Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-23T11:56:18.555Z Has data issue: false hasContentIssue false

Paleobiology of Predators, Parasitoids, and Parasites: Death and Accomodation in the Fossil Record of Continental Invertebrates

Published online by Cambridge University Press:  21 July 2017

Conrad C. Labandeira*
Affiliation:
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0121 and Department of Entomology, University of Maryland, College Park, Maryland 20742 USA
Get access

Abstract

Carnivory is the consumption of one animal by another animal; among invertebrates in terrestrial and freshwater ecosystems this type of feeding can take three forms: predation, parasitoidism, and parasitism. Differences among these three functional modes involve (i) whether the duration of feeding on the prey item is quick or there is an accommodation, coevolutionary or otherwise, between the carnivore and the host prey; (ii) whether the prey or host is killed; (iii) whether single or multiple prey or host items are consumed during the carnivore's lifespan, and (iv) the relative sizes of the carnivore and its prey or host. Uniformitarian and nonuniformitarian evidence directly relating to the history of carnivory can be found in exceptionally preserved deposits from the mid-Paleozoic to the Recent, but such evidence is relatively rare because carnivores are the least represented trophic group in ecosystems. Six types of paleobiological data provide evidence for carnivory: taxonomic affiliation, fossil structural and functional attributes, organismic damage, gut contents, coprolites, and indications of mechanisms for predator avoidance.

Only 12 invertebrate phyla have become carnivorous in the continental realm. Six are lophotrochozoans (Acanthocephala, Rotifera, Platyhelminthes, Nemertinea, Mollusca, and Annelida) and six are ecdysozoans (Nematoda, Nematomorpha, Tardigrada, Onychophora, Pentastoma, and Arthropoda). Most of these groups have poor continental fossil records, but the two most diverse—nematodes and arthropods—have comparatively good representation. The record of arthropods documents (i) the presence of predators among primary producers, herbivores, and decomposers in early terrestrial ecosystems; (ii) the addition later in the fossil record of the more accommodationist strategies of parasitoids and parasites interacting with animal hosts; (iii) the occurrence of simpler food-web structures in terrestrial ecosystems prior to parasitoid and parasite diversification; and (iv) a role for mass extinction in the degradation of food-web structure that ultimately affected carnivory. Future research should explore how different modes of carnivory have brought about changes in ecosystem structure through time. Despite numerous caveats and uncertainties, trace fossils left by predators on skeletons of their prey remain one of the most promising research directions in paleoecology and evolutionary paleobiology.

Type
Section II: Patterns
Copyright
Copyright © 2002 by The Paleontological Society 

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

Abel, O. 1935. Vorzeitliche Lebensspuren. Gustav Fischer, Jena, 644 p.Google Scholar
Abele, L. G., Kim, W., and Felgenhauer, B. E. 1989. Molecular evidence for inclusion of the phylum Pentastomida in the Crustacea. Molecular Biology and Evolution, 6:685691.Google Scholar
Abrahamson, W. G., and Weis, A. E. 1997. The Evolutionary Ecology of a Tritrophic-level Interaction: Goldenrod, the Stem Gall maker and its Natural Enemies. Princeton University Press, Princeton, NJ.Google Scholar
Aguinaldo, A. M. A., Turbeville, J. M., Linford, L. S., Rivera, M. C., Garey, J. R., Raff, R. A., and Lake, J. A. 1997. Evidence for a clade of nematodes, arthropods and other moulting animals. Nature, 387:489493.Google Scholar
Allison, C. W. 1975. Primitive fossil flatworm from Alaska: new evidence bearing on ancestry of the Metazoa. Geology, 3:649652.2.0.CO;2>CrossRefGoogle Scholar
Almond, J. E. 1985. The Silurian-Devonian fossil record of the Myriapoda. Philosophical Transactions of the Royal Society of London, B, 309:227237.Google Scholar
Andres, D. 1989. Phosphatisierte Fossilien aus dem unteren Ordoviz von Südschweden. Berliner Geowissenschaftliche Abhandlungen A, 106:919.Google Scholar
Arnett, R. H. Jr. 2000. American Insects, Second Edition. CRC Press, Boca Raton, FL, 1003 p.Google Scholar
Askew, R. W. 1979. Parasitic Insects. Heinemann, London, 316 p.Google Scholar
Bachofen-Echt, A. 1934. Beobachtungen über im Bernstein vorkommende Spinnengewebe. Biologia Generalis, 10:179184.Google Scholar
Balashov, Y. S. 1999. Evolution of haematophagy in insects and ticks. Entomological Review, 79:943954.Google Scholar
Banks, H. P., and Colthart, B. J. 1993. Plant-animal-fungal interactions in Early Devonian trimerophytes from Gaspé, Canada. American Journal of Botany, 80:9921001.CrossRefGoogle Scholar
Bänziger, H. 1975. Skin-piercing blood-sucking moths I: ecological and ethological studies on Calpe eustrigata (Lepid., Noctuidae). Acta Tropica, 32:125144.Google ScholarPubMed
Barnes, R. S. K., Calow, P., and Olive, P. J. W. 1993. The Invertebrates: A New Synthesis, Second Edition. Blackwell, London, 488 p.Google Scholar
Barthel, K. W., Swinburne, N. H. M., and Conway Morris, S. 1994. Solnhofen: A Study in Mesozoic Palaeontology. Cambridge University Press, Cambridge, UK, 236 p.Google Scholar
Bechley, G. 1996. Morphologische Untersuchungen am Flügelgeäder der Rezenten Libellen und deren Stammgruppenvertreter (Insecta; Pterygota; Odonata). Petalura, 2:1402.Google Scholar
Bechley, G., Nel, A., Martínez-Delclòs, X., Jarzembowsi, E., Coram, R., Martill, D., Fleck, G., Escuillié, F., Wissha, M. M., and Maisch, M. 2001. A revision and phylogenetic study of Mesozoic Aeschnoptera, with description of numerous new taxa (Insecta: Odonata: Anisoptera). Neue Paläontologische Abhandlungen, 4:1219.Google Scholar
Bertolani, R., and Grimaldi, D. 2000. A new eutardigrade (Tardigrada: Milnesiidae) in amber from the Upper Cretaceous (Turonian) of New Jersey, p. 103110. In Grimaldi, D. A. (ed.), Studies on Fossils in Amber, with Particular Reference to the Cretaceous of New Jersey. Backhuys, Leiden.Google Scholar
Betz, O. 1996. Function and evolution of the adhesion-capture apparatus of Stenus species (Coleoptera, Staphylinidae). Zoomorphology, 116:1534.CrossRefGoogle Scholar
Blackwell, A., Mordue, A. J., Young, M. R., and Mordue, W. 1992. Morphology of the antennae of two species of biting midge: Culicoides impuncatatus (Goetghebuer) and Culicoides nubeculosus (Meigen) (Diptera, Ceratopogonidae). Journal of Morphology, 213:85103.Google Scholar
Blaxter, M. L., De Ley, P., Garey, J. R., Liu, L. X., Scheldeman, P., Vierstraete, A., Vanfleteren, J. R., Mackey, L. Y., Dorris, M., Frisse, L. M., Vida, J. T., and Thomas, W. K. 1998. A molecular framework for the phylum Nematoda. Nature, 392:7175.Google Scholar
Borkent, A. 1995. Biting Midges in the Cretaceous Amber of North America (Diptera: Ceratopogonidae). Backhuys, Leiden, 237 p.Google Scholar
Bown, T. M., Hasiotis, S. T., Genise, J. F., Maldonado, F., and Brouwers, E. M. 1997. Trace fossils of Hymenoptera and other insects and paleoenvironments of the Claron Formation (Paleocene and Eocene), southwestern Utah. United States Geological Survey Bulletin, 2153-C:4158.Google Scholar
Brauckmann, C., and Zessin, W. 1989. Neue Meganeuridae aus dem Namurium von Hagen-Vorhalle (BRD) und die Phylogenie der Meganisoptera (Insecta, Odonata). Deutsche Entomologische Zeitschrift, N.F., 36:177215.CrossRefGoogle Scholar
Brown, S. J. 1989. Pathological consequences of feeding by hematophagous arthropods: comparison of feeding strategies, p. 415. In Jones, C. J. and Williams, R. E. (eds.), Physiological Interactions between Hematophagous Arthropods and Their Vertebrate Hosts. Entomological Society of America, Lanham, MD.Google Scholar
Brues, C. T. 1972. Insects, Food and Ecology. Dover Publications, New York, 466 p.Google Scholar
Brusca, R. C. 2000. Unraveling the history of arthropod biodiversification. Annals of the Missouri Botanical Garden, 87:1325.Google Scholar
Brusca, R. C., and Brusca, G. J. 1990. Invertebrates. Sinauer, Sunderland, MA, 922 p.Google Scholar
Buatois, L., Mángano, M. G., Genise, J. F., and Taylor, T. N. 1998. The ichnologic record of the continental invertebrate invasion: evolutionary trends in environmental expansion, ecospace utilization, and behavioral complexity. Palaios, 13:217240.Google Scholar
Capasso, L. 1993. Fossil mosquitoes and the spread of infectious diseases in man's ancestors. Journal of Paleopathology, 3:171201.Google Scholar
Carpenter, F. M. 1971. Adaptations among Paleozoic insects, p. 12361251. In Yochelson, E. (ed.), Proceedings of the First North American Paleontological Convention. Allen Press, Lawrence, KS.Google Scholar
Carpenter, F. M. 1992. Volume 3: Superclass Hexapoda. In Kaesler, R. L., Brosius, E., Keim, J., and Priesner, J. (eds.), Treatise on Invertebrate Paleontology, Part R, Arthropoda 4. Geological Society of America and University of Kansas, Lawrence, KS, 655 p.Google Scholar
Chalwatzis, N., Hauf, J., Van Der Peer, P. Y., Kinzelbach, R., and Zimmermann, F. K. 1996. 18S ribosomal RNA genes of insects: primary structure of the genes and molecular phylogeny of the Holometabola. Annals of the Entomological Society of America, 89:789803.Google Scholar
Clausen, C. P. 1940. Entomophagous Insects. McGraw-Hill, New York, 688 p.Google Scholar
Cockerell, T. D. A. 1917. Glossina and the extinction of Tertiary mammals. Nature, 103:265.Google Scholar
Cohen, A. S. 1989. The taphonomy of gastropod shell accumulations in large lakes: an example from Lake Tanganyika, Africa. Paleobiology, 15:2645.CrossRefGoogle Scholar
Conway Morris, S. 1981. Parasites and the fossil record. Parasitology, 82:489509.CrossRefGoogle Scholar
Conway Morris, S. 1998. The Crucible of Creation: The Burgess Shale and the Rise of Animals. Oxford University Press, Oxford, UK, 242 p.Google Scholar
Conway Morris, S., and Crompton, D. W. T. 1982. The origins and evolution of the Acanthocephala. Biological Reviews, 57:85115.Google Scholar
Cooper, K. W. 1964. The first fossil tardigrade: Beorn leggi Cooper, from Cretaceous amber. Psyche, 71:4148.CrossRefGoogle Scholar
Corbet, P. S. 1999. Dragonflies—Behaviour and Ecology of Odonata. Harley, Colchester, UK, 829 p.Google Scholar
Cronin, J. T., and Abrahamson, W. G. 2001. Do parasitoids diversify in response to host-plant shifts by herbivorous insects? Ecological Entomology, 26:347355.Google Scholar
Crowson, R. A. 1981. The Biology of the Coleoptera. Academic Press, New York, 802 p.Google Scholar
Dampf, A. 1910. Palaeopsylla klebsiana n. sp., ein fossiler Flöh aus dem baltischen Bernstein. Schriften der Physikalisch-ökonomischen Gesellschaft zu Köningsberg in Prussia, 51:248259.Google Scholar
Deangelis, D. L. 1992. Dynamics of Nutrient Cycling and Food Webs. Chapman & Hall, London, 270 p.CrossRefGoogle Scholar
Desportes, C. 1942. Forcipomyia velox Winn, et Sycorax silacea Curtis, vecteurs d'Icosiella neglecta (Diesing): filaire commune de la grenouille verte. Annales de Parasitologie, 19:5368.Google Scholar
Dewel, R. A., and Dewel, W. C. 1998. The place of tardigrades in arthropod evolution, p. 109123. In Fortey, R. A. and Thomas, R. H. (eds.), Arthropod Relationships. Chapman and Hall, London.Google Scholar
Downes, J. A. 1971a. Feeding and mating in the insectivorous Ceratopogoninae (Diptera). Memoirs of the Entomological Society of Canada, 104:162.Google Scholar
Downes, J. A. 1971b. The ecology of blood-sucking Diptera: an evolutionary perspective, p. 232258. In Fallis, A. M. (ed.), Ecology and Physiology of Parasites. University of Toronto Press, Toronto.Google Scholar
Downes, J. A., and Colless, D. H. 1967. Mouthparts of the biting and blood-sucking type in Tanyderidae and Chironomidae (Diptera). Nature, 214:13551356.CrossRefGoogle Scholar
Dubinin, B. V. 1948. Discovery of a Pleistocene louse (Anoplura) and nematodes during the study of corpses of Indigirsk fossil gophers. Doklady Akademiia Nauk SSSR, N.S., 62:417420 [in Russian].Google Scholar
Dubinina, M. N. 1972. Nematoda Alfortia edentatus (Looss, 1900) iz Kischechinka Verkhne pleistotsenovoi loshadi. Parasitologiya, 6:441443 [in Russian].Google Scholar
Dudley, R. 1998. Atmospheric oxygen, giant Paleozoic insects and the evolution of aerial locomotor performance. Journal of Experimental Biology, 201:10431050.Google Scholar
Dunlop, J. A. 1994. The paleobiology of the Writhlington trigontarbid arachnid. Proceedings of the Geologists' Association, 105:287296.Google Scholar
Dunne, J. A., Williams, R. J., and Martinez, N. D. 2002. Network topology and species loss in food webs: robustness increases with connectance. Santa Fe Institute Working Paper, 02-03-013:117.Google Scholar
Durante, M. V., and Maucci, W. 1972. Descrizione di Hybisibius (Isohyps.) basalovoi sp. nov. e altre notizie su tardigradi del Veronese. Memorie del Museo Civico di Storia Naturale di Verona, 20:275281.Google Scholar
Durden, C. J. 1988. Hamilton insect fauna, p. 117124. In Mapes, G. and Mapes, R. H. (eds.), Regional Geology and Paleontology of Upper Paleozoic Hamilton Quarry Area in Southeastern Kansas. Geological Society of America, Lawrence, KS.Google Scholar
Edwards, D., Selden, P. A., Richardson, J. B., and Axe, L. 1995. Coprolites as evidence for plant-animal interaction in Siluro-Devonian terrestrial ecosystems. Nature, 377:329331.Google Scholar
Eggleton, P., and Belshaw, R. 1992. Insect parasitoids: an evolutionary overview. Proceedings of the Royal Society of London, B, 337:120.Google Scholar
Eisenbeis, G, and Wichard, W. 1988. An Atlas on the Biology of Soil Arthropods. Springer Verlag, Berlin, 437 p.Google Scholar
Erwin, T. L., and Erwin, L. J. M. 1976. Relationships of predaceous beetles to tropical forest wood decay, Part II: The natural history of Neotropical Eurycoleus macularis Chevrolat (Carabidae: Lebiini) and its implications in the evolution of ectoparasitoidism. Biotropica, 8:215224.CrossRefGoogle Scholar
Erwin, T. L., and Scott, J. C. 1980. Seasonal and size patterns, trophic structure and richness of Coleoptera in the tropical arboreal ecosystem: the fauna of the tree Luehea seemannii Triana and Planch in the Canal Zone of Panama. Coleopterists Bulletin, 34:305322.Google Scholar
Evanoff, E., McIntosh, W. C., and Morphey, P. C. 2001. Stratigraphic summary and 40Ar/39Ar geochronology of the Florissant Formation, Colorado, p. 116. In Evanoff, E., Gregory-Wodzicki, K. M., and Johnson, K. R. (eds.), Fossil Flora and Stratigraphy of the Florissant Formation, Colorado. Proceedings of the Denver Museum of Nature and Science, 4.Google Scholar
Evans, H. C. 1989. Mycopathogens of insects of epigeal and aerial habitats, p. 205238. In Wilding, N., Collins, N. M., Hammond, P. M., and Webber, J. F. (eds.), Insect-Fungus Interactions. Academic Press, London.Google Scholar
Fang, Q. Q., McKeever, S., and French, F. E. 1999. Cladistic analysis of tabanids (Diptera: Tabanidae) using microscopic characters of the mouthparts. Memoirs on Entomology International, 14:355366.Google Scholar
Fischer, D. C. 1979. Evidence for subaerial activity of Euproops danae (Merostomata, Xiphosurida), p. 379447. In Nitecki, M. H. (ed.), Mazon Creek Fossils. Academic Press, New York.CrossRefGoogle Scholar
Foelix, R. F. 1996. Biology of Spiders, Second Edition. Oxford University Press, New York, 330 p.Google Scholar
Fortey, R. A., and Thomas, R. H. 1998. Arthropod Relationships. Chapman & Hall, London, 383 p.CrossRefGoogle Scholar
Franzen, J. L. 1985. Exceptional preservation of Eocene vertebrates in the lake deposit at Grube Messel (West Germany). Philosophical Transactions of the Royal Society of London, B, 311:181186.Google Scholar
Frey, D. G. 1964. Remains of animals in Quaternary lake and bog sediments and their interpretation. Archiv für Hydrobiologie, Beihefte, 2:1114.Google Scholar
Fuchs, G.-V. 1975. Die Gewinnung von Pollen und Nektar bei Käfern. Natur und Museum, 104:4554.Google Scholar
Fullard, J. H., and Napeoleone, N. 2001. Diel flight periodicity and the evolution of auditory defences in the Macrolepidoptera. Animal Behaviour, 62:349368.CrossRefGoogle Scholar
Gaston, K. J., and Hudson, E. 1994. Regional patterns of diversity and estimates of global insect species richness. Biodiversity and Conservation, 3:493500.Google Scholar
Gauld, I., and Bolton, B. 1988. The Hymenoptera. Oxford University Press, Oxford, UK, 332 p.Google Scholar
Genise, J. F. 1995. Upper Cretaceous trace fossils in permineralized plant remains from Patagonia, Argentina. Ichnos, 3:287299.Google Scholar
Gerhard, S., and Rietschel, W. 1968. Ein Stuck Bernstein und seine Einschlusse. Natur und Museum, 98:515520.Google Scholar
Glukhova, V. M. 1989. Blood-sucking midges of the genera Culicoides and Forcipomyia (Ceratopogonidae). Fauna of the USSR, 139,3,5a, 408. Nauka, Leningrad.Google Scholar
Godfray, H. C. J. 1994. Parasitoids: Behavioral and Evolutionary Ecology. Princeton University Press, Princeton, NJ, 473 p.Google Scholar
Godfray, H. C. J., Lewis, O. T., and Memmott, J. 1999. Studying insect diversity in the tropics. Philosophical Transactions of the Royal Society of London, B, 354:18111824.CrossRefGoogle ScholarPubMed
Göpfert, M. C., and Wasserthal, L. T. 1999. Hearing with the mouthparts: behavioural response and the structural basis of ultrasound perception in acherontiine hawkmoths. Journal of Experimental Biology, 202:909918.Google Scholar
Gordon, M. S., and Olson, E. C. 1995. Invasions of the Land: The Transitions of Organisms from Aquatic to Terrestrial Life. Columbia University Press, New York, 312 p.Google Scholar
Gould, S. J. 1995. Of tongue worms, velvet worms, and water bears. Natural History, 103:115.Google Scholar
Gradstein, F., and Ogg, J. 1996. A Phanerozoic time scale. Episodes, 19:35.Google Scholar
Graham, J. B., Dudley, R., Aguilar, N., and Gans, C. 1995. Implications of the late Palaeozoic oxygen pulse for physiology and evolution. Nature, 375:117120.Google Scholar
Gray, J. 1988. Evolution of the freshwater ecosystem: the fossil record. Palaeogeography, Palaeoclimatology, Palaeoecology, 61:1214.Google Scholar
Gray, J., and Shear, W. 1992. Early life on land. American Scientist, 80:444456.Google Scholar
Greene, A. 1975. Biology of the five species of Cychrini (Coleoptera: Carabidae) n the steppe region of southeastern Washington. Melanderia, 19:143.Google Scholar
Greenslade, P. J. M., and Whalley, P. E. S. 1986. The systematic position of Rhyniella praecursor Hirst & Maulik (Collembola), the earliest known hexapod, p. 319323. In Dallai, R. (ed.), Second International Seminar on Apterygota. University of Siena, Italy.Google Scholar
Grimaldi, D. A. 1992. Vicariance biogeography, geographic extinctions, and the North American tsetse flies, p. 178204. In Novacek, M. J. and Wheeler, Q. D. (eds.), Extinction and Phylogeny. Columbia University Press, New York.Google Scholar
Gronenberg, W. 1996. The trap-jaw mechanism in the dacetine ants Daceton armigerum and Strumigenys sp. Journal of Experimental Biology, 99:20212033.CrossRefGoogle Scholar
Hafner, M. S., and Nadler, S. A. 1988. Phylogenetic trees support the coevolution of parasites and their hosts. Nature, 332:258259.Google Scholar
Haas, F., and Kukalová-Peck, J. 2001. Dermaptera hindwing structure and folding: new evidence for familial, ordinal and superordinal relationships within the Neoptera (Insecta). European Journal of Entomology, 98:445509.Google Scholar
Hagen, K. S. 1987. Nutritional ecology of terrestrial insect predators, p. 533577. In Slansky, F. Jr., and Rodriguez, J. G. (eds.), Nutritional Ecology of Insects, Mites, Spiders, and Related Invertebrates. John Wiley & Sons, New York.Google Scholar
Hannibal, J. T., and Feldmann, R. M. 1988. Millipeds from late Paleozoic limestones at Hamilton, Kansas, p. 125131. In Mapes, G. and Mapes, R. H. (eds.), Regional Geology and Paleontology of Upper Paleozoic Hamilton Quarry Area in Southeastern Kansas. Geological Society of America, Lawrence, KS.Google Scholar
von Heyden, C. 1860. Mermis antiqua, ein fossiler Eingeweidewurm. Entomologische Zeitung, 21:38.Google Scholar
von Heyden, C. 1862. Gliederthiere aus der Braunkohle des Niederrhein's der Wetterau und der Rohn. Palaeontographica, 10:6282.Google Scholar
Hong, Y.-C. 1984. New fossil insects of Laiyang Group from Laiyang Basin, Shandong Province. Professional Papers in Stratigraphy and Palaeontology, 11:3141 [in Chinese with English abstract].Google Scholar
Hopkin, S. P. 1997. The Biology of the Springtails (Insecta: Collembola). Oxford University Press, Oxford, UK, 330 p.Google Scholar
Houston, T. F. 1987. Fossil brood cells of stenotritid bees (Hymenoptera: Apoidea) from the Pleistocene of South Australia. Transactions of the Royal Society of South Australia, 3:9397.Google Scholar
Iturralde-Vinent, M. A. 2001. Geology of the amber-bearing deposits of the Greater Antilles. Caribbean Journal of Science, 37:141167.Google Scholar
Janzen, D. H. 1971. Seed predation by animals. Annual Review of Ecology and Systematics, 2:465492.Google Scholar
Janzen, D. H. 1978. The ecology and evolutionary biology of seed chemistry as related to seed predation. In Harborne, J. B. (ed.), Biochemical Aspects of Plant and Animal Coevolution. Annals and Proceedings of the Phytochemical Society of Europe, 15:163206.Google Scholar
Jarzembowski, E. A. 1976. Report of Easter Field Meeting: the Lower Tertiaries of the Isle of Wight, 27–31. III. 1975. Tertiary Research, 1:1116.Google Scholar
Jarzembowski, E. A. 1984. Early Cretaceous insects from southern England. Modern Geology, 9:7193.Google Scholar
Jarzembowski, E. A. 1989. Cretaceous insect extinction. Mesozoic Research, 2:2528.Google Scholar
Jarzembowski, E. A. 1994. Fossil cockroaches or pinnule insects? Proceedings of the Geologists' Association, 105:305311.Google Scholar
Jell, P. A., and 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, 3:111205.Google Scholar
Juniper, B. E. 1986. The path to plant carnivory, p. 195218. In Juniper, B. E. and Southwood, T. R. E. (eds.), Insects and the Plant Surface. Edward Arnold, London.Google Scholar
Jeram, A. J. 1990. Book-lungs in a Lower Carboniferous scorpion. Nature, 343:360361.Google Scholar
Jervis, M. A., and Kidd, N. A. C. 1986. Host-feeding strategies in hymenopteran parasitoids. Biological Reviews, 61:395434.CrossRefGoogle Scholar
Kalugina, N. S. 1991. New Mesozoic Simuliidae and Leptoconopidae and blood-sucking origin in lower dipterans. Paleontologicheskii Zhurnal, 1991:6970 [in Russian].Google Scholar
Kelley, P. H., and Hansen, T. A. 1996. Naticid gastropod prey selectivity through time and the hypothesis of escalation. Palaios, 11:437445.Google Scholar
Kevan, P. G., Chaloner, W. G., and Savile, D. B. O. 1975. Interrelationships of early terrestrial arthropods and plants. Palaeontology, 18:391417.Google Scholar
Kim, K. C. 1988. Evolutionary parallelism in Anoplura and eutherian mammals. In Service, M. W. (ed), Biosystematics of Haematophagous Insects. Systematics Association Special Volume, 37:91114. Oxford University Press, Oxford, UK.Google Scholar
Kinzelbach, R. K., and Pohl, H. 1994. The fossil Strepsiptera (Insecta: Strepsiptera). Annals of the Entomological Society of America, 87:5970.Google Scholar
Klompen, H., and Grimaldi, D. A. 2001. First Mesozoic record of a parasitiform mite: a larval argasid tick in Cretaceous amber (Acari: Ixodida: Argasidae). Annals of the Entomological Society of America, 94:1015.Google Scholar
Korn, W. 1943. Die Muskulatur des Kopfes und des Thorax von Myrmeleon europaeus und ihre Metamorphose. Zoologischer Jahrbücher (Anatomie), 68:273330.Google Scholar
Kozur, H. 1970. Fossile Hirudinea aus dem Oberjura von Bayern. Lethaia, 3:225232.Google Scholar
Kristensen, N. P. 1999. Phylogeny of endopterygote insects, the most successful lineage of living organisms. European Journal of Entomology, 96:237253.Google Scholar
Kukalová-Peck, J. 1990. Fossil history and the evolution of hexapod structures, p. 141179. In Naumann, I. D., Carne, P. B., Lawrence, J. U. F., Nielsen, E. S., Spradbery, J. P., Taylor, R. W., Whitten, M. J., and Littlejohn, M. J. (eds.), The Insects of Australia, Second Edition, 1. Cornell University Press, Ithaca, NY.Google Scholar
Kutscher, M., and Koteja, J. 2000a. Trace fossils in Bitterfeld amber: excrements or detritus? Polskie Pismo Entomologiczne, 69:175178.Google Scholar
Kutscher, M., and Koteja, J. 2000b. Coccids and aphids (Hemiptera: Coccinea, Aphidinea), prey of ants (Hymenoptera: Formicidae): evidence from Bitterfeld amber. Polskie Pismo Entomologiczne, 69:179185.Google Scholar
Labandeira, C. C. 1990. Use of a Phenetic Analysis of Recent hexapod Mouthparts for the Distribution of Hexapod Food Resource Guilds in the Fossil Record. Ph.D. Dissertation, University of Chicago, 1186 p.Google Scholar
Labandeira, C. C. 1994. A compendium of fossil insect families. Milwaukee Public Museum Contributions in Biology and Geology, 88:171.Google Scholar
Labandeira, C. C. 1997a. Insect mouthparts: ascertaining the paleobiology of insect feeding strategies. Annual Review of Ecology and Systematics, 28:153193.CrossRefGoogle Scholar
Labandeira, C. C. 1997b. Permian pollen eating. Science, 277:14221423.Google Scholar
Labandeira, C. C. 1998a. The role of insects in Late Jurassic to middle Cretaceous ecosystems, p. 105124. In Lucas, S. G., Kirkland, J. I., and Estep, J. W. (eds.), Lower and Middle Cretaceous Terrestrial Ecosystems. New Mexico Museum of Natural History and Science Bulletin, 14.Google Scholar
Labandeira, C. C. 1998b. Early history of arthropod and vascular plant associations. Annual Review of Earth and Planetary Sciences, 26:329377.Google Scholar
Labandeira, C. C. 2000. The paleobiology of pollination and its precursors, p. 233269. In Gastaldo, R. and DiMichele, W. (eds.), Phanerozoic Terrestrial Ecosystems. Paleontological Society Papers, 6.Google Scholar
Labandeira, C. C. 2002a. The history of associations between plants and animals, p. 2674, 248–261. In Herrera, C. and Pellmyr, O. (eds.), Plant-Animal Interactions. Blackwell Science, Oxford, UK.Google Scholar
Labandeira, C. C. 2002b. The paleobiology of middle Eocene plant-insect associations from the Pacific Northwest: a preliminary report. Rocky Mountain Geology, 37 (in press).Google Scholar
Labandeira, C. C., and Sepkoski, J. J. Jr. 1993. Insect diversity in the fossil record. Science, 261:310315.Google Scholar
Labandeira, C. C., Beall, B. S., and Hueber, F. M. 1988. Early insect diversification: evidence from a Lower Devonian bristletail from Quebec. Science, 242:913916.CrossRefGoogle Scholar
Labandeira, C. C., Johnson, K. R., and Wilf, P. 2002. Impact of the terminal Cretaceous event on plant-insect associations. Proceedings of the National Academy of Sciences, USA, 99:20612066.CrossRefGoogle ScholarPubMed
Lakshminarayana, K. V., Mani, M. S., and Eichler, W. D. 1984. On the relationship of the fossil flea Saurophthirus longipes Ponomarenko (Siphonaptera: Insecta). Records of the Zoological Survey of India, 81:4348.Google Scholar
Lambrecht, F. L. 1993. Tsetse flies and trypanosomiasis during the American Tertiary. National Geographic Society Research Reports, 21:241249.Google Scholar
Larsson, S. G. 1978. Baltic amber—a palaeobiological study. Entomonograph, 1:1192.Google Scholar
Lehane, M. J. 1991. Biology of Blood-Sucking Insects. Harper-Collins, London, 288 p.CrossRefGoogle Scholar
Lewis, R. E., and Grimaldi, D. A. 1997. A pulicid flea in Miocene amber from the Dominican Republic (Insecta: Siphonaptera: Pulicidae). American Museum Novitates, 3205:19.Google Scholar
Little, C. 1983. The Colonisation of Land: Origins and Adaptations of Terrestrial Animals. Cambridge University Press, Cambridge, UK, 290 p.Google Scholar
Little, C. 1990. The Terrestrial Invasion: An Ecophysiological Approach to the Origins of Land Animals. Cambridge University Press, Cambridge, UK, 304 p.Google Scholar
Lyal, C. H. C. 1985. Phylogeny and classification of the Psocodea, with special reference to the lice (Psocodea: Phthiraptera). Systematic Entomology, 10:145165.Google Scholar
Lyal, C. H. C. 1987. Co-evolution of trichodectid lice (Insecta: Phthiraptera) and their mammalian hosts. Journal of Natural History, 21:128.Google Scholar
Malyshev, S. I. 1968. Genesis of the Hymenoptera and the Phases of their Evolution. Methuen, London, 319 p.Google Scholar
Manton, S. M. 1977. The Arthropoda: Habits, Functional Morphology and Evolution. Oxford University Press, Oxford UK, 527 p.Google Scholar
McKeever, S., Hagan, D. V., and Grogan, W. L. 1991. Comparative study of mouthparts of ten species of predaceous midges of the tribe Ceratopogonini (Diptera: Ceratopogonidae). Annals of the Entomological Society of America, 84:93106.Google Scholar
Menge, A. 1866. Ueber ein Rhipidipteron und einiger andere im Bernstein eingeschlossene Tiere. Schriften der Naturforschenden Gesellschaft in Danzig, N.F., 1:18.Google Scholar
Merritt, R. W., and Cummins, K. W. (eds.). 1984. An Introduction to the Aquatic Insects. Second edition. Kendall/Hunt, Dubuque, IA, 722 p.Google Scholar
Mikuilás, R., Dvorak, Z., and Pek, I. 1998. Lamniporichnus vulgaris igen. et isp. nov.: traces of insect larvae in stone fruits of hackberry (Celtis) from the Miocene and Pliocene of the Czech Republic. Journal of the Czech Geological Society, 43:277280.Google Scholar
Mikulic, D. G., Briggs, D. E. G., and Kluessendorf, J. E. 1985. A Silurian soft-bodied biota. Science, 228:715717.Google Scholar
Moussa, M. T. 1970. Nematode fossil trails from the Green River Formation (Eocene) in the Uinta Basin, Utah. Journal of Paleontology, 44:304307.Google Scholar
Müller, K. J., Walossek, D., and Zakharov, A. 1995. “Orsten” type phosphatized soft-integument preservation and a new record from the Middle Cambrian Kuonamka Formation in Siberia. Neues Jahrbuch für Geologie und Paläontologie Abhandlungen, 197:101118.Google Scholar
Naumann, I. D., Carne, P. B., Lawrence, J. F., Nielsen, E. S., Spradbery, J. P., Taylor, R. W., Whitten, M. J., and Littlejohn, M. J. (eds.). 1990. The Insects of Australia: A Textbook for Students and Research Workers, Second Edition. Cornell University Press, Ithaca, NY, 1137 p.Google Scholar
North, F. J. 1931. Insect-life in the coal forests, with special reference to South Wales. Transactions of the Cardiff Naturalists Society, 62:1644.Google Scholar
Panfilov, D. V. 1968. Kalligrammatids (Neuroptera, Kalligrammatidae) in Jurassic deposits of Karatau, p. 166175. In Rohdendorf, B. B. (ed.), Jurassic Insects of Karatau. Izdatel'stvo Nauka, Moscow [in Russian].Google Scholar
Papier, F., Nel, A., Grauvogel-Stamm, L., and Gall, J.-C. 1997. La plus ancienne sauterelle Tettigoniidae, Orthoptera (Trias, NE France): Mimétisme ou exaptation? Paläontologische Zeitschrift, 71:7177.CrossRefGoogle Scholar
Pawlowski, J., Szadziewski, R., Kmieciak, D., Fahrni, J., and Bittar, G. 1996. Phylogeny of the infraorder Culicomorpha (Diptera: Nematocera) based on 28S RNA gene sequences. Systematic Entomology, 21:167178.Google Scholar
Petrunkevitch, A. 1942. A study of amber spiders. Transactions of the Connecticut Academy of Arts and Sciences, 34:119464.Google Scholar
Pierce, W. D. 1960. Silicified Turbellaria from Calico Mountains nodules. Bulletin of the South California Academy of Sciences, 59:138143.Google Scholar
Poinar, G. O. Jr. 1977. Fossil nematodes from Mexican amber. Nematologica, 23:232238.Google Scholar
Poinar, G. O. Jr. 1984a. Fossil evidence of nematode parasitism. Revue de Nématologie, 7:201203.Google Scholar
Poinar, G. O. Jr. 1984b. First fossil record of parasitism by insect parasitic Tylenchida (Allantonematidae: Nematoda). Journal of Parasitology, 70:306308.Google Scholar
Poinar, G. O. Jr. 1984c. Heydenius dominicus n. sp. (Nematoda: Mermithidae), a fossil parasite from the Dominican amber. Journal of Nematology, 16:371375.Google Scholar
Poinar, G. O. Jr. 1985. Fossil evidence of insect parasitism by mites. International Journal of Acarology, 11:3738.Google Scholar
Poinar, G. O. Jr. 1987. Fossil evidence of spider parasitism by Ichneumonidae. Journal of Arachnology, 14:399400.Google Scholar
Poinar, G. O. Jr. 1988. Hair in Dominican amber: evidence for Tertiary land mammals in the Antilles. Experientia, 44:8889.Google Scholar
Poinar, G. O. Jr. 1991. The mycetophagous and entomophagous stages of Iotonchium californicum n. sp. (Iotonchiidae: Tylenchida). Revue de Nématologie, 14:565580.Google Scholar
Poinar, G. O. Jr. 1993. Insects in amber. Annual Review of Entomology, 46:145159.Google Scholar
Poinar, G. O. Jr. 1996. Fossil velvet worms in Baltic and Dominican amber: onychophoran evolution and biogeography. Science, 273:13701371.Google Scholar
Poinar, G. O. Jr. 1999a. Chrysomelidae in fossil resin: behavioural inferences, p. 116. In Cox, M. L. (ed.), Advances in Chrysomelidae Biology 1. Backhuys, Leiden.Google Scholar
Poinar, G. O. Jr. 1999b. Paleochordodes protus n. g., n. sp. (Nematomorpha, Chordodidae), parasites of a fossil cockroach, with a critical examination of other fossil hairworms and helminthes of extant cockroaches (Insecta: Blattaria). Invertebrate Biology, 118:109115.Google Scholar
Poinar, G. O. Jr. 2000. Heydenius araneus n. sp. (Nematoda: Mermithidae), a parasite of a fossil spider, with an examination of fossil helminths from extant spiders. Invertebrate Biology, 119:388393.Google Scholar
Poinar, G. O. Jr., and Brodzhinsky, J. 1985. Fossil evidence of nematode (Tylenchida) parasitism in Staphylinidae (Coleoptera). Nematologica, 32:353355.Google Scholar
Poinar, G. O. Jr., and Grimaldi, D. A. 1990. Fossil and extant macrochelid mites (Acari: Macrochelidae) phoretic on drosophilid flies (Diptera: Drosophilidae). Journal of the New York Entomological Society, 98:8892.Google Scholar
Poinar, G. O. Jr., and Milki, R. 2001. Lebanese Amber: The Oldest Insect Ecosystem in Fossilized Resin. Oregon State University Press, Corvallis, OR, 96 p.Google Scholar
Poinar, G. O. Jr., and Miller, J. C. 2002. First fossil record of endoparasitism of adult ants (Formicidae: Hymenoptera) by Braconidae (Hymenoptera). Annals of the Entomological Society of America, 95:4143.Google Scholar
Poinar, G. O. Jr., and Poinar, R. 1999. The Amber Forest: A Reconstruction of a Vanished World. Princeton University Press, Princeton, NJ, 239 p.Google Scholar
Poinar, G. O. Jr., and Ricci, C. 1992. Bdelloid rotifers in Dominican amber: evidence for parthenogenetic continuity. Experientia, 48:408410.Google Scholar
Poinar, G. O. Jr., Acra, A., and Acra, F. 1994a. Animal-animal parasitism in Lebanese amber. Medical Science Research, 22:159.Google Scholar
Poinar, G. O. Jr., Acra, A., and Acra, F. 1994b. Earliest fossil nematode (Mermithidae) in Cretaceous Lebanese amber. Fundamental and Applied Nematology, 17:475477.Google Scholar
Poinar, G. O. Jr., Treat, A. E., and Southcott, R. V. 1991. Mite parasitism of moths: examples of paleosymbiosis in Dominican amber. Experientia, 47:210212.Google Scholar
Ponomarenko, A. G. 1976. A new insect from the Cretaceous of Transbaikalia, a possible parasite of pterosaurians. Paleontological Journal, 1976:339343.Google Scholar
Ponomarenko, A. G. (ed.). 1988. The Cretaceous Biocoenotic Crisis in the Evolution of Insects. USSR Academy of Sciences, Moscow, 230 p. [in Russian].Google Scholar
Ponomarenko, A. G. 1996. Evolution of continental aquatic ecosystems. Paleontological Journal, 30:705709.Google Scholar
Popham, E. J. 1962. The anatomy related to the feeding habits of Arixenia and Hemimerus (Dermaptera). Proceedings of the Zoological Society of London, 139:429450.Google Scholar
Preston-Mafham, R., and Preston-Mafham, K. 1993. The Encyclopedia of Land Invertebrate Behaviour. Massachusetts Institute of Technology Press, Cambridge, MA, 320 p.Google Scholar
Price, P. W. 1980. Evolutionary Biology of Parasites. Princeton University Press, Princeton, NJ.Google Scholar
Price, P. W. 1997. Insect Ecology, Third Edition. Wiley, New York, 874 p.Google Scholar
Proctor, M., Yeo, P., and Lack, A. 1996. The Natural History of Pollination. Timber Press, Portland, OR, 479 p.Google Scholar
Pruvost, P. 1919. Le faune continentale du terrain houiller du Nord de la France. Mémorie pour Servir à l'Explication de la Carte Géologique Détaillé de la France. Imprimerie Nationale, Paris, 584 p.Google Scholar
Pritchard, G. 1976. Further observations on the functional morphology of the head and mouthparts of dragonfly larvae (Odonata). Quaestiones Entomologicae, 12:89114.Google Scholar
Rasnitsyn, A. P. 1975. Hymenoptera Apocrita of the Mesozoic. Transactions of the Paleontological Institute, 147:1134 [in Russian].Google Scholar
Rasnitsyn, A. P. 1988. An outline of evolution of the hymenopterous insects (order Vespida). Oriental Insects, 22:115145.Google Scholar
Rasnitsyn, A. P. 1992. Strashila incredibilis, a new enigmatic mecopteroid insect with possible siphonapteran affinities from the Upper Jurassic of Siberia. Psyche, 99:323333.Google Scholar
Rasnitsyn, A. P., and Zherikhin, V. V. 1999. First fossil chewing louse from the Lower Cretaceous of Baissa, Transbaikalia (Insecta, Pediculida = Phthiraptera, Saurodectidae fam. n.). Russian Entomological Journal, 8:253255.Google Scholar
Richards, O. W., and Davies, R. G. 1977. Imms' General Textbook of Entomology, Tenth Edition. Chapman and Hall, London, 1354 p.Google Scholar
Richter, G. 1992. Fossilized gut contents: analysis and interpretation, p. 285289. In Schaal, S. and Ziegler, W. (eds.), Messel—An Insight into the History of Life and of the Earth. Oxford University Press, Oxford, UK.Google Scholar
Richter, G., and Baszio, S. 2001. First proof of planctivory/insectivory in a fossil fish: Thaumaturus intermedius from the Eocene Lake Messel (FRG). Palaeogeography, Palaeoclimatology, Palaeoecology, 173:7585.Google Scholar
Richter, R., and Storch, G. 1980. Beiträge zur Ernährungsbiologie eozäner Fledermause aus der “Grube Messel.” Natur und Museum, 110:353367.Google Scholar
Riek, E. F. 1970. Lower Cretaceous fleas. Nature, 227:746747.Google Scholar
Ritzkowski, S. 1997. K-Ar-Alterbestimmungen der bernsteinführenden Sedimente des Samlandes (Paläogen, Bezirk Kaliningrad). Metalla Bochum, 66:1923.Google Scholar
Rohdendorf, B. B., and Rasnitsyn, A. P. 1980. Historical Development of the Class Insecta. USSR Academy of Sciences, Moscow [in Russian].Google Scholar
Rolfe, W. D. I. 1985. Aspects of the Carboniferous terrestrial arthropod community, p. 303316. In Dutro, J. T. Jr. and Pfefferkorn, H. W. (eds.), Comptes Rendus de Neuvième Congrès International de Stratigraphie et de Géologie du Carbonifère, 5. Southern Illinois University Press, Carbondale, IL.Google Scholar
Rolfe, W. D. I., Schram, F. R., Pacaud, G., Sotty, D., and Secretan, S. 1982. A remarkable Stephanian biota from Montceau-les-Mines, France. Journal of Paleontology, 56:426428.Google Scholar
Roth, B. 1986. Land mollusks (Gastropoda: Pulmonata) from early Tertiary Bozeman Group, Montana. Proceedings of the California Academy of Sciences, 44:237267.Google Scholar
Rowley, W. A., and Cornford, M. 1972. Scanning electron microscopy of the pit of the maxillary palp of selected species of Culicoides. Canadian Journal of Zoology, 50:12071210.Google Scholar
Ruiz-Trillo, I., Riutort, M., Littlewood, D. T. J., Lherniou, E. A., and Baguña, J. 1999. Acoel flatworms: earliest extant bilaterian metazoans, not members of Platyhelminthes. Science, 283:19191923.Google Scholar
Samways, M. J., Osborn, R., and Saunders, T. L. 1997. Mandible form relative to the main food type in ladybirds (Coleoptera: Coccinellidae). Biocontrol Science and Technology, 7:275286.Google Scholar
Schlee, D., and Glöckner, W. 1978. Bernstein. Stuttgarter Beiträge zur Naturkunde (C), 8:172.Google Scholar
Schmid-Hempel, P. 1998. Parasites in Social Insects. Princeton University Press, Princeton, NJ, 409 p.Google Scholar
Schmidt, W., Schurmann, M., and Teichmüller, M. 1958. Bisz-Spuren an Früchten des Miozän-Waldes der niederrheinischen Braunkohlen-formation. Fortschritte in der Geologie von Rheinland und Westfalen, 2:563572.Google Scholar
Schoonhoven, L. M., Jermy, T., and Van Loon, J. J. A. 1998. Insect-Plant Biology: From Physiology to Evolution. Chapman and Hall, London, 409 p.Google Scholar
Schram, F. R. 1973. Pseudocoelomates and a nemertine from the Illinois Pennsylvanian. Journal of Paleontology, 47:985989.Google Scholar
Schram, F. R. 1986. Crustacea. Oxford University Press, Oxford, UK, 606 p.Google Scholar
Schram, F. R., Feldman, R. M., and Copeland, M. J. 1978. The Late Devonian Palaeopalaemonidae and the earliest decapod crustaceans. Journal of Paleontology, 52:13751387.Google Scholar
Schuh, R. T., and Slater, J. A. 1995. True Bugs of the World (Hemiptera: Heteroptera). Cornell University Press, Ithaca, NY, 337 p.Google Scholar
Sciacchitano, J. 1955. Su un gordio fossile. Monitore Zoologico Italiano, 63:5761.Google Scholar
Scott, A. C., and Taylor, T. N. 1983. Plant/animal interactions during the Upper Carboniferous. Botanical Review, 49:259307.Google Scholar
Scourfield, D. J. 1926. On a new type of crustacean from the Old Red Sandstone (Rhynie Chert Bed, Aberdeenshire)—Lepidocaris rhyniensis gen. and sp. nov. Philosophical Transactions of the Royal Society of London, 214:153187.Google Scholar
Scudder, S. H. 1895. Revision of the American fossil cockroaches with descriptions of new forms. Bulletin of the United States Geological Survey, 124:1176.Google Scholar
Selden, P. A. 1996. Fossil mesothele spiders. Nature, 379:498499.Google Scholar
Sharov, A. G. 1973. Morphological features and way of life of Palaeodictyoptera, p. 4963. In Bei-Benko, G. Y. (ed.), 24th Annual Lectures in Memory of N. A. Kholodkovskogo, 24. Academy of Sciences, Moscow.Google Scholar
Shear, W. A., and Bonamo, P. M. 1988. Devonobiomorpha, a new order of centipeds (Chilopoda) from the Middle Devonian of Gilboa, New York state, USA, and the phylogeny of centiped orders. American Museum Novitates, 2927:130.Google Scholar
Shear, W. A., and Kukalová-Peck, J. 1990. The ecology of Paleozoic terrestrial arthropods: the fossil evidence. Canadian Journal of Zoology, 68:18071834.Google Scholar
Shear, W. A., and Selden, P. A. 2001. Rustling in the undergrowth: animals in early terrestrial ecosystems, p. 2951. In Gensel, P. G. and Edwards, D. (eds.), Plants Invade the Land: Evolutionary and Environmental Perspectives. Columbia University Press, New York.Google Scholar
Shear, W. A., Schawaller, W., and Bonamo, P. M. 1989b. Record of Palaeozoic pseudoscorpions. Nature, 341:527529.CrossRefGoogle Scholar
Shear, W. A., Palmer, J. M., Coddington, J. A., and Bonamo, P. M. 1989a. A Devonian spinneret: early evidence of spiders and silk use. Science, 246:479481.Google Scholar
Sih, A. 1987. Nutritional ecology of aquatic insect predators, p. 579607. In Slansky, F. Jr. and Rodriguez, J. G. (eds.), Nutritional Ecology of Insects, Mites, Spiders, and Related Invertebrates. John Wiley & Sons, New York.Google Scholar
Smit, F. G. A. M. 1972. On some adaptive structures in Siphonaptera. Folia Parasitologica, 19:517.Google Scholar
Smith, J. J. B. 1985. Feeding mechanisms, p. 3385. In Kerkut, G. A. and Gilbert, L. E. (eds.), Comprehensive Insect Physiology, Biochemistry, and Pharmacology, 4. Pergamon Press, Oxford, UK.Google Scholar
Snodgrass, R. E. 1952. A Textbook of Arthropod Anatomy. Cornell University Press, Ithaca, NY, 363 p.Google Scholar
Solem, A., and Yochelson, E. L. 1979. North American Paleozoic land snails, with a summary of other Paleozoic nonmarine snails. United States Geological Survey Professional Paper, 1072:142.Google Scholar
Southcott, R. V., and Lange, R. T. 1971. Acarine and other microfossils from the Maslin Eocene, South Australia. Records of the South Australian Museum, 16:121.Google Scholar
Storch, G., and Richter, G. 1992. The ant-eater Eurotamandua: a South American in Europe, p. 209215. In Schaal, S. and Ziegler, W. (eds.), Messel—An Insight into the History of Life and of the Earth. Oxford University Press, Oxford, UK.Google Scholar
Størmer, L. 1963. Gigantoscorpio willsi, a new scorpion from the Lower Carboniferous of Scotland and its associated preying microorganisms. Skrifter Utgitt a v Det Norske Videnskaps-Akademi i Oslo, 8:1171.Google Scholar
Størmer, L. 1969. Oldest known terrestrial arachnids. Science, 164:12761277.Google Scholar
Stricker, S. A. 1983. SEM and polarization microscopy of nemertean stylets. Journal of Morphology, 175:153169.Google Scholar
Sukacheva, I. D. 1982. Historical development of the order Phryganeida. Transactions of the Paleontological Institute, 197:1111 [in Russian].Google Scholar
Szadziewsi, R. 1996. Biting midges from Lower Cretaceous amber of Lebanon and Upper Cretaceous amber of Taimyr (Diptera, Ceratopogonidae). Studia Dipterologica, 3:2386.Google Scholar
Tasch, P. 1957. Flora and fauna of the Rhynie Chert: a paleoecological reevaluation of published evidence. University of Wichita Bulletin, 36:124.Google Scholar
Tasnádi-Kubacska, A. 1962. Paläopathologie, Vol. 2. Gustav Fischer, Jena, Germany, 269 p.Google Scholar
Taylor, T. N. 1981. Pollen and pollen organ evolution in early seed plants, p. 125. In Niklas, K. J. (ed.), Paleobotany, Paleoecology, and Evolution, 2. Praeger, New York.Google Scholar
Thompson, I. D., and Jones, D. S. 1980. A possible onychophoran from the Middle Pennsylvanian Mazon Creek of northern Illinois. Journal of Paleontology, 54:588596.Google Scholar
Thulborn, R. A. 1991. Morphology, preservation and palaeobiological significance of dinosaur coprolites. Palaeogeography, Palaeoclimatology, Palaeoecology, 83:341366.Google Scholar
Tillier, S., Masselot, M., and Tillier, A. 1996. Phylogenetic relationships of the pulmonate gastropods from rRNA sequences, and tempo and age of the stylommatophoran radiation, p. 267284. In Taylor, J. (ed.), Origin and Evolutionary Radiation of the Mollusca. Oxford University Press, Oxford, UK.Google Scholar
Tracey, S., Todd, J. A., and Erwin, D. H. 1993. Mollusca: Gastropoda, p. 131167. In Benton, M. J. (ed.), The Fossil Record 2. Chapman & Hall, London.Google Scholar
Traub, R., and Starcke, H. 1980. Fleas. Balkema, Rotterdam, 420 p.Google Scholar
Vermeij, G. J. 1987. Evolution and Escalation: An Ecological History of Life. Princeton University Press, Princeton, NJ, 527 p.Google Scholar
Vermeij, G. J., and Lindberg, D. R. 2000. Delayed herbivory and the assembly of marine benthic ecosystems. Paleobiology, 26:419430.Google Scholar
Vinson, S. B., and Barbosa, P. 1987. Interrelationships of nutritional ecology of parasitoids, p. 673695. In Slansky, F. Jr. and Rodriguez, J. G. (eds.), Nutritional Ecology of Insects, Mites, Spiders, and Related Invertebrates. John Wiley & Sons, New York.Google Scholar
Voigt, E. 1938. Ein fossiler Saitenwurm (Gordius tenuifibrosus n. sp.) aus der eozänen Braunkohle des Geiseltales. Nova Acta Leopoldina, N.F., 5:351360.Google Scholar
Voigt, E. 1952. Ein Haareinschluss mit Phthirapteren-Eiern im Bernstein. Mitteilungen der Geologische Staatsinstitut in Hamburg, 21:5974.Google Scholar
Voigt, E. 1957. Ein parasitischer Nematode in fossiler Coleopteren—Muskulatur aus der eozänen Braunkohle des Geiseltales bei Halle (Saale). Paläontologische Zeitschrift, 31:3539.Google Scholar
von Koenigswald, W., Richter, G. G., and Storch, G. 1981. Nachweis von Hornschuppen bein Eomanis waldi aus der ‘Grube Messel’ bein Darmstadt (Mammalia, Pholidota). Senckenbergiana Lethaea, 61:291298.Google Scholar
Waage, J. K. 1979. The evolution of insect/vertebrate associations. Biological Journal of the Linnean Society, 12:187224.Google Scholar
Walker, E. M. 1932. Prognathism and hypognathism in insects. Canadian Entomologist, 44:223229.Google Scholar
Walossek, D., and Müller, K. J. 1994. Pentastomid parasites from the Lower Paleozoic of Sweden. Transactions of the Royal Society of Edinburgh, Earth Sciences, 85:137.Google Scholar
Walossek, D., Repetski, J. E., and Müller, K. J. 1994. Heymonsicambria taylori n. sp. (Articulata: Pentastomida) from Upper Cambrian/Lower Ordovician boundary beds of Newfoundland, Canada. Canadian Journal of Earth Sciences, 31:16641671.Google Scholar
Walter, D., and Proctor, H. 1999. Mites: Ecology, Evolution and Behaviour. University of New South Wales Press, Sydney, 322 p.Google Scholar
Weitschat, W., and Wichard, W. 1998. Atlas der Pflanzen und Tiere im Baltischen Bernstein. Friedrich Pfeil, Munich, 256 p.Google Scholar
Wheeler, Q. D. 1990. Insect diversity and cladistic constraints. Annals of the Entomological Society of America, 83:10311047.Google Scholar
Whitfield, J. B. 1998. Phylogeny and evolution of host-parasitoid interactions in Hymenoptera. Annual Review of Entomology, 43:129151.Google Scholar
Wiegmann, B. M., Mitter, C., and Farrell, B. 1993. Diversification of carnivorous parasitic insects: extraordinary radiation or specialized dead end? American Naturalist, 142:738754.Google Scholar
Wilson, E. O. 1992. The Diversity of Life. Harvard University Press, Cambridge, MA, 424 p.Google Scholar
Wills, M. 1993. Miscellanea, p. 555560. In Benton, M., (ed.), The Fossil Record 2. Chapman and Hall, London.Google Scholar
Wirth, W. W., and Hubert, A. A. 1989. The Culicoides of Southeast Asia (Diptera, Nematocera). Memoirs of the American Entomological Institute, 44:1508.Google Scholar
Withycombe, C. L. 1922. Notes on the biology of some British Neuroptera (Planipennia). Transactions of the Entomological Society of London, 1922:501594.Google Scholar
Yeates, D. K., and Greathead, D. 1997. the evolutionary pattern of host use in the Bombyliidae (Diptera): a diverse family of parasitoid flies. Biological Journal of the Linnean Society, 60:149185.Google Scholar
Zherikhin, V. V. 1989. Oligocene seed beetles and acorn weevils (Coleoptera: Bruchidae, Curculionidae) from the Bol'shoy Svetlovodnaya River (Northern Primorye). Cenozoic of the Far East, 1989:145150.Google Scholar
Zumpt, F. 1965. Myiasis in Man and Animals in the Old World. Butterworth, London.Google Scholar