Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-23T16:24:30.684Z Has data issue: false hasContentIssue false

Predators and Predation in Paleozoic Marine Environments

Published online by Cambridge University Press:  21 July 2017

Carlton E. Brett
Affiliation:
Department of Geology, University of Cincinnati, Cincinnati, Ohio 45221-0013 USA
Sally E. Walker
Affiliation:
Department of Geology, University of Georgia, Athens, Georgia 30602-2501 USA
Get access

Abstract

The Paleozoic body fossil record of potential benthic predators includes nautiloid and ammonoid cephalopods, phyllocarids, decapods, and several lineages of gnathostomes. The latter group, in particular, radiated rapidly during the Devonian. In the pelagic realm, predator-prey interactions involving cephalopods and some nektonic arthropods probably appeared in the Ordovician. Again, evidence indicates intensification of pelagic predation, much of it by arthrodires and sharks on other fishes, during the Devonian radiation of gnathostomes.

Trace fossils provide direct evidence of predatory attack from the Ediacarian and Early Cambrian onward, but with a substantial increase in the Siluro-Devonian. Brachiopod and molluscan shells and trilobite exoskeletons show evidence of healed bite marks and peeling from the Cambrian onward, but with an increased frequency in the Devonian. Predatory drill holes with stereotypical position and prey-species preference are found in brachiopods (Cambrian onward) and mollusks (Ordovician onward); boreholes also show increased frequency in the middle Paleozoic. Certain of these boreholes are tentatively attributable to platyceratid gastropods.

Hard-shelled benthic organisms with thicker, more spinose skeletons may have had a selective advantage as durophagous predators increased. Brachiopods, gastropods, trilobites, and crinoids show an abrupt increase in spinosity beginning in the Siluro-Devonian. But spinosity decreases after the early Carboniferous. Late Paleozoic benthos may have taken refuge in smaller size and resistant, thick-walled skeletons, as well as endobenthic and cementing modes of life. Conversely, in the pelagic realm, external armor was reduced, while more efficient, fast-swimming modes of life (e.g., in sharks) increased in the post-Devonian.

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

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 and repair of shell breakage induced by durophages in Late Ordovician brachiopods. Journal of Paleontology, 60:273285.CrossRefGoogle Scholar
Alexander, R. R. 1986b. Frequency of sublethal shell-breakage in articulates through geologic time, p. 159166. In Racheboeuf, P. R. and Emig, C. (eds.), Les Brachiopodes Fossiles et Actuels. Université de Bretagne Occidentale. Biostratigraphie du Paleozoique.Google Scholar
Alexander, R. R. 2001. Functional morphology and biomechanics of articulate brachiopod shells, p. 145170. In Carlson, S. J. and Sandy, M. (eds.), Brachiopods Ancient and Modern: A Tribute to G. Arthur Cooper. Paleontological Society Papers, 7.Google Scholar
Aronson, R. B. 1991. Escalating predation on crinoids in the Devonian: Negative community-level evidence. Lethaia, 24:123128.CrossRefGoogle Scholar
Ausich, W. I., and Bother, D. J. 1982. Tiering in suspension-feeding communities on soft substrata throughout the Phanerozoic. Science, 216:173174.CrossRefGoogle ScholarPubMed
Ausich, W. I., and Gurrola, R. A. 1979. Two boring organisms in Lower Mississippian community of southern Indiana. Journal of Paleontology, 53:335344.Google Scholar
Ausich, W. I., Kammer, T. W., and Baumiller, T. K. 1994. Demise of the middle Paleozoic crinoid fauna: A single extinction event or rapid faunal turnover. Paleobiology, 20:345361.CrossRefGoogle Scholar
Babcock, L. E. 1993. Trilobite malformations and the fossil record of behavioral symmetry. Journal of Paleontology, 67:217229.Google Scholar
Babcock, L. E. In press. Trilobites in Paleozoic predator-prey systems and their role in reorganization of early Paleozoic ecosystems. In Kelley, P. H., Kowalewski, M., and Hansen, T. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology, Kluwer Academic/Plenum Publishers, New York.Google Scholar
Babcock, L. E., and Robison, R. A. 1989. Preferences of Paleozoic predators. Nature, 337:695696.Google Scholar
Bambach, R. K. 1993. Seafood through time: Changes in biomass, energetics, and productivity in the marine ecosystem. Paleobiology, 19:372397.Google Scholar
Bandel, K. 1985. Cephalopod morphology and function, p. 190. In Bottjer, D. J., Hickman, C. S., and Ward, P. D. (eds.), Mollusks: Notes for a Short Course. University of Tennessee Department of Geological Sciences, Studies in Geology 13.Google Scholar
Basch, L. V., and Engle, J. M. 1989. Aspects of the ecology and behaviour of the stomatopod Hemisquilla ensigera californiensis (Gonodactyloidea: Hemisquillidae), p. 199212. In Ferrero, E. (ed.), Biology of Stomatopods. Selected Symposia and Monographs, U. Z. I., Volume 3.Google Scholar
Bassett, M. G., and Bryant, C. 1993. The micromorphic rhynchonelloidean brachiopod Lambdarina from the type Dinantian. Journal of Paleontology, 67:518527.CrossRefGoogle Scholar
Baumiller, T. K. 1990. Non-predatory drilling of Mississippian crinoids by platyceratid gastropods. Palaeontology, 33:743748.Google Scholar
Baumiller, T. K. 1996. Boreholes in Devonian blastoids and their implications for boring platyceratids. Lethaia, 26:4147.Google 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.Google Scholar
Bengtson, S. 1994. The advent of animal skeletons, p. 412425. In Bengtson, S. (ed.), Early Life on Earth. Columbia University Press, New York.Google Scholar
Bengtson, S., and Zhao, Y. 1992. Predatorial borings in late Precambrian mineralized exoskeletons. Science, 257:367369.Google Scholar
Benton, M. J. 1997. Vertebrate Palaeontology, 2nd edition. Chapman and Hall, London, New York, 452 p.Google Scholar
Bergström, J. 1973. Organisation, life and systematics of trilobites. Fossils and Strata, 21:169.Google Scholar
Bishop, G. A. 1975. Traces of predation, p. 267281. In Frey, R. (ed.), The Study of Trace Fossils. Springer-Verlag.Google Scholar
Blake, D. B., and Guensburg, T. E. 1992. Predatory asteroids and the fate of articulate brachiopods. Lethaia, 23:429430.CrossRefGoogle Scholar
Blake, D. B., and Guensburg, T. E. 1994. Predation by the Ordovician asteroid Promopalaeaster on a pelecypod. Lethaia, 27:235238.CrossRefGoogle Scholar
Bond, P. N., and Saunders, W. B. 1989. Sublethal shell injury and shell repair in Upper Mississippian ammonoids. Paleobiology, 15:414428.CrossRefGoogle Scholar
Bordeaux, Y. L., and Brett, C. E. 1990. Substrate specific associations on Middle Devonian brachiopods: Implications for paleobiology. Historical Biology, 4:203220.CrossRefGoogle Scholar
Bottjer, D. J. 1985. Bivalve paleoecology, p. 112137. In Bottjer, D. J., Hickman, C. S., and Ward, P. D. (eds.), Mollusks: Notes for a Short Course. University of Tennessee Department of Geological Sciences, Studies in Geology 13.Google Scholar
Boucot, A. J. 1981. Principles of Benthic Marine Paleoecology. Academic Press, NewYork, 463 p.Google Scholar
Boucot, A. J. 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier, Amsterdam, 725 p.Google Scholar
Bowsher, A. L. 1955. Origin and adaptation of platyceratid gastropods. University of Kansas Paleontological Contributions, Mollusca, 5:111.Google Scholar
Boyd, D. W., and Newell, N. D. 1972. Taphonomy and diagenesis of a Permian fossil assemblage from Wyoming. Journal of Paleontology, 46:117.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. 1992. Predation, 4.13.1, Marine, p. 368372. In Briggs, D. E. G. and Crowther, P. W. (eds.), Paleobiology: A Synthesis. Blackwell Scientific, Oxford.Google Scholar
Brett, C. E. In press. Durophagous predation on Paleozoic marine benthic assemblages. In Kelley, P. H., Kowalewski, M., and Hansen, T. (eds.), Predator-Prey Interactions in the Fossil Record. Topics in Geobiology, Kluwer Academic/Plenum Publishers, New York.Google Scholar
Brett, C. E., and Cottrell, J. F. 1982. Substrate specificity of the Devonian tabulate coral Pleurodictyum. Lethaia, 15:248263.Google Scholar
Briggs, D. E. G., Erwin, D. H., and Collier, F. J. 1994. The Fossils of the Burgess Shale. Smithsonian Institution Press, Washington, D.C. and London.Google Scholar
Bromley, R. G. 1981. Concepts in ichnotaxonomy illustrated by small, round holes in shells. Acta Geologica Hispanica, 16:5564.Google Scholar
Brunton, H. 1966. Predation and shell damage in a Visean brachiopod fauna. Palaeontology, 9:355359.Google Scholar
Brunton, C. H. C., and Champion, C. 1974. A Lower Carboniferous brachiopod fauna from the Manifold Valley, Staffordshire. Palaeontology, 17:107111.Google Scholar
Bucher, W. H. 1938. A shell-boring gastropod in a Dalmanella bed of upper Cincinnatian age. American Journal of Science, 36:17.Google Scholar
Buehler, E. J. 1969. Cylindrical borings in Devonian shell. Journal of Paleontology, 43:1291.Google Scholar
Cameron, B. 1967. Oldest carnivorous gastropod borings found in Trentonian (Middle Ordovician) brachiopods. Journal of Paleontology, 41:147150.Google Scholar
Carriker, M. 1969. Penetration of calcium carbonate substrates by lower plants and invertebrates. American Zoologist, 9:6291020.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. J., and Yochelson, E. L. 1968. Recent gastropod boreholes and Ordovician cylindrical borings. US Geological Survey Professional Paper, B593:126.Google Scholar
Chatterton, B., and Whitehead, H. L. 1987. Predatory borings in inarticulate brachiopods Artiotrema from the Silurian of Oklahoma. Lethaia, 20:6774.CrossRefGoogle Scholar
Clarke, J. M. 1921. Organic dependence and disease: Their origin and significance. New York State Museum Bulletin 221/222. 113 p.Google Scholar
Clarkson, E. N. K. 1996. Invertebrate Palaeontology and Evolution, 3rd edition. Chapman and Hall, London, 434 p.Google Scholar
Connell, J. H. 1970. A predator-prey system in the marine intertidal region: I Balanus glandula and several predatory species of Thais. Ecological Monographs, 40:4978.Google Scholar
Conway Morris, S. 1977. Fossil priapulid worms. Special Papers in Palaeontology, 20:195.Google Scholar
Conway Morris, S. 2001. Significance of early shells, p. 3140. In Briggs, D. E. G. and Crowther, P. R. (eds.), Paleobiology II. Blackwell Science, Oxford.CrossRefGoogle Scholar
Conway Morris, S., and Bengtson, S. 1994. Cambrian predators: Possible evidence from boreholes. Journal of Paleontology, 68:123.Google Scholar
Conway Morris, S., and Jenkins, R. J. F. 1985. Healed injuries in Early Cambrian trilobites from South Australia. Alcheringa, 9:167177.Google Scholar
Conway Morris, S., and Robison, R. A. 1986. Middle Cambrian priapulids and other soft-bodied fossils from Utah and Spain. University of Kansas Paleontological Contributions, Paper 117:117.Google Scholar
Conway Morris, S., and Robison, R. A. 1988. More soft-bodied animals and algae from the Middle Cambrian of Utah and British Columbia. University of Kansas Paleontological Contributions, Paper 122:148.Google Scholar
Cooper, G. A. 1956. New Pennsylvanian brachiopods. Journal of Paleontology, 30:521530.Google Scholar
Donovan, S. T., and Gale, A. S. 1990. Predatory asteroids and the decline of the articulate brachiopods. Lethaia, 23:7786.Google Scholar
Ebbestad, J. O. R. 1998. Multiple attempted predation in the Middle Ordovician gastropod Bucania gracillima. Geologiska Förenings I Stockholm Forhandlingar, 120:2733.Google 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, 7:10071019.Google Scholar
Elliot, D. K., and Brew, D. C. 1988. Cephalopod predation on a Desmoinesian brachiopod from the Naco Formation, central Arizona. Journal of Paleontology, 62:145147.Google Scholar
Fenton, C. L., and Fenton, M. A. 1931. Some borings of Paleozoic age. American Midland Naturalist, 12:522528.Google Scholar
Fenton, C. L., and Fenton, M. A. 1932. Orientation and injury in the Devonian genus Atrypa. American Midland Naturalist, 13:6370.CrossRefGoogle Scholar
Fortey, R. A., and Owens, R. M. 1999. Feeding habits in trilobites. Palaeontology, 42:429465.Google Scholar
Gould, S. J. 1989. Wonderful Life: The Burgess Shale and the Nature of History. Norton, New York, London, 347 p.Google Scholar
Gould, S. J. 2002. The Structure of Evolutionary Theory. Belknap-Harvard Press, 1433 p.Google Scholar
Gould, S. J., and Vrba, E. S. 1982. Exaptation—a missing term in the science of form. Paleobiology, 8:415.Google Scholar
Grant, R. E. 1966. Spine arrangement and life habits of the productoid brachiopod Waagenoconcha. Journal of Paleontology, 40:10631069.Google Scholar
Grant, R. E. 1988. The Family Cardiarinidae (Late Paleozoic rhynchonellid Brachiopoda). Senckenbergiana Lethaia, 69:121135.Google Scholar
Hansen, M. C., and Mapes, R. H. 1990. A predator-prey relationship between sharks and cephalopods in the late Paleozoic, p. 189192. In Boucot, A. J., Evolutionary Paleobiology of Behavior and Coevolution. Elsevier, Amsterdam.Google Scholar
Häntzshel, W., El-Baz, F., and Amstutz, G. C. 1968. Coprolites: An annotated bibliography. Memoir of the Geological Society of America, 108:1132.Google Scholar
Harper, E. M. 1991. The role of predation in the evolution of cementation in bivalves. Palaeontology, 34:455460.Google Scholar
Hlavin, W. J. 1973. A review of the vertebrate fauna of the Upper Devonian (Famennian) Cleveland Shale: Arthrodira. , Boston University, Boston, Mass.Google Scholar
Hlavin, W. J. 1990. Arthrodire-ctenacanth shark, p. 192195. In Boucot, A. J., Evolutionary Paleobiology of Behavior and Coevolution. Elsevier, Amsterdam.Google Scholar
Hof, C. H. J. 1998. Fossil Stomatopods (Crustacea: Malacostraca) and their phylogenetic impact. Journal of Natural History, 32:15671576.Google Scholar
Hoffmeister, A. P., Kowalewski, M., Bambach, R. K., and Baumiller, T. K. 2001a. Intense drilling predation on the brachiopod Cardiarina cordata (Cooper, 1956) from the Pennsylvanian of New Mexico. Geological Society of America Abstracts with Programs, 33:A248.Google Scholar
Hoffmeister, A. P., Kowalewski, M., Bambach, R. K., and Baumiller, T. K. 2001b. Evidence for predatory drilling in late Paleozoic brachiopods and bivalve mollusks from West Texas. North American Paleontological Convention, Paleobios, 21 (Supplement 2):6667.Google Scholar
Hoffmeister, A. P., Kowalewski, M., Bambach, R. K., and Baumiller, T. K. 2002. A boring history of drilling predation on the Paleozoic brachiopod Composita. Geological Society of America Abstracts with Programs, 34:A116.Google Scholar
Jenner, R. A., Hof, C. H. J., and Schram, F. R. 1998. Palaeo- and archaeostomatopods (Hoplocarida: Crustacea) from the Bear Gulch Limestone, Mississippian (Namurian), of Central Montana. Contributions to Zoology, 67:155185.Google Scholar
Kabat, A. R. 1990. Predatory ecology of naticid gastropods with a review of shell boring predation. Malacologia, 32:155–103.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
Kesling, R.V., and Chilman, R. B. 1975. Strata and Megafossils of the Middle Devonian Silica Formation. University of Michigan Museum of Paleontology, Papers on Paleontology 8, 408 p.Google Scholar
Kloc, G. J. 1987. Coprolites containing ammonoids from the Devonian of New York. Geological Society of America, Abstracts with Programs, 19:23.Google Scholar
Kloc, G. J. 1992. Spine function in the odontopleurid trilobites Leonaspis and Dicranurus from the Devonian of Oklahoma. North American Paleontological Convention Abstracts and Program. Paleontological Society Special Publication 6:167.Google Scholar
Kluessendorf, J., and Doyle, P. 2000. Pohlsepia mazonensis, an early ‘octopus’ from the Carboniferous of Illinois, USA. Palaeontology, 43:919926.Google Scholar
Knoll, A. S., Bambach, R. K., Grotzinger, J. P., and Canfield, D. 1996. Comparative Earth history and Late Permian mass extinction. Science, 273:452457.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. W., Torello, F. F., Mello, L. H. C., and Gilardi, R. P. 2000. Drill holes in shells of Permian benthic invertebrates. Journal of Paleontology, 74:532543.Google Scholar
Lehmann, U. 1976. The Ammonites: Their Life and Their World. Cambridge University Press, Cambridge, 245 p.Google Scholar
Lehmann, U. 1988. On the dietary habits and locomotion of fossil cephalopods, p. 633640. In Wiedmann, J. and Kullmann, J. (eds.), Cephalopods Present and Past. 2nd International Cephalopod Symposium, E. Schweizerbart'sche Verlagsbuchhandlung, Stuttgart, Germany.Google Scholar
Leighton, L. R. 2001a. New examples of Devonian predatory boreholes and the influence of brachiopod spines on predator success. Palaeogeography, Palaeoclimatology, Palaeoecology, 165:7191.CrossRefGoogle Scholar
Leighton, L. R. 2001b. New directions in the paleoecology of Paleozoic brachiopods, p. 185206. In Carlson, S. J. and Sandy, M. (eds.), Brachiopods Ancient and Modern: A Tribute to G. Arthur Cooper, Paleontological Society Papers 7.Google Scholar
Lescinsky, H. L., and Benninger, L. 1994. Pseudoborings and predatory traces: Artifacts of pressure dissolution in fossiliferous shales. Palaios, 9:599604.Google Scholar
Liljedal, L. 1985. Ecological aspects of a silicified bivalve fauna from the Silurian of Gotland. Lethaia, 18:5356.Google Scholar
Malzahn, E. 1968. Uber neue Funde von Janassa bituminosa (Schloth.) im neiderrheinisschen Zechstein. Geologisch Jarhbuch, 85:6796.Google Scholar
Mapes, R. H., and Benstock, E. J. 1988. Color pattern on the Carboniferous bivalve Streblochondria? Newell. Journal of Paleontology, 62:439441.Google Scholar
Mapes, R. H., and Hansen, M. C. 1984. Pennsylvanian shark-cephalopod predation: A case study. Lethaia, 17:175183.Google Scholar
McMenamin, M. 1986. The Garden of Ediacara. Palaios, 1:178182.Google Scholar
McMenamin, M., and McMenamin, D. L. S. 1990. The Emergence of Animals: The Cambrian Breakthrough. Columbia University Press, New York, 217 p.CrossRefGoogle Scholar
Mehl, J. 1984. Radula und Fangarme bei Michelinoceras sp. asu dem Silur von Bolivien. Paläeontologica Zeitschrift, 58:211229.Google Scholar
Meyer, D. L. 1985. Evolutionary implications of predation on Recent comatulid crinoids. Paleobiology, 11:154164.CrossRefGoogle Scholar
Meyer, D. L., and Ausich, W. I. 1983. Biotic interactions among Recent crinoids and among fossil crinoids, p. 377427. In Tevesz, M. J. S. and McCall, P. L. (eds.), Biotic Interactions in Recent and Fossil Benthic Communities. Plenum Press, New York.Google Scholar
Miller, R. H., and Sundberg, F. A. 1984. Boring Late Cambrian organisms. Lethaia, 17:185190.CrossRefGoogle Scholar
Morris, P. G. 1994. Some new Carboniferous cardiniarid brachiopods from the Middle Limestones (Visean, Chadian) of North Staffordshire, England. Neues Jahrbuch für Geologie und Palaontologie Monatshefte, 5:267276.Google Scholar
Moy-Thomas, J. A., and Miles, R. S. 1971. Palaeozoic Fishes. W. B. Saunders Co., Philadelphia, 259 p.Google Scholar
Nedin, C. 1999. Anomalocaris predation on non-mineralized and mineralized trilobites. Geology, 27:987990.Google Scholar
Nixon, M. 1988. The feeding mechanisms and diets of cephalopods—Living and fossil, p. 642652. In Wiedmann, J. and Kullmann, J. (eds.), Cephalopods Present and Past. 2nd International Cephalopod Symposium, E. Schweizerbart'sche Verlagsbuchhandlung, Stuttgart, Germany.Google Scholar
Paine, R. T. 1974. Intertidal community structure: Experimental studies on the relationship between dominat competitor and its principal predator. Oecologia, 15:93120.Google Scholar
Patterson, C. 1965. The phylogeny of the chimaeroids. Philosophical Transactions of the Royal Society, Series B, 249:101209.Google Scholar
Peel, J. S. 1984. Attempted predation and shell repair in Euomphalopterus (Gastropoda) from the Silurian of Gotland. Lethaia, 32:163168.Google Scholar
Richards, R. P., and Shabica, C. W. 1969. Cylindrical living burrows in Ordovician dalmanellid brachiopod shells. Journal of Paleontology, 43:838841.Google Scholar
Rodriguez, R. P., and Gutschick, R. C. 1970. Late Devonian-Early Mississippian ichnofossils from western Montana and northern Utah, p. 407438. In Crimes, J. P. and Harper, L. C. (eds.), Trace Fossils. Geological Journal, Special Issue 3. Seel House, Liverpool.Google Scholar
Rohr, D. M. 1976. Silurian predator borings in the brachiopod Dicaelosia from the Canadian Arctic. Journal of Paleontology, 65:687688.Google Scholar
Rollins, H. B., and Brezinski, D. K. 1988. Reinterpretation of crinoid-platyceratid interaction. Lethaia, 21:189292.Google Scholar
Rudkin, D. M. 1985. Exoskeleton abnormalities in four trilobites. Canadian Journal of Earth Sciences, 22:479483.Google Scholar
Rudwick, M. J. S. 1970. Living and Fossil Brachiopods. Hutchinson University Library, London, 199 p.Google Scholar
Saunders, W. B., and Ward, P. D. 1987. Ecology, distribution, and population characteristics of Nautilus , p. 137162. In Saunders, W. B. and Landman, N. H. (eds.), Nautilus: The Biology and Paleobiology of a Living Fossil. Plenum Press, New York.Google Scholar
Schaefer, B., and Williams, M. 1977. Relationships of fossil and living elasmobranchs. American Zoologist, 17:293302.Google Scholar
Schindel, D. E., Vermeij, G. J., and Zipser, E. 1982. Frequencies of repaired shell fractures among the Pennsylvanian gastropods of north-central Texas. Journal of Paleontology, 56:729740.Google Scholar
Schram, F. R. 1977. Palaeozoogeography of Late Paleozoic and Triassic Malacostraca. Systematic Zoology, 26:367379.CrossRefGoogle Scholar
Schram, F. R. 1982. The fossil record and evolution of the Crustacea, p. 93147. In Abele, L. G. (ed.), The Biology of the Crustacea: Systematics, the Fossil Record, and Biogeography. Academic Press, Inc., New York.Google Scholar
Schram, F. R. 1984. Upper Pennsylvanian arthropods from black shales of Iowa and Nebraska. Journal of Paleontology, 58:197209.Google Scholar
Selden, P. A. 1984. Autecology of Silurian eurypterids. In Bassett, M. G. and Lawson, J. D. (eds.), Autecology of Silurian Organisms. Special Papers in Palaeontology, 32:3954 Google Scholar
Selden, P. A. 1992. Biomechanics, p. 322326. In Briggs, D. E. G. and Crowther, P. R. (eds.), Paleobiology: A Synthesis. Blackwell Scientific Publications, Oxford.Google Scholar
Sheehan, P. M., and Lesperance, P. J. 1978. Effect of predation on 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 Foraminferal Research, 1:2029.CrossRefGoogle Scholar
Smith, S. A., Thayer, C. W., and Brett, C. E. 1985. Predation in the Paleozoic: gastropod-like drill holes in Devonian brachiopods. Science, 230:10331037.Google Scholar
Spencer, W. K., and Wright, C. W. 1966. Asterozoans, p. U1U157. In Moore, R. C. (ed.), Treatise on Invertebrate Paleontology, Pt. U. Geological Society of America and University of Kansas Press.Google Scholar
Sprinkle, J. 1973. Morphology and Evolution of blastozoan echinoderms. Museum of Comparative Zoology, Harvard University Publication, 284.Google Scholar
Stanley, S. M. 1970. Relation of shell form to life habits of the Bivalvia (Mollusca). Memoir of the Geological Society of America, 125:1296.Google Scholar
Stanley, S. M. 1977. Rates, trends, and patterns of evolution in the Bivalvia, p. 209250. In Hallam, A. (ed.), Patterns of Evolution as Illustrated by the Fossil Record. Elsevier, Amsterdam.Google Scholar
Stensiö, E. 1969. Les cyclostomes fossiles ou ostracodermes, p. 71692. In Piveteau, J. (ed.), Traité des Paléontologie 4. Masson, Paris.Google Scholar
Tasch, P. 1980. Paleobiology of the Invertebrates: Data Retrieval from the Fossil Record. John Wiley, New York, 975 p.Google Scholar
Thayer, C. W. 1983. Sediment-mediated biological disturbance and the evolution of the marine benthos, p. 479595. In Tevesz, M. J. S. and McCall, P. J. (eds.), Biotic Interactions in Recent and Fossil Benthic Communities. Plenum Press, New York.Google Scholar
Vermeij, G. J. 1977. The Mesozoic marine revolution: Evidence from snails, predators, and grazers. Paleobiology, 3:245258.CrossRefGoogle Scholar
Vermeij, G. J. 1983. Shell breaking predation through time, p. 649669. In Tevesz, M. J. S. and McCall, P. L. (eds.), Biotic Interactions in Recent and Fossil Benthic Communities. Plenum Press, New York.Google Scholar
Vermeij, G. J. 1987. Evolution and Escalation. Princeton University Press, Princeton, NJ, 527 p.CrossRefGoogle Scholar
Vermeij, G. J. 1995. Economics, volcanoes, and Phanerozoic revolutions. Paleobiology, 21:125152.CrossRefGoogle Scholar
Vermeij, G. J., Schindel, D. E., and Zipser, E. 1981. Predation through geological time: Evidence from gastropod shell repair. Science, 214(27):10241026.Google Scholar
Waters, J. W., and Maples, C. 1991. Mississippian pelmatozoan community reorganization: A predation-mediated faunal change. Paleobiology, 17:400410.Google Scholar
Williams, M. E. 1990. Feeding behavior in Cleveland Shale fishes, p. 273287. In Boucot, A. J., Evolutionary Paleobiology of Behavior and Coevolution. Elsevier, Amsterdam.Google Scholar
Zangerl, R. 1981. Chondrichthyes, 1: Paleozoic elasmobranches. Handbook of Paleoichthyology, 3A:1115.Google Scholar
Zangerl, R., and Richardson, E. 1963. Paleoecological history of two Pennsylvanian black shales. Fieldiana Geological Memoir, 4, 352 p.Google Scholar
Zangerl, R., Woodland, B. G., Richardson, E. S., and Zachry, D. L. 1969. Early diagenetic phenomena in the Fayetteville Black Shale (Mississippian) in Arkansas. Sedimentary Geology, 3:87120.Google Scholar