokhlova, I. S., Krasnov, B. R., Shenbrot, G. I. & Degen, A. A. (2001). Body mass and environment: a study in Negev rodents. Israel Journal of Zoology, 47, 1–14.CrossRefGoogle Scholar
Khokhlova, I. S., Krasnov, B. R., Kam, M., Burdelova, N. V. & Degen, A. A. (2002). Energy cost of ectoparasitism: the flea Xenopsylla ramesis on the desert gerbil Gerbillus dasyurus. Journal of Zoology, 258, 349–354.CrossRefGoogle Scholar
Khokhlova, I. S., Spinu, M., Krasnov, B. R. & Degen, A. A. (2004a). Immune response to fleas in a wild desert rodent: effect of parasite species, parasite burden, sex of host and host parasitological experience. Journal of Experimental Biology, 207, 2725–2733.CrossRefGoogle Scholar
Khokhlova, I. S., Spinu, M., Krasnov, B. R. & Degen, A. A. (2004b). Immune responses to fleas in two rodent species differing in natural prevalence of infestation and diversity of flea assemblages. Parasitology Research, 94, 304–311.CrossRefGoogle Scholar
Khokhlova, I. S., Hovhanyan, A., Krasnov, B. R. & Degen, A. A. (2007). Reproductive success in two species of desert fleas: density-dependence and host effect. Journal of Experimental Biology, 210, 2121–2127.CrossRefGoogle ScholarPubMed
Khrustselevsky, V. P., Sokolova, A. A. & Balabas, N. G. (1971). Materials on the reproduction of Xenopsylla gerbilli in the Moyynkum Desert. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 436–439 (in Russian).Google Scholar
Khudyakov, I. S. (1965). Fleas (Aphaniptera) of the coastal zone of southern Primorie Region. Entomological Review, 44, 117–122 (in Russian).Google Scholar
Kiefer, M., Klimaszewski, S. M. & Krumpál, M. (1982). Zoogeographical regionalization of Mongolia on the basis of flea fauna (Siphonaptera). Polskie Pismo Entomologiczne, 52, 13–29.Google Scholar
Kiefer, M., Krumpál, M., Cendsuren, N., Lobachev, V. S. & Chotolchu, N. (1984). Checklist, distribution and bibliography of Mongolian Siphonaptera. In Erforschung biologischer Ressourcen der mongolischen Volksrepublik, vol. 4, ed. Stubbe, M., Hilbig, W. & Dawaa, N.. Halle, Germany: Wissenschaftliche Beiträge Universität Halle-Wittenberg, pp. 91–123.Google Scholar
Kilpatrick, A. M. & Ives, A. R. (2003). Species interactions can explain Taylor's power law for ecological time series. Nature, 422, 65–68.CrossRefGoogle ScholarPubMed
Kim, K. C. (1985a). Evolution and host associations of Anoplura. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 197–232.Google Scholar
Kim, K. C. (1985b). Evolutionary relationships of parasitic arthropods and mammals. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 3–82.Google Scholar
King, C. M. (1976). The fleas of a population of weasels in Wytham Wood, Oxford. Journal of Zoology, 180, 525–535.CrossRefGoogle Scholar
King, C. M. & Moody, J. E. (1982). The biology of the stoat (Mustela erminea) in the National Parks of New Zealand. VII. Fleas. New Zealand Journal of Zoology, 9, 141–144.CrossRefGoogle Scholar
Kings, R. C. & Teasly, M. (1980). Insect oogenesis: some generalities and their bearing on the ovarian development of fleas. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 337–340.Google Scholar
Kingsolver, J. G. (1987). Mosquito host choice and the epidemiology of malaria. American Naturalist, 130, 811–827.CrossRefGoogle Scholar
Kiriakova, A. N., Koptzev, L. A. & Koptzeva, Z. G. (1970). Annual number of generations of Xenopsylla fleas in the northern Kyzylkum Desert. Parazitologiya, 6, 528–536 (in Russian).Google Scholar
Kirillova, N. Y., Kirillov, A. A. & Ivashkina, V. A. (2006). Ectoparasites of the edible dormouse Glis glis L. of the Samarskaya Luka Peninsula (Russia). Polish Journal of Ecology, 54, 387–390.Google Scholar
Kirk, W. D. J. (1991). The size relationship between insects and their hosts. Ecological Entomology, 16, 351–359.CrossRefGoogle Scholar
Kisielewska, K. (1970). Ecological organization of intestinal helminth groupings in Clethrionomys glareolus (Schreb.) (Rodentia). III. Structure of helminth groupings in C. glareolus populations of various forest biocoenoses in Poland. Acta Parasitologica, 18, 163–176.Google Scholar
Klasing, K. C. (1998). Nutritional modulation of resistance to infectious diseases. Poultry Science, 77, 1119–1125.CrossRefGoogle ScholarPubMed
Klassen, G. J. (1992). Coevolution: a history of the macroevolutionary approach to studying host–parasite associations. Journal of Parasitology, 78, 573–587.CrossRefGoogle ScholarPubMed
Kleiber, M. (1961). The Fire of Life: An Introduction to Animal Energetics. New York: John Wiley.Google Scholar
Klein, J. (1990). Immunology. Oxford, UK: Blackwell Science.
Klein, J. M. (1966). Données écologiques et biologiques sur Synopsyllus fonquerniei Wagner et Roubaud, 1932 (Siphonaptera) puce du rat péridomestique, dans la région de Tananarive. Cahiers ORSTOM, Série Entomologie Médicale et Parasitologie, 4, 3–29.Google Scholar
Klein, J. M., Simonkovich, E., Alonso, J. M. & Baranton, G. (1975). Observations écologiques dans une zone epizootique de peste en Mauritanie. II. Les puces de rongeurs (Insecta, Siphonaptera). Cahiers ORSTOM, Série Entomologie Médicale et Parasitologie, 13, 29–39.Google Scholar
Klein, S. L. & Nelson, R. J. (1998a). Sex and species differences in cell-mediated immune responses in voles. Canadian Journal of Zoology, 76, 1394–1398.CrossRefGoogle Scholar
Klein, S. L. & Nelson, R. J. (1998b). Adaptive immune responses are linked to the mating system of arvicoline rodents. American Naturalist, 151, 59–67.CrossRefGoogle Scholar
Klompen, J. S. H., Black, W. C., Keirans, J. E. & Oliver, J. H. (1996). Evolution of ticks. Annual Review of Entomology, 41, 141–161.CrossRefGoogle ScholarPubMed
Knopf, P. M. & Coghlan, R. L. (1989). Maternal transfer of resistance to Schistosoma mansoni. Journal of Parasitology, 75, 398–404.CrossRefGoogle ScholarPubMed
Knülle, W. (1967). Physiological properties and biological implications of the water vapour sorption mechanism in larvae of the oriental rat flea, Xenopsylla cheopis (Roths). Journal of Insect Physiology, 13, 333–357.CrossRefGoogle Scholar
Koehler, P. G., Leppla, N. C. & Patterson, S. (1990). Circadian rhythm in the cat flea Ctenocephalides felis (Siphonaptera: Pulicidae). In Chronobiology: Its Role in Clinical Medicine, General Biology, and Agriculture, part B, ed. Hayes, D. K., Pauly, J. E. & Reiter, R. J.. New York: Wiley-Liss, pp. 661–665.Google Scholar
Kolpakova, S. A. (1950). Migration of fleas from the burrows of the great gerbil. Ectoparasites, 2, 115–128 (in Russian).Google Scholar
Kondrashkina, K. I. & Dudnikova, A. F. (1962). Oxygen consumption in fleas parasitic on the ground squirrels as a physiological test of their viability. In Particularly Dangerous and Natural Diseases, ed. Anonymous. Moscow, USSR: Medgiz, pp. 63–69 (in Russian).Google Scholar
Kondrashkina, K. I. & Dudnikova, A. F. (1968). Oxygen consumption in fleas parasitic on the Norway rats. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 87–91 (in Russian).Google Scholar
Kondrashkina, K. I. & Gerasimova, N. G. (1971). Dependence of the metabolic rate of two common fleas parasitic on the great gerbil Rhombomys opimus on their physiological conditions. Problems of Particularly Dangerous Diseases, 17, 32–37 (in Russian).Google Scholar
Konkova, K. V. & Timofeeva, A. A. (1970). Studies of fleas (Siphonaptera) of the Sakhalin and Kuril Islands. In Vectors of Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 371–390 (in Russian).Google Scholar
Konnov, N. P., Demchenko, T. A., Anisimov, P. I., Kondrashkina, K. I. & Lukyanova, A. D. (1986). Pathogenic effect of the plague microbe on the flea Xenopsylla cheopis and ultrastructure of the agent at various periods of its stay in the vector. Parazitologiya, 20, 19–22 (in Russian).Google Scholar
Korallo, N. P., Vinarski, M. V., Krasnov, B. R., et al. (2007). Are there general rules governing parasite diversity? Small mammalian hosts and gamasid mite assemblages. Diversity and Distributions, 13, 353–360.CrossRefGoogle Scholar
Korneeva, L. A. & Sadovenko, E. V. (1990). Patterns of development of the fleas of the great gerbils in the laboratory. In Advantages of Medical Entomology and Acarology in the USSR, ed. Medvedev, G. S.. Leningrad, USSR: All-Union Entomological Society and Zoological Institute, Academy of Sciences of the USSR, pp. 12–14 (in Russian).Google Scholar
Korovin, E. P. (1961). Vegetation of Middle Asia and Southern Kazakhstan. Tashkent, USSR: Academy of Sciences of the Uzbek SSR Press (in Russian).Google Scholar
Korpimaki, E., Tolonen, P. & Bennett, G. F. (1995). Blood parasites, sexual selection and reproductive success of European kestrels. Ecoscience, 2, 335–343.CrossRefGoogle Scholar
Koshkin, S. M. (1966). Materials on flea fauna in Sovetskaya Gavan. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 242–248 (in Russian).Google Scholar
Kosminsky, R. B. (1960). Reproduction of fleas parasitic on mice in the field and experimental conditions. In Parasitological Problems: Proceedings of the 3rd Parasitological Conference of the Ukrainian SSR, ed. Anonymous. Kiev, USSR: Insitute of Zoology of the Academy of Sciences of the Ukrainian SSR, pp. 326–328 (in Russian).Google Scholar
Kosminsky, R. B. (1961). Study of bionomics of the fleas parasitic on the house mice. Unpublished Ph.D. thesis, Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, Stavropol, USSR (in Russian).
Kosminsky, R. B. (1962). Characteristics of the life cycles of fleas parasitic on the house mice in the buildings of a steppe settlement in the Stavropol Region. In Problems of Ecology, vol. 4, ed. Anonymous. Kiev, USSR: Kiev University Press, pp. 65–66 (in Russian).Google Scholar
Kosminsky, R. B. (1965). Feeding and reproduction in fleas parasitic on the house mice in the field and under experiment. Zoologicheskyi Zhurnal, 44, 1372–1375 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1974a). On gonotrophic activity of some fleas belonging to the genus Ctenophthalmus Kolenati (Siphonaptera). In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 154–156 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1974b). On biology of Amphipsylla rossica Wagn., 1912 (Ceratophyllidae, Siphonaptera): a flea with all-year-round activity. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 156–158 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1975a). Age-related changes in the imago of Ctenophthalmus wagneri Tifl, 1928 (Ctenophthalmidae, Siphonaptera). Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 44, 96–100 (in Russian).Google Scholar
Kosminsky, R. B. & Guseva, A. A. (1975b). Feeding behavior and reproduction of Ctenophthalmus wagneri Tifl., 1928 (Ctenophthalmidae, Siphonaptera) under experimental conditions. Parazitologiya, 9, 265–270 (in Russian).Google Scholar
Kosminsky, R. B. & Udovitskaya, E. Y. (1975). Upper temperature boundary of life of imago of some species of fleas (Siphonaptera). Proceedings of the Academy of Sciences of the USSR [Doklady Akademii Nauk SSSR], 222, 500–503 (in Russian).Google Scholar
Kosminsky, R. B., Avetisyan, G. A. & Talybov, A. N. (1970). Annual cycle of Ctenophthalmus wladimiri Isajeva-Gurvich, 1948: common flea species of the common vole in the south-east of Trans-Caucasian Upland. In Vectors of Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 79–107 (in Russian).Google Scholar
Kosminsky, R. B., Bryukhanova, L. V., Darskaya, N. F., et al. (1974). Annual cycle of Ceratophyllus (Nosopsyllus) consimilis Wagn., 1898 (Ceratophyllidae, Siphonaptera) in the Stavropol Upland. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 152–154 (in Russian).Google Scholar
Kosoy, M. Y., Regnery, R., Tzianabos, T., et al. (1997). Distribution, diversity, and host specificity of Bartonella in rodents from the southern United States. American Journal of Tropical Medicine and Hygiene, 57, 578–588.CrossRefGoogle Scholar
Kosoy, M. Y., Mandel, E., Green, D., Marston, E. L. & Childs, J. E. (2004). Prospective studies of Bartonella of rodents. I. Demographic and temporal patterns in population dynamics. Vector-Borne and Zoonotic Diseases, 4, 285–95.CrossRefGoogle ScholarPubMed
Kotler, B. P., Brown, J. S. & Subach, A. (1993). Mechanisms of species coexistence of optimal foragers: temporal partitioning by two species of sand dune gerbils. Oikos, 67, 548–556.CrossRefGoogle Scholar
Kotti, B. K. & Kovalevsky, I. (1995). Fleas parasitic on small mammals in the area between the Amur and Bureya Rivers. Zoologicheskyi Zhurnal, 74, 70–76 (in Russian).Google Scholar
Kouki, J. & Hanski, I. (1995). Population aggregation facilitates coexistence of many competing carrion fly species. Oikos, 72, 223–227.CrossRefGoogle Scholar
Kozlovskaya, O. L. (1958). Flea (Aphaniptera) fauna of rodents of the valley of the Ussury River in the Khabarovsk Region. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 109–116 (in Russian).Google Scholar
Kozlovskaya, O. L. & Demidova, A. A. (1958). Data on the ecology of fleas parasitic on the field mouse in the Khabarovsk Region. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 59–64 (in Russian).Google Scholar
Krämer, F. & Mencke, N. (2001). Flea Biology and Control: The Biology of the Cat Flea – Control and Prevention with Imidacloprid in Small Animals. New York: Springer-Verlag.CrossRefGoogle Scholar
Krampitz, H. E. (1964). Über Vorkomen und Verhalten von Haemococcidien der Gattung Hepatozoon Miller, 1908 (Protozoa, Adeloidea) in mittel- und südeuropäischen Säugern. Acta Tropica, 21, 114–154.Google Scholar
Krampitz, H. E. (1981). Development of Hepatozoon erhardovae Krampitz, 1964, (Protozoa: Haemogregarinidae) in experimental mammalian and arthropod hosts. II. Sexual development in fleas and sporozoite indices in xenodiagnosis. Transactions of the Royal Society of Tropical Medicine and Hygiene, 75, 155–157.CrossRefGoogle ScholarPubMed
Krasnov, B. R. & Khokhlova, I. S. (2001). The effect of behavioural interactions on the exchange of flea (Siphonaptera) between two rodent species. Journal of Vector Ecology, 26, 181–190.Google ScholarPubMed
Krasnov, B. R. & Knyazeva, T. V. (1983). Ectoparasite exchange between the midday jird, Meriones meridianus, and the house mouse, Mus musculus, under experiment. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 243–244 (in Russian).Google Scholar
Krasnov, B. R. & Shenbrot, G. I. (2002). Coevolutionary events in history of association of jerboas (Rodentia: Dipodidae) and their flea parasites. Israel Journal of Zoology, 48, 331–350.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S., Degen, A. A. & Rogovin, K. V. (1996a). On the biology of Sundevall's jird (Meriones crassus Sundevall) in Negev Highlands, Israel. Mammalia, 60, 375–391.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Ivanitskaya, E. Y. (1996b). Spatial structure of rodent community in the Ramon erosion cirque, Negev Highlands (Israel). Journal of Arid Environments, 32, 319–327.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Medvedev, S. G., Vatschenok, V. S. & Khokhlova, I. S. (1997). Host–habitat relations as an important determinant of spatial distribution of flea assemblages (Siphonaptera) on rodents in the Negev Desert. Parasitology, 114, 159–173.CrossRefGoogle ScholarPubMed
Krasnov, B. R., Shenbrot, G. I., Medvedev, S. G., Khokhlova, I. S. & Vatschenok, V. S. (1998). Habitat-dependence of a parasite–host relationship: flea assemblages in two gerbil species of the Negev Desert. Journal of Medical Entomology, 35, 303–313.CrossRefGoogle ScholarPubMed
Krasnov, B. R., Hastriter, M., Medvedev, S. G., et al. (1999). Additional records of fleas (Siphonaptera) on wild rodents in the southern part of Israel. Israel Journal of Zoology, 45, 333–340.Google Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2001a). The effect of temperature and humidity on the survival of pre-imaginal stages of two flea species (Siphonaptera: Pulicidae). Journal of Medical Entomology, 38, 629–637.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2001b). Development rates of two Xenopsylla flea species in relation to air temperature and humidity. Medical and Veterinary Entomology, 15, 249–258.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Oguzoglu, I. & Burdelova, N. V. (2002a). Host discrimination by two desert fleas using an odour cue. Animal Behaviour, 64, 33–40.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2002b). The effect of substrate on survival and development of two species of desert fleas (Siphonaptera: Pulicidae). Parasite, 9, 135–142.CrossRefGoogle Scholar
Krasnov, B. R., Burdelova, N. V., Shenbrot, G. I. & Khokhlova, I. S. (2002c). Annual cycles of four flea species (Siphonaptera) in the central Negev Desert. Medical and Veterinary Entomology, 16, 266–276.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Fielden, L. J. & Burdelova, N. V. (2002d). Time to survival under starvation in two flea species (Siphonaptera: Pulicidae) at different air temperatures and relative humidities. Journal of Vector Ecology, 27, 70–81.Google Scholar
Krasnov, B. R., Khokhlova, I. S. & Shenbrot, G. I. (2002e). The effect of host density on ectoparasite distribution: an example with a desert rodent parasitized by fleas. Ecology, 83, 164–175.CrossRefGoogle Scholar
Krasnov, B. R., Burdelov, S. A., Khokhlova, I. S. & Burdelova, N. V. (2003a). Sexual size dimorphism, morphological traits and jump performance in seven species of desert fleas (Siphonaptera). Journal of Zoology, 261, 181–189.CrossRefGoogle Scholar
Krasnov, B. R., Sarfati, M., Arakelyan, M. S., et al. (2003b). Host-specificity and foraging efficiency in blood-sucking parasite: feeding patterns of a flea Parapulex chephrenis on two species of desert rodents. Parasitology Research, 90, 393–399.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S. & Shenbrot, G. I. (2003c). Density-dependent host selection in ectoparasites: an application of isodar theory to fleas parasitizing rodents. Oecologia, 134, 365–373.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Burdelova, N. V., Mirzoyan, N. S. & Degen, A. A. (2004a). Fitness consequences of density-dependent host selection in ectoparasites: testing reproductive patterns predicted by isodar theory in fleas parasitizing rodents. Journal of Animal Ecology, 73, 815–820.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Burdelov, S. A. & Fielden, L. J. (2004b). Metabolic rate and jumping performance in seven species of desert fleas. Journal of Insect Physiology, 50, 149–156.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2004c). Relationships between parasite abundance and the taxonomic distance among a parasite's host species: an example with fleas parasitic on small mammals. International Journal for Parasitology, 34, 1289–1297.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I. & Khokhlova, I. S. (2004d). Sampling fleas: the reliability of host infestation data. Medical and Veterinary Entomology, 18, 232–240.CrossRefGoogle Scholar
Krasnov, B. R., Mouillot, D., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2004e). Geographical variation in host specificity of fleas (Siphonaptera) parasitic on small mammals: the influence of phylogeny and local environmental conditions. Ecography, 27, 787–797.CrossRefGoogle Scholar
Krasnov, B. R., Poulin, R., Shenbrot, G. I., Mouillot, D. & Khokhlova, I. S. (2004f). Ectoparasitic ‘jacks-of-all-trades’: relationship between abundance and host specificity in fleas (Siphonaptera) parasitic on small mammals. American Naturalist, 164, 506–515.Google Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Degen, A. A. (2004g). Flea species richness and parameters of host body, host geography and host ‘milieu’. Journal of Animal Ecology, 73, 1121–1128.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Degen, A. A. (2004h). Relationship between host diversity and parasite diversity: flea assemblages on small mammals. Journal of Biogeography, 31, 1857–1866.CrossRefGoogle Scholar
Krasnov, B. R., Poulin, R., Shenbrot, G. I., Mouillot, D. & Khokhlova, I. S. (2005a). Host specificity and geographic range in haematophagous ectoparasites. Oikos, 108, 449–456.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2005b). Diversification of ectoparasite assemblages and climate: an example with fleas parasitic on small mammals. Global Ecology and Biogeography, 14, 167–175.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S., Hawlena, H., Khokhlova, I. S. & Shenbrot, G. I. (2005c). Sex-biased parasitism, seasonality and sexual size dimorphism in desert rodents. Oecologia, 146, 209–217.CrossRefGoogle Scholar
Krasnov, B. R., Khokhlova, I. S., Arakelyan, M. S. & Degen, A. A. (2005d). Is a starving host tastier? Reproduction in fleas parasitizing food limited rodents. Functional Ecology, 19, 625–631.CrossRefGoogle Scholar
Krasnov, B. R., Burdelova, N. V., Khokhlova, I. S., Shenbrot, G. I. & Degen, A. A. (2005e). Pre-imaginal interspecific competition in two flea species parasitic on the same rodent host. Ecological Entomology, 30, 146–155.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Mouillot, D., Khokhlova, I. S. & Poulin, R. (2005f). What are the factors determining the probability of discovering a flea species (Siphonaptera)?Parasitology Research, 97, 228–237.CrossRefGoogle Scholar
Krasnov, B. R., Mouillot, D., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2005g). Abundance patterns and coexistence processes in communities of fleas parasitic on small mammals. Ecography, 28, 453–464.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S., Khokhlova, I. S., Shenbrot, G. I. & Hawlena, H. (2005h). Abundance and distribution of fleas on desert rodents: linking Taylor's power law to ecological specialization and epidemiology. Parasitology, 131, 825–837.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Miklisova, D. & Morand, S. (2005i). Distribution of fleas (Siphonaptera) among small mammals: mean abundance predicts prevalence via simple epidemiological model. International Journal for Parasitology, 35, 1097–1101.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2005j). Nested pattern in flea assemblages across the host's geographic range. Ecography, 28, 475–484.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Mouillot, D., Khokhlova, I. S. & Poulin, R. (2005k). Spatial variation in species diversity and composition of flea assemblages in small mammalian hosts: geographic distance or faunal similarity?Journal of Biogeography, 32, 633–644.CrossRefGoogle Scholar
Krasnov, B. R., Mouillot, D., Khokhlova, I. S., Shenbrot, G. I. & Poulin, R. (2005l). Covariance in species diversity and facilitation among non-interactive parasite taxa: all against the host. Parasitology, 131, 557–568.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M. & Morand, S. (2006a). Age-dependent flea (Siphonaptera) parasitism in rodents: a host's life history matters. Journal of Parasitology, 92, 242–248.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S., Mouillot, D., et al. (2006b). Resource predictability and host specificity in fleas: the effect of host body mass. Parasitology, 133, 81–88.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Mouillot, D., Khokhlova, I. S. & Poulin, R. (2006c). Ecological characteristics of flea species relate to their suitability as plague vectors. Oecologia, 149, 474–481.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2006d). Is abundance a species attribute? An example with haematophagous ectoparasites. Oecologia, 150, 132–140.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Miklisova, D. & Morand, S. (2006e). Host specificity, parasite community size and the relation between abundance and its variance. Evolutionary Ecology, 20, 75–91.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Khokhlova, I. S., et al. (2006f). Aggregation and species coexistence in fleas parasitic on small mammals. Ecography, 29, 159–168.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S., Hawlena, H. & Degen, A. A. (2006g). Temporal variation in parasite infestation of a host individual: does a parasite-free host remain uninfested permanently?Parasitology Research, 99, 541–545.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M. & Morand, S. (2006h). Are ectoparasite communities structured? Species co-occurrence, temporal variation and null models. Journal of Animal Ecology, 75, 1330–1339.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Khokhlova, I. S., et al. (2006i). Relationships between local and regional species richness in flea communities of small mammalian hosts: saturation and spatial scale. Parasitology Research, 98, 403–413.CrossRefGoogle Scholar
Krasnov, B. R., Morand, S. & Poulin, R. (2006j). Patterns of macroparasite diversity in small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 197–232.CrossRefGoogle Scholar
Krasnov, B. R., Stanko, M., Miklisova, D. & Morand, S. (2006k). Habitat variation in species composition of flea assemblages on small mammals in central Europe. Ecological Research, 21, 460–469.CrossRefGoogle Scholar
Krasnov, B. R., Korine, C., Burdelova, N. V., Khokhlova, I. S. & Pinshow, B. (2007a). Between-host phylogenetic distance and feeding efficiency in haematophagous ectoparasites: rodent fleas and a bat host. Parasitology Research, 101, 365–371.CrossRefGoogle Scholar
Krasnov, B. R., Hovhanyan, A., Khokhlova, I. S. & Degen, A. A. (2007b). Density-dependence and feeding success in haematophagous ectoparasites. Parasitology, 134, 1379–1386.CrossRefGoogle Scholar
Krasnov, B. R., Shenbrot, G. I., Khokhlova, I. S. & Poulin, R. (2007c). Geographic variation in the ‘bottom-up’ control of diversity: fleas and their small mammalian hosts. Global Ecology and Biogeography, 16, 179–186.CrossRefGoogle Scholar
Krebs, C. J. (1994). Ecology: The Experimental Analysis of Distribution and Abundance, 4th edn. New York: HarperCollins.Google Scholar
Krebs, C. J. (1996). Population cycles revisited. Journal of Mammalogy, 77, 8–24.CrossRefGoogle Scholar
Kristan, D. M. (2002). Maternal and direct effects of the intestinal nematode Heligmosomoides polygyrus on offspring growth and susceptibility to infection. Journal of Experimental Biology, 205, 3967–3977.Google ScholarPubMed
Kristensen, N. P. (1975). The phylogeny of hexapod ‘orders’: a critical review of recent accounts. Zeitschrift für zoologische Systematik und Evolutionsforschung, 13, 1–44.CrossRefGoogle Scholar
Kristensen, N. P. (1981). Phylogeny of insect orders. Annual Review of Entomology, 26, 135–157.CrossRefGoogle Scholar
Krivokhatsky, V. A. (1984). Seasonal changes in the distribution of fleas in burrow passages as indicator of their migration activity. Parazitologiya, 18, 150–153 (in Russian).Google Scholar
Krylov, D. G. (1986). On fauna and ecology of fleas parasitic on small mammals from the Moscow Region, Russian Federation, USSR. Parazitologiya, 20, 356–363 (in Russian).Google Scholar
Kucheruk, V. V. (1983). Mammal burrows: their structure, topology and use. Fauna and Ecology of Rodents, 15, 5–54 (in Russian).Google Scholar
Kucheruk, V. V., Kulik, I. L. & Dubrovsky, Y. A. (1972). The great gerbil as a desert life form. Fauna and Ecology of Rodents, 11, 5–70 (in Russian).Google Scholar
Kulakova, Z. G. (1962). On the role of fleas in transmission of the tick-borne encephalitis (experimental results). Bulletin of the Moscow Naturalist Society, Series Biology, 67, 144–145 (in Russian).Google Scholar
Kulakova, Z. G. (1964). Feeding of Xenopsylla gerbilli caspica and some other flea species. Ectoparasites, 4, 205–219 (in Russian).Google Scholar
Kumar, R. & Kumar, R. (1996). Cross-resistance to Hyalomma anatolicum anatolicum ticks in rabbits immunized with midgut antigens of Hyalomma dromedarii. Indian Journal of Animal Sciences, 66, 657–661.Google Scholar
Kunitskaya, N. T. (1960). Study of the reproductive organs of female fleas and identification of their physiological age. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 29, 688–701 (in Russian).Google Scholar
Kunitskaya, N. T. (1970). Structure of the ovaries in fleas. Parazitologiya, 4, 444–450 (in Russian).Google Scholar
Kunitskaya, N. T., Gauzshtein, D. M., Kunitsky, V. N., Rodionov, I. A. & Filimonov, V. I. (1965a). Feeding activity of fleas parasitic on the great gerbil in experiments. In Proceedings of the 4th Scientific Conference on Natural Focality and Prophylaxis of Plague, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 135–137 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1965b). On the reproduction of fleas parasitic on the great gerbil. In Proceedings of the 4th Scientific Conference on Natural Focality and Prophylaxis of Plague, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 137–138 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1969). Reproduction and age structure of populations of fleas from genera Coptopsylla and Paradoxopsyllus in the southern Balkhash region. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 74–76 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N., Gauzshtein, D. M., Morozova, I. V. & Savelova, N. M. (1971). Age composition of imago in populations of Xenopsylla gerbilli and Xenopsylla hirtipes in the southern Pri-Balkhashie. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 387–389 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1974). On fecundity of fleas Xenopsylla gerbilli in experiments. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 323–325 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N. & Gauzshtein, D. M. (1977). Phenological age of fleas and an attempt to analyze age composition of natural populations of Xenopsylla gerbilli Wagn. Parazitologiya, 11, 202–210 (in Russian).Google Scholar
Kunitskaya, N. T., Kunitsky, V. N., Gauzshtein, D. M. & Savelova, N. M. (1979). Spatial distribution of the larvae of fleas parasitic on the great gerbil in the host's burrow. Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 107–110 (in Russian).Google Scholar
Kunitsky, V. N. (1961). On the environmental conditions of fleas parasitic on gerbils in the southeastern Azerbaijan SSR. Zoologicheskyi Zhurnal, 40, 848–858 (in Russian).Google Scholar
Kunitsky, V. N. (1970). Essay on the comparative ecology of fleas parasitic on gerbils in southwestern Azerbaijan. In Vectors of Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 153–227 (in Russian).Google Scholar
Kunitsky, V. N. & Kunitskaya, N. T. (1962). Fleas of southwestern Azerbaijan. Proceedings of the Azerbaijanian Anti-Plague Station, 3, 156–169 (in Russian).Google Scholar
Kunitsky, V. N., Kunitskaya, N. T., Gauzshtein, D. M. & Podkovyrova, T. S. (1963). Reproduction and development of Ceratophyllus laeviceps Wagn. in natural and experimental conditions. In Proceedings of Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Anonymous. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 126–128 (in Russian).Google Scholar
Kunitsky, V. N., Gauzshtein, D. M. & Kunitskaya, N. T. (1971a). On the effect of the soil moisture of the longevity of fleas under low ambient temperatures. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 395–397 (in Russian).Google Scholar
Kunitsky, V. N., Volkov, V. M., Lelikova, Z. F., et al. (1971b). Changes in the numbers of fleas in the artificially decreased populations of the great gerbil in the northeastern part of the Caspian Lowland. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 392–395 (in Russian).Google Scholar
Kunitsky, V. N., Volkov, V. M., Lelikova, Z. F. & Agunkova, O. S. (1974). On annual number of generations of Xenopsylla skrjabini in the Caspian Lowlands. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 328–3330 (in Russian).Google Scholar
Kuris, A. M., Blaustein, A. R. & Aho, J. J. (1980). Hosts as islands. American Naturalist, 116, 570–586.CrossRefGoogle Scholar
Kusiluka, L. J. M., Kambarage, D. M., Matthewman, R. W., Daborn, C. J. & Harrison, L. J. S. (1995). Prevalence of ectoparasites of goats in Tanzania. Journal of Applied Animal Research, 7, 69–74.CrossRefGoogle Scholar
Kuznetsov, A. A. & Matrosov, A. N. (2003). The use of individual marking of fleas (Siphonaptera) for studies of their dispersal by hosts. Zoologicheskyi Zhurnal, 82, 964–972 (in Russian).Google Scholar
Kuznetsov, A. A., Matrosov, A. N., Nikitin, P. N. & Eigelis, S. Y. (1993). Method of individual marking of fleas and the results of the application of this method for the study of the dispersal of ectoparasites of gerbils in the Volga-Ural Sands. Problems of Particularly Dangerous Diseases, 73, 58–64 (in Russian).Google Scholar
Kuznetsov, A. A., Matrosov, A. N., Chyong, L. T. V. & Dat, D. T. (1999). Movements of the synantropous rats and their fleas in the settlements of the southern Vietnam. Problems of Particularly Dangerous Diseases, 79, 59–65 (in Russian).Google Scholar
Kyriazakis, I., Tolkamp, B. J. & Hutchings, M. R. (1998). Towards a functional explanation for the occurrence of anorexia during parasitic infections. Animal Behaviour, 56, 265–274.CrossRefGoogle ScholarPubMed
Labandeira, C. C. (1997). Insect mouthparts: ascertaining the paleobiology of insect feeding strategies. Annual Review of Ecology and Systematics, 28, 153–193.CrossRefGoogle Scholar
Labunets, N. F. (1967). Zoogeographic characteristics of the western Khangay. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 27, 231–240 (in Russian).Google Scholar
Lafferty, K. D. & Kuris, A. M. (2000). Parasite–host modelling meets reality: adaptive peaks and their ecological attributes. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R., Morand, S. & Skorping, A.. Amsterdam, the Netherlands: Elsevier Science, pp. 9–26.Google Scholar
Lafferty, K. D. & Kuris, A. M. (2002). Trophic strategies, animal diversity and body size. Trends in Ecology and Evolution, 17, 507–513.CrossRefGoogle Scholar
Lambrechts, M. M. & Santos, Dos A. (2000). Aromatic herbs in Corsican blue tit nests: the ‘Potpourri’ hypothesis. Acta Oecologica, 21, 175–178.CrossRefGoogle Scholar
Lang, J. D. (1996). Factors affecting the seasonal abundance of ground squirrel and wood rat fleas (Siphonaptera) in San Diego County, California. Journal of Medical Entomology, 33, 790–804.CrossRefGoogle Scholar
Lapchin, L. & Guillemaud, T. (2005). Asymmetry in host and parasitoid diffuse coevolution: when the Red Queen has to keep a finger in more than one pie. Frontiers in Zoology, 2, 4. doi:10.1186/1742-9994-2-4.CrossRefGoogle Scholar
Lareschi, M. (2006). The relationship of sex and ectoparasite infestation in the water rat Scapteromys aquaticus (Rodentia: Cricetidae) in La Plata, Argentina. Revista de Biologia Tropical, 54, 673–679.CrossRefGoogle Scholar
Larrivee, D. H., Benjamini, E., Feingold, B. F. & Shimizu, M. (1964). Histologic studies of guinea pig skin: different stages of allergic reactivity to flea bites. Experimental Parasitology, 15, 491–502.CrossRefGoogle ScholarPubMed
Larsen, K. S. (1995). Laboratory rearing of the squirrel flea Ceratophyllus sciurorum sciurorum with notes on its biology. Entomologia Experimentalis et Applicata, 76, 241–245.CrossRefGoogle Scholar
Larson, O. R. (1973). North Dakota fleas. IV. Cold tolerance in the bird flea Ceratophillus idius (Jordan and Rothschild). Proceedings of the North Dakota Academy of Sciences, 26, 51–55.Google Scholar
Lauer, D. M. & Sonenshine, D. E. (1978). Bionomics of the squirrel flea, Orchopeas howardi (Siphonaptera: Ceratophyllidae), in laboratory and field colonies of the southern flying squirrel, Glaucomys volans, using radiolabelling technique. Journal of Medical Entomology, 15, 1–10.CrossRefGoogle Scholar
Launay, H. (1989). Facteurs écologiques influençant la répartition et la dynamique des populations de Xenopsylla cunicularis Smit, 1957 (Insecta: Siphonaptera) puce inféodée au lapin de garenne, Oryctolagus cuniculus (L.). Vie et Milieu, 39, 111–120.Google Scholar
Lavoipierre, M. M. J., Radovsky, F. J. & Budwiser, P. D. (1979). The feeding process of a tungid flea, Tunga monositus (Siphonaptera: Tungidae), and its relationship to the host inflammatory and repair response. Journal of Medical Entomology, 15, 187–217.CrossRefGoogle Scholar
Lawrence, W. & Foil, L. D. (2002). The effect of diet upon pupal development and cocoon formation by the cat flea (Siphonaptera: Pulicidae). Journal of Vector Ecology, 27, 39–43.Google Scholar
Lawton, J. H. & Hassell, M. P. (1981). Asymmetrical competition in insects. Nature, 289, 793–795.CrossRefGoogle Scholar
Layene, J. N. (1954). The biology of the red squirrel, Tamiasciurus hudsonicus loquax (Bangs), in central New York. Ecological Monographs, 24, 227–267.CrossRefGoogle Scholar
Layene, J. N. (1963). A study of the parasites of the Florida mouse, Peromyscus floridanus, in relation to host and environmental factors. Tulane Studies in Zoology and Botany, 11, 1–27.Google Scholar
Lee, C. Y., Alexander, P. S., Yang, V. V. C. & Yu, J. Y. L. (2001). Seasonal reproductive activity of male formosan wood mice (Apodemus semotus): relationships to androgen levels. Journal of Mammalogy, 82, 700–708.2.0.CO;2>CrossRefGoogle Scholar
Lee, P. L. M. & Clayton, D. H. (1995). Population biology of swift (Apus apus) ectoparasites in relation to host reproductive success. Ecological Entomology, 20, 43–50.CrossRefGoogle Scholar
Lee, S. E., Johnstone, I. P., Lee, R. P. & Opdebeeck, J. P. (1999). Putative salivary allergens of the cat flea, Ctenocephalides felis. Veterinary Immunology and Immunopathology, 69, 229–237.CrossRefGoogle ScholarPubMed
Lee, W. B. & Houston, D. C. (1993). The effect of diet quality on gut anatomy in British voles (Microtinae). Journal of Comparative Physiology B, 163, 337–339.CrossRefGoogle Scholar
Leeson, H. S. (1936). Further experiments upon the longevity of Xenopsylla cheopis Roths. (Siphonaptera). Parasitology, 28, 403–409.CrossRefGoogle Scholar
Legendre, P., Galzin, R. & Harmelin-Vivien, M. L. (1997). Relating behavior to habitat: solutions to the fourth-corner problem. Ecology, 78, 547–562.Google Scholar
Legendre, P., Desdevises, Y. & Bazin, E. (2002). A statistical test for host–parasite coevolution. Systematic Biology, 51, 217–234.CrossRefGoogle ScholarPubMed
Lehane, M. (2005). The Biology of Blood-Sucking in Insects, 2nd edn. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Lehmann, T. (1992). Reproductive activity of Synosternus cleopatrae (Siphonaptera: Pulicidae) in relation to host factors. Journal of Medical Entomology, 29, 946–952.CrossRefGoogle ScholarPubMed
Lehmann, T. (1993). Ectoparasites: direct impact on host fitness. Parasitology Today, 9, 8–13.CrossRefGoogle ScholarPubMed
Lehmann, T. (1994). Reinfestation analysis to estimate ectoparasite population size, emergence and mortality. Journal of Medical Entomology, 31, 257–264.CrossRefGoogle ScholarPubMed
Leonov, Y. A. (1958). Fleas parasitic on rodents in the southern part of Primorie (Far East). Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 147–152 (in Russian).Google Scholar
Lepage, D. (2006). Avibase: The World Bird Database. Available online at www.bsc-eoc.org/avibase/avibase.jsp?pg=home&lang=EN
Letov, G. S., Emelianova, N. D., Letova, G. I. & Sulimov, A. D. (1966). Rodents and their ectoparasites in the settlements of Tuva. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 270–276 (in Russian).Google Scholar
Letova, G. I., Letov, G. S. & Mamontova, E. V. (1969). Parasitological description of the Mungun-Taigin part of the Altai plague focus. Problems of Particularly Dangerous Infections, 10, 55–60 (in Russian).Google Scholar
Levenbook, L. (1985). Storage proteins. In Comprehensive Insect Biochemistry, Physiology and Pharmacology, ed. Gilbert, L. I. & Kerkut, G.. Oxford, UK: Pergamon Press, pp. 307–346.Google Scholar
Levine, J. M. (1999). Indirect facilitation: evidence and predictions from a riparian community. Ecology, 80, 1762–1769.CrossRefGoogle Scholar
Levine, J. M. & Rees, M. (2002). Coexistence and relative abundance in annual plant assemblages: the roles of competition and colonization. American Naturalist, 160, 452–467.CrossRefGoogle ScholarPubMed
Levins, R. (1968). Evolution in Changing Environments. Princeton, NJ: Princeton University Press.Google Scholar
Lewis, R. E. (1998). Résumé of the Siphonaptera (Insecta) of the world. Journal of Medical Entomology, 35, 377–389.CrossRefGoogle Scholar
Lewis, R. E. & Eckerlin, R. P. (2004). A new species of Hystrichopsylla Taschenberg, 1880 (Siphonaptera: Hystrichopsyllidae) from Guatemala. Proceedings of the Entomological Society of Washington, 106, 757–760.Google Scholar
Lewis, R. E. & Grimaldi, D. (1997). A pulicid flea in Miocene amber from the Dominican Republic (Insecta: Siphonaptera: Pulicidae). American Museum Novitates, 3205, 1–9.Google Scholar
Lewis, R. E. & Haas, G. E. (2001). A review of the North American Catallagia Rothschild, 1915, with the description of a new species (Siphonaptera: Ctenophthalmidae: Neopsyllinae: Phalacropsyllini). Journal of Vector Ecology, 26, 51–69.Google Scholar
Lewis, R. E. & Lewis, J. H. (1990). An annotated checklist of the fleas (Siphonaptera) of the Middle East. Fauna of Saudi Arabia, 11, 251–277.Google Scholar
Lewis, R. E. & Stone, E. (2001). Psittopsylla mexicana, a new genus and species of bird flea from Chihuahua, Mexico (Siphonaptera: Ceratophyllidae: Ceratophyllinae). Journal of the New York Entomological Society, 109, 360–366.CrossRefGoogle Scholar
Lewis, R. E., Lewis, J. H. & Maser, C. (1988). The Fleas of the Pacific Northwest. Corvallis, OR: Oregon State University Press.Google Scholar
Li, B.-G., Zhang, P., Watanabe, K. & Tan, C.-L. (2002). Does allogrooming serve a hygienic function in the Sichuan snub-nosed monkey (Rhinopithecus roxellana)?Acta Zoologica Sinica, 48, 705–715 (in Chinese).Google Scholar
Li, W., Wang, Z.-Y., Wang, C., Zhang, Z.-J. & Ye, R.-Y. (2004). Observations on the life cycle of Neopsylla teratura Rothschild, 1913 in the laboratory. Endemic Diseases Bulletin, 19, 8–9 (in Chinese).Google Scholar
Li, Z.-L. & Ma, L.-M. (1999). Aggregation degree of distribution of two fleas, Citellophilus tesquorum sungaris (Jordan) and Neopsylla bidentatiformis Wagner, on host body. Entomological Knowledge, 36, 89–91 (in Chinese).Google Scholar
Li, Z.-L. & Zhang, Y.-X. (1997). Analysis on the yearly dynamics relation between body flea index and population of Citellus dauricus. Acta Entomologica Sinica, 40, 166–170 (in Chinese).Google Scholar
Li, Z.-L. & Zhang, Y.-X. (1998). The yearly dynamics relationship between burrow nest flea index and population of Citellus dauricus. Acta Entomologica Sinica, 41, 77–81 (in Chinese).Google Scholar
Li, Z.-L., Zhang, W.-R. & Ma, L.-M. (1995). Analysis of the relations among flea index, populations of Meriones unguiculatus and meteorological factors. Acta Entomologica Sinica, 38, 442–447 (in Chinese).Google Scholar
Li, Z.-L., Zhang, W.-R. & Yan, W.-L. (2000). Studies on dynamics of body and burrow nest fleas of Meriones unguiculatus. Acta Entomologica Sinica, 43, 58–63 (in Chinese).Google Scholar
Li, Z.-L., Yang, Y. & Chen, S.-G. (2001a). An analysis on the population dynamics of fleas in the man-made plague focuses of Harbin suburbs. Acta Entomologica Sinica, 44, 507–511 (in Chinese).Google Scholar
Li, Z.-L., Liu, T.-C. & Niu, Y. (2001b). Studies on dynamics of body and burrow fleas of Microtus brandti and succession of their community. Acta Entomologica Sinica, 44, 327–331 (in Chinese).Google Scholar
Li, Z.-L., Yu, G.-J. & Chen, D. (2002). Statespace model between Nosopsyllus laeviceps kuzenkovi and population of Meriones unguiculatus. Acta Entomologica Sinica, 45, 132–133 (in Chinese).Google Scholar
Liao, H.-R. & Lin, D.-H. (1993). Laboratorial observation on some biological characters of two rat fleas in south China. Endemic Diseases Bulletin, 8, 61–64 (in Chinese).Google Scholar
Lighton, J. R. B. (1991). Measurements in insects. In Concise Encyclopedia on Biological and Biomedical Measurement Systems, ed. Payne, C. A.. Oxford, UK: Pergamon Press, pp. 201–208.Google Scholar
Lighton, J. R. B., Fielden, L. J. & Rechav, Y. (1993). Discontinuous ventilation in a non-insect, the tick Amblyomma marmoreum (Acari, Ixodidae): characterization and metabolic modulation. Journal of Experimental Biology, 180, 229–245.Google Scholar
Liker, A., Markus, M., Vazar, A., Zemankovics, E. & Rózsa, L. (2001). Distribution of Carnus hemapterus in a starling colony. Canadian Journal of Zoology, 79, 574–580.CrossRefGoogle Scholar
Linardi, P. M. & Guimarães, L. R. (2000). Sifonápteros do Brasil. São Paulo, Brazil: Museu de Zoologia da Universidade de São Paulo.Google Scholar
Linardi, P. M., Botelho, J. R. & Cunha, H. C. (1985). Ectoparasites of rodents of the urban region of Belo Horizonte, MG. II. Variations of the infestation indices in Rattus norvegicus norvegicus. Memórias do Instituto Oswaldo Cruz, 80, 227–232.CrossRefGoogle Scholar
Linardi, P. M., Gomes, A. F., Botelho, J. R. & Lopes, C. M. L. (1994). Some ectoparasites of commensal rodents from Huambo, Angola. Journal of Medical Parasitology, 31, 754–756.Google ScholarPubMed
Linardi, P. M., DeMaria, M. & Botelho, J. R. (1997). Effects of larval nutrition on the postembryonic development of Ctenocephalides felis felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 34, 494–497.CrossRefGoogle Scholar
Lindén, M. (1991). Divorce in great tits: chance or choice? An experimental approach. American Naturalist, 138, 1039–1048.CrossRefGoogle Scholar
Lindsay, L. R. & Galloway, T. D. (1997). Seasonal activity and temporal separation of four species of fleas (Insecta: Siphonaptera) infesting Richardson's ground squirrels, Spermophilus richardsoni (Rodentia: Sciuridae), in Manitoba, Canada. Canadian Journal of Zoology, 75, 1310–1322.CrossRefGoogle Scholar
Lindsay, L. R. & Galloway, T. D. (1998). Reproductive status of four species of fleas (Insecta: Siphonaptera) on Richardson's ground squirrel (Rodentia: Sciuridae) in Manitoba, Canada. Journal of Medical Entomology, 35, 423–430.CrossRefGoogle Scholar
Linley, J. R., Benton, A. H. & Day, J. F. (1994). Ultrastructure of the eggs of seven flea species (Siphonaptera). Journal of Medical Entomology, 31, 813–827.CrossRefGoogle Scholar
Linsdale, J. M. (1947). The California Ground Squirrel. Berkeley, CA: University of California Press.Google Scholar
Linsdale, J. M. & Davis, B. S. (1956). Taxonomic appraisal and occurrence of fleas at the Hastings Reservation in Central California. University of California Publications in Zoology, 54, 293–370.Google Scholar
Linsdale, J. M. & Tevis, L. P. (1951). The Dusky-Footed Wood Rat. Berkeley, CA: University of California Press.Google Scholar
Litvinova, E. A. (2004). On biology and ecology of Ctenophthalmus congeneroides Wagner, 1929 (Siphonaptera: Hystrichopsyllidae): one of the most abundant fleas of rodents in the Primorie Region. A. I. Kurentsov's Annual Memorial Meetings, 15, 104–107 (in Russian).Google Scholar
Liu, J., Li, S.-J., Amin, O. M. & Zhang, Y.-M. (1993). Blood-feeding of the gerbil flea Nosopsyllus laeviceps kuzenkovi (Yagubyants), vector of plague in Inner Mongolia, China. Medical and Veterinary Entomology, 7, 54–58.Google Scholar
Liu, Z.-Y., Wu, H.-Y., Li, G.-Z., et al. (1986). Fauna Sinica. Insecta. Siphonaptera. Beijing: Science Press (in Chinese).Google Scholar
Lively, C. M. (1989). Adaptation by a parasitic nematode to local populations of its snail host. Evolution, 50, 1663–1671.CrossRefGoogle Scholar
Lloyd, M. (1967). Mean crowding. Journal of Animal Ecology, 36, 1–30.CrossRefGoogle Scholar
Lo, C. M., Morand, S. & Galzon, R. (1998). Parasite diversity/host age and size relationship in three coral-reef fish from French Polynesia. International Journal for Parasitology, 28, 1695–1708.CrossRefGoogle ScholarPubMed
Lochmiller, R. L. & Dabbert, C. B. (1993). Immunocompetence, environmental stress, and the regulation of animal populations. Trends in Comparative Biochemistry and Physiology, 1, 823–855.Google Scholar
Lochmiller, R. L. & Deerenberg, C. (2000). Trade-offs in evolutionary immunology: just what is the cost of immunity?Oikos, 88, 87–98.CrossRefGoogle Scholar
Lochmiller, R. L., Vestey, M. R. & Boren, J. C. (1993). Relationship between protein nutritional status and immunocompetence in northern bobwhite chicks. Auk, 110, 503–510.CrossRefGoogle Scholar
Lombardero, M. J., Ayres, M. P., Hofstetter, R. W., Moser, J. C. & Lepzig, K. D. (2003). Strong indirect interactions of Tarsonemus mites (Acarina: Tarsonemidae) and Dendroctonous frontalis. Oikos, 102, 243–252.CrossRefGoogle Scholar
Lomnicki, A. (1988). Population Ecology of Individuals. Princeton, NJ: Princeton University Press.Google ScholarPubMed
López-Sepulcre, A. & Kokko, H. (2005). Territorial defense, territory size, and population regulation. American Naturalist, 166, 317–329.CrossRefGoogle ScholarPubMed
Lorange, E. A., Race, B. L., Sebbane, F. & Hinnebusch, B. J. (2005). Poor vector competence of fleas and the evolution of hypervirulence in Yersinia pestis. Journal of Infectious Diseases, 191, 1907–1912.CrossRefGoogle ScholarPubMed
Loreau, M. (2000). Are communities saturated? On the relationship of S, J and T diversity. Ecology Letters, 3, 73–76.CrossRefGoogle Scholar
Losos, J. B., Leal, M., Glor, R. E., et al. (2003). Niche lability in the evolution of a Caribbean lizard community. Nature, 424, 542–545.CrossRefGoogle ScholarPubMed
Lott, D. F. (1991). Intraspecific Variation in the Social Systems of Wild Vertebrates. Cambridge, UK: Cambridge University Press.Google Scholar
Louw, J. P., Horak, I. G. & Braack, L. E. O. (1993). Fleas and lice on scrub hares (Lepus saxatilis). Onderstepoort Journal of Veterinary Research, 60, 95–101.Google Scholar
Louw, J. P., Horak, I. G., Horak, M. L. & Braack, L. E. O. (1995). Fleas, lice and mites on scrub hares (Lepus saxatilis) in Northern and Eastern Transvaal and in KwaZulu-Natal, South Africa. Onderstepoort Journal of Veterinary Research, 62, 133–137.Google ScholarPubMed
Lu, L. & Wu, H. (2001). The molecular phylogeny of some species of the bidentatiformis group of the genus Neopsylla based on 16s rRNA gene. Acta Entomologica Sinica, 44, 548–554 (in Chinese).Google Scholar
Lu, L. & Wu, H. (2002). The variation of rDNA ITS2 sequences in nine species of Neopsylla (Siphonaptera: Ctenophthalmidae). Acta Parasitologica et Medica Entomologica Sinica, 9, 106–113 (in Chinese).Google Scholar
Lu, L. & Wu, H. (2005). Morphological phylogeny of Geusibia Jordan, 1932 (Siphonaptera: Leptopsyllidae) and the host–parasite relationships with pikas. Systematic Parasitology, 61, 65–78.Google Scholar
Luchetti, A., Mantovani, B., Pampiglione, S. & Trentini, M. (2005). Molecular characterization of Tunga trimamillata and T. penetrans (Insecta, Siphonaptera, Tungidae): taxonomy and genetic variability. Parasite, 12, 123–129.CrossRefGoogle Scholar
Ludwig, D. (1999). Is it meaningful to estimate a probability of extinction?Ecology, 80, 298–310.CrossRefGoogle Scholar
Lukashevich, E. D. & Mostovsky, M. B. (2003). Haematophagous insects in the fossil record. Paleontologicheskyi Zhurnal, 37, 48–56 (in Russian).Google Scholar
Lundqvist, L. (1988). Reproductive strategies of ectoparasites on small mammals. Canadian Journal of Zoology, 66, 774–781.CrossRefGoogle Scholar
Lundqvist, L. & Brinck-Lindroth, G. (1990). Patterns of coexistence: ectoparasites on small mammals in northern Fennoscandia. Holarctic Ecology, 13, 39–49.Google Scholar
Ma, L.-M. (1983). Distribution of fleas in the hair coat of the host. Acta Entomologica Sinica, 26, 409–412 (in Chinese).Google Scholar
Ma, L.-M. (1988). Abundance of fleas in relation to population fluctuations of their hosts. Acta Entomologica Sinica, 31, 50–54 (in Chinese).Google Scholar
Ma, L.-M. (1989). The distribution of fleas on the host body in relation to temperature and the number of fleas. Acta Entomologica Sinica, 32, 68–73 (in Chinese).Google Scholar
Ma, L.-M. (1990). Observations on longevity of adult Neopsylla bidentatiformis Wagner and Citellophilus tesquorum sungaris (Jordan) under different conditions. Entomological Knowledge, 27, 358–359 (in Chinese).Google Scholar
Ma, L.-M. (1993a). Resistance of fleas Neopsylla bidentatiformis and Citellophilus (Citellophilus) tesquorum sungaris to low temperature. Endemic Diseases Bulletin, 8, 71–73 (in Chinese).Google Scholar
Ma, L.-M. (1993b). The sex ratios of some fleas in north China. Acta Entomologica Sinica, 36, 63–66 (in Chinese).Google Scholar
Ma, L.-M. (1994a). Observation on survival of fleas and their hosts under high temperature in field of northern China. Acta Zoologica Sinica, 40, 100–104 (in Chinese).Google Scholar
Ma, L.-M. (1994b). Laboratory studies on fleas Neopsylla bidentatiformis and Citellophilus tesquorum sungaris attacking and leaving from hosts. Acta Entomologica Sinica, 36, 63–66 (in Chinese).Google Scholar
Ma, L.-M. (1995). Activity of Neopsylla bidentatiformis Wagner and Citellophilus tesquorum sungaris Jordan. Entomological Knowledge, 32, 225–227 (in Chinese).Google Scholar
Ma, L.-M. (1997). The monstrosities of fleas reared in laboratory. Entomological Journal of East China, 6, 107–109 (in Chinese).Google Scholar
Ma, L.-M. (2000). Body length of fleas in relation to some factors, and influence of host nutrition on fleas. Acta Parasitologica et Medica Entomologica Sinica, 7, 235–240 (in Chinese).Google Scholar
Ma, L.-M. (2002). Relationship of hunger tolerance to some environmental and physiologic factors in fleas Neopsylla bidentatiformis and Citellophilus tesquorum sungaris. Acta Parasitologica et Medica Entomologica Sinica, 9, 246–248 (in Chinese).Google Scholar
MacArthur, R. H. (1955). Fluctuations in animal populations, and a measure of community stability. Ecology, 36, 533–536.CrossRefGoogle Scholar
MacArthur, R. H. (1972). Geographical Ecology. New York: Harper & Row.Google Scholar
MacArthur, R. H. & Levins, R. (1964). Competition, habitat selection and character displacement in a patchy environment. Proceedings of the National Academy of Sciences of the USA, 51, 1207–1210.CrossRefGoogle Scholar
MacArthur, R. H. & Levins, R. (1967). The limiting similarity of convergence and divergence of coexisting species. American Naturalist, 101, 377–385.CrossRefGoogle Scholar
MacArthur, R. H. & Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacInnis, A. J. (1976). How parasites find hosts: some thoughts on the inception of host–parasite integration. In Ecological Aspects of Parasitology, ed. Kennedy, C. R.. Amsterdam, the Netherlands: North Holland, pp. 3–20.Google Scholar
Madhavi, R. & Anderson, R. M. (1985). Variability in the susceptibility of the fish host, Poecilia reticulata, to infection with Gyrodactylus bullatarudis (Monogenea). Parasitology, 91, 531–544.CrossRefGoogle Scholar
Main, A. J. (1983). Fleas (Siphonaptera) on small mammals in Connecticut, USA. Journal of Medical Entomology, 20, 33–39.CrossRefGoogle ScholarPubMed
Maizels, R. M., Balic, A., Gomez-Escobar, N., et al. (2004). Helminth parasites: masters of regulation. Immunological Reviews, 201, 89–116.CrossRefGoogle Scholar
Makundi, R. H. & Kilonzo, B. S. (1994). Seasonal dynamics of rodent fleas and its implication on control strategies in Lushoto district, north-eastern Tanzania. Journal of Applied Entomology, 118, 165–171.CrossRefGoogle Scholar
Manhert, V. (1972). Zum Auftreten von Kleinsäuger-Flöhen auf ihren Wirten in Abhängigkeit von Jahreszeit und Höhenstufen. Oecologia, 8, 400–418.CrossRefGoogle Scholar
Mans, B. J., Louw, A. I. & 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, 1695–1705.CrossRefGoogle ScholarPubMed
Mappes, T., Mappes, J. & Kotiaho, J. (1994). Ectoparasites, nest-site choice and breeding success in the pied flycatcher. Oecologia, 98, 147–149.CrossRefGoogle ScholarPubMed
Margalit, Y. & Shulov, A. S. (1972). Effect of temperature on development of prepupa and pupa of the rat flea, Xenopsylla cheopis Rothschild. Journal of Medical Entomology, 9, 117–125.CrossRefGoogle ScholarPubMed
Margolis, L., Esch, G. W., Holmes, J. C., Kuris, A. M. & Schad, G. A. (1982). The use of ecological terms in parasitology (report of an ad hoc committee of the American Society of Parasitologists). Journal of Parasitology, 68, 131–133.CrossRefGoogle Scholar
Marie, J.-L., Fournier, P.-E., Rolain, J.-M., et al. (2006). Molecular detection of Bartonella quintana, B. elizabethae, B. koehlerae, B. doshiae, B. taylorii, and Rickettsia felis in rodent fleas collected in Kabul, Afghanistan. American Journal of Tropical Medicine and Hygiene, 74, 436–439.Google ScholarPubMed
Marshall, A. G. (1981a). The Ecology of Ectoparasitic Insects. London: Academic Press.Google Scholar
Marshall, A. G. (1981b). Sex ratio in ectoparasitic insects. Ecological Entomology, 6, 155–174.CrossRefGoogle Scholar
Martin, T. E., M⊘ller, A. P., Merino, S. & Clobert, J. (2001). Does clutch size evolve in response to parasites and immunocompetence?Proceedings of the National Academy of Sciences of the USA, 98, 2071–2076.CrossRefGoogle ScholarPubMed
Mashek, H., Licznerski, B. & Pincus, S. (1997). Tungiasis in New York. International Journal of Dermatology, 36, 276–278.Google ScholarPubMed
Mashtakov, V. I. (1969). Density dynamics of gerbils and their fleas in different landscape-ecological regions. Problems of Particularly Dangerous Infections, 8, 100–105 (in Russian).Google Scholar
Maslennikova, Z. P., Bibikova, V. A. & Morozova, I. V. (1967). Abundance of Xenopsylla fleas in winter micropopulations in the association with the abundance of the great gerbil. In Proceedings of the 5th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 176–178 (in Russian).Google Scholar
Matejusová, I., Morand, S. & Gelnar, M. (2000). Nestedness in assemblages of gyrodactylids (Monogenea: Gyrodactylidea) parasitising two species of cyprinid: with reference to generalists and specialists. International Journal for Parasitology, 30, 1153–1158.CrossRefGoogle ScholarPubMed
Matthews, J. W. (2004). Effects of site and species characteristics on nested patterns of species composition in sedge meadows. Plant Ecology, 174, 271–278.CrossRefGoogle Scholar
May, R. M. & Anderson, R. M. (1978). Regulation and stability of host–parasite population interactions. II. Destabilizing processes. Journal of Animal Ecology, 47, 455–461.CrossRefGoogle Scholar
May, R. M. & Anderson, R. M. (1979). Population biology of infectious diseases. II. Nature, 280, 455–461.CrossRefGoogle Scholar
Mayevsky, M. P., Bazanova, L. P. & Popkov, A. F. (1999). Winter survival of the causative agent of plaque in the long-tailed ground squirrel in the Tuva natural focus. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 68, 55–58 (in Russian).Google Scholar
Mazgajski, T. D., Kedra, A. H., Modlińska, E. & Samborski, J. (1997). Siphonaptera influence the condition of starling Sturnus vulgaris nestlings. Acta Ornithologica, 32, 185–190.Google Scholar
McCallum, H. & Dobson, A. P. (1995). Detecting disease and parasite threats to endangered species and ecosystems. Trends in Ecology and Evolution, 10, 190–194.CrossRefGoogle ScholarPubMed
McCoy, G. W. & Mitzmain, M. B. (1909). The regional distribution of fleas on rodents. Parasitology, 2, 297–304.CrossRefGoogle Scholar
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, 361–375.CrossRefGoogle Scholar
McKenzie, A. A. (1990). The ruminant dental grooming apparatus. Zoological Journal of the Linnean Society, 99, 117–128.CrossRefGoogle Scholar
McKinney, P. & McDonald, L. C. (2001). Tungiasis. Medical Journal, 2, 3–10.Google Scholar
McTier, T. L., George, J. E. & Bennet, S. N. (1981). Resistance and cross-resistance of guinea pigs to Dermacentor andersoni Stiles, D. variabilis (Say), Amblyomma americanum (Linnaeus) and Ixodes scapularis Say. Journal of Parasitogy, 67, 813–822.CrossRefGoogle ScholarPubMed
Mead-Briggs, A. R. (1964). The reproductive biology of the rabbit flea Spilopsyllus cuniculi (Dale) and the dependence of this species upon the breeding of its host. Journal of Experimental Biology, 41, 371–402.Google Scholar
Mead-Briggs, A. R. & Rudge, A. J. B. (1960). Breeding of the rabbit flea, Spilopsyllus cuniculi (Dale): requirement of a ‘factor’ from a pregnant rabbit for ovarian maturation. Nature, 187, 1136–1137.CrossRefGoogle Scholar
Mead-Briggs, A. R. & Vaughan, J. A. (1969). Some requirements for mating in the rabbit flea, Spilopsyllus cuniculi (Dale). Journal of Experimental Biology, 51, 495–511.Google Scholar
Mead-Briggs, A. R., Vaughan, J. A. & Rennison, B. D. (1975). Seasonal variation in numbers of the rabbit flea on the wild rabbit. Parasitology, 70, 103–118.CrossRefGoogle ScholarPubMed
Mears, S., Clark, F., Greenwood, M. & Larsen, K. S. (2002). Host location, survival and fecundity of the Oriental rat flea Xenopsylla cheopis (Siphonaptera: Pulicidae) in relation to black rat Rattus rattus (Rodentia: Muridae) host sex and age. Bulletin of Entomological Research, 92, 375–384.CrossRefGoogle Scholar
Medvedev, S. G. (1989a). Structure of the head capsule in fleas (Siphonaptera). I. Entomological Review, 68, 1–18.Google Scholar
Medvedev, S. G. (1989b). Ecological characteristics and distribution of fleas of the family Ischnopsyllidae (Siphonaptera). Parasitological Collection, 36, 21–43 (in Russian).Google Scholar
Medvedev, S. G. (1990). Evolution of fleas parasitic on Chiroptera. Parazitologiya, 24, 457–465 (in Russian).Google Scholar
Medvedev, S. G. (1994). Morphological basis of classification of the order Siphonaptera. Entomological Review, 73, 22–43 (in Russian).Google Scholar
Medvedev, S. G. (1996). Geographical distribution of families of fleas (Siphonaptera). Entomological Review, 76, 978–992.Google Scholar
Medvedev, S. G. (1997a). Host–parasite relations in fleas (Siphonaptera). I. Entomological Review, 77, 318–337.Google Scholar
Medvedev, S. G. (1997b). Host–parasite relations in fleas (Siphonaptera). II. Entomological Review, 77, 511–521.Google Scholar
Medvedev, S. G. (1998). Fauna and host–parasite relations of fleas (Siphonaptera) in the Palaearctic. Entomological Review, 78, 292–308.Google Scholar
Medvedev, S. G. (2000a). Fauna and host–parasite associations of fleas (Siphonaptera) in different zoogeographical regions of the world. I. Entomological Review, 80, 409–435.Google Scholar
Medvedev, S. G. (2000b). Fauna and host–parasite associations of fleas (Siphonaptera) in different zoogeographical regions of the world. II. Entomological Review, 80, 640–655.Google Scholar
Medvedev, S. G. (2001). Morphological structure of thoracic and abdominal ctenidia of fleas (Siphonaptera). Parazitologiya, 35, 291–306 (in Russian).Google Scholar
Medvedev, S. G. (2002). Specific features of the distribution and host associations of fleas (Siphonaptera). Entomological Review, 82, 1165–1177.Google Scholar
Medvedev, S. G. (2003a). Morphological adaptations of fleas (Siphonaptera) to parasitism. I. Entomological Review, 82, 40–62 (in Russian).Google Scholar
Medvedev, S. G. (2003b). Morphological adaptations of fleas (Siphonaptera) to parasitism. II. Entomological Review, 82, 820–835 (in Russian).Google Scholar
Medvedev, S. G. (2004). Morphological adaptations of fleas (Siphonaptera) to parasitism. III. Entomological Review, 83, 313–333 (in Russian).Google Scholar
Medvedev, S. G. (2005). An Attempted System Analysis of the Evolution of the Order of Fleas (Siphonaptera), Lectures in Memoriam N. A. Kholodkovsky, No. 57. St Petersburg, Russia: Russian Entomological Society and Zoological Institute of Russian Academy of Sciences (in Russian).Google Scholar
Medvedev, S. G. & Lobanov, A. L. (1999). Information-analytical system of the World fauna of fleas (Siphonaptera): results and prospects. Entomological Review, 79, 654–665.Google Scholar
Medvedev, S. G. & Krasnov, B. R. (2006). Fleas: permanent satellites of small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 161–177.CrossRefGoogle Scholar
Medvedev, S. G., Khabilov, T. K. & Rybin, S. N. (1984). On the biology of the bat fleas (Ischnopsyllidae: Siphonaptera) in the Middle Asia and Kazakhstan. Parazitologiya, 18, 140–149 (in Russian).Google Scholar
Medvedev, S. G., Lobanov, A. L. & Lyangouzov, I. A. (2005). World Database of Fleas (Nov 2004 Version). In Species 2000 and ITIS Catalogue of Life: 2005 Annual Checklist, ed. Bisby, F. A., Ruggiero, M. A., Wilson, K. L., et al.Reading, MA: Species 2000 (CD-ROM).Google Scholar
Medzykhovsky, G. A. (1971). On the method of measurement of the duration of uninterrupted stay of fleas on a host. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 404–406 (in Russian).Google Scholar
Mellanby, K. (1933). The influence of the temperature and humidity on the pupation of Xenopsylla cheopis. Bulletin of Entomological Research, 24, 197–202.CrossRefGoogle Scholar
Méndez, E. (1977). Mammalian–siphonapteran associations, the environment, and biogeography of mammals of southwestern Colombia. Quaestiones Entomologicae, 13, 91–182.Google Scholar
Meng, F.-X., Feng, Y.-L., Chen, J.-Q., Song, X.-P. & Liu, Q.-Y. (2006). The sex ratio and adult eclosion of Xenopsylla cheopis in laboratory. Chinese Journal of Vector Biology and Control, 17, 15–16 (in Chinese).Google Scholar
Menier, K. (2003). Infestation of dairy goats with Pulex irritans. Veterinary Record, 153, 128–128.Google ScholarPubMed
Merila, J. & Allander, K. (1995). Do great tits (Parus major) prefer ectoparasite-free roost sites? – An experiment. Ethology, 99, 53–60.CrossRefGoogle Scholar
Merino, S. & Potti, J. (1996). Weather dependent effects of nest ectoparasites on their bird hosts. Ecography, 19, 107–113.CrossRefGoogle Scholar
Merino, S., Minguez, E. & Belliure, B. (1999). Ectoparasite effects on nestling European storm-petrels. Waterbirds, 22, 297–301.CrossRefGoogle Scholar
Metzger, M. E. & Rust, M. K. (1992). Egg production and emergence of adult cat fleas (Siphonaptera: Pulicidae) exposed to different photoperiods. Journal of Medical Entomology, 33, 651–655.CrossRefGoogle Scholar
Metzger, M. E. & Rust, M. K. (1997). Effect of temperature on cat flea (Siphonaptera: Pulicidae) development and overwintering. Journal of Medical Entomology, 34, 173–178.CrossRefGoogle ScholarPubMed
Meyer, K. F. (1947). The prevention of plague in the light of newer knowledge. Annals of the New York Academy of Sciences, 48, 429–467.CrossRefGoogle Scholar
Michelsen, V. (1997). A revised interpretation of the mouthparts in adult fleas (Insecta, Siphonaptera). Zoologischer Anzeiger, 235, 217–223.Google Scholar
Miklisova, D. & Stanko, M. (1992). Negative binomial distribution as a model for fleas on small rodents. Biologia, Bratislava, 52, 647–652.Google Scholar
Mikulin, M. A. (1956). Data on fleas of the Middle Asia. II. Fauna and some characteristic of geographic distribution of fleas parasitic on the great gerbil in deserts of the southern Trans-Balkhash Desert. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 2, 95–107 (in Russian).Google Scholar
Mikulin, M. A. (1958). Data on fleas of the Middle Asia and Kazakhstan. V. Fleas of the Tarbagatai Mountains. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 4, 227–240 (in Russian).Google Scholar
Mikulin, M. A. (1959a). Data on fleas of the Middle Asia and Kazakhstan. VIII. Fleas of the Akmolinsk Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 5, 237–245 (in Russian).Google Scholar
Mikulin, M. A. (1959b). Data on fleas of the Middle Asia and Kazakhstan. X. Fleas of the eastern Balkhash Desert, Trans-Alakul Desert and Sungorian Gates. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 6, 205–220 (in Russian).Google Scholar
Milazzo, C., Goüy de Bellocq, J., Cagnin, M., et al. (2003). Helminths and ectoparasites of Rattus rattus and Mus musculus from Sicily, Italy. Comparative Parasitology, 70, 199–204.CrossRefGoogle Scholar
Milinski, M. (1990). Parasites and host decision-making. In Parasitism and Host Behaviour, ed. Barnard, C. J. & Behnke, J. M.. London: Taylor & Francis, pp. 95–116.Google Scholar
Milinski, M. & Bakker, T. C. M. (1990). Female sticklebacks use male coloration in mate choice and hence avoid parasitized males. Nature, 344, 330–333.CrossRefGoogle Scholar
Miller, D. H. & Benton, A. H. (1970). Cold tolerance of some adult fleas (Ceratophyllidae: Siphonaptera). Canadian Field Naturalist, 84, 396–397.Google Scholar
Miller, R. A. (1996). The aging immune system: primer and prospectus. Science, 273, 70–74.CrossRefGoogle ScholarPubMed
Mineur, Y. S., Prasol, D. J., Belzung, C. & Crusio, W. E. (2003). Agonistic behavior and unpredictable chronic mild stress in mice. Behavior Genetics, 33, 513–519.CrossRefGoogle Scholar
Mironov, A. N. & Pasyukov, V. V. (1987). Observations on the construction of cocoons by fleas Nosopsyllus fasciatus. Parazitologiya, 21, 10–15 (in Russian).Google Scholar
Mitchell-Jones, A. J., Amori, G., Boganowicz, W., et al. (1999). The Atlas of European Mammals. London: T. & A.D. Poyser.Google Scholar
Miyamoto, K. & Hashimoto, Y. (2000). Outbreak of the sparrow flea bite cases in Hokkaido, Japan. Medical Entomology and Zoology, 51, 111 (in Japanese).CrossRefGoogle Scholar
Moeller, D. A. (2004). Facilitative interactions among plants via shared pollinators. Ecology, 85, 3289–3301.CrossRefGoogle Scholar
Mohr, C. O. (1958). Relation to mean number of fleas to prevalence of infestation on rats. American Journal of Tropical Medicine and Hygiene, 7, 519–522.CrossRefGoogle ScholarPubMed
Moll, A. A. & Leary, O' S. B. (1945). Plague in the Americas: Historical and Quasi-Epidemiological Survey. Washington, DC: The Pan American Sanitary Bureau.Google Scholar
M⊘ller, A. P. (1989). Parasites, predators and nest boxes: facts and artefacts in the nest boxes studies of birds?Oikos, 56, 421–423.CrossRefGoogle Scholar
M⊘ller, A. P. (1991). Ectoparasite loads affect optimal clutch size in swallows. Functional Ecology, 5, 351–359.CrossRefGoogle Scholar
M⊘ller, A. P. (1993). Parasites differentially increase the degree of fluctuating asymmetry in secondary sexual characters. Journal of Evolutionary Biology, 5, 691–699.CrossRefGoogle Scholar
M⊘ller, A. P. (1997). Parasitism and the evolution of host life history. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 105–127.Google Scholar
M⊘ller, A. P. & Lope, F. (1999). Senescence in a short-lived migratory bird: age-dependent morphology, migration, reproduction and parasitism. Journal of Animal Ecology, 68, 163–171.CrossRefGoogle Scholar
M⊘ller, A. P., Christe, P. & Garamszegi, L. Z. (2005). Coevolutionary arms races: increased host immune defense promotes specialization by avian fleas. Journal of Evolutionary Biology, 18, 46–59.CrossRefGoogle Scholar
Moore, J. (2002). Parasites and the Behavior of Animals. New York: Oxford University Press.Google Scholar
Moore, S. L. & Wilson, K. (2002). Parasites as a viability cost of sexual selection in natural populations of mammals. Science, 297, 2015–2018.CrossRefGoogle ScholarPubMed
Mooring, M. S. (1995). The effect of tick challenge on grooming rate by impala. Animal Behaviour, 50, 377–392.CrossRefGoogle Scholar
Mooring, M. S. & Hart, B. L. (1995). Costs of allogrooming in impala: distraction from vigilance. Animal Behaviour, 49, 1414–1416.CrossRefGoogle Scholar
Mooring, M. S. & Hart, B. L. (1997). Self grooming in impala mothers and lambs: testing the body size and tick challenge principles. Animal Behaviour, 53, 925–934.CrossRefGoogle Scholar
Mooring, M. S. & Samuel, W. M. (1999). Premature loss of winter hair in free-ranging moose (Alces alces) infested with winter ticks (Dermacentor albipictus) is correlated with grooming rate. Canadian Journal of Zoology, 77, 148–156.CrossRefGoogle Scholar
Mooring, M. S., Benjamin, J. E., Harte, C. R. & Herzog, N. B. (2000). Testing the interspecific body size principle in ungulates: the smaller they come, the harder they groom. Animal Behaviour, 60, 35–45.CrossRefGoogle ScholarPubMed
Mooring, M. S., Reisig, D. D., Niemeyer, J. M. & Osborne, E. R. (2002). Sexually and developmentally dimorphic grooming: a comparative survey of the Ungulata. Ethology, 108, 911–934.CrossRefGoogle Scholar
Mooring, M. S., Blumstein, D. T. & Stoner, C. J. (2004). The evolution of parasite-defence grooming in ungulates. Biological Journal of the Linnean Society, 81, 17–37.CrossRefGoogle Scholar
Morand, S. (1996). Life-history traits in parasitic nematodes: a comparative approach for the search of invariants. Functional Ecology, 10, 210–218.CrossRefGoogle Scholar
Morand, S. (2000). Wormy world: comparative tests of theoretical hypotheses on parasite species richness. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R., Morand, S. & Skorping, A.. Amsterdam, the Netherlands: Elsevier Science, pp. 63–79.Google Scholar
Morand, S. & Guégan, J.-F. (2000). Distribution and abundance of parasite nematodes: ecological specialization, phylogenetic constraints or simply epidemiology?Oikos, 88, 563–573.CrossRefGoogle Scholar
Morand, S. & Harvey, P. H. (2000). Mammalian metabolism, longevity and parasite species richness. Proceedings of the Royal Society of London B, 267, 1999–2003.CrossRefGoogle ScholarPubMed
Morand, S. & Poulin, R. (1998). Density, body mass and parasite species richness of terrestrial mammals. Evolutionary Ecology, 12, 717–727.CrossRefGoogle Scholar
Morand, S. & Poulin, R. (2000). Nematode parasite species richness and the evolution of spleen size in birds. Canadian Journal of Zoology, 78, 1356–1360.CrossRefGoogle Scholar
Morand, S. & Poulin, R. (2003). Phylogenies, the comparative method and parasite evolutionary ecology. Advances in Parasitology, 54, 281–302.CrossRefGoogle ScholarPubMed
Morand, S., Pointier, J.-P., Borel, G. & Theron, A. (1993). Pairing probability of schistosomes related to their distribution among the host population. Ecology, 74, 2444–2449.CrossRefGoogle Scholar
Morand, S., Poulin, R., Rohde, K. & Hayward, C. (1999). Aggregation and species coexistence of ectoparasites of marine fishes. International Journal for Parasitology, 29, 663–672.CrossRefGoogle ScholarPubMed
Morand, S., Hafner, M. S., Page, R. D. M. & Reed, D. L. (2000). Comparative body size relationships in pocket gophers and their chewing lice. Biological Journal of the Linnean Society, 70, 239–249.CrossRefGoogle Scholar
Morand, S., Goüy de Bellocq, J., Stanko, M. & Miklisova, D. (2004). Is sex-biased ectoparasitism related to sexual size dimorphism in small mammals of central Europe?Parasitology, 129, 505–510.CrossRefGoogle ScholarPubMed
Moret, Y. & Schmid-Hempel, P. (2000). Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 1166–1168.CrossRefGoogle Scholar
Morlan, H. B. (1955). Mammal fleas of Santa Fe County, New Mexico. Texas Reports on Biology and Medicine, 13, 93–125.Google ScholarPubMed
Morozkina, E. A., Lysenko, L. S. & Kafarskaya, D. G. (1970). Materials on the ecology of fleas of the red marmot in the eastern Pamir Mountains. In Vectors of Particularly Dangerous Diseases and their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 337–341 (in Russian).Google Scholar
Morozkina, E. A., Lysenko, L. S. & Kafarskaya, D. G. (1971). Fleas of the red marmot (Marmota caudata) and other animals inhabiting the Gissar Ridge. Problems of Particularly Dangerous Infections, 17, 38–44 (in Russian).Google Scholar
Morozov, Y. A. (1974). The level of flea infestation in the great gerbils of different age. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 337–338 (in Russian).Google Scholar
Morozov, Y. A., Rapoport, L. P. & Kovtun, I. P. (1972). Parasitological contacts between the great gerbils (Rhombomys opimus Licht.) and the yellow ground squirrels (Citellus fulvus Licht.) in the Tchu-Moyynkum Desert. Parazitologiya, 6, 334–337 (in Russian).Google Scholar
Morris, D. W. (1987a). Ecological scale and habitat use. Ecology, 68, 362–369.CrossRefGoogle Scholar
Morris, D. W. (1987b). Spatial scale and the cost of density-dependent habitat selection. Evolutionary Ecology, 1, 379–388.CrossRefGoogle Scholar
Morris, D. W. (1988). Habitat-dependent population regulation and community structure. Evolutionary Ecology, 2, 253–269.CrossRefGoogle Scholar
Morris, D. W. (1990). Temporal variation, habitat selection and community structure. Oikos, 59, 303–312.CrossRefGoogle Scholar
Morris, D. W. (2003). Toward an ecological synthesis: a case for habitat selection. Oecologia, 136, 1–13.CrossRefGoogle ScholarPubMed
Morrison, M. L., Marcot, B. G. & Mannan, R. W. (1992). Wildlife–Habitat Relationships: Concepts and Applications. Madison, WI: University of Wisconsin Press.Google Scholar
Morrone, J. J. & Acosta, R. (2006). A synopsis of the fleas (Insecta: Siphonaptera) parasitizing New World species of Soricidae (Mammalia: Insectivora). Zootaxa, 1354, 1–30.Google Scholar
Morrone, J. J. & Gutiérrez, A. (2004). Do fleas (Insecta: Siphonaptera) parallel their mammal host diversification in the Mexican transition zone?Journal of Biogeography, 32, 1315–1325.CrossRefGoogle Scholar
Morrone, J. J., Acosta, R. & Gutiérrez, A. (2000). Cladistics, biogeography, and host relationships of the flea subgenus Ctenophthalmus (Alloctenus), with the description of a new Mexican species (Siphonaptera: Ctenophthalmidae). Journal of the New York Entomological Society, 108, 1–12.CrossRefGoogle Scholar
Morsy, T. A., Kady, El G. A., Salama, M. M. & Sabry, A. H. (1993). The seasonal abundance of Gerbillus pyramidum and their flea ectoparasites in Al Arish, North Sinai Governorate, Egypt. Journal of the Egyptian Society of Parasitology, 23, 269–276.Google ScholarPubMed
Moser, B. A., Koehler, P. G. & Patterson, R. S. (1991). Effect of larval diet on cat flea (Siphonaptera: Pulicidae) developmental times and adult emergence. Journal of Economic Entomology, 84, 1257–1261.CrossRefGoogle ScholarPubMed
Moskalenko, V. V. (1958). On the effect of temperature on the behaviour of fleas after deaths of their host. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 181–184 (in Russian).Google Scholar
Moskalenko, V. V. (1963a). On the longevity of several flea species parasitic on rodents in the Primorie Region. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 6, 166–169 (in Russian).Google Scholar
Moskalenko, V. V. (1963b). On the effect of temperature on reproduction of some flea species from the Primorie Region under laboratory conditions. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 6, 162–165 (in Russian).Google Scholar
Moskalenko, V. V. (1966). On the frequency of feeding in fleas from the Primorie Region. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 349–354 (in Russian).
Mouillot, D. & Poulin, R. (2004). Taxonomic partitioning shedding light on the diversification of parasite communities. Oikos, 104, 205–207.CrossRefGoogle Scholar
Mouillot, D., Krasnov, B. R., Gaston, K., Shenbrot, G. I. & Poulin, R. (2006). Conservatism of host specificity in parasites. Ecography, 29, 596–602.CrossRefGoogle Scholar
Moura, M. O., Bordignon, M. & Graciolli, G. (2003). Host characteristics do not affect community structure of ectoparasites on the fishing bat Noctilio leporinus (L., 1758) (Mammalia: Chiroptera). Memórias do Instituto Oswaldo Cruz, 98, 811–815.CrossRefGoogle Scholar
Moynahan, E. J. (1987). Kawasaki disease: a novel feline virus transmitted by fleas. Lancet, i (8526), 195.CrossRefGoogle Scholar
Muehlen, M., Heukelbach, J., Wilcke, T., et al. (2003). Investigations on the biology, epidemiology, pathology and control of Tunga penetrans in Brazil. II. Prevalence, parasite load and topographic distribution of lesions in the population of a traditional fishing village. Parasitology Research, 90, 449–455.CrossRefGoogle ScholarPubMed
Muirhead-Thomson, R. C. (1968). Ecology of Insect Vector Populations. New York: Academic Press.Google Scholar
Mulenga, A., Sugimoto, C. & Onuma, M. (2000). Issues in tick vaccine development: identification and characterization of potential candidate vaccine antigens. Microbes and Infection, 2, 1353–1361.CrossRefGoogle ScholarPubMed
Mules, M. W. (1940). Notes on the life history and artificial breeding of the Australian ‘stickfast’ flea Echidnophaga mirmecobii Rothschild. Australian Journal of Experimental Biological and Medical Science, 18, 385–390.CrossRefGoogle Scholar
Muller, G. H., Kirk, R. W., Scott, D. W., et al. (2001). Muller and Kirk's Small Animal Dermatology, 6th edn. Philadelphia, PA: Saunders College Publishing.Google Scholar
Muñoz, L., Milenko, A. & Casanueva, M. E. (2003). Prevalencia e intensidad de ectoparasitos asociados a Tadarida brasiliensis (Geoffroy Saint-Hilaire, 1824) (Chiroptera: Molossidae) en Concepción. Gayana, 67, 1–8.Google Scholar
Murzakhmetova, K. (1958). On studying physiological activity of fleas parasitic on gerbils. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 4, 223–226 (in Russian).Google Scholar
Myalkovskaya, S. A. (1983). Some ecological characteristics of fleas parasitic on the pygmy ground squirrels in Dagestan. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 254–255 (in Russian).Google Scholar
Nakazawa, K., Oisi, I. & Kumi, S. (1957). Seasonal prevalence of rodent fleas. Medical Entomology and Zoology, 8, 11–13 (in Japanese).CrossRefGoogle Scholar
Naumov, R. L. & Gutova, V. P. (1984). Experimental study of the participation of gamasid mites and fleas in circulation of the tick-borne encephalitis virus (a review). Parazitologiya, 18, 106–115 (in Russian).Google Scholar
Navarro, C., Marzal, A., Lope, F. & M⊘ller, A. P. (2003). Dynamics of an immune response in house sparrows Passer domesticus in relation to time of day, body condition and blood parasite infection. Oikos, 101, 291–298.CrossRefGoogle Scholar
Nayak, N. C. & Bhowmik, M. K. (1990). Goat flea (order Siphonaptera) as a possible vector for the transmission of caprine mycoplasmal polyarthritis with septicaemia. Preventive Veterinary Medicine, 9, 259–266.CrossRefGoogle Scholar
Nazarova, I. V. (1981). Fleas of the Volga–Kama Region. Moscow, USSR: Nauka (in Russian).Google Scholar
Neal, E. G. & Roper, T. J. (1991). The environmental impact of badgers (Meles meles) and their setts. Symposia of the Zoological Society of London, 63, 89–106.Google Scholar
Nee, S. (1994). How populations persist. Nature, 367, 123–124.CrossRefGoogle Scholar
Nee, S., Hassell, M. P. & May, R. M. (1997). Two-species metapopulation models. In Metapopulation Biology: Ecology, Genetics, and Evolution, ed. Hanski, I. & Gilpin, M. E.. San Diego, CA: Academic Press, pp. 123–147.Google Scholar
Nekola, J. C. & White, P. S. (1999). The distance decay of similarity in biogeography and ecology. Journal of Biogeography, 26, 867–878.CrossRefGoogle Scholar
Nelson, R. J. & Demas, G. E. (1996). Seasonal changes in immune function. Quarterly Review of Biology, 71, 511–548.CrossRefGoogle ScholarPubMed
Nelzina, E. N., Danilova, G. M. & Tchernova, G. I. (1963). Abundance and spatial distribution of micropopulations of hematophagous arthropods in microhabitats of Citellus pygmaeus. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 32, 45–54 (in Russian).Google Scholar
Němec, F. (1993). Flea community inhabiting nests of the common vole (Microtus arvalis Pallas, 1779) in west Bohemian farmland. Folia Musei Rerum Naturalium Bohemiae Occidentalis Zoologica, 38, 1–37.Google Scholar
Neter, J., Wasserman, W. & Kutner, M. H. (1990). Applied Linear Statistical Models: Regression, Analysis of Variance, and Experimental Design. Homewood, IL: Richard D. Irwin.
Neuhaus, P. (2003). Parasite removal and its impact on litter size and body condition in Columbian ground squirrels (Spermophilus columbianus). Proceedings of the Royal Society of London B, 270, S213–S215.CrossRefGoogle Scholar
Newton, I. (1998). Population Limitation in Birds. London: Academic Press.Google Scholar
Nijssen, A. (1985). Body temperature regulation by care of the fur in rats Rattus norvegicus albinos. Netherlands Journal of Zoology, 35, 423–437.CrossRefGoogle Scholar
Nikitina, N. A. (1961). Results of marking and recapturing of small mammals in the Komi Soviet Republic. Bulletin of the Moscow Naturalist Society, Series Biology, 66, 15–25 (in Russian).Google Scholar
Nikitina, N. A. & Nikolaeva, G. (1979). Study of the ability of some rodents to get rid of fleas. Zoologicheskyi Zhurnal, 58, 931–933 (in Russian).Google Scholar
Nikitina, N. A. & Nikolaeva, G. (1981). Ability of rodents to clean themselves of specific and non-specific fleas. Zoologicheskyi Zhurnal, 60, 165–167 (in Russian).Google Scholar
Nikulshin, S. V. (1980). Descriptions of the annual cycles of fleas (Aphaniptera) parasitic on Citellus musicus from the Baksan Valley. Parazitologiya, 14, 134–141 (in Russian).Google Scholar
Nikulshin, S. V. & Shinkareva, V. N. (1983). Seasonal patterns of ecology of fleas Citellophilus tesquorum in the Enikol Tract. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 260–261 (in Russian).Google Scholar
Nilsson, J. A. (2003). Ectoparasitism in marsh tits: costs and functional explanations. Behavioral Ecology, 14, 175–181.CrossRefGoogle Scholar
Njau, B. C. & Nyindo, M. (1987). Detection of immune response in rabbits infested with Rhipicephalus appendiculatus andRhipicephalus evertsi evertsi. Research in Veterinary Science, 43, 217–221.Google Scholar
Njunwa, K. J., Mwaiko, G. L., Kilonzo, B. S. & Mhina, J. I. (1989). Seasonal patterns of rodents, fleas and plague status in the Western Usambara Mountains, Tanzania. Medical and Veterinary Entomology, 3, 17–22.CrossRefGoogle ScholarPubMed
Nordling, D., Andersson, M., Zohari, S. & Gustafsson, L. (1998). Reproductive effort reduces specific immune response and parasite resistance. Proceedings of the Royal Society of London B, 265, 1291–1298.CrossRefGoogle Scholar
Norman, R., Bowers, R. G., Begon, M. & Hudson, P. J. (1999). Persistence of tick-borne virus in the presence of multiple host species: tick reservoirs and parasite mediated competition. Journal of Theoretical Biology, 200, 111–118.CrossRefGoogle ScholarPubMed
Nosil, P. (2002). Transition rates between specialization and generalization in phytophagous insects. Evolution, 56, 1701–1706.CrossRefGoogle ScholarPubMed
Novokreshchenova, N. S. (1960). Materials on ecology of fleas of Citellus pygmaeus in relation to their epizootological importance. Proceedings of the Anti-Plague ‘Microb’ Institute, 4, 444–456 (in Russian).Google Scholar
Novokreshchenova, N. S. (1962). Materials on comparative ecology of three species of fleas parasitic on the great gerbil. In Particularly Dangerous and Natural Diseases, ed. ,Anonymous. Moscow, USSR: Medgiz, pp. 53–63 (in Russian).Google Scholar
Novokreshchenova, N. S. & Kuznetsova, G. S. (1964). Ecological characteristic of fleas of the great gerbil in regions with permanent plague epizootics. Zoologicheskyi Zhurnal, 43, 1638–1647 (in Russian).Google Scholar
Novokreshchenova, N. S., Soldatkin, I. S. & Levoshina, A. I. (1968). Comparative frequency of blood meals in various flea species, measured in the laboratory conditions using radioactive indicators. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 49–54 (in Russian).Google Scholar
Novokreshchenova, N. S., Zagniborodova, E. N., Zabegalova, M. N., et al. (1975). Multiannual dynamics of density in fleas parasitic on the great gerbils in Turkmenistan. Problems of Particularly Dangerous Diseases, 43/44, 84–90 (in Russian).Google Scholar
Novozhilova, E. N. (1977). Ectoparasites of small mammals and inhabitants of their burrows in the Pre-Polar Ural. Proceedings of the Komi Branch of the Academy of Sciences of the USSR, 34, 125–139 (in Russian).Google Scholar
Nuriev, K. K., Rapoport, L. P., Shokputov, T. M., Orlova, L. M. & Duisenbiev, D. M. (2004). Materials on rodent fleas from mountain regions of southern Kazakhstan. Quarantinable and Zoonotic Infections in Kazakhstan, 9, 66–70 (in Russian).Google Scholar
Oberdorff, T., Hugueny, B., Compin, A. & Belkessam, D. (1998). Non-interactive fish communities in the coastal streams of north-western France. Journal of Animal Ecology, 67, 472–484.CrossRefGoogle Scholar
Brien, O' E. L. & Dawson, R. D. (2005). Perceived risk of ectoparasitism reduces primary reproductive investment in tree swallows Tachycineta bicolor. Journal of Avian Biology, 36, 269–275.CrossRefGoogle Scholar
Donnell, O' M. J. & Machin, J. (1988). Water vapor absorption by terrestrial organisms. Advances in Comparative and Environmental Physiology, 2, 47–90.CrossRefGoogle Scholar
Oguge, N., Rarieya, M. & Ondiaka, P. (1997). A preliminary survey of macroparasite community of rodents of Kahawa, central Kenya. Belgian Journal of Zoology, 127, S113–S118.Google Scholar
Hare, O' N. (2005). Asking the right questions. New Zealand Listener, 197 (Jan. 29–Feb. 4), No. 3377.Google Scholar
Olsen, N. J. & Kovacs, W. J. (1996). Gonadal steroids and immunity. Endocrine Review, 17, 369–384.Google ScholarPubMed
Olsson, K. & Allander, K. (1995). Do fleas, and/or old nest material, influence nest-site preference in hole-nesting passerines? Ethology, 101, 160–170.CrossRefGoogle Scholar
Olsufiev, N. G. (1975). Taxonomy, Microbiology and Laboratory Diagnostics of the Tularaemia Pathogen. Moscow, USSR: Meditsina (in Russian).Google Scholar
Olsufiev, N. G. & Dunaeva, T. N. (1960). Epizootology (natural focality) of tularaemia. In Tularemia, ed. Olsufiev, N. G. & Rudnev, G. P.. Moscow, USSR: Meditsina, pp. 136–206 (in Russian).Google Scholar
Oparina, O., Starozhitskaya, G. S. & Samurov, M. A. (1989). Human blood-sucking ability of fleas parasitic on rodents and small gerbils. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 58, 23–25 (in Russian).Google Scholar
Opdebeeck, J. P. & Slacek, B. (1993). An attempt to protect cats against infestation with Ctenocephalides felis felis using gut membrane antigens as a vaccine. International Journal for Parasitology, 23, 1063–1067.CrossRefGoogle ScholarPubMed
Oppliger, A., Richner, H. & Christe, P. (1994). Effect of an ectoparasite on lay date, nest-site choice, desertion, and hatching success in the great tit (Parus major). Behavioral Ecology, 5, 130–134.CrossRefGoogle Scholar
Orell, M., Rytkönen, S. & Ilomäki, K. (1993). Do pied flycatchers prefer nest boxes with old nest material?Annales Zoologici Fennici, 30, 313–316.Google Scholar
Osacar-Jimenez, J. J., Lucientes-Curdi, J. & Calvete-Margolles, C. (2001). Abiotic factors influencing the ecology of wild rabbit fleas in north-eastern Spain. Medical and Veterinary Entomology, 15, 157–166.CrossRefGoogle ScholarPubMed
Osbrink, W. L. & Rust, M. (1984). Fecundity and longevity of the adult cat flea Ctenocephalides felis felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 21, 727–731.CrossRefGoogle Scholar
Osbrink, W. L. & Rust, M. (1985). Cat flea (Siphonaptera: Pulicidae): factors influencing host finding behaviour in the laboratory. Annals of the Entomological Society of America, 78, 29–34.CrossRefGoogle Scholar
Ostfeld, R. & Keesing, F. (2000). The function of biodiversity in the ecology of vector-borne zoonotic diseases. Canadian Journal of Zoology, 78, 2061–2078.CrossRefGoogle Scholar
Overal, W. L. (1980). Host-relations of the batfly Megistopoda aranea (Diptera: Streblidae) in Panama. University of Kansas Scientific Bulletin, 52, 1–20.Google Scholar
Öztürk, L., Pelin, Z., Karadeniz, D., et al. (1999). Effects of 48 hours' sleep deprivation on human immune profile. Sleep Research Online, 2, 107–111.Google Scholar
Pacala, S. W. & Dobson, A. P. (1988). The relation between the number of parasites/host and host age: population dynamic causes and maximum likelihood estimation. Parasitology, 96, 197–210.CrossRefGoogle ScholarPubMed
Pacejka, A. J., Gratton, C. M. & Thompson, C. F. (1998). Do potentially virulent mites affect house wren (Troglodytes aedon) reproductive success?Ecology, 79, 1797–1806.CrossRefGoogle Scholar
Page, R. M. D. (1990). Component analysis: a valiant failure?Cladistics, 6, 119–136.CrossRefGoogle Scholar
Page, R. M. D. (1993a). Parasites, phylogeny and cospeciation. International Journal for Parasitology, 23, 499–506.CrossRefGoogle Scholar
Page, R. M. D. (1993b). Genes, organisms, and areas: the problem of multiple lineages. Systematic Biology, 42, 77–84.CrossRefGoogle Scholar
Page, R. M. D. (1994a). Parallel phylogenies: reconstructing the history of host–parasite assemblages. Cladistics, 10, 155–173.CrossRefGoogle Scholar
Page, R. M. D. (1994b). Maps between trees and cladistic analysis of historical associations among genes, organisms, and areas. Systematic Biology, 43, 58–77.Google Scholar
Page, R. D. M. (ed.) (2003). Tangled Trees: Phylogeny, Cospeciation, and Coevolution. Chicago, IL: University of Chicago Press.Google Scholar
Page, R. D. M. & Charleston, M. D. (1998). Trees within trees: phylogeny and historical associations. Trends in Ecology and Evolution, 13, 356–359.CrossRefGoogle ScholarPubMed
Palombit, R. A., Cheney, D. L. & Seyfarth, R. M. (2001). Female–female competition for male ‘friends’ in wild chacma baboons, Papio cynocephalus ursinus. Animal Behaviour, 61, 1159–1171.CrossRefGoogle Scholar
Pampiglione, S., Trentini, M., Fioravanti, M. L., Onore, G. & Rivasi, F. (2003). Additional description of a new species of Tunga (Siphonaptera) from Ecuador. Parasite, 10, 9–15.CrossRefGoogle ScholarPubMed
Panchenko, G. M. (1971). Ecological study of larvae of the rat fleas of Siberia and Far East. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 229–231 (in Russian).Google Scholar
Paramonov, B. B., Emelianova, N. D., Zarubina, V. N. & Kontrimavitchus, V. L. (1966). Materials for the study of ectoparasites of rodents and shrews of the Kamchatka Peninsula. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 333–341 (in Russian).Google Scholar
Park, T. (1948). Experimental studies of interspecific competition. I. Competition between populations of the flour beetles, Tribolium confusum Duval and Tribolium castaneum Herbst. Ecological Monographs, 18, 265–308.CrossRefGoogle Scholar
Parman, D. C. (1923). Biological notes on the hen flea, Echidnophaga gallinacea. Journal of Agricultural Research, 23, 1007–1009.Google Scholar
Parola, R. & Raoult, D. (2006). Tropical rickettsioses. Clinics in Dermatology, 24, 191–200.CrossRefGoogle ScholarPubMed
Parola, P., Davoust, B. & Raoult, D. (2005). Tick- and flea-borne rickettsial emerging zoonoses. Veterinary Research, 36, 469–492.CrossRefGoogle ScholarPubMed
Parsons, P. A. (1990). Fluctuating asymmetry: an epigenetic measure of stress. Biological Reviews, 17, 391–421.Google Scholar
Pascal, M., Beaucournu, J. C. & Lorvelec, O. (2004). An enigma: the lack of Siphonaptera on wild rats and mice on densely populated tropical islands. Acta Parasitologica, 49, 168–172.Google Scholar
Paterson, A. M. & Banks, J. (2001). Analytical approaches to measuring cospeciation of host and parasites: through a glass, darkly. International Journal for Parasitology, 31, 1012–1022.CrossRefGoogle ScholarPubMed
Paterson, A. M. & Gray, R. D. (1997). Host–parasite co-speciation, host switching, and missing the boat. In Host–Parasite Evolution: General Principles and Avian Models, ed. Clayton, D. H. & Moore, J.. Oxford, UK: Oxford University Press, pp. 236–250.Google Scholar
Paterson, A. M., Gray, R. D. & Wallis, G. P. (1993). Parasites, petrels and penguins: does louse presence reflect seabird phylogeny?International Journal for Parasitology, 23, 515–526.CrossRefGoogle Scholar
Paterson, A. M., Wallis, G. P., Wallis, L. J. & Gray, R. D. (2000). Seabird and louse coevolution: Complex histories revealed by 12S rRNA sequences and reconciliation analyses. Systematic Biology, 38, 144–153.Google Scholar
Patrick, M. J. (1991). Distribution of enteric helminthes in Glaucomys volans L. (Sciuridae): a test for competition. Ecology, 72, 755–758.CrossRefGoogle Scholar
Patterson, B. D. (1990). On the temporal development of nested subset patterns of species composition. Oikos, 59, 330–342.CrossRefGoogle Scholar
Patterson, B. D. & Atmar, W. (1986). Nested subsets and the structure of insular mammalian faunas and archipelagos. Biological Journal of the Linnean Society, 28, 65–82.CrossRefGoogle Scholar
Patterson, B. D. & Brown, J. H. (1991). Regionally nested patterns of species composition in granivorous rodent assemblages. Journal of Biogeography, 18, 395–402.CrossRefGoogle Scholar
Pauller, O. F. & Tchipizubova, P. A. (1958). Ecology of fleas of the Daurian ground squirrel in the Trans-Baikalia. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 161–179 (in Russian).Google Scholar
Pauller, O. F., Elshanskaya, N. I. & Shvetsova, I. V. (1966). Ecological and faunistical review of mammalian and bird ectoparasites in the tularemia focus of the Selenga River delta. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 322–332 (in Russian).Google Scholar
Peach, W. J., Fowler, J. A. & Greenwood, M. T. (1987). Seasonal variation in the infestation of starlings Sturnus vulgaris by fleas (Siphonaptera). Bird Study, 34, 251–252.CrossRefGoogle Scholar
Pennings, S. C. & Silliman, B. R. (2005). Linking biogeography and community ecology: latitudinal variation in plant–herbivore interaction strength. Ecology, 86, 2310–2319.CrossRefGoogle Scholar
Perlman, S. J. & Jaenike, J. (2001). Competitive interactions and persistence of two nematode species that parasitize Drosophila recens. Ecology Letters, 4, 577–584.CrossRefGoogle Scholar
Perry, J. N. (1988). Some models for spatial variability of animal species. Oikos, 51, 124–130.CrossRefGoogle Scholar
Perry, J. N. & Taylor, L. R. (1986). Stability of real interacting populations in space and time: implications, alternatives and negative binomial kc. Journal of Animal Ecology, 55, 1053–1068.CrossRefGoogle Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Peterson, A. T., Soberon, J. & Sanchez-Cordero, V. (1999). Conservatism of ecological niches in evolutionary time. Science, 285, 1265–1267.CrossRefGoogle ScholarPubMed
Petit, C., Hossaert-McKey, M., Perret, P., Blondel, J. & Lambrechts, M. M. (2002). Blue tits use selected plants and olfaction to maintain an aromatic environment for nestlings. Ecology Letters, 5, 585–589.CrossRefGoogle Scholar
Peus, F. (1968). Über die beiden Bernstein-Flöhe (Insecta, Siphonaptera). Paläontologische Zeitschrift, 42, 62–72.CrossRefGoogle Scholar
Peus, F. (1970). Zur Kenntnis der Flöhe Deutschlands (Insecta, Siphonaptera). IV. Faunistik und Ökologie der Säugeteierflöhe. Zoologische Jahrbücher, Abteilung für Systematic Ökologie und Geographie der Tiere, 99, 400–418.Google Scholar
Piersma, T. & Drent, J. (2003). Phenotypic flexibility and the evolution of organismal design. Trends in Ecology and Evolution, 18, 228–233.CrossRefGoogle Scholar
Pigage, H. K. & Larson, O. R. (1983). The detection of glycerol in overwintering purple martin fleas, Ceratophyllus idius. Comparative Biochemistry and Physiology A, 75, 593–595.CrossRefGoogle Scholar
Pigage, H. K., Pigage, J. C. & Tillman, J. F. (2005). Fleas associated with the northern pocket gopher (Thomomys talpoides) in Elbert County, Colorado. Western North American Naturalist, 65, 210–214.Google Scholar
Pilgrim, R. L. C. & Galloway, T. D. (2004). Descriptions of flea larvae (Siphonaptera: Ceratophyllidae, Leptopsyllidae) found in nests of the house martin, Delichon urbica (Aves: Hirundinidae), in Great Britain. Journal of Natural History, 38, 473–502.CrossRefGoogle Scholar
Pimm, S. L. (1979). The structure of food webs. Theoretical Population Biology, 16, 144–158.CrossRefGoogle ScholarPubMed
Pimm, S. L. & Rosenzweig, M. L. (1981). Competitors and habitat use. Oikos, 37, 1–6.CrossRefGoogle Scholar
Pollitzer, R. (1960). A review of recent literature on plague. Bulletin of the World Health Organization, 23, 313–400.Google ScholarPubMed
Pollitzer, R. & Meyer, K. F. (1961). The ecology of plague. In Studies of Disease Ecology, ed. May, J. F.. New York: Hafner, pp. 433–590.Google Scholar
Ponomarenko, A. G. (1976). The new insect from the Cretaceous of the Trans-Baikalia was a probable parasite of pterosaurs. Paleontologicheskyi Zhurnal, 3, 102–106 (in Russian).Google Scholar
Ponomarenko, A. G. (1986). Scarabaeiformis incertae sedis. Transactions of the Joint Soviet–Mongolian Paleontological Expedition, 28, 110–112 (in Russian).Google Scholar
Popov, V. N. & Verzhutsky, D. B. (1988). Characteristics of intrapopulation aggregations of the long-tailed ground squirrel (Citellus undulatus Pall.) under density decline. Bulletin of the Moscow Naturalist Society, Series Biology, 93, 47–50 (in Russian).Google Scholar
Popova, A. S. (1968). Flea fauna of the Moyynkum Desert. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 402–406 (in Russian).Google Scholar
Poulin, R. (1993). The disparity between observed and uniform distibutions: a new look at parasite aggregation. International Journal for Parasitology, 23, 937–944.CrossRefGoogle Scholar
Poulin, R. (1995a). Clutch size and egg size in free-living and parasitic copepods: a comparative analysis. Evolution, 49, 325–336.CrossRefGoogle Scholar
Poulin, R. (1995b). Phylogeny, ecology, and the richness of parasite communities in vertebrates. Ecological Monographs, 65, 283–302.CrossRefGoogle Scholar
Poulin, R. (1996a) Sexual inequalities in helminth infections: a cost of being male?American Naturalist, 147, 289–295.CrossRefGoogle Scholar
Poulin, R. (1996b). Richness, nestedness and randomness in parasite infracommunity structure. Oecologia, 105, 545–551.CrossRefGoogle Scholar
Poulin, R. (1997). Parasite faunas of freshwater fish: the relationship between richness and the specificity of parasites. International Journal for Parasitology, 27, 1091–1098.CrossRefGoogle ScholarPubMed
Poulin, R. (1998). Large-scale patterns of host use by parasites of freshwater fishes. Ecology Letters, 1, 118–128.CrossRefGoogle Scholar
Poulin, R. (1999). Body size vs. abundance among parasite species: positive relationships?Ecography, 22, 246–250.CrossRefGoogle Scholar
Poulin, R. (2003). The decay of similarity with geographical distance in parasite communities of vertebrate hosts. Journal of Biogeography, 30, 1609–1615.CrossRefGoogle Scholar
Poulin, R. (2005). Relative infection levels and taxonomic distances among the host species used by a parasite: insights into parasite specialization. Parasitology, 130, 109–115.CrossRefGoogle ScholarPubMed
Poulin, R. (2006). Variation in infection parameters among populations within parasite species: intrinsic properties versus local factors. International Journal for Parasitology, 36, 877–885.CrossRefGoogle ScholarPubMed
Poulin, R. (2007a). Evolutionary Ecology of Parasites: From Individuals to Communities, 2nd edn. Princeton, NJ: Princeton University Press.Google Scholar
Poulin, R. (2007b). Are there general laws in parasite ecology?Parasitology, 134, 763–776.CrossRefGoogle Scholar
Poulin, R. & Guégan, J.-F. (2000). Nestedness, antinestedness, and relationship between prevalence and intensity in ectoparasite assemblages of marine fish: a spatial model of species co-existence. International Journal for Parasitology, 30, 1147–1152.CrossRefGoogle Scholar
Poulin, R. & Hamilton, W. J. (1997). Ecological correlates of body size and egg size in parasitic Ascothoracida and Rhizocephala (Crustacea). Acta Oecologica, 18, 621–635.CrossRefGoogle Scholar
Poulin, R. & Morand, S. (2004). Parasite Biodiversity. Washington, DC: Smithsonian Institution Press.Google Scholar
Poulin, R. & Mouillot, D. (2003). Parasite specialization from a phylogenetic perspective: a new index of host specificity. Parasitology, 126, 473–480.CrossRefGoogle ScholarPubMed
Poulin, R. & Mouillot, D. (2004a). The relationship between specialization and local abundance: the case of helminth parasites of birds. Oecologia, 140, 372–378.CrossRefGoogle Scholar
Poulin, R. & Mouillot, D. (2004b). The evolution of taxonomic diversity in helminth assemblages of mammalian hosts. Evolutionary Ecology, 18, 231–247.CrossRefGoogle Scholar
Poulin, R. & Mouillot, D. (2005a). Combining phylogenetic and ecological information into a new index of host specificity. Journal of Parasitology, 91, 511–514.CrossRefGoogle Scholar
Poulin, R. & Mouillot, D. (2005b). Host specificity and the probability of discovering species of helminth parasites. Parasitology, 130, 709–715.CrossRefGoogle Scholar
Poulin, R. & Valtonen, E. T. (2001). Nested assemblages resulting from host-size variation: the case of endoparasite communities in fish hosts. International Journal for Parasitology, 31, 194–1204.CrossRefGoogle ScholarPubMed
Poulin, R. & Valtonen, E. T. (2002). The predictability of helminth community structure in space: a comparison of fish populations from adjacent lakes. International Journal for Parasitology, 30, 1235–1243.CrossRefGoogle Scholar
Poulin, P., Brodeur, J. & Moore, J. (1994). Parasite manipulation of host behaviour: should hosts always lose?Oikos, 70, 479–484.CrossRefGoogle Scholar
Poulin, R., Mouillot, D. & George-Nascimento, M. (2003). The relationship between species richness and productivity in metazoan parasite communities. Oecologia, 137, 277–285.CrossRefGoogle ScholarPubMed
Poulin, R., Krasnov, B. R. & Morand, S. (2006a). Patterns of host specificity in parasites exploiting small mammals. In Micromammals and Macroparasites: From Evolutionary Ecology to Management, ed. Morand, S., Krasnov, B. R. & Poulin, R.. New York: Springer-Verlag, pp. 233–256.CrossRefGoogle Scholar
Poulin, R., Krasnov, B. R., Shenbrot, G. I., Mouillot, D. & Khokhlova, I. S. (2006b). Evolution of host specificity in fleas: is it directional and irreversible?International Journal for Parasitology, 36, 185–191.CrossRefGoogle Scholar
Prasad, R. S. (1969). Influence of host on fecundity of the Indian rat flea, Xenopsylla cheopis (Roths.). Journal of Medical Entomology, 6, 443–447.CrossRefGoogle Scholar
Prasad, R. S. (1972). Different site selections by the rat fleas Xenopsyllsa cheopis and Xenopsyllsa astia (Siphonaptera, Pulicidae). Entomologist's Gazette, 108, 63–64.Google Scholar
Prasad, R. S. (1973). Studies on host–flea relationships. II. Sex hormones of the host and fecundity of rat fleas Xenopsylla astia (Rothschild) and Xenopsylla cheopis (Rothschild) (Siphonaptera). Indian Journal of Medical Research, 61, 38–44.Google Scholar
Prasad, R. S. (1976). Studies on host–flea relationships. IV. Progesterone and cortisone do not influence reproductive potentials of rat fleas Xenopsylla astia (Rothschild) and Xenopsylla cheopis (Rothschild) (Siphonaptera). Parasitology Research, 50, 81–46.Google Scholar
Prasad, R. S. (1987). Host dependency among haematophagous insects: a case study on flea–host association. Proceedings of the Indian Academy of Sciences, 96, 349–360.CrossRefGoogle Scholar
Prasad, R. S. & Kamala Bai, M. (1976). Studies on host specificity in fleas. In Insect and Host Specificity: Proceedings of the Symposium on Problems of Host Specificity, ed. Ananthakrishnan, T. N.. Madras, India: Loyola College, pp. 111–115.Google Scholar
Price, T., Lovette, I. J., Bermingham, E., Gibbs, H. L. & Richman, A. D. (2000). The imprint of history on communities of North American and Asian warblers. American Naturalist, 156, 354–367.CrossRefGoogle ScholarPubMed
Prokopiev, V. N. (1969). Morphological types of cocoons and mechanism of emergence of imago fleas. In Proceedings of the 4th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute and Kainar, pp. 182–184 (in Russian).Google Scholar
Promislow, D. E. L. (1992). Costs of sexual selection in natural populations of mammals. Proceedings of the Royal Society of London B, 247, 203–210.CrossRefGoogle Scholar
Puchala, P. (2004). Detrimental effects of larval blow files (Protocalliphora azurea) on nestlings and breeding success of tree sparrows (Passer montanus). Canadian Journal of Zoology, 82, 1285–1290.CrossRefGoogle Scholar
Pullen, S. R. & Meola, R. W. (1995). Survival and reproduction of the cat flea (Siphonaptera: Pulicidae) fed human blood on an artificial membrane system. Journal of Medical Entomology, 32, 467–4670.CrossRefGoogle ScholarPubMed
Pulliam, H. R. (2000). On the relationship between niche and distribution. Ecology Letters, 3, 349–361.CrossRefGoogle Scholar
Punsky, E. E. & Zagniborodova, E. N. (1964). Role of fleas Xenopsylla gerbilli gerbilli in maintenance and transmission of the pathogen of erysepeloid. Proceedings of the Turkmenian Scientific Institute of Epidemiology and Hygiene, 6, 345–347 (in Russian).Google Scholar
Pushnitsa, F. A., Shevchenko, S. F., Mironov, A. N., et al. (1978). Present status of abundance and habitat distribution of the ground squirrels and their fleas in the eastern parts of the Rostov Region. Problems of Particularly Dangerous Diseases, 60, 48–52 (in Russian).Google Scholar
Qi, Y.-M. (1990a). Fine structures of the reproductive system of three flea species: development of female genitalia. Acta Entomologica Sinica, 33, 182–188 (in Chinese).Google Scholar
Qi, Y.-M. (1990b). Fine structures of the reproductive system of three flea species: development of male genitalia. Acta Entomologica Sinica, 33, 403–411 (in Chinese).Google Scholar
Qian, T.-J., Gong, Z.-D. & Guo, X.-G. (2000). Sex ratio analysis on dominant flea species of the flea community in the foci of human plague in Yunnan. Medical Journal of Dali College, 9, 1–3 (in Chinese).Google Scholar
Radovsky, F. J. (1985). Evolution of mammalian mesostigmatid mites. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 441–504.Google Scholar
Randolph, S. E. (1977). Changing spatial relationships in a population of Apodemus sylvaticus with the onset of breeding. Journal of Animal Ecology, 46, 653–676.CrossRefGoogle Scholar
Randolph, S. E. (1994). Density-dependent acquired resistance to ticks in natural hosts, independent of concurrent infection with Babesia microti. Parasitology, 108, 413–419.CrossRefGoogle ScholarPubMed
Rapoport, E. H. (1982). Areography: Geographic Strategies of Species. London: Pergamon Press.Google Scholar
Rapoport, L. P., Morozov, Y. A. & Korneyev, G. A. (1976). Intensity of parasitological contacts in the colonies of the great gerbils under different population densities of animals and their fleas. Parazitologiya, 10, 392–396 (in Russian).Google Scholar
Rapoport, L. P., Kondratenko, L. P., Orlova, L. M., et al. (2007). Fleas in the Moyynkum Desert and their epozootological role. Zoologicheskyi Zhurnal, 86, 44–51 (in Russian).Google Scholar
Rasnitsyn, A. P. (1992). Strashila incredibilis, a new enigmatic mecopteroid insect with possible siphonapteran affinities from upper Jurassic of Siberia. Psyche, 99, 323–333.CrossRefGoogle Scholar
Rasnitsyn, A. P. (2002a). Infraclass Gryllones Laicharting, 1781: the grylloneans. In History of Insects, ed. Rasnitsyn, A. P. & Quicke, D. L. J.. Dordrecht, the Netherlands: Kluwer, pp. 254–324.CrossRefGoogle Scholar
Rasnitsyn, A. P. (2002b). Order Pulicida Billbergh, 1820: the fleas. In History of Insects, ed. Rasnitsyn, A. P. & Quicke, D. L. J.. Dordrecht, the Netherlands: Kluwer, pp. 239–242.CrossRefGoogle Scholar
Rassokhina, O. S., Starozhitskaya, G. S. & Knyazeva, T. V. (1985). Fecundity of four flea species in the laboratory colonies. Parazitologiya, 19, 488–490 (in Russian).Google Scholar
Raszl, S. M., Cabral, D. D. & Linardi, P. M. (1998). Xenopsylla cheopis on dogs from Brazil: first report. Arquivo Brasileiro de Medicina Veterinária e Zootecnia, 50, 211–212 (in Portugese).Google Scholar
Raszl, S. M., Cabral, D. D. & Linardi, P. M. (1999). Notas sobre Sifonápteros (Pulicidae, Tungidae e Rhopalopsyllidae) de carnívoros domésticos brasileiros. Revista Brasiliera de Entomologia, 43, 95–97.Google Scholar
Ratovonjato, J., Duchemin, J. B. & Chanteau, S. (2000). Optimized method for rearing fleas (Xenopsylla cheopis and Synopsyllus fonquerniei). Archives de l'Institut Pasteur de Madagascar, 66, 75–77.Google Scholar
Rausher, M. D. (1993). The evolution of habitat preference: avoidance and adaptation. In Evolution of Insect Pests: Patterns of Variation, ed. Kim, K. C. & McPherson, B. A.. New York: John Wiley, pp. 259–283.Google Scholar
Ravkin, Y. S. & Sapegina, V. F. (1990). Fleas of rodents of the southern taiga of the Pri-Angarie. Bulletin of the Siberian Branch of the Academy of Sciences of the USSR, Series Biological Sciences, 3, 63–68 (in Russian).Google Scholar
Rea, J. G. & Irwin, S. W. B. (1994). The ecology of host-finding behaviour and parasite transmission: past and future perspectives. Parasitology, 109, S31–S39.CrossRefGoogle ScholarPubMed
Rechav, Y. (1992). Naturally acquired resistance to ticks: a global view. Insect Science and its Application, 13, 495–504.Google Scholar
Rechav, Y. & Dauth, J. (1987). Development of resistance in rabbits to immature stages of the ixodid tick Rhipicephalus appendiculatus. Medical and Veterinary Entomology, 1, 177–183.CrossRefGoogle ScholarPubMed
Rechav, Y. & Fielden, L. J. (1995). The effect of host resistance on the metabolic rate of engorged females of Rhipicephalus evertsi evertsi. Medical and Veterinary Entomology, 9, 289–292.CrossRefGoogle ScholarPubMed
Rechav, Y & Fielden, L. J. (1997). The effect of various host species on the feeding performance of immature stages of the tick Hyalomma truncatum (Acari: Ixodidae). Experimental and Applied Acarology, 21, 551–559.CrossRefGoogle Scholar
Rechav, Y., Heller-Haupt, A. & Varma, M. G. R. (1989). Resistance and cross-resistance in guinea-pigs and rabbits to immature stages of ixodid ticks. Medical and Veterinary Entomology, 3, 333–336.CrossRefGoogle ScholarPubMed
Redford, K. H. & Eisenberg, J. F. (1992). Mammals of the Neotropics: The Southern Cone, vol. 2, Chile, Argentine, Uruguay, Paraguay. Chicago, IL: University of Chicago Press.Google Scholar
Reichardt, T. R. & 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. Journal of Medical Entomology, 31, 105–113.CrossRefGoogle Scholar
Reichman, O. J. & Smith, S. C. (1990). Burrows and burrowing behavior by mammals. In Current Mammology, vol. 2, ed. Genoways, H. H.. New York: Plenum Press, pp. 197–244.Google Scholar
Reiczigel, J. & Rózsa, L. (1998). Host-mediated site-segregation of ectoparasites: an individual-based simulation study. Journal of Parasitology, 84, 491–498.CrossRefGoogle Scholar
Reinhold, K. (1999). Energetically costly behaviour and the evolution of resting metabolic rate in insects. Functional Ecology, 13, 217–224.CrossRefGoogle Scholar
Reiss, M. J. (1986). Sexual dimorphism in body size: are larger species more dimorphic? Journal of Theoretical Biology, 121, 163–172.
Reiss, M. J. (1989). The Allometry of Growth and Reproduction. Cambridge, UK: Cambridge University Press.
Reitblat, A. G. & Belokopytova, A. M. (1974). On cannibalism and predatory habits of flea larvae. Zoologicheskyi Zhurnal, 53, 135–137 (in Russian).Google Scholar
Řehaček, J. (1961). Transmission of tick-borne encephalitis virus by fleas. Journal of Hygiene, Epidemiology and Immunology (Prague), 5, 282–285 (in Russian).Google ScholarPubMed
Rementsova, M. M. (1962). Brucellosis in Wildlife. Alma-Ata, USSR: Kainar (in Russian).Google Scholar
Rendell, W. B. & Verbeek, N. A. M. (1996). Old nest material in nest boxes of tree swallows: effects on nest-site choice and nest building. Auk, 113, 319–328.CrossRefGoogle Scholar
Rensch, B. (1960). Evolution above the Species Level. New York: Columbia University Press.Google Scholar
Reshetnikova, P. I. (1959). Flea fauna of the Kustanai Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 6, 261–265 (in Russian).Google Scholar
Ribeiro, J. M. C. (1987). Role of saliva in blood feeding in arthropods. Annual Review of Entomology, 32, 463–478.CrossRefGoogle ScholarPubMed
Ribeiro, J. M. C. (1995). Blood-feeding arthropods: live syringes or invertebrate pharmacologists?Infectious Agents and Disease – Reviews Issues and Commentary, 4, 143–152.Google ScholarPubMed
Ribeiro, J. M. C. (1996). Common problems of arthropod vectors of disease. In The Biology of Disease Vectors, ed. Beaty, B. J. & Marquardt, W. C.. Niwot, CO: University of Colorado Press, pp. 25–33.Google Scholar
Ribeiro, J. M. C., Vaughan, J. A. & Farhang-Azad, A. (1990). Characterization of the salivary apyrase activity of three rodent flea species. Comparative Biochemistry and Physiology B, 95, 215–218.CrossRefGoogle ScholarPubMed
Richards, P. A. & Richards, A. G. (1969). Acanthae: a new type of cuticular process in the proventriculus of Mecoptera and Siphonaptera. Zoologische Jahrbücher, Abteilung für Anatomie und Ontogenie der Tiere, 86, 158–176.Google Scholar
Richner, H. (1996). Flohzirkus im Vogelnest: Wirt–Parasiten-Interaktionen in der Brutzeit. Ornithologische Beobachter, 93, 103–110.Google Scholar
Richner, H. (1998). Host–parasite interactions and life-history evolution. Zoology, 101, 333–344.Google Scholar
Richner, H. & Heeb, P. (1995). Are clutch and brood size patterns in birds shaped by ectoparasites?Oikos, 73, 435–441.CrossRefGoogle Scholar
Richner, H., Oppliger, A. & Christe, P. (1993). Effect of an ectoparasite on reproduction in great tits. Journal of Animal Ecology, 62, 703–710.CrossRefGoogle Scholar
Ricotta, C. (2004). A parametric diversity measure combining the relative abundances and taxonomic distinctiveness of species. Diversity and Distributions, 10, 143–146.CrossRefGoogle Scholar
Riddoch, B. J., Greenwood, M. T. & Ward, R. D. (1984). Aspects of the population structure of the sand martin flea, Ceratophyllus styx, in Britain. Journal of Natural History, 18, 475–484.CrossRefGoogle Scholar
Riek, E. F. (1970). Lower Cretaceous fleas. Nature, 227, 746–747.CrossRefGoogle ScholarPubMed
Ritter, R. C. & Epstein, A. N. (1974). Saliva lost by grooming: major item in rats' water economy. Behavioral Biology, 11, 581–585.CrossRefGoogle Scholar
Robbins, R. G. & Faulkenberry, G. D. (1982). A population model for fleas of the gray-tailed vole, Microtus canicaudus Miller. Entomological News, 93, 70–74.Google Scholar
Roberts, M. G., Smith, G. & Grenfell, B. T. (1995). Matematical models for macroparasites of wildlife. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 177–208.CrossRefGoogle Scholar
Robson, D. S. (1972). Appendix: Statistical tests of significance. Journal of Theoretical Biology, 34, 350–352.Google Scholar
Rödl, P. (1979). Investigation of the transfer of fleas among small mammals using radioactive phosphorus. Folia Parasitologica, 26, 265–274.Google ScholarPubMed
Rodrigues, A. F. S. F. & Daemon, E. (2004). Ixodideos e sifonapteros em Cerdocyon thous L. (Carnivora, Canidae) procedentes da zona da Mata Mineira, Brasil.Arquivos do Instituto Biologico São Paulo, 71, 371–372.Google Scholar
Rodríguez, Z., Moreira, E. C., Linardi, P. M. & Santos, H. A. (1999). Notes on the bat flea Hormopsylla fosteri (Siphonaptera: Ischnopsyllidae) infesting Molossops abrasus (Chiroptera). Memórias do Instituto Oswaldo Cruz, 94, 727–728.CrossRefGoogle Scholar
Rodríguez-Gironés, M. A. & Santamaría, L. (2006). A new algorithm to calculate the nestedness temperature of presence–absence matrices. Journal of Biogeography, 33, 924–935.CrossRefGoogle Scholar
Roehrig, J. T., Piesman, J., Hunt, A. R., et al. (1992). The hamster immune-response to tick-transmitted Borrelia burgdorferi differs from the response to needle-inoculated, cultured organisms. Journal of Immunology, 149, 3648–3653.Google ScholarPubMed
Rogovin, K. A., Randall, J., Kolosova, I. & Moshkin, M. (2003). Social correlates of stress in adult males of the great gerbil, Rhombomys opimus, in years of high and low population densities. Hormones and Behavior, 43, 132–139.CrossRefGoogle ScholarPubMed
Rogowitz, G. L. & Chappell, M. A. (2000). Energy metabolism of eucalyptus-boring beetles at rest and during locomotion: gender makes a difference. Journal of Experimental Biology, 203, 1131–1139.Google ScholarPubMed
Rohde, K. (1979). A critical evaluation of intrinsic and extrinsic factors responsible for niche restriction in parasites. American Naturalist, 114, 648–671.CrossRefGoogle Scholar
Rohde, K. (1985). Increased viviparity of marine parasites at high latitudes. Hydrobiologia, 137, 197–201.CrossRefGoogle Scholar
Rohde, K. (1992). Latitudinal gradients in species diversity: the search for the primary cause. Oikos, 65, 514–527.CrossRefGoogle Scholar
Rohde, K. (1994). Niche restriction in parasites: proximate and ultimate causes. Parasitology, 109, S69–S84.CrossRefGoogle ScholarPubMed
Rohde, K. (1996). Rapoport's rule is a local phenomenon and cannot explain latitudinal gradients in species diversity. Biodiversity Letters, 3, 10–13.CrossRefGoogle Scholar
Rohde, K. (1998). Is there a fixed number of niches for endoparasites of fish?International Journal for Parasitology, 28, 1861–1865.CrossRefGoogle Scholar
Rohde, K. (1999). Latitudinal gradients in species diversity and Rapoport's rule re-visited: a review of recent work and what can parasites teach us about the causes of the gradients?Ecography, 22, 593–613.CrossRefGoogle Scholar
Rohde, K., Hayward, C. & Heap, M. (1995). Aspects of the ecology of metazoan ectoparasites of marine fishes. International Journal for Parasitology, 25, 945–970.CrossRefGoogle ScholarPubMed
Rohde, K., Worthen, W., Heap, M., Hugueny, B. A. & Guégan, J.-F. (1998). Nestedness in assemblages of metazoan ecto- and endoparasites of marine fish. International Journal for Parasitology, 28, 543–549.CrossRefGoogle ScholarPubMed
Rolff, J. (2002). Bateman's principle and immunity. Proceedings of the Royal Society of London B, 269, 867–872.CrossRefGoogle ScholarPubMed
Roman, E. & Pichot, J. (1975). Fleas of mammals in bird nests during winter. Bulletin Mensuel de la Société Linnéenne de Lyon, 44, 53–57.CrossRefGoogle Scholar
Ronquist, F. (1995). Reconstructing the history of host–parasite associations using generalized parsimony. Cladistics, 11, 73–89.CrossRefGoogle Scholar
Ronquist, F. (2001). TreeFitter, Version 1.0. Available online at www.ebc.uu.se/systzoo/research/treefitter/treefitter.html.Google Scholar
Ronquist, F. & Liljeblad, J. (2001). Evolution of the gall wasp–host plant association. Evolution, 55, 2503–2522.Google ScholarPubMed
Roper, T. J., Ostler, J. R., Schmid, T. K. & Christian, S. F. (2001). Sett use in European badgers Meles meles. Behaviour, 138, 173–187.CrossRefGoogle Scholar
Roper, T. J., Jackson, T. P., Conradt, L. & Bennett, N. C. (2002). Burrow use and the influence of ectoparasites in Brants' whistling rat Parotomys brantsii. Ethology, 108, 557–564.CrossRefGoogle Scholar
Rosenfeld, J. S. (2002). Functional redundancy in ecology and conservation. Oikos, 98, 156–162.CrossRefGoogle Scholar
Rosenzweig, M. L. (1981). A theory of habitat selection. Ecology, 62, 327–335.CrossRefGoogle Scholar
Rosenzweig, M. L. (1989). Habitat selection as a source of biological diversity. Evolutionary Ecology, 1, 315–330.CrossRefGoogle Scholar
Rosenzweig, M. L. (1991). Habitat selection and population interactions: the search for mechanism. American Naturalist, 137, 5–28.CrossRefGoogle Scholar
Rosenzweig, M. L. (1992). Species diversity gradients: we know more and less than we thought. Journal of Mammalogy, 73, 715–730.CrossRefGoogle Scholar
Rosenzweig, M. L. (1995). Species Diversity in Space and Time. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Rosenzweig, M. L. & Abramsky, Z. (1985). Detecting density-dependent habitat selection. American Naturalist, 126, 405–417.CrossRefGoogle Scholar
Rosenzweig, M. L. & Ziv, Y. (1999). The echo pattern of species diversity: pattern and processes. Ecography, 22, 614–628.CrossRefGoogle Scholar
Rosenzweig, M. L., Abramsky, Z., Kotler, B. P. & Mitchell, W. A. (1985). Can interaction coefficients be determined from census data?Oecologia, 66, 194–198.CrossRefGoogle Scholar
Rossin, A. & Malizia, A. I. (2002). Relationship between helminth parasites and demographic attributes of a population of the subterranean rodent Ctenomys talarum (Rodentia: Octodontidae). Journal of Parasitology, 88, 1268–1270.CrossRefGoogle Scholar
Rothschild, N. C. (1915a). A synopsis of the British Siphonaptera. Entomological Monthly Magazine, 51, 49–112.Google Scholar
Rothschild, N. C. (1915b). On Neopsylla and some allied genera of Siphonaptera. Ectoparasites, 1, 30–44.Google Scholar
Rothschild, M. (1965a). Fleas. Scientific American, 213, 44–53.CrossRefGoogle Scholar
Rothschild, M. (1965b). The rabbit flea and hormones. Endeavour, 24, 162–168.Google Scholar
Rothschild, M. (1969). Notes of fleas: with the first record of the mermithid nematode from the order. Proceedings of the British Entomological and Natural History Society, 1, 1–8.Google Scholar
Rothschild, M. (1973). Note given at meeting on Pulex irritans found in Viking pit excavations. Proceedings of the Royal Entomological Society of London A, 38, 29.Google Scholar
Rothschild, M. (1975). Recent advances in our knowledge of the order Siphonaptera. Annual Review of Entomology, 20, 241–259.CrossRefGoogle ScholarPubMed
Rothschild, M. (1992). Neosomy in fleas, and the sessile life-style. Journal of Zoology, 226, 613–629.CrossRefGoogle Scholar
Rothschild, M. & Clay, T. (1952). Fleas, Flukes and Cuckoos: A Study of Bird Parasites, 3rd edn. London: Collins.Google Scholar
Rothschild, M. & Ford, R. (1966). Hormones of the vertebrate host controlling ovarian regression and copulation of the rabbit flea. Nature, 211, 261–266.CrossRefGoogle ScholarPubMed
Rothschild, M. & Ford, R. (1969). Does a pheromone-like factor from the nestling rabbit stimulate impregnation and maturation in the rabbit flea?Nature, 221, 1169–1170.CrossRefGoogle ScholarPubMed
Rothschild, M. & Ford, R. (1972). Breeding cycle of the flea Cediopsylla simplex is controlled by breeding cycle of host. Science, 178, 625–626.CrossRefGoogle ScholarPubMed
Rothschild, M. & Ford, R. (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). Journal of Zoology, 170, 87–137.CrossRefGoogle Scholar
Rothschild, M. & Hinton, H. E. (1968). Holding organs on the antennae of male fleas. Proceedings of the Royal Entomological Society of London A, 43, 105–107.CrossRefGoogle Scholar
Rothschild, M. & Neville, C. (1967). Fleas: insects which fly with their legs. Proceedings of the Royal Entomological Society of London C, 32, 9–10.Google Scholar
Rothschild, M. & Schlein, J. (1975). The jumping mechanism of Xenospylla cheopis. I. Exoskeletal structures and musculature. Philosophical Transactions of the Royal Society of London B, 271, 457–489.CrossRefGoogle Scholar
Rothschild, M., Ford, B. & Hughes, M. (1970). Maturation of the rabbit flea (Spilopsyllus cuniculi) and the oriental flea (Xenopsylla cheopis): some effects of mammalian hormones on development and impregnation. Transactions of the Zoological Society of London, 32, 105–188.Google Scholar
Rothschild, M., Schlein, J., Parker, K. & Sternberg, S. (1972). Jump of the oriental rat flea Xenopsylla cheopis (Roths.). Nature, 239, 45–48.CrossRefGoogle Scholar
Rothschild, M., Schlein, J., Parker, K., Neville, C. & Sternberg, S. (1973). The flying leap of the flea. Scientific American, 229, 92–100.CrossRefGoogle Scholar
Rothschild, M., Schlein, J., Parker, K., Neville, C. & Sternberg, S. (1975). The jumping mechanism of Xenospylla cheopis. III. Execution of the jump and activity. Philosophical Transactions of the Royal Society of London B, 271, 499–515.CrossRefGoogle Scholar
Roulin, A., Jeanmonod, J. & Blanc, T. (1997). Green plant material on common buzzard's (Buteo buteo) nests during the rearing of chicks. Alauda, 65, 251–257.Google Scholar
Roulin, A., Riols, C., Dijkstra, C. & Ducrest, A. (2001). Female plumage spottiness and parasite resistance in the barn owl (Tyto alba). Behavioral Ecology, 12, 103–110.CrossRefGoogle Scholar
Rousset, F., Thomas, F., MeeÛs, T. & Renaud, F. (1996). Inference of parasite-induced host mortality from distribution of parasite loads. Ecology, 77, 2203–2211.CrossRefGoogle Scholar
Roy, B. A. (2001). Patterns of association between crucifers and their flower-mimic pathogens: host jumps are more common than coevolution or cospeciation. Evolution, 55, 41–53.CrossRefGoogle ScholarPubMed
Rózsa, L., Rekasi, J. & Reiczigel, J. (1996). Relationship of host coloniality to the population ecology of avian lice (Insecta: Phthiraptera). Journal of Animal Ecology, 65, 242–248.CrossRefGoogle Scholar
Rudenchik, Y. V., Soldatkin, I. S., Klimova, Z. I. & Severova, Z. A. (1967). On the relationships between abundance of fleas and abundance of the great gerbils. In Proceedings of the 5th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 181–183 (in Russian).Google Scholar
Rudolph, D. & Knülle, W. (1982). Novel uptake systems for atmospheric water vapor among insects. Journal of Experimental Zoology, 222, 321–333.CrossRefGoogle Scholar
Ruffer, D. G. (1965). Burrows and burrowing behavior of Onychomys leucogaster. Journal of Mammalogy, 46, 241–247.CrossRefGoogle Scholar
Rust, M. K. (1992). Influence of photoperiod on egg production of cat fleas (Siphonaptera: Pulicidae) infesting cats. Journal of Medical Entomology, 29, 242–245.CrossRefGoogle ScholarPubMed
Rust, M. K. (1994). Interhost movement of adult cat fleas (Siphonaptera: Pulicidae). Journal of Medical Entomology, 31, 486–489.CrossRefGoogle Scholar
Ryba, J., Rodl, P., Bartos, L., Daniel, M. & Cerny, V. (1986). Some features of the ecology of fleas inhabiting the nests of the suslik (Citellus citellus (L.). I. Population dynamics, sex ratio, feeding, reproduction. Folia Parasitologica, 33, 265–275.Google ScholarPubMed
Ryba, J., Rodl, P., Bartos, L., Daniel, M. & Cerny, V. (1987). Some features of the ecology of fleas inhabiting the nests of the suslik (Citellus citellus (L.). II. The influence of mesostigmatid mites on fleas. Folia Parasitologica, 34, 61–68.Google Scholar
Ryckman, R. E. (1971). Plague vector studies. I. The rate of transfer of fleas among Citellus, Rattus and Sylvilagus under field conditions in southern California. Journal of Medical Entomology, 8, 535–540.CrossRefGoogle ScholarPubMed
Rytkönen, S., Lehtonen, R. & Orell, M. (1998). Breeding great tits Parus major avoid nestboxes infested with fleas. Ibis, 140, 687–690.CrossRefGoogle Scholar
Rzhevskaya, A. E., Rapoport, L. P., Orlova, L. M., Nuriev, K. K. & Suslova, L. P. (1991). Fleas in the eastern Kyzylkum Desert and their epizootic importance. Parazitologiya, 25, 504–511 (in Russian).Google Scholar
Sabilaev, A. S., Davydova, V. N. & Pole, D. S. (2003). About rodents' flea fauna on the left bank of the Ili River. Quarantinable and Zoonotic Infections in Kazakhstan, 7, 148–149 (in Russian).Google Scholar
Saino, N., Calza, S. & M⊘ller, A. P. (1998). Effects of a dipteran ectoparasite on immune response and growth trade-offs in barn swallow, Hirundo rustica, nestlings. Oikos, 81, 217–228.CrossRefGoogle Scholar
Sakaguti, K. & Jameson, E. W. (1962). The Siphonaptera of Japan. Pacific Insects Monographs, 3, 1–169.Google Scholar
Salas, V. & Herrera, E. A. (2004). Intestinal helminths of capybaras, Hydrochoerus hydrochaeris, from Venezuela. Memórias do Instituto Oswaldo Cruz, 99, 563–566.CrossRefGoogle ScholarPubMed
Salkeld, D. J. & Stapp, P. (2006). Seroprevalence rates and transmission of plague (Yersinia pestis) in mammalian carnivores. Vector-Borne and Zoonotic Diseases, 6, 231–239.CrossRefGoogle Scholar
Samarina, G. P., Alekseev, A. N. & Shiranovich, P. I. (1968). Study of fecundity of the rat fleas (Xenopsylla cheopis Rothschild and Ceratophyllus fasciatus Bosc.) when fed on different host species. Zoologicheskyi Zhurnal, 47, 261–268 (in Russian).Google Scholar
Samurov, M. A. (1985). Life history and the prognosis of abundance of fleas parasitizing gerbils of the genus Meriones in the Volga–Ural Sands. Unpublished Ph.D. thesis, Institute of Zoology of Academy of Science of the Kazakh SSR, Alma-Ata, USSR (in Russian).Google Scholar
Samurov, M. A. & Ageyev, V. S. (1983). Annual number of generations of fleas Xenopsylla conformis Wagn. (Siphonaptera, Pulicidae) in the Volga–Ural Sands. Entomological Review, 62, 226–269 (in Russian).Google Scholar
Samurov, M. A. & Yakunin, B. M. (1979). On the annual number of generations of fleas Ceratophyllus laeviceps in the Volga–Ural Sands. In Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 1, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 128–130 (in Russian).Google Scholar
Samurov, M. A. & Yakunin, B. M. (1980). Age composition and cycle of reproductive activity in female fleas Ceratophyllus laeviceps Wagn. (Siphonaptera) in the Volga–Ural Sands. Entomological Review, 59, 510–512 (in Russian).Google Scholar
Sapegina, V. F. (1976). Distribution of fleas parasitic on small mammals and birds in the southern taiga of the Pri–Irtyshie Region. Parazitologiya, 10, 397–400 (in Russian).Google Scholar
Sapegina, V. F. (1988). Fleas of small mammals and birds in the forest-park area of the city of Novosibirsk. Parazitologiya, 22, 132–136 (in Russian).Google ScholarPubMed
Sapegina, V. F. & Kharitonova, N. N. (1969). On the ability of bird fleas to transmit virus of the Omsk haemorhaggic fever in the experiment. In Migrating Birds and their Role in the Circulation of Arboviruses, ed. Tcherepanov, A. I.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 263–267 (in Russian).Google Scholar
Sapegina, V. F., Yudin, B. S. & Dudareva, G. V. (1980a). Materials on the biology of fleas of the Taimyr and Gydanskyi Penunsulae. In Parasitic Insects and Ticks of Siberia, ed. Davydova, M. S.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 225–231 (in Russian).Google Scholar
Sapegina, V. F., Ravkin, Y. S., Lukianova, I. V. & Sebeleva, G. G. (1980b). Fleas of the forest zone of western Siberia. In Problems of Zoogeography and Faunal History, ed. Belyshev, B. F. & Ravkin, Y. S.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 94–166 (in Russian).Google Scholar
Sapegina, V. F., Lukianova, I. V. & Fomin, B. N. (1981a). Fleas of small mammals in northern foothills of the Altai Mountains and Upper Ob River Region. In Biological Problems of Natural Foci, ed. Maximov, A. A.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 167–176 (in Russian).Google Scholar
Sapegina, V. F., Yudin, B. S. & Yudina, S. A. (1981b). Fleas of small mammals in the northern taiga of the southern Taimyr Peninsula. Bulletin of the Siberian Branch of the Academy of Sciences of the USSR, Series Biological Sciences, 1, 96–104 (in Russian).Google Scholar
Sapegina, V. F., Vartapetov, L. G. & Pokrovskaya, I. V. (1990). The fleas of small mammals in the northern taiga of western Siberia. Parazitologiya, 24, 56–62 (in Russian).Google ScholarPubMed
Sarfati, M., Krasnov, B. R., Ghazaryan, L., et al. (2005). Energy costs of blood digestion in a host-specific haematophagous parasite. Journal of Experimental Biology, 208, 2489–2496.CrossRefGoogle Scholar
Sasal, P., Trouvé, S., Müller-Graf, C. & Morand, S. (1999). Specificity and host predictability: a comparative analysis among monogenean parasites of fish. Journal of Animal Ecology, 68, 437–444.CrossRefGoogle Scholar
Saxena, V. K. (1987). Rodent–ectoparasite association in selected biotopes of Mirzapur and Varanasi districts of Uttar Pradesh. Journal of Communicable Diseases, 19, 310–316.Google ScholarPubMed
Saxena, V. K. (1999). Mesostigmatid mite infestations of rodents in diverse biotopes of central and southern India. Journal of Parasitology, 85, 147–149.CrossRefGoogle ScholarPubMed
Scalon, O. I. (1981). On fleas from eastern Mongolia with description of male and female of Echidnophaga tiscadaea Smit, 1967 (Siphonaptera). Parazitologiya, 15, 280–287 (in Russian).Google Scholar
Schall, J. J. (1990). The ecology of lizard malaria. Parasitology Today, 6, 264–269.CrossRefGoogle ScholarPubMed
Scharf, W. C. (1991). Geographic distribution of Siphonaptera collected from small mammals on Lake Michigan islands. Great Lakes Entomologist, 24, 39–43.Google Scholar
Scharf, W. C. (1998). Fleas (Siphonaptera) from migrating owls: passengers on the journey. Michigan Birds and Natural History, 5, 167–171.Google Scholar
Scheidt, V. J. (1988). Flea allergy dermatitis. In The Veterinary Clinics of North America: Small Animal Practice, ed. White, S. D.. Philadelphia, PA: Saunders College Publishing, pp. 1023–1042.Google Scholar
Schelhaas, D. P. & Larson, O. R. (1989). Cold hardiness and winter survival in the bird flea, Ceratophyllus idius. Journal of Insect Physiology, 35, 149–153.CrossRefGoogle Scholar
Schemmer, K. R. & Halliwell, R. E. (1987). Efficacy of alum-precipitated flea antigen for hyposensitization of flea-allergic dogs. Seminars in Veterinary Medicine and Surgery (Small Animal), 2, 195–198.Google ScholarPubMed
Schlein, Y. (1980). Morphological similarities between the skeletal structures of Siphonaptera and Mecoptera. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 359–367.Google Scholar
Schluter, D. (1984). A variance test for detecting species associations, with some example applications. Ecology, 65, 998–1005.CrossRefGoogle Scholar
Schmid-Hempel, P. (2003). Variation in immune defence as a question of evolutionary ecology. Proceeding of the Royal Society of London B, 270, 357–366.CrossRefGoogle ScholarPubMed
Schmid-Hempel, P. & Ebert, D. (2003). On the evolutionary ecology of specific immune defence. Trends in Ecology and Evolution, 18, 27–32.CrossRefGoogle Scholar
Schmidt-Nielsen, K. (1990). Animal Physiology: Adaptation and Environment, 4th edn. Cambridge, UK: Cambridge University Press.Google Scholar
Schoener, T. W. (1974). Competition and the form of the habitat shift. Theoretical Population Biology, 6, 265–307.CrossRefGoogle ScholarPubMed
Schofield, S. & Torr, S. J. (2002). A comparison of feeding behaviour of tsetse and stable flies. Medical and Veterinary Entomology, 16, 177–185.CrossRefGoogle ScholarPubMed
Schönrogge, K., Gardner, M. G., Elmes, G. W., et al. (2006). Host propagation permits extreme local adaptation in a social parasite of ants. Ecological Letters, 9, 1032–1040.CrossRefGoogle Scholar
Schradin, C. & Pillay, N. (2005). Demography of the striped mouse (Rhabdomys pumilio) in the succulent karoo. Mammalian Biology, 70, 84–92.CrossRefGoogle Scholar
Schwan, T. G. (1975). Flea reinfestation on the California meadow vole (Microtus californicus). Journal of Medical Entomology, 21, 760.CrossRefGoogle Scholar
Schwan, T. G. (1986). Seasonal abundance of fleas (Siphonaptera) on grassland rodents in Lake Nakuru National Park, Kenya, and potential for plague transmission. Bulletin of Entomological Research, 76, 633–648.CrossRefGoogle Scholar
Schwan, T. G. (1993). Sex ratio and phoretic mites of fleas (Siphonaptera, Pulicidae and Hystrichopsyllidae) on the Nile grass rat (Arvicanthis niloticus) in Kenya. Journal of Medical Entomology, 30, 122–135.CrossRefGoogle Scholar
Schwan, T. G. & Schwan, V. R. (1980). Observations on the fleas Xenopsylla debilis and Xenopsylla difficilis (Siphonaptera) infesting the gerbil Tatera nigricauda in Southern Kenya. African Journal of Ecology, 18, 267–272.CrossRefGoogle Scholar
Scofield, A., Riera, M. D. F., Eliseri, C., Massard, C. L. & Linardi, P. M. (2005). Ocorrência de Rhopalopsyllus lutzi lutzi (Baker) (Siphonaptera, Rhopalopsyllidae) em Canis familiaris (Linnaeus) de zona rural do município de Piraí, Rio de Janeiro, Brasil. Revista Brasileira de Entomologia, 49, 159–161.CrossRefGoogle Scholar
Seal, S. C. & Bhattacharji, L. M. (1961). Epidemiological studies of plague in Calcutta. I. Bionomics of two species of rat fleas and distribution, densities and resistance of rodents in relation to the epidemiology of plague in Calcutta. Indian Journal of Medical Research, 49, 974–1007.Google Scholar
Segerman, J. (1995). Siphonaptera of Southern Africa: handbook for the identification of fleas. Publications of the South African Institute for Medical Research, 57, 1–264.Google Scholar
Serzhan, O. S. (2002). Pathways of evolution of the faunistic complexes of rodent fleas in Kazakhstan, Middle Asia and adjacent regions and their role in the endemism of plague. Quarantinable and Zoonotic Diseases in Kazakhstan, 6, 83–90 (in Kazakh).Google Scholar
Serzhan, O. S. & Ageyev, V. S. (2000). Geographical distribution and host complexes of plague-infected fleas in relation to some problems of paleogenesis of plague enzootics. Quarantinable and Zoonotic Diseases in Kazakhstan, 2, 183–192 (in Russian).Google Scholar
Sevenster, J. G. (1996). Aggregation and coexistence. I. Theory and analysis. Journal of Animal Ecology, 65, 297–307.CrossRefGoogle Scholar
Sgonina, K. (1935). Die Reizphysiologie des Igelflöhs (Archaeopsylla erinacei Bouché) und seiner Larve. Zeitschrift für Parasitenkunde, 7, 539–571.CrossRefGoogle Scholar
Shafi, M. M., Ali, R., Ghazi, R. R. & Noor, U. N. (1988). Flea index studies of synanthropic rats in Karachi, Pakistan. Acta Parasitologica Polonica, 33, 185–194.Google Scholar
Shaftesbury, A. D. (1934). The Siphonaptera (fleas) of North Carolina, with special reference to sex ratios. Journal of the Elisha Mitchell Scientific Society, 49, 247–263.Google Scholar
Shargal, E., Kronfeld-Schor, N. & Dayan, T. (2000). Population biology and spatial relationships of coexisting spiny mice (Acomys) in Israel. Journal of Mammalogy, 81, 1046–1052.2.0.CO;2>CrossRefGoogle Scholar
Sharif, M. (1949). Effects of constant temperature and humidity on the development of the larvae and the pupae of the three Indian species of Xenopsylla (Insecta: Siphonaptera). Philosophical Transactions of the Royal Society of London B, 233, 581–633.CrossRefGoogle Scholar
Shaw, D. J. & Dobson, A. P. (1995). Patterns of macroparasite abundance and aggregation in wildlife populations: a quantitative review. Parasitology, 111, S111–S127.CrossRefGoogle ScholarPubMed
Shaw, D. J., Grenfell, B. T. & Dobson, A. P. (1998). Patterns of macroparasite aggregation in wildlife host populations. Parasitology, 117, 597–610.CrossRefGoogle ScholarPubMed
Shaw, J. L. & Moss, R. (1990). Effect of the caecal nematode Trichostrongylus tenius on egg-laying by captive red grouse. Research in Veterinary Science, 48, 253–258.Google Scholar
Shaw, S. E., Kenny, M. J., Tasker, S. & Birtles, R. J. (2004). Pathogen carriage by the cat flea Ctenocephalides felis (Bouché) in the United Kingdom. Veterinary Microbiology, 102, 183–188.CrossRefGoogle ScholarPubMed
Shchedrin, V. I. (1974). Morphological and histochemical data on the blood digestion in some flea species: vectors of plague. Unpublished Ph.D. thesis, All-Union Scientific Anti-Plague Institute ‘Microb’, Saratov, USSR (in Russian).Google Scholar
Shchedrin, V. I., Loktev, N. A. & Lunina, E. A. (1974). Morphological and histochemical data on digestion in fleas P. irritans. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 281–283 (in Russian).Google Scholar
Sheldon, B. C. & Verhulst, S. (1996). Ecological immunology: costly parasite defenses and trade-offs in evolutionary ecology. Trends in Ecology and Evolution, 11, 317–321.CrossRefGoogle Scholar
Shenbrot, G. I. & Krasnov, B. R. (2002). Can interaction coefficients be determined from census data? Testing two estimation methods with Negev Desert rodents. Oikos, 99, 47–58.CrossRefGoogle Scholar
Shenbrot, G. I., Krasnov, B. R. & Khokhlova, I. S. (1994). On the biology of Gerbillus henleyi (Rodentia: Gerbillidae) in the Negev Highlands, Israel. Mammalia, 58, 581–589.CrossRefGoogle Scholar
Shenbrot, G. I., Sokolov, V. E., Geptner, V. G. & Kovalskaya, Y. M. (1995). The Mammals of Russia and Adjacent Regions: Dipodoidea.Moscow, Russia: Nauka (in Russian).Google Scholar
Shenbrot, G. I., Krasnov, B. R. & Khokhlova, I. S. (1997). On the biology of Wagner's gerbil Gerbillus dasyurus (Wagner, 1842) (Rodentia: Gerbillidae) in the Negev Highlands, Israel. Mammalia, 61, 467–486.Google Scholar
Shenbrot, G. I., Krasnov, B. R. & Rogovin, K. A. (1999a). Spatial Ecology of Desert Rodent Communities. New York: Springer-Verlag.CrossRefGoogle Scholar
Shenbrot, G. I., Krasnov, B. R. & Khokhlova, I. S. (1999b). Notes on the biology of the bushy-tailed jird, Sekeetamys calurus, in the central Negev, Israel. Mammalia, 63, 374–377.Google Scholar
Shenbrot, G. I., Krasnov, B. R., Khokhlova, I. S., Demidova, T. & Fielden, L. J. (2002). Habitat-dependent differences in architecture and microclimate of the Sundevall's jird (Meriones crassus) burrows in the Negev Desert, Israel. Journal of Arid Environments, 51, 265–279.CrossRefGoogle Scholar
Shenbrot, G. I., Krasnov, B. R. & Lu, L. (2007). Geographic range size and host specificity in ectoparasites: A case study with Amphipsylla fleas and rodent hosts. Journal of Biogeography, 34, 1679–1690.CrossRefGoogle Scholar
Shepherd, R. C. H. & Edmonds, J. W. (1979). The distribution of the stickfast fleas, Echidnophaga myrmecobii Rothschild and E. perilis Jordan, on the wild rabbit, Oryctolagus cuniculus (L.). Australian Journal of Zoology, 27, 261–271.CrossRefGoogle Scholar
Shevchenko, V. L., Samurov, M. A., Kaimashnikov, V. I. & Polyakov, V. K. (1971). Some patterns of variation in the abundance of the midday jirds and fleas Xenopsylla conformis in the Volga–Ural Sands. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 449–450 (in Russian).Google Scholar
Shevchenko, V. L., Grazhdanov, A. K., Zharinova, L. K. & Andreeva, T. A. (1976). Abilities of avian fleas Frontopsylla frontalis alatau Fed., 1946 to infect rodents with plague. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 45, 49–52 (in Russian).Google ScholarPubMed
Shields, W. M. & Crook, J. R. (1987). Barn swallow coloniality: a net cost for group breeding in the Adirondacks. Ecology, 68, 1373–1386.CrossRefGoogle Scholar
Shinozaki, Y., Sh, T. iibashi, Yoshizawa, K., et al. (2004). Ectoparasites of the Pallas squirrel, Callosciurus erythraeus, introduced to Japan. Medical and Veterinary Entomology, 18, 61–63.CrossRefGoogle ScholarPubMed
Shorrocks, B. (1996). Local diversity: a problem with too many solutions. In The Genesis and Maintenance of Biological Diversity, ed. Hochberg, M., Clobert, J. & Barbault, R.. Oxford, UK: Oxford University Press, pp. 104–122.Google Scholar
Shorrocks, B. & Rosewell, J. (1986). Guild size in drosophilids: a simulation model. Journal of Animal Ecology, 55, 527–541.CrossRefGoogle Scholar
Shryock, J. A. & Houseman, R. M. (2005). A comparison of fecal protein content in male and female cat fleas, Ctenocephalides felis (Bouché) (Siphonaptera: Pulicidae). Florida Entomologist, 88, 335–337.CrossRefGoogle Scholar
Shubber, A. H., Lloyd, S. & Soulsby, E. J. L. (1981). Infection with gastrointestinal helminths: effect of lactation and maternal transfer of immunity. Parasitology Research, 65, 181–189.Google ScholarPubMed
Shudo, E. & Iwasa, Y. (2001). Inducible defense against pathogens and parasites: optimal choice among multiple options. Journal of Theoretical Biology, 209, 233–247.CrossRefGoogle ScholarPubMed
Shulov, A. & Naor, D. (1964). Experiments on the olfactory responses and host specificity of the Oriental rat flea (Xenopsylla cheopis). Parasitology, 54, 225–231.CrossRefGoogle Scholar
Shurin, J. B. & Allen, E. G. (2001). Effects of competition, predation, and dispersal on species richness at local and regional scales. American Naturalist, 158, 624–637.CrossRefGoogle ScholarPubMed
Shurin, J. B., Havel, J. E., Leibold, M. A. & Pinel-Alloul, B. (2000). Local and regional zooplankton species richness: a scale-independent test for saturation. Ecology, 81, 3062–3073.CrossRef
Shutler, D. & Campbell, A. A. (2007). Experimental addition of greenery reduces flea loads in nests of a non-greenery using species, the tree swallow Tachycineta bicolor. Journal of Avian Biology, 38, 7–12.CrossRefGoogle Scholar
Shutler, D., Petersen, S. D., Dawson, R. D. & Campbell, A. (2003). Sex ratios of fleas (Siphonaptera: Ceratophyllidae) in nests of tree swallows (Passeriformes: Hirundinidae) exposed to different chemicals. Environmental Entomology, 32, 1045–1048.CrossRefGoogle Scholar
Shutler, D., Mullie, A. & Clark, R. G. (2004). Tree swallow reproductive investment, stress, and parasites. Canadian Journal of Zoology, 82, 442–448.CrossRefGoogle Scholar
Shwartz, E. A., Berendyaeva, E. L. & Grebenyuk, R. V. (1958). Fleas parasitic on rodents in the Frunze Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 4, 255–261 (in Russian).Google Scholar
Siemann, E. (1998). Experimental tests of effects of plant productivity and diversity on grassland arthropod diversity. Ecology, 79, 2057–2070.CrossRefGoogle Scholar
Silverman, J. & Appel, A. G. (1994). Adult cat flea (Siphonaptera, Pulicidae) excretion of host blood proteins in relation to larval nutrition. Journal of Medical Entomology, 31, 265–271.CrossRefGoogle ScholarPubMed
Silverman, J. & Rust, M. K. (1983). Some abiotic factors affecting the survival of the cat flea, Ctenocephalides felis (Siphonaptera: Pulicidae). Environmental Entomology, 12, 490–495.CrossRefGoogle Scholar
Silverman, J. & Rust, M. K. (1985). Extended longevity of the pre-emerged adult of the cat flea (Siphonaptera: Pulicidae) and factors stimulating emergence from the pupal cocoon. Annals of the Entomological Society of America, 78, 763–768.CrossRefGoogle Scholar
Silverman, J., Rust, M. K. & Reierson, D. A. (1981). Influence of temperature and humidity on survival and development of the cat flea, Ctenocephalides felis (Siphonaptera: Pulicidae). Journal of Medical Entomology, 18, 78–83.CrossRefGoogle Scholar
Silvertown, J., Franco, M. & Harper, J. L. (eds.) (1997). Plant Life Histories: Ecology, Phylogeny and Evolution. New York: Cambridge University Press.Google Scholar
Simiczyjew, B. & Margas, W. (2001). Ovary structure in the bat flea Ischnopsyllus spp. (Siphonaptera: Ischnopsyllidae): phylogenetic implications. Zoologica Poloniae, 46, 5–14.Google Scholar
Šimková, A., Desdevises, Y., Gelnar, M. & Morand, S. (2000). Coexistence of nine gill ectoparasites (Dactylogyrus: Monogenea) parasitizing the roach (Rutilus rutilus L.): history and present ecology. International Journal for Parasitology, 30, 1077–1088.CrossRefGoogle Scholar
Šimková, A., Desdevises, Y., Gelnar, M. & Morand, S. (2001). Morphometric correlates of host specificity in Dactylogyrus species (Monogenea) parasites of European cyprinid fish. Parasitology, 123, 169–177.CrossRefGoogle ScholarPubMed
Šimková, A., Kadlec, D., Gelnar, M. & Morand, S. (2002). Abundance–prevalence relationship of gill congeneric ectoparasites: testing the core–satellite hypothesis and ecological specialization. Parasitology Research, 88, 682–686.Google Scholar
Šimková, A., Sitko, J., Okulewicz, J. & Morand, S. (2003). Occurrence of intermediate hosts and structure of digenean communities of the black-headed gull, Larus ridibundus (L.). Parasitology, 126, 69–78.CrossRefGoogle Scholar
Šimková, A., Verneau, O., Gelnar, M. & Morand, S. (2006). Specificity and specialization of congeneric monogeneans parasitizing cyprinid fish. Evolution, 60, 1023–1037.CrossRefGoogle ScholarPubMed
Sinelshchikov, V. A. (1956). Study of flea fauna of the Pavlodar Region. Proceedings of the Middle Asian Scientific Anti-Plague Institute, 2, 147–153 (in Russian).Google Scholar
Singh, S. K. & Girschick, H. J. (2003). Tick–host interactions and their immunological implications in tick-borne diseases. Current Science, 85, 1284–1298.Google Scholar
Skinner, J. D. & Smithers, H. N. (1990). The Mammals of the Southern African Subregion, 2nd edn. Pretoria, South Africa: University of Pretoria Press.Google Scholar
Skorping, A., Read, A. F. & Keymer, A. E. (1991). Life history covariation in intestinal nematodes of mammals. Oikos, 60, 365–372.CrossRefGoogle Scholar
Skuratowicz, W. (1960). Pchly (Aphaniptera) ptaków i ssaków Bialowieskiego Parku Narodowego. Annales Zoologici (Warszawa), 19, 1–32.Google Scholar
Skuratowicz, W. (1967). Pchly – Siphonaptera (Aphaniptera): Klucze do Oznacznia Owadow Polski 29. Warsaw: Panstwowe Wydawnictwo Naukowe.Google Scholar
Slomczyński, R., Kaliński, A., Wawrzyńiak, J., et al. (2006). Effects of experimental reduction in nest micro-parasite and macro-parasite loads on nestling hemoglobin level in blue tits Parus caeruleus. Acta Oecologica, 30, 223–227.CrossRefGoogle Scholar
Slonov, M. N. (1965). On the biology of a flea Ceratophyllus tamias Wagn., 1927. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 34, 485–487 (in Russian).Google ScholarPubMed
Smetana, A. (1965). On the transmission of tick-borne encephalitis by fleas. Acta Virologica, 9, 375–379.Google ScholarPubMed
Smit, F. G. A. M. (1954). Lopper: Danmarks Fauna, 60. Copenhagen, Denmark: G. E. C. Gads Forlag.Google Scholar
Smit, F. G. A. M. (1962a). Neotunga euloidea gen. n., sp. n. (Siphonaptera, Pulicidae). Bulletin of the British Museum of Natural History, Entomology, 12, 365–378.CrossRefGoogle Scholar
Smit, F. G. A. M. (1962b). Siphonaptera collected from moles and their nests at Wilp, Netherlands, by Jhr. W. C. Van Heurn. Tijdschrift voor Entomologie, 105, 29–44.Google Scholar
Smit, F. G. A. M. (1972). On some adaptive structures in Siphonaptera. Folia Parasitologica, 19, 5–17.Google ScholarPubMed
Smit, F. G. A. M. (1974). Siphonaptera collected by Dr J. Martens in Nepal. Senckenbergiana Biologica, 55, 357–398.Google Scholar
Smit, F. G. A. M. (1977). An unusual form of the stick-tight flea Echidnophaga gallinacea. Revue Zoologique Africaine, 91, 198–199.Google Scholar
Smit, F. G. A. M. (1978). Fossil ‘fleas’. Flea News, 14, 1–2.Google Scholar
Smit, F. G. A. M. (1982). Classification of the Siphonaptera. In Synopsis and Classification of Living Organisms, vol. 2, ed. Parker, S.. New York: McGraw-Hill, pp. 557–563.Google Scholar
Smit, F. G. A. M. (1987). An Illustrated Catalogue of the Rothschild Collection of Fleas (Siphonaptern) in the British Museum (Natural History), vol. 7, Malacopsylloidea (Malacopsyllidae and Rhopalopsyllidae). London: Oxford University Press.
Smith, A. (1951). The effect of relative humidity on the activity of the tropical rat flea Xenopsylla cheopis (Roths.). Bulletin of Entomological Research, 42, 585–600.CrossRefGoogle Scholar
Smith, A. (1980). Lack of interspecific interactions of Everglades rodents on two spatial scales. Acta Theriologica, 25, 61–70.CrossRefGoogle Scholar
Smith, A., Telfer, S., Burthe, S., Bennett, M. & Begon, M. (2005). Trypanosomes, fleas and field voles: ecological dynamics of a host–vector–parasite interaction. Parasitology, 131, 355–365.CrossRefGoogle ScholarPubMed
Smith, F. A., Brown, J. H., Haskell, J. P., et al. (2004). Similarity of mammalian body size across the taxonomic hierarchy and across space and time. American Naturalist, 163, 672–691.CrossRefGoogle ScholarPubMed
Smith, S. A. & Clay, M. E. (1985). Morphology of the antennae of the bat flea Myodopsylla insignis (Siphonaptera: Ischnopsyllidae). Journal of Medical Entomology, 22, 64–71.CrossRefGoogle Scholar
Smits, J. E., Bortolotti, G. R. & Tella, J. L. (1999). Simplifying the phytohaemagglutinin skin-testing technique in studies of avian immunocompetence. Functional Ecology, 13, 567–572.CrossRefGoogle Scholar
Snodgrass, R. E. (1944). The feeding apparatus of biting and sucking insects affecting man and animals. Smithsonian Miscellaneous Collections, 104, 1–107.Google Scholar
Sobey, W. R., Menzies, W. & Conolly, D. (1974). Myxomatosis: some observations on breeding the European rabbit flea Spilopsyllus cuniculi (Dale) in an animal house. Journal of Hygiene, 71, 453–465.CrossRefGoogle Scholar
Sokolova, A. A. & Popova, A. S. (1969). On the biology of fleas Coptopsylla lamellifer. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 90–92 (in Russian).Google Scholar
Sokolova, A. A., Balabas, N. G. & Trofimenko, I. P. (1971). Data on reproduction of Coptopsylla lamellifer in the Moyynkum Desert. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 416–417 (in Russian).Google Scholar
Soliman, S., Marzouk, A. S., Main, A. J. & Montasser, A. A. (2001). Effect of sex, size, and age of commensal rat hosts on the infestation parameters of their ectoparasites in a rural area of Egypt. Journal of Parasitology, 87, 1307–1316.CrossRefGoogle Scholar
Soloshenko, I. Z. (1958). Role of haematophagous arthropods in the maintenance of the epizootics of leptospiroses in the foci of these diseases. In Proceedings of the 10th Conference on Parasitological Problems and Diseases with Natural Focality, ed. Anonymous. Moscow–Leningrad, USSR, pp. 139–140 (in Russian).Google Scholar
Soloshenko, I. Z. (1962). Role of haematophagous arthropods in transmission and maintanence of pathogenous leptospires. II. Relationships between haematophagous arthropods and leptospiroses. Journal of Microbiology [Zhurnal Mikrobiologii], 4, 31–34 (in Russian).Google Scholar
Solovieva, A. V., Alania, I. I. & Kosminsky, R. B. (1976). On the ecology of fleas Ctenophthalmus (Euctenophthalmus) strigosus Rostigaev et Zolotova, 1964 (Ctenophthalmidae, Siphonaptera) in southern Trans-Caucasus. Problems of Particularly Dangerous Diseases, 51, 46–49 (in Russian).Google Scholar
Sorci, G. (1996). Patterns of haemogregarine load, aggregation and prevalence as a function of host age in the lizard Lacerta vivipara. Journal of Parasitology, 82, 676–678.CrossRefGoogle ScholarPubMed
Sorci, G., Defraipont, M. & Clobert, J. (1997). Host density and ectoparasite avoidance in the common lizard (Lacerta vivipara). Oecologia, 11, 183–188.CrossRefGoogle Scholar
Soshina, Y. F. (1973). The rate of flea infestation in common myomorph rodents in the forest zone of the Krym Mountains. Parazitologiya, 7, 31–35 (in Russian).Google Scholar
Sosnina, E. F. (1967a). The dependence of the infestation and species composition of rodent ectoparasites on the host's habitat (on the example of Rattus turkestanicus). Wiadomosci Parazytologiczne, 13, 637–641.Google Scholar
Sosnina, E. F. (1967b). An attempt of biocoenotical analysis of the assemblages of arthropods collected from rodents. Parasitological Collection, 23, 61–69 (in Russian).Google Scholar
Sotnikova, A. N. & Soldatov, G. M. (1969). Extraction of the virus of the tick-borne encephalitis from fleas Ceratophyllus tamias Wagn. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 33, 622–624 (in Russian).Google Scholar
Southwood, T. R. E. (1966). Ecological Methods. London: Chapman & Hall.Google Scholar
Souza, W. P. (1994). Patterns and processes in communities of helminth parasites. Trends in Ecology and Evolution, 9, 52–57.CrossRefGoogle Scholar
Sreter-Lancz, Z., Tornyai, K., Szell, Z., Sreter, T. & Marialigeti, K. (2006). Bartonella infections in fleas (Siphonaptera: Pulicidae) and lack of Bartonellae in ticks (Acari: Ixodidae) from Hungary. Folia Parasitologica, 53, 313–316.CrossRefGoogle ScholarPubMed
Srivastava, D. (1999). Using local–regional richness plots to test for species saturation: pitfalls and potentials. Journal of Animal Ecology, 68, 1–16.CrossRefGoogle Scholar
Stanko, M. (1987). Siphonaptera of small mammals in the northern part of the Krupina Plain. Stredné Slovensko, Zborník Stredoslovenského Múzea, Banská Bystrica, 6, 108–117 (in Slovak).Google Scholar
Stanko, M. (1988). Fleas (Siphonaptera) of small mammals in eastern part of the Volovské Vrchy Mountains. Acta Rerum Naturalium Musei Nationalis Slovaci Bratislava, 34, 29–40 (in Slovak).Google Scholar
Stanko, M. (1994). Fleas synusy (Siphonaptera) of small mammals from the central part of the East-Slovakian lowlands. Biologia, Bratislava, 49, 239–246.Google Scholar
Stanko, M., Miklisova, D., Goüy de Bellocq, J. & Morand, S. (2002). Mammal density and patterns of ectoparasite species richness and abundance. Oecologia, 131, 289–295.CrossRefGoogle ScholarPubMed
Stanko, M., Krasnov, B. R. & Morand, S. (2006). Relationship between host density and parasite distribution: inferring regulating mechanisms from census data. Journal of Animal Ecology, 75, 575–583.CrossRefGoogle ScholarPubMed
Stanko, M., Krasnov, B. R., Miklisova, D. & Morand, S. (2007). Simple epidemiological model predicts the relationships between prevalence and abundance in ixodid ticks. Parasitology, 134, 59–68.CrossRefGoogle ScholarPubMed
Starikov, V. P. & Sapegina, V. F. (1987). Ectoparasites of small mammals in the forest-steppe Trans-Ural Region. In Ecology and Geography of Arthropods in Siberia, ed. Tcherepanov, A. I.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 76–83 (in Russian).Google Scholar
Stark, H. E. (2002). Population dynamics of adult fleas (Siphonaptera) on hosts and in nests of the California vole. Journal of Medical Entomology, 39, 818–824.CrossRefGoogle ScholarPubMed
Stark, H. E. & Miles, V. I. (1962). Ecological studies of wild rodent plague in the San Francisco Bay area of California. VI. The relative abundance of the certain flea species and their host relationships on coexisting wild and domestic rodents. American Journal of Tropical Medicine and Hygiene, 11, 525–534.CrossRefGoogle ScholarPubMed
Starozhitskaya, G. S. (1968). Effect of gonotrophic cycle of fleas on the duration of their uninterrupted stay on a host. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 59–64 (in Russian).Google Scholar
Starozhitskaya, G. S. (1970). Diapause in fleas of the genus Xenopsylla and its epizootological importance. Problems of Particularly Dangerous Diseases, 13, 148–155 (in Russian).Google Scholar
Statzner, B., Dolédec, S. & Hugueny, B. (2004). Biological trait composition of European stream invertebrate communities: assessing the effects of various trait filter types. Ecography, 27, 470–488.CrossRefGoogle Scholar
Stearns, S. C. (1992). The Evolution of Life Histories. New York: Oxford University Press.Google Scholar
Steele, W. K., Pilgrim, R. L. C. & Palma, R. L. (1997). Occurrence of the flea Glaciopsyllus antarcticus and avian lice in central Dronning Maud Land. Polar Biology, 18, 292–294.CrossRefGoogle Scholar
Stenseth, N. C, Samia, N. I., Viljugrein, H., et al. (2006). Plague dynamics are driven by climate variation. Proceedings of the National Academy of Sciences of the USA, 103, 13110–13115.CrossRefGoogle ScholarPubMed
Stepanova, N. A. & Mitropolsky, O. V. (1971). Relationships between co-occurring fleas Xenopsylla hirtipes and Xenopsylla gerbilli. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 418–420 (in Russian).Google Scholar
Stepanova, N. A. & Mitropolsky, O. V. (1977). Spatial distribution of two sympathric species of fleas parasitic on the great gerbil in the Kyzyl-Kum Desert. Parazitologiya, 11, 147–152 (in Russian).Google Scholar
Stevens, G. C. (1989). The latitudinal gradient in geographical range: how so many species coexist in the tropics. American Naturalist, 133, 240–256.CrossRefGoogle Scholar
Stevenson, H. L., Bai, Y., Kosoy, M. Y., et al. (2003). Detection of novel Bartonella strains and Yersinia pestis in pairie dogs and their fleas (Siphonaptera: Ceratophyllidae and Pulicidae) using multiplex polymerase chain reaction. Journal of Medical Entomology, 40, 329–337.CrossRefGoogle ScholarPubMed
Stevenson, H. L., Labruna, M. B., Montenieri, J. A., et al. (2005). Detection of Rickettsia felis in a New World flea species, Anomiopsyllus nudata (Siphonaptera: Ctenophthalmidae). Journal of Medical Entomology, 42, 163–167.CrossRefGoogle Scholar
Stewart, M. A. & Evans, F. C. (1941). A comparative study of rodent and burrow flea populations. Proceedings of the Society for Experimental Biology and Medicine, 47, 140–142.CrossRefGoogle Scholar
Stewart, P. D. & MacDonald, D. W. (2003). Badgers and badger fleas: strategies and counter-strategies. Ethology, 109, 751–764.CrossRefGoogle Scholar
Stireman, J. O. (2005). The evolution of generalization? Parasitoid flies and the perils of inferring host range evolution from phylogenies. Journal of Evolutionary Biology, 18, 325–336.CrossRefGoogle ScholarPubMed
Stone, L. & Roberts, A. (1991). Conditions for a species to gain advantage from the presence of competitors. Ecology, 72, 1964–1972.CrossRefGoogle Scholar
Strahan, R. (ed.) (1983). The Complete Book of Australian Mammals. North Ryde, Australia: Collins, Angus & Robertson.Google Scholar
Streilein, J. W. (1990). Skin associated lymphoid tissues (SALT): the next generation. In Skin Immune System (SIS), ed. Bos, J. D.. Boca Raton, FL: CRC Press, pp. 25–48.Google Scholar
Studdert, V. P. & Arundel, J. H. (1988). Dermatitis of the pinnae of cats in Australia associated with the European rabbit flea (Spilopsyllus cuniculi). Veterinary Record, 123, 624–625.Google Scholar
Štys, P. & Bilinski, S. M. (1990). Ovariole types and the phylogeny of hexapods. Biology Review, 65, 401–429.CrossRefGoogle Scholar
Sukhanova, V. I., Tchernikina, M. A., Sosnovtseva, V. P., et al. (1978). Multiannual dynamics of abundance in fleas parasitic on the great gerbil in northern Turkmenistan. Problems of Particularly Dangerous Diseases, 60, 53–57 (in Russian).Google Scholar
Sukhdeo, M. V. K. (1997). Earth's third environment: the worm's eye view. BioScience, 47, 141–149.Google Scholar
Sukhdeo, M. V. K. (2000). Inside the vertebrate host: ecological strategies by parasites living in the third environment. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R., Morand, S. & Skorping, A.. Amsterdam, the Netherlands: Elsevier Science, pp. 43–62.Google Scholar
Sukhdeo, M. V. K. & Bansemir, A. D. (1996). Critical resources that influence habitat selection decisions by gastrointestinal helminth parasites. International Journal for Parasitology, 26, 483–498.CrossRefGoogle ScholarPubMed
Sukhdeo, M. V. K., Sukhdeo, S. C. & Bansemir, A. D. (2002). Interactions between intestinal nematodes and vertebrate hosts. In The Behavioural Ecology of Parasites, ed. Lewis, E. E., Campbell, J. F. & Sukhdeo, M. V. K.. Wallingford, UK: CAB International, pp. 223–242.CrossRefGoogle Scholar
Suleimenov, B. M. (2004). Mechanism of Plague Enzootic. Almaty, Kazakhstan: Almaty (in Russian).Google Scholar
Suntsov, V. V. & Suntsova, N. I. (2003). Origin and genesis of natural and anthropogenic plague foci: ecological, geographical and social aspects. In Ecological and Epizootological Aspects of Plague in Vietnam, ed. Korzun, L. P. & Suntsov, V. V.. Moscow (Russia), Ho Chi Minh, Buonmathuot (Vietnam): GEOS, pp. 109–149 (in Russian).Google Scholar
Suntsov, V. V. & Suntsova, N. I. (2006). Plague: Origin and Evolution of Epizootic System (Ecological, Geographical and Social Aspects). Moscow, Russia: KMK Scientific Press (in Russian).Google Scholar
Suntsov, V. V., Vi, Li Thi K. & Suntsova, N. I. (1992a). Some features of the flea (Insecta, Siphonaptera) fauna of small mammals in Vietnam. Zoologicheskyi Zhurnal, 71, 88–94 (in Russian).Google Scholar
Suntsov, V. V., Huong, L. T. & Suntsova, N. I. (1992b). Notes on fleas (Siphonaptera) in the plague foci on the Tay Nguyen Plateau (Vietnam). Parazitologiya, 26, 516–520 (in Russian).Google Scholar
Suter, P. R. (1964). Biologie von Echidnophaga gallinacea (Westw.) und Vergleich mit andern Verhaltenstypen bei Flöhen. Acta Tropica, 21, 193–238.Google Scholar
Sutherland, W. J. (1983). Aggregation and the ‘ideal free’ distribution. Journal of Animal Ecology, 52, 821–828.CrossRefGoogle Scholar
Sutherland, W. J. (1996). From Individual Behaviour to Population Ecology. Oxford, UK: Oxford University Press.Google Scholar
Sviridov, G. G. (1963). Application of radioactive isotopes for the study of some problems of flea ecology. II. The contact of animals and intensity of the exchange of ectoparasites in the population of Rhombomis opimus. Zoologicheskyi Zhurnal, 42, 546–550 (in Russian).Google Scholar
Syrvatcheva, N. G. (1964). Data on flea fauna of the Kabardino-Balkarian ASSR. Proceedings of the Armenian Anti-Plague Station, 3, 389–405 (in Russian).Google Scholar
Szidat, L. (1940). Beiträge zum Aubfau eines natürlichen Systems der Trematoden. I. Die Entwicklung von Echinocercaria choanophila U. Szidat zu Cathaemasia hians und die Ableitung der Fasciolidae von den Echinostomidae. Zeitschrift für Parasitenkunde, 11, 239–283.CrossRefGoogle Scholar
Tabor, S. P., Williams, D. F., Germano, D. J. & Thomas, R. E. (1993). Fleas (Siphonaptera) infesting giant kangaroo rats (Dipodomys ingens) on the Elkhorn and Carrizo plains, San Luis Obispo County, California. Journal of Medical Entomology, 30, 291–294.CrossRefGoogle Scholar
Takahashi, K., Tuno, N. & Kagaya, T. (2005). The relative importance of spatial aggregation and resource partitioning on the coexistence of mycophagous insects. Oikos, 109, 125–134.CrossRefGoogle Scholar
Talybov, A. N. (1974). Some data on the lifespan of fleas parasitic on the common vole in the Trans-Caucasus mountains. In Particularly Dangerous Diseases in Caucasus: Proceedings of the 3rd Scientific–Practical Conference of the Anti-Plague Establishments of Caucasus on Natural Focality, Epidemiology and Prophylaxis of Particularly Dangerous Diseases, 14–16 May 1974, ed. Pilipenko, V. G.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 183–185 (in Russian).Google Scholar
Talybov, A. N. (1975). Life expectancy of Ctenophthalmus wladimiri Is.-Gurv., 1948 (Siphonaptera, Ctenophthalmidae) under laboratory conditions. Parazitologiya, 9, 354–358 (in Russian).Google ScholarPubMed
Talybov, A. N. (1976). Development of the pre-imaginal phases of flea Ctenophthalmus wladimiri Is.-Gurv., 1948. Parazitologiya, 10, 320–324 (in Russian).Google Scholar
Tanitovsky, V. A., Bidashko, F. G., Grazhdanov, A. K. & Dauletova, S. B. (2004). Species structure and number of fleas parasitizing small mammals in the middle part of the Ural River valley. Quarantinable and Zoonotic Infections in Kazakhstan, 9, 76–80 (in Russian).Google Scholar
Tarasevich, L. N., Tagiltsev, A. A. & Malkov, G. B. (1969). Results of virological examination of ixodid ticks and fleas in the southern Omsk Region. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 38, 705–707 (in Russian).Google Scholar
Tarshis, I. B. (1956). Feeding techniques for blood-sucking arthropods. Proceedings of the 10th International Congress of Entomology, 3, 767–784.Google Scholar
Taylor, J. & Purvis, A. (2003). Have mammals and their chewing lice diversified in parallel? In Tangled Trees: Phylogeny, Cospeciation, and Coevolution, ed. Page, R. D. M.. Chicago, IL: University of Chicago Press, pp. 240–261.Google Scholar
Taylor, L. H., Mackinnon, M. J. & Read, A. F. (1998). Virulence of mixed-clone and single-clone infections of the rodent malaria Plasmodium chabaudi. Evolution, 52, 583–591.CrossRefGoogle ScholarPubMed
Taylor, L. R. (1961). Aggregation, variance and the mean. Nature, 189, 732–735.CrossRefGoogle Scholar
Taylor, L. R. & Taylor, R. A. J. (1977). Aggregation, migration and population dynamics. Nature, 265, 415–421.CrossRefGoogle Scholar
Taylor, L. R. & Woiwod, I. P. (1980). Temporal stability as a density-dependent species characteristic. Journal of Animal Ecology, 49, 209–224.CrossRefGoogle Scholar
Taylor, L. R., Woiwod, I. P. & Perry, J. N. (1979). The negative binomial as a dynamic ecological model and density-dependence of k. Journal of Animal Ecology, 48, 289–304.CrossRefGoogle Scholar
Taylor, R. A. J., Lindquist, R. K. & Shipp, J. L. (1998). Variation and consistency in spatial distribution as measured by Taylor's power law. Environmental Entomology, 27, 191–201.CrossRefGoogle Scholar
Taylor, S. D., Cruz, Dittmar K., Porter, M. L. & Whiting, M. F. (2005). Characterization of the long-wavelength opsin from Mecoptera and Siphonaptera: does a flea see?Molecular Biology and Evolution, 22, 1165–1174.CrossRefGoogle ScholarPubMed
Tchernova, N. A. (1971). Reproduction of Xenopsylla skrjabini and their preferences for different elements of the great gerbil burrow in the Mangyshlak Peninsula. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 443–444 (in Russian).Google Scholar
Tchimanina, B. M. & Kozlovskaya, O. L. (1971a). Experimental study of the role of fleas Ctenophthalmus congeneroides Wagn. and Neopsylla bidentatiformis Wagn. in the circulation of the tick-borne encephalitis virus. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 235–236 (in Russian).Google Scholar
Tchimanina, B. M. & Kozlovskaya, O. L. (1971b). Experimental study of the circulation of the tick-borne encephalitis virus in the nests of the forest voles via fleas Frontopsylla elata botis, Jord., 1929. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 237–238 (in Russian).Google Scholar
Tchumakova, I. V. & Kozlov, M. P. (1983). Quantitative parameters of mortality and survival in fleas Nosopsyllus consimilis at different stages of the metamorphosis. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 280–281 (in Russian).Google Scholar
Tchumakova, I. V., Tokanev, F. I. & Kozlov, M. P. (1978). Dependence of the reproduction capacity of fleas (Aphaniptera) on the recurrence of mating. Parazitologiya, 12, 292–296 (in Russian).Google Scholar
Tchumakova, I. V., Kozlov, M. P. & Belokopytova, A. (1981). Estimation of the dependence of the reproduction of rodent fleas (Siphonaptera) on feeding by experimental breeding of fleas on different hosts. Entomological Review, 60, 562–569 (in Russian).Google Scholar
Tchumakova, I. V., Ermolova, N. V. & Shaposhnikova, L. I. (2002). Principles for prediction of population densities of fleas parasitic on rodents. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 72, 45–48 (in Russian).Google Scholar
Telfer, S., Bown, K. J., Sekules, R., et al. (2005). Disruption of a host–parasite system following the introduction of an exotic host species. Parasitology, 130, 661–668.CrossRefGoogle ScholarPubMed
Tella, J. L. (2002). The evolutionary transition to coloniality promotes higher blood parasitism in birds. Journal of Evolutionary Biology, 15, 32–41.CrossRefGoogle Scholar
Tella, J. L., Scheuerlein, A. & Ricklefs, R. E. (2002). Is cell-mediated immunity related to the evolution of life-history strategies in birds?Proceedings of the Royal Society of London B, 269, 1059–1066.CrossRefGoogle ScholarPubMed
Tellam, R. L., Smith, D., Kemp, D. H. & Willadsen, P. (1992). Vaccination against ticks. In Animal Parasite Control Utilizing Biotechnology, ed. Yong, W. K.. Boca Raton, FL: CRC Press, pp. 303–331.Google Scholar
Tenquist, J. D. & Charleston, W. A. G. (2001). A revision of the annotated checklist of ectoparasites of terrestrial mammals in New Zealand. Journal of the Royal Society of New Zealand, 31, 481–542.CrossRefGoogle Scholar
Teplinskaya, T. A., Labunetz, N. F. & Kuliev, M. T. (1983). Seasonal dynamics of age structure and reproduction of Xenopsylla conformis in the Caspian natural plague focus. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 277–278 (in Russian).Google Scholar
Terborgh, J. W. & Faaborg, J. (1980). Saturation of bird communities in the West Indies. American Naturalist, 116, 178–195.CrossRefGoogle Scholar
Hofstede, H. M. & Fenton, M. B. (2005). Relationships between roost preferences, ectoparasite density and grooming behaviour of neotropical bats. Journal of Zoology, 266, 333–340.CrossRefGoogle Scholar
Theodor, O. & Costa, M. (1967). A Survey of the Parasites of Wild Mammals and Birds in Israel. vol. 1, wEctoparasites. Jerusalem: Israel Academy of Science and Humanities.Google Scholar
Théron, A. & Combes, C. (1995). Asynchrony of infection timing, habitat preference, and sympatric speciation of schistosome parasites. Evolution, 49, 372–375.CrossRefGoogle ScholarPubMed
Thomas, C. D. & Hanski, I. (1997). Butterfly metapopulations. In Metapopulation Biology: Ecology, Genetics, and Evolution, ed. Hanski, I. & Gilpin, M. E.. San Diego, CA: Academic Press, pp. 359–386.Google Scholar
Thomas, K. & Shutler, D. (2001). Ectoparasites, nestling growth, parental feeding rates, and begging intensity of tree swallows. Canadian Journal of Zoology, 79, 346–353.CrossRefGoogle Scholar
Thomas, R. (1988). A review of flea collection records from Onychomys leucogaster with observations on the role of grasshopper mice in the epizootiology of wild rodent plague. Great Basin Naturalist, 48, 83–95.Google Scholar
Thomas, R. (1996). Fleas and the agents they transmit. In The Biology of Disease Vectors, ed. Beaty, B. J. & Marquardt, W. C.. Niwot, CO: University of Colorado Press, pp. 146–159.Google Scholar
Thompson, C. F. & Neill, A. J. (1991). House wrens do not prefer clean nestboxes. Animal Behaviour, 42, 1022–1024.CrossRefGoogle Scholar
Thompson, C. W., Hillgarth, N., Leu, M. & McClure, H. E. (1997). High parasite load in house finches (Carpodacus mexicanus) is correlated with reduced expression of a sexually selected trait. American Naturalist, 149, 270–294.CrossRefGoogle Scholar
Thompson, J. N. (1994). The Coevolutionary Process. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Thompson, J. N. (2005). The Geographic Mosaic of Coevolution. Chicago, IL: University of Chicago Press.Google Scholar
Tian, J. (1995). Niches of 27 flea species in the natural focus of plague in Jianchuan, Yunnan Province. Endemic Diseases Bulletin, 10, 27–32 (in Chinese).Google Scholar
Tiflov, V. E. (1959). Role of fleas in epizootology of tularemia. Proceedings of the Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, 2, 363–392 (in Russian).Google Scholar
Tiflov, V. E. (1964). Destiny of the bacterial cultures in the organism of a flea. Ectoparasites, 4, 181–198 (in Russian).Google Scholar
Tiflov, V. E. & Ioff, I. G. (1932). Observations on the biology of fleas. Herald of Microbiology and Epidemiology [Vestnik Mikrobiologii i Epidemiologii], 11, 95–117 (in Russian).Google Scholar
Tillyard, R. J. (1926). The Insects of Australia and New Zealand. Sydney Australia: Angus and Robertson.Google Scholar
Tillyard, R. J. (1935). The evolution of the scorpion-flies and their derivatives (order Mecoptera). Annals of the Entomological Society of America, 28, 1–45.CrossRefGoogle Scholar
Timi, J. T. & Poulin, R. (2003). Parasite community structure within and across host populations of a marine pelagic fish: how repeatable is it?International Journal for Parasitology, 33, 1353–1362.CrossRefGoogle Scholar
Tipton, V. J. & Machado-Allison, C. E. (1972). Fleas of Venezuela. Brigham Young University Scientific Bulletin, Biological Series, 17, 1–115.Google Scholar
Tipton, V. J. & Méndez, E. (1966). The fleas (Siphonaptera) of Panama. In Ectoparasites of Panama, ed. Wenzel, R. L. & Tipton, V. J.. Chicago, IL: Field Museum of Natural History, pp. 289–385.Google Scholar
Tipton, V. J. & Méndez, E. (1968). New species of fleas (Siphonaptera) from Cerro Potosi, Mexico, with notes on ecology and host–parasite relationships. Pacific Insects, 10, 177–214.Google Scholar
Toft, C. A. & Karter, A. J. (1990). Parasite–host coevolution. Trends in Ecology and Evolution, 5, 326–329.CrossRefGoogle ScholarPubMed
Tofts, R. & Silvertown, J. (2000). A phylogenetic approach to community assembly from a local species pool. Proceedings of the Royal Society of London B, 267, 363–369.CrossRefGoogle ScholarPubMed
Tokeshi, M. (1999). Species Coexistence: Ecological and Evolutionary Perspectives. Oxford, UK: Blackwell Science.Google Scholar
Tokmakova, E. G., Verzhutsky, D. B. & Bazanova, L. P. (2006). Formation of the proventriculus blockage, alimentary activity and mortality in fleas Amphipsylla primaris primaris infected with Yersinia pestis. Parazitologiya, 40, 215–224 (in Russian).Google Scholar
Trager, W. (1939). Acquired immunity to ticks. Journal of Parasitology, 25, 57–81.CrossRefGoogle Scholar
Tränkle, S. B. (1989). Wirtspecifizität und Wanderaktivität des Katzenflohes Ctenocephalides felis (Bouché). Unpublished M.Sc. thesis, Albert Ludwigs Universität, Freiburg im Beisgau, Germany.Google Scholar
Traub, R. (1972a). The Gunong Benom Expedition 1967. Ⅻ. Notes on zoogeography, convergent evolution and taxonomy of fleas (Siphonaptera), based on collection from Gunong Benom and elsewhere in South-East Asia. 2. Convergent evolution. Bulletin of the British Museum Natural History, Zoology, 23, 309–387.Google Scholar
Traub, R. (1972b). The relationship between the spines, combs and other skeletal features of fleas (Siphonaptera) and the vestiture, affinities and habits of their hosts. Journal of Medical Entomology, 9, 601.Google Scholar
Traub, R. (1980). The zoogeography and evolution of some fleas, lice and mammals. In Fleas: Proceedings of the International Conference on Fleas, Ashton Wold, Peterborough, UK, 21–25 June 1977, ed. Traub, R. & Starcke, H.. Rotterdam, the Netherlands: A. A. Balkema, pp. 93–172.Google Scholar
Traub, R. (1985). Coevolution of fleas and mammals. In Coevolution of Parasitic Arthropods and Mammals, ed. Kim, K. C.. New York: John Wiley, pp. 295–437.Google Scholar
Traub, R., Wisseman, C. L. & Farhang-Azad, A. (1978). The ecology of murine typhus: a critical review. Tropical Diseases Bulletin, 75, 237–317.Google ScholarPubMed
Traub, R., Rothschild, M. & Haddow, J. F. (1983). The Ceratophyllidae: Key to the Genera and Host Relationships.Cambridge, UK: Cambridge University Press.Google Scholar
Tripet, F. & Richner, H. (1997a). Host responses to ectoparasites: food compensation by parent blue tits. Oikos, 78, 557–561.CrossRefGoogle Scholar
Tripet, F. & Richner, H. (1997b). The coevolutionary potential of a ‘generalist’ parasite, the hen flea Ceratophyllus gallinae. Parasitology, 115, 419–427.CrossRefGoogle Scholar
Tripet, F. & Richner, H. (1999a). Density-dependent processes in the population dynamics of a bird ectoparasite Ceratophyllus gallinae. Ecology, 80, 1267–1277.CrossRefGoogle Scholar
Tripet, F. & Richner, H. (1999b). Demography of the hen flea Ceratophyllus gallinae in blue tit Parus caeruleus nests. Journal of Insect Behavior, 12, 159–174.CrossRefGoogle Scholar
Tripet, F., Christe, P. & M⊘ller, A. P. (2002a). The importance of host spatial distribution for parasite specialization and speciaton: a comparative study of bird fleas (Siphonaptera: Ceratophyllidae). Journal of Animal Ecology, 71, 735–748.CrossRefGoogle Scholar
Tripet, F., Glaser, M. & Richner, H. (2002b). Behavioural responses to ectoparasites: time-budget adjustments and what matters to blue tits Parus caeruleus infested by fleas. Ibis, 144, 461–469.CrossRefGoogle Scholar
Tripet, F., Jacot, A. & Richner, H. (2002c). Larval competition affects the life histories and dispersal behaviour of an avian ectoparasite. Ecology, 83, 935–945.CrossRefGoogle Scholar
Trivers, R. L. & Willard, D. E. (1973). Natural selection of parental ability to vary sex ratio of offspring. Science, 179, 90–92.CrossRefGoogle ScholarPubMed
Trudeau, W. L., Fernandez-Caldas, E. & Fox, R. W. (1993). Allergenicity of the cat flea (Ctenocephalides felis felis). Clinical and Experimental Allergy, 23, 377–383.CrossRefGoogle Scholar
Trukhachev, N. N. (1971). Effect of host on the offspring of fleas Xenopsylla cheopis. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 423–425 (in Russian).Google Scholar
Tschirren, B., Fitze, P. S. & Richner, H. (2003). Sexual dimorphism in susceptibility to parasites and cell-mediated immunity in great tit nestlings. Journal of Animal Ecology, 72, 839–845.CrossRefGoogle Scholar
Tschirren, B., Richner, H. & Schwabl, H. (2004). Ectoparasite-modulated deposition of maternal androgens in great tit eggs. Proceedings of the Royal Society of London B, 271, 1371–1375.CrossRefGoogle ScholarPubMed
Tschirren, B., Saladin, V., Fitze, P. S., Schwabl, H. & Richner, H. (2005). Maternal yolk testosterone does not modulate parasite susceptibility or immune function in great tit nestlings. Journal of Animal Ecology, 74, 675–682.CrossRefGoogle Scholar
Tschirren, B., Bischoff, L. L., Saladin, V. & Richner, H. (2007a). Host condition and host immunity affect parasite fitness in a bird–ectoparasite system. Functional Ecology, 21, 372–378.CrossRefGoogle Scholar
Tschirren, B., Fitze, P. S. & Richner, H. (2007b). Maternal modulation of natal dispersal in a passerine bird: an adaptive strategy to cope with parasitism?American Naturalist, 169, 87–93.CrossRefGoogle Scholar
Uchikawa, K., Sato, A. & Kugimoto, M. (1967). A report on the flea fauna on the Oki Islands. Medical Entomology and Zoology, 18, 14–17 (in Japanese).CrossRefGoogle Scholar
Ulmanen, I. & Myllymäki, A. (1971). Species composition and numbers of fleas (Siphonaptera) in a local population of the field vole, Microtus agrestis (L.). Annales Zoologici Fennici, 8, 374–384.Google Scholar
Uvarov, B. P. (1931). Insects and climate. Transactions of the Royal Entomological Society of London, 79, 1–247.CrossRefGoogle Scholar
Vainikka, A., Jokinen, E. I., Kortet, R. & Taskinen, J. (2004). Gender- and season-dependent relationships between testosterone, oestradiol and immune functions in wild roach. Journal of Fish Biology, 64, 227–240.CrossRefGoogle Scholar
Valone, T. J. & Hoffman, C. D. (2002). Effects of regional pool size on local diversity in small-scale annual plant communities. Ecology Letters, 5, 477–480.CrossRefGoogle Scholar
Valtonen, E. T., Pulkkinen, K., Poulin, R. & Julkunen, M. (2001). The structure of parasite component communities in brackish water fishes of the northeastern Baltic Sea. Parasitology, 122, 471–481.CrossRefGoogle ScholarPubMed
Vandermeer, J. (1990). Indirect and diffuse interactions: complicated cycles in a population embedded in a large community. Journal of Theoretical Biology, 142, 429–442.CrossRefGoogle Scholar
Vansulin, S. A. (1961). Ecology of fleas parasitic on the great gerbils. In Proceedings of the Interdisciplinary Conference Dedicated to the 40th Anniversary of the Kazakh Soviet Socialist Republic, ed. Anonymous. Alma-Ata, USSR: Kainar, pp. 47–49 (in Russian).Google Scholar
Vansulin, S. A. (1965). On the ecology of fleas (Aphaniptera) of the great gerbil. Entomological Review, 44, 307–314 (in Russian).Google Scholar
Vansulin, S. A. & Volkova, L. A. (1962). Fur structure in Rhombomys opimus Licht. and its effect on the abundance of fleas parasitizing these rodents in different seasons. Zoologicheskyi Zhurnal, 41, 147–150 (in Russian).Google Scholar
Vuren, D. (1996). Ectoparasites, fitness, and social behaviour of yellow-bellied marmots. Ethology, 102, 686–694.CrossRefGoogle Scholar
Varma, M. G. R. & Page, R. J. C. (1966). The epidemiology of louping ill in Ayshire, Scotland: ectoparasites of small mammals. I. (Siphonaptera). Journal of Medical Entomology, 3, 331–335.CrossRefGoogle Scholar
Varma, M. G. R., Hellerhaupt, A., Trinder, P. K. E. & Langi, A. O. (1990). Immunization of guinea-pigs against Rhipicephalus appendiculatus adult ticks using homogenates from unfed immature ticks. Immunology, 71, 133–138.Google ScholarPubMed
Vashchenok, V. S. (1966a). Histological description of the oogenesis in fleas Echidnophaga oschanini Wagn. (Pulicidae, Aphaniptera). Zoologicheskyi Zhurnal, 45, 1821–1831 (in Russian).Google Scholar
Vashchenok, V. S. (1966b). Morphophysiological changes in the organism of fleas Echidnophaga oschanini Wagn. (Aphaniptera, Pulicidae) during feeding and reproduction. Entomological Review, 45, 715–727 (in Russian).Google Scholar
Vashchenok, V. S. (1967a). Gonotrophic relationships in fleas Ceratophyllus consimilis Wagn. (Aphaniptera, Ceratophyllidae). Parasitological Collection, 13, 222–235 (in Russian).Google Scholar
Vashchenok, V. S. (1967b). On the ecology of fleas Echidnophaga oshanini Wagn. (Pulicidae, Aphaniptera) in the Tuva ASSR. Parazitologiya, 1, 27–35 (in Russian).Google Scholar
Vashchenok, V. S. (1974). Activity of blood digestion in fleas. In Proceedings of the 7th Meeting of the All-Union Entomological Society, ed. Anonymous. Leningrad, USSR: Zoological Institute of the Academy of Sciences of the USSR, p. 209 (in Russian).Google Scholar
Vashchenok, V. S. (1979). Maintanence of the causative agent of the yersiniosis in fleas Xenopsylla cheopis. Parazitologiya, 13, 19–25 (in Russian).Google Scholar
Vashchenok, V. S. (1984). Fleas and agents of bacterial diseases of humans and animals. Parasitological Collection, 32, 79–123 (in Russian).Google Scholar
Vashchenok, V. S. (1988). Fleas: Vectors of Pathogens Causing Diseases in Humans and Animals. Leningrad, USSR: Nauka (in Russian).Google Scholar
Vashchenok, V. S. (1993). Factors regulating egg production in fleas Leptopsylla segnis (Leptopsyllidae: Siphonaptera). Parazitologiya, 27, 382–388 (in Russian).Google Scholar
Vashchenok, V. S. (1995). The dependence of the egg-laying activity in the fleas Leptopsylla segnis (Siphonaptera: Leptopsyllidae) on their abundance on a host. Parazitologiya, 29, 267–271.Google Scholar
Vashchenok, V. S. (2001). Age changes of fecundity in fleas Leptopsylla segnis (Siphonaptera: Leptopsyllidae). Parazitologiya, 35, 460–464 (in Russian).Google Scholar
Vashchenok, V. S. (2006). Species composition, host preferences and niche differentiation in fleas (Siphonaptera) parasitic on small mammals in the Ilmen–Volkhov Lowland. Parazitologiya, 40, 425–437 (in Russian).Google Scholar
Vashchenok, V. S. & Solina, L. T. (1969). On blood digestion in fleas Xenopsylla cheopis Roths. (Aphaniptera, Pulicidae). Parazitologiya, 3, 451–460 (in Russian).Google Scholar
Vashchenok, V. S. & Solina, L. T. (1972). Age-related changes in the fat tissue of female fleas Xenopsylla cheopis. Zoologicheskyi Zhurnal, 60, 79–85 (in Russian).Google Scholar
Vashchenok, V. S. & Tchirov, P. A. (1976). Histological study of fleas Ceratophyllus consimilis Wagn. infected with the causative agent of listeriosis (Listeria monocytogenes). Parazitologiya, 10, 61–66 (in Russian).Google Scholar
Vashchenok, V. S. & Tretiakov, K. A. (2003). Seasonal dynamic of flea (Siphonaptera) abundance on Clethrionomys glareolus in the northern part of the Novgorod Region. Parazitologiya, 37, 177–189 (in Russian).Google Scholar
Vashchenok, V. S. & Tretiakov, K. A. (2004). Seasonal dynamic of flea (Siphonaptera) abundance on the common shrew Sorex araneus in the northern part of the Novgorod Region. Parazitologiya, 38, 503–514 (in Russian).Google Scholar
Vashchenok, V. S. & Tretiakov, K. A. (2005). Seasonal dynamic of flea (Siphonaptera) abundance on the pygmy woodmouse Apodemus uralensis in the northern part of the Novgorod Region. Parazitologiya, 39, 270–277 (in Russian).Google Scholar
Vashchenok, V. S., SolinaL, T. L, T. & Zhirnov, A. E. (1976). Digestion of blood of different animals by fleas Xenopsylla cheopis. Parazitologiya, 10, 544–549 (in Russian).Google Scholar
Vashchenok, V. S., Bryukhanova, L. V. & Shchedrin, V. I. (1985). Characteristics of feeding and digestion in fleas. Parasitological Collection, 33, 134–148 (in Russian).Google Scholar
Vashchenok, V. S., Karandina, R. S. & Bryukhanova, L. V. (1988). Amount of blood consumed by different flea species in the experiments. Parazitologiya, 22, 312–320 (in Russian).Google Scholar
Vashchenok, V. S., Sheikin, A. O. & Serzhanov, O. S. (1992). Morphophysiological characteristics of fleas Xenopsylla gerbilli during the autumn–winter diapause. Parasitological Collection, 37, 5–15 (In Russian).Google Scholar
Vasiliev, G. I. (1961). Observations on flea breeding in the laboratory. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 2, 97–99 (in Russian).Google Scholar
Vasiliev, G. I. (1966). On ectoparasites and their hosts in relation to the plague epizootic in the Bajan-Khongor Aimak (Mongolian People's Republic). Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 26, 277–281 (in Russian).Google Scholar
Vasiliev, G. I. (1971). Fleas of the long-tailed ground squirrel (species composition, ecology, epizootological importance for plague). Unpublished Ph.D. thesis, Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, Irkutsk, USSR (in Russian).
Vasiliev, G. I. & Zhovty, I. F. (1961). An attempt to investigate the rules of the distribution of micropopulations of fleas in a microhabitat. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 1, 88–91 (in Russian).Google Scholar
Vasiliev, G. I. & Zhovty, I. F. (1971). On the annual cycle of Oropsylla asiatica Wagn., 1929 (Siphonaptera) parasitic on Spermophilus ungulatus in the Siberian Cis-Baikalia. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 9, 227–229 (in Russian).Google Scholar
Vaughan, J. A. & Coombs, M. E. (1979). Laboratory breeding of the European rabbit flea, Spilopsyllus cuniculi (Dale). Journal of Hygiene, 83, 521–530.CrossRefGoogle Scholar
Vaughan, J. A. & Mead-Briggs, A. R. (1970). Host-finding behaviour of the rabbit flea, Spilopsyllus cuniculi with special reference to the significance of urine as an attractant. Parasitology, 61, 397–409.CrossRefGoogle Scholar
Vaughan, J. A., Jerse, A. E. & Farhang-Azad, A. (1989). Rat leucocyte's response to the bites of rat fleas (Siphonaptera: Pulicidae). Journal of Medical Entomology, 26, 449–453.CrossRefGoogle Scholar
Vázquez, D. P. & Aizen, M. A. (2003). Null model analyses of specialization in plant–pollinator interactions. Ecology, 84, 2493–2501.CrossRefGoogle Scholar
Vázquez, D. P. & Stevens, R. D. (2004). The latitudinal gradient in niche breadth: concepts and evidence. American Naturalist, 164, E1–E19.CrossRefGoogle ScholarPubMed
Vázquez, D. P., Poulin, R., Krasnov, B. R. & Shenbrot, G. I. (2005). Species abundance patterns and the distribution of specialization in host–parasite interaction networks. Journal of Animal Ecology, 74, 946–955.CrossRefGoogle Scholar
Verts, B. J. (1961). Observations on the fleas (Siphonaptera) of some small mammals in northwestern Illinois. American Midland Naturalist, 66, 471–476.CrossRefGoogle Scholar
Verzhutsky, D. B., Zonov, G. B. & Popov, V. V. (1990). Epizootological importance of flea accumulation in the aggregations of female long-tailed ground squirrels in the Tuva plague focus. Parazitologiya, 24, 186–92 (in Russian).Google Scholar
Via, S. (2001). Sympatric speciation in animals: the ugly duckling grows up. Trends in Ecology and Evolution, 16, 381–390.CrossRefGoogle ScholarPubMed
Vidal-Martinez, V. M. & Poulin, R. (2003). Spatial and temporal repeatability in parasite community structure of tropical fish hosts. Parasitology, 127, 387–398.CrossRefGoogle ScholarPubMed
Viitala, J., Hakkarainen, H. & Ylönen, H. (1994). Different dispersal in Clethrionomys and Microtus. Annales Zoologici Fennici, 31, 411–415.Google Scholar
Violovich, N. A. (1969). Landscape and geographic distribution of fleas. In Biological Regionalization of the Novosibirsk Region, ed. Maximov, A. A.. Novosibirsk, USSR: Nauka, Siberian Branch, pp. 211–221 (in Russian).Google Scholar
Visser, M., Rehbein, S. & Wiedemann, C. (2001). Species of flea (Siphonaptera) infesting pets and hedgehogs in Germany. Journal of Veterinary Medicine B, 48, 197–202.CrossRefGoogle ScholarPubMed
Vobis, M., D'Haese, J., Mehlhorn, H., Heukelbach, J., Mencke, N. & Feldmeier, H. (2005). Molecular biological investigations of Brazilian Tunga sp. isolates from man, dogs, cats, pigs and rats. Parasitological Research, 96, 107–112.CrossRefGoogle ScholarPubMed
Volfertz, A. A. & Kolpakova, S. A. (1946). On epizootology of tularemia: role of fleas Ctenophthalmus orientalis Wagn. in epizootology of tularemia. Medical Parasitology and Parasitic Diseases [Meditsinskaya Parazitologiya i Parazitarnye Bolezni], 16, 83–84 (in Russian).Google Scholar
Volis, S., Mendlinger, S., Olswig-Whittaker, L., Safriel, U. N. & Orlovsky, N. (1998). Phenotypic variation and stress resistance in core and peripheral populations of Hordeum spontaneum. Biodiversity and Conservation, 7, 799–813.CrossRefGoogle Scholar
Vysotskaya, S. O. (1967). Biocoenotical relationships between ectoparasites of rodents and other inhabitants of their nests. Parasitological Collection, 23, 19–60 (in Russian).Google Scholar
Waage, J. K. (1979). The evolution of insect/vertebrate associations. Biological Journal of the Linnean Society, 12, 187–224.CrossRefGoogle Scholar
Wade, S. E. & Georgi, J. R. (1988). Survival and reproduction of artificially fed cat fleas, Ctenocephalides felis Bouché (Siphonaptera: Pulicidae). Journal of Medical Entomology, 25, 186–190.CrossRefGoogle Scholar
Waeber, P. O. & Hemelrijk, C. K. (2003). Female dominance and social structure in Alaotran gentle lemurs. Behaviour, 140, 1235–1246.CrossRefGoogle Scholar
Wagner, J. (1929). About new species of Palaearctic fleas (Aphaniptera). II. AnnualReports of the Zoological Museum of the Academy of Sciences of the USSR, 30, 531–547 (in Russian).Google Scholar
Wakelin, D. (1996). Immunity to Parasites: How Parasitic Infections Are Controlled, 2nd edn. Cambridge, UK: Cambridge University Press.Google Scholar
Walker, M., Steiner, S., Brinkhof, M. W. G. & Richner, H. (2003). Induced responses of nestling great tits reduce hen flea reproduction. Oikos, 102, 67–74.CrossRefGoogle Scholar
Wall, R. & Shearer, D. (2001). Veterinary Ectoparasites: Biology, Pathology and Control, 2nd edn. Oxford: Blackwell Science.CrossRefGoogle Scholar
Walshe, B. M. (1948). The oxygen requirements and thermal resistance of chironomid larvae from flowing and from still waters. Journal of Experimental Biology, 25, 35–44.Google Scholar
Walton, D. W. & Hong, H.-K. (1976). Fleas of small mammals from the endemic haemorrhagic fever zones of Kyonggi and Kangwon Provinces of the Repubic of Korea. Korean Journal of Parasitology, 14, 17–24.CrossRefGoogle ScholarPubMed
Walton, D. W. & Tun, U. M. (1978). Fleas of small mammals from Rangoon, Burma. Southeast Asian Journal of Tropical Medicine and Public Health, 9, 369–377.Google ScholarPubMed
Wang, G.-L., Xi, N., Adily, S., et al. (2004a). Studies on some characteristics of bioecology and morphology of Vermipsylla alakurt. Endemic Diseases Bulletin, 19, 25–27 (in Chinese).Google Scholar
Wang, G.-L., Xi, N., Dang, X.-S., et al. (2004b). Pathogen identification of vermipsyllosis of domestic animal in Bazhou, Xinjiang. Chinese Journal of Veterinary Parasitology, 12, 2–3 (in Chinese).Google Scholar
Warburg, A., Saraiva, E., Lanzaro, G. C., Titus, R. G. & Neva, F. (1994). Saliva of Lutzomyia longipalpis sibling species differs in its composition and capacity to enhance leishmaniasis. Philosophical Transactions of the Royal Society of London B, 345, 223–230.CrossRefGoogle ScholarPubMed
Ward, S. A. (1992). Assessing functional explanations of host specificity. American Naturalist, 139, 883–891.CrossRefGoogle Scholar
Warwick, R. M. & Clarke, K. R. (2001). Practical measures of marine biodiversity based on relatedness of species. Oceanography and Marine Biology, 39, 207–231.Google Scholar
Watkins, R. A., Moshier, S. E. & Pinter, A. J. (2006). The flea Megabothris abantis: an invertebrate host of Hepatozoon sp. and a likely definitive host in Hepatozoon infections of the montane vole, Microtus montanus. Journal of Wildlife Diseases, 42, 386–390.CrossRefGoogle Scholar
Watt, C., Dobson, A. P. & Grenfell, B. T. (1995). Glossary. In Ecology of Infectious Diseases in Natural Populations, ed. Grenfell, B. T. & Dobson, A. P.. Cambridge, UK: Cambridge University Press, pp. 510–521.CrossRefGoogle Scholar
Watts, M. M., Pascoe, D. & Carroll, K. (2002). Population responses of the freshwater amphipod Gammarus pulex (L.) to an environmental estrogen, 17 alpha-ethinylestradiol. Environmental Toxicology and Chemistry, 21, 445–450.CrossRefGoogle Scholar
Webb, C. T., Brooks, C. P., Gage, K. L. & Antolin, M. F. (2006). Classic flea-borne transmission does not drive plague epizootics in prairie dogs. Proceedings of the National Academy of Sciences of the USA, 103, 6236–6241.CrossRefGoogle Scholar
Webb, D. R., Porter, W. P. & Mcclure, P. A. (1990). Development of insulation in juvenile rodents: functional compromise in insulation. Functional Ecology, 4, 251–256.CrossRefGoogle Scholar
Webber, L. A. & Edman, J. D. (1972). Anti-mosquito behaviour of ciconiiform birds. Animal Behaviour, 20, 228–232.CrossRefGoogle Scholar
Wedekind, C. (1992). Detailed information about parasites revealed by sexual ornamentation. Proceedings of the Royal Society of London B, 247, 169–174.CrossRefGoogle Scholar
Wegner, Z. (1970). Lice (Anoplura) of small mammals caught in Dobrogea (Roumania). SocietƷʝii de Þiinʝe Biologia din Republica SocialistƷ România, Zoologia, 20, 305–314.Google Scholar
Wenk, P. (1953). Der Kopf von Ctenocephalus canis (Curt.) (Aphaniptera). Zoologische Jahrbücher, Abteilung für Anatomie und Ontogenie der Tiere, 73, 103–164.Google Scholar
Wenzel, R. L. & Tipton, V. J. (1966). Some relationships between mammal hosts and their ectoparasites. In Ectoparasites of Panama, ed. Wenzel, R. L. & Tipton, V. J.. Chicago, IL: Field Museum of Natural History, pp. 677–723.Google Scholar
Wesołowski, T. & Stańska, M. (2001). High ectoparasite loads in hole-nesting birds: a nestbox bias?Journal of Avian Biology, 32, 281–285.CrossRefGoogle Scholar
Wessels, W. (1998). Gerbillidae from the Miocene and Pliocene of Europe. Mitteilungen bayerische Staatssammlung für Paläontologie und historische Geologie, 38, 187–207.Google Scholar
Whitehead, M. D., Burton, H. R., Bell, P. J., Arnould, J. P. Y. & Rounsevell, D. E. (1991). A further contribution on the biology of the Antarctic flea, Glaciopsyllus antarcticus (Siphonaptera: Ceratophyllidae). Polar Biology, 11, 379–383.CrossRefGoogle Scholar
Whiteman, N. K. & Parker, P. G. (2004). Body condition and parasite load predict territory ownership in the Galapagos hawk. Condor, 106, 915–921.CrossRefGoogle Scholar
Whiteman, N. K. & Parker, P. G. (2005). Using parasites to infer host population history: a new rationale for parasite conservation. Animal Conservation, 8, 175–181.CrossRefGoogle Scholar
Whiting, M. F., Carpenter, J. C., Wheeler, Q. D. & Wheeler, W. C. (1997). The Strepsiptera problem: phylogeny of the holometabolous insect orders inferred from 18S and 28S ribosomal DNA sequences and morphology. Systematic Biology, 46, 1–68.Google ScholarPubMed
Whiting, M. F. (2002a). Phylogeny of the holometabolous insect orders: molecular evidence. Zoologica Scripta, 31, 3–15.CrossRefGoogle Scholar
Whiting, M. F. (2002b). Mecoptera is paraphyletic: multiple genes and phylogeny of Mecoptera and Siphonaptera. Zoologica Scripta, 31, 93–104.CrossRefGoogle Scholar
Whiting, M. F., Whiting, A. S. & Hastriter, M. W. (2003). A comprehensive phylogeny of Mecoptera and Siphonaptera. Entomologische Abhandlungen, 61, 169.Google Scholar
Widmann, O. (1922). Extracts from the diary of Otto Widmann. Transactions of the Academy of Science of St Louis, 24, 1–77.Google Scholar
Wikel, S. K. (1984). Immunomodulation of host responses to ectoparasite infestation: an overview. Veterinary Parasitology, 14, 321–339.CrossRefGoogle ScholarPubMed
Wikel, S. K. (ed.) (1996). The Immunology of Host–Ectoparasitic Arthropod Relationships. Wallingford, UK: CAB International.Google Scholar
Wikel, S. K. & Alarcon-Chaidez, F. J. (2001). Progress toward molecular characterization of ectoparasite modulation of host immunity. Veterinary Parasitology, 101, 275–287.CrossRefGoogle ScholarPubMed
Willadsen, P. (1980). Immunity to ticks. Advances in Parasitology, 18, 293–313.CrossRefGoogle ScholarPubMed
Willadsen, P. (1987). Immunological approaches to the control of ticks. International Journal for Parasitology, 17, 671–677.CrossRefGoogle ScholarPubMed
Willadsen, P. (2001). The molecular revolution in the development of vaccines against ectoparasites. Veterinary Parasitology, 101, 353–367.CrossRefGoogle ScholarPubMed
Willadsen, P. (2006). Vaccination against ectoparasites. Parasitology, 133, S9–S25.CrossRefGoogle ScholarPubMed
Willadsen, P., Bird, P. E., Cobon, G. & Hungerford, J. (1995). Commercialization of a recombinant vaccine against Boophilus microplus. Parasitology, 110, 543–550.CrossRefGoogle Scholar
Williams, B. (1991). Adaptations to endoparasitism in the larval integument and respiratory system of the flea Uropsylla tasmanica Rothschild (Siphonaptera, Pygiopsyllidae). Australian Journal of Zoology, 39, 77–90.CrossRefGoogle Scholar
Williams, B. (1993). Reproductive success of cat fleas, Ctenocephalides felis, on calves as unusual hosts. Medical and Veterinary Entomology, 7, 94–98.CrossRefGoogle ScholarPubMed
Williams, R. T. (1971). Observations on the behaviour of the European rabbit flea, Spilopsyllus cuniculi (Dale), on a natural population of wild rabbits, Oryctolagus cuniculus (L.), in Australia. Australian Journal of Zoology, 19, 41–51.CrossRefGoogle Scholar
Williams, R. T. & Paper, I. (1971). Observations on the dispersal of the European rabbit flea, Spilopsyllus cuniculi (Dale), through a natural population of wild rabbits, Oryctolagus cuniculus (L.). Australian Journal of Zoology, 19, 129–140.CrossRefGoogle Scholar
Willmann, R. (1981a). Das Exoskelett der männlichen Genitalien der Mecoptera (Insecta). I. Morphologie. Zeitschrift für zoologische Systematik und Evolutionsforschung, 19, 96–150.CrossRefGoogle Scholar
Willmann, R. (1981b). Das Exoskelett der männlichen Genitalien der Mecoptera (Insecta). II. Die phylogenetischen Beziehungen der Schnabelfliegen-Familien. Zeitschrift für zoologische Systematik und Evolutionsforschung, 19, 153–174.CrossRefGoogle Scholar
Wilson, D. E. & Reeder, D. M. (eds.) (2005). Mammal Species of the World: A Taxonomic and Geographic Reference, 3rd edn. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Wilson, K., Bj⊘rnstad, O. N., Dobson, A. P., et al. (2001). Heterogeneities in macroparasite infections: patterns and processes. In The Ecology of Wildlife Diseases, ed. Hudson, P. J., Rizzoli, A., Grenfell, B. T., Heesterbeek, H. & Dobson, A. P.. Oxford, UK: Oxford University Press, pp. 6–44.Google Scholar
Wilson, N. A. (1961). The ectoparasites (Ixodides, Anoplura and Siphonaptera) of Indiana Mammals. Unpublished Ph.D. thesis, Purdue University, West Lafayette, IN.
Wilson, N. A. & Durden, L. A. (2003). Ectoparasites of terrestrial vertebrates inhabiting the Georgia Barrier Islands, USA: an inventory and preliminary biogeographical analysis. Journal of Biogeography, 30, 1207–1220.CrossRefGoogle Scholar
Wilson, N. A., Telford, S. R. & Forrester, D. J. (1991). Ectoparasites of a population of urban gray squirrels in northern Florida. Journal of Medical Entomology, 28, 461–464.CrossRefGoogle ScholarPubMed
Windsor, D. A. (1990). Heavenly hosts. Nature, 348, 104.CrossRefGoogle Scholar
Windsor, D. A. (1995). Equal rights for parasites. Conservation Biology, 9, 1–2.CrossRefGoogle Scholar
Winkel, W. (1975a). Vergleichend-brutbiologische Untersuchungen an fünf Meisenarten (Parus spp.) in einem niedersächsischen Aufforstungsgebeit mit japanischer Lärche Larix leptolepis. Die Vogelwelt, 96, 41–63.Google Scholar
Winkel, W. (1975b). Vergleichend-brutbiologische Untersuchungen an fünf Meisenarten (Parus spp.) in einem niedersächsischen Aufforstungsgebeit mit japanischer Lärche Larix leptolepis. Die Vogelwelt, 96, 104–114.Google Scholar
Withers, P. C. (1992). Comparative Animal Physiology. Fort Worth, TX: Saunders College Publishing.Google Scholar
Witt, L. H., Linardi, P. M., Meckes, O., et al. (2004). Blood-feeding of Tunga penetrans males. Medical and Veterinary Entomology, 18, 439–441.CrossRefGoogle ScholarPubMed
Worthen, W. B. (1996). Community composition and nested-subsets analyses: basic descriptors for community ecology. Oikos, 76, 417–426.CrossRefGoogle Scholar
Worthen, W. B. & Rohde, K. (1996). Nested subsets analyses of colonization-dominated communities: metazoan ectoparasites of marine fishes. Oikos, 75, 471–478.CrossRefGoogle Scholar
Wright, D. H., Patterson, B. D., Mikkelson, G. M., Cutler, A. & Atmar, W. (1998). A comparative analysis of nested subset patterns of species composition. Oecologia, 113, 1–20.CrossRefGoogle Scholar
Xun, H. & Qi, Y.-M. (2004). Histochemistry of three enzymes in newly emerged and engorged adults of rat fleas Monopsyllus anisus (Rothschild) and Leptopsylla segnis (Schönherr). Acta Entomologica Sinica, 47, 444–448 (in Chinese).Google Scholar
Xun, H. & Qi, Y.-M. (2005). Histochemistry of fat and nonspecific esterase in newly emerged and sucked adults of rat fleas Monopsyllus anisus (Rothschild) and Leptopsylla segnis (Schoenherr). Acta Entomologica Sinica, 48, 829–832 (in Chinese).Google Scholar
Yadav, A., Khajuria, J. K. & Devi, J. (2006). Cat flea infestation in goats. Indian Veterinary Journal, 83, 439–440.Google Scholar
Yakunin, B. M. & Kunitskaya, N. T. (1980). On the effect of high relative humidity on reproduction and longevity of fleas Xenopsylla skrjabini. In Problems of Natural Focality of Plague: Proceedings of the 4th Soviet–Mongol Conference of Specialists from the Anti-Plague Establishments, vol. 1, ed. Golubinsky, E. P.. Irkutsk, USSR: Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, pp. 88–89 (in Russian).Google Scholar
Yakunin, B. M., Zolotova, S. I., Serzhanov, O. S., et al. (1971). On density dynamics and reproduction of Pulex irritans. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 450–452 (in Russian).Google Scholar
Yakunin, B. M., Tchernova, N. A. & Kunitskaya, N. T. (1979). Annual number of generations of fleas Xenopsylla skrjabini in the Mangyshlak Peninsula (Aphaniptera). Parazitologiya, 13, 510–515 (in Russian).Google Scholar
Yamauchi, T. (2005). Human dermatitis caused by the house-martin flea, Ceratophyllus farreni chaoi (Siphonaptera: Ceratophyllidae) in Shimane Prefecture, Japan. Medical Entomology and Zoology, 56, 49–52 (in Japanese).CrossRefGoogle Scholar
Yensen, E., Baird, C. R. & Sherman, P. W. (1996). Larger ectoparasites of the Idaho ground squirrel (Spermophilus brunneus). Great Basin Naturalist, 56, 237–246.CrossRefGoogle Scholar
Yeruham, I. & Koren, O. (2003). Severe infestation of a she-ass with the cat flea Ctenocephalides felis felis (Bouché, 1835). Veterinary Parasitology, 115, 365–367.CrossRefGoogle Scholar
Yeruham, I., Rosen, S. & Hadani, A. (1989). Mortality in calves, lambs and kids caused by severe infestation with the cat flea Ctenocephalides felis felis (Bouché, 1835) in Israel. Veterinary Parasitology, 30, 351–356.CrossRefGoogle Scholar
Ying, B., Kosoy, M. Y, Maupin, G. O., Tsuchiya, K. R. & Gage, K. L. (2002). Genetic and ecological characteristics of Bartonella communities in rodents in southern China. American Journal of Tropical Medicine and Hygiene, 66, 622–627.CrossRefGoogle Scholar
Yinon, U., Shulov, A. & Margalit, Y. (1967). The hygroreaction of the larvae of the Oriental rat flea, Xenopsylla cheopis Rothsch. (Siphonaptera: Pulicidae). Parasitology, 57, 315–319.CrossRefGoogle Scholar
Yudin, B. S., Krivosheev, V. G. & Belyaev, V. G. (1976). Small Mammals of the Northern Far East. Novosibirsk, USSR: Nauka, Siberian Branch (in Russian).Google Scholar
Yue, B.-S., Zou, F.-D., Sun, Q.-Z. & Li, J. (2002). Mating behavior of the cat flea, Ctenocephalides felis Bouché (Siphonaptera: Pulicidae) and male response to female extract on an artificial feeding system. Acta Entomologica Sinica, 9, 29–34 (in Chinese).Google Scholar
Yurgenson, I. A. & Maksimov, V. N. (1981). Effect of air temperature and relative humidity on the pre-imaginal development of fleas Ctenophthalmus teres (Siphonaptera). Parazitologiya, 15, 38–46 (in Russian).Google Scholar
Yushchenko, G. V. (1965). On the problem of the pseudotuberculosis research in the USSR. Proceedings of the Scientific Institute of Vaccines and Sera and Tomsk Medicine Institute, 16, 167–173 (in Russian).Google Scholar
Zagniborodova, E. N. (1960). Fauna and ecology of fleas in western Turmenistan. In Problems of Natural Foci and Epizootology of Plague in Turkmenistan, ed. Fenyuk, B. K.. Saratov, USSR: Turkmenian Anti-Plague Station and All-Union Scientific Anti-Plague Institute ‘Microb’, pp. 320–334 (in Russian).Google Scholar
Zagniborodova, E. N. (1965). Epizootological importance of migrating fleas of the great gerbil in Turkmenistan. Proceedings of the Academy of Sciences of the Turkmenian SSR, Biology, 5, 65–70 (in Russian).Google Scholar
Zagniborodova, E. N. (1968). Long-term study of the ecology of fleas of the great gerbil in the southern part of the central Kara-Kum Desert. In Rodents and their Ectoparasites, ed. Fenyuk, B. K.. Saratov, USSR: Saratov University Press, pp. 78–86 (in Russian).Google Scholar
Zahavi, A. (1977). The cost of honesty (further remarks on the handicap principle). Journal of Theoretical Biology, 67, 603–605.CrossRefGoogle Scholar
Zahn, A. & Rupp, D. (2004). Ectoparasite load in European vespertilionid bats. Journal of Zoology, 262, 383–391.CrossRefGoogle Scholar
Zakson-Aiken, M., Gregory, L. M. & Shoop, W. L. (1996). Reproductive strategies of the cat flea (Siphonaptera: Pulicidae): parthenogenesis and autogeny?Journal of Medical Entomology, 33, 395–397.CrossRefGoogle ScholarPubMed
Zatsarinina, G. V. (1972). Salvic alakurt (Dorcadia dorcadia Roth., Aphaniptera, Vermipsillidae): the pest of the deer breeding. Proceedings of the Entomological Sector [Trudy Entomologisheskogo Sektora], 3, 1–108 (in Russian).Google Scholar
Zavala-Velazquez, J. E., Ruiz-Sosa, J. A., Sanchez-Elias, R. A., Becerra-Carmona, G. & Walker, D. H. (2000). Rickettsia felis rickettsiosis in Yucatan. Lancet, 356, 1079–1080.CrossRefGoogle ScholarPubMed
Zeigler, R. & Ibrahim, M. M. (2001). Formation of lipid reserves in fat body and eggs of the yellow fever mosquito, Aedes aegypti. Journal of Insect Physiology, 47, 623–627.CrossRefGoogle Scholar
Zenuto, R. R., Antinuchi, C. D. & Busch, C. (2002). Bioenergetics of reproduction and pup development in a subterranean rodent (Ctenomys talarum). Physiological and Biochemical Zoology, 75, 469–478.CrossRefGoogle Scholar
Zhang, T., Yu, X.-M. & Zhang, S.-Y. (2005). Investigation on the community composition of the small mammals and the parastic fleas in Xingning, Guangdong. Chinese Journal of Vector Biology and Control, 16, 446–447 (in Chinese).Google Scholar
Zhang, Y., Jin, S., Quan, G., et al. (1997). Distribution of Mammalian Species in China. Beijing: China Forestry Publishing House.Google Scholar
Zhang, Y.-Z., Gong, Z.-D., Feng, X.-G., et al. (2002). Study on the relationship between fleas and hosts in Mt. Baicaoling, Yunnan Province, China. Endemic Diseases Bulletin, 17, 22–23 (in Chinese).Google Scholar
Zhao, L., Jin, H.-L., She, R.-P., et al. (2006). A rodent model for allergic dermatitis induced by flea antigens. Veterinary Immunology and Immunopathology, 114, 285–296.CrossRefGoogle ScholarPubMed
Zhovty, I. F. (1963). Some contradictory questions of ecology of the rodent fleas in association with their epidemiological importance. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 6, 96–104 (in Russian).Google Scholar
Zhovty, I. F. (1967). Effect of rodents' life history on ecological conditions of their fleas. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 27, 195–210 (in Russian).Google Scholar
Zhovty, I. F. (1970). Essays on the ecology of fleas parasitic on rodents in Siberia and Far East. II. Fleas of marmots. In Vectors of Particularly Dangerous Diseases and Their Control, ed. Tiflov, V. E.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 253–283 (in Russian).Google Scholar
Zhovty, I. F. & Kopylova, O. A. (1957). Fleas of the Daurian pika in the period of the increase of its density. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 15, 293–298 (in Russian).Google Scholar
Zhovty, I. F. & Leonov, Y. A. (1958). Abundance of fleas on the Norway rats in the settlements of the southern part of the Primorie Region (Far East) and some patterns of its variation. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 75–89 (in Russian).Google Scholar
Zhovty, I. F. & Peshkov, B. I. (1958). Observations on the overwintering of fleas parasitic on the grey marmots in the Trans-Baikalia. Proceedings of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 17, 27–32 (in Russian).Google Scholar
Zhovty, I. F. & Vasiliev, G. I. (1962a). Temperature conditions of rodents' fur as an environment for fleas. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 4, 152–156 (in Russian).Google Scholar
Zhovty, I. F. & Vasiliev, G. I. (1962b). On the self-cleaning from fleas in rodents. Transactions of the Irkutsk State Scientific Anti-Plague Institute of Siberia and Far East, 4, 156–160 (in Russian).Google Scholar
Zhovty, I. F., Netchaeva, L. K., Koshkin, S. M., et al. (1983). Patterns of seasonal density fluctuations in populations of the rat fleas in the Primorie Region (Far East) and a search for the criteria for their prognosis. In Prophylaxis of Diseases in the Natural Foci, ed. Taran, I. F.. Stavropol, USSR: Scientific Anti-Plague Institute of Caucasus and Trans-Caucasus, pp. 234–235 (in Russian).Google Scholar
Zolotova, S. I. (1968). Comparative ecological descriptions of fleas of the great gerbil, Xenopsylla gerbilli minax Jord., 1926 and Ctenophthalmus dolichus Ioff, 1953. Unpublished Ph.D. thesis, The Middle Asian Scientific Anti-Plague Institute, Alma-Ata, USSR (in Russian).Google Scholar
Zolotova, S. I. & Afanasieva, O. V. (1969). Materials on the ecology of fleas of the great gerbil. IV. Duration of pre-imaginal development in Ctenophthalmus dolichus. In Proceedings of the 6th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, vol. 2, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 66–68 (in Russian).Google Scholar
Zolotova, S. I. & Iskhanova, Z. A. (1979). Relationships between fleas Xenopsylla gerbilli minax and X. skrjabini in the area of overlapping of their geographic ranges. In Proceedings of the 10th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 99–101 (in Russian).Google Scholar
Zolotova, S. I. & Varshavskaya, P. N. (1974). Age composition of imago in the population of Xenopsylla skrjabini in the northern Cis-Aral Region. In Proceedings of the 8th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 316–318 (in Russian).Google Scholar
Zolotova, S. I., Pavlova, A. E. & Yakunin, B. M. (1971). Longevity of Pullex irritans after feeding on a non-specific host. In Proceedings of the 7th Scientific Conference of the Anti-Plague Establishments of the Middle Asia and Kazakhstan, ed. Aikimbaev, M. A.. Alma-Ata, USSR: The Middle Asian Scientific Anti-Plague Institute, pp. 377–378 (in Russian).Google Scholar
Zolotova, S. I., Bibikova, V. I. & Murzakhmetova, K. (1979). On the fecundity of fleas Xenopsylla gerbilli minax parasitic on the great gerbil (Aphaniptera). Parazitologiya, 13, 497–502.Google Scholar
Zuk, M. & McKean, K. A. (1996). Sex differences in parasite infections: patterns and processes. International Journal for Parasitology, 26, 1009–1024.CrossRefGoogle ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Boris R. Krasnov, Ben-Gurion University of the Negev, Israel
  • Book: Functional and Evolutionary Ecology of Fleas
  • Online publication: 14 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511542688.021
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
  • Boris R. Krasnov, Ben-Gurion University of the Negev, Israel
  • Book: Functional and Evolutionary Ecology of Fleas
  • Online publication: 14 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511542688.021
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
  • Boris R. Krasnov, Ben-Gurion University of the Negev, Israel
  • Book: Functional and Evolutionary Ecology of Fleas
  • Online publication: 14 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511542688.021
Available formats
×