Hostname: page-component-84b7d79bbc-lrf7s Total loading time: 0 Render date: 2024-07-30T18:31:03.772Z Has data issue: false hasContentIssue false

The Fossil Record of Predation: An Overview of Analytical Methods

Published online by Cambridge University Press:  21 July 2017

Michal Kowalewski*
Affiliation:
Department of Geological Sciences, Virginia Polytechnic Institute and State University, Blacksburg, Virginia 24060 USA
Get access

Abstract

Paleontological research on predation has been expanding rapidly in scope, methods, and goals. The growing assortment of research strategies and goals has led to increasing differences in sampling strategies, types of data collected, definition of variables, and even reporting style. This methodological overview serves as a starting point for erecting some general methodological guidelines for studying the fossil record of predation. I focus here on trace fossils left by predators in the skeleton of their prey, arguably one of the most powerful sources of direct data on predator-prey interactions available in the fossil record. A critical survey of sampling protocols (data collecting strategy, sieve size, and sample size) and analytical approaches (predation intensity metrics, strategies for evaluating behavioral selectivity of predators, and taphonomic tests) reveals that various approaches can be fruitful depending on logistic circumstances and scientific goals of paleoecological projects. Despite numerous caveats and uncertainties, trace fossils left by predators on skeletons of their prey remain one of the most promising directions of research in paleoecology and evolutionary paleobiology.

Type
Section I: Methods
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

Alexander, R. R. 1981. Predation scars preserved in Chesterian brachiopods: probable culprits and evolutionary consequences for the articulates. Journal of Paleontology, 55:192203.Google Scholar
Alexander, R. R. 1986a. Resistance to repair of shell breakage induced by durophages in Late Ordovician brachiopods. Journal of Paleontology, 60:273285.Google Scholar
Alexander, R. R. 1986b. Frequency of sublethal shell-breakage in articulate brachiopod assemblages through geologic time. In Racheboeuf, P. R. and Emig, C. C. (eds.), Les Brachiopodes Fossiles et Actuels, First International Brachiopod Congress Proceedings, Biostratigraphie du Paleozoique, 4:159166.Google Scholar
Alexander, R. R., and Dietl, G. P. 2001. Shell repair frequencies in New Jersey bivalves: a recent baseline for tests of escalation with Tertiary, Mid-Atlantic congeners. Palaios, 16:354371.Google Scholar
Alexander, R. R., and Dietl, G. P. In press. The Phanerozoic history of shell-breaking predation on marine bivalves and gastropods. In Kelley, P. H., Kowalewski, M., and Hansen, T. A. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology Series, Plenum Press/Kluwer, New York.Google Scholar
Allmon, W. D., Nieh, J. C., and Norris, R. D. 1990. Drilling and peeling of turritelline gastropods since the Late Cretaceous. Palaeontology, 33:595611.Google Scholar
Alpert, S. P., and Moore, J. N. 1975. Lower Cambrian trace fossil evidence for predation on trilobites. Lethaia, 8:223230.Google Scholar
Ambrose, R. F. 1986. Effects of Octopus predation on motile invertebrates in a rocky subtidal community. Marine Ecology Progress Series, 30:261273.Google Scholar
Anderson, L. C. 1992. Naticid gastropod predation on corbulid bivalves: Effects of physical factors, morphological features, and statistical artifacts. Palaios, 7:602620.Google Scholar
Anderson, L. C., Geary, D. H., Nehm, R. H., and Allmon, W. D. 1991. A comparative study of naticid gastropod predation on Varicorbula caloosae and Chione cancellata, Plio-Pleistocene of Florida, U.S.A. Palaeogeography, Palaeoclimatology, Palaeoecology, 85:283290.Google Scholar
Andrews, P., and Fernandez Jalvo, Y. 1998. 101 uses for fossilized faeces. Nature, 393:629630.Google Scholar
Arnold, A. J., d'Escrivan, F., and Parker, W. C. 1985. Predation and avoidance responses in the Foraminifera of the Galapagos hydrothermal mounds. Journal of Foraminiferal Research, 15:3842.Google Scholar
Arnold, J. M., and Arnold, K. O. 1969. Some aspects of hole-boring predation by Octopus vulgaris. American Zoologist, 9:991996.Google Scholar
Arua, I., and Hoque, M. 1989. Fossil predaceous gastropod borings from Nigeria. Palaeogeography, Palaeoclimatology, Palaeoecology, 73:175183.Google Scholar
Ausich, W. I., and Gurrola, R. A. 1979. Two boring organisms in a Lower Mississippian community of southern Indiana. Journal of Paleontology, 53:335344.Google Scholar
Babcock, L. E. 1993. Trilobite malformations and the fossil record of behavioral asymmetry. Journal of Paleontology, 67:217229.Google Scholar
Babcock, L. E., and Robinson, R. A. 1989. Preferences of Palaeozoic predators. Nature, 337:695696.Google Scholar
Bauk, W., and Radwaski, A. 1996. Stomatopod predation upon gastropods from the Korytnica Basin, and from other classical Miocene localities in Europe. Acta Geologica Polonica, 46:279304.Google Scholar
Bambach, R. K. 2002. Supporting predators: Changes in the global ecosystem inferred from changes in predator diversity. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Bambach, R. K., and Kowalewski, M. 2000. How to count fossils. Geological Society of America Abstracts with Programs, 32:A-95.Google Scholar
Baumiller, T. K. 1990. Non-predatory drilling of Mississippian crinoids by platyceratid gastropods. Palaeontology, 33:743748.Google Scholar
Baumiller, T. K. 1993. Boreholes in Devonian blastoids and their implications for boring by platyceratids. Lethaia, 26:4147.Google Scholar
Baumiller, T. K. 1996. Boreholes in the Middle Devonian blastoid Heteroschisma and their implications for gastropod drilling. Palaeogeography, Palaeoclimatology, Palaeoecology, 123:343351.Google Scholar
Baumiller, T. K. 2002. Multi-snail infestation of Devonian crinoids and the nature of platyceratid-crinoid interactions. Acta Palaeontologica Polonica, 47:133139.Google Scholar
Baumiller, T. K., and Gahn, F. 2002. Fossil record of parasitism on marine invertebrates with special emphasis on the platyceratid-crinoid interactions. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Baumiller, T. K., and Macurda, D. B. Jr. 1995. Borings in Devonian and Mississippian blastoids (Echinodermata). Journal of Paleontology, 69:10841089.Google Scholar
Baumiller, T. K., Leighton, L. R., and Thompson, D. 1999. Boreholes in brachiopods of the Fort Payne Formation (Lower Mississippian, central USA). Palaeogeography, Palaeoclimatology, Palaeoecology, 147:283289.Google Scholar
Becker, M. A., Meier, J., and Slattery, W. 1999. Spiral coprolites from the Upper Cretaceous Wenonah-Mt. Laurel and Navesink formations in the northern coastal plain of New Jersey. Northeastern Geology and Environmental Sciences, 21:181187.Google Scholar
Behrensmeyer, A. K., and Hook, R. W. 1992, Paleoenvironmental contexts and taphonomic modes, p. 15136. In Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H. D., and Wing, S. L. (eds.), Terrestrial Ecosystems through Time. The University of Chicago Press, Chicago.Google Scholar
Bengtson, S. 2002. Origins and early evolution of predation. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Bengtson, S., and Yue, Z. 1992. Predatorial borings in late Precambrian mineralized exoskeletons. Science, 257:367369.Google Scholar
Bennington, J. B., and Rutheford, S. D. 1999. Precision and reliability in paleocommunity comparisons based on cluster-confidence intervals; how to get more statistical bang for your sampling buck. Palaios, 14:506515.Google Scholar
Berg, J. C. 1976. Ontogeny of predatory behavior in snails (Prosobranchia: Naticidae). Nautilus, 90:14.Google Scholar
Berg, J. C., and Nishenko, S. 1975. Stereotypy of predatory boring behavior of Pleistocene naticid gastropods. Paleobiology, 1:258260.Google Scholar
Bishop, G. A. 1975. Traces of predation, p. 261281. In Frey, R. W. (ed.), The Study of Trace Fossils. Springer-Verlag, New York.Google Scholar
Bishop, G. A. 1977. Pierre feces: A scatological study of the Dakoticancer assemblage, Pierre Shale (Upper Cretaceous) of South Dakota. Journal of Sedimentary Petrology, 47:129139.Google Scholar
Bookstein, F. L. 1991. Morphometric Tools for Landmark Data. Cambridge University Press, New York, 435 p.Google Scholar
Botton, M. L. 1984. Diet and food preference of the adult horseshoe crab Limulus polyphemus in Delaware Bay, New Jersey, USA. Marine Biology, 81:199207.Google Scholar
Boucot, A. J. 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier, Amsterdam, 725 p.Google Scholar
Brandt, D. S., Meyer, D. L., and Lask, P. B. 1995. Isotelus (Trilobita) “hunting burrow” from Upper Ordovician strata, Ohio. Journal of Paleontology, 69:10791083.Google Scholar
Brett, C. E. In press. Predation in Paleozoic marine communities. In Kelley, P. H., Kowalewski, M., and Hansen, T. A. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology Series, Plenum Press/Kluwer, New York.Google Scholar
Brett, C. E., and Walker, S. E. 2002. Predators and predation in Paleozoic marine environments. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Bromley, R. G. 1970. Borings as trace fossils and Entobia cretacea Portlock, as an example. Geological Journal Special Issue, 3:4990.Google Scholar
Bromley, R. G. 1981. Concepts in ichnotaxonomy illustrated by small round holes in shells. Acta Geologica Hispanica, 16:5564.Google Scholar
Bromley, R. G. 1993. Predation habits of octopus past and present and a new ichnospecies, Oichnus ovalis. Geological Society of Denmark Bulletin, 40:167173.Google Scholar
Bromley, R. G. 1996. Trace Fossils; Biology, Taphonomy and Applications. Chapman and Hall, 361 p.Google Scholar
Brunton, H. 1966. Predation and shell damage in a Visean brachiopod fauna. Palaeontology, 9:355359.Google Scholar
Bunn, H. T., and Kroll, E. M. 1986. Systematic butchery by Plio/Pleistocene hominids at Olduvai George, Tanzania. Current Anthropology, 27:431452.Google Scholar
Cadée, G. C. 1968. Molluscan biocoenoses and thanatocoenoses in the Ria de Arosa, Galicia, Spain. Zoologische Verhandelingen, Leiden, 95:1121.Google Scholar
Cadée, G. C. 1994. Eider, shelduck, and other predators, the main producers of shell fragments in the Wadden Sea; palaeoecological implications. Palaeontology, 37:181202.Google Scholar
Cadée, G. C. 2000. Herring gulls feeding on a Recent invader in the Wadden Sea, Ensis directus. In The Evolutionary Biology of the Bivalvia. Geological Society Special Publications, 177:459464.Google Scholar
Cadée, G. C., Walker, S. E., and Flessa, K. W. 1997. Gastropod shell repair in the intertidal of Bahia la Choya (N. Gulf of California). Palaeogeography, Palaeoclimatology, Palaeoecology, 136:6778.Google Scholar
Carpenter, K. 2000. Evidence of predatory behavior by carnivorous dinosaurs. Gaia, 15:135144.Google Scholar
Carriker, M. R. 1943. On the structure and function of the proboscis in the common oyster drill, Urosalpinx cinerea Say. Journal of Morphology, 73:441506.Google Scholar
Carriker, M. R. 1957. Preliminary study of behavior of newly hatched oyster drills Urosalpinx cinerea (Say). Journal of the Elisha Mitchell Scientific Society, 73:328351.Google Scholar
Carriker, M. R. 1961. Comparative functional morphology of the boring mechanisms in gastropods. American Zoologist, 1:263266.Google Scholar
Carriker, M. R. 1981. Shell penetration and feeding by naticacean and muricacean predatory gastropods: a synthesis. Malacologia, 20:403422.Google Scholar
Carriker, M. R., and Van Zandt, D. 1972a. Predatory behavior of a shell-boring muricid gastropod, p. 157244. In Winn, H. E. and Olla, B. L. (eds.), Behavior of Marine Animals: Current Perspectives in Research, Vol. 1 Invertebrates. Plenum Press, New York.Google Scholar
Carriker, M. R., and Van Zandt, D. 1972b. Regeneration of the accessory boring organ of muricid gastropods after excision. Transactions of American Microscope Society, 91:455466.Google Scholar
Carriker, M. R., and Yochelson, E. L. 1968. Recent gastropod boreholes and Ordovician cylindrical borings. U.S. Geological Survey Professional Paper, 593-B:126.Google Scholar
Carrion, J. S., Riquelme, J. A., Navarro, C., and Munuera, M. 2001. Pollen in hyaena coprolites reflects late glacial landscape in southern Spain. Palaeogeography, Palaeoclimatology, Palaeoecology, 176:193205.Google Scholar
Cate, A. S., and Evans, I. 1994. Taphonomic significance of the biomechanical fragmentation of live molluscan shell material by a bottom-feeding fish (Pogonias cromis) in Texas coastal bays. Palaios, 9:254274.Google Scholar
de Cauwer, G. 1985. Gastropod predation on corbulid bivalves; paleoecology or taphonomy? Annales du Société Zoologique de Belgique, 115:183196.Google Scholar
Chatterton, B. D. E., and Whitehead, H. L. 1987. Predatory borings in inarticulate brachiopod Artiotreta from the Silurian of Oklahoma. Lethaia, 20:6774.Google Scholar
Checa, A. 1993. Non-predatory shell damage in Recent deep-endobenthic bivalves from Spain. Palaeogeography, Palaeoclimatology, Palaeoecology, 100:309331.Google Scholar
Chin, K. 1997. What did dinosaurs eat? Coprolites and other direct evidence of dinosaur diet, p. 371382. In Farlow, J. O. and Brett-Surman, M. K. (eds.), The Complete Dinosaur. Indiana University Press, Bloomington.Google Scholar
Chin, K. 2002. Analyses of coprolites produced by carnivorous vertebrates. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Chin, K., Tokaryk, T. T., Erickson, G. M., and Calk, L. 1998. A king-sized theropod coprolite. Nature, 393:680682.Google Scholar
Colbath, S. L. 1985. Gastropod predation and depositional environment of two molluscan communities from the Miocene Astoria Formation at Beverly Beach State Park, Oregon. Journal of Paleontology, 59:849869.Google Scholar
Conway Morris, S., and Bengtson, S. 1994. Cambrian predators: Possible evidence from boreholes. Journal of Paleontology, 68:123.Google Scholar
Cooper, G. A., and Grant, R. E. 1972–1976. Permian brachiopods of West Texas, I–V. Smithsonian Contributions to Paleobiology, 14:1231; 15:233–793; 19:795–1298; 19:1300–1921; 21:1923–2285; 21:2288–2607; 24:2609–3159.Google Scholar
Coy, C. E. 1995. The first record of spiral coprolites from the Dinosaur Park Formation (Judith River Group, Upper Cretaceous) southern Alberta, Canada. Journal of Paleontology 69:11911194.Google Scholar
Crossland, M. R., Alford, R. A., and Collins, J. D. 1991. Population dynamics of an ectoparasitic gastropod, Hypermastus sp. (Eulimidae), on the sand dollar, Arachnoides placenta (Echinoidea). Australian Journal of Marine and Freshwater Research, 42:6976.Google Scholar
Culver, S. J., and Lipps, J. H. In press. Predation on and by foraminifera. In Kelley, P. H., Kowalewski, M., and Hansen, T. A. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology Series, Plenum Press/Kluwer, New York, in press.Google Scholar
Dawkins, R. 1976. The selfish gene. Oxford University Press, New York, 224 p.Google Scholar
Day, J. A. 1969. Feeding of the cymatiid gastropod, Argobuccinum argus, in relation to the structure of the proboscis and secretions of the proboscis gland. American Zoologist, 9:909919.Google Scholar
Degner, E. 1928. Über das Fleisch und Kalkbeddurfnis von Cepaea nemoralis. Archiv für Molluskenk., 60:209213.Google Scholar
Dietl, G. P. 2000. Successful and unsuccessful predation of the gastropod Nucella lapillus (Muricidae) on the mussel Mytilus edulis from Maine. Veliger, 43:319329.Google Scholar
Dietl, G. P., and Alexander, R. R. 1995. Borehole site and prey size stereotypy in naticid predation on Euspira (Lunatia) heros Say and Neverita (Polinices) duplicata Say from the southern New Jersey coast. Journal of Shellfish Research, 14:307314.Google Scholar
Dietl, G. P., and Alexander, R. R. 2000. Post-Miocene shift in stereotypic naticid predation on confamilial prey from the Mid-Atlantic shelf: Coevolution with dangerous prey. Palaios, 15:414429.Google Scholar
Dietl, G. P., and Kelley, P. H. 2001. Mid-Paleozoic latitudinal predation gradient: Distribution of brachiopod ornamentation reflects shifting Carboniferous climate. Geology, 29:111114.Google Scholar
Dietl, G. P., and Kelley, P. H. 2002. The fossil record of predator-prey arms races: Coevolution and escalation hypotheses. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Dietl, G. P., Alexander, R. R., and Bien, W. F. 2000. Escalation in Late Cretaceous–early Paleocene oysters (Gryphaeidae) from the Atlantic Coastal Plain. Paleobiology, 26:215237.Google Scholar
Dietl, G., Alexander, R. R., Kelley, P. H., and Hansen, T. A. 2001. Stereotypy of naticid predation on bivalves since the Cretaceous: trends, controlling factors, and implications for escalation. PaleoBios 21 (supplement to number 2):78.Google Scholar
Dodson, P., and Wexlar, D. 1979. Taphonomic investigations of owl pellets. Paleobiology, 5:279284.Google Scholar
Dryden, I. L., and Mardia, K. V. 1998. Statistical Shape Analysis. Wiley, New York, 347 p.Google Scholar
Farlow, J. O., and Holtz, T. R. 2002. The fossil record of predation in dinosaurs. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation, Paleontological Society Special Papers, 8 (this volume).Google Scholar
Feyerabend, P. K. 1978. Against method: Outline of an Anarchistic Theory of Knowledge. Verso, London, 339 p.Google Scholar
Fischer, P. H. 1922. Sur les gastéropodes perceurs. Journal de Conchyliologie, 67:156.Google Scholar
Fischer, P. H. 1962a. Perforations de fossiles Pré-Tertiaries attribuées a des gastéropodes prédateurs. Journal de Conchyliologie, 102:6878.Google Scholar
Fischer, P. H. 1962b. Action des gastéropodes perceurs sur des Mesalia de L'Étage Lutétien. Journal de Conchyliologie, 102:9597.Google Scholar
Fischer, P. H. 1963. Corbules fossiles perforées par des gastéropodes prédateurs. Journal de Conchyliologie, 103:2931.Google Scholar
Fischer, P. H. 1966. Perforations de fossiles Tertiaires par des gastéropodes prédateurs. Journal de Conchyliologie, 106:6696.Google Scholar
Frey, R. W., Howard, J. D., and Hong, J. S. 1986. Naticid gastropods may kill solenid bivalves without boring: Ichnologic and taphonomic consequences. Palaios, 1:610612.Google Scholar
Fujita, S. 1916. On the boring of pearl oyster by Octopus (Polypus) vulgaris Lamarck. Dobytsugaku Zasshi, 28:250257.Google Scholar
Geary, D. H., Allmon, W. D., and Reaka-Kudla, M. J. 1991. Stomatopod predation on fossil gastropods from the Plio-Pleistocene of Florida. Journal of Paleontology, 65:355360.Google Scholar
Gilinsky, N. L., and Bennington, B. 1994. Estimating numbers of whole individuals from collections of body parts: A taphonomic limitation of the paleontological record. Paleobiology, 20:245258.Google Scholar
Gittenberger, E. 1999. Predatory bore-holes in shells of terrestrial snails: Roth has priority. Basteria, 63:164.Google Scholar
Gibson, M. A., and Watson, J. B. 1989. Predatory and non-predatory borings in echinoids from the Upper Ocala Formation (Eocene), north-central Florida, U.S.A. Palaeogeography, Palaeoclimatology, Palaeoecology, 71:309321.Google Scholar
Guerra, A., and Nixon, M. 1987. Crab and mollusc shell drilling by Octopus vulgaris (Mollusca: Cephalopoda) in the Ria de Vigo (north-west Spain). Journal of Zoology, London, 211:515523.Google Scholar
Hagadorn, J. W., and Boyajian, G. E. 1997. Subtle changes in mature predator-prey systems; an example from Neogene Turritella (Gastropoda). Palaios, 12:372379.Google Scholar
Hageman, S. A., and Kaessler, R. L. 2002. Fusulinids: Predation damage and repair of tests from the Upper Pennsylvanian of Kansas. Journal of Paleontology, 76:181184.Google Scholar
Hagstrom, K. M. 1996. Effects of compaction and wave-induced forces on the preservation and macroevolutionary perception of naticid predator-prey interactions. , 63 p.Google Scholar
Hansen, T. A., and Kelley, P. H. 1995. Spatial variation in naticid gastropod predation in the Eocene of North America. Palaios, 10:268278.Google Scholar
Harper, E. M., Forsythe, G. T. W., and Palmer, T. 1998. Taphonomy and the Mesozoic marine revolution: Preservation state masks the importance of boring predators. Palaios, 13:352360.Google Scholar
Harper, E. M., Forsythe, G. T. W., and Palmer, T. 1999. A fossil record full of holes: The Phanerozoic history of drilling predation: Comment. Geology, 27:959.Google Scholar
Harper, E. M., and Wharton, D. S. 2000. Boring predation and Mesozoic articulate brachiopods. Palaeogeography, Palaeoclimatology, Palaeoecology, 158:1524.Google Scholar
Haynes, G. 2002. Archeological methods for reconstructing human predation on terrestrial vertebrates. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Henderson, W. G., Anderson, L. C., and McGimsey, C. R. 2002. Stratigraphic, taxonomic, and taphonomic criteria to distinguish shell-rich chenier and archaeological deposits of the Louisiana chenier plain. Palaios, in press.Google Scholar
Hirsch, G. C. 1915. Die Ernährungsbiologie fleischfressender Gastropoden (Murex, Natica, Pterotrachea, Pleurobranchea, Tritonium). I Teil. Makroskopischer Bau, Nahrungsaufnahme, Verdauung, Sekretion. Zoologische Jahrbuch, 35:357504.Google Scholar
Hirsch, K. F., Kihm, A. J., and Zelenitsky, D. K. 1997. New eggshell of ratite morphotype with predation marks from the Eocene of Colorado. Journal of Vertebrate Paleontology, 17:360369.Google Scholar
Hoffman, A., and Martinell, J. 1984. Prey selection by naticid gastropods in the Pliocene of Emporda (Northeast Spain). Neues Jahrbuch für Geologie und Paläontologie, Monatshaft, 1984:393399.Google Scholar
Hoffman, A., Pisera, A., and Ryszkiewicz, M. 1974. Predation by muricid and naticid gastropods on the Lower Tortonian mollusks from the Korytnica Clays. Acta Geologica Polonica, 24:249260.Google Scholar
Hoffmeister, A. P. 2002. Quantitative analysis of drilling predation patterns in the fossil record: Ecological and evolutionary implications. Unpublished Ph.D. dissertation, Virginia Polytechnic Institute and State University, Blacksburg, 293 p.Google Scholar
Hoffmeister, A.P., and Kowalewski, M. 2001. Spatial and environmental variation in the fossil record of drilling predation: A case study from the Miocene of Central Europe. Palaios, 16:566579.Google Scholar
Hoffmeister, A. P., Kowalewski, M., Bambach, R. K., and Baumiller, T. K. Submitted. Intense drilling predation on the brachiopod Cardiarina cordata Cooper 1956 from the Pennsylvanian of New Mexico. Lethaia.Google Scholar
Hughes, R. N., and Hughes, P. I. 1971. A study on the gastropod Cassis tuberosa (L.) preying upon sea urchins. Journal of Experimental Marine Biology and Ecology, 7:305314.Google Scholar
Hughes, R. N., and Hughes, H. P. I. 1981. Morphological and behavioural aspects of feeding in the Cassidea (Tonnacea, Mesogastropoda). Malacologia, 20:385402.Google Scholar
Jacobsen, A. R. 1997. Tooth marks, p. 738739. In Currie, P. J. and Padian, K. (eds.), Encyclopedia of Dinosaurs. Academic Press, San Diego.Google Scholar
Jacobsen, A. R. 1998. Feeding behavior of carnivorous dinosaurs as determined by tooth marks on dinosaur bones. Historical Biology, 13:1726.Google Scholar
Jensen, S. 1990. Predation by Early Cambrian trilobites on infaunal worms; evidence from the Swedish Mickwitzia Sandstone. Lethaia, 23:2942.Google Scholar
Kabat, A. R. 1990. Predatory ecology of naticid gastropods with a review of shell boring predation. Malacologia, 32:155193.Google Scholar
Kaplan, P., and Baumiller, T. K. 2000. Taphonomic inferences on boring habit in the Richmondian Onniella meeki epibole. Palaios, 15:499510.Google Scholar
Kaplan, P., and Baumiller, T. K. 2001. A misuse of Occam's Razor that trims more than just the fat. Palaios, 16:525528.Google Scholar
Kase, T., Johnston, P. A., Seilacher, A., and Boyce, J. B. 1998. Alleged mosasaur bite marks on Late Cretaceous ammonites are limpet (patellogastropod) home scars. Geology, 26:947950.Google Scholar
Kauffman, E. G., and Kesling, R. V. 1960. An Upper Cretaceous ammonite bitten by a mosasaur (South Dakota). Contributions from the Museum of Paleontology, University of Michigan, 15:193248.Google Scholar
Kelley, P. H. 1988. Predation by Miocene gastropods of the Chesapeake Group: Stereotyped and predictable. Palaios, 3:436448.Google Scholar
Kelley, P. H., and Hansen, T. A. 1993. Evolution of the naticid gastropod predator-prey system: An evaluation of the hypothesis of escalation. Palaios, 8:358375.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
Kelley, P. H., and Hansen, T. A. In press. The fossil record of drilling predation on bivalves and gastropods. In Kelley, P. H., Kowalewski, M., and Hansen, T. A. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology Series, Plenum Press/Kluwer, New York.Google Scholar
Kelley, P. H., Hansen, T. A., Graham, S. E., and Huntoon, A. G. 2001. Temporal patterns in the efficiency of naticid gastropod predators during the Cretaceous and Cenozoic of the United States Coastal Plain. Palaeogeography, Palaeoclimatology, Palaeoecology, 166:165176.Google Scholar
Kitchell, J. A. 1986. The Evolution of predator-prey behavior: naticid gastropods and their molluscan prey, p. 88110. In Nitecki, M. H. and Kitchell, J. A. (eds.), Evolution of Animal Behavior: Paleontological and Field Approaches. Oxford University Press, New York.Google Scholar
Kitchell, J. A., Boggs, C. H., Kitchell, J. F., and Rice, J. A. 1981. Prey selection by naticid gastropods: Experimental tests and application to the fossil record. Paleobiology, 7:533552.Google Scholar
Kohn, A. J., and Arua, I. 1999. An Early Pleistocene molluscan assemblage from Fiji: Gastropod faunal composition, paleoecology and biogeography. Palaeogeography, Palaeoclimatology, Palaeoecology, 146:99145.Google Scholar
Kornicker, L. S., Wise, C. D., and Wise, J. M. 1961. Factors affecting the distribution of opposing mollusk valves. Journal of Sedimentary Petrology, 33:703712.Google Scholar
Kosuge, S., and Hayashi, S. 1967. Notes on the feeding habits of Capulus dilatatus A. Adams, 1860 (Gastropoda). Scientific Reports of Yokosuka City Museum, 13:4554.Google Scholar
Kowalewski, M. 1990. A hermeneutic analysis of the shell-drilling gastropod predation on mollusks in the Korytnica Clays (Middle Miocene; Holy Cross Mountains; Central Poland). Acta Geologica Polonica, 40:183213.Google Scholar
Kowalewski, M. 1993. Morphometric analysis of predatory drillholes. Palaeogeography, Palaeoclimatology, Palaeoecology, 102:6988.Google Scholar
Kowalewski, M., and Flessa, K. W. 1994. A predatory drillhole in Glottidia palmeri Dall (Brachiopoda; Lingulidae) from Recent tidal flats of northeastern Baja California, Mexico. Journal of Paleontology, 68:14031405.Google Scholar
Kowalewski, M., and Flessa, K. W. 2000. Seasonal predation by migratory shorebirds recorded in shells of lingulid brachiopods from Baja California, Mexico. Bulletin of Marine Science, 66:405416.Google Scholar
Kowalewski, M., and Nebelsick, J. H. In press. Predation and parasitism on recent and fossil echinoids. In Kelley, P. H., Kowalewski, M., and Hansen, T. A. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology Series, Plenum Press/Kluwer, New York.Google Scholar
Kowalewski, M., Flessa, K. W., and Marcot, J. D. 1997. Predatory scars in the shells of a Recent lingulid brachiopod: Paleontological and ecological implications. Acta Palaeontologica Polonica, 42:497532.Google Scholar
Kowalewski, M., Dulai, A., and Fürsich, F. T. 1998. A fossil record full of holes: The Phanerozoic history of drilling predation. Geology, 26:10911094.Google Scholar
Kowalewski, M., Simões, M. G., Torello, F. F., Mello, L. H. C., and Ghilardi, R. P. 2000. Drill holes in shells of Permian benthic invertebrates. Journal of Paleontology, 74:532543.Google Scholar
Kowalewski, M., Gürs, K., Nebelsick, J. H., Oschmann, W., Piller, W., and Hoffmeister, A. P. 2002. Multivariate hierarchical analyses of Miocene mollusks of Europe: Paleogeographic, biostratigraphic, and paleoecological implications. Geological Society of America Bulletin, 114:239256.Google Scholar
Kusmer, K. D. 1990. Taphonomy of owl pellet deposition. Journal of Paleontology, 64:629637.Google Scholar
Labandeira, C. 2002. Paleobiology of predators, parasitoids, and parasites: Death and accommodation in the fossil record of continental invertebrates. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Lawrence, J. M., and Vasquez, J. 1996. The effects of sublethal predation on the biology of echinoderms. Oceanologica Acta, 19:431440.Google Scholar
Leighton, L. R. 1999. Possible latitudinal predation gradient in middle Paleozoic oceans. Geology, 27:4750.Google Scholar
Leighton, L. R. 2001. New example of Devonian predatory boreholes and the influence of brachiopod spines on predator success. Palaeogeography, Palaeoclimatology, Palaeoecology, 165:5369.Google Scholar
Leighton, L. R. 2002. Inferring predation intensity in the marine fossil record. Paleobiology, 28: in press.Google Scholar
Lescinsky, H. L., and Benninger, L. 1994. Pseudo-borings and predator traces; artifacts of pressure-dissolution in fossiliferous shales. Palaios, 9:599604.Google Scholar
Lever, J., Kessler, A., Van Overbeeke, P., and Thijssen, R. 1961. Quantitative beach research, II, The ‘hole effect’: A second mode of sorting lamellibranch valves on sandy beaches. Netherlands Journal of Sea Research, 1:339358.Google Scholar
Lipps, J. H. 1988. Predation on foraminifera by coral reef fish: Taphonomy and evolutionary implications. Palaios, 3:315326.Google Scholar
Lipps, J. H., and Culver, S. J., 2002. The trophic role of marine microorganisms through geologic time. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation, Paleontological Society Special Papers, 8 (this volume).Google Scholar
Lockley, M. G., and Madsen, J. Jr. 1992. Permian vertebrate trackways from the Cedar Mesa Sandstone of Eastern Utah: Evidence for predator-prey interaction. Ichnos, 2:147153.Google Scholar
Lyman, R. L. 1994. Vertebrate Taphonomy. Cambridge University Press, Cambridge, 524 p.Google Scholar
Marcus, L. F., Corti, M., Loy, A., Naylor, G. J. P., and Slice, D. E. (eds.). 1996. Advances in Morphometrics. NATO ASI Series A: Life Sciences, 284, Plenum, New York, 587 p.Google Scholar
Martill, D. M. 1990. Predation on Kosmoceras by semionotid fish in the Middle Jurassic lower Oxford Clay of England. Palaeontology, 33:739742.Google Scholar
Matsukuma, A. 1977. Notes on Genkaimurex varicosa (Kuroda 1953) (Prosobranchia: Neogastropoda). Venus, 36:8188.Google Scholar
Matsukuma, A. 1978. Fossil boreholes made by shell-boring predators or commensals, Part I: Boreholes of capulid gastropods. Venus, 37:2945.Google Scholar
Mayhew, D. F. 1977. Avian predators as accumulators of fossil mammal material. Boreas, 6:2531.Google Scholar
McRoberts, C. A. 2001. Triassic bivalves and the initial marine Mesozoic revolution; a role for predators? Geology, 29:359362.Google Scholar
Merle, D. 2000. Premiere étude taphonomique de la predation affectant de grands mollusques benthiques dans l'Eocene de Gan (Pyrenées-Atlantiques, France). Comptes Rendus de l'Academie des Sciences, Serie II. Sciences de la Terre et des Planetes, 330:217220.Google Scholar
Miller, R. H., and Sundberg, F. A. 1984. Boring Late Cambrian organisms. Lethaia, 17:185190.Google Scholar
Mordan, P. B. 1977. Factors affecting the distribution and abundance of Aegopinella and Nesovitrea (Pulmonata: Zonitidae) at Monks Wood National Nature Reserve, Huntingdonshire. Biological Journal of the Linnean Society, 9:5972.Google Scholar
Morton, B., and Chan, K. 1997. The first report of shell-boring predation by a representative of the Nassariidae (Gastropoda). Journal of Molluscan Studies, 63:476478.Google Scholar
Moy-Thomas, J. A., and Miles, R. S. 1977. Palaeozoic Fishes, 2nd ed. Saunders, Philadelphia.Google Scholar
Nebelsick, J. H. 1999. Taphonomic legacy of predation on echinoids, p. 347352. In Candia Carnevali, M. D. and Bonasoro, F. (eds.), Echinoderm Research 1998. Balkema, Rotterdam.Google Scholar
Nebelsick, J. H., and Kampfer, S. 1994. Taphonomy of Clypeaster humilis and Echinodiscus auritus from the Red Sea, p. 803808. In David, B. A., Guille, A., Féral, J. P., and Roux, M. (eds.), Echinoderms through Time. Balkema, Rotterdam.Google Scholar
Nebelsick, J. H., and Kowalewski, M. 1999. Drilling predation on Recent clypeasteroid echinoids from the Red Sea. Palaios, 14:127144.Google Scholar
Nedin, C. 1999. Anomalocaris predation on nonmineralized and mineralized trilobites. Geology, 27:987990.Google Scholar
Negus, M. 1975. An analysis of boreholes drilled by Natica catena (Da Costa) in the valves of Donax vittatus (Da Costa). Proceedings of the Malacological Society of London, 41:353356.Google Scholar
Neumann, C. 2000. Evidence of predation on Cretaceous sea stars from northwest Germany. Lethaia, 33:6570.Google Scholar
Nixon, M. 1979. Hole boring in shells by Octopus vulgaris in the Mediterranean. Malacologia, 18:431444.Google Scholar
Oji, T., Ogaya, C., and Sato, T. 2001. Shell fragments in fossil shell beds as a result of durophagous predation: Results from field observation and tumbling experiments. Geological Society of America Abstracts with Programs, 33:A307.Google Scholar
Orr, V. 1962. The drilling habit of Capulus danieli (Crosse) (Mollusca Gastropoda). Veliger, 5:6367.Google Scholar
Ørstan, A. 1999. Drill holes in land snail shells from western Turkey. Schriften zur Malakozoologie aus dem Haus der Natur, 13:3136.Google Scholar
Peterson, C. H., and Black, R. 1995. Drilling by buccinid gastropods of the genus Cominella in Australia. Veliger, 38:3742.Google Scholar
Pickerill, R. K., and Blissett, D. 1999. A predatory Rusophycus burrow from the Cambrian of southern New Brunswick, eastern Canada. Atlantic Geology, 35:179183.Google Scholar
Ponder, W. F., and Taylor, J. D. 1992. Predatory shell drilling by two species of Austroginella (Gastropoda; Marginellidae). Journal of Zoology, 228:317328.Google Scholar
Pratt, B. R. 1998. Probable predation on Upper Cambrian trilobites and its relevance for the extinction of soft-bodied Burgess Shale-type animals. Lethaia, 31:7388.Google Scholar
Pyke, G. H. 1984. Optimal Foraging Theory: A critical review. Annual Reviews of Ecology and Systematics, 15:523575.Google Scholar
Reyment, R. A. 1963. Bohrlöcher bei Ostrakoden (sowie einige paläoethologische Bemerkungen). Paläontologische Zeitschrift, 37:283291.Google Scholar
Reyment, R. A. 1966a. Preliminary observation on gastropod predation in the western Niger Delta. Palaeogeography, Palaeoclimatology, Palaeoecology, 2:81102.Google Scholar
Reyment, R. A. 1966b. Studies on Nigerian Upper Cretaceous and Lower Tertiary Ostracoda, Part 3: Stratigraphical, paleoecological and biometrical conclusions. Stockholm Contributions in Geology, 14:1151.Google Scholar
Reyment, R. A. 1967. Paleoethology and fossil drilling gastropods. Transactions of the Kansas Academy of Science, 70:3350.Google Scholar
Reyment, R. A. 1971. Introduction to Quantitative Paleoecology. Elsevier, Amsterdam, 226 p.Google Scholar
Richards, R. P., and Shabica, C. W. 1969. Cylindrical living burrows in Ordovician dalmanellid brachiopod beds. Journal of Paleontology, 43:838841.Google Scholar
Richter, G., and Baszio, S. 2001. Traces of a limnic food web in the Eocene Lake Messel; a preliminary report based on fish coprolite analyses. Palaeogeography, Palaeoclimatology, Palaeoecology, 166:345368.Google Scholar
Rinaldi, A. C. 1994. Frequency and distribution of Vitreolina philippi (De Rayneval and Ponzi, 1854) (Prosobranchia, Eulimidae) on two regular echinoid species found along the southern coast of Sardinia. Bolletino Malacologia, 30:2932.Google Scholar
Robba, E., and Ostinelli, F. 1975. Studi paleoecologici sul Pliocene Ligure, I: Testimoniaze di predazione sui molluschi Pliocenici di Albenga. Rivista Italiana Paleontologia Stratigraphia, 81:309372.Google Scholar
Rohr, D. M. 1976. Silurian predator borings in the brachiopod Dicaelosia from the Canadian Arctic. Journal of Paleontology, 50:11751179.Google Scholar
Rohr, D. M. 1991. Borings in the shell of an Ordovician (Whiterockian) gastropod. Journal of Paleontology, 65:687688.Google Scholar
Roopnarine, P. D., and Beussink, A. 1999. Extinction and naticid predation of the bivalve Chione von Mühlfeld in the Late Neogene of Florida. Palaeontologia Electronica, 2(1), 24, 719 KB.Google Scholar
Roughley, T. 1925. The perils of the oyster. Australian Museum Magazine, 2:277284.Google Scholar
Roy, K., Miller, D. J., and LaBarbera, M. 1994. Taphonomic bias in analyses of drilling predation: Effects of gastropod drill holes on bivalve shell strength. Palaios, 9:413421.Google Scholar
Savazzi, E., and Reyment, R. A. 1989. Subaerial hunting behavior in Natica gualteriana (naticid gastropod). Palaeogeography, Palaeoclimatology, Palaeoecology, 74:355364.Google Scholar
Schindel, D. E., Vermeij, G. E., and Zipser, E. 1982. Frequencies of repaired shell fractures among the Pennsylvanian gastropods of North-Central Texas. Journal of Paleontology, 56:729740.Google Scholar
Schoener, T. W. 1979. Inferring properties of predation and other injury-producing agents from injury frequencies. Ecology, 60:11101115.Google Scholar
Sheehan, P. M., and Lespérance, P. 1978. Effects of predation on the population dynamics of a Devonian brachiopod. Journal of Paleontology, 52:812817.Google Scholar
Signor, P. W. III, and Brett, C. E. 1984. The mid-Paleozoic precursor to the Mesozoic marine revolution. Paleobiology, 10:229245.Google Scholar
Sliter, W. V. 1971. Predation on benthic foraminifers. Journal of Foraminiferal Research, 1:2029.Google Scholar
Sliter, W. V. 1975. Foraminiferal life and residue assemblages from Cretaceous slope deposits. Geological Society of America Bulletin, 86:897906.Google Scholar
Smith, S. A., Thayer, C. W., and Brett, C. E. 1985. Predation in the Paleozoic: Gastropod-like drillholes in Devonian brachiopods. Science, 230:10331035.Google Scholar
Spencer, W.K., and Wright, C. W. 1966. Asterozoans, p. U4U107. In Moore, R. (ed.), Treatise on Invertebrate Paleontology, Pt. U, Echinodermata 3, Vol. 1. Geological Society of America and University of Kansas Press, New York.Google Scholar
Sohl, N. F. 1969. The fossil record of shell boring by snails. American Zoologist, 9:725734.Google Scholar
Sohn, E. G., and Chatterjee, S., 1979. Freshwater ostracodes from Late Triassic coprolites in Central India. Journal of Paleontology, 53:578586.Google Scholar
Stallibrass, S. 1984. The distinction between the effects of small carnivores and humans on post-glacial faunal assemblages, p. 259269. In Grigson, C. and Clutton-Brock, J. (eds.), Animals and Archaeology: 4. Husbandry in Europe. British Archaeological Reports International Series, 227.Google Scholar
Stallibrass, S. 1990. Canid damage to animal bones: Two current lines of evidence, p. 151165. In Robinson, D. E. (ed.), Experimentation and Reconstruction in Environmental Archaeology. Oxbow Books, Oxford.Google Scholar
Stewart, J. D., and Carpenter, K. 1990. Examples of vertebrate predation on cephalopods in the Late Cretaceous of the Western Interior, p. 203207. In Boucot, A. J. (ed.), Evolutionary Paleobiology of Behavior and Coevolution. Elsevier, Amsterdam.Google Scholar
Stewart, J. D., and Carpenter, K. 1999. Examples of vertebrate predation on cephalopods in the Late Cretaceous of the Western Interior. Bulletin of the Southern California Paleontological Society, 31:6673.Google Scholar
Stewart, K. M., Leblanc, L., Matthiesen, D. P., and West, J. 1999. Microfaunal remains from a modern East African raptor roost; patterning and implications for fossil bone scatters. Paleobiology, 25:483503.Google Scholar
Stuckenrath, R. 1977. Radiocarbon: Some notes from Merlin's diary. Annals of the New York Academy of Sciences, 288:181188.Google Scholar
Taylor, J. D., Morris, N. J., and Taylor, C. N. 1980. Food specialization and the evolution of predatory prosobranch gastropods. Palaeontology, 23:375409.Google Scholar
Taylor, J. D., Cleevely, R. J., and Morris, N. J. 1983. Predatory gastropods and their activities in the Blackdown Greensand (Albian) of England. Palaeontology, 26:521553.Google Scholar
Thomas, R. D. K. 1976. Gastropod predation on sympatric Neogene species of Glycymeris (Bivalvia) from the eastern United States. Journal of Paleontology, 50:488499.Google Scholar
Todd, L. C., and Rapson, D. J. 1988. Long bone fragmentation and interpretation of faunal assemblages; approaches to comparative analysis. Journal of Archaeological Science, 15:307325.Google Scholar
Tsujita, C. J., and Westermann, G. E. G. 2001. Were limpets or mosasaurs responsible for the perforations in the ammonite Placenticeras? Palaeogeography, Palaeoclimatology, Palaeoecology, 169:245270.Google Scholar
Van Valkenburgh, B., and Hertel, F. 1993. Tough times at La Brea: Tooth breakage in large carnivores of the Late Pleistocene. Science, 261:456459.Google Scholar
Van Valkenburgh, B., and Jenkins, I. 2002. Evolutionary patterns in the history of Permo-Triassic and Cenozoic synapsid predators. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Vermeij, G. J. 1977. The Mesozoic marine revolution: Evidence from snails, predators, and grazers. Paleobiology, 3:245258.Google Scholar
Vermeij, G. J. 1980. Drilling predation of bivalves in Guam: Some paleoecological implications. Malacologia, 19:329334.Google Scholar
Vermeij, G. J. 1982. Gastropod shell form, breakage, and repair in relation to predation by the crab Calappa , Malacologia, 23:112.Google Scholar
Vermeij, G. J. 1983. Traces and trends of predation, with special reference to bivalved animals. Palaeontology, 26:455465.Google Scholar
Vermeij, G. J. 1987. Evolution and Escalation. Princeton University Press, Princeton, New Jersey, 527 p.Google Scholar
Vermeij, G. J. 2002. Evolution in the consumer age: Predators and the history of life. In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Vermeij, G. J., and Dudley, E. C. 1982. Shell repair and drilling in some gastropods from the Ripley Formation (Upper Cretaceous) of the Southeastern U.S.A. Cretaceous Research, 3:397403.Google Scholar
Vermeij, G. J., and Lindberg, D. R. 2000. Delayed herbivory and the assembly of marine ecosystems. Paleobiology, 26:419430.Google Scholar
Vermeij, G. J., Zipser, E., and Dudley, E. C. 1980. Predation in time and space: Peeling and drilling in terebrid gastropods. Paleobiology, 6:352364.Google Scholar
Vermeij, G. J., Schindel, D. E., and Zipser, E. 1981, Predation through geological time: Evidence from gastropod shell repair. Science, 214:10241026.Google Scholar
Villa, P., and Mahieu, E. 1991. Breakage patterns of human long bones. Journal of Human Evolution, 21:2748.Google Scholar
Wächtler, V. W. 1927. Zür biologie der Raublungenschnecke Poiretia (Glandina). Algira Brug. Zoologischen Anzeiger, 72:191197.Google Scholar
Walker, S. E. 1989. Hermit crabs as taphonomic agents. Palaios, 4:439452.Google Scholar
Walker, S. E., and Brett, C. E. 2002. Post-Paleozoic patterns in marine predation: Was there a Mesozoic and Cenozoic marine predatory revolution? In Kowalewski, M. and Kelley, P. H. (eds.), The Fossil Record of Predation. Paleontological Society Special Papers, 8 (this volume).Google Scholar
Walker, S. E., and Yamada, S. B. 1993. Implications for the gastropod fossil record of mistaken crab predation on empty mollusc shells. Palaeontology, 36:735741.Google Scholar
Walker, S. E., and Voigt, J. R. 1994. Paleoecologic and taphonomic potential of deepsea gastropods. Palaios, 9:4859.Google Scholar
Warén, A, 1980. Revision of the genera Thyca, Stilifer, Scalenostoma, Mucronalia and Echineulima (Mollusca, Prosobranchia, Eulimidae). Zoologica Scripta, 9:187210.Google Scholar
Warén, A. 1981. Eulimid gastropods parasitic on echinoderms in the New Zealand region. New Zealand Journal of Zoology, 8:313324.Google Scholar
Warén, A., Norris, D. R., and Tempelado, J. T. 1994. Description of four new eulimid gastropods parasitic on irregular sea urchins. Veliger, 37:141154.Google Scholar
Warén, A., and Crossland, M.R. 1991. Revision of Hypermastus Pilsbry, 1899 and Turveria Berry, 1956 (Gastropoda: Prosobranchia: Eulimidae), two genera parasitic on sand dollars. Records of Australian Museum, 43:85112.Google Scholar
Wiltse, W. I. 1980. Predation by juvenile Polinices duplicates (Say) on Gemma gemma (Totton). Journal of Experimental Marine Biology and Ecology, 42:187199.Google Scholar
Wilson, J. B. 1967. Palaeoecological studies on shell-beds and associated sediments in the Solway Firth. Scottish Journal of Geology, 3:329371.Google Scholar
Wilson, M. V. H. 1987. Predation as a source of fish fossils in Eocene lake sediments. Palaios, 2:497504.Google Scholar
Wilson, M. A., and Palmer, T. J. 2001. Domiciles, not predatory borings: A simpler explanation of the holes in Ordovician shells analyzed by Kaplan and Baumiller, 2000. Palaios, 16:524525.Google Scholar
Woelke, C. E. 1957. The flatworm Pseudostylochus ostreophagus Hyman, a predator of oysters. Proceedings of the National Shellfisheries Association, 47:6267.Google Scholar
Wood, R. In press. Predation in ancient reef-builders. In Kelley, P. H., Kowalewski, M., and Hansen, T. A. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology Series, Plenum Press/Kluwer, New York.Google Scholar
Wodinsky, J. 1969. Penetration of the shell and feeding on gastropods by Octopus. American Zoologist, 9:9971010.Google Scholar
Yochelson, E. L., Dockery, D., and Wolf, H. 1983. Predation on sub-Holocene scaphopod mollusks from southern Louisiana. United States Geological Survey Professional Paper, 1282:113.Google Scholar
Yonge, C. M. 1964. Rock borers. Sea Frontiers, 10:106116.Google Scholar
Young, D. K. 1969. Okadaia elegans, a tube-boring nudibranch mollusc from the central and west Pacific. American Zoologist, 9:903907.Google Scholar
Zar, J. H. 1999. Biostatistical analysis, 4th ed. Prentice Hall, Upper Saddle River, 663 p.Google Scholar
Ziegelmeier, E. 1954. Beobachtungen über den Nahrungserwerb bei der Naticide Lunatia nitida Donovan (Gastropoda Prosobranchia). Helgoländer wissentschaftliche Meeresuntersuchungen, 5:133.Google Scholar
Zilch, A. 1959. Euthyneura, p. 1701. In Schindewolf, O. H. (ed.), Handbuch der Paläozoologie, 6. Berlin.Google Scholar
Zotnik, M. 2001. Size-related changes in predatory behaviour of naticid gastropods from the middle Miocene Korytnica Clays, Poland. Acta Palaeontologica Polonica, 46:8797.Google Scholar
Zuschin, M., and Stanton, R. J. Jr. 2001. Experimental measurement of shell strength and its taphonomic interpretation. Palaios, 16:161170.Google Scholar