Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-24T08:13:27.481Z Has data issue: false hasContentIssue false

Evolution in the Consumer Age: Predators and the History of Life

Published online by Cambridge University Press:  21 July 2017

Geerat J. Vermeij*
Affiliation:
Department of Geology, University of California at Davis, One Shields Avenue, Davis, California 95616 USA
Get access

Abstract

Three properties of predation make this form of consumption an important agency of evolution: universality (all species have predators), high frequency (encounters of prey with predators test both parties often), and imperfection (many predatory attacks fail, enabling antipredatory selection to take place). On long time scales, predators have two principal effects: they influence their victims' phenotypes, and prey species that are highly vulnerable to all phases of predatory attacks are evolutionarily restricted to environments where predators are rarely encountered. Although predator and prey can affect each other's behavior and morphology on timescales commensurate with individual lifespans, predators have the evolutionary upper hand over the long run, especially in the expression of sensory capacities, locomotor performance, and the application of force. Only in passive defenses (armor, toxicity, large body size) does escalation favor the prey. In a review of methods for inferring predation in the geological past, I argue against the use of whole assemblages, which combine species of contrasting adaptive type, Instead, I strongly favor species-level and clade-level approaches (including examples of clade replacement) in which comparisons among places and among time intervals are made within the same adaptive types and the same physical environments. The available evidence, much of which comes from studies of shell drilling and shell breakage, points to temporal increases in both predator power and prey defenses. Escalation between species and their enemies, including predators, has proceeded episodically against a backdrop of generally increasing productivity and increasing top-down evolutionary control by high-energy predators during the Phanerozoic, the consumer age.

Type
Section III: Processes
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

Abler, W. L. 1992. The serrated teeth of tyrannosaurid dinosaurs, and biting structures in other animals. Paleobiology, 18:161183.CrossRefGoogle Scholar
Agrawal, A. A. 2001. Phenotypic plasticity in the interactions and evolution of species. Science, 294:321326.CrossRefGoogle ScholarPubMed
Alexander, R. R. 1986. Frequency of sublethal shell-breakage in articulate brachiopod assemblages through geologic time, p. 159166. In Racheboeuf, P. R. and Emig, C. C. (eds.), Biostratigraphie du Paléozoïque 4, Les Brachiopodes fossiles et actuels.Google Scholar
Alexander, R. R. 1990. Mechanical strength of shells of selected extant articulate brachiopods: implications for Paleozoic morphologic trends. Historical Biology, 3:169188.CrossRefGoogle Scholar
Alexander, R. R., and Dietl, G. P. 2001. Latitudinal trends in naticid predation on Anadara ovalis (Bruguière, 1789) and Divalinga quadrisulcata (Orbigny, 1842) from New Jersey to the Florida Keys. American Malacological Bulletin, 16:179194.Google Scholar
Allmon, W. D. 1996. Systematics and evolution of Cenozoic American Turritellidae (Mollusca: Gastropoda), I: Paleocene and Eocene coastal plain species related to “Turritella mortoni Conrad” and “Turritella humerosa Conrad.” Palaeontographica Americana, 59:7134.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
Ambrose, S. H. 2001. Paleolithic technology and human evolution. Science, 291:17481753.CrossRefGoogle ScholarPubMed
Appleton, R. D., and Palmer, A. R. 1988. Water-borne stimuli released by predatory crabs and damaged prey induce more predator-resistant shells in a marine gastropod. Proceedings of the National Academy of Sciences of the United States of America, 85:43874391.CrossRefGoogle Scholar
Babcock, L. E., and Robison, R. A. 1989. Preferences of Palaeozoic predators. Nature, 337:695696.CrossRefGoogle Scholar
Barker, R. T. 1983. The deer flees, the wolf pursues: incongruencies in predator-prey coevolution, p. 350382. In Futuyma, D. J. and Slatkin, M. (eds.), Coevolution. Sinauer, Sunderland, MA.Google Scholar
Baluk, W., and Radwanski, A. 1996. Stomatopod predation upon gastropods from the Korytnica basin, and from other classical Miocene localities in Europe. Acta Geologica Polonica, 76:279304.Google Scholar
Bambach, R. K. 1993. Seafood through time: changes in biomass, energetics, and productivity in the marine ecosystem. Paleobiology, 19:372397.CrossRefGoogle Scholar
Bambach, R. K. 1999. Energetics in the global marine fauna: a connection between terrestrial diversification and change in the marine biosphere. Géobios, 32:131144.CrossRefGoogle 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).CrossRefGoogle 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 Mastoids and their implications for boring by platyceratids. Lethaia, 26:4147.CrossRefGoogle Scholar
Baumiller, T. K. 1996. Boreholes in the Middle Devonian blastoid Heteroschisma and their implications for gastropod drilling. Palaeogeography, Palaeoclimatology, Palaeoecology, 123:343351.CrossRefGoogle Scholar
Baumiller, T. K., and Macurda, D. B. 1995. Borings in Devonian and Mississippian blastoids (Echinodermata). Journal of Paleontology 69:10841089.CrossRefGoogle Scholar
Baumiller, T. K., Leighton, L. R., and Thompson, D. L. 1999. Boreholes in Mississippian spiriferide brachiopods and their implications for Paleozoic gastropod drilling. Palaeogeography, Palaeoclimatology, Palaeoecology, 147:283289.CrossRefGoogle Scholar
Bengtson, S., and Zhao, Y. 1992. Predatorial borings in Late Precambrian mineralized exoskeletons. Science, 257:367369.CrossRefGoogle ScholarPubMed
Blake, D. B. 1990. Adaptive zones of the class Asterioidea (Echinodermata). Bulletin of Marine Science, 46:701718.Google Scholar
Blake, D. B., and Guensburg, T. E. 1994. Predation by the Ordovician asteroid Promopalaeaster on a pelecypod. Lethaia, 27:235239.CrossRefGoogle Scholar
Boyd, D. W., and Newell, N. D. 1972. Taphonomy and diagenesis of a Permian fossil assemblage from Wyoming. Journal of Paleontology, 46:114.Google Scholar
Boyle, P. R., and Knobloch, D. 1981. Hole boring of crustacean prey by the octopus Eledone cirrhosa (Mollusca, Cephalopoda). Journal of Zoology, London, 193:110.CrossRefGoogle Scholar
Briggs, D. E. G. 1994. Giant predators from the Cambrian of China. Science, 264:12831284.CrossRefGoogle ScholarPubMed
Brock, R. E., and Smith, L. D. 1998. Recovery of claw size and function following autotomy in Cancer productus (Decapoda: Brachyura). Biological Bulletin, 194:5362.CrossRefGoogle ScholarPubMed
Burness, G. P., Diamond, J., and Flannery, T. 2001. Dinosaurs, dragons, and dwarfs: the evolution of maximal body size. Proceedings of the National Academy of Sciences, USA, 98:1451814521.CrossRefGoogle ScholarPubMed
Butterfield, N. J. 1997. Plankton ecology in the Proterozoic–Phanerozoic transition. Paleobiology, 23:247262.CrossRefGoogle 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., Walker, S. E., and Flessa, K. W. 1997. Gastropod shell repair in the intertidal of Bahía La Cholla (N. Gulf of California). Palaeogeography, Palaeoclimatology, Palaeoecology, 136:6778.CrossRefGoogle Scholar
Chamberlain, J. A. Jr. 1991. Cephalopod locomotor design and evolution: the constraints of jet propulsion, p. 5798. In Rayner, J. M. V. and Wootton, R. J. (eds.), Biomechanics in Evolution. Cambridge University Press, Cambridge.Google Scholar
Chen, J-Y., Ramskold, L., and Zhou, G-Q. 1994. Evidence for monophyly and arthropod affinity of Cambrian giant predators. Science, 264:13041308.CrossRefGoogle ScholarPubMed
Collins, D. 1996. The “evolution” of Anomalocaris and its classification in the arthropod class Dinocarida (nov.) and order Radiodonta (nov.). Journal of Paleontology, 70:280293.CrossRefGoogle Scholar
Conway Morris, S. 1979. The Burgess Shale (Middle Cambrian) fauna. Annual Review of Ecology and Systematics, 10:327349.CrossRefGoogle Scholar
Conway Morris, S., and Bengtson, S. 1994. Cambrian predators: possible evidence for boreholes. Journal of Paleontology, 68:123.CrossRefGoogle Scholar
Coovert, G. A., and Coovert, P. K. 1995. Revision of the supraspecific classification of marginelliform gastropods. Nautilus, 109:43110.Google Scholar
Currie, P. J. 1997. Theropods, p. 216233. In Farlow, J. O. and Brett-Surman, M. M. (eds.), The Complete Dinosaur. Indiana University Press, Bloomington, IN.Google Scholar
Daniel, T.L., Helmuth, B.S., Saunders, W. B., and Ward, P. D. 1997. Septal complexity in ammonoid cephalopods increased mechanical risk and limited depth. Paleobiology, 23:470481.CrossRefGoogle Scholar
Dayton, P. K., Rosenthal, R. J., Mahen, L. C., and Antezana, T. 1977. Population structure and foraging biology of the predaceous Chilean asteroid Meyenaster gelatinosus and the escape biology of its prey. Marine Biology, 39:361370.CrossRefGoogle Scholar
Dietl, G. P., and Alexander, R. R. 1998. Shell repair frequencies in whelks and moon snails from Delaware and southern New Jersey. Malacologia, 39:151165.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, 15:414429.2.0.CO;2>CrossRefGoogle 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, 26:215237.2.0.CO;2>CrossRefGoogle Scholar
Dodge, R., and Scheel, D. 1999. Remains of the prey—recognizing the midden piles of Octopus dofleini (Wülker). Veliger, 32:260266.Google Scholar
Ebbestad, J. O. R. 1998. Multiple attempted predation in the Middle Ordovician gastropod Bucania gracillima. GFF, 120:2733.CrossRefGoogle Scholar
Ebbestad, J. O. R., and Peel, J. S. 1997. Attempted predation and shell repair in Middle and Upper Ordovician gastropods from Sweden. Journal of Paleontology, 71:10071019.CrossRefGoogle Scholar
Erickson, G. M., Van Kirk, S. D., Su, G., Levenston, M. E., Caler, W. E., and Carter, D. R. 1996. Bite-force estimation of Tyrannosaurus rex from tooth-marked bones. Nature, 382:706708.CrossRefGoogle Scholar
Fürsich, F. T., and Jablonski, D. 1984. Late Triassic naticid drillholes: carnivorous gastropods gain a major adaptation but fail to radiate. Science, 224:7880.CrossRefGoogle Scholar
Gabbott, S. E., Aldridge, R. J., and Theron, J. R. 1995. A giant conodont with preserved muscle tissue from the Upper Ordovician of South Africa. Nature, 374:800803.CrossRefGoogle Scholar
Garvie, C. L. 1991. Two new species of Muricinae from the Cretaceous and Paleocene of the Gulf Coastal Plain, with comments on the genus Odontopolys Gabb, 1860. Tulane Studies in Geology and Paleontology, 24:8792.Google Scholar
Geary, D. H., Allmon, W. D., and Reaka-Kudla, M. L. 1991. Stomatopod predation on fossil gastropods from the Plio-Pleistocene of Florida. Journal of Paleontology, 65:355360.CrossRefGoogle Scholar
Gordillo, S., and Muchástegui, S. N. 1998. Estratégias de depredación del gastrópodo perforador Trophon geversianus (Pallas) (Muricoidea: Trophoninae). Malacologia, 39:8391.Google Scholar
Greene, H. W. 1997. Snakes: The Evolution of Mystery in Nature. University of California Press, Berkeley.CrossRefGoogle Scholar
Guerrero, R., Pedros-Alió, C., Esteve, I., Mas, J., Chase, D., and Margulis, L. 1986. Predatory prokaryotes: Predation and primary consumption evolved in bacteria. Proceedings of the National Academy of Sciences of the United States of America, 83:21382142.CrossRefGoogle ScholarPubMed
Hagadorn, J. W., and Boyajian, G. E. 1997. Subtle changes in mature predator-prey systems: an example from Neogene Turritella (Gastropoda). Palaios, 12:372379.CrossRefGoogle Scholar
Hansen, T. A., and Kelley, P. H. 1995. Spatial variation of naticid gastropod predation in the Eocene of North America. Palaios, 10:268278.CrossRefGoogle Scholar
Harper, E. M. 1991. The role of predation in the evolution of cementation in bivalves. Palaeontology, 34:455460.Google Scholar
Harper, E. M. 1994. Are conchiolin sheets in corbulid bivalves primarily defensive? Palaeontology, 37:551578.Google Scholar
Harper, E. M., and Skelton, P. W. 1993a. A defensive value of the thickened periostracum in the Mytiloidea. Veliger, 36:3642.Google Scholar
Harper, E. M., and Skelton, P. W. 1993b. The Mesozoic marine revolution and epifaunal bivalves. Scripta Geologica, Special Issue 2:127153.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.CrossRefGoogle 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, 16:566569.2.0.CO;2>CrossRefGoogle Scholar
Jacobs, D. K. 1992. Shape, drag, and power in ammonoid swimming. Paleobiology, 18:203220.CrossRefGoogle Scholar
Jones, E. C. 1971. Isistius brasiliensis, a squaloid shark, the probable cause of crater wounds on fishes and cetaceans. Fisheries Bulletin, 69:791798.Google Scholar
Juanes, F., and Smith, L. D. 1995. The ecological consequences of limb damage and loss in decapod crustaceans: a review and prospectus. Journal of Experimental Marine Biology and Ecology, 193:197223.CrossRefGoogle Scholar
Kabat, A. R. 1990. Predatory ecology of naticid gastropods with a review of shell boring predation. Malacologia, 32:155193.Google Scholar
Kardon, G. 1998. Evidence from the fossil record of an antipredatory exaptation: conchiolin layers in corbulid bivalves. Evolution, 52:6879.CrossRefGoogle ScholarPubMed
Kelley, P. H. 1989. Evolutionary trends within bivalve prey of Chesapeake Group naticid gastropods. Historical Biology, 2:139156.CrossRefGoogle Scholar
Kelley, P. H. 1992. Coevolutionary patterns of naticid gastropods of the Chesapeake Group: an example of coevolution? Journal of Paleontology, 66:794800.CrossRefGoogle 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.CrossRefGoogle Scholar
Kelley, P. H., and Hansen, T. A. 1996. Recovery of the naticid gastropod predator-prey system from the Cretaceous-Tertiary and Eocene-Oligocene extinctions, p. 373386. In Hart, M. B. (ed.), Biotic Recovery from Mass Extinction Events. Geological Society Special Publication 102.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.CrossRefGoogle Scholar
Kier, P. M. 1981. A bored Cretaceous echinoid. Journal of Paleontology, 55:656659.Google Scholar
Kirby, M. X. 2001. Differences in growth rate and environment between Tertiary and Quaternary Crassostrea oysters. Paleobiology, 27:84103.2.0.CO;2>CrossRefGoogle Scholar
Knoll, A. H., and Bambach, R. K. 2000. Directionality in the history of life: diffusion from the left wall or repeated scaling of the right? Paleobiology, 26 (Supplement to No. 4):114.CrossRefGoogle Scholar
Kojumdgieva, E. 1975. Les gastéropodes perceurs et leurs victimes du Miocene de Bulgarie du nord-ouest. Bulgarian Academy of Sciences, Bulletin of the Geological Institute (Series Paleontology), 25:524.Google Scholar
Kowalewski, M., Dulai, A., and Fürsich, F. T. 1998. A fossil record full of holes: the Phanerozoic history of drilling predation. Geology, 26:10911094.2.3.CO;2>CrossRefGoogle Scholar
Kowalewski, M., Simões, M. G., Torello, F. F., Mello, L. H. C., and Ghilardi, R. P. 2000. Drillholes in shells of Permian benthic invertebrates. Journal of Paleontology, 74:532543.2.0.CO;2>CrossRefGoogle Scholar
Kroger, B. 2000. Schalenverletzungen an Jurassischen Ammoniten—ihre paläontologische und paläoökologische Aussagefähigkeit. Berliner Geowissenschaftliche Abhandlungen (E), 33:196.Google Scholar
Landman, N. H., and Waage, K. M. 1986. Shell abnormalities in scaphitid ammonites. Lethaia, 19:211224.CrossRefGoogle Scholar
Lee, S. Y., and Seed, R. 1992. Ecological implications of cheliped size in crabs: some data from Carcinus maenas and Liocarcinus holsatus. Marine Ecology Progress Series, 84:151160.CrossRefGoogle 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.CrossRefGoogle Scholar
Leonard, G. H., Bertness, M. D., and Yund, P. O. 1999. Crab predation, waterborne cues, and inducible defenses in the blue mussel, Mytilus edulis. Ecology, 80:114.CrossRefGoogle Scholar
Lindberg, D. R., and Ponder, W. F. 2001. The influence of classification on the evolutionary interpretation of structure—a re-evaluation of the evolution of the pallial cavity of gastropod molluscs. Organisms, Diversity, and Evolution, 1:273299.CrossRefGoogle Scholar
Margulis, L. 1981. Symbiosis in cell evolution: life and its environment on the early Earth. W. H. Freeman, San Francisco.Google Scholar
Merle, D. 2000. Première étude taphonomique de la predation affectant de grands mollusques benthiques dans l'Eocene de Gan (Pyrénées Atlantiques, France). Comptes Rendus de l'Académie des Sciences de Paris, Sciences de la Terre et des Planètes, 330:217220.Google Scholar
Merle, D., and Pacaud, J-M. 2001. The first record of Poirieria subcristata (d'Orbigny, 1850) (Muricidae: Muricinae) in the early Cuisian of the Paris Basin (Celles-sur-Aisne, Aizy Formation), with comments on the sculptural evolution of some Palaeocene and Eocene Poirieria and Paziella. Tertiary Research, in press.Google Scholar
Monks, N. 2000. Mid-Cretaceous heteromorph ammonite shell damage. Journal of Molluscan Studies, 66:283285.CrossRefGoogle 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
Palmer, A. R. 1982. Predation and parallel evolution: recurrent parietal plate reduction in balanomorph barnacles. Paleobiology, 8:3144.CrossRefGoogle Scholar
Palmer, A. R. 1990. Effect of crab effluent and scent of damaged conspecifics on feeding, growth, and shell morphology of the Atlantic dogwhelk (Nucella lapillus (L.)). Hydrobiologia, 193:155182.CrossRefGoogle Scholar
Pan, H.-Z. 1991. Lower Turonian gastropod ecology and biotic interaction in Helicaulax community from western Tarim Basin, southern Xinjiang, China. Paleoecology of China, 1:266280.Google Scholar
Pether, J. 1995. Belichnus new ichnogenus, a ballistic trace on mollusc shells from the Holocene of the Benguela region, South Africa. Journal of Paleontology, 69:171181.CrossRefGoogle Scholar
Ponder, W. F., and Taylor, J. D. 1992. Predatory shell drilling by two species of Austroginella (Gastropoda: Marginellidae). Journal of Zoology, London, 228:317328.CrossRefGoogle Scholar
Purnell, M. A. 1995. Microwear on conodont elements and macrophagy in the first vertebrates. Nature, 374:798800.CrossRefGoogle Scholar
Rayfield, E. J., Norman, D. B., Horner, C. C., Harner, J. R., Smith, P. M., Thomason, J. J., and Upchurch, P. 2001. Carnial design and function in a large theropod dinosaur. Nature, 409:10331037.CrossRefGoogle Scholar
Riedel, F. 1995. An outline of cassoidean phylogeny (Mollusca: Gastropoda). Contributions to Tertiary and Quaternary Geology, 32:97132.Google Scholar
Riedel, F. 2000. Ursprung und Evolution der “höheren” Caenogastropoda. Berliner Geowissenschaftliche Abhandlungen (E), 32:1240.Google Scholar
Rudebeck, G. 1950–1951. The choice of prey and methods of hunting of predatory birds with special reference to their selective effect. Oikos, 2:6588; 3:200–231.CrossRefGoogle Scholar
Saunders, W. B., Work, D. M., and Nikolaeva, S. V. 1999. Evolution of complexity in Paleozoic ammnonoid sutures. Science, 286:760763.CrossRefGoogle ScholarPubMed
Schindel, D. E., Vermeij, G. J., and Zipser, E. 1982. Frequencies of repaired shell fractures among the Pennsylvanian gastropods from north-central Texas. Journal of Paleontology, 56:729740.Google Scholar
Schmidt, N. 1989. Paleobiological implications of shell repair in Recent marine gastropods from the northern Gulf of California. Historical Biology, 3:127139.CrossRefGoogle Scholar
Signor, P. W. III, and Brett, C. E. 1984. The mid-Paleozoic precursor to the Mesozoic marine revolution. Paleobiology, 10:229245.CrossRefGoogle Scholar
Smith, L. D. 1992. The impact of limb autotomy on mate competition in blue crabs Callinectes sapidus Rathbun. Oecologia, (Berlin) 89:494501.CrossRefGoogle ScholarPubMed
Smith, L. D., and Jennings, J. A. 2000. Induced defensive responses in the bivalve Mytilus edulis to predators with different attack modes. Marine Biology, 136:461469.CrossRefGoogle Scholar
Smith, L. D., and Palmer, A. R. 1994. Effects of manipulated diet on size and performance of brachyuran crab claws. Science, 264:710712.CrossRefGoogle ScholarPubMed
Smith, S. A., Thayer, C. W., and Brett, C. E. 1985. Predation in the Paleozoic: gastropod-like drillholes in Devonian brachiopods. Science, 230:10331035.CrossRefGoogle ScholarPubMed
Stone, H. M. I. 1998. On predator deterrence by pronounced shell ornament in epifaunal bivalves. Palaeontology, 41:10511068.Google Scholar
Sues, H. D. 1991. Venom-conducting teeth in a Triassic reptile. Nature, 351:141143.CrossRefGoogle 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
Tintori, A. 1998. Fish biodiversity in the marine Norian (Late Triassic) of northern Italy: the first neopterygian radiation. Italian Journal of Zoology, 65 (supplement):193199.CrossRefGoogle Scholar
Van Valkenburgh, B. 1991. Iterative evolution of hypercarnivory in canids (Mammalia: Carnivora). Paleobiology, 17:340362.CrossRefGoogle 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.CrossRefGoogle Scholar
Vannier, J., and Chen, J.-Y. 2000. The Early Cambrian colonization of pelagic niches exemplified by Isoxys (Arthropoda). Lethaia, 33:295311.CrossRefGoogle Scholar
Vermeij, G. J. 1975. Evolution and distribution of left-handed and planispiral coiling in snails. Nature, 254:419420.CrossRefGoogle Scholar
Vermeij, G. J. 1977. The Mesozoic marine revolution: evidence from snails, predators and grazers. Paleobiology, 3:245258.CrossRefGoogle Scholar
Vermeij, G. J. 1979. Shell architecture and causes of death in Micronesian reef snails. Evolution, 33:686696.CrossRefGoogle ScholarPubMed
Vermeij, G. J. 1980. Drilling predation of bivalves in Guam: some paleoecological implications. Malacologia, 19:329334.Google Scholar
Vermeij, G. J. 1982a. Unsuccessful predation and evolution. American Naturalist, 120:701720.CrossRefGoogle Scholar
Vermeij, G. J. 1982b. Gastropod shell form, repair, and breakage in relation to predation by the crab Calappa. Malacologia, 23:112.Google Scholar
Vermeij, G. J. 1987. Evolution and Escalation, an Ecological History of Life. Princeton University Press, Princeton.CrossRefGoogle Scholar
Vermeij, G. J. 1989. The origin of skeletons. Palaios, 5:585589.CrossRefGoogle Scholar
Vermeij, G. J. 1993. A Natural History of Shells. Princeton University Press, Princeton.Google Scholar
Vermeij, G. J. 1994. The evolutionary interaction among species: selection, escalation, and coevolution. Annual Review of Ecology and Systematics, 25:219236.CrossRefGoogle Scholar
Vermeij, G. J. 2001. Innovation and evolution at the edge: origins and fates of gastropods with a labral tooth. Biological Journal of the Linnean Society, 72:461508.CrossRefGoogle Scholar
Vermeij, G. J., and Zipser, E. 1986. Burrowing performance of some tropical Pacific gastropods. Veliger, 29:200206.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.CrossRefGoogle Scholar
Vermeij, G. J., Schindel, D. E., and Zipser, E. 1981. Predation through geological time: evidence from gastropod shell repair. Science, 214:10241026.CrossRefGoogle ScholarPubMed
Ward, P. 1986. Cretaceous ammonite shell shapes. Malacologia, 27:328.Google Scholar
Wainwright, P. C., and Bellwood, D. R. In Press. Ecomorphology of feeding in coral reef fishes, p. 3355. In Sale, P. F. (ed.), Coral reef fishes: dynamics and diversity in a complex ecosystem. Academic Press, San Diego.Google Scholar
Zipser, E., and Vermeij, G. J. 1980. Survival after non-lethal shell damage in the gastropod Conus sponsalis. Micronesica, 16:229234.Google Scholar
Zullo, V. A. 1984. New genera and species of balanoid barnacles from the Oligocene and Miocene of North Carolina. Journal of Paleontology, 58:13121338.Google Scholar