Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-29T03:35:59.875Z Has data issue: false hasContentIssue false

Paleobiogeography and Evolutionary History of Paleozoic Lacustrine Faunas

Published online by Cambridge University Press:  21 July 2017

Lisa E. Park
Affiliation:
Department of Geology, University of Akron, Akron, Ohio 44325-4101, USA
Elizabeth H. Gierlowski-Kordesch
Affiliation:
Department of Geological Sciences, Ohio University, Athens, Ohio 45701-2979, USA
Get access

Abstract

Lakes are important archives for continental records of paleoenvironmental as well as paleoclimatic change. They also record a unique macroevolutionary pattern that occurred when faunas invaded the continental realm. In order to document that pattern, we compiled a database of over ninety lake basins from the Neoproterozoic to the Permian. Each basin was evaluated based upon its sedimentology and paleontology and, where appropriate, was classified into one of three types: underfilled, balanced-filled, and overfilled, sensu Carroll and Bohacs (1999). The faunal elements from each were recorded at the species, generic, class, and phylum levels.

Looking at this critical time in lacustrine evolution, several patterns emerged. For the Neoproterozoic through Silurian time, there is a paucity of documented lake deposits and lake faunal records. Lakes during this time were oligotrophic and their nutrient cycling regimes were primitive. It is not until the establishment of land plants in the Silurian that lakes begin to respond with higher diversities and more complex physical and chemical conditions. During the Devonian-Carboniferous periods, diversity was on the rise as trophic levels became more complex. Globally, CO2 increased while marine Sr decreased, coinciding with the peaking of diversity within lakes. Most lakes of the Devonian and Carboniferous formed along continental margins or in tectonic basins with occasional connection to the marine realm. The faunas from these types of lakes were commonly comprised of mixed marine and freshwater elements and were far more diverse than other, more inland lakes. This “estuary effect” created a gateway or filter through which faunas invaded the continental realm.

The early history of lake faunas is one of opportunity and amelioration. The feedback loops created by the establishment of vascular plants altered the nutrient cycle on land and in lakes. All trophic levels were established early but became increasingly complex throughout the Paleozoic, as roles changed and faunal elements became established. Groups invading the continents via the “estuary effect” did so numerous times before establishing themselves permanently. This was linked with the episodic reestablishment in marine-freshwater connections along these continental margins. In general, the macroevolutionary history of lake faunas demonstrates a dramatically different diversification pattern than that of the marine, and further study is necessary to understand the intricacies of these patterns and whether or not they continue through the Mesozoic and Cenozoic eras.

Type
Research Article
Copyright
Copyright © 2005 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

Algeo, T. J., and Scheckler, S. E. 1998. Terrestrial-marine teleconnections in the Devonian: Links between the evolution of land plants, weathering processes, and marine anoxic events. Philosophical Transactions of the Royal Society of London B, 353:113130.CrossRefGoogle Scholar
Almond, J. E. 1985. The Silurian-Devonian fossil record of the Myriapoda. Philosophical Transactions of the Royal Society, London, B309:227237.Google Scholar
Anderson, R. R. 1997. Keweenawan Supergroup clastic rocks in the Midcontinent rift of Iowa, p. 211230. In Ojakangas, R. W., Dickas, A. B., and Green, J. C., (eds.), Middle Proterozoic to Cambrian Rifting, Central North America. Geological Society of America Special Paper 312.Google Scholar
Archer, A. W., Calder, J. H., Gibling, M. E., Naylor, R. D., Reid, D. R., and Wightman, W. G. 1995. Invertebrate trace fossils and agglutinated foraminifera as indicators of marine influence within the classic Carboniferous section at Joggins, Nova Scotia, Canada. Canadian Journal of Earth Sciences, 32:20272039.CrossRefGoogle Scholar
Armstrong, M., and Paterson, I. B. 1970. The Lower Old Red Sandstone of the Strathmore Region. Institute of Geological Sciences Report, No. 70/13. HMSO.Google Scholar
Awramik, S. M., Kidder, D. L., and Zieg, J. 1993. Collenia undosa, Charles Walcott, and lacustrine stromatolites. Geological Society of America Abstracts with Programs, 25, no. 6:A357.Google Scholar
Babcock, L., Miller, M. F., Isbell, J. L., Collinson, J. W., and Hasiotis, S. T. 1998. Paleozoic-Mesozoic crayfish from Antarctica: Earliest evidence of freshwater decapod crustaceans. Geology, 26:539542.2.3.CO;2>CrossRefGoogle Scholar
Baird, D. 1952. Revision of the Pennsylvanian and Permian footprints Limnopus, Allopus, and Baropus . Journal of Paleontology, 26:832840.Google Scholar
Baird, G. C., Sroka, S. D., Shavica, C. W., and Beard, T. L. 1985. Mazon Creek-type fossil assemblages in the U.S. midcontinent Pennsylvanian: Their recurrent character and palaeonenvironmental significance. Philosophical Transactions of the Royal Society, London, B311:8799.Google Scholar
Barlow, J. A. (ed). 1975. The age of the Dunkard. Proceedings of the First I.C. White Memorial Symposium, West Virginia Geological and Economic Survey, Morgantown, 352 p.Google Scholar
Barnes, R. S. K. 1989. What, if anything, is a brackish-water fauna? Transactions of the Royal Society of Edinburgh: Earth Sciences, 80:235240.CrossRefGoogle Scholar
Barr, S. M., and Peterson, K. C. A. 1998. Field relationships and petrology of the Late Devonian Fisset Brook Formation in the Cheticamp area, western Cape Breton Island, Nova Scotia. Atlantic Geology, 34:121134.CrossRefGoogle Scholar
Barron, E. J., and Fawcett, P. J. 1995. The climate of Pangaea: A review of climate model simulations of the Permian, p. 3752. In Scholle, P. A., Peryt, T. M. and Ulmer-Scholle, D. S., (eds.), Permian of Northern Pangea: Paleogeography, Paleoclimates, Stratigraphy. Springer-Verlag, Berlin.CrossRefGoogle Scholar
Behrensmeyer, A. K., Damuth, J. D., Dimichele, W. A., Potts, R., Sues, H.-D., and Wing, S. L. (eds.). 1992. Terrestrial Ecosystems Through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. University of Chicago Press, Chicago, 568 p.Google Scholar
Bension, K. C., and Goldstein, R. H. 2000. Sedimentology of ancient saline pans: An example from the Permian Opeche Shale, Williston basin, North Dakota, USA. Journal of Sedimentary Research, 70:159169.CrossRefGoogle Scholar
Bension, K. C., and Goldstein, R. H. 2001. Evaporites and siliciclastics of the Permian Nippewalla Group of Kansas, USA: A case for non-marine deposition in saline lakes and saline pans. Sedimentology, 48:165188.CrossRefGoogle Scholar
Bension, K. C., Goldstein, R. H., Wopenka, B., Burruss, R. C., and Pasteris, J. D. 1998. Extremely acid Permian lakes and groundwaters in North America. Nature, 392:911914.CrossRefGoogle Scholar
Berner, R. A., and Rye, D. M. 1992. Calculations of the Phanerozoic strontium isotope record of the oceans from a carbon cycle model. American Journal of Science, 292:136148.CrossRefGoogle Scholar
Bohacs, K. M., Carroll, A. R., Neal, J. E., and Mankiewicz, P. J. 2000. Lake-basin type, source potential, and hydrocarbon character: An integrated sequence-stratigraphic-geochemical framework, p. 333. In Gierlowski-Kordesch, E. H., and Kelts, K. R., (eds.), Lake Basins Through Space and Time. American Association of Petroleum Geologists Studies in Geology, No. 46.Google Scholar
Bohacs, K. M., Carroll, A. R., and Neal, J. E. 2003. Lessons from large lake systems: Thresholds, nonlinearity, and strange attractors. Geological Society of America Special Paper, 370:7590.Google Scholar
Boy, J. A., and Schindler, T. 2000. Okostratigraphische Bioevents im Grenzbereich Stephanium/Autunium (höchtes Karbon) des Saar-Nahe-Beckens (SW Deutschland) und benachbarter Gebiete. Neues Jahrbuch für Geologie und Palaeontologie Abhandlungen, 216:89152.CrossRefGoogle Scholar
Brand, U. 1994. Continental hydrology and climatology of the Carboniferous Joggins Formation (lower Cumberland Group) at Joggins, Nova Scotia: Evidence from the geochemistry of bivalves. Palaeogeography, Palaeoclimatology, Palaeoecology, 106:307321.CrossRefGoogle Scholar
Brookfield, M. E. 2000. Temporary desert lake deposits, lower Permian (Rotliegendes), southern Scotland, U.K., p. 6774. In Gierlowski-Kordesch, E. H., and Kelts, K. R., (eds.), Lake Basins Through Space and Time. American Association of Petroleum Geologists Studies in Geology, no. 46.Google Scholar
Buatois, L. A., and Mángano, M. G. 1993. Ecospace utilization, paleo-environmental trends, and the evolution of early nonmarine biotas. Geology, 21:595598.2.3.CO;2>CrossRefGoogle Scholar
Buatois, L. A., and Mángano, M. G. 1994. Lithofacies and depositional processes from a Carboniferous lake, Sierra de Narváez, northwest Argentina. Sedimentary Geology, 93:2549.CrossRefGoogle Scholar
Buatois, L. A., and Mángano, M. G. 2004. Animal-substrate interactions in freshwater environments: Applications of ichnology in facies and sequence stratigraphic analysis of fluvio-lacustrine successions. The Application of Ichnology to Palaeoenvironmental and Stratigraphic Analysis. Geological Society, London, Special Publications, 228:311333.CrossRefGoogle Scholar
Buatois, L. A., Limarino, C. O., and Césari, S. N. 1994. Carboniferous lacustrine deposits from the Paganzo basin, Argentina, p. 135140. In Gierlowski-Kordesch, E., and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge.Google Scholar
Buatois, L. A., Jalfin, G., and Acenolaza, F. G. 1997. Permian nonmarine invertebrate trace fossils from southern Patagonia, Argentina: Ichnologic signatures of substrate consolidation and colonization sequences. Journal of Paleontology, 71:324336.CrossRefGoogle Scholar
Buatois, L. A., Mángano, M. G., Genise, J. F., and Taylor, T. N. 1998. The ichnologic record of the continental invertebrate invasion: Evolutionary trends in environmental expansion, ecospace utilization, and behavioral complexity. Palaios, 13:217240.CrossRefGoogle Scholar
Calver, M. A. 1965. Coal measures invertebrate faunas, p. 147177. In Murchison, D. and Westoll, T. S., (eds.), Coal and Coal-Bearing Strata. Elsevier, New York.Google Scholar
Carasco, B. 1989. Lacustrine sedimentation in a Permian intermontane basin: The Villé graben (Vosges, France). Palaeogeography, Palaeoclimatology, Palaeoecology, 70:179186.CrossRefGoogle Scholar
Carroll, A. R., and Bohacs, K. M. 1999. Stratigraphic classification of ancient lakes: Balancing tectonic and climatic controls. Geology, 27:99102.2.3.CO;2>CrossRefGoogle Scholar
Carroll, A. R., and Bohacs, K. M. 2001. Lake-type controls on petroleum source rock potential in nonmarine basins. American Association of Petroleum Geologists Bulletin, 85:10331054.Google Scholar
Carter, D. C., and Pickerill, R. K. 1985. Algal swamp, marginal and shallow evaporitic lacustrine lithofacies from the Late Devonian-Early Carboniferous Albert Formation, southeastern New Brunswick, Canada. Maritime Sediments and Atlantic. Geology, 21:6986.Google Scholar
Cassle, C. F. 2005. Petrographic analysis of Late Pennsylvanian limestones within the northern Appalachian basin, USA. . Athens, OH, Ohio University, 227 p.Google Scholar
Caster, K. E., and Brooks, H. K. 1956. New fossils from the Canadian-Chazyan (Ordovician) hiatus in Tennessee. Bulletin of American Paleontology, 36:157199.Google Scholar
Chlupá, I. 1995. Lower Cambrian arthropods from the Paseky Shale (Barrandian area, Czech Republic). Journal of the Czech Geological Survey, 40:936.Google Scholar
Cohen, A. S. 2003. Paleolimnology: The History and Evolution of Lake Systems. Oxford University Press, Oxford, 500 p.CrossRefGoogle Scholar
Di Michele, W. A. 2001. Carboniferous coal-swamp forests. p. 7982. In Briggs, D. E. G. and Crowther, P. R., (eds.), Paleobiology II. Blackwell Publishing, Oxford.CrossRefGoogle Scholar
Driese, S. G. 1985. Interdune pond carbonates, Weber Sandstone (Pennsylvanian-Permian), northern Utah and Colorado. Journal of Sedimentary Petrology, 55:187195.Google Scholar
Duncan, W. I., and Buxton, N. W. K. 1995. New evidence for evaporitic Middle Devonian lacustrine sediments with hydrocarbon source potential on the East Shetland Platform, North Sea. Journal of the Geological Society, London, 152:251258.CrossRefGoogle Scholar
Eager, R. M. C., and Belt, E. S. 2003. Succession, palaeoecology, evolution, and speciation of Pennsylvanian non-marine bivalves, northern Appalachian Basin, USA. Geological Journal, 38:109143.CrossRefGoogle Scholar
Eggleston, J. R. 1994. The facies and depositional environment of an Upper Pennsylvanian limestone, northern Appalachian Basin, p. 297319. In Lomando, A. J., Schreiber, B. C., and Harris, P. M., (eds.), Lacustrine Reservoirs and Depositional Systems. SEPM Core Workshop, no. 19.CrossRefGoogle Scholar
Englund, K. J., Arndt, H. H., and Henry, T. W. (eds.). 1979. Proposed Pennsylvanian System Stratotype, Virginia and West Virginia. Field Trip No. 1, Ninth International Congress of Carboniferous Stratigraphy and Geology, American Geological Institute Selected Guidebook Series, No. 1.Google Scholar
Fayers, S. R., and Trewin, N. H. 2004. A review of the palaeoenvironments and biota of the Windyfield chert. Transactions of the Royal Society of Edinburgh, 94:325340.CrossRefGoogle Scholar
Frank, T. D., Lyons, T. W., and Lohmann, K. C. 1997. Isotopic evidence for the paleoenvironmental evolution of the Mesoproterozoic Helena Formation, Belt Supergroup, Montana, U.S.A. Geochimica et Cosmochimia Acta, 61:50235041.CrossRefGoogle Scholar
Friend, P. F., and Moody-Stuart, M. 1970. Carbonate deposition on the river floodplains of the Wood Bay Formation (Devonian) of Spitzbergen. Geological Magazine, 107:181195.CrossRefGoogle Scholar
Gand, G., Kerp, H., Parsons, C., and Martinez-Garcia, E. 1997. Palaeoenvironmental and stratigraphic aspects of animal traces and plant remains in Spanish Permian red beds (Peña Sagra, Cantabrian Mountains, Spain). Geobios, 30:295318.CrossRefGoogle Scholar
Gaupp, R., Gast, R., and Forster, C. 2000. Late Permian playa lake deposits of the southern Permian basin (central Europe), p. 7586. In Gierlowski-Kordesch, E. H., and Kelts, K. R., (eds.), Lake Basins Through Space and Time. American Association of Petroleum Geologists Studies in Geology, no. 46.Google Scholar
Gierlowski, E. H. 1978. On the paleogeographic distribution of vascular plants in the Devonian. . Chicago, University of Chicago, 60 p.Google Scholar
Gierlowski, E. H. and Kelts, K., (eds.). 1994. Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge, 427 p.Google Scholar
Gierlowski, E. H. and Kelts, K., (eds.). 2000. Lake Basins Through Space and Time. American Association of Petroleum Geologists Studies in Geology, no. 46. 648 p.CrossRefGoogle Scholar
Gierlowski, E. H., and Park, L. E. 2004. Comparing species diversity in the modern and fossil records of lakes. Journal of Geology, 112:703717.CrossRefGoogle Scholar
Gray, J. (ed.). 1988a. Aspects of freshwater paleoecology and biogeography. Palaeogeography, Palaeoclimatology, Palaeoecology, 62:1623.CrossRefGoogle Scholar
Gray, J. 1988b. Evolution of the freshwater ecosystem: The fossil record. Palaeogeography, Palaeoclimatology, Palaeoecology, 62:1214.CrossRefGoogle Scholar
Hamblin, A. P. 1992. Half-graben lacustrine sedimentary rocks of the lower Carboniferous Strathlorne Formation, Horton Group, Cape Breton Island, Nova Scotia, Canada. Sedimentology, 39:263284.CrossRefGoogle Scholar
Hamblin, A. P., and Rust, B. R. 1989. Tectono-sedimentary analysis of alternate-polarity half-graben basin-fill successions: The Late Devonian-Early Carboniferous Horton Group, Caper Breton Island, Nova Scotia. Basin Research, 2:239255.CrossRefGoogle Scholar
Hasiotis, S. T. 2003. Complex ichnofossils of solitary and social soil organisms: Understanding their evolution and roles in terrestrial paleoecosystems. Palaeogeography, Palaeoclimatology, Palaeoecology, 192:259320.CrossRefGoogle Scholar
Hembree, D. I., Martin, L. D., and Hasiotis, S. T. 2004. Amphibian burrows and ephemeral ponds of the lower Permian Speiser Shale, Kansas: Evidence for seasonality in the midcontinent. Palaeogeography, Palaeoclimatology, Palaeoecology, 203:127152.CrossRefGoogle Scholar
Hieshima, G. B., and Pratt, L. M. 1991. Sulfur/carbon ratios and extractable organic matter of the middle Proterozoic Nonesuch Formation, North American midcontinent rift. Precambrian Research, 54:6579.CrossRefGoogle Scholar
Higgs, R. 1994. Lake Bude (Upper Carboniferous), southwest England, p. 121125. In Gierlowski-Kordesch, E. and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge, UK.Google Scholar
Hook, R. W., and Ferm, J. C. 1985. A depositional model for the Linton tetrapod assemblage (Westphalian D, upper Carboniferous) and its palaeoenvironmental significance. Philosophical Transactions of the Royal Society, London, B311:101109.Google Scholar
Hook, R. W., and Baird, D. 1988. An overview of the Upper Carboniferous fossil deposit at Linton, Ohio. Ohio Journal of Science, 88:5560.Google Scholar
Jeram, A. J., Selden, P. A., and Edwards, D. 1990. Land animals in the Silurian: Arachnids and myriapods from Shropshire, England. Science, 250:658661.CrossRefGoogle ScholarPubMed
Johnson, E. W., Briggs, D. E. G., Suthren, R. J., Wright, J. L., and Tunnicliff, S. P. 1994. Nonmarine arthropod traces from the subaerial Ordovician Borrowdale Volcanic Group, English Lake District. Geological Magazine, 131:395406.CrossRefGoogle Scholar
Joyce, D., Lunt, D. H., Bills, R., Turner, G. F., Katongo, C., Duftner, N., Sturmbuaer, C., and Seehausen, O. 2005. An extant cichlid fish radiation emerged in an extinct Pleistocene lake. Nature, 435:9095.CrossRefGoogle Scholar
Kaaya, C. Z., and Kreuser, T. 1994. Permian lakes in East Africa (Tanzania), p. 8387. In Gierlowski-Kordesch, E. and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge.Google Scholar
Keighley, D. G., and Pickerill, R. K. 1997. Systematic ichnology of the Mabou and Cumberland Groups (Carboniferous) of western Cape Breton Island, eastern Canada, 1: Burrows, pits, trails, and coprolites. Atlantic Geology, 33:181215.CrossRefGoogle Scholar
Keighley, D. G., and Pickerill, R. K. 1998. Systematic ichnology of the Mabou and Cumberland Groups (Carboniferous) of western Cape Breton Island, eastern Canada, 2: Surface markings. Atlantic Geology, 34:83112.CrossRefGoogle Scholar
Keighley, D. G., and Pickerill, R. K. 2003. Ichnocoenoses from the Carboniferous of eastern Canada and their implications for the recognition of ichnofacies in nonmarine strata. Atlantic Geology, 39:122.CrossRefGoogle Scholar
Kelts, K. 1988. Environments of deposition of lacustrine petroleum source rocks: An introduction, p. 326. In Fleet, A. J., Kelts, K., and Talbot, M. R., (eds.), Lacustrine Petroleum Source Rocks, Geological Society (London) Special Publication, no. 40. Blackwell Scientific Publications, Oxford.Google Scholar
Kerp, H. 2000. The modernization of landscapes during the late Paleozoic-early Mesozoic, p. 79113. In Gastaldo, R. A. and DiMichele, W. A., (eds.), Phanerozoic Terrestrial Ecosystems. Paleontological Society Papers, 6.Google Scholar
Kidder, D. L., and Gierlowski-Kordesch, E. H., 2005. Impact of grassland radiation on the nonmarine silica cycle and Miocene diatomite. Palaios, 20:198206.CrossRefGoogle Scholar
Knox, L. W., and Gordon, E. A. 1999. Ostracodes as indicators of brackish water environments in the Catskill magnafacies (Devonian) of New York State. Palaeogeography, Palaeoclimatology, Palaeoecology, 148:922.CrossRefGoogle Scholar
Koren', T. N., Popov, L. E., Degtjarev, N. K., Kovalevsky, N. G., and Modzalevskaya, T. L. 2003. Kazakhstan in the Silurian, p. 323343. In Landing, E. and Johnson, M. E., (eds.), Silurian Lands and Seas: Paleogeography Outside of Laurentia. New York State Museum Bulletin, 493.Google Scholar
Legler, B., Schneider, J. W., Gand, G., and Koerner, F. 2004. Playa and sabhka environments from northern Germany and southern France, p. 6482. Freiberg University Workshop and IGCP 469 Central Meeting.Google Scholar
Limarino, C. O., and Césari, S. N. 1988. Paleoclimatic significance of the lacustrine Carboniferous deposits in northwest Argentina. Palaeogeography, Palaeoclimatology, Palaeoecology, 65; 115131.CrossRefGoogle Scholar
Loftus, G. W. F., and Greensmith, J. T. 1988. The lacustrine Burdiehouse Limestone Formation-A key to the deposition of the Dinantian oil shales of Scotland, p. 219234. In Fleet, A. J., Kelts, K., and Talbot, M. R., (eds.), Lacustrine Petroleum Source Rocks, Geological Society (London) Special Publication, no. 40. Oxford, Blackwell Scientific Publications.Google Scholar
Logan, G. H., Summons, R. E., and Hayes, J. M. 1997. An isotopic biogeochemical study of the Neoproterozoic and Early Cambrian sediments from the Centralian superbasin, Australia. Geochemica et Cosmochemica Acta, 61:53915409.CrossRefGoogle ScholarPubMed
Love, L. G. 1960. Assemblages of small spores from the lower Oil-Shale Group of Scotland. Proceedings of the Royal Society of Edinburgh, 67:99126.Google Scholar
Maples, C. G., and Archer, A. W. 1989. The potential of Paleozoic nonmarine trace fossils for paleoecological interpretations. Palaeogeography, Palaeoclimatology, Palaeoecology, 73:185195.CrossRefGoogle Scholar
Martin, R. E. 1996. Secular increase in nutrient levels through the Phanerozoic: Implications for productivity, biomass, and diversity of the marine biosphere. Palaios, 11:209219.CrossRefGoogle Scholar
Martin, W. D. 1998. Geology of the Dunkard Group (Upper Pennsylvanian-lower Permian) in Ohio, West Virginia, and Pennsylvania. Ohio Division of the Geological Survey Bulletin, 73, 49 p.Google Scholar
Mc Comas, G. A., and Mapes, R. H. 1988. Fauna associated with the Pennsylvanian floral zones of the 7–11 Mine, Columbiana County, northeastern Ohio. Ohio Journal of Science, 88:5355.Google Scholar
Mikulá_, R. 1995. Trace fossils from the Paseky Shale (Early Cambrian, Czech Republic). Journal of the Czech Geological Survey, 40:3744.Google Scholar
Miller, J. M. G. 1994. The Neoproterozoic Konnarock Formation, southwestern Virginia, USA: Glaciolacustrine facies in a continental rift, p. 4759. In Deynoux, M., Miller, J. M. G., Domack, E. W., Eyles, N., Fairchild, I. J., and Young, G. M., (eds.), Earth's Glacial Record. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Miller, M. F. 1984. Distribution of biogenic structures in Paleozoic nonmarine and marine-margin sequences: An actualistic model. Journal of Paleontology, 58:550570.Google Scholar
Miller, M. F. 1991. Morphology and palaeoenvironmental distribution of Paleozoic Spirophyton and Zoophycos: Implications for the Zoophycos ichnofacies. Palaios, 6:410425.CrossRefGoogle Scholar
Miller, M. F., and Collinson, J. W. 1994. Late Paleozoic post-glacial inland sea filled by fine-grained turbidites: Mackellar Formation, Central Transantarctic Mountains, p. 215233. In Deynoux, M., Miller, J. M. G., Domack, E. W., Eyles, N., Fairchild, I. J., and Young, G. M., (eds.), Earth's Glacial Record. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Miller, M. F., and Johnson, K. G. 1981. Spirophyton in alluvial-tidal facies of the Catskill deltaic complex: Possible biological control of ichnofossil distribution. Journal of Paleontology, 55:10161027.Google Scholar
Miller, M. F., and Labandeira, C. C. 2002. The slow crawl across the salinity divide: Delayed colonization of freshwater ecosystems by invertebrates. GSA Today, 12:410.2.0.CO;2>CrossRefGoogle Scholar
Miller, M. F., and Woodrow, D. L. 1991. Shoreline deposits of the Catskill Deltaic Complex, Schoharie Valley, New York, p. 153177. In Landing, E. and Brett, C. E., (eds.), Dynamic stratigraphy and depositional environments of the Hamilton Group (Middle Devonian) in New York State, Part II. New York State Museum Geological Survey, 469.Google Scholar
Miller, M. F., Kemp, N. R., Mc Dowell, T., Shyr, Y., and Smail, S. E. 2002. Hardly used habitats: Dearth and distribution of burrowing in Paleozoic and Mesozoic stream and lake deposits. Geology, 30:527530.2.0.CO;2>CrossRefGoogle Scholar
Murphy, J. B., and Rice, R. J. 1998. Stratigraphy and depositional environment of the Horton Group in the St. Marys basin, central mainland Nova Scotia. Atlantic Geology, 34:125.Google Scholar
Mykura, W. 1983. Old Red Sandstone, p. 205242. In Craig, G. Y., (ed.), Geology of Scotland. Scottish Academic Press, Edinburgh.Google Scholar
Nadon, G. C., Gierlowski-Kordesch, E. H., and Smith, J. P. 1998. Sedimentology and provenance of Carboniferous and Permian rocks of Athens County, southeastern Ohio. Ohio Geological Survey Guidebook, no. 15:123.Google Scholar
Nilsen, T. H., and Mc Laughlin, R. J. 1985. Comparison of tectonic framework and depositional patterns of the Hornelen strike-slip basin of Norway and the Ridge and Little Sulphur Creek strike-slip basins of California, p. 79103. In Biddle, K. T. and Christie-Blick, N., (eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. SEPM Special Publication, no. 37.CrossRefGoogle Scholar
Nitecki, M. H. (ed.). 1979. Mazon Creek Fossils. Academic Press, New York, 581 p.Google Scholar
Park, L. E. 2002. Phylogenetic congruence between hard and soft part data sets: How taphonomy affects ostracod phylogenies. 46th Annual Meeting, Abstracts, Palaeontological Association, Cambridge, 51:33.Google Scholar
Parnell, J. 1994. Late Proterozoic, northwest Scotland, U.K., p. 6364. In Gierlowski-Kordesch, E. and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge, UK, Cambridge University Press.Google Scholar
Piasecki, S., Christiansen, F. G., and Stemmerik, L. 1990. Depositional history of an Upper Carboniferous organic-rich lacustrine shale from East Greenland. Bulletin of Canadian Petroleum Geology, 38:273287.Google Scholar
Picard, M. D., and High, L. R. Jr. 1972. Criteria for recognizing lacustrine rocks, p. 108145. In Rigby, J. K. and Hamblin, W. K., (eds.), Recognition of Ancient Sedimentary Environments, SEPM Special Publication, no. 16.Google Scholar
Pickerill, R. K., Carter, D. C., and St. Peter, C. 1985. The Albert Formation: Oil shales, lakes, fans and deltas. Geological Association of Canada, Field Excursion, no. 6, 75 p.Google Scholar
Pickerill, R. K. 1990. Nonmarine Paleodictyon from the Carboniferous Albert Formation of southern New Brunswick. Atlantic Geology, 26:157163.CrossRefGoogle Scholar
Pollard, J. E., Steel, R. J., and Undersrud, E. 1982. Facies sequences and trace fossils in lacustrine/fan delta deposits, Hornelen basin (M. Devonian), western Norway. Sedimentary Geology, 32:6387.CrossRefGoogle Scholar
Ponomarenko, A. G. 1996. Evolution of continental aquatic ecosystems. Paleontological Journal, 30:705709.Google Scholar
Post, R. T. 2001. Sedimentology and tectonic significance of Cambrian stratigraphy, Muncho Lake, northern British Columbia: Evidence for the initiation of the Kechika Trough. , Sudbury, Ontario, Laurentian University, 262 p.Google Scholar
Pratt, L. M., Summons, R. E., and Hieshima, G. B. 1991. Sterane and Triterpane biomarkers in the Precambrian Nonesuch Formation, North American midcontinent rift. Geochimica et Cosmochimica Acta, 17:505509.Google Scholar
Raymond, A., Parker, W. C., and Barrett, S. B. 1985. Early Devonian phytogeography, p. 129167. In Tiffney, B. H., (ed.), Geological Factors and the Evolution of Plants. Yale University Press, New Haven, Connecticut.Google Scholar
Retallack, G. J. 2000. Ordovician life on land and early Paleozoic global change, p. 2145. In Gastaldo, R. A. and DiMichele, W. A., (eds.), Phanerozoic Terrestrial Ecosystems, Paleontological Society Papers, 6.Google Scholar
Rice, C. M., and Trewin, N. H. 1988. A Lower Devonian gold-bearing hot spring system, Rhynie, Scotland. Transactions of the Institute of Mining and Metallurgy, 97:B141B144.Google Scholar
Rice, C. M., Ashcroft, W. A., Batten, D. J., Boyce, A. J., Caulfield, J. B. D., Fallick, A. E., Hole, M. J., Jones, E., Pearson, M. J., Rogers, G., Saxton, J. M., Stuart, F. M., Trewin, N. H., and Turner, G. 1995. A Devonian auriferous hot spring system, Rhynie, Scotland. Journal of the Geological Society, London, 152:229250.CrossRefGoogle Scholar
Rice, C. M., Trewin, N. H., and Anderson, L. I. 2002. Geological setting of the Early Devonian Rhynie cherts, Aberdeenshire, Scotland: An early terrestrial hot spring system. Journal of the Geological Society, London, 159:203214.CrossRefGoogle Scholar
Richardson, E. S. Jr., and Johnson, R. G. 1971. The Mazon Creek faunas. Proceedings of the North American Paleontological Convention (1969), Part I:12221235.Google Scholar
Roberts, S. M. 1986. Belt Supergroup: A guide to Proterozoic rocks of western Montana and adjacent areas. Montana Bureau of Mines and Geology Special Publication 94, 311 p.Google Scholar
Rolfe, W. D. I. 1980. Early invertebrate terrestrial faunas, p. 117157. In Panchen, A. L., (ed.), The Terrestrial Environment and the Origin of Land Vertebrates. Systematics Association Special Volume, 15.Google Scholar
Ronchi, A., and Santi, G. 2003. Nonmarine biota from the lower Permian of the central Southern Alps (Orobic and Collio basins, N. Italy): A key to the paleoenvironment. Geobios, 36:749760.CrossRefGoogle Scholar
Schäfer, A., and Stapf, K. R. G. 1978. Permian Saar-Nahe basin and Recent Lake Constance (Germany): Two environments of lacustrine algal carbonates, p. 83107. In Matter, A. and Tucker, M. E., (eds.), Modern and Ancient Lake Sediments, International Association of Sedimentologists Special Publication, no. 2.CrossRefGoogle Scholar
Schwartzman, D. W., and Volk, T. 1989. Biotic enhancement of weathering and the habitability of Earth. Nature, 340:457460.CrossRefGoogle Scholar
Scotese, C. R. 2005. PALEOMAP Project (www.scotese.com).Google Scholar
Selley, R. C. 1965. Diagnostic characters of fluviatile sediments of the Torridonian Formation (Precambrian) of northwest Scotland. Journal of Sedimentary Petrology, 35:366380.CrossRefGoogle Scholar
Shear, W. A., Bonamo, P. M., Grierson, J. D., Rolfe, W. D. I., Smith, E. L., and Norton, R. A. 1984. Early land animals in North America: Evidence from Devonian age arthropods from Gilboa, New York. Science, 224:492494.CrossRefGoogle ScholarPubMed
Shear, W. A., and Kukalová-Peck, J. 1990. The ecology of Paleozoic terrestrial arthropods: The fossil evidence. Canadian Journal of Zoology, 68:18071834.CrossRefGoogle Scholar
Sko_Ek, V. 1994. The Stephanian B lake, Bohemian basin, Czech Republic, p. 8990. In Gierlowski-Kordesch, E. and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge, UK.Google Scholar
Smith, R. M. H. 1990. Alluvial paleosols and pedofacies sequences in the Permian lower Beaufort of the southwestern Karoo basin, South Africa. Journal of Sedimentary Petrology, 60:258276.Google Scholar
Smith, R. M. H. 1993. Sedimentology and ichnology of floodplain paleosurfaces in the Beaufort Group (late Permian), Karoo sequence, South Africa. Palaios, 8:339357.CrossRefGoogle Scholar
Solle, G. 1935. Die Devonische Ostracoden Spitzbergens I. Leperditiidae. Skrifter Svalbard og Ishavet, 64, 61 p.Google Scholar
Southgate, P. N., Lambert, I. B., Donnelly, T. H., Henry, R., Etminan, H., and Weste, G. 1989. Depositional environments and diagenesis in Lake Parakeelya: A Cambrian alkaline playa from the Officer basin, South Australia. Sedimentology, 36:10911112.CrossRefGoogle Scholar
Stapf, K. R. G. 1989. Biogene fluvio-lakustrine Sedimentation im Rotliegend des permokarbonen Saar-Nahe-Beckens (SW-Deutschland). Facies, 20:169198.CrossRefGoogle Scholar
Stollhofen, H. 1998. Facies architecture variations and siesmogenic structures in the Carboniferous-Permian Saar-Nahe basin (SW Germany): Evidence for extension-related transfer fault activity. Sedimentary Geology, 119:4783.CrossRefGoogle Scholar
Stollhofen, H., and Stanistreet, I. G. 1994. Interaction between bimodal volcanism, fluvial sedimentation, and basin development in the Permo-Carboniferous Saar-Nahe basin (south-west Germany). Basin Research, 6:245267.CrossRefGoogle Scholar
Suszek, T. 1997. Petrography and sedimentation of the middle Proterozoic (Keweenawan) Nonesuch Formation, western Lake Superior region, Midcontinent rift system, p. 195210. In Ojakangas, R. W., Dickas, A. B. and Green, J. C., (eds.), Middle Proterozoic to Cambrian Rifting, Central North America. Geological Society of America Special Paper, 312.Google Scholar
Szulc, J., and Cwizewicz, M. 1989. The lower Permian freshwater carbonates of the Slawkow graben, southern Poland: Sedimentary facies context and stable isotope study. Palaeogeography, Palaeoclimatology, Palaeoecology, 70:107120.CrossRefGoogle Scholar
Thomson, K. S., Sutton, M., and Thomas, B. 2003. A larval Devonian lungfish. Nature, 426:833834.CrossRefGoogle ScholarPubMed
Tiebert, N. E., and Scott, D. B. 1999. Ostracodes and agglutinated foraminifera as indicators of paleoenvironmental change in an early Carboniferous brackish bay, Atlantic Canada. Palaios, 14:246260.CrossRefGoogle Scholar
Toutin-Morin, N. 1994. Lacustrine and palustrine carbonates in the Permian of eastern Provenance (France), p. 9196. In Gierlowski-Kordesch, E. and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge, UK.Google Scholar
Trewin, N. H. 2000. The ichnogenus Undichna, with examples from the Permian of the Falkland Islands. Palaeontology, 43:979997.CrossRefGoogle Scholar
Trewin, N. H., and Davidson, R. J. 1996. An Early Devonian lake and its associated biota in the Midland Valley of Scotland. Transactions of the Royal Society of Edinburgh: Earth Sciences, 86:233246.CrossRefGoogle Scholar
Trewin, N. H., Macdonald, D. I. M., and Thomas, C. G. C. 2002. Stratigraphy and sedimentology of the Permian of the Falkland Islands: Lithostratigraphic and palaeoenvironmental links with South Africa. Journal of the Geological Society, London, 259:519.CrossRefGoogle Scholar
Tunbridge, I. P. 1984. Facies model for a sandy ephemeral stream and clay playa complex; the Middle Devonian Trentishoe Formation of North Devon, U.K. Sedimentology, 31:697715.CrossRefGoogle Scholar
Turner, B. R., and Smith, D. B. 1997. A playa deposit of pre-Yellow Sands age (upper Rotliegend/Weissliegend) in the Permian of northeast England. Sedimentary Geology, 114:305319.CrossRefGoogle Scholar
Valero Garcés, B. L. 1994a. The Permian lacustrine Basque basin (western Pyrenees): An example for high environmental variability in small and shallow carbonate lakes developed in closed systems, p. 101105. In Gierlowski-Kordesch, E. and Kelts, K., (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge, UK.Google Scholar
Valero Garcés, B. L. 1994b. Carbonate lacustrine episodes in the continental Permian Aragón-Béarn basin (western Pyrenees), p. 107119. In Gierlowski-Kordesch, E. and Kelts, K. (eds.), Global Geological Record of Lake Basins, Volume 1. Cambridge University Press, Cambridge, UK.Google Scholar
Valero Garcés, B. L., and Aguilar, J. G. 1992. Shallow carbonate lacustrine facies models in the Permian of the Aragon-Bearn basin (western Spanish-French Pyrenees). Carbonates and Evaporites, 7:94107.CrossRefGoogle Scholar
Valero Garcés, B. L., Gierlowski-Kordesch, E. H., and Bragonier, W. A. 1997. Pennsylvanian continental cyclothem development: No evidence of direct climate control in the Upper Freeport Formation (Allegheny Group) of Pennsylvania (northern Appalachian Basin). Sedimentary Geology, 109:305319.CrossRefGoogle Scholar
Van De Poll, H.W., Gibling, M. R., and Hyde, R. S. 1995. Chapter 5-Upper Paleozoic rocks, p. 449455. In Williams, H., (ed.), Geology of the Appalachian-Caledonian Orogen in Canada and Greenland, Geology of Canada No. 6. Geological Survey of Canada, Ottawa, Canada.Google Scholar
Wartes, M. A., Carroll, A. R., Greene, T. J., Cheng, K., and Ting, H. 2000. Permian lacustrine deposits of northwest China, p. 123132. In Gierlowski-Kordesch, E. H. and Kelts, K. R., (eds.), Lake Basins Through Space and Time. American Association of Petroleum Geologists Studies in Geology, no. 46.Google Scholar
Weir, J., and Leitch, D. 1936. The zonal distribution of the nonmarine lamellibranches in the coal measures of Scotland. Transactions of the Royal Society of Edinburgh, 58:697751.Google Scholar
Wetzel, R. G. 2001. Limnology: Lake and River Ecosystems. 3rd Edition. Academic Press, San Diego, 1006 p.Google Scholar
White, A. H., and Youngs, B.C. 1980. Cambrian alkali playa-lacustrine sequece in the northeastern Officer basin, South Australia. Journal of Sedimentary Petrology, 50:12791286.Google Scholar
White, C. E., and Barr, S. M. 1998. Stratigraphy and tectonic significance of the Lower to Middle Devonian McAdams Lake Formation, Cape Breton Island, Nova Scotia. Atlantic Geology, 34:133145.CrossRefGoogle Scholar
Williams, E. G. 1960. Marine and freshwater fossiliferous beds in the Pottsville and Allegheny Groups of western Pennsylvania. Journal of Paleontology, 34:908922.Google Scholar
Wilson, E. O. 1999. The Diversity of Life. W.W. Norton and Co., New York, 424 p.Google Scholar
Yemane, K. 1993. Contribution of Late Permian palaeogeography in maintaining a temperate climate in Gondwana. Nature, 361:5154.CrossRefGoogle Scholar
Yemane, K., Siegenthaler, C., and Kelts, K. 1989. Lacustrine environment during Lower Beaufort (upper Permian) Karoo deposition in northern Malawi. Palaeogeography, Palaeoclimatology, Palaeoecology, 70:165178.CrossRefGoogle Scholar
Yemane, K., Kahr, G., and Kelts, K. 1996. Imprints of post-glacial climates and palaeogeography in the detrital clay mineral assemblages of an upper Permian fluviolacustrine Gondwana deposit from northern Malawi. Palaeogeography, Palaeoclimatology, Palaeoecology, 125:2749.CrossRefGoogle Scholar
Zhang, Z., Sun, K., and Yin, J. 1997. Sedimentology and sequence stratigraphy of the Shanxi Formation (Lower Permian) in the northwestern Ordos Basin, China: An alternative sequence model for fluvial strata. Sedimentary Geology, 112:123136.CrossRefGoogle Scholar
Zhang, G., Buatois, L. A., Mángano, M. G., and Acenolaza, F. G. 1998. Sedimentary facies and environmental ichnology of a ?Permian playa-lake complex in western Argentina. Palaeogeography, Palaeoclimatology, Palaeoecology, 138:221243.CrossRefGoogle Scholar
Zhao, X., and Tang, Z. 2000. Lacustrine deposits of the upper Permian Pingdiquan Formation in the Kelameili area of the Junggar basin, Xinjiang, China, p. 109122. In Gierlowski-Kordesch, E. H. and Kelts, K. R., (eds.), Lake Basins Through Space and Time. American Association of Petroleum Geologists Studies in Geology, no. 46.Google Scholar
Ziegler, A. M. 1990. Phytogeographic patterns and continental configurations during the Permian Period, p. 363379. In McKerrow, W. S. and Scotese, C. R., (eds.), Palaeozoic palaeogeography and biogeography. Geological Society of London Memoir, 12.Google Scholar
Ziegler, A. M., Barrett, S. F., and Scotese, C. R. 1981. Palaeoclimate, sedimentation and continental accretion. Philosophical Transactions of the Royal Society of London, A301:253264.Google Scholar