Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-06-09T12:02:16.201Z Has data issue: false hasContentIssue false

Invertebrate Trace Fossils: The Backbone of Continental Ichnology

Published online by Cambridge University Press:  17 July 2017

Stephen T. Hasiotis
Affiliation:
Department of Geological Sciences, Campus Box 250, Boulder, CO 80309-0250 and U.S. Geological Survey, Box 25046, M.S. 919, DFC, Denver, CO 80225-0046
Thomas M. Bown
Affiliation:
U.S. Geological Survey, Box 25046, M.S. 919, DFC, Denver, CO 80225-0046

Extract

The purpose of this chapter is to evoke new concepts, provide guidelines and new frontiers for future research, and demonstrate that invertebrate traces actually comprise the “backbone” of continental (as well as marine) ichnology. Invertebrate organisms that inhabit the continental, nonmarine realm include some of the most diverse and populous classes in the animal kingdom. For example, both the Insecta and Crustacea exhibit burrowing behaviors unique to subaqueous freshwater and subaerial systems. Because of the sheer biomass of burrowing pupae, larvae, juvenile, and adult stages of these organisms, invertebrates dependent on the position of the water table form the basis for ecological niche-partitioning of depositional systems within all of the continental realm. A distinction must be made here between continental and marine ichnocoenoses because they represent distinctly different styles of living. These in turn dictate different behavioral and genetic responses of the organisms that inhabit them. Burrow architectures that occur in both continental and marine ichnocoenoses can be differentiated by subtle differences in morphology that are due to convergence of the burrowing mechanisms of the respective organisms.

Type
Research Article
Copyright
Copyright © 1992 Paleontological Society 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abel, O. 1935. Vorzeitliche Lebensspuren. Jena, Gustav Fisher, 644 p.Google Scholar
Ahlbrandt, T.S., Andrews, S., and Gwynne, D.T. 1978. Bioturbation of eolian deposits. Journal of Sedimentary Petrology, 48:839848.Google Scholar
Ahlbrandt, T.S., and Fryberger, S.G. 1982. Introduction to Eolian Deposits. In Scholle, P.A. and Spearing, D., eds., AAPG Memior 31.Google Scholar
Allen, J.R.L. 1974. Studies in fluviatile sedimentation: Implications of pedogenic carbonate units, Lower Old Red Sandstone, Anglo-Welsh outcrop. Geological Journal, 9:181208.Google Scholar
Anderson, J.M. 1977. The organization of soil animal communities. In Lohm, U. and Persson, A., eds., Soil Organisms as Components of Ecosystems. Proceedings, 6th Int. Coll. Soil Zool. Ecol. Bull., 25:1523.Google Scholar
Archer, A.W., and Maples, C.G. 1984. Trace-fossil distribution across a marine-to-nonmarine gradient in the Pennsylvanian of southwestern Indiana. Journal of Paleontology, 54:448466.Google Scholar
Ardo, P. 1957. Studies in the marine shore dune ecosystem with special reference to the dipterous fauna. Opuscula Entomology, Supplement, 14:1255.Google Scholar
Basan, P.B. 1978. Trace Fossil Concepts. SEPM Short Course No. 5, Oklahoma City, 181 p.Google Scholar
Basan, P.B., and Frey, R.W. 1977. Actual-paleontology and neoichnology of salt marshes near Sapelo Island, Georgia, p. 4170. In Crimes, T.P., and Harper, J.C., eds., Trace Fossils II. Seel House Press, Liverpool, UK.Google Scholar
Batra, S.W.T. 1966. Nesting behavior of Halictus scabiosae in Switzerland (Hymenoptera, Halictidae). Insectes Socianx, 13:8792.Google Scholar
Batra, S.W.T. 1968. Nests and social behavior of Halictine bees of India (Hymenoptera: Halictidae). Indian Journal of Entomology, 28:375393.Google Scholar
Behnke, F.L. 1977. A natural history of termites. Charles Scribner's Sons, New York, 118 p.Google Scholar
Bown, T.M. 1982. Ichnofossils and rhizoliths of the nearshore fluvial Jebel Qatroni Formation (Oligocene), Fayum Province, Egypt. Palaeogeography, Palaeoclimatology, Palaeoecology, 40:255309.Google Scholar
Bown, T.M., and Kraus, M.J. 1983. Ichnofossils of the Alluvial Willwood Formation (Lower Eocene), Bighorn Basin, Northwest Wyoming, U.S.A. Palaeogeography, Palaeoclimatology, Palaeoecology, 43:95128.Google Scholar
Bown, T.M., and Kraus, M.J. 1987. Integration of channel and floodplain suites, I. Developmental sequence and lateral relations of alluvial paleosols. Journal of Sedimentary Petrology, 57:587601.Google Scholar
Bown, T.M., and Kraus, M.J. 1988. Geology and paleoenvironment of the Oligocene Jebel Qatrani Formation and adjacent rocks, Fayum Depression, Egypt. U.S. Geological Survey Profeseeional Paper 1452, 60 p.CrossRefGoogle Scholar
Bown, T.M., and Larriestra, C.N. 1990. Sedimentary paleoenvironments of fossil playrrhine localities, Miocene Pinturas Formation, Santa Cruz Province, Argentina, p. 87119. In Fleagle, J.G. and Rosenberger, A.L., eds., The Platyrrhine Fossil Record. Academic Press, New York.Google Scholar
Bown, T.M., and Laza, J.H. 1990. A Miocene termite nest from southern Argentina and its paleoclimatological implications. Ichnos, 1:7382.Google Scholar
Bown, T.M., and Ratcliffe, B.C. 1988. The origin of Chubutolithes Ihering, ichnofossils from the Eocene and Oligocene of Chubut Province, Argentina. Journal of Paleontology, 62:163167.Google Scholar
Bromley, R.G. 1990. Trace Fossils: Biology and taphonomy. Special Topics in Paleontology, No. 3. Unwin Hyman, Ltd., 280 p.Google Scholar
Bromley, R.G., and Asgaard, U. 1979. Triassic freshwater ichnocoenoses from Carlsberg Fjord, East Greenland. Palaeogeography, Palaeoclimatology, Palaeoecology 28:3980.Google Scholar
Bromley, R.G., Pemberton, S.G., and Rahmani, R. 1984. A Cretaceous woodground: the Teredolites ichnofacies. Journal of Paleontology, 58:488498.Google Scholar
Brown, R.W. 1934. Celliforma spirifer, the fossil larval chambers of mining bees. Washington Academy of Science Journal, 24:532539.Google Scholar
Chamberlain, C.K. 1975. Recent Lebensspuren in nonmarine aquatic environments, p. 431458. In Frey, R.W., ed., The Study of Trace Fossils. Springer Verlag, New York.Google Scholar
Cloudsley-Thompson, J.L. 1987. The fauna of sand dunes at Alvor (Portugal) and Lanzarote (Canary Isles). Entomolo. Rec., 99:3536.Google Scholar
Cohen, A.S. 1984. Effects of zoobenthic standing crop on laminae preservation in tropical lake sediment, lake Turkana, East Africa. Journal of Paleontology, 59:499510.Google Scholar
Collinson, J.D. 1986a. Alluvial Sediments, p. 2062. In Reading, H.G., ed., Sedimentary Environments and Facies, 2nd. ed. Elsevier, New York.Google Scholar
Collinson, J.D. 1986b. Deserts, p. 95112. In Reading, H.G., ed., Sedimentary Environments and Facies, 2nd. ed. Elsevier, New York.Google Scholar
Crawford, C.S. 1991. Animal adaptions and ecological processes in desert dunefields. Journal of Arid Environments, 21:245260.Google Scholar
Crawford, C.S., and Seely, M.K. 1987. Assemblages of surface-active arthropods in the Namib dunefield and associated habitats. Revue de Zoologique Africa, 101:397421.Google Scholar
Crimes, T.P., and Harper, J.C., eds. 1970. Trace Fossils. Geological Journal Special Issue No. 3, 547 p.Google Scholar
Crimes, T.P., and Harper, J.C., eds. 1977. Trace Fossils 2. Geological Journal Special Issue No. 9, 351 p.Google Scholar
Curran, H.A., ed. 1985. Biogenic structures: Their use in interpreting depositional environments. SEPM Special Publication No. 35, 347 p.Google Scholar
Driscoll, F.G. 1986. Ground Water and Wells, 2nd edition. Johnson Division Publishers, St. Paul, Minnesota, 1089 p.Google Scholar
Dubiel, R.F., Blodgett, R.H., and Bown, T.M. 1987. Lungfish burrows in the Upper Triassic Chinle and Dolores formations, Colorado Plateau. Journal of Sedimentary Petrology, 57:512521.Google Scholar
Dubiel, R.F., Blodgett, R.H., and Bown, T.M. 1988. Lungfish burrows in the Upper Triassic Chinle and Dolores Formations, Colorado Plateau—Reply. Journal of Sedimentary Petrology, 58:367369.Google Scholar
Dubiel, R.F., Blodgett, R.H., and Bown, T.M. 1989. Lungfish burrows in the Upper Triassic Chinle and Dolores Formations, Colorado Plateau—Reply. Journal of Sedimentary Petrology, 59:876878.Google Scholar
Dubiel, R.F., Parrish, J.T., Parrish, J.M., and Good, S.C. 1991. The Pangean Megamonsoon—evidence from the Upper Triassic Chinle Formation, Colorado Plateau. Palaios, 6:347370.Google Scholar
Dubiel, R.F., Skipp, G., and Hasiotis, S.T. 1992. Continental depositional environments and tropical paleosols in the Upper Triassic Chinle Formation, Eagle Basin, western Colorado. In Flores, R.M., ed., Mesozoic of the Western Interior. SEPM Theme Meeting, Ft. Collins, CO., 46 p.Google Scholar
Ekdale, A. A., Bromley, R.G., and Pemberton, S.G. 1984. Ichnology: The Use of Trace Fossils in Sedimentology and Stratigraphy. SEPM Publication, Tulsa, Oklahoma, 317 p.Google Scholar
Ekdale, A. A., and Picard, M.D. 1985. Trace fossils in a Jurassic eolianite, Entrada Sandstone, Utah, U.S.A., p. 312. In Curran, H.A., ed., Biogenic structures: Their use in interpreting depositional environments. SEPM Special Publication.Google Scholar
Ekdale, A. A., and Pollard, J., eds. 1991. Ichnofabrics and Ichnofacies. Palaios, Special Issue, 6:343 p.Google Scholar
Elliott, T. 1986. Siliciclastic Shorelines, p. 155188. In Reading, H.G., ed., Sedimentary environments and facies, 2nd. ed. Elsevier, New York.Google Scholar
Evans, H.E. 1958. Studies on the nesting behavior of digger wasps of the tribe Spencini. Part I: Genus Priononyx Dahlbom. Annals of the Entomological Society of America, 51:177186.Google Scholar
Farrow, G.E. 1971. Back-reef and lagoonal environments of Aldabra Atoll distinguished by their crustacean burrows. Symp. Zool. Soc. London, 28:455500.Google Scholar
Fisher, J.B. 1982. Effects of marcobenthos on the chemical diagenesis of freshwater sediments, p. 170220. In McCall, P.L. and Tevesz, M.J.S., eds., Animal-sediment relations: The biogenic alteration of sediments. Plenum Press, New York.Google Scholar
Fisher, J.B., Lick, W., McCall, P.L., and Robbins, J. A. 1980. Vertical mixing of lake sediments by tubificid oligochaetes. Journal of Geophysical Research, 85:39974006.Google Scholar
Fouch, T.D., and Dean, W.E. 1982. Lacustrine and associated clastic depositional environments, p. 87114. In Scholle, P.A. and Spearing, D., eds., AAPG Memior 31.Google Scholar
Frey, R.W., ed. 1975. The Study of Trace Fossils. Springer Verlag, New York Inc., 562 p.Google Scholar
Frey, R.W., ed. 1978. Behavioral and ecological implications of trace fossils, p. 4366. In Basan, P.B., ed., Trace Fossil Concepts. SEPM Short Course No. 5, Oklahoma City.Google Scholar
Frey, R.W., and Mayou, T.V. 1971. Decapod burrows in Holocene barrier island beaches and washover fans, Georgia. Senckenbergiana Maritima, 3:5377.Google Scholar
Frey, R.W., and Pemberton, S.G. 1987. Psilonichnus ichnocoenose, and its relationship to adjacent marine and nonmarine ichnocoenoses along the Georgia coast. Bulletin of Canadian Petroleum Geology, 33:72115.Google Scholar
Frey, R.W., and Fagerstrom, J.A. 1984. Morphological, ethological, and environmental significance of the ichnogenera Scoyenia and Ancorichnus . Journal of Paleontology, 58:511528.Google Scholar
Fürsich, F.T. 1981. Invertebrate trace fossils form the Upper Jurassic of Portugal. Comm. Geol. Portugal, 67:153168.Google Scholar
Fürsich, F.T., and Mayr, H. 1981. Non-marine Rhizocorallium (trace fossil) from the Upper Freshwater Molasse (Upper Miocene) of southern Germany. Neus Jahrbuch für Geologie und Palaontologiel, Monatsh., 1981:321333.Google Scholar
Gebo, D.L., and Rasmussen, D.T. 1985. The earliest fossil pangolin (Pholidota: Manidae) from Africa. Journal of Mammalogy, 66:538541.Google Scholar
Genise, J.F., and Bown, T.M. 1990. The constructor of the ichnofossil Chubutolithes . Journal of Paleontology, 64:482483.Google Scholar
Genise, J.F., and Bown, T.M. 1991. A reassessment of the ichnofossil Chubutolithes gaimanensis Bown and Ratcliffe: Reply. Journal of Paleontology, 65:705706.Google Scholar
Genise, J.F., and Bown, T.M. In press. Paleoichnology and paleoenvironmental reconstructions of Santacruzian (Miocene) fossil vertebrate localities, Provincia Santa Cruz, Argentina. Ichnos. Google Scholar
Gibbard, P.L., and Stuart, A.J. 1974. Trace fossils from proglacial lake sediments. Boreas, 3:6974.CrossRefGoogle Scholar
Gibbard, P.L., and Dreimanis, A. 1978. Trace Fossils from late Pleistocene glacial lake sediments in southwestern Ontario, Canada. Canadian Journal of Earth Sciences, 15:19671976.Google Scholar
Gillette, D.D., and Lockley, M.G., eds. 1989. Dinosaur Tracks and Traces. Cambridge University Press, New York. 454 p.Google Scholar
Gillmore, C.W. 1926. Fossil footprints from the Grand Canyon. Smithsonian Miscellaneous Collections, 77:41 p.Google Scholar
Glennie, K.W., and Evamy, B.D. 1968. Dikaka: Plants and plant-root structures associated with eolian sand. Palaeogeography, Palaeoclimatology, Palaeoecology, 4:7787.Google Scholar
Goldring, R., Bosence, D.W.J., and Blake, T. 1978. Estuarine sedimentation in the Eocene of southern England. Sedimentology, 25:861876.Google Scholar
Gunn, D.L., and Kennedy, J.S. 1936. Apparatus for investigating the reactions of land arthropods to humidity. Journal of Experimental Biology 13:450459.Google Scholar
Hanley, J.H., Steidtman, J.R., and Toots, H. 1971. Trace fossils from the Casper Sandstone (Permian) southern Laramie Basin, Wyoming and Colorado. Journal of Sedimentary Petrology, 41:10651068.Google Scholar
Häntzschel, W. 1965. Vestiga invertebratorum et Problematica. Fossilium Catalogus, 1: Animalia, pars 108, 142 p.Google Scholar
Hasiotis, S.T. 1990a. Identification of the architectural and surficial burrow morphologies of ancient lungfish and crayfish burrows: Their importance to ichnology. The Australasian Institute of Mining and Metallurgy Pacific Rim Congress 90, 3:529536.Google Scholar
Hasiotis, S.T. 1990b. Upper Triassic Chinle Formation of southeastern Utah, U.S.A.: Crayfish burrows as floodplain and water table indicators. The Canadian Paleontology and Biostratigraphy Seminar, Queen's University, Ontario, Canada, 1990:2627.Google Scholar
Hasiotis, S.T. 1991. Paleontology, ichnology, and paleoecology of the Upper Triassic Chinle Formation of the Canyonlands, southeastern Utah. Unpublished Master's thesis, State University of New York at Buffalo, Amherst, New York, 350 p.Google Scholar
Hasiotis, S.T. 1992a, Notes on the burrow morphologies and nesting behaviors of adults and juveniles of Procambarus clarkii and Procambarus acutus acutus . Proceedings to the 8th International Freshwater Crayfish Symposium, Romaire, R., ed., Baton Rouge, Louisiana, 14 p.Google Scholar
Hasiotis, S.T. 1992b. New fossil occurrences and paleoecological implications from the Upper Triassic Chinle Formation of the Canyonlands vicinity, southeastern, Utah: Palaios, 41 p.Google Scholar
Hasiotis, S.T. In press. New observations on the burrow morphologies and burrowing behavior of the crayfish Cambarus diogenes diogenes Girard. American Midland Naturalist.Google Scholar
Hasiotis, S.T., Aslan, A., and Bown, T.M. In press a. Construction, architecture, and paleoecology of the Early Eocene nonmarine ichnofossil Scaphichnium hamatum . Ichnos. Google Scholar
Hasiotis, S.T., and Hannigan, R. 1990. Identification of possible lungfish burrows in the Upper Triassic Chinle Formation of southeastern Utah: Periodic Acid Schiff (PAS) Reaction and mucus burrow linings. Rochester Academy of Science Symposium, p. 18.Google Scholar
Hasiotis, S.T., and Hannigan, R. In press. A modified Periodic Acid Schiff (PAS) staining technique for the identification of mucus linings in ichnofabric. Ichnos. Google Scholar
Hasiotis, S.T., and Mitchell, C.E. 1989. Lungfish burrows of the Upper Triassic Chinle and Dolores formations, Colorado Plateau—Discussion: New Evidence suggests origin by a decapod crustacean. Journal of Sedimentary Petrology, 59:871875.Google Scholar
Hasiotis, S.T., and Mitchell, C.E. 1990. Preliminary report on the comparison of Triassic and Holocene crayfish burrow morphologies. Northeastern Section Geological Society of America Abstracts with Programs, 22:22.Google Scholar
Hasiotis, S.T., and Mitchell, C.E. In press b. A comparison of crayfish burrow morphologies: Triassic and Holocene fossil, paleo- and neo-ichnological evidence, and the identification of their burrowing signatures. Ichnos. Google Scholar
Hasiotis, S.T., and Mitchell, C.E. In press b. Application of burrow morphologic evaluations: Permian and Holocene lungfish burrow morphologies and their comparison to Upper Triassic Chinle Formation burrow localities in southeastern Utah and western Colorado: Ichnos.Google Scholar
Hill, G.W., and Hunter, R.E. 1976. Interaction of biological and geological processes in the beach and nearshore environments, northern Padre Island, Texas. SEPM Special Publication, 24:169187.Google Scholar
Hitchcock, E. 1858. Ichnology of New England. A report of the sandstone of the Connecticut Valley especially its footprints. Boston, W. White, 220 p.Google Scholar
Hole, F.D. 1981. Effects of animals on soil. Geoderma, 25:75112.Google Scholar
Hobbs, H.H. Jr. 1981. The crayfishes of Georgia. Smithsonian Contributions to Zoology No. 166, 166 p.Google Scholar
Hobbs, H.H. Jr., and Whiteman, M. 1991. Notes on the burrows, behavior, and color of the crayfish Fallicambarus (F.) devastator (Decapoda: Cambaridae). Southwestern Naturalist, 36:127135.Google Scholar
Horowitz, P.H.J., and Richardson, A.M.M. 1986. An ecological classification of the burrows of Australian freshwater crayfish. Australian Journal of Marine and Freshwater Research, 37:237242.Google Scholar
Howard, J.D., and Frey, R.W. 1975. Regional animal-sediment characteristics of Georgia estuaries. Senckenbergiana Maritima, 7:33103.Google Scholar
Howden, H.F. 1955. Biology and taxonomy of North American beetles of the subfamily Geotrupinae with revisions of the genera Bolbocerosoma, Eucanthus, Geotrupes, and Peltotrupes (Scarabaeidae). Proceedings, U.S. National Museum, 104:151319.Google Scholar
Imbrie, J., and Newell, N.D., eds. 1964. Approaches to Paleoecology. John Wiley and Sons, Inc., 432 p.Google Scholar
Jenny, H. 1941. Factors of Soil Formation. McGraw-Hill Publishers, New York, 281 p.Google Scholar
Klappa, C.F. 1980. Rhizoliths in terrestrial carbonates: Classification, recognition, genesis and significance. Sedimentology, 27:613629.Google Scholar
Kocurek, G. 1981. Significance of interdune deposits and bounding surfaces in aeolian dune sands. Sedimentology, 28:753780.Google Scholar
Kraus, M.J. 1987. Integration of channel and floodplain suites, II. Vertical relations of alluvial paleosols. Journal of Sedimentary Petrology, 57:602612.Google Scholar
Lockley, M.G. 1991. Tracking Dinosaurs: A New Look at an Ancient World. Cambridge University Press, New York, 238 p.Google Scholar
Lockley, M.G., Rindsberg, A.K., and Zeiler, R.M. 1987. The paleoenvironmental significance of the nearshore Curvolithos ichnofacies. Palaios, 2:255262.Google Scholar
McKee, E.D. 1934. An investigation of the light-colored, cross-bedded sandstones of Canyon De Chelly, Arizona. American Journal of Science, 28:219233.Google Scholar
McKee, E.D. 1944. Tracks that go uphill. Plateau, 16:6172.Google Scholar
McKee, E.D. 1979. A study of global sand seas. U.S. Geological Survey Professional Paper 1052, 429 p.Google Scholar
McLachlan, A. 1991. Ecology of coastal dune fauna. Journal of Arid Environments, 21:229243.Google Scholar
Martin, L.D., and Bennet, D.K. 1977. The burrows of the Miocene Beaver Paleocaster, western Nebraska, U.S.A. Palaeogeography, Palaeoclimatology, Palaeoecology, 22:173193.Google Scholar
Merrill, R.D. 1984. Ophiomorpha and other nonmarine trace fossils from the Eocene Ione Formation, California. Journal of Paleontology, 58:542549.Google Scholar
Mieras, B., Sageman, B.B., and Kauffman, E.G. in press. Trace fossil distribution patterns in Cretaceous facies of the Western Interior Basin, North America. Geological Association of Canada Special Paper.Google Scholar
Miller, M.F. 1984. Distribution of biogenic structures in Paleozoic nonmarine and marine-margin sequences: An actualistic model. Journal of Paleontology, 54:550570.Google Scholar
Miller, M.F., Ekdale, A.A., and Picard, M.D., eds. 1984. Trace fossils and paleoenvironments: Marine carbonate, marginal marine terrigenous, and continental terrigenous settings. Journal of Paleontology Special Issue, 54: 598 p.Google Scholar
Miller, M.F., Ekdale, A.A., and Johnson, K.G. 1981. Spirophyton in alluvial-tidal facies of the Catskill delta complex: possible biological control of ichnological distribution. Journal of Paleontology, 55:10161027.Google Scholar
Miller, W., and Whitcomb, N.J., and Brown, N.A. 1981. Paleoecological aspects of a freshwater ephemeral stream, New Hope Valley, Orange County, North Carolina. Southeastern Geology, 22:149158.Google Scholar
Moussa, M.T. 1970. Nematode trails from the Green River Formation(Eocene) in the Uinta Basin, Utah. Journal of Paleontology, 44:304307.Google Scholar
Payne, C.E., ed. 1977. Seismic Stratigraphy: Applications to hydrocarbon exploration. AAPG Memoir 26, 516 p.Google Scholar
Pemberton, S.G., and Wightman, D.M. 1987. Brackish water trace fossil suites: Examples from the Lower Cretaceous Mannville Group, p. 183190. In Currie, P.M. and Koster, E.H., eds., Fourth Symposium on Mesozoic Terrestrial Ecosystems, Short Papers.Google Scholar
Picard, M.D. 1977. Stratigraphic analysis of the Navajo Sandstone: A discussion. Journal of Sedimentary Petrology, 47:475483.Google Scholar
Pienkowski, G. 1985. Early Liassic trace fossil assemblages from the Holy Cross Mountains, Poland: Their distribution in continental and marginal marine environments, p. 3751. In Curren, H.A., ed., Biogenic structures: Their use in interpreting depositional environments. SEPM Special Publication No. 35.Google Scholar
Plint, A.G., and Pickerill, R.K. 1985. Non-marine Teredolites from the Middle Eocene of southern England. Lethaia, 18:339345.Google Scholar
Posamentier, H.W., Jervey, M.T., and Vail, P.R. 1988. Eustatic controls on clastic deposition—Conceptual framework, p. 109124. In Wilgus, C.K., Posamentier, H., Ross, C.A., and Kendall, C. G. St. G., eds., Sea Level Changes: An Integrated Approach. SEPM Special Publication No. 42.Google Scholar
Posamentier, H.W., Jervey, M.T., and Vail, P.R. 1988. Eustatic controls on clastic deposition—Sequence and systems tract models, p. 124154. In Wilgus, C.K., Posamentier, H., Ross, C.A., and Kendall, C. G. St. G., eds., Sea Level Changes: An Integrated Approach. SEPM Special Publication No. 42.Google Scholar
Pryor, W.A. 1967. Biogenic directional features on several recent pointbars. Sedimentary Geology, 1:235245.Google Scholar
Rafes, P.M. 1960. The life forms of insects inhabiting the Naryn sands of the semidesert Transvolga region. Entomological Review, 38:1931.Google Scholar
Ransom, M.D., Smeck, N.E., and Bigham, J.M. 1987. Stratigraphy and genesis of polygenetic soils on the Illinoisan till plain of southwestern Ohio. Soil Science Society of America Journal, 51:135141.Google Scholar
Ratcliffe, B.C., and Fagerstrom, J.A. 1980. Invertebrate lebensspuren of Holocene floodplains: Their morphology, origin, and paleoecological significance. Journal of Paleontology, 54:614630.Google Scholar
Retallack, G J. 1977. Triassic paleosols in the upper Narrabeen Group of New South Wales, Part I, Features of the paleosols. Geological Society of Australia Journal, 23:383399.Google Scholar
Retallack, G J. 1984. Trace fossils of burrowing beetles and bees in an Oligocene paleosol, Badlands National Park, South Dakota. Journal of Paleontology, 58:571592.Google Scholar
Richter, R. 1927. Die fossilien Fahrten und Bauten der Wurmer, ein Uberblick uber ihre biologischen Grundformen und deren geologishce Bedeutung. Palaontologisch Zeitschrift, 9:193240.Google Scholar
Risk, Mj., and Szczuczko, R.B. 1977. A method for staining trace fossils. Journal of Sedimentary Petrology, 47:855859.Google Scholar
Ryer, T.A. 1981. Deltaic coals of Ferron Sandstone member of Mancos Shale: Predicative model for Cretaceous coal-bearing strata of Western Interior. American Association of Petroleum Geologists, Bulletin, 65:23232340.Google Scholar
Sands, W.A. 1987. Fossil invertebrates: Ichnocoenoses of probable termite origin from Laetoli, p. 409433. In Leakey, M.D. and Harris, J.M., eds., Laetoli: A Pliocene Site in Northern Tanzania. Clarendon Press, Oxford.Google Scholar
Sarjeant, W.A.S., ed. 1983. Terrestrial Trace Fossils. Hutchinson Ross Publishing Company, Benchmark Papers in Geology, 76:415 p.Google Scholar
Schaller, F. 1968. Soil Animals. University Michigan Press, Ann Arbor, 144 p.Google Scholar
Schuchert, C. 1918. On the Carboniferous of the Grand Canyon. American Journal of Science, 45:350.Google Scholar
Seilacher, A. 1953. Uber die Methoden der Palichnologie. 1. Studien zur Palichnologie. Neues Jahrbuch für Geologie und Palaontologie, Abhandlungen, 96:421452.Google Scholar
Seilacher, A. 1967. Bathymetry of trace fossils. Marine Geology, 5:413428.Google Scholar
Seilacher, A. 1978. Use of trace fossils for recognizing depositional environments, p. 167181. In Basan, P.B., ed., Trace Fossil Concepts, SEPM Short Course No. 5, Oklahoma City.Google Scholar
Seilacher, A., and Goldring, R. 1971. Limulid undertracks and their sedimentological implications. Neues Jahrbuch für Geologie und Palaontologie, Abhandlungen, 37:422442.Google Scholar
Shorey, H.H., and Gyrisco, G.G. 1960. Effects of soil temperature and moisture on the vertical distribution of European chafer larvae. Annals of the Entomological Society of America, 53:666670.Google Scholar
Silvey, J.K.G. 1936. An investigation of the burrowing inner-beach insects of some fresh-water lakes. Michigan Academy of Sciences and Arts, Letters and Papers. 21:655696.Google Scholar
Skiba, U., and Wainwright, M. 1984. Nitrogen transformation in coastar sands and dune soils. Journal of Arid Environments, 7:18.Google Scholar
Smith, N.D., and Hein, F.J. 1971 Biogenic reworking of fluvial sediments by staphylinid beetles. Journal of Sedimentary Petrology, 41:598602.Google Scholar
Stanley, K.O., and Fagerstrom, J.A. 1974. Miocene invertebrate trace fossils from a braided river environment, western Nebraska, U.S.A. Palaeogeography, Palaeclimatology, Palaeoecology, 15:6382.Google Scholar
Stewart, D.J. 1978. Ophiomorpha: a marine indicator? Proceedings of the Geologists Association, 89:3341.Google Scholar
Stewart, J.H., Poole, F.G., and Wilson, R.F. 1972. Stratigraphy and origin of the Chinle Formation and related Upper Triassic strata in the Colorado Plateau region. U.S. Geological Survey Professional Paper 690, 336 p.Google Scholar
Tasch, P. 1976. Jurassic nonmarine trace fossils (Transantarctic Mountains) and the food web. Journal of Paleontology, 50:754758.Google Scholar
Thompson, W.O. 1949. Lyons Sandstone of Colorado Front Range. American Association of Petroleum Geologists, Bulletin, 33:5272.Google Scholar
Thorpe, J. 1949. Effects of certain animals that live in soils. Scientific Monthly, 68:180191.Google Scholar
Toots, H. 1967. Invertebrate burrows in the non-marine Miocene of Wyoming. Wyoming University Contributions to Geology, 6:9396.Google Scholar
Tevesz, M.J.S., and McCall, P.L. 1982. Geological Significance of aquatic nonmarine trace fossils, p. 257285. In McCall, P.L. and Tevesz, M.J.S., eds., Animal-Sediment Relations: The Biogenic Alteration of Sediments. Plenum Press, New York.Google Scholar
Turner, B.R. 1978. Trace fossils from the Upper Triassic fluviatile Molteno Formation of the Karoo (Gondawana) Supergroup, Lesotho. Journal of Paleontology, 52:959963.Google Scholar
Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F., Loutit, T.S., and Hardenbol, J. 1988. An overview of the fundamentals of sequence stratigraphy and key definitions, p. 3946. In Wilgus, C.K., Posamentier, H., Ross, C.A., and Kendall, C. G. St. G., eds., Sea Level Changes: An Integrated Approach. SEPM Special Publication No. 42.Google Scholar
Van Wagoner, J.C., Mitchum, R.M., Campion, K.M., and Rahmanian, V.D. 1990. Siliciclastic Sequence Stratigraphy in well logs, cores, and outcrops: Concepts for high-resolution correlation of time and facies. AAPG Methods in Exploration Series, No. 7, 55 p.Google Scholar
Van Wagoner, J.C., Nummedal, D., Jones, C.R., Taylor, D.R., Jennette, D.C., and Riley, G.W. 1991. Sequence stratigraphy applications to shelf sandstone reservoirs: Outcrop to subsurface examples. AAPG Field Conference Guidebook, Sept. 21-28.Google Scholar
Voorhies, M.R. 1975. Vertebrate burrows, p. 325350. In Frey, R.W., ed., The Study of Trace Fossils, Springer Verlag, New York.Google Scholar
Walker, R.G., and Harms, J.C. 1972. Eolian origin of flagstone beds, Lyons Sandstone (Permian), type area, Boulder County, Colorado. Mountain Geologist, 9:279288.Google Scholar
Warburg, M.R. 1992. Reproductive patterns in three isopod species from the Negev Desert. Journal of Arid Environments, 22:7385.Google Scholar
Weimer, R.J., and Hoyt, J.H. 1964. Burrows of Callianassa major Say, geologic indicators of littoral and shallow neritic environments. Journ. Paleont., 38:761767.Google Scholar
White, C.D. 1929. Flora of the Hermit shale, Grand Canyon, Arizona. Carnegie Institute of Washington Publication, 405, 221 p.Google Scholar
Williams, F.X. 1956. Life History studies of Pepsis and Hemipepsis wasps in California (Hymenoptera, Pompilidae). Annals of the Entomological Society of America, 49:447466.Google Scholar
Willis, E.R., and Roth, L.M. 1950. Humidity reactions of Tribolium castaneum (Herbst). Journal of Experimental Zoology, 115:567587.Google Scholar
Willis, E.R., and Roth, L.M. 1962. Soil and moisture relations of Scaptocoris divergins Troeschner (Hemiptera: Cynidae). Annals of the Entomological Society of America, 55:2132.Google Scholar