Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-lvtdw Total loading time: 0 Render date: 2024-08-18T19:24:02.274Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2012

John Bridge
Affiliation:
State University of New York, Binghamton
Robert Demicco
Affiliation:
State University of New York, Binghamton
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2008

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

Aber, J. S., Croot, D. G., & Fenton, M. M. 1989. Glaciotectonic Landforms and Structures. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Abrahams, A. D. 1986. Hillslope Processes. Boston: Allen and Unwin.Google Scholar
Abrahams, A. D., & Parsons, A. J. 1994. Geomorphology of Desert Environments. London: Chapman and Hall.CrossRefGoogle Scholar
Adams, A. E., & MacKenzie, W. S. 1998. A Colour Atlas of Carbonate Sediments and Rocks under the Microscope. London: Manson Publishing.Google Scholar
Adams, A. E., MacKenzie, W. S., & Guilford, C. 1984. Atlas of Sedimentary Rocks under the Microscope. New York: John Wiley and Sons.Google Scholar
Adams, J. E., & Rhodes, M. L. 1960. Dolomitization by seepage refluxion. American Association of Petroleum Geologists Bulletin 44, 1912–1921.Google Scholar
Ahlbrandt, T. S., & Fryberger, S. G. 1980. Eolian deposits in the Nebraska Sand Hills. United States Geological Survey Professional Paper1120A.
Ahlbrandt, T. S., & Fryberger, S. G.1981. Sedimentary features and significance of interdune deposits. In Recent and Ancient Nonmarine Depositional Environments: Models for Exploration (ed. Ethridge, F. G. and Flores, R. M.), pp. 293–314. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Aigner, T., & Reineck, H. -E. 1982. Proximality trends in modern storm sands from the Helgoland Bight (North Sea) and their implications for basin analysis. Senckenbergiana Maritima 14, 183–215.Google Scholar
Aitken, J. F., & Flint, S. S. 1995. The application of high-resolution sequence stratigraphy to fluvial systems: a case study from the Upper Carboniferous Breathitt Group, eastern Kentucky, USA. Sedimentology 42, 3–30.CrossRefGoogle Scholar
Alexander, C. R, Davis, R. A., & Henry, V. J. 1998. Tidalites: Processes and Products. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Alexander, C. R., Nittrouer, C. A., DeMaster, D. J., Park, Y. A., & Park, S. C. 1991. Macrotidal mudflats of the southwestern Korean coast: a model for interpretation of intertidal deposits. Journal of Sedimentary Petrology 61, 805–824.Google Scholar
Alexander, J., Bridge, J. S., Cheel, R. J., & Leclair, S. F. 2001. Bed forms and associated sedimentary structures formed under supercritical water flows over aggrading sand beds. Sedimentology 48, 133–152.CrossRefGoogle Scholar
Alexander, J., Bridge, J. S., Leeder, M. R., Collier, R. E. L., & Gawthorpe, R. L. 1994. Holocene meander belt evolution in an active extensional basin, southwestern Montana. Journal of Sedimentary Research B64, 542–559.Google Scholar
Alexander, J., & Leeder, M. R. 1987. Active tectonic control on alluvial architecture. In Recent Developments in Fluvial Sedimentology (ed. Ethridge, F. G., Flores, R. M., & Harvey, M. D.), pp. 243–252. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Alexander, J., & Morris, S. 1994. Observations on the experimental, nonchannelised, high-concentration turbidity currents and variations in deposits around obstacles. Journal of Sedimentary Research 64A, 899–909.Google Scholar
Algeo, T. J., Scheckler, S. E., & Maynard, J. B. 2001. Effects of the Middle to Late Devonian spread of vascular plants on weathering regimes, marine biotas, and global climate. In Plants Invade the Land (ed. Gensel, P. G. & Edwards, D.), pp. 216–236. New York: Columbia University Press.CrossRefGoogle Scholar
Algeo, T. J., & Wilkinson, B. H. 1988. Periodicity of mesoscale Phanerozoic sedimentary cycles and the role of Milankovitch orbital modulation. Journal of Geology 96, 313–322.CrossRefGoogle Scholar
Allen, G. P. 1991. Sedimentary processes and facies in the Gironde estuary: a recent model for macrotidal estuarine systems. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 29–40. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Allen, J. R. L. 1965. A review of the origin and characteristics of recent alluvial sediments. Sedimentology 5, 89–191.CrossRefGoogle Scholar
Allen, J. R. L. 1966. On bedforms and paleocurrents. Sedimentology 6, 153–190.CrossRefGoogle Scholar
Allen, J. R. L. 1970. Physical Processes of Sedimentation. London: George Allen and Unwin.Google Scholar
Allen, J. R. L. 1973. Features of cross-stratified units due to random and other changes in bed forms. Sedimentology 20, 189–202.CrossRefGoogle 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, 181–208.CrossRefGoogle Scholar
Allen, J. R. L. 1978. Studies in fluviatile sedimentation: an exploratory quantitative model for architecture of avulsion-controlled alluvial suites. Sedimentary Geology 21, 129–147.CrossRefGoogle Scholar
Allen, J. R. L. 1979a. A model for the interpretation of wave-ripple marks using their wavelength, textural composition and shape. Journal of the Geological Society of London 136, 673–682.CrossRefGoogle Scholar
Allen, J. R. L. 1979b. Studies in fluviatile sedimentation: an elementary geometrical model for the connectedness of avulsion-related channel sand bodies. Sedimentary Geology 24, 253–267.CrossRefGoogle Scholar
Allen, J. R. L. 1980. Sand waves: a model of origin and internal structure. Marine Geology 26, 281–328.Google Scholar
Allen, J. R. L. 1982a. Sedimentary Structures: Their Character and Physical Basis. Amsterdam: Elsevier Science Publishers.Google Scholar
Allen, J. R. L. 1982b. Mud drapes in sand-wave deposits: a physical model with application to the Folkestone beds (early Cretaceous, southeast England). Philosophical Transactions of the Royal Society of London 306A, 291–345.CrossRefGoogle Scholar
Allen, J. R. L. 1985. Principles of Physical Sedimentology. London: George Allen and Unwin.Google Scholar
Allen, J. R. L. 1991. The Bouma division A and the possible duration of turbidity currents. Journal of Sedimentary Petrology 61, 291–295.Google Scholar
Allen, J. R. L., & Banks, N. L. 1972. An interpretation and analysis of recumbent-folded deformed cross bedding. Sedimentology 19, 257–283.CrossRefGoogle Scholar
Allen, J. R. L., & Pye, K. 1992. Saltmarshes. Cambridge: Cambridge University Press.Google Scholar
Allen, P. A. 1981. Some guidelines in reconstructing ancient sea conditions from wave ripple marks. Marine Geology 43, M59–67.CrossRefGoogle Scholar
Allen, P. A. 1985. Hummocky cross-stratification is not produced purely under progressive gravity waves. Nature 313, 562–564.CrossRefGoogle Scholar
Allen, P. A. 1997. Earth Surface Processes. Oxford: Blackwells.CrossRefGoogle Scholar
Allen, P. A., & Allen, J. R. 2005. Basin analysis, 2nd edn. Oxford: Blackwells.Google Scholar
Allen, P. A., & Densmore, A. L. 2000. Sediment flux from an uplifting fault block. Basin Research 12, 367–380.CrossRefGoogle Scholar
Allen, P. A., & Homewood, P. 1984. Evolution and mechanics of a Miocene tidal sand wave. Sedimentology 31, 63–81.CrossRefGoogle Scholar
Alley, R. B. 1991. Deforming-bed origin for southern Laurentide till sheets. Journal of Glaciology 37, 67–76.CrossRefGoogle Scholar
Alley, R. B., Lawson, D. E., Evenson, E. B., Larson, G. J., & Baker, G. S. 2003. Stabilizing feedbacks in glacier bed erosion. Nature 424, 758–760.CrossRefGoogle ScholarPubMed
Alley, R. B., Lawson, D. E., Evenson, E. B., Strasser, J. C., & Larson, G. J. 1998. Glaciohydraulic supercooling: a freeze-on mechanism to create stratified, debris-rich basal ice. 2. Theory. Journal of Glaciology 44, 563–569.CrossRefGoogle Scholar
Alsharhan, A. S., & Kendall, C. G.St, C. 2003. Holocene coastal carbonates and evaporites of the southern Arabian Gulf and their ancient analogues. Earth-Science Reviews 61, 191–243.CrossRefGoogle Scholar
American Society of Civil Engineers Task Force on Bed Forms in Alluvial Channels 1966. Nomenclature for bedforms in alluvial channels. Journal of the Hydraulics Division ASCE, 92, 51–64.
Amos, C. L., Li, M. Z., & Choung, K.-S. 1996. Storm-generated, hummocky stratification on the outer-Scotian shelf. Geomarine Letters 16, 85–94.Google Scholar
Anadón, P., Cabrera, L., & Kelts, K. 1991. Lacustrine Facies Analysis. Oxford: Blackwells.CrossRefGoogle Scholar
Anbar, A. D., & Knoll, A. H. 2002. Proterozoic ocean chemistry and evolution: a bioinorganic bridge?Science 297, 1137–1192.CrossRefGoogle ScholarPubMed
Anderson, J. B., & Ashley, G. M. 1991. Glacial Marine Sedimentation: Palaeoclimatic Significance. Boulder, CO: Geological Society of America.Google Scholar
Anderson, M. G., Walling, D. E., & Bates, P. D. 1996. Floodplain Processes. Chichester: Wiley.Google Scholar
Anderson, M. P. 1997. Characterization of geological heterogeneity. In Subsurface Flow and Transport: A Stochastic Approach (ed. Dagan, G. & Neuman, S. P.), pp. 23–43. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Anderson, R. S. 1987. A theoretical model of aeolian impact ripples. Sedimentology 34, 943–956.CrossRefGoogle Scholar
Anderson, R. S., & Bunas, K. L. 1993. Grain size segregation and stratigraphy in aeolian ripples modeled with a cellular automaton. Nature 365, 740–743.CrossRefGoogle Scholar
Archer, A. W. 1998. Hierarchy of controls on cyclic rhythmite deposition: Carboniferous basins of eastern and mid-continental USA. In Tidalites: Processes and Products (ed. Alexander, C. R., Davis, R. A., & Henry, V. J.), pp. 59–68. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Arnott, R. W. C., & Hand, B. M. 1989. Bedforms, primary structures and grain fabric in the presence of sediment rain. Journal of Sedimentary Petrology 59, 1062–1069.Google Scholar
Arnott, R. W., & Southard, J. B. 1990. Experimental study of combined-flow bed configurations in fine sands, and some implications for stratification. Journal of Sedimentary Petrology 60, 211–219.Google Scholar
Aronson, J. L., & Hower, J. 1976. Mechanism of burial metamorphism of argillaceous sediment: 2, radiogenic argon evidence. Geological Society of America Bulletin 87, 738–744.2.0.CO;2>CrossRefGoogle Scholar
Arthur, M. A., Dean, W. E., & Stow, D. A. V. 1984. Models for the deposition of Mesozoic–Cenozoic fine-grained organic-carbon-rich sediment in the deep sea. In Fine-Grained Sediments: Deep-Water Processes and Facies (ed. Stow, D. A. V. & Piper, D. J. W.), pp. 527–560. London: Geological Society of London.Google Scholar
Arthur, M. A., & Sageman, B. B. 1994. Marine black shales: depositional mechanisms and environments of ancient deposits. Annual Review of Earth and Planetary Sciences 22, 499–551.CrossRefGoogle Scholar
Ashley, G. M. 1988. The hydrodynamics and sedimentology of a back-barrier lagoon–salt marsh system, Great Sound, New Jersey. Marine Geology 82, 1–132.Google Scholar
Ashley, G. M. 1990. Classification of large-scale subaqueous bed-forms: a new look at an old problem. Journal of Sedimentary Petrology 60, 160–172.Google Scholar
Ashley, G. M.2002. Glaciolacustrine environments. In Modern and Past Glacial Environments (ed. Menzies, J.), pp. 335–359. Oxford: Butterworth-Heinemann.CrossRef
Ashley, G. M., Shaw, J., & Smith, N. D. 1985. Glacial Sedimentary Environments. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ashley, G. M., Wellner, R. W., Esker, D., & Sheridan, R. E. 1991. Clastic sequences developed during late Quaternary glacio-eustatic sea-level fluctuations on a passive margin: example from the inner continental shelf near Barnegat Inlet, New Jersey. Geological Society of America Bulletin 103, 1607–1621.2.3.CO;2>CrossRefGoogle Scholar
Ashworth, P. J., Best, J. L., & Jones, M. 2004. Relationship between sediment supply and avulsion frequency in braided rivers. Geology 32, 21–24.CrossRefGoogle Scholar
Aslan, A., & Blum, M. D. 1999. Contrasting styles of Holocene avulsion, Texas Gulf Coastal Plain, USA. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 193–209 J. Oxford: Blackwells.CrossRefGoogle Scholar
Assereto, R., & Folk, R. L. 1980. Diagenetic fabrics of aragonite, calcite, and dolomite in an ancient peritidal-spelean environment: Triassic Calcare Rosso, Lombardia, Italy. Journal of Sedimentary Petrology 50, 371–394.Google Scholar
Atwater, B. F., & Hemphill-Haley, E. 1997. Recurrence intervals for great earthquakes of the past 3,500 years at northeastern Willapa Bay, Washington. United States Geological Survey Professional Paper 1576.
Atwater, B. F., Musumi-Rokaku, S., Satake, K. et al. 2005. The Orphan Tsunami of 1700: Japanese clues to a parent earthquake in North America. United States Geological Survey Professional Paper 1707.
Augustinius, P. G. E. F. 1989. Cheniers and chenier plains: a general introduction. Marine Geology 90, 219–229.CrossRefGoogle Scholar
Augustinius, P. G. E. F., Hazelhoff, L., & Kroon, A. 1989. The chenier coast of Suriname; modern and geological development. Marine Geology 90, 269–281.CrossRefGoogle Scholar
Ausich, W. I., & Bottjer, D. J. 1982. Tiering in suspension-feeding communities on soft substrata throughout the Phanerozoic. Science 216, 173–174.CrossRefGoogle ScholarPubMed
Ausich, W. I., & Bottjer, D. J.1985. Phanerozoic tiering in suspension-feeding communities on soft substrata: implications for diversity. In Phanerozoic Diversity Patterns (ed. Valentine, J. W.), pp. 255–274. New Jersey: Princeton University Press.Google Scholar
Autin, W. J., Burns, S. F., Miller, R. T., Saucier, R. T., & Snead, J. I. 1991. Quaternary geology of the lower Mississippi valley. In Quaternary Nonglacial Geology: Conterminous U.S. The Geology of North America, pp. 547–582. Boulder, CO: Geological Society of America.Google Scholar
Baas, J. H., Kesteren, W., & Postma, G. 2004. Deposits of depletive high-density turbidity currents: a flume analogue of bed geometry, structure and texture. Sedimentology 51, 1053–1088.CrossRefGoogle Scholar
Badiozamani, K. 1973. The dorag dolomitization model – application to the Middle Ordovician of Wisconsin. Journal of Sedimentary Petrology 43, 965–984.Google Scholar
Baeuerlein, E. 2004. Biomineralization: Progress in Biology, Molecular Biology and Application. Weinheim: Wiley-VCH.CrossRefGoogle Scholar
Bagnold, R. A. 1941. The Physics of Blown Sand and Desert Dunes. London: Methuen.Google Scholar
Bagnold, R. A. 1946. Motion of waves in shallow water. Interaction between waves and sand bottoms. Proceedings of the Royal Society of London 187A, 1–16.Google Scholar
Bagnold, R. A. 1954. Experiments on the gravity-free dispersion of large solid spheres in a Newtonian fluid under shear. Proceedings of the Royal Society of London 187A, 49–63.CrossRefGoogle Scholar
Bagnold, R. A. 1956. The flow of cohesionless grains in fluids. Philosophical Transactions of the Royal Society 249A, 235–297.CrossRefGoogle Scholar
Bagnold, R. A. 1960. Some aspects of the shape of river meanders. United States Geological Survey Professional Paper 282E, pp. 135–144.Google Scholar
Bagnold, R. A. 1962. Auto-suspension of transported sediment: turbidity currents. Philosophical Transactions of the Royal Society 265A, 315–319.Google Scholar
Bagnold, R. A.1963. Mechanics of marine sedimentation. In The Sea: Volume 3 (ed. Hill, M. N.), pp. 507–528. New York: Wiley.Google Scholar
Bagnold, R. A. 1966. An approach to the sediment transport problem from general physics. United States Geological Survey Professional Paper 442-I.
Bagnold, R. A. 1973. The nature of saltation and of “bed-load” transport in water. Philosophical Transactions of the Royal Society 332A, 473–504.Google Scholar
Baker, V. R. 1973. Palaeohydrology and sedimentology of Lake Missoula flooding in eastern Washington. Geological Society of America Special Paper 144.
Baker, V. R. 1990. Geological fluvial geomorphology. Geological Society of America Bulletin 100, 1157–1167.2.3.CO;2>CrossRefGoogle Scholar
Baker, V. R., Kochel, R. C., & Patton, P. C. 1988. Flood Geomorphology. New York: Wiley.Google Scholar
Baker, V. R., & Nummedal, D. 1978. The Channeled Scabland. Washington, D.C.: NASA Planetary Geology Program.Google Scholar
Baldwin, B., & Butler, C. O. 1985. Compaction curves. American Association of Petroleum Geologists Bulletin 69, 622–626.Google Scholar
Ball, M. M. 1967. Carbonate sand bodies of Florida and the Bahamas. Journal of Sedimentary Petrology 37, 556–591.Google Scholar
Banerjee, I., & McDonald, B. C. 1975. Nature of esker sedimentation. In Glaciofluvial and Glaciolacustrine Sediments (ed. Jopling, A. V. & McDonald, B. C.), pp. 132–154. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Barnes, M. A., Barnes, W. C., & Bustin, R. M. 1990. Chemistry and diagenesis of organic matter in sediments and fossil fuels. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 189–204. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Barnes, R. S. K., & Hughes, R. N. 1982. An Introduction to Marine Ecology. Oxford: Blackwells.Google Scholar
Barry, J. M. 1997. Rising Tide: The Great Mississippi Flood of 1927 and How It Changed America. New York: Simon and Schuster.Google Scholar
Barwis, J. H. 1978. Sedimentology of some South Carolina tidal creek point bars, and a comparison with their fluvial counterparts. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 129–160. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Bates, C. C. 1953. Rational theory of delta formation. American Association of Petroleum Geologists Bulletin 37, 2119–2162.Google Scholar
Bathurst, R. G. C. 1975. Carbonate Sediments and Their Diagenesis, 2nd edn. Amsterdam: Elsevier.Google Scholar
Bathurst, R. G. C. 1980. Lithification of carbonate sediments. Science Progress Oxford 66, 451–471.Google Scholar
Bathurst, R. G. C. 1987. Diagenetically enhanced bedding in argillaceous platform limestones: stratified cementation and selective compaction. Sedimentology 34, 749–778.CrossRefGoogle Scholar
Batten, K. L., Narbonne, G. M., & James, N. P. 2004. Paleoenvironments and growth of early Neoproterozoic calcimicrobial reefs: platformal Little Dal Group, northwestern Canada. Precambrian Research 133, 249–269.CrossRefGoogle Scholar
Bear, J. 1972. Dynamics of Fluids in Porous Media. New York: Dover Publications.Google Scholar
Beard, B. L., Johnson, C. M., Cox, L.et al. 1999. Iron isotope biosignatures. Science 285, 1889–1892.CrossRefGoogle ScholarPubMed
Beaumont, C. 1981. Foreland basins. Geophysical Journal of the Royal Astronomical Society 65, 291–329.CrossRefGoogle Scholar
Beaumont, C., Fullsack, P., & Hamilton, J. 1992. Erosional control of active compressional orogens. In Thrust Tectonics (ed. McClay, K. R.), pp. 1–18. London: Chapman and Hall.CrossRefGoogle Scholar
Beaumont, C., Quinlan, G., & Hamilton, J. 1988. Orogeny and stratigraphy: numerical models of the Paleozoic in the eastern interior of North America. Tectonics 7, 389–416.CrossRefGoogle Scholar
Behrensmeyer, A. K. 1987. Miocene fluvial facies and vertebrate taphonomy in northern Pakistan. In Recent Developments in Fluvial Sedimentology (ed. Ethridge, F. G., Flores, R. M., & Harvey, M. D.), pp. 169–176. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H. -D., & Wing, S. L. 1992. Terrestrial Ecosystems Through Time. Chicago: University of Chicago Press.Google Scholar
Behrensmeyer, A. K., & Hook, R. W. 1992. Paleoenvironmental contexts and taphonomic modes. In Terrestrial Ecosystems Through Time (ed. Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A.et al.), pp. 15–136. Chicago: University of Chicago Press.Google Scholar
Bekker, A., Holland, H. D., Wang., P. -L., Rumble, D. III. et al. 2004. Dating the rise of atmospheric oxygen. Nature 427, 117–120.CrossRefGoogle ScholarPubMed
Belderson, R. H., Johnson, M. A., & Kenyon, N. H. 1982. Bedforms. In Offshore Tidal Sands (ed. Stride, A. H.), pp. 25–57. London: Chapman and Hall.CrossRefGoogle Scholar
Bengtson, S. 1999. Early Life on Earth: Nobel Symposium No. 84. New York: Columbia University Press.Google Scholar
Benito, G., Baker, V. R., & Gregory, K. J. 1998. Palaeohydrology and Environmental Change. Chichester: Wiley.Google Scholar
Benjamin, T. B. 1968. Gravity currents and related phenomena. Journal of Fluid Mechanics 31, 209–248.CrossRefGoogle Scholar
Benn, D. I. 1994. Fluted moraine and till genesis below a temperate valley glacier: Slettmarkbreen, Jotunheimen, southern Norway. Sedimentology 41, 279–292.CrossRefGoogle Scholar
Benn, D. I., & Evans, D. J. A. 1998. Glaciers and Glaciation. London: Arnold.Google Scholar
Benn, D. I., & Evans, D. J. A.2006. Subglacial megafoods: outrageous hypothesis or just outrageous? In Glacier Science and Environmental Change (ed. Knight, P. G.), pp. 42–50. Oxford: Blackwells.CrossRefGoogle Scholar
Bennett, R. H., Bryant, W. R., & Hulbert, M. H. 1991. Microstructure of Fine-Grained Sediments: From Mud to Shale. New York: Springer-Verlag.CrossRefGoogle Scholar
Bennett, S. J., & Best, J. L. 1995. Mean flow and turbulence structure over fixed, two-dimensional dunes: implications for sediment transport and dune stability. Sedimentology 42, 491–514.CrossRefGoogle Scholar
Bennett, S. J., & Best, J. L.1996. Mean flow and turbulence structure over fixed ripples and the ripple–dune transition. In Coherent Flow Structures in Open Channels (ed. Ashworth, P. J., Bennett, S. J., Best, J. L., & McLelland, S. J.), pp. 281–304. Chichester: Wiley.Google Scholar
Bennett, S. J., & Bridge, J. S. 1995. An experimental study of flow, bedload transport and bed topography under conditions of erosion and deposition and comparison with theoretical models. Sedimentology 42, 117–146.CrossRefGoogle Scholar
Bennett, S. J., Bridge, J. S., & Best, J. L. 1998. The fluid and sediment dynamics of upper-stage plane beds. Journal of Geophysical Research 103, 1239–1274.CrossRefGoogle Scholar
Bennett, S. J., & Simon, A. 2004. Riparian Vegetation and Fluvial Geomorphology. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Berelson, W. M., & Heron, S. D. 1985. Correlations between Holocene flood tidal delta and barrier island inlet sequences: Back Sound–Shackleford Banks, North Carolina. Sedimentology 32, 215–222.CrossRefGoogle Scholar
Berendsen, H. J. A., & Stouthamer, E. 2001. Paleogeographic Development of the Rhine–Meuse Delta, the Netherlands. Assen: Koninklijke Van Gorcum.Google Scholar
Berger, A., & Loutre, M.-F. 1991. Insolation values for the climate of the last 10 million years. Quaternary Science Reviews 10, 297–317.CrossRefGoogle Scholar
Bergman, K. M., & Snedden, J. W. 1999. Isolated Shallow Marine Sand Bodies: Sequence Stratigraphic Analysis and Sedimentological Interpretation. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Bernard, H. A., LeBlanc, R. J., & Major, C. F. Jr. 1962. Recent and Pleistocene geology of southwest Texas, field excursion 3. Geology of the Gulf Coast and Central Texas and Guidebook of Excursion, pp. 175–225. Houston: Houston Geological Society.Google Scholar
Berné, S., Auffret, J. P., & Walker, P. 1988. Internal structures of subtidal sandwaves revealed by high-resolution seismic reflection. Sedimentology 35, 5–20.CrossRefGoogle Scholar
Berné, S., Castaing, P., LeDrezen, E., & Lericolais, G. 1993. Morphology, internal structure and reversal of asymmetry of large subtidal dunes in the entrance to Gironde estuary (France). Journal of Sedimentary Petrology 63, 780–793.Google Scholar
Berné, S., Lericolais, G., Marsset, T., Bourillet, J., & Batist, M. 1998. Erosional offshore sand ridges and lowstand shorelines: examples from tide- and wave-dominated environments of France. Journal of Sedimentary Research 68, 540–555.CrossRefGoogle Scholar
Berner, R. A. 1980. Early Diagenesis: A Theoretical Approach. Princeton, New Jersey: Princeton University Press.Google Scholar
Berner, R. A. 1991. A model for atmospheric CO2 over Phanerozoic time. American Journal of Science 291, 339–376.CrossRefGoogle Scholar
Berner, R. A. 1994. GEOCARB II: A revised model of atmospheric CO2 over Phanerozoic time. American Journal of Science 294, 56–91.CrossRefGoogle Scholar
Berner, R. A.2001. The effect of the rise of land plants on atmospheric CO2 during the Paleozoic. In Plants Invade the Land (ed. Gensel, P. G. & Edwards, D.), pp. 173–178. New York: Columbia University Press.CrossRefGoogle Scholar
Berner, E. K., & Berner, R. A. 1996. Global Environment: Water, Air, and Geochemical Cycles. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Berner, R. A., & Holdren, G. R. Jr. 1977. Mechanisms of feldspar weathering: some observational evidence. Geology 5, 369–372.2.0.CO;2>CrossRefGoogle Scholar
Berner, R. A., & Holdren, G. R. Jr. 1979. Mechanism of feldspar weathering: II. Observations of feldspars from soils. Geochimica et Cosmochimica Acta 43, 1173–1186.CrossRefGoogle Scholar
Berner, R. A., Holdern, G. R. Jr., & Schott, J. 1985. Surface layers on dissolving silicates. Geochimica et Cosmochimica Acta 49, 1657–1658.CrossRefGoogle Scholar
Berner, R. A., & Kothavala, Z. 2001. GEOCARB III: a revised model of atmospheric CO2 over Phanerozoic time: American Journal of Science 301, 182–204.CrossRefGoogle Scholar
Best, J. L. 1992a. On the entrainment of sediment and initiation of bed defects: insights from recent developments within turbulent boundary layer research. Sedimentology 39, 797–811.CrossRefGoogle Scholar
Best, J. L. 1992b. Sedimentology and event timing of a catastrophic volcaniclastic mass flow, Volcan Hudson, Southern Chile. Bulletin of Volcanology 54, 299–318.CrossRefGoogle Scholar
Best, J. L.1993. On the interactions between turbulent flow structure, sediment transport and bedform development: some considerations from recent experimental research. In Turbulence: Perspectives on Flow and Sediment Transport (ed. Clifford, N. J., French, J. R., & Hardisty, J.), pp. 61–92. Chichester: Wiley.Google Scholar
Best, J. L.1996. The fluid dynamics of small-scale alluvial bedforms. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 67–125. Chichester: Wiley.Google Scholar
Best, J. L. 2005. The fluid dynamics of river dunes: a review and some future research directions, Journal of Geophysical Research, 110, F04S02, doi:10.1029/2004JF000218.CrossRefGoogle Scholar
Best, J. L., & Ashworth, P. J. 1997. Scour in large braided rivers and the recognition of sequence stratigraphic boundaries. Nature 387, 275–277.CrossRefGoogle Scholar
Bhattacharya, J. P., & Giosan, L. 2003. Wave-influenced deltas: geomorphological implications for facies reconstruction. Sedimentology 50, 187–210.CrossRefGoogle Scholar
Bhattacharya, J. P., & Walker, R. G. 1992. Deltas. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 157–177. Ottawa: Geological Association of Canada.Google Scholar
Birkeland, C. 1997. Life and Death of Coral Reefs. New York: Chapman and Hall.CrossRefGoogle Scholar
Birkeland, P. W. 1999. Soils and Geomorphology, 3rd edn. New York: Oxford University Press.Google Scholar
Bierkens, M. F. P., & Weerts, H. J. T. 1994. Application of indicator simulation to modeling the lithological properties of a complex confining layer. Geodema 62, 265–284.CrossRefGoogle Scholar
Bjørlykke, K. 2003. Compaction (consolidation) of sediments. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 161–168. Dordrecht: Kluwer Academic Publishers.Google Scholar
Bjørkum, P. A. 1996. How important is pressure in causing dissolution of quartz in sandstones?Journal of Sedimentary Petrology 66, 147–154.Google Scholar
Bjornsson, H. 1992. Jokulhlaups in Iceland: prediction, characteristics, and simulation. Annals of Glaciology 16, 95–106.CrossRefGoogle Scholar
Black, K. S., Paterson, D. M., & Cramp, A. 1998. Sedimentary Processes in the Intertidal Zone. London: Geological Society of London.Google Scholar
Black, K. S., Tolhurst, T. J., Paterson, D. M., & Hagerthey, S. E. 2002. Working with natural cohesive sediments. Journal of Hydraulic Engineering ASCE 128, 2–8.CrossRefGoogle Scholar
Black, M. 1933. The algal sediments of Andros Island, Bahamas. Philosophical Transactions of the Royal Society 122B, 165–191.Google Scholar
Blakey, R. C., Havholm, K. G., & Jones, L. S. 1996. Stratigraphic analysis of eolian interactions with marine and fluvial deposits, Middle Jurassic Page Sandstone and Carmel Formation, Colorado Plateau, USA. Journal of Sedimentary Research 66, 324–342.Google Scholar
Blatt, H. 1982. Sedimentary Petrology. San Francisco: W. H. Freeman.Google Scholar
Blatt, H., Middleton, G., & Murray, R. 1980. Origin of Sedimentary Rocks, 2nd edn. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Blum, M. D., & Price, D. M. 1998. Quaternary alluvial plain construction in response to glacio-eustatic and climatic controls, Texas Gulf Coastal Plain. In Relative Role of Eustasy, Climate and Tectonism on Continental Rocks (ed. Shanley, K. J. & McCabe, P. J.), pp. 31–48. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Blum, M. D., & Törnqvist, T. E. 2000. Fluvial responses to climate and sea-level change: a review and look forward. Sedimentology 47 (Supplement 1), 2–48.CrossRefGoogle Scholar
Boardman, R. S., Cheetham, A. H., & Rowell, A. J. 1987. Fossil Invertebrates. Oxford: Blackwell Scientific.Google Scholar
Boersma, J. R. 1970. Distinguishing features of wave-ripple cross-stratification and morphology. Unpublished Ph.D. thesis, Utrecht University, the Netherlands.
Boersma, J. R., & Terwindt, J. H. J. 1981. Neap–spring tide sequences of intertidal shoal deposits in a mesotidal estuary. Sedimentology 28, 151–170.CrossRefGoogle Scholar
Bohacs, K., & Suter, J. 1997. Sequence stratigraphic distribution of coaly rocks: fundamental controls and paralic examples. American Association of Petroleum Geologists Bulletin 81, 1612–1639.Google Scholar
Boles, J. R., & Franks, S. G. 1979. Clay diagenesis in Wilcox sandstones of southwest Texas: implications of smectite diagenesis on sandstone cementation. Journal of Sedimentary Petrology 49, 55–70.Google Scholar
Bond, G. C., & Lotti, R. 1995. Iceberg discharges into the North Atlantic on millennial time scales during the last glaciation. Science 267, 1005–1010.CrossRefGoogle ScholarPubMed
Bondevik, S., Mangerud, J., Dawson, S., Dawson, A., & Lohne, O. 2005. Evidence for three North Sea tsunamis at the Shetland Islands between 8000 and 1500 years ago. Quaternary Science Reviews 24, 1757–1775.CrossRefGoogle Scholar
Boothroyd, J. C. 1985. Tidal inlets and tidal deltas. In Coastal Sedimentary Environments (ed. Davis, R. A.), pp. 445–533. New York: Springer-Verlag.CrossRefGoogle Scholar
Boothroyd, J. C., Friedrich, N. E., & McGinn, S. R. 1985. Geology of microtidal coastal lagoons: Rhode Island. Marine Geology 63, 35–76.CrossRefGoogle Scholar
Borchert, H., & Muir, R. O. 1964. Salt Deposits; The Origin, Metamorphism and Deformation of Evaporites. Princeton, NJ: Van Nostrand.Google Scholar
Bosence, D. W. J. 1989. Biogenic carbonate production in Florida Bay. Bulletin of Marine Science 44, 419–433.Google Scholar
Bosence, D. W. J., & Allison, P. A. 1995. Marine Paleoenvironmental Analysis from Fossils. London: Geological Society of London.Google Scholar
Bosence, D. W. J., Rowlands, R. J., & Quine, M. L. 1985. Sedimentology and budget of a Recent carbonate mound, Florida Keys. Sedimentology 32, 3137–343.CrossRefGoogle Scholar
Bosscher, H., & Schlager, W. 1992. Computer simulation of reef growth. Sedimentology 39, 503–512.CrossRefGoogle Scholar
Bossellini, A. 1984. Progradation geometries of carbonate platforms: examples from the Triassic of the Dolomites, northern Italy. Sedimentology 31, 1–24.CrossRefGoogle Scholar
Boucot, A. J., & Gray, J., 2001. A critique of Phanerozoic climatic models involving changes in the CO2 content of the atmosphere. Earth-Science Reviews 56, 1–159.CrossRefGoogle Scholar
Boulton, G. S. 1972. Modern Arctic glaciers as depositional models for former ice sheets. Quarterly Journal of the Geological Society of London 128, 361–393.CrossRefGoogle Scholar
Boulton, G. S.1974. Processes and patterns of glacial erosion. In Glacial Geomorphology (ed. Coates, D. R.), pp. 41–87. Binghamton, NY: Publications in Geomorphology.Google Scholar
Boulton, G. S. 1976. The origin of glacially fluted surfaces: observations and theory. Journal of Glaciology 17, 287–309.CrossRefGoogle Scholar
Boulton, G. S. 1979. Processes of glacier erosion on different substrata. Journal of Glaciology 23, 15–37.CrossRefGoogle Scholar
Boulton, G. S.1987. A theory of drumlin formation by subglacial sediment deformation. In Drumlin Symposium (ed. Menzies, J. & Rose, J.), pp. 25–80. Rotterdam: Balkema.Google Scholar
Boulton, G. S.1990. Sedimentary and sea level changes during glacial cycles and their control on glacimarine facies architecture. In Glacimarine Environments (ed. Dowdeswell, J. A. & Scourse, J. D.), pp. 15–52. London: Geological Society of London.Google Scholar
Boulton, G. S. 1996a. Theory of glacial erosion, transport and deposition as a consequence of subglacial sediment deformation. Journal of Glaciology 42, 43–62.CrossRefGoogle Scholar
Boulton, G. S. 1996b. The origin of till sequences by subglacial sediment deformation beneath mid-latitude ice sheets. Annals of Glaciology 22, 75–84.CrossRefGoogle Scholar
Boulton, G. S.2006. Glaciers and their coupling with hydraulic and sedimentary processes. In Glacier Science and Environmental Change (ed. Knight, P. G.), pp. 3–22. Oxford: Blackwells.CrossRefGoogle Scholar
Boulton, G. S., & Hindmarsh, R. C. A. 1987. Sediment deformation beneath glaciers: rheology and geological consequences. Journal of Geophysical Research 92B, 9059–9082.CrossRefGoogle Scholar
Boulton, G. S., Dobbie, K. E., & Zatsepin, S. 2001. Sediment deformation beneath glaciers and its coupling to the subglacial hydraulic system. Quaternary International 86, 3–28.CrossRefGoogle Scholar
Bouma, A. H. 1962. Sedimentology of Some Flysch Deposits. Amsterdam: Elsevier.Google Scholar
Bouma, A. H., Coleman, J. R., Stetling, C. E., & Kohl, B. 1989. Influence of relative sea level on the construction of the Mississippi Fan. Geo-Marine Letters 9, 161–170.CrossRefGoogle Scholar
Bouma, A. H., Normark, W. R., & Barnes, N. E. 1985a. Submarine Fans and Related Turbidite Systems. New York: Springer-Verlag.CrossRefGoogle Scholar
Bouma, A. H., Stetling, C. E., & Coleman, J. M. 1985b. Mississippi Fan, Gulf of Mexico. In Submarine Fans and Related Turbidite Systems (ed. Bouma, A. H., Normark, W. R., & Barnes, N. E.), pp. 143–150. New York: Springer-Verlag.CrossRefGoogle Scholar
Bourgeois, J., Hansen, T. A., Wiberg, P. L., & Kauffman, E. G. 1988. A tsunami deposit at the Cretaceous–Tertiary boundary in Texas. Science 241, 567–570.CrossRefGoogle ScholarPubMed
Bowden, K. F. 1983. Physical Oceanography of Coastal Waters. Chichester: Ellis Horwood.Google Scholar
Bowen, A. J. 1969. Rip currents, 1: theoretical investigations. Journal of Geophysical Research 74, 5467–5478.CrossRefGoogle Scholar
Bowen, A. J., & Inman, D. L. 1969. Rip currents, 2: laboratory and field observations. Journal of Geophysical Research 74, 5479–5490.CrossRefGoogle Scholar
Bowen, A. J., Normark, W. R., & Piper, D. J. W. 1984. Modelling of turbidity currents of Navy Submarine Fan, California Continental Borderland. Sedimentology 31, 169–185.CrossRefGoogle Scholar
Boyce, J. I., & Eyles, N. 1991. Drumlins carved by deforming till streams below the Laurentide Ice Sheet. Geology 19, 787–790.2.3.CO;2>CrossRefGoogle Scholar
Boyd, R., Dalrymple, R. W., & Zaitlin, B. A. 1992. Classification of clastic coastal depositional environments. Sedimentary Geology 80, 139–150.CrossRefGoogle Scholar
Boyd, R., Suter, J., & Penland, S. 1989. Sequence stratigraphy of the Mississippi delta. Transactions of the Gulf Coast Association of Geological Societies 39, 331–340.Google Scholar
Bradley, R. S. 1999. Paleoclimatology. San Diego, CA: Academic Press.Google Scholar
Brady, N. C., & Weil, R. R. 2002. The Nature and Properties of Soils, 13th edn. Upper Saddle River, NJ: Prentice-Hall.Google Scholar
Braitsch, O. 1971. Salt Deposits; Their Origin and Composition. New York: Springer-Verlag.CrossRefGoogle Scholar
Branney, M., & Kokelaar, P. 2003. Pyroclastic Density Currents and the Sedimentation of Ignimbrites. London: Geological Society of London.Google Scholar
Brayshaw, A. C., Davies, G. W., & Corbett, P. W. M. 1996. Depositional controls on primary permeability and porosity at the bedform scale in fluvial reservoir sandstones. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 374–394. Chichester: Wiley.Google Scholar
Brett, C. E., & Baird, G. C. 1985. Carbonate-shale cycles in the Middle Devonian of New York: an evaluation of models for the origin of limestones in terrigenous shelf sequences. Geology 13, 324–327.2.0.CO;2>CrossRefGoogle Scholar
Brewer, R. 1976. Fabric and Mineral Analysis of Soils, 2nd edn. New York: Krieger.Google Scholar
Bricker, O. P. 1971. Carbonate Cements. Baltimore, MA: The Johns Hopkins University Press.Google Scholar
Bridge, J. S. 1978. Paleohydraulic interpretation using mathematical models of contemporary flow and sedimentation in meandering channels. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 723–742. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Bridge, J. S. 1984. Large-scale facies sequences in alluvial overbank environments. Journal of Sedimentary Petrology 54, 583–588.Google Scholar
Bridge, J. S. 1985. Paleochannel patterns inferred from alluvial deposits: a critical evaluation. Journal of Sedimentary Petrology 55, 579–589.Google Scholar
Bridge, J. S. 1992. A revised model for water flow, sediment transport, bed topography and grain size sorting in natural river bends. Water Resources Research 28, 999–1023.CrossRefGoogle Scholar
Bridge, J. S.1993a. The interaction between channel geometry, water flow, sediment transport and deposition in braided rivers. In Braided Rivers (ed. Best, J. L. & Bristow, C. S.), pp. 13–72. London: Geological Society of London.Google Scholar
Bridge, J. S. 1993b. Description and interpretation of fluvial deposits: a critical perspective. Sedimentology 40, 801–810.CrossRefGoogle Scholar
Bridge, J. S. 1997. Thickness of sets of cross strata and planar strata as a function of formative bed-wave geometry and migration, and aggradation rate. Geology 25, 971–974.2.3.CO;2>CrossRefGoogle Scholar
Bridge, J. S.1999. Alluvial architecture of the Mississippi Valley: predictions using a 3D simulation model. In Floodplains: Interdisciplinary Approaches (ed. Marriott, S. B. & Alexander, J.), pp. 269–278. London: Geological Society of London.Google Scholar
Bridge, J. S.2000. The geometry, flow and sedimentary processes of Catskill rivers and coasts. In New Perspectives on the Old Red Sandstone (ed. Friend, P. F. & Williams, B. P. J.), pp. 85–108. London: Geological Society of London.Google Scholar
Bridge, J. S. 2003. Rivers and Floodplains: Forms, Processes, and Sedimentary Record. Oxford: Blackwells.Google Scholar
Bridge, J. S. 2006. Fluvial Facies Models. In Facies Models Revisited (ed. Posamentier, H. W. and Walker, R. G.), pp. 85–170. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Bridge, J. S.2007. Numerical modelling of alluvial deposits: recent developments. In Analogue and Numerical Forward Modelling of Sedimentary Systems; from Understanding to Prediction (ed. Boer, P. L., Postima, G., Kukla, P., Zwan, K. van des, & Burgess, P.). Oxford: Blackwells.Google Scholar
Bridge, J. S., & Bennett, S. J. 1992. A model for the entrainment and transport of sediment grains of mixed sizes, shapes and densities. Water Resource Research 28, 337–363.CrossRefGoogle Scholar
Bridge, J. S., & Best, J. L. 1988. Flow, sediment transport and bedform dynamics over the transition from upper-stage plane beds: implications for the formation of planar laminae. Sedimentology 35, 753–763.CrossRefGoogle Scholar
Bridge, J. S., & Gabel, S. L. 1992. Flow and sediment dynamics in a low sinuosity, braided river: Calamus River, Nebraska Sandhills. Sedimentology, 39, 125–142.CrossRefGoogle Scholar
Bridge, J. S. & Karssenberg, D. 2005. Simulation of flow and sedimentary processes, including channel bifurcation and avulsion, on alluvial fans. 8th International Conference on Fluvial Sedimentology, Delft. The Netherlands, August 7–12, 2005.Google Scholar
Bridge, J. S., & Leeder, M. R. 1979. A simulation model of alluvial stratigraphy. Sedimentology 26, 617–644.CrossRefGoogle Scholar
Bridge, J. S., & Lunt, I. A. 2006. Depositional models of braided rivers. In Braided Rivers II (ed. Smith, G. H. Sambrook, Best, J. L., Bristow, C. S., & Petts, G.), pp. 11–50. Oxford: Blackwells.CrossRefGoogle Scholar
Bridge, J. S., & Mackey, S. D. 1993a. A theoretical study of fluvial sandstone body dimensions. In Geological Modeling of Hydrocarbon Reservoirs (ed. Flint, S. S. & Bryant, I. D.), pp. 213–236. Oxford: Blackwells.Google Scholar
Bridge, J. S., & Mackey, S. D.1993b. A revised alluvial stratigraphy model. In Alluvial Sedimentation (ed. Marzo, M. & Puidefabregas, C.), pp. 319–337. Oxford: Blackwells.CrossRefGoogle Scholar
Bridge, J. S., & Tye, R. S. 2000. Interpreting the dimensions of ancient fluvial channel bars, channels, and channel belts from wireline-logs and cores. American Association of Petroleum Geologists Bulletin 84, 1205–1228.Google Scholar
Bridge, J. S., & Willis, B. J. 1994. Marine transgressions and regressions recorded in Middle Devonian shore-zone deposits of the Catskill clastic wedge. Geological Society of America Bulletin 106, 1440–1458.2.3.CO;2>CrossRefGoogle Scholar
Bridges, P. H., & Leeder, M. R. 1976. Sedimentary model for intertidal mudflat channels with examples from the Solway Firth, Scotland. Sedimentology 23, 533–552.CrossRefGoogle Scholar
Brierley, G. J., Ferguson, R. J., & Woolfe, K. J. 1997. What is a fluvial levee?Sedimentary Geology 114, 1–9.CrossRefGoogle Scholar
Bristow, C. S., Bailey, S. D., & Lancaster, N. 2000. The sedimentary structures of linear sand dunes. Nature 406, 56–59.CrossRefGoogle ScholarPubMed
Bristow, C. S., Lancaster, N., & Duller, G. A. T. 2005. Combining ground penetrating radar surveys and optical dating to determine dune migration in Namibia. Journal of the Geological Society of London 162, 315–321.CrossRefGoogle Scholar
Bristow, C. S., Pugh, J., & Goodall, T. 1996. Internal structure of aeolian dunes in Abu Dhabi determined using ground-penetrating radar. Sedimentology 43, 995–1003.CrossRefGoogle Scholar
Bristow, C. S., Skelly, R. L., & Ethridge, F. G. 1999. Crevasse splays from the rapidly aggrading, sand-bed, braided Niobrara River, Nebraska: effect of base-level rise. Sedimentology, 46, 1029–1047.CrossRefGoogle Scholar
Brodzikowski, K., & Loon, A. J. 1991. Glacigenic Sediments. Amsterdam: Elsevier.Google Scholar
Broecker, W. A., & Denton, G. H. 1989. The role of ocean–atmosphere reorganizations in glacial cycles. Geochimica et Cosmochimica Acta 53, 2465–2501.CrossRefGoogle Scholar
Broecker, W. A., & Takahashi, T. 1966. Calcium carbonate production on the Bahama Bank. Journal of Geophysical Research 71, 1575–1602.CrossRefGoogle Scholar
Bromley, R. G. 1996. Trace Fossils: Biology, Taphonomy and Applications, 2nd edn. London: Chapman and Hall.CrossRefGoogle Scholar
Brookfield, M. E. 1977. The origin of bounding surfaces in ancient aeolian sandstones. Sedimentology 24, 303–332.CrossRefGoogle Scholar
Brookfield, M. E.1992. Eolian Systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 143–156. St. John's Newfoundland: Geological Association of Canada.Google Scholar
Brookfield, M. E., & Ahlbrandt, T. S. 1983. Eolian Sediments and Processes. Amsterdam: Elsevier.Google Scholar
Brown, T. M., & Kraus, M. J. 1987. Integration of channel floodplain suite, I. Developmental sequence and lateral relations of alluvial paleosols. Journal of Sedimentary Petrology 57, 587–601.Google Scholar
Brown, T. M., & Ratcliffe, B. C. 1988. The origin of Chubutolithes lhering, ichnofossils for the Eocene and Oligocene of Chubut Province, Argentine. Journal of Paleontology 62, 163–167.Google Scholar
Brunsden, D., & Prior, D. B. 1984. Slope Instability. New York: Wiley.Google Scholar
Bryant, E. 2001. Tsunami: The Underrated Hazard. Cambridge: Cambridge University Press.Google Scholar
Bryant, I. D., & Flint, S. 1993. Quantitative clastic reservoir geological modeling: problems and perspectives. In Geological Modeling of Hydrocarbon Reservoirs (ed. Flint, S. & Bryant, I. D.), pp. 3–20. Oxford: Blackwells.Google Scholar
Buatois, L. A., Mangano, M. G., Genise, J. F., & Taylor, T. N. 1998. The ichnologic record of the continental invertebrate invasion: evolutionary trends in environmental expansion, ecospace utilization, and behavioral complexity. Palaios 13, 217–240.CrossRefGoogle Scholar
Bull, W. B. 1991. Geomorphic Responses to Climate Change. New York: Oxford University Press.Google Scholar
Bullock, P., Fedoroff, N., Jungerius, A., Stoops, G., & Tursina, T. 1988. Handbook of Soil Thin Section Description. Albrighton: Waine Research.Google Scholar
Burbank, D. W. 1992. Causes of recent Himalayan uplift deduced from deposited patterns in the Ganges basin. Nature 357, 680–683.CrossRefGoogle Scholar
Burbank, D. W., & Anderson, R. S. 2001. Tectonic Geomorphology. Oxford: Blackwells.Google Scholar
Burley, S. D. 2003. Cathodoluminescence (applied to the study of sedimentary rocks). In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 102–106. Dordrecht: Kluwer Academic Publishers.Google Scholar
Burley, S. D., Kantorowicz, J. D., & Waugh, B. 1985. Clastic diagenesis. In Sedimentology: Recent Developments and Applied Aspects (ed. Brenchley, P. J. & Williams, B. P. J.), pp. 189–228. London, Blackwell Scientific Publications.Google Scholar
Burley, S. D., & Worden, R. H. 2003. Sandstone Diagenesis: Recent and Ancient. Malden, MA: Blackwells.CrossRefGoogle Scholar
Burnett, W., & Riggs, S. R. 1990. Phosphate Deposits of the World, Volume 3, Neogene to Modern Phosphorites. Cambridge: Cambridge University Press.Google Scholar
Burst, J. F. 1965. Subaqueously formed shrinkage cracks in clay. Journal of Sedimentary Petrology 35, 348–353.CrossRefGoogle Scholar
Busby, C., & Ingersoll, R. V. 1995. Tectonics of Sedimentary Basins. Oxford: Blackwells.Google Scholar
Bustillo, M. A. 2003. Silcrete. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 659–660. Dordrecht: Kluwer Academic Publishers.Google Scholar
Butler, G. P. 1969. Modern evaporite deposition and geochemistry of coexisting brines, the sabkha, Trucial Coast, Arabian Gulf. Journal of Sedimentary Petrology 39, 70–89.Google Scholar
Campbell, C. S. 1990. Rapid granular flows. Annual Review of Fluid Mechanics 22, 57–92.CrossRefGoogle Scholar
Campbell, C. S. 2001. Granular flows in the elastic limit. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 83–89. Oxford: Blackwells.CrossRefGoogle Scholar
Campbell, N. A., Reece, J. B., & Mitchell, L. G. 1999. Biology, 5th edn. Park, Menlo, CA: Addison Wesley Longman.Google Scholar
Cao, S., & Knight, D. W. 1998. Design of hydraulic geometry of alluvial channels. Journal of Hydraulic Engineering, ASCE 124, 484–492.CrossRefGoogle Scholar
Carballo, J. D., Land, L. S., & Miser, D. E. 1987. Holocene dolomitization of supratidal sediments by active tidal pumping, Sugarloaf Key, Florida. Journal of Sedimentary Petrology 57, 153–165.Google Scholar
Carle, S. F., Labolle, E. M., Weissmann, G. S. et al. 1998. Conditional simulation of hydrofacies architecture: a transition probability/Markov approach. In Hydrogeologic Models of Sedimentary Aquifers (ed. Fraser, G. S. & Davis, J. M.), pp. 147–170. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Carling, P. A. 1981. Sediment transport by tidal currents and waves: observations from a sandy intertidal zone (Burry Inlet, South Wales). In Holocene Marine Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 65–80. Oxford: Blackwells.CrossRefGoogle Scholar
Carling, P. A. 1999. Subaqueous gravel dunes. Journal of Sedimentary Research 69, 534–545.CrossRefGoogle Scholar
Carling, P. A., & Dawson, M. R. 1996. Advances in Fluvial Dynamics and Stratigraphy. Chichester: Wiley.Google Scholar
Carling, P. A. & Shvidchenko, A. B. 2002. A consideration of the dune : antidune transition in fine gravel. Sedimentology, 49, 1,269–1,282.CrossRefGoogle Scholar
Carpenter, A. B. 1978. Origin and chemical evolution of brines in sedimentary basins. In Thirteenth Annual Forum on the Geology of Industrial Minerals (ed. Johnson, K. S. & Russell, J. A.), pp. 60–77. Tulsa, OK: Oklahoma Geological Survey.Google Scholar
Carson, M. A., & Kirkby, M. J. 1972. Hillslope Form and Process. Cambridge: Cambridge University Press.Google Scholar
Carter, D. J. T. 1982. Prediction of wave height and period for a constant wind velocity using the JONSWAP results. Ocean Engineering 9, 17–33.CrossRefGoogle Scholar
Carter, L., & McCave, I. N. 1997. The sedimentary regime beneath the deep Western Boundary Current inflow to the southwest Pacific Ocean. Journal of Sedimentary Research 67, 1005–1017.Google Scholar
Carter, R. M. 1975. A discussion and classification of subaqueous mass-transport with particular application to grain flow, slurry-flow, and fluxoturbidites. Earth-Science Reviews 11, 145–177.CrossRefGoogle Scholar
Carter, R. W. G., & Woodroffe, C. D. 1994. Coastal Evolution: Late Quaternary Shoreline Morphodynamics. Cambridge: Cambridge University Press.Google Scholar
Cas, R. A. F., & Wright, J. V. 1987. Volcanic Successions: Modern and Ancient. London: Allen & Unwin.CrossRefGoogle Scholar
Cathles, L. M. 1993. A discussion of flow mechanisms responsible for alteration and mineralization in the Cambrian aquifers of the Ouachita–Arkoma Basin–Ozark system. In Diagenesis and Basin Development (ed. Horbury, A. D. & Robinson, A. G.), pp. 99–112. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Cayeux, L. 1935. Les roches sédimentaires de France: roches carbonatées. Paris: Masson. [English translation by Carozzi, A. V. 1970. Sedimentary Rocks of France: Carbonate Rocks. Darien, CN: Hafner Publishing.]Google Scholar
Cerling, T. E. 1999. Stable carbon isotopes in paleosol carbonates. In Palaeoweathering, Palaeosurfaces and Related Continental Deposits (ed. Thiry, M. & Simon-Coicon, R.), pp. 43–60. Oxford: Blackwells.Google Scholar
Cerling, T. E., & Craig, H. 1994. Geomorphology and in-situ cosmogenic isotopes. Annual Review of Earth and Planetary Sciences 22, 273–317.CrossRefGoogle Scholar
Chafetz, H. S. 1986. Marine pisoids: a product of bacterially induced precipitation. Journal of Sedimentary Petrology 56, 812–817.Google Scholar
Chafetz, H. S., & Folk, R. L. 1984. Travertines, depositional morphology and the bacterially constructed constituents. Journal of Sedimentary Petrology 54, 289–316.Google Scholar
Chang, H. H. 1988. Fluvial Processes in River Engineering. New York: Wiley.Google Scholar
Chave, K. E. 1954. Aspects of the biogeochemistry of magnesium 1. Calcareous marine organisms. Journal of Geology 62, 266–283.CrossRefGoogle Scholar
Chave, K. E., Smith, S. V., & Roy, R. J. 1972. Carbonate production by coral reefs. Marine Geology 12, 123–140.CrossRefGoogle Scholar
Cheel, R. J., & Leckie, D. A. 1992. Coarse-grained storm beds of the Upper cretaceous Chungo Member (Wapiabi Formation), southern Alberta, Canada. Journal of Sedimentary Petrology 62, 933–945.Google Scholar
Chester, R. 2003. Marine Geochemistry, 2nd edn. Oxford: Blackwells.Google Scholar
Chilingarian, G. V., & Wolf, K. H. 1988. Diagenesis. Amsterdam: Elsevier.Google Scholar
Choquette, P. W., & Pray, L. C. 1970. Geological nomenclature and classification of porosity in sedimentary carbonates. American Association of Petroleum Geologists Bulletin 54, 207–250.Google Scholar
Christie-Blick, N., Sohl, L. E., & Kennedy, M. J. 1999. Considering a Neoproterozoic snowball Earth. Science 284, 1087.CrossRefGoogle Scholar
Chuhan, F. A., Bjørlykke, K., & Lowrey, C. J. 2000. The role of provenance in illitization of deeply buried reservoir sandstones from Haltenbanken and north Viking Graben, offshore Norway. Marine and Petroleum Geology 17, 673–689.CrossRefGoogle Scholar
Chuhan, F. A., Bjørlykke, K., & Lowrey, C. J. 2001. Closed-system burial diagenesis in reservoir sandstones: examples from the Garn Formation at Haltenbanken Area, offshore mid-Norway. Journal of Sedimentary Petrology 71, 15–26.CrossRefGoogle Scholar
Church, M. 1996. Channel morphology and typology. In River Flows and Channel Forms (ed. Petts, G. & Calow, P.), pp. 185–202. Oxford: Blackwell Science Limited.Google Scholar
Clark, C. D., & Meehan, R. T. 2001. Subglacial bedform morphology of the Irish Ice Sheet reveals major configuration changes during growth and decay. Journal of Quaternary Science 16, 483–496.CrossRefGoogle Scholar
Clark, J. D., & Pickering, K. T. 1996. Submarine Channels: Processes and Architecture. London: Vallis Press.Google Scholar
Clark, J. D., Kenyon, N. H., & Pickering, K. T. 1992. Quantitative analysis of the geometry of submarine channels: implications for the classification of submarine fans. Geology 20, 633–636.2.3.CO;2>CrossRefGoogle Scholar
Clarke, G. K. C., Collins, S. G., & Thompson, D. E. 1984. Flow, thermal structure, and subglacial conditions of a surge-type glacier. Canadian Journal of Earth Sciences 21, 232–240.CrossRefGoogle Scholar
Clarkson, E. N. K. 1998. Invertebrate Palaeontology and Evolution, 4th edn. Oxford: Blackwell Scientific.Google Scholar
Clemmensen, L. B. 1989. Preservation of interdraa and draa plinth deposits by the lateral migration of large linear draas (Lower Permian Yellow Sands, northeast England). Sedimentary Geology 65, 139–151.CrossRefGoogle Scholar
Clemmensen, L. B., Olsen, H., & Blakey, R. C. 1989. Erg-margin deposits in the Lower Jurassic Moenave Formation and Wingate Sandstone, southern Utah. Geological Society of America Bulletin 101, 759–773.2.3.CO;2>CrossRefGoogle Scholar
Clevis, Q., Boer, P., & Nijman, W. 2004a. Differentiating the effect of episodic tectonism and eustatic sea-level fluctuations in foreland basins filled by alluvial fans and axial deltaic systems: insights from a three-dimensional stratigraphic forward model. Sedimentology 51, 809–835.CrossRefGoogle Scholar
Clevis, Q., Boer, P., & Wachter, M. 2003. Numerical modeling of drainage basin evolution and three-dimensional alluvial fan stratigraphy. Sedimentary Geology 163, 85–110.CrossRefGoogle Scholar
Clevis, Q., Jager, G., Nijman, W., & Boer, P. L. 2004b. Stratigraphic signatures of translation of thrust-sheet top basins over low-angle detachment faults. Basin Research 16, 145–163.CrossRefGoogle Scholar
Clifton, H. E. 1969. Beach lamination: nature and origin. Marine Geology 7, 553–559.CrossRefGoogle Scholar
Clifton, H. E., & Dingler, J. R. 1984. Wave-formed structures and palaeoenvironmental reconstruction. Marine Geology 60, 165–198.CrossRefGoogle Scholar
Clifton, H. E., Hunter, R. E., & Phillips, R. L. 1971. Depositional structures and processes in the non-barred, high energy nearshore. Journal of Sedimentary Petrology 41, 651–670.Google Scholar
Clout, J. M. F., & Simonson, B. M. 2005. Precambrian iron formations and iron formation-hosted iron ore deposits. Economic Geology 100th Anniversary Issue, 643–679.Google Scholar
Cohen, A. S., 2003. Paleolimnology: The History and Evolution of Lake Systems. Oxford: Oxford University Press.Google Scholar
Cole, J. M., Rasbury, E. T., Montañez, I. P.et al. 2005. Petrographic and trace element analysis of uranium-rich tufa calcite, middle Miocene Barstow Formation, California, USA. Sedimentology 51, 433–453.CrossRefGoogle Scholar
Coles, D. F. 1956. The law of the wake in turbulent boundary layer. Journal of Fluid Mechanics 1, 191–226.CrossRefGoogle Scholar
Colella, A., & Prior, D. B. 1990. Coarse-Grained Deltas. Oxford: Blackwells.CrossRefGoogle Scholar
Coleman, J. M. 1969. Brahmaputra River: channel processes and sedimentation. Sedimentary Geology 3, 129–239.CrossRefGoogle Scholar
Coleman, J. M. 1988. Dynamic changes and processes in the Mississippi River delta. Geological Society of America Bulletin 100, 999–1015.2.3.CO;2>CrossRefGoogle Scholar
Coleman, J. M., & Gagliano, S. M. 1964. Cyclic sedimentation in the Mississippi River deltaic plain. Transactions of the Gulf Coast Association of Geological Societies 14, 67–80.Google Scholar
Coleman, J. M., & Gagliano, S. M.1965. Sedimentary structures: Mississippi River deltaic plain. In Primary Sedimentary Structures and Their Hydrodynamic Interpretation (ed. Middleton, G. V.), pp. 133–148. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Coleman, J. M., Gagliano, S. M., & Webb, J. E. 1964. Minor sedimentary structures in a prograding distributary. Marine Geology 1, 240–258.CrossRefGoogle Scholar
Coleman, J. M., & Prior, D. B. 1982. Deltaic environments. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D. R.), pp. 139–178. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Coleman, J. M., Prior, D. B., & Lindsay, J. F. 1983. Deltaic influences on shelfedge instability processes. In The Shelfbreak: Critical Interface on Continental Margins (ed. Stanley, D. J. & Moore, G. T.), pp. 121–137. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Coleman, J. M., & Wright, L. D. 1975. Modern river deltas: variability of processes and sand bodies. In Deltas, Models for Exploration (ed. Broussard, M. L.), pp. 99–149. Houston, TX: Houston Geological Society.Google Scholar
Collins, M. B., Amos, C. L., & Evans, G. 1981. Observations of some sediment transport processes over intertidal flats, The Wash, UK. In Holocene Marine Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 81–98. Oxford: Blackwells.CrossRefGoogle Scholar
Coniglio, M., & Dix, G. R. 1992. Carbonate slopes. In Facies Models, Response to Sea Level Changes (ed. Walker, R. G. & James, N. P.), pp. 349–373. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Cook, H. E., Hine, A. C., & Mullins, H. T. 1983. Platform Margin and Deep Water Carbonates. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Cook, H. E., & Mullins, H. T. 1983. Basin margin environments. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 540–617. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Cook, P. J., & Shergold, J. H. 1986. Phosphate Deposits of the World, Volume 1, Proterozoic and Cambrian Phosphorites. Cambridge: Cambridge University Press.Google Scholar
Cooke, R. U., & Warren, A. 1973. Geomorphology in Deserts. London: Batsford.Google Scholar
Cooke, R. U., Warren, A., & Goudie, A. 1993. Desert Geomorphology. London: UCL Press.Google Scholar
Corney, R. K. T., Peakall, J., Parsons, D. R.et al. 2006. The orientation of helical flow in curved channels. Sedimentology 53, 249–257.CrossRefGoogle Scholar
Costa, J. E., & Wieczorek, G. F. 1987. Debris Flows/Avalanches: Process, Recognition, and Mitigation. Boulder, CO: Geological Society of America.Google Scholar
Costa, J. E., & Williams, G. P. 1984. Debris Flow Dynamics. United States Geological Survey Open File Report 84–606 (VHS Videotape). Reston, VA: United States Geological Survey.Google Scholar
Cotter, E., & Link, J. E. 1993. Deposition and diagenesis of Clinton ironstones (Silurian) in the Appalachian Foreland basin of Pennsylvania. Geological Society of America Bulletin 105, 911–922.2.3.CO;2>CrossRefGoogle Scholar
Coulthard, T. J., Macklin, M. G., & Kirkby, M. J. 2002. A cellular model of Holocene upland river basin and alluvial fan evolution. Earth Surface Processes and Landforms 27, 269–288.CrossRefGoogle Scholar
Crabaugh, M., & Kocurek, G. 1993. Entrada sandstone: an example of a wet aeolian system. In The Dynamics and Environmental Context of Aeolian Sedimentary Systems (ed. Pye, K.), pp. 103–126. London: Geological Society of London.Google Scholar
Craft, J. H., & Bridge, J. S. 1987. Shallow-marine sedimentary processes in the Late Devonian Catskill Sea, New York State. Geological Society of America Bulletin 98, 338–355.2.0.CO;2>CrossRefGoogle Scholar
Crawford, D. A., & Mader, C. L. 1998. Modeling asteroid impact and tsunami. Science of Tsunami Hazards 16, 21–30.Google Scholar
Crevello, P. D., & Schlager, W. 1980. Carbonate debris sheets and turbidites, Exuma Sound, Bahamas. Journal of Sedimentary Petrology 50, 1121–1148.Google Scholar
Crick, R. E. 1986. Origin, Evolution, and Modern Aspects of Biomineralization in Plants and Animals. New York: Plenum Press.Google Scholar
Cross, T. A. 1990. Quantitative Dynamic Stratigraphy. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Crowley, T. J., & Berner, R. A. 2001. CO2 and climate change. Science 289, 270–277.CrossRefGoogle Scholar
Crowley, T. J., & North, G. R. 1991. Paleoclimatology. New York: Oxford University Press.Google Scholar
Csanady, G. T. 1978. Water circulation and dispersal mechanisms. In Lakes: Chemistry Geology and Physics (ed. Lerman, A.), pp. 21–64. New York: Springer-Verlag.CrossRefGoogle Scholar
Curray, J. R., Emmel, F. J., & Crampton, P. J. S. 1969. Holocene history of a strand plain, lagoonal coast, Narayit, Mexico. In Coastal Lagoons – A Symposium (ed. Castanares, A. A. & Phleger, F. B.), pp. 63–100. Mexico City: Universidad Nacional Autónoma.Google Scholar
Dade, W. B., & Huppert, H. E. 1994. Predicting the geometry of channelized deep-sea turbidites. Geology 22, 645–648.2.3.CO;2>CrossRefGoogle Scholar
Dade, W. B., Nowell, A. R. M., & Jumars, P. A. 1992. Predicting the erosion resistance of muds. Marine Geology 105, 285–297.CrossRefGoogle Scholar
Dalrymple, R. W. 1984. Morphology and internal structure of sand waves in the Bay of Fundy. Sedimentology 31, 365–382.CrossRefGoogle Scholar
Dalrymple, R. W.1992. Tidal depositional systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 195–218. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Dalrymple, R. W., Baker, E. K., Hughes, M., & Harris, P. T. 2003. Geomorphology and sedimentology of the muddy, tide-domonated, Fly River delta, Papua New Guinea. In Tropical Deltas of South-East Asia – Sedimentology, Stratigraphy, and Petroleum Geology (ed. Sidi, F. H., Nummedal, D., Imbert, P., Darman, H., & Posamentier, H.), pp. 147–174. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Dalrymple, R. W., Boyd, R., & Zaitlin, B. A. 1994. Incised Valley Systems: Origin and Sedimentary Sequences. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Dalrymple, R. W., Knight, R. J., Zaitlin, B. A., & Middleton, G. V. 1990. Dynamics and facies model of a macrotidal sand-bar complex, Cobequid Bay–Salmon River Estuary (Bay of Fundy). Sedimentology 37, 577–612.CrossRefGoogle Scholar
Dalrymple, R. W., Makino, Y., & Zaitlin, B. A. 1991. Temporal and spatial patterns of rhythmite deposition on mudflats in the macrotidal Cobequid Bay–Salmon River Estuary, Bay of Fundy. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 137–160. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Dalrymple, R. W., & Rhodes, R. N. 1995. Estuarine dunes and barforms. In Geomorphology and Sedimentology of Estuaries (ed. Perillo, G. M.), pp. 359–422. Amsterdam: Elsevier.Google Scholar
Dalrymple, R. W., Zaitlin, B. A., & Boyd, R. 1992. Estuarine facies models: Conceptual basis and stratigraphic implications. Journal of Sedimentary Petrology 62, 1130–1146.CrossRefGoogle Scholar
Damuth, J. E., & Embley, R. W. 1981. Mass-transport processes on Amazon cone: western equatorial Atlantic. American Association of Petroleum Geologists Bulletin 65, 629–643.Google Scholar
Damuth, J. E., Flood, R. D., Kowsmann, R. O., Belderson, R. H., & Gorini, M. A. 1988. Anatomy and growth pattern of Amazon deep-sea fans as revealed by long-range side-scan sonar (GLORIA) and high-resolution seismic studies. American Association of Petroleum Geologists Bulletin 72, 885–911.Google Scholar
Damuth, J. E., Flood, R. D., Pirmez, C., & Manley, P. L. 1995. Architectural elements and depositional processes of Amazon deep-sea fan imaged by sidescal sonar (GLORIA), bathymetric swath mapping (SeaBeam), high-resolution seismic, and piston-core data. In Atlas of Deep Water Environments: Architectural Style in Turbidite Systems (ed. Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, F. Ricci, & Smith, R. D. A.), pp. 105–121. London: Chapman and Hall.CrossRefGoogle Scholar
Damuth, J. E., Jacobi, R. D., & Hayes, D. E. 1983a. Sedimentary processes in Northwest Pacific Basin revealed by echo-character mapping studies. Geological Society of America Bulletin 94, 381–395.2.0.CO;2>CrossRefGoogle Scholar
Damuth, J. E., Kolla, V., Flood, R. D.et al. 1983b. Distributary channel meandering and bifurcation patterns on the Amazon deep-sea fan as revealed by long-range side-scan sonar (GLORIA). Geology 11, 94–98.2.0.CO;2>CrossRefGoogle Scholar
Damuth, J. E., Kowsmann, R. O., Flood, R. D., Belderson, R. H., & Gorini, M. A. 1983c. Age relationships of distributary channels on Amazon Deep-Sea fan: implications for fan growth pattern. Geology 11, 470–473.2.0.CO;2>CrossRefGoogle Scholar
Davidson, J. F., Harrison, D., & Carvalho, Guedes J. R. F. 1977. On the liquidlike behaviour of fluidized beds. Annual Review of Fluid Mechanics 9, 55–86.CrossRefGoogle Scholar
Davidson-Arnott, R. G. D., & Greenwood, B. 1974. Bedforms and structures associated with bar topography in the shallow water wave environment, Kouchibouguac Bay, New Brunswick, Canada. Journal of Sedimentary Petrology 44, 698–704.Google Scholar
Davies, G. R. 1970. Algal-laminated sediments, Gladstone Embayment, Shark Bay, Western Australia. In Carbonate Sedimentation and Environments, Shark Bay, Western Australia (ed. Logan, B. W., Davies, G. R., Read, J. F., & Cebulski, D. E.), pp. 169–205. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Davies, P. J., & Till, R. 1968. Stained dry cellulose peels of ancient and recent impregnated carbonate sediments. Journal of Sedimentary Petrology 38, 234–237.CrossRefGoogle Scholar
Davies, S., Hampson, G., Flint, S., & Elliott, T. 1999. Continent-scale sequence stratigraphy of the Namurian, Upper Carboniferous and its applications to reservoir prediction. In Petroleum Geology of Northwest Europe: Proceedings 5th Conference (ed. Fleet, A. J. & Boldy, A. A. R.), pp. 757–770. London: The Geological Society.Google Scholar
Davis, R. A. 1985. Coastal Sedimentary Environments. New York: Springer-Verlag.CrossRefGoogle Scholar
Davis, R. A., & Clifton, H. E. 1987. Sea-level change and the preservation of wave-dominated and tide-dominated coastal sequences. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 167–178. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Davis, R. A., & Fitzgerald, D. M. 2004. Beaches and Coasts. Oxford: Blackwells.Google Scholar
Davis, R. A., & Flemming, B. W. 1995. Stratigraphy of a combined wave- and tide-dominated intertidal sand body: Martens Plate, East Frisian Wadden Sea, Germany. In Tidal Signatures in Modern and Ancient Sediments (ed. Flemming, B. W. & Bartholoma, A.), pp. 121–132. Oxford: Blackwells Publication.CrossRefGoogle Scholar
Dawson, A. G. 1992. Ice Age Earth: Late Quaternary Geology and Climate. London: Routledge.Google Scholar
Dawson, A. G. 1994. Geomorphological effects of tsunami run-up and backwash. Geomorphology 10, 83–94.CrossRefGoogle Scholar
Dawson, A. G. 1999. Linking tsunami deposits, submarine slides and offshore earthquakes. Quaternary International 60, 119–126.CrossRefGoogle Scholar
Dawson, A. G., Foster, I. K. L., Shi, S., Smith, D. E., & Long, D. 1991. The identification of tsunami deposits in coastal sediment sequences. Science of Tsunami Hazards 9, 73–82.Google Scholar
Dean, W. E., & Fouch, T. D. 1983. Lacustrine. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 97–130. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Boer, P. L. 1979. Convolute lamination in modern sands of the estuary of the Oosterschelde, the Netherlands, formed as the result of entrapped air. Sedimentology 26, 283–294.CrossRefGoogle Scholar
Boer, P. L., & Smith, D. G. 1994. Orbital Forcing and Cyclic Sequences. Oxford: Blackwells.CrossRefGoogle Scholar
Boer, P. L., Gelder, A., & Nio, S. D. 1988. Tide-influenced Sedimentary Environments and Facies. Boston: D. Reidel Publishing Company.CrossRefGoogle Scholar
Decho, A. W., Visscher, P. T., & Reid, R. P. 2005. Production and cycling of natural microbial exopolymers (EPS) within a marine stromatolite. Palaeogeography, Palaeoclimatology, Palaeoecology, 219 71–86.CrossRefGoogle Scholar
Deer, W. A., Howie, R. A., & Zussman, J. 1962. Rock Forming Minerals, Volume 3: Sheet Silicates. New York: Wiley.Google Scholar
Defant, A. 1958. Ebb and Flow. Ann Arbor, MI: University of Michigan Press.Google Scholar
Deffeyes, K. S., Lucia, F. J., & Weyl, P. K. 1965. Dolomitization of Recent and Plio-Pleistocene sediments by marine evaporite waters on Bonaire, Netherlands Antilles. In Dolomitization and Limestone Diagenesis, A Symposium (ed. Pray, L. C. & Murray, R. C.), pp. 71–80. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Degens, E. T., & Ross, D. A. 1969. Hot Brines and Recent Heavy Metal Deposits in the Red Sea: A Geochemical and Geophysical Account. New York: Springer-Verlag.CrossRefGoogle Scholar
Demarest, J. M. II, & Kraft, J. C. 1987. Stratigraphic record of Quaternary sea levels: implications for more ancient strata. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 223–239. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Demicco, R. V. 1983. Wavy and lenticular-bedded carbonate ribbon rocks of the Upper Cambrian Conococheague Limestone, western Maryland. Journal of Sedimentary Petrology 53, 1121–1132.Google Scholar
Demicco, R. V. 1998. CYCOPATH 2D – a two-dimensional model of cyclic sedimentation on carbonate platforms. Computers & Geosciences 24, 405–423.CrossRefGoogle Scholar
Demicco, R. V., & Gierlowski-Kordesch, E. 1986. Facies sequences of a semi-arid closed basin: the Lower Jurassic East Berlin Formation of the Hartford Basin, New England, U.S.A. Sedimentology, 33, 107–118.CrossRefGoogle Scholar
Demicco, R. V., & Hardie, L. A. 1994. Sedimentary Structures and Early Diagentic Features of Shallow Marine Carbonate Deposits. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Demicco, R. V., & Hardie, L. A. 2002. The “carbonate factory” revisited: a reexamination of sediment production functions used to model deposition on carbonate platforms. Journal of Sedimentary Research 72, 849–857.CrossRefGoogle Scholar
Demicco, R. V., Lowenstein, T. K., Hardie, L. A., & Spencer, R. J. 2005. Model of seawater composition for the Phanerozoic. Geology 33, 877–880.CrossRefGoogle Scholar
Demicco, R. V., & Mitchell, R. W. III 1982. Facies of the Great American Bank in the central Appalachians. In Central Appalachian Geology, Field Trip Guidebook (ed. Lyttle, P. T.), pp. 171–266. Falls Church, VA: American Geological Institute.Google Scholar
Mowbray, T. 1983. The genesis of lateral accretion deposits in recent intertidal mudflat channels, Solway Firth, Scotland. Sedimentology 30, 425–435.CrossRefGoogle Scholar
Mowbray, T., & Visser, M. J. 1984. Reactivation surfaces in subtidal channel deposits, Oosterschelde, southwest Netherlands. Journal of Sedimentary Petrology 54, 811–824.Google Scholar
Raaf, J. F. M., Boersma, J. R., & Gelder, A. 1977. Wave-generated structures and sequences from a shallow marine succession, Lower Carboniferous, County Cork, Ireland. Sedimentology 24, 451–483.CrossRefGoogle Scholar
Deutsch, C. V. 2002. Geostatistical Reservoir Modeling. New York: Oxford University Press.Google Scholar
Deutsch, C., & Cockerham, P. 1994. Practical considerations in the application of simulated annealing to stochastic simulation. Mathematical Geology 26, 67–82.CrossRefGoogle Scholar
Deutsch, C. V., & Wang, L. 1996. Hierarchical object-based stochastic modeling of fluvial reservoirs. Mathematical Geology 28, 857–880.CrossRefGoogle Scholar
D'Hondt, S., Rutherford, S., & Spivak, A. J. 2002. Metabolic activity of subsurface life in deep-sea sediments. Science 295, 2067–2070.CrossRefGoogle ScholarPubMed
D'Hondt, S., Jørgensen, B. B., Miller, D. J.et al. 2004. Distributions of microbial activities in deep subseafloor sediments. Science 306, 2216–2221.CrossRefGoogle ScholarPubMed
Dickinson, W. R. 1976. Plate Tectonic Evolution of Sedimentary Basins. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Dickinson, W. R.1985. Interpreting provenance relations from detrital modes of sandstones. In Provenance of Arenites (ed. Zuffa, G. G.), pp. 333–361. Dordrecht: Reidel.CrossRefGoogle Scholar
Dickinson, W. R. 1995. Forearc basins. In Tectonics of Sedimentary Basins (ed. Busby, C. & Ingersoll, R.), pp. 221–262. Oxford: Blackwell Science.Google Scholar
Dickson, J. A. D. 1966. Carbonate identification and genesis as revealed by staining. Journal of Sedimentary Petrology 36, 491–505.Google Scholar
Dickson, J. A. D. 2002. Fossil echinoderms as monitor of the Mg/Ca ratio of Phanerozoic oceans. Science 298, 1222–1224.CrossRefGoogle ScholarPubMed
Diem, B. 1984. Analytical method for estimating paleowave climate and water depth from wave ripple marks. Sedimentology 32, 705–720.CrossRefGoogle Scholar
Dietrich, W. E., & Dunne, T. 1993. The channel head. In Channel Network Hydrology (ed. Beven, K. & Kirkby, M. J.), pp. 175–219. Chichester: Wiley.Google Scholar
Dill, R. F., Kendall, C. G.St., C., & Shinn, E. A. 1989. Giant Subtidal Stromatolites and Related Sedimentary Features, Lee Stocking Island, Exumas, Bahamas. IGC Field TripT 373, Washington, D.C.: American Geophysical Union.Google Scholar
Dill, R. F., Shinn, E. A., Jones, A. T., Kelly, K., & Steinen, R. P. 1986. Giant subtidal stromatolites forming in normal salinity water. Nature 324, 55–58.CrossRefGoogle Scholar
Dingler, J. R. 1979. The threshold of grain motion under oscillatory flow in a laboratory wave channel. Journal of Sedimentary Petrology 49, 287–294.Google Scholar
Domack, E. W., & Ishman, S. 1993. Oceanographic and physiographic controls on modern sedimentation within Antarctic fjords. Geological Society of America Bulletin 105, 1175–1189.2.3.CO;2>CrossRefGoogle Scholar
Dominguez, J. M. L. 1996. The São Francisco strandplain: a paradigm for wave-dominated deltas? In Geology of Siliciclastic Shelf Seas (ed. Baptist, M. & Jacobs, P.), pp. 217–231. London: Geological Society of London.Google Scholar
Dominguez, J. M. L., Martin, L., & Bittencourt, A. C. S. P. 1987. Sea-level history and Quaternary evolution of river mouth-associated beach-ridge plains along the east-southeast Brazilian coast: a summary. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 115–127. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRef
Donovan, R. N., & Foster, R. J. 1972. Subaqueous shrinkage cracks from the Caithness Flagstone series (Middle Devonian) of NE Scotland. Journal of Sedimentary Petrology 42, 309–317.Google Scholar
Donovan, S. K. 1994. Insects and other arthropods as trace-makers in non-marine environments and paleoenvironments. In The Paleobiology of Trace Fossils (ed. Donovan, S. K.), pp. 200–220. Baltimore, MA: The Johns Hopkins University Press.Google Scholar
Dott, R. H. Jr. 1992. Eustasy: the ups and downs of a major geological concept. Boulder, CO: Geological Society of America.
Dott, R. H., & Bourgeois, J. 1982. Hummocky stratification: significance of its variable bedding sequences. Geological Society of America Bulletin 93, 663–680.2.0.CO;2>CrossRefGoogle Scholar
Dove, P. M., DeYoreo, J. J., & Weiner, S. 2003. Biomineralization. Chantilly, VA: The Mineralogical Society of America.Google Scholar
Doveton, J. H. 1994. Theory and applications of vertical variability measures from Markov chain analysis. In Stochastic Modeling and Geostatistics (ed. Yarus, J. M. & Chambers, R. L.), pp. 55–64. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Dowdeswell, J. A., Elverhoi, A., Andrews, J. T., & Hebbeln, D. 1999. Asynchronous deposition of ice-rafted layers in the Nordic seas and North Atlantic Ocean. Nature 400, 348–351.CrossRefGoogle Scholar
Dowdeswell, J. A., Maslin, M. A., Andrews, J. T., & McCave, I. N. 1995. Iceberg production, debris rafting, and the extent and thickness of Heinrich layers (H-1, H-2) in North Atlantic sediments. Geology 23, 301–304.2.3.CO;2>CrossRefGoogle Scholar
Dowdeswell, J. A., & Cofaigh, O C. 2002. Glacier-influenced Sedimentation on High-latitude Continental Margins. London: Geological Society of London.Google Scholar
Dowdeswell, J. A., O Cofaigh, C., Taylor, J. et al. 2002. On the architecture of high-latitude continental margins: the influence of ice-sheet and sea-ice processes in the Polar North Atlantic. In Glacier-influenced Sedimentation on High-latitude Continental Margins (ed. Dowdeswell, J. A. & Cofaigh, C. O), pp. 33–54. London: Geological Society of London.Google Scholar
Dowdeswell, J. A., & Scourse, J. D. 1990. Glacimarine Environments. London: Geological Society of London.Google Scholar
Dowdeswell, J. A., Whittington, R. J., & Marienfeld, P. 1994. The origin of massive diamicton facies by iceberg rafting and scouring, Scoresby Sund, East Greenland. Sedimentology 41, 21–35.CrossRefGoogle Scholar
Dravis, J. J. 1979. Rapid and widespread generation of Recent oolitic hardgrounds on a high energy Bahamian Platform, Eleuthera Bank, Bahamas. Journal of Sedimentary Petrology 49, 195–208.Google Scholar
Dravis, J. J. 1983. Hardened subtidal stromatolites. Science 219, 385–386.CrossRefGoogle ScholarPubMed
Drever, J. J. 1994. The effect of land plants on weathering rates of silicate minerals. Geochimica et Cosmochimica Acta 58, 2325–2332.CrossRefGoogle Scholar
Drever, J. J. 1997. The Geochemistry of Natural Waters, 3rd edn. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Drever, J. J., & Stillings, L. L. 1997. The role of organic acids in mineral weathering. Colloids and Surfaces A: Physicochemical and Engineering Aspects 120, 167–181.CrossRefGoogle Scholar
Driemanis, A. 1988. Tills: their genetic terminology and classification. In Genetic Classification of Glacigenic Deposits (ed. Goldthwait, R. P. & Matsch, C. L.), pp. 17–83. Rotterdam: Balkema.Google Scholar
Driese, S. G., & Mora, C. I. 2001. Diversification of Siluro-Devonian plant traces in paleosols and influence on estimates of paleoatmospheric CO2 levels. In Plants Invade the Land (ed. Gensel, P G. & Edwards, D.), pp. 237–253. New York: Columbia University Press.CrossRefGoogle Scholar
Driese, S. G., Mora, C. I., & Elrick, J. M. 2000. The paleosol record of increasing plant diversity and depth of rooting and changes in atmospheric pCO2 in the Siluro-Devonian. In Phanerozoic Terrestrial Ecosystems: A Short Course (ed. Gastaldo, R. A. & DiMichele, W. A.), pp. 47–61. Lawrence, KA: The Paleontological Society.Google Scholar
Droser, M. L., & Bottjer, D. L. 1986. A semiquantitative field classification of ichnofabric. Journal of Sedimentary Petrology 56, 558–559.CrossRefGoogle Scholar
Droser, M. L., & Bottjer, D. L. 1989. Ichnofabric of sandstones deposited in high energy nearshore environments: measurement and utilization. Palaios 4, 598–604.CrossRefGoogle Scholar
Droz, L., & Bellaiche, G. 1985, Rhone deep-sea fan: morphostructure and growth pattern. American Association of Petroleum Geologists Bulletin 69, 460–479.Google Scholar
Droz, L., & Bellaiche, G.1991. Seismic facies and geological evolution of the central portion of the Indus fan. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 383–402. New York: Springer-Verlag.CrossRefGoogle Scholar
Duc, A. W., & Tye, R. S. 1987. Evolution and stratigraphy of a regressive/backbarrier complex: Kaiwah Island, South Carolina. Sedimentology 34, 237–251.CrossRefGoogle Scholar
Duke, W. L. 1985. Hummocky cross-stratification, tropical hurricanes, and intense winter storms. Sedimentology 32, 167–194.CrossRefGoogle Scholar
Duke, W. L. 1990. Geostrophic circulation or shallow marine turbidity currents? The dilemma of paleoflow patterns in storm-influenced prograding shoreline systems. Journal of Sedimentary Petrology 60, 870–883.CrossRefGoogle Scholar
Duke, W. L., Arnott, R. W., & Cheel, R. J. 1991. Shelf sandstones and hummocky cross-stratification: new insights on a stormy debate. Geology 19, 625–628.2.3.CO;2>CrossRefGoogle Scholar
Dumas, S., Arnott, R. W. C., & Southard, J. B. 2005. Experiments on oscillatory-flow and combined-flow bed forms: implications for interpreting parts of the shallow-marine sedimentary record. Journal of Sedimentary Research 75, 501–513.CrossRefGoogle Scholar
Dunham, R. J. 1969a. Early vadose silt in Townsend mound (reef), New Mexico. In Depositional Environments in Carbonate Rocky, a Symposium (ed. Friedman, G. M.), pp. 139–182. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Dunham, R. J.1969b. Vadose pisolite in the Capitan reef (Permian), New Mexico and Texas. In Depositional Environments in Carbonate Rocky, a Symposium (ed. Friedman, G. M.), pp. 182–190. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Dunlop, P., & Clarke, C. D. 2006. The morphological characteristics of ribbed moraine. Quaternary Science Reviews 25, 1668–1691.CrossRefGoogle Scholar
Dunne, T. 1980. Formation and controls of channel networks. Progress in Physical Geography 4, 211–239.CrossRefGoogle Scholar
Dury, G. H. 1985. Attainable standards of accuracy in the retrodiction of palaeodischarge from channel dimensions. Earth Surface Processes and Landforms 10, 205–213.CrossRefGoogle Scholar
Dutton, S. P., Flanders, W. A., & Barton, M. D. 2003. Reservoir characterization of a Permian deep-water sandstone, East Ford field, Delaware basin, Texas. American Association of Petroleum Geologists Bulletin 87, 609–627.CrossRefGoogle Scholar
Dutton, S. P., White, C. D., Willis, B. J., & Novakovic, D. 2002. Calcite cement distribution and its effect on fluid flow in a deltaic sandstone, Frontier Formation, Wyoming. American Association of Petroleum Geologists Bulletin 86, 2007–2021.Google Scholar
Dyer, K. R. 1989. Sediment processes in estuaries: future research requirements. Journal of Geophysical Research 94C, 14,327–14,339.CrossRefGoogle Scholar
Dyer, K. R., & Huntley, D. A. 1999. The origin, classification and modeling of sand banks and ridges. Continental Shelf Research 19, 1285–1330.CrossRefGoogle Scholar
Dzulyński, S., & Kotlarczyk, J. 1962. On load-casted ripples. Annales de la Société Géologique de Pologne 32, 148–159.Google Scholar
Dzulyński, S., & Walton, E. K. 1965. Sedimentary Features of Flysch and Graywackes. Amsterdam: Elsevier.Google Scholar
Easterbrook, D. J. 1999. Surface Processes and Landforms, 2nd edn. New York: Macmillan Publishing Company.Google Scholar
Edmonds, D. A., & Slingerland, R. 2007. Mechanics of middle ground bar formation: implications for the formation of delta distributary channel networks. Journal of Geophysical Research – Earth Surface 112, F02034, doi:10.1029/2006 JF000574.Google Scholar
Edwards, D. E., Leeder, M. R., Best, J. L., & Pantin, H. M. 1994. An experimental study of reflected density currents and the interpretation of certain turbidites. Sedimentology 41, 437–461.CrossRefGoogle Scholar
Ehrenberg, S. N., & Nadeau, P. H. 2005. Sandstone vs. carbonate petroleum reservoirs: a global perspective on porosity–depth and porosity–permeability relationships. American Association of Petroleum Geologists Bulletin 89, 435–445.CrossRefGoogle Scholar
Einsele, G. 1992. Sedimentary Basins, Evolution, Facies and Sediment Budget. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Einsele, G., Ricken, W., & Seilacher, A. 1991. Cycles and Events in Stratigraphy. Berlin: Springer-Verlag.Google Scholar
Eisma, D. 1997. Intertidal Deposits. Boca Raton, FL: CRC Press.Google Scholar
Ekdale, A. A., Bromley, R. G., & Pemberton, S. G. 1984. Ichnology: The Use of Trace Fossils in Sedimentology and Stratigraphy. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ekdale, A. A., & Pollard, J. 1991. Ichnofabric and ichnofacies. Palios 6, 197–343.CrossRefGoogle Scholar
Elfeki, A., & Dekking, M. 2001. A Markov chain model for subsurface characterization: theory and application. Mathematical Geology 33, 569–589.CrossRefGoogle Scholar
Elliott, T. 1974. Interdistributary bay sequences and their genesis. Sedimentology 21, 611–622.CrossRefGoogle Scholar
Elrick, M., & Snider, A. C. 2002. Deep-water stratigraphic cyclicity and carbonate mud mound development in the Middle Cambrian Marjum Formation, House Range, Utah. Sedimentology 49, 1021–1047.CrossRefGoogle Scholar
Embley, R. W. 1976. New evidence for occurrence of debris flow deposits in the deep sea. Geology 4, 371–374.2.0.CO;2>CrossRefGoogle Scholar
Embley, R. W. 1980. The role of mass transport in the distribution and character of deep-ocean sediments with special reference to the North Atlantic. Marine Geology 38, 23–50.CrossRefGoogle Scholar
Embry, A. F., & Klovan, J. E. 1971. A Late Devonian reef tract on north-eastern Banks Island, N.W.T. Bulletin of Canadian Petroleum Geology 19, 730–781.Google Scholar
Emery, D., & Myers, K. J. (eds.) 1996. Sequence Stratigraphy. Oxford: Blackwells.CrossRefGoogle Scholar
Emery, K. O. 1952. Continental shelf sediments off southern California. Geological Society of America Bulletin 63, 1105–1108.CrossRefGoogle Scholar
Enos, P. 1974. Surface Sediment Facies Map of the Florida–Bahamas Plateau. Boulder, CO: Geological Society of America.Google Scholar
Enos, P.1983. Shelf. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 267–295. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Enos, P. 1991. Sedimentary parameters for computer modeling. Kansas Geological Survey Bulletin 233, 63–99.Google Scholar
Enos, P., & Moore, C. H. 1983. Fore-reef slope. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 507–538. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Enos, P., & Perkins, R. D. 1977. Quaternary Sedimentation in South Florida. Boulder, CO: Geological Society of America.Google Scholar
Enos, P., & Perkins, R. D. 1979. Evolution of Florida Bay from island stratigraphy. Geological Society of America Bulletin 90, 59–83.2.0.CO;2>CrossRefGoogle Scholar
Enos, P., & Sawatsky, L. H. 1981. Pore networks in Holocene carbonate sediments. Journal of Sedimentary Petrology 51, 961–985.Google Scholar
Erickson, M. C., Masson, D. S., Slingerland, R., & Swetland, D. W. 1990. Numerical simulation of circulation and sediment transport in the Late Devonian Catskill Sea. In Quantitative Dynamic Stratigraphy (ed. Cross, T. A.), pp. 293–305. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Ernst, W. G. 2000. Earth Systems: Processes and Issues. Cambridge: Cambridge University Press.Google Scholar
Esteban, M. 1974. Caliche textures and “Microcodium.”Bollettino della Società Geologica Italiana 92, 105–125.Google Scholar
Esteban, M., & Klappa, C. F. 1983. Subaerial exposure. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 1–54. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Esteban, M., & Pray, L. C. 1983. Pisoids and pisolitic facies (Permian), Guadalupe Mountains, New Mexico and West Texas. In Coated Grains (ed. Peryt, T. M.), pp. 503–537. New York: Springer-Verlag.CrossRefGoogle Scholar
Ethridge, F. G., & Schumm, S. A. 1978. Reconstructing paleochannel morphologic and flow characteristics: methodology, limitations and assessment. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 703–721. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Ethridge, F. G., Skelly, R. L., & Bristow, C. S. 1999. Avulsion and crevassing in the sandy, braided Niobrara River: complex response to base-level rise and aggradation. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 179–191. Oxford: Blackwells.CrossRefGoogle Scholar
Eugster, H. P. 1969. Inorganic bedded cherts from the Magadi area, Kenya. Contributions to Mineralogy and Petrology 22, 1–31.CrossRefGoogle Scholar
Eugster, H. P., & Hardie, L. A. 1978. Saline lakes. In Lakes: Chemistry Geology Physics (ed. Lerman, A.), pp. 237–293. New York: Springer-Verlag.CrossRefGoogle Scholar
Evamy, B. D. 1962. The application of a chemical staining technique to a study of dedolomitization. Sedimentology 2, 164–170.CrossRefGoogle Scholar
Evans, D. J. A. 2003. Glacial Landsystems. London: Arnold.Google Scholar
Evans, D. J. A., Phillips, E. R., Hiemstra, J. F., & Auton, C. A. 2006a. Subglacial till: formation, sedimentary characteristics and classification. Earth-Science Reviews 78, 115–176.CrossRefGoogle Scholar
Evans, D. J. A., Rea, B. R., Hiemstra, J. F., & Cofaigh, O C. 2006b. A critical assessment of subglacial mega-floods: a case study of glacial sediments and landforms in south-central Alberta, Canada. Quaternary Science Reviews 25, 1638–1667.CrossRefGoogle Scholar
Evans, G. 1965. Intertidal flat sediments and their environments of deposition in the Wash. Quarterly Journal of the Geology Society of London 121, 209–245.CrossRefGoogle Scholar
Ewers, W. E., & Morris, R. C. 1981. Studies of the Dales Gorge Member of the Brockman Iron Formation, Western Australia. Economic Geology 76, 1929–1953.CrossRefGoogle Scholar
Eyles, C. H., Eyles, N., & Miall, A. D. 1985. Models of glaciomarine sedimentation and their application to the ancient glacial record. Palaeogeography, Palaeoclimatology, Palaeoecology 51, 15–84.CrossRefGoogle Scholar
Eyles, N., & Eyles, C. H. 1992. Glacial depositional systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 73–100. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Eyles, N., Sladen, J., & Gilroy, S. 1982. A depositional model for stratigraphic complexes and facies superimposition in lodgement tills. Boreas 11, 317–333.CrossRefGoogle Scholar
Faugeres, J.-C., Mezerais, M.-L., & Stow, D. A. V. 1993. Contourite drift types and their distribution in the North and South Atlantic ocean basins. Sedimentary Geology 82, 189–206.CrossRefGoogle Scholar
Farquhar, J., Bao, H., & Thiemens, M. 2000. Atmospheric influence of Earth's earliest sulfur cycle. Science 289, 756–758.CrossRefGoogle ScholarPubMed
Farrell, K. M. 1987. Sedimentology and facies architecture of overbank deposits of the Mississippi River, False River Region, Louisiana. In Recent Developments in Fluvial Sedimentology (ed. Ethridge, F. G., Flores, R. M., & Harvey, M. D.), pp. 111–120. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Farrell, K. M. 2001. Geomorphology, facies architecture, and high resolution, non-marine sequence stratigraphy in avulsion deposits, Cumberland marshes, Saskatchewan. Sedimentary Geology 139, 93–150.CrossRefGoogle Scholar
Faure, G. 1991. Principles and Applications of Inorganic Geochemistry, 2nd edn. New York: MacMillan Publishing Company.Google Scholar
Fawthrop, N. P. 1996. Modelling hydrological processes for river management. In River Flows and Channel Forms (ed. Petts, G. & Calow, P.), pp. 51–76. Oxford: Blackwell Science Ltd.Google Scholar
Felix, M. 2001. A two-dimensional numerical model for a turbidity current. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 71–81. Oxford: Blackwells.CrossRefGoogle Scholar
Field, M. E. 1980. Sand bodies on coastal plain shelves: Holocene record of the U.S. Atlantic shelf off Maryland. Journal of Sedimentary Petrology 50, 505–528.Google Scholar
Figueredo, A. G. J., Sanders, J. E., & Swift, D. J. P. 1982. Storm graded layers on inner continental shelves: examples from south Brazil and the Atlantic coast of the central United States. Sedimentary Geology 31, 171–190.CrossRefGoogle Scholar
Fischer, A. G. 1964. The Lofer cyclothems of the Alpine Triassic. Kansas State Geological Survey Bulletin 169, 7–149.Google Scholar
Fischer, A. G. 1982. Long-term climatic oscillations recorded in stratigraphy. In Climate in Earth History (ed. Geophysics Study Committee), pp. 97–104. Washington, D.C.: National Academy Press.Google Scholar
Fisher, R. V., & Schminke, H. U. 1984. Pyroclastic Rocks. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Fisher, R. V., & Smith, G. A. 1991. Sedimentation in Volcanic Settings. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Fisher, W. L., Brown, L. F., Scott, A. J., & McGowen, J. H. 1969. Delta Systems in the Exploration for Oil and Gas. Austin, TX: Bureau of Economic Geology, University of Texas.Google Scholar
Fisk, H. N. 1961. Bar finger sands of the Mississippi delta. In Geometry of Sandstone Bodies – A Symposium (ed. Peterson, J. A. & Osmond, J. C.), pp. 29–52. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Fisk, H. N., McFarlan, E. Jr., Kolb, C. R., & Wilbert, L. J. Jr. 1954. Sedimentary framework of the modern Mississippi delta. Journal of Sedimentary Petrology 24, 76–99.CrossRefGoogle Scholar
Fitzgerald, D. M. 1988. Shoreline erosional–depositional processes associated with tidal inlets. In Hydrodynamics and Sediment Dynamics of Tidal Inlets (ed. Aubrey, D. G. & Weishar, L.), pp. 186–225. Berlin: Springer-Verlag.Google Scholar
Fitzgerald, D. M.1993. Origin and stability of tidal inlets in Massachusetts. In Formation and Evolution of Multiple Tidal Inlets (ed. Aubrey, D. G. & Geise, F. J.), pp. 1–61. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Fitzgerald, M. G., Mitchum, R. M. Jr., Uliana, M. A., & Biddle, K. T. 1990. Evolution of San Jorge Basin. American Association of Petroleum Geologists Bulletin 74, 879–920.Google Scholar
Fitzpatrick, E. A. 1984. Micromorphology of Soils. London: Chapman and Hall.CrossRefGoogle Scholar
Flemming, B. W. 1980. Sand transport and bedform patterns on the continental shelf between Durban and Port Elizabeth, southeast African continental margin. Sedimentary Geology 26, 179–205.CrossRefGoogle Scholar
Flemming, B. W., & Bartholoma, A. 1995. Tidal Signatures in Modern and Ancient Sediments. Oxford: Blackwells.CrossRefGoogle Scholar
Flemings, P. B., & Jordan, T. E. 1989. A synthetic stratigraphic model of foreland basin development. Journal of Geophysical Research 94, 3851–3866.CrossRefGoogle Scholar
Flemings, P. B., & Jordan, T. E. 1990. Stratigraphic modeling of foreland basins: interpreting thrust deformation and lithospheric rheology. Geology 18, 430–434.2.3.CO;2>CrossRefGoogle Scholar
Flint, R. F. 1971. Glacial and Quaternary Geology. New York: Wiley.Google Scholar
Flint, S. S., Aitken, J., & Hampson, G. 1995. Application of sequence stratigraphy to coal-bearing coastal plain successions: implications for the UK coal measures. In European Coal Geology (ed. Whateley, M. K. G. & Spears, D. A.), pp. 1–16. London: Geological Society of London.Google Scholar
Flood, R. D., & Damuth, J. E. 1987. Quantitative characteristics of sinuous distributary channels on the Amazon deep-sea fan. Geological Society of America Bulletin 98, 728–738.2.0.CO;2>CrossRefGoogle Scholar
Flood, R. D., Manley, P. C., Kowsman, R. O., Appi, C. J., & Pirmez, C. 1991. Seismic facies and Late Quaternary growth of Amazon submarine fan. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 415–433. New York: Springer-Verlag.CrossRefGoogle Scholar
Flügel, E. 2004. Microfacies of Carbonate Rocks: Analysis, Interpretation and Application. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Folk, R. L. 1959. Practical petrographic classification of limestones. American Association of Petroleum Geologists Bulletin 43, 1–38.Google Scholar
Folk, R. L. 1965a. Petrology of Sedimentary Rocks. Austin, TX: Hemphil's.Google Scholar
Folk, R. L.1965b. Some aspects of recrystalization in ancient limestones. In Dolomitization and Limestone Diagenesis; A Symposium (ed. Pray, L. C. & Murray, R. C.), pp. 14–48. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Folk, R. L., & McBride, E. F. 1976. The Caballos Novaculite revisited. Part I: origin of the novaculite members. Journal of Sedimentary Petrology 46, 659–669.Google Scholar
Forristall, G. S., Hamilton, R. C., & Cardone, V. J. 1977. Continental shelf currents in Tropical Storm Delia: observation and theory. Journal of Physical Oceanography 7, 532–546.2.0.CO;2>CrossRefGoogle Scholar
Fortin, D., & Langley, S. 2005. Formation and occurrence of biogenic iron-rich minerals. Earth-Science Reviews 72, 1–19.CrossRefGoogle Scholar
Fouch, T. D., & Dean, W. E., 1982. Lacustrine environments. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D.), pp. 87–114. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Frakes, L. A. 1979. Climates Throughout Geologic History. Amsterdam: Elsevier.Google Scholar
Frakes, L. A., Francis, J. E., & Syktus, J. I. 1992. Climate Modes of the Phanerozoic. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Francis, J. E., & Smith, M. P. 2002. Paleoclimate reconstruction using fossils and lithologic indicators. Palaeogeography, Palaeoclimatology, Palaeoecology 182, 1–143.CrossRefGoogle Scholar
Francis, P. W. 1993. Volcanoes: A Planetary Perspective. Oxford: Oxford University Press.Google Scholar
Franseen, E. K., Watney, W. L., Kendall, C. G., C.St., & Ross, W. 1991. Sedimentary Modeling: Computer Simulations and Methods for Improved Parameter Definition. Lawrence, KA: Kansas Geological Survey.Google Scholar
Fraser, G. S., & DeCelles, P. G. 1992. Geomorphic controls on sediment accumulation at margins of foreland basins. Basin Research 4, 233–252.CrossRefGoogle Scholar
Frazier, D. E. 1967. Recent deltaic deposits of the Mississippi delta: their development and chronology. Transactions of the Gulf Coast Association of Geological Societies 17, 287–315.Google Scholar
Fredsoe, J., & Deigaard, R. 1992. Mechanics of Coastal Sediment Transport. Singapore: World Scientific.CrossRefGoogle Scholar
Freeze, R. A., & Cherry, J. A. 1979. Groundwater. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Frey, R. W., Howard, J. D., Han, S. J., & Park, B. K. 1989. Sediments and sedimentary sequences on a modern macrotidal flat, Inchon, Korea. Journal of Sedimentary Petrology 59, 28–44.Google Scholar
Friedman, G. M. 1959. Identification of carbonate minerals by staining methods. Journal of Sedimentary Petrology 29, 87–97.Google Scholar
Friedman, G. M., & Sanders, J. E. 1967. Origin and occurrence of dolostones. In Carbonate Rocks, Part A: Origin, Occurrence, and Classification (ed. Chilingar, G. V., Bissell, H. J., & Fairbridge, R. W.), pp. 167–348. Amsterdam: Elsevier.Google Scholar
Friedman, S. J. 1998. Rock-avalanche elements of the Shadow Valley Basin, Eastern Mojave Desert, California: processes and problems. Journal of Sedimentary Research 67, 792–804.Google Scholar
Frostick, L. E., & Reid, I. 1987. Desert Sediment: Ancient and Modern. London: Geological Society of London.Google Scholar
Fryberger, S. G. 1993. A review of aeolian bounding surfaces, with examples from the Permian Minnelusa Formation, USA. In Characterisation of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 167–197. London: Geological Society of London.Google Scholar
Fryberger, S. G., Ahlbrandt, T. S., & Andrews, S. 1979. Origin, sedimentary features and significance of low-angle eolian “sand sheet” deposits, Great Sand Dunes National Monument and vicinity, Colorado. Journal of Sedimentary Petrology 49, 733–746.CrossRefGoogle Scholar
Füchtbauer, H., & Hardie, L. A. 1976. Experimentally determined homogeneous distribution coefficients for precipitated magnesium calcites: application to marine carbonate cements. Geological Society of America Abstracts with Program 8, 877.Google Scholar
Füchtbauer, H., & Hardie, L. A. 1980. Comparison of experimental and natural magnesium calcites. IAS International Meeting, Bochum. Bochum: International Association of Sedimentologists. pp. 167–169.Google Scholar
Galloway, W. E. 1975. Process framework for describing the morphologic and stratigraphic evolution of deltaic depositional systems. In Deltas, Models for Exploration (ed. Broussard, M. L.), pp. 87–98. Houston, TX: Houston Geological Society.Google Scholar
Galloway, W. E.1981. Depositional architecture of Cenozoic Gulf coastal plain fluvial systems. In Recent and Ancient Nonmarine Depositional Environments: Models for Exploration (ed. Ethridge, F. G. & Flores, R. M.), pp. 127–155. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Galloway, W. E. 1989a. Genetic stratigraphic sequences in basin analysis I: architecture and genesis of flooding-suface bounded depositional units. American Association of Petroleum Geologists Bulletin 73, 125–142.Google Scholar
Galloway, W. E. 1989b. Genetic stratigraphic sequences in basin analysis II: application to northwest Gulf of Mexico Cenozoic basin. American Association of Petroleum Geologists Bulletin 73, 143–154.Google Scholar
Garcia-Castellanos, D. 2002. Interplay between lithospheric flexure and river transport in foreland basins. Basin Research 14, 89–104.CrossRefGoogle Scholar
Garrels, R. M., & MacKenzie, F. T. 1971. Evolution of Sedimentary Rocks. New York: W. W. Norton.Google Scholar
Garret, P. 1977. Biological communities and their sedimentary record. In Sedimentation on the Modern Carbonate Tidal Flats of Northwest Andros Island (ed. Hardie, L. A.), pp. 124–158. Baltimore, MA: The Johns Hopkins University Press.
Garven, G., Ge, S., Person, M. A., & Sverjensky, D. A. 1993. Genesis of stratabound ore deposits in the mid-continent of North America. 1. The role of regional groundwater flow. American Journal of Science 293, 497–568.CrossRefGoogle Scholar
Gawthorpe, R. L., & Leeder, M. R. 2000. Tectono-sedimentary evolution of active extensional basins. Basin Research 12, 195–218.CrossRefGoogle Scholar
Gaynor, G. C., & Swift, D. J. P. 1988. Shannon Sandstone depositional model; sand ridge formation on the Campanian western interior shelf. Journal of Sedimentary Petrology 58, 868–880.Google Scholar
Gebelein, C. D. 1974. Guidebook for Modern Bahamian Platform Environments. St. George, Bermuda: Geological Society of America.Google Scholar
Gebelein, C. D., Steinen, R. P., Garrett, P. et al. 1980. Subsurface dolomitization beneath the tidal flats of Central West Andros Island, Bahamas. In Concepts and Models of Dolomitization (ed. Zenger, D. H., Dunham, J. B., & Ethington, R. L.), pp. 31–49. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Genise, J. F., & Brown, T. M. 1994a. New Miocene scarabeid and hymenopterous nests and early Miocene (Santacrucian) paleoenvironments. Ichnos 3, 107–117.CrossRefGoogle Scholar
Genise, J. F., & Brown, T. M. 1994b. New trace fossils of termites (insecta: isoptera) from the late Eocene–early Miocene of Egypt, and the reconstruction of ancient isopteran behavior. Ichnos 3, 155–183.CrossRefGoogle Scholar
Genise, J. F., Mangano, M. G., Buatois, L. A., Laza, J. H., & Verde, M. 2000. Insect trace fossil associations in paleosols: the Coprinisphaera ichnofacies. Palaios 15, 49–64.2.0.CO;2>CrossRefGoogle Scholar
Gensel, P. G., & Edwards, D. 2001. Plants Invade the Land. New York: Columbia University Press.CrossRefGoogle Scholar
George, G., & Berry, J. K. 1993. A new lithostratigraphy and depositional model for the Upper Rotleigend of the UK sector of the southern North Sea. In Characterisation of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 291–320. London: Geological Society of London.Google Scholar
Ghibaudo, G. 1992. Subaqueous sediment gravity flow deposits: practical criteria for their field description and classification. Sedimentology 39, 423–454.CrossRefGoogle Scholar
Ghosh, B., & Lowe, D. R. 1993. The architecture of deep-water channel complexes, Cretaceous Venado Sandstone Member, Sacramento Valley, California. In Advances in the Sedimentary Geology of the Great Valley Group, Sacramento Valley, California (ed. Graham, S. A. & Lowe, D. R.), pp. 51–65. Los Angeles, CA: Pacific Section, Society of Economic Paleontologists and Mineralogists.Google Scholar
Gibling, M. R., & Bird, D. J. 1994. Late Carboniferous cyclothems and alluvial paleovalleys in the Sydney Basin, Nova Scotia. Geological Society of America Bulletin 106, 105–117.2.3.CO;2>CrossRefGoogle Scholar
Gilbert, G. K. 1885. The Topographic Features of Lake Shores. Washington, D.C.: United States Geological Survey.Google Scholar
Giles, M. R. 1997. Diagenesis: A Quantitative Perspective, Implications for Basin Modelling and Rock Property Prediction. Dordrecht: Kluwer Academic Publishers.Google Scholar
Giles, M. R., & Boer, R. B. 1990. Origin and significance of redistributional secondary porosity. Marine and Petroleum Geology 7, 378–397.CrossRefGoogle Scholar
Gill, W. D., & Kuenen, P. H. 1958. Sand volcanoes on slumps in the Carboniferous of County Clare, Ireland. Quarterly Journal of the Geology Society of London 113, 441–460.CrossRefGoogle Scholar
Ginsburg, R. N. 1967. Stromatolites. Science 157, 339–340.CrossRefGoogle ScholarPubMed
Ginsburg, R. N. 2001. Subsurface Geology of a Prograding Carbonate Platform Margin, Great Bahama Bank: Results of the Bahamas Drilling Project. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ginsburg, R. N., Isham, L. B., Bein, S. J., & Kuperburg, J. 1954. Laminated Algal Sediments of South Florida and Their Recognition in the Fossil Record. Coral Gables, FL: University of Miami Marine Laboratory.Google Scholar
Ginsburg, R. N., & James, N. P. 1974. Holocene carbonate sediments of continental margins. In The Geology of Continental Margins (ed. Burke, C. A. & Drake, C. L.), pp. 137–155. New York: Springer-Verlag.CrossRefGoogle Scholar
Giosan, L., & Bhattacharya, J. P. 2005. River Deltas – Concepts, Models, and Examples. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Glen, J. W. 1952. Experiments on the deformation of ice. Journal of Glaciology 2, 111–114.CrossRefGoogle Scholar
Glen, J. W. 1955. The creep of polycrystalline ice. Proceedings of the Royal Society of London 228A, 519–538.CrossRefGoogle Scholar
Glenn, C. R., & Garrison, R. E. 2003. Phosphorites. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 257–263. Dordrecht: Kluwer Academic Publisher.Google Scholar
Glennie, K. W. 1970. Desert Sedimentary Environments. Amsterdam: Elsevier.Google Scholar
Glennie, K. W.1986. Early Permian Rotleigend. In Introduction to the Petroleum Geology of the North Sea (ed. Glennie, K. W.), pp. 63–86. Oxford: Blackwells.Google Scholar
Goldhammer, R. K. 1997. Compaction and decompaction algorithms for sedimentary carbonates. Journal of Sedimentary Petrology 67, 26–35.CrossRefGoogle Scholar
Goldhammer, R. K., Dunn, P. A., & Hardie, L. A. 1987. High frequency glacio-eustatic sea level oscillations with Milankovitch characteristics recorded in Middle Triassic platform carbonates in northern Italy. American Journal of Science 287, 853–892.CrossRefGoogle Scholar
Goldhammer, R. K., Dunn, P. A., & Hardie, L. A. 1990. Depositional cycles, composite sea-level changes, cycle stacking patterns, and the hierarchy of stratigraphic forcing: examples from Alpine Triassic platform carbonates. Geological Society of America Bulletin 102, 535–562.2.3.CO;2>CrossRefGoogle Scholar
Goldhammer, R. K., & Harris, M. T. 1989. Eustatic controls on the stratigraphy and geometry of the Latemar buildup (Middle Triassic): the Dolomites of northern Italy. In Controls on Carbonate Platform and Basin Development (ed. Crevello, P. D., Wilson, J. L., Sarg, J. F., & Read, J. F.), pp. 323–338. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Goldrich, S. S. 1938. A study in rock weathering. Journal of Geology 46, 17–58.CrossRefGoogle Scholar
Goldsmith, J. R., & Graf, D. L. 1958. Relation between lattice constants and composition of the Ca–Mg carbonates. American Mineralogist 43, 84–101.Google Scholar
Goldsmith, J. R., Graf, D. L., & Joensuu, O. I. 1955. The occurrence of magnesium calcites in nature. Geochimica et Cosmochimica Acta 7, 212–230.CrossRefGoogle Scholar
Gosse, J. C., & Phillips, F. M. 2001. Terrestrial in situ cosmogenic nuclides: theory and applications. Quaternary Sciences Reviews 20, 1475–1560.CrossRefGoogle Scholar
Goudie, A. S. 1973. Duricrusts in Tropical and Subtropical Landscapes. Oxford: Clarendon Press.Google Scholar
Goudie, A. S.1999. Wind erosional landforms: yardangs and pans. In Aeolian Environments, Sediments and Landforms (ed. Goudie, A. S., Livingstone, I., & Stokes, S.), pp. 167–180. Chichester: Wiley.Google Scholar
Goudie, A. S., Livingstone, I., & Stokes, S. 1999. Aeolian Environments, Sediments and Landforms. Chichester: Wiley.Google Scholar
Gould, H. R. 1970. The Mississippi delta complex. In Deltaic Sedimentation Modern and Ancient (ed. Morgan, J. P. & Shaver, R. H.), pp. 3–30. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Graf, D. L., & Goldsmith, J. R. 1956. Some hydrothermal syntheses of dolomite and protodolomite. Journal of Geology 64, 173–186.CrossRefGoogle Scholar
Grant, W. D., & Madsen, O. S. 1979. Combined wave and current interaction with a rough bottom. Journal of Geophysical Research 84, 1797–1808.CrossRefGoogle Scholar
Grass, A. J. 1970. Initial instability of fine sand beds. Journal of the Hydraulic Division of the ASCE 96, 619–632.Google Scholar
Greeley, R., & Iverson, J. D. 1985. Wind as a Geological Process: On Earth, Mars, Venus and Titan. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Green, M. O., Rees, J. M., & Pearson, N. D. 1990. Evidence for the influence of wave–current interactions in a tidal boundary layer. Journal of Geophysical Research 95C, 9629–9644.CrossRefGoogle Scholar
Greenwood, B., & Mittler, P. R. 1985. Vertical sequence and lateral transitions in the facies of a barred nearshore environment. Journal of Sedimentary Petrology 55, 366–375.Google Scholar
Greenwood, B., & Sherman, D. J. 1986. Hummocky cross-stratification in the surf zone: flow parameters and bedding genesis. Sedimentology 33, 33–45.CrossRefGoogle Scholar
Greenwood, B., & Sherman, D. J. 1993. Waves, currents, sediment flux and morphological response in a barred nearshore system. Marine Geology 60, 31–61.CrossRefGoogle Scholar
Gregg, J. M., & Sibley, D. F. 1984. Epigenetic dolomitization and the origin of xenotopic dolomite textures. Journal of Sedimentary Petrology 54, 908–931.Google Scholar
Gregory, K. J. 1983. Background to Palaeohydrology. Chichester: Wiley.Google Scholar
Gregory, K. J., Lewin, J., & Thornes, J. B. 1987. Palaeohydrology in Practice. Chichester: Wiley.Google Scholar
Gregory, K. J., Starkel, L., & Baker, V. R. 1996. Continental Palaeohydrology. Chichester: Wiley.Google Scholar
Griffing, D. H., Bridge, J. S. & Hotton, C. L. 2000. Coastal–fluvial paleoenvironments and plant ecology of the early Devonian (Emsian), Gaspe Bay, Canada. In New Perspectives on the Old Red Sandstone (ed. Friend, P. F., & Williams, B. P. J.), pp. 61–84. London: Geological Society of London.Google Scholar
Griffiths, C. M., Dyt, C., Paraschivoiu, E., & Liu, K. 2001. SEDSIM in hydrocarbon exploration. In Geologic Modeling and Simulation: Sedimentary Systems (ed. Merriam, D. F. & Davis, J. C.), pp. 71–97. New York: Kluwer Academic/Plenum Publishers.CrossRefGoogle Scholar
Grotzinger, J. P., & James, N. P. 2000. Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Grotzinger, J. P., Jordan, T. H., Press, F., & Siever, R. 2007. Understanding Earth, 5th edn. New York: W. H. Freeman and Company.Google Scholar
Grotzinger, J. P., & Knoll, A. H. 1999. Stromatolites in Precambrian Carbonates: evolutionary mileposts or environmental dipsticks. Annual Review of Earth and Planetary Sciences 27, 313–358.CrossRefGoogle ScholarPubMed
Grover, G. Jr., & Read, J. F. 1978. Fenestral and associated vadose diagenetic fabrics of tidal flat carbonates, Middle Ordovician New Market Limestone, southwestern Virginia. Journal of Sedimentary Petrology 48, 453–473.Google Scholar
Grunau, H. R. 1965. Radiolarian cherts and associated rocks in space and time. Eclogae Geologicae Helvetiae 58, 157–208.Google Scholar
Gruszczynski, M., Rudowski, S., Semil, J., Slominski, J., & Zrobek, J. 1993. Rip currents as a geological tool. Sedimentology 40, 217–236.CrossRefGoogle Scholar
Gupta, S. 1997. Himalayan drainage patterns and the origin of fluvial megafans in the Ganges foreland basin. Geology 25, 11–14.2.3.CO;2>CrossRefGoogle Scholar
Gupta, S., & Cowie, P. 2000. Processes and controls in the stratigraphic development of extensional basins. Basin Research 12, 185–241.CrossRefGoogle Scholar
Gustard, A. 1996. Analysis of river regimes. In River Flows and Channel Forms (ed. Petts, G. & Calow, P.), pp. 32–50. Oxford: Blackwell Science Ltd.Google Scholar
Guza, R. T., & Bowen, A. J. 1975. The resonant instabilities of long waves obliquely incident on a beach. Journal of Geophysical Research 80, 4529–4534.CrossRefGoogle Scholar
Guza, R. T., & Davis, R. E. 1974. Excitation of edge waves by waves incident on a beach. Journal of Geophysical Research 79, 1285–1291.CrossRefGoogle Scholar
Haff, P. K. 1983. Grain flow as a fluid-mechanical phenomenon. Journal of Fluid Mechanics 134, 401–430.CrossRefGoogle Scholar
Hallet, B. 1996. Glacial quarrying: a simple theoretical model. Annals of Glaciology 22, 1–8.CrossRefGoogle Scholar
Halley, R. B., & Evans, C. C. 1983. The Miami Limestone. A Guide to Selected Outcrops and Their Interpretation (with a Discussion of Diagenesis in the Formation). Miami, FL: Miami Geological Society.Google Scholar
Halley, R. B., Harris, P. M., & Hine, A. C., 1983. Bank margin. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 463–506. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Hambrey, M. 1994. Glacial Environments. London: UCL Press.Google Scholar
Hambrey, M., & Harland, W. B. 1981. Earth's Pre-Pleistocene Glacial Record. Cambridge: Cambridge University Press.Google Scholar
Hampson, G. J., Elliott, T., & Davies, S. J. 1997. The application of sequence stratigraphy to Upper Carboniferous fluvio-deltaic strata of the onshore UK and Ireland: implications for the southern North Sea. Journal of the Geological Society of London 154, 719–733.CrossRefGoogle Scholar
Hampson, G. J., Davies, S. J., Elliott, T., Flint, S. S., & Stollhofen, H. 1999. Incised valley fill sandstone bodies in Upper Carboniferous fluvio-deltaic strata: recognition and reservoir characterization of Southern North Sea analogues. In Petroleum Geology of Northwest Europe (ed. Fleet, A. J. & Boldy, S. A. R.), pp. 771–788. London: The Geological Society.Google Scholar
Hampton, M. A. 1972. The role of subaqueous debris flow in generating turbidity currents. Journal of Sedimentary Petrology 42, 775–793.Google Scholar
Hampton, M. A. 1975. Competence of fine-grained debris flows. Journal of Sedimentary Petrology 45, 834–844.Google Scholar
Handford, C. F., Kendall, A. C., Prezbindowski, D. R., Dunham, J. B., & Logan, B. W. 1984. Salina-margin tepees, pisolites, and aragonite cements, Lake MacLeod, Western Australia: their significance in interpreting ancient analogs. Geology 12, 523–527.2.0.CO;2>CrossRefGoogle Scholar
Handford, C. R., Loucks, R. G., & Davies, G. R. 1985. Depositional and Diagenetic Spectra of Evaporites – A Core Workshop. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Hanna, S. R. 1969. The formation of longitudinal sand dunes by large helical eddies in the atmosphere. Journal of Applied Meteorology 8, 874–883.2.0.CO;2>CrossRefGoogle Scholar
Hannington, M. D., Jonasson, I. R., Herzig, P. M., & Petersen, S. 1995. Physical and chemical processes of seafloor mineralization at mid-ocean ridges. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions (ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S., & Thomson, R. E.), pp. 115–157. Washington: D.C.: American Geophysical Union.CrossRefGoogle Scholar
Hanor, J. S. 1994. Origin of saline fluids in sedimentary basins. In Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins (ed. Parnell, J.), pp. 151–174. London: Geological Society of London.Google Scholar
Hanor, J. S., & McIntosh, J. C. 2006. Are secular variations in seawater chemistry reflected in the compositions of basinal brines?Journal of Geochemical Exploration 89, 153–156.CrossRefGoogle Scholar
Häntschel, W. 1975. Trace Fossils and Problematica. Treatise on Invertebrate Paleontology, Part W Supplement 1. Boulder, CO: Geological Society of America; and Lawrence, KA: The University of Kansas.Google Scholar
Harbaugh, J. W., Watney, W. L., Rankey, E. C.et al. 1999. Numerical Experiments in Stratigraphy: Recent Advances in Stratigraphic and Sedimentologic Computer Simulations. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Harbor, J. M. 1992. Numerical modeling of the development of U-shaped valleys by glacial erosion. Geological Society of America Bulletin 104, 1364–1375.2.3.CO;2>CrossRefGoogle Scholar
Hardie, L. A. 1977. Sedimentation on the Modern Carbonate Tidal Flats of Northwest Andros Island, Bahamas. Baltimore, MA: The Johns Hopkins University Press.Google Scholar
Hardie, L. A. 1984. Evaporites: marine or nonmarine?American Journal of Science 284, 193–240.CrossRefGoogle Scholar
Hardie, L. A. 1987. Dolomitization: a critical view of some current views. Journal of Sedimentary Petrology 57, 166–183.CrossRefGoogle Scholar
Hardie, L. A. 1990. The roles of rifting and hydrothermal CaCl2 brines in the origin of potash evaporites: an hypothesis. American Journal of Science 290, 43–106.CrossRefGoogle Scholar
Hardie, L. A. 1991. On the significance of evaporites. Annual Review of Earth and Planetary Sciences 19, 131–168.CrossRefGoogle Scholar
Hardie, L. A. 1996. Secular variation in seawater chemistry: an explanation for the coupled secular variation in the mineralogies of marine limestones and potash evaporites over the past 600 m.y. Geology 34, 279–283.2.3.CO;2>CrossRefGoogle Scholar
Hardie, L. A.2003a Evaporites. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 257–263. Dordrecht: Kluwer Academic Publishing.Google Scholar
Hardie, L. A.2003b Sabkha, salt flat, salina. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 584–585. Dordrecht: Kluwer Academic Publishing.Google Scholar
Hardie, L. A., & Eugster, H. P. 1970. The evolution of closed basin brines. Mineralogical Society of America Special Paper 3, 273–290.Google Scholar
Hardie, L. A., & Lowenstein, T. K. 2004. Did the Mediterranean Sea dry out during the Miocene? A reassessment of the evaporite evidence from DSDP Legs 13 and 42 A cores. Journal of Sedimentary Research 74, 453–461.CrossRefGoogle Scholar
Hardie, L. A., Lowenstein, T. K., & Spencer, R. J. 1985. The problem of distinguishing primary from secondary evaporites. In Proceedings of the Sixth International Symposium on Salt (ed. Schreiber, B. C. & Harner, H. L.), pp. 11–39. Alexandria, VA: The Salt Institute.Google Scholar
Hardie, L. A., & Shinn, E. A. 1986. Carbonate depositional environments modern and ancient, part 3: tidal flats. Colorado School of Mines Quarterly 81, 1–74.Google Scholar
Hardie, L. A., Smoot, J. P., & Eugster, H. P. 1978. Saline lakes and their deposits, a sedimentological approach. In Modern and Ancient Lake Sediments (ed. Matter, A. & Tucker, M. E.), pp. 7–41. Oxford: Blackwells.CrossRefGoogle Scholar
Harms, J. C., Southard, J. B., & Walker, R. G. 1982. Structures and Sequences in Clastic Rocks. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Harris, N. B. 1989. Diagenetic quartzarenite and destruction of secondary porosity: an example from the Middle Jurassic Brent sandstone of northwest Europe. Geology 17, 361–364.2.3.CO;2>CrossRefGoogle Scholar
Harris, P. M. 1979. Facies Anatomy and Diagenesis of a Bahamian Ooid Shoal. Sedimenta 7. Miami, FL: Comparative Sedimentology Laboratory, University of Miami.Google Scholar
Harris, P. T., Pattiaratchi, C. B., Cole, A. R., & Keene, J. B. 1992. Evolution of subtidal sandbanks in Moreton bay, eastern Australia. Marine Geology 103, 225–247.CrossRefGoogle Scholar
Harris, P. T., Pattiaratchi, C. B., Collins, M. B., & Dalrymple, R. W. 1995. What is a bedload parting? In Tidal Signatures in Modern and Ancient Sediments (ed. Flemming, B. W. & Bartholoma, A.), pp. 3–18. Oxford: Blackwells.CrossRefGoogle Scholar
Harris, P. T., Pattiaratchi, C. B., Keene, J. B.et al. 1996. Late Quaternary deltaic and carbonate sedimentation in the Gulf of Papua foreland basin: response to sea-level change. Journal of Sedimentary Research B66, 801–819.Google Scholar
Harrison, W. J., & Summa, L. L. 1991. Paleohydrology of the Gulf of Mexico Basin. American Journal of Science 291, 109–176.CrossRefGoogle Scholar
Hart, B. S., & Plint, A. G. 1989. Gravelly shoreface deposits: a comparison of modern and ancient facies sequences. Sedimentology 36, 551–557.CrossRefGoogle Scholar
Hartley, A. J., & Prosser, D. J. 1995. Characterization of Deep Marine Clastic Systems. London: Geological Society of London.Google Scholar
Hartmann, D. L. 1994. Global Physical Climatology. San Diego, CA: Academic Press.Google Scholar
Harvie, C. E., Möller, N., & Weare, J. H. 1984. The prediction of mineral solubilities in natural waters: the Na–K–Mg–Ca–H–Cl–SO4–OH–HCO3–CO3–CO2–H2O system to high ionic strengths at 25 °C. Geochimica et Cosmochimica Acta 48, 723–752.CrossRefGoogle Scholar
Harvie, C. E., Weare, J. H., Hardie, L. A., & Eugster, H. P. 1980. Evaporation of sea water: calculated mineral sequences. Science 208, 498–500.Google Scholar
Hasiotis, S. T. 2002. Continental Trace Fossils. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google 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, 259–320.CrossRefGoogle Scholar
Hasiotis, S. T. 2004. Reconnaissance of Upper Jurassic Morrison Formation ichnofossils, Rocky Mountain Region, USA: paleoenvironmental, stratigraphic, and paleoclimatic significance of terrestrial and freshwater ichnocoenoses. Sedimentary Geology 167, 177–268.CrossRefGoogle Scholar
Hättestrand, C., & Kleman, J. 1999. Ribbed moraine formation. Quaternary Science Reviews 18, 43–61.CrossRefGoogle Scholar
Havholm, K. G., Blakey, R. C., Capps, M. et al. 1993. Aeolian genetic stratigraphy: an example from the Middle Jurassic Page Sandstone, Colorado Plateau. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 87–107. Oxford: Blackwells.CrossRefGoogle Scholar
Hawley, N. 1981. Flume experiments on the origin of flaser bedding. Sedimentology 28, 699–712.CrossRefGoogle Scholar
Hayes, M. O. 1967. Hurricanes as Geological Agents: Case Studies of Hurricanes Carla, 1961, and Cindy, 1963. Austin, TX: Bureau of Economic Geology.Google Scholar
Hayes, M. O.1975. Morphology of sand accumulation in estuaries: an introduction to the symposium. In Estuarine Research: Volume II Geology and Engineering (ed. Cronin, L E.), pp. 3–22. London: Academic Press.Google Scholar
Hayes, M. O.1979. Barrier island morphology as a function of tidal and wave regime. In Barrier Islands – From the Gulf of St Lawrence to the Gulf of Mexico (ed. Leatherman, S. P.), pp. 1–27. London: Academic Press.Google Scholar
Heckel, P. H., Gibling, M. R., & King, N. R. 1998. Stratigraphic model for glacial–eustatic Pennsylvanian cyclotherms in highstand nearshore detrital regimes. Journal of Geology 106, 373–383.CrossRefGoogle Scholar
Heezen, B. C., Ericson, D. B., & Ewing, M. 1954. Further evidence for a turbidity current following the 1929 Grand banks earthquake. Deep-Sea Research 1, 193–202.CrossRefGoogle Scholar
Heiken, G., & Wohletz, K. 1985. Volcanic Ash. Berkeley: University of California Press.Google Scholar
Heikoop, J. M., Tsujita, C. J., Risk, M. J., Tomascik, T., & Mah, A. J. 1996. Modern iron ooids from a shallow-marine volcanic setting: Mahengetang, Indonesia. Geology 24, 759–762.2.3.CO;2>CrossRefGoogle Scholar
Heller, P. L., & Paola, C. 1992. The large scale dynamics of grain-size variation in alluvial basins, 2. Application to syntectonic conglomerate. Basin Research 4, 91–102.CrossRefGoogle Scholar
Heller, P. L., & Paola, C. 1996. Downstream changes in alluvial architecture: an exploration of controls on channel-stacking patterns. Journal of Sedimentary Research B66, 297–306.Google Scholar
Hequette, A., & Hill, P. R. 1993. Storm-generated currents and offshore sediment transport on a sandy shoreface, Tibjak Beach, Canadian Beaufort Sea. Marine Geology 113, 283–304.CrossRefGoogle Scholar
Hesp, P., Hyde, R., Hesp, V., & Zengyu, Q. 1989. Longitudinal dunes can move sideways. Earth Surface Processes and Landforms 14, 447–451.CrossRefGoogle Scholar
Hesse, R. 1990a. Origin of chert: diagenesis of biogenic siliceous sediments. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 227–251. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Hesse, R.1990b. Silica diagenesis: origin of inorganic and replacement cherts. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 253–275. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Hills, J. G., & Mader, C. L. 1997. Tsunami produced by the impacts of small asteroids. Annals of New York Academy of Sciences 822, 381–394.CrossRefGoogle Scholar
Hindmarsh, R. C. A. 1998. Drumlinization and drumlin-forming instabilities: viscous till mechanisms. Journal of Glaciology 44, 293–314.CrossRefGoogle Scholar
Hindmarsh, R. C. A, Dunlop, P., Clark, C. D. 2003. Modeling the geomorphological effects of till redistribution; assessing a dynamic theory for Rogen moraine formation and drumlin formation. In 16th Congress of the International Union for Quaternary Research.Reno, NV: INQUA.Google Scholar
Hine, A. C. 1975. Bedform distribution and migration patterns on tidal deltas in the Chatham Harbor estuary, Cape Cod, Massachusetts. In Estuarine Research: Volume II, Geology and Engineering (ed. Cronin, L. E.), pp. 235–252. London: Academic Press.Google Scholar
Hine, A. C. 1977. Lily Bank, Bahamas: history of an active oolite sand shoal. Journal of Sedimentary Petrology 47, 1554–1581.Google Scholar
Hine, A. C., & Neumann, A. C. 1977. Shallow carbonate bank margin growth and structure, Little Bahama Bank, Bahamas. American Association of Petroleum Geologists Bulletin 61, 376–406.Google Scholar
Hinnov, L. 2000. New perspectives on orbitally forced stratigraphy. Annual Review of Earth and Planetary Sciences 28, 419–475.CrossRefGoogle Scholar
Hoffman, P. 1973. Recent and ancient algal stromatolites: seventy years of pedagogic cross-pollination. In Evolving Concepts in Sedimentology (ed. Ginsburg, R. N.), pp. 178–191. Baltimore, MA: The Johns Hopkins University Press.
Hoffman, P. 1974. Shallow and deepwater stromatolites in Lower Proterozoic platform-to-basin facies change, Great Slave Lake, Canada. American Association of Petroleum Geologists Bulletin 58, 856–867.Google Scholar
Hoffman, P. 1976. Stromatolite morphogenesis in Shark Bay, Western Australia. In Stromatolites (ed. Walter, M. R.), pp. 261–271. Amsterdam: Elsevier.CrossRef
Hoffman, P. F., Kaufman, A. J., Halverson, G. P., & Schrag, D. P. 1998. A Neoproterozoic snowball Earth. Science 281, 1342–1346.CrossRefGoogle ScholarPubMed
Hoffman, P. F., & Schrag, D. P. 2002. The snowball Earth hypothesis: testing the limits of global change. Terra Nova 14, 129–155.CrossRefGoogle Scholar
Hogg, A. J., & Huppert, H. E. 2001. Two-dimensional and axisymmetric models for compositional and particle-driven gravity currents in uniform ambient flows. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 121–134. Oxford: Blackwells.CrossRefGoogle Scholar
Holland, H. D. 1972. The geologic history of seawater – an attempt to solve the problem. Geochimica et Cosmochimica Acta 36, 637–651.CrossRefGoogle Scholar
Holland, H. D.1994. Early Proterozoic atmospheric change. In Early Life on Earth, Nobel Symposium No. 84 (ed. Bengston, S.), pp. 237–244. New York: Columbia University Press.Google Scholar
Holland, H. D. 2005. Sea level, sediments and the composition of sea water. American Journal of Science 305, 220–239.CrossRefGoogle Scholar
Hollister, C. D. 1993. The concept of deep-sea contourites. Sedimentary Geology 82, 5–15.CrossRefGoogle Scholar
Hollister, C. D., & McCave, I. N. 1984. Sedimentation under deep-sea storms. Nature 309, 220–225.CrossRefGoogle Scholar
Holman, R. A. 1983. Edge waves and the configuration of the shoreline. In CRC Handbook of Coastal Processes and Erosion (ed. Komar, P. D.), pp. 21–34. Boca Raton, FL: CRC Press.Google Scholar
Holman, R. A., & Sallenger, A. H. 1993. Sand bar generation: a discussion of the Duck experimental series. Journal of Coastal Research, Special Issue 15, 76–92.Google Scholar
Holmlund, P. 1988. Internal geometry and evolution of moulins, Storglaciaren, Sweden. Journal of Glaciology 34, 242–248.CrossRefGoogle Scholar
Homewood, P., & Allen, P. A. 1981. Wave-, tide-, and current-controlled sandbodies of Miocene molasse, western Switzerland. American Association of Petroleum Geologists Bulletin 65, 2534–2545.Google Scholar
Honji, H., Kaneko, A., & Matsunaga, N. 1980. Flows above oscillatory ripples. Sedimentology 27, 225–229.CrossRefGoogle Scholar
Hooke, R. LeB. 1998. Principles of Glacier Mechanics. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Hooke, R. L., & Jennings, C. E. 2006. On the formation of the tunnel valleys of the southern Laurentide ice sheet. Quaternary Science Reviews 25, 1364–1372.CrossRefGoogle Scholar
Horbury, A. D., & Robinson, A. G. 1993. Diagenesis and Basin Development. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Hori, K., Saito, Y., Zhang, Q., & Wang, P. 2002. Architecture and evolution of the tide-dominated Changjiang (Yangtze) River delta, China. Sedimentary Geology 146, 249–264.CrossRefGoogle Scholar
Horita, J., Zimmermann, H., & Holland, H. D. 2002. Chemical evolution of seawater during the Phanerozoic: implications from the record of marine evaporites. Geochimica et Cosmochimica Acta 66, 3733–3756.CrossRefGoogle Scholar
Horowitz, D. H. 1982. Geometry and origin of large-scale deformation structures in some ancient windblown sand deposits. Sedimentology 29, 155–180.CrossRefGoogle Scholar
Horton, R. E. 1945. Erosional development of streams and their drainage basins: hydrophysical approach to quantitative morphology. Geological Society of America Bulletin 56, 275–370.CrossRefGoogle Scholar
House, M. R., & Gale, A. S. 1995. Orbital Forcing Timescales and Cyclostratigraphy. London: Geological Society of London.Google Scholar
Howard, A. D. 1967. Drainage analysis in geologic interpretation: a summary. American Association of Petroleum Geologists Bulletin 51, 2246–2259.Google Scholar
Howard, A. D. 1987. Modeling fluvial systems: rock-, gravel- and sand-bed channels. In River Channels: Environment and Process (ed. Richards, K. S.), pp. 69–94. Oxford: Blackwells.Google Scholar
Howard, J. D., & Frey, R. W. 1975. Estuaries of the Georgia coast, U.S.A.: sedimentology and biology, II. Regional animal-sediment characteristics of Georgia Estuaries. Senckenbergiana Maritima 7, 33–103.Google Scholar
Howard, J. D., & Reineck, H.-E. 1972. Georgia coastal region, Sapelo Island, U.S.A.: sedimentology and biology, IV. Physical and biogenic sedimentary structures of the nearshore shelf. Senckenbergiana Maritima 4, 47–79.Google Scholar
Howard, J. D., & Reineck, H.-E. 1981. Depositional facies of high-energy beach-to-offshore sequence, comparison with low energy sequence. American Association of Petroleum Geologists Bulletin 65, 807–830.Google Scholar
Hower, J., Eslinger, E. V., Hower, M. E., & Perry, E. A. 1976. Mechanism of burial metamorphism of argillaceous sediment: 1. Mineralogical and chemical evidence. Geological Society of America Bulletin 87, 725–737.2.0.CO;2>CrossRefGoogle Scholar
Hoyt, J. H., & Henry, V. J. 1967. Influence of island migration on barrier island sedimentation. Geological Society of America Bulletin 78, 77–86.CrossRefGoogle Scholar
Hoyt, J. H., & Henry, V. J. 1971. Origin of capes and shoals along the southeastern coast of the United States. Geological Society of America Bulletin 82, 59–66.CrossRefGoogle Scholar
Hsü, K. J. 1972a. Origin of saline giants: a critical review after the discovery of the Mediterranean evaporite. Earth-Science Reviews 8, 371–396.CrossRefGoogle Scholar
Hsü, K. J. 1972b. When the Mediterranean dried up. Scientific American 227, 26–36.CrossRefGoogle Scholar
Hsü, K. J., & Jenkyns, H. C. 1974. Pelagic Sediments: On Land and Under the Sea. Oxford: Blackwells.Google Scholar
Hubbard, D. K., Oertel, G., & Nummedal, D. 1979. The role of waves and tidal currents in the development of tidal-inlet sedimentary structures and sand body geometry: examples from North Carolina, South Carolina and Georgia. Journal of Sedimentary Petrology 49, 1073–1092.Google Scholar
Hulbe, C. L., MacAyeal, D. R., Denton, G. H., Kleman, J., & Lowell, T. V. 2004. Catastrophic ice shelf breakup as the source of Heinrich event icebergs. Paleoceanography19, PA1004, doi: 10.1029/2003PA000890.CrossRef
Hulscher, S. 1996. Tide-induced large-scale regular bedform patterns in three-dimensional shallow water model. Journal of Geophysical Research 101, 20,727–20,744.CrossRefGoogle Scholar
Hunt, D., & Gawthorpe, R. L. 2000. Sedimentary Responses to Forced Regressions. London: Geological Society of London.Google Scholar
Hunter, R. E. 1977. Basic types of stratification in small eolian dunes. Sedimentology 24, 361–387.CrossRefGoogle Scholar
Hunter, R. E. 1985. Subaqueous sand flow cross strata. Journal of Sedimentary Petrology 55, 886–894.Google Scholar
Hunter, R. E., & Clifton, H. E. 1982. Cyclic deposits and hummocky cross-stratification of probable storm origin in the Upper Cretaceous rocks of the Cape Sebastian area, south-western Oregon. Journal of Sedimentary Petrology 52, 127–143.Google Scholar
Hunter, R. E., Clifton, H. E., & Phillips, R. L. 1979. Depositional processes, sedimentary structures and predicted vertical sequences in barred nearshore systems, southern Oregon coast. Journal of Sedimentary Petrology 49, 711–726.Google Scholar
Hunter, R. E., & Kocurek, G. 1986. An experimental study of subaqueous slipface deposition. Journal of Sedimentary Petrology 56, 387–394.CrossRefGoogle Scholar
Huntley, D. A., & Bowen, A. J. 1973. Field observations of edge waves. Nature 243, 160–161.CrossRefGoogle Scholar
Huntley, D. A., Guza, R. T., & Thornton, D. B. 1981. Field observations of surf beat, 1: progressive edge waves. Journal of Geophysical Research 83, 1913–1920.Google Scholar
Huppert, H. E. 1998. Quantitative modeling of granular suspension flows. Philosophical Transactions of the Royal Society of London 356A, 2471–2496.Google Scholar
Hutcheon, I. E. 1990. Aspects of diagenesis in coarse-grained siliciclastic rocks. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 165–176. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Huthnance, J. M. 1982a. On one mechanism forming linear sand banks. Estuarine, Coastal and Shelf Science 14, 79–99.CrossRefGoogle Scholar
Huthnance, J. M. 1982b. On the formation of sand banks of finite extent. Estuarine, Coastal and Shelf Science 15, 277–299.CrossRefGoogle Scholar
Hyndman, D., & Hyndman, D. 2006. Natural Hazards and Disasters. Belmont, CA: Thomson Brooks/Cole.Google Scholar
Iida, K., & Iwasaki, T. 1983. Tsunamis: Their Science and Engineering. Dordrecht: Reidel.CrossRefGoogle Scholar
Ikeda, S., & Parker, G. 1989. River Meandering. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Illing, L. V., Wells, A. J., & Taylor, J. C. M. 1965. Penecontemporary dolomite in the Persian Gulf. In Dolomitization and Limestone Diagenesis, A Symposium (ed. Pray, L. C. & Murray, R. C.), pp. 89–111. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Imboden, D. M., & Wüest, A. 1995. Mixing mechanisms in lakes. In Physics and Chemistry of Lakes, 2nd edn. (ed. Lerman, A., Imboden, D., & Gat, J.), pp. 83–138. New York: Springer-Verlag.CrossRefGoogle Scholar
Imbrie, J., Hays, J. D., Martinson, D. G. et al. 1984. The orbital theory of Pleistocene climate: support from a revised chronology of the marine δ18O record. In Milankovitch and Climate: Understanding the Response to Astronomical Forcing (ed. Berger, A., Imbrie, J., Hays, J., & Kukla, G.), pp. 269–305. Dordrecht: Reidel.CrossRefGoogle Scholar
Imbrie, J., & Imbrie, K. P. 1979. Ice Ages: Solving the Mystery. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Imbrie, J., Mix, A. C., & Martinson, D. G. 1993. Milankovitch theory viewed from Devils Hole. Nature 363, 531–533.CrossRefGoogle Scholar
Imperato, D. P., Sexton, W. J., & Hayes, M. O. 1988. Stratigraphy and sediment characteristics of a mesotidal ebb-tidal delta, North Edisto Inlet, South Carolina. Journal of Sedimentary Petrology 58, 950–958.Google Scholar
Immenhauser, A., Hillgartner, H., & Benthum, E. 2005. Microbial–foraminiferal episodes in the Early Aptian of the southern Tethyan margin: ecological significance and possible relation to oceanic anoxia. Sedimentology 52, 77–99.CrossRefGoogle Scholar
Inden, R. F., & Moore, C. H. 1983. Beach. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 211–265. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Iverson, N. R. 1991. Potential effects of subglacial water-pressure fluctuations on quarrying. Journal of Glaciology 37, 27–36.CrossRefGoogle Scholar
Iverson, R. M. 1997. The physics of debris flows. Reviews of Geophysics 35, 245–296.CrossRefGoogle Scholar
Jackson, R. G. 1975. Hierarchical attributes and a unifying model of bedforms composed of cohesionless sediment and produced by shearing flow. Geological Society of America Bulletin 86, 1,523–1,533.2.0.CO;2>CrossRefGoogle Scholar
Jackson, R. G. 1976. Sedimentological and fluid-dynamic implications of the turbulent bursting phenomenon in geophysical flows. Journal of Fluid Mechanics 77, 531–560.CrossRefGoogle Scholar
Jaeger, H. M., Nagel, S. R., & Behringer, R. P. 1996. The physics of granular materials. Physics Today 49, 32–36.CrossRefGoogle Scholar
Jahren, A. H. 2004. Factors of soil formation: biota. In Encyclopedia of Soils in the Environment (ed. Hillel, D., Rosenzweig, C., Powlson, D.et al.), pp. 507–512. New York: Academic Press.Google Scholar
James, N. P., & Choquette, P. W. 1983. Diagenesis 6. Limestones – the sea floor diagenetic environment. Geoscience Canada 11, 161–194.Google Scholar
James, N. P. & MacIntyre, I. G. 1985. Carbonate depositional environments, modern and ancient. Part 1 – reefs, zonation, depositional facies and diagenesis. Quarterly Journal of the Colorado School of Mines 80, 70pp.Google Scholar
James, H. L. 1954. Sedimentary facies of iron formations. Economic Geology 49, 235–293.CrossRefGoogle Scholar
James, N. P. 1981. Megablocks of calcified algae in the Cow Head Breccia, western Newfoundland: vestiges of a Cambro-Ordovician margin. Geological Society of America Bulletin 42, 799–811.2.0.CO;2>CrossRefGoogle Scholar
James, N. P.1983. Reefs. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 345–462. Tulsa, OK: American Association of Petroleum.Google Scholar
James, N. P., Bone, Y., Collins, L. B., & Kyser, T. K. 2001. Surficial sediments of the Great Australian Bight: facies dynamics and oceanography on a vast cool-water carbonate shelf. Journal of Sedimentary Research 71, 5549–567.CrossRefGoogle Scholar
James, N. P., & Bourque, P. A. 1992. Reefs and mounds. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 323–345. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Choquette, P. W. 1988. Paleokarst. New York: Springer-Verlag.CrossRefGoogle Scholar
James, N. P., & Choquette, P. W.1990a. Limestones – introduction. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 9–12. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Choquette, P. W.1990b. Limestones – the sea floor diagenetic environment. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 13–34. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Choquette, P. W.1990c. Limestones – the meteoric diagenetic environment. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 35–74. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Clarke, A. D. 1997. Cool Water Carbonates. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
James, N. P., & Ginsburg, R. N. 1979. The Seaward Margin of the Belize Reef. Oxford: Blackwells.Google Scholar
James, N. P., & Kendall, A. C. 1992. Introduction to carbonate and evaporite facies models. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 265–275. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Jenkins, J. G., & Savage, S. B. 1983. A theory for the rapid flow of identical, smooth, nearly elastic spherical particles. Journal of Fluid Mechanics 130, 187–202.CrossRefGoogle Scholar
Jenkyns, H. C., Forster, A., Schouten, S., & Damste, J. S. S. 2004. High temperatures in the Late Cretaceous Arctic Ocean. Nature 432, 888–892.CrossRefGoogle ScholarPubMed
Johnson, A. M. 1970. Physical Processes in Geology. San Francisco: W. H. Freeman.Google Scholar
Johnson, D. D., & Beaumont, C. 1995. Preliminary results from a planform kinematic model of orogen evolution, surface processes and the development of clastic foreland basin stratigraphy. In Stratigraphic Evolution of Foreland Basins (ed. Dorobek, S. L. & Ross, G. M.), pp. 3–24. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Johnson, M. A., Kenyon, N. H., Belderson, R. H., & Stride, A. H. 1982. Sand transport. In Offshore Tidal Sands (ed. Stride, A. H.), pp. 58–94. London: Chapman and Hall.CrossRefGoogle Scholar
Johnson, N. M., & Driess, S. J. 1989. Hydrostratigraphic interpretation using indicator geostatistics. Water Resources Research 25, 2501–2510.CrossRefGoogle Scholar
Jones, B., & Desrochers, A. 1992. Shallow platform carbonates. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 277–301. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Jones, B., & Luth, R. W. 2002. Dolostones from Grand Cayman, British West Indes. Journal of Sedimentary Research 72, 559–569.CrossRefGoogle Scholar
Jones, L. S., & Schumm, S. A. 1999. Causes of avulsion: an overview. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 171–178. Oxford: Blackwells.CrossRefGoogle Scholar
Jopling, A. V., & McDonald, B. C. 1975. Glaciofluvial and Glaciolacustrine Sediments. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Jordan, T. E. 1981. Thrust loads and foreland basin evolution, Cretaceous western United States. American Association of Petroleum Geologists Bulletin 65, 2506–2620.Google Scholar
Jordan, T. E., & Flemings, P. B. 1990. From geodynamical models to basin fill – a stratigraphic perspective. In Quantitative Dynamic Stratigraphy (ed. Cross, T. A.), pp. 149–163. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Jordan, T. E., & Flemings, P. B. 1991. Large-scale stratigraphic architecture, eustatic variation, and unsteady tectonism: a theoretical evaluation. Journal of Geophysical Research 96, 6681–6699.CrossRefGoogle Scholar
Jorgenson, B. B. 2000. Bacteria and marine biogeochemistry. In Marine Geochemistry (ed. Schulz, H. D. & Zabel, M.), pp. 173–207. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Julien, P. Y. 2002. River Mechanics. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jullien, R., Meakin, P., & Pavlovitch, A. 2002. Three-dimensional model for particle-size segregation by shaking. Physical Review Letters 69, 640–643.CrossRefGoogle Scholar
Kaiser, J. 2004. Wounding Earth's fragile skin. Science 304, 1616–1618.CrossRefGoogle ScholarPubMed
Kamb, B. 1987. Glacier surge mechanism based on linked cavity configuration of the basal water conduit system. Journal of Geophysical Research 92, 9083–9100.CrossRefGoogle Scholar
Kamb, B. 1991. Rheological nonlinearity and flow instability in the deforming bed mechanism of ice stream motion. Journal of Geophysical Research 96B, 16,585–16,595.CrossRefGoogle Scholar
Kamb, B., Raymond, C. F., Harrison, W. D.et al. 1985. Glacier surge mechanism: 1982–1983 surge of Variegated Glacier, Alaska. Science 227, 469–479.CrossRefGoogle ScholarPubMed
Kanfoush, S. L., Hodell, D. A., Charles, C. D.et al. 2000. Millennial-scale instability of the Antarctic Ice sheet during the last glaciation. Science 288, 1815–1819.CrossRefGoogle ScholarPubMed
Karssenberg, D., Törnqvist, T. E., & Bridge, J. S. 2001. Conditioning a process-based model of sedimentary architecture to well data. Journal of Sedimentary Research 71, 868–879.CrossRefGoogle Scholar
Karssenberg, D., Dalman, R., Weltje, G. J., Postma, G., & Bridge, J. S. 2004. Numerical modelling of delta evolution by nesting high and low resolution process-based models of sedimentary basin filling. In Joint EURODELTA/EUROSTRATAFORM Meeting, Venice (Italy), pp. 1–37.
Katz, M. E., Finkel, Z. V., Grzebyk, D., Knoll, A. H., & Falkowski, P. G. 2004. Evolutionary trajectories and biogeochemical impacts of marine eukaryotic phytoplankton. Annual Review of Ecology, Evolution, and Systematics 35, 523–556.CrossRefGoogle Scholar
Keen, T. R., & Glenn, S. M. 1994. A coupled hydrodynamic–bottom boundary layer model of Ekman flow on stratified continental shelves. Journal of Physical Oceanography 24, 1732–1749.2.0.CO;2>CrossRefGoogle Scholar
Kelly, D. S., Karson, J. A., Früh-Green, G. L.et al. 2005. A serpentine-hosted ecosystem: the Lost City hydrothermal field. Science 307, 1428–1434.CrossRefGoogle Scholar
Kemp, P. H., & Simons, R. R. 1982. The interaction between waves and a turbulent current: waves propagating with the current. Journal of Fluid Mechanics 116, 227–250.CrossRefGoogle Scholar
Kendall, A. C. 1992. Evaporites. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 375–409. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Kendall, A. C., & Harwood, G. M. 1996. Marine evaporites: arid shorelines and basins. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 281–324. Oxford: Blackwell Science.Google Scholar
Kendall, C. G. St. C., & Skipwith, P. A. D'E. 1968. Recent algal mats of a Persian Gulf lagoon. Journal of Sedimentary Petrology 38, 1040–1058.Google Scholar
Kendall, C. G. St. C., & Skipwith, P. A. D'E. 1969. Geomorphology of a Recent shallow-water carbonate province: Khor al Bazam, Trucial Coast, southwestern Persian Gulf. Geological Society of America Bulletin 80, 865–892.CrossRefGoogle Scholar
Kendall, C. G. St. C., & Warren, J. 1987. A review of the origin and setting of tepees and their associated fabrics. Sedimentology 34, 1007–1027.CrossRefGoogle Scholar
Kennard, J. M., & James, N. P. 1986. Thrombolites and stromatolites: two distinct types of microbial structures. Palios 1, 492–503.CrossRefGoogle Scholar
Kennedy, J. F. 1963. The mechanics of dunes and antidunes in erodible-bed channels. Journal of Fluid Mechanics 16, 521–544.CrossRefGoogle Scholar
Kenyon, N. H., Amir, A., & Cramp, A. 1995. Geometry of the younger sediment bodies on the Indus Fan. In Atlas of Deep Water Environments: Architectural Style in Turbidite Systems (ed. Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, F. Ricci, & Smith, R. D. A.), pp. 89–93. London: Chapman and Hall.CrossRefGoogle Scholar
Kiessling, W., Flügel, E., & Golonka, J. 2002. Phanerozoic Reef Patterns. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Kim, J., Dong, H., Seabaugh, J., Newell, S. W., & Eberl, D. D. 2004. Role of microbes in the smectite-to-illite reaction. Science 303, 830–832.CrossRefGoogle ScholarPubMed
King, L. H. 1993. Till in the marine environment. Journal of Quaternary Science 8, 347–358.CrossRefGoogle Scholar
Kinsman, D. J. 1966. Gypsum and anhydrite of Recent age, Trucial Coast, Persian Gulf. In Proceedings of the Second Salt Symposium (ed. Rau, J. L.), Vol. 1, pp. 302–326. Cleveland, OH: Northern Ohio Geological Society.Google Scholar
Kinsman, D. J. J., & Park, R. K. 1976. Algal belt and coastal sabkha evolution, Trucial Coast, Persian Gulf. In Stromatolites (ed. Walter, M. R.), pp. 421–433. Amsterdam: Elsevier.Google Scholar
Kirkby, M. J. 1994. Thresholds and instability in stream head hollows: a model of magnitude and frequency for wash processes. In Process Models and Theoretical Geomorphology (ed. Kirkby, M. J.), pp. 294–314. Chichester: Wiley.Google Scholar
Kirschvink, J. L. 1992. Late Proterozoic low latitude glaciation: the snowball Earth. In The Proterozoic Biosphere (ed. Schopf, J. W. & Klein, C.), pp. 51–52. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Klappa, C. F. 1978. Biolithogenesis of Microcodium; elucidation. Sedimentology 25, 489–522.CrossRefGoogle Scholar
Klein, C., & Beukes, N. J. 1992. Proterozoic iron formations. In Proterozoic Crustal Evolution (ed. Condie, K. C.), pp. 383–418. Amsterdam: Elsevier.Google Scholar
Klein, G. deV. 1970. Depositional and dispersal dynamics of intertidal sand bars. Journal of Sedimentary Petrology 40, 1095–1127.Google Scholar
Kleinhans, M. G. 2004. Sorting in grains flows at the lee side of dunes. Earth-Science Reviews 65, 75–102.CrossRefGoogle Scholar
Kleinhans, M. G. 2005. Grain-size sorting in grainflows at the lee side of deltas. Sedimentology 52, 291–311.CrossRefGoogle Scholar
Kleypas, J. A., Buddemeier, R. W., Archer, D.et al. 1999. Geochemical consequences of increased atmospheric carbon dioxide on coral reefs. Science 284, 118–120.CrossRefGoogle ScholarPubMed
Knauth, L. P. 2003. Siliceous sediments. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 660–666. Dordrecht: Kluwer Academic Publishers.Google Scholar
Kneller, B. 1996. Beyond the turbidite paradigm: physical models for deposition of turbidites and their implication for reservoir prediction. In Characterisation of Deep Marine Clastic Systems (ed. Hartley, A. & Prosser, D. J.), pp. 29–46. London: Geological Society of London.Google Scholar
Kneller, B. C., & Branney, M. J. 1995. Sustained high density turbidity currents and the deposition of massive sands. Sedimentology 42, 607–616.CrossRefGoogle Scholar
Kneller, B. C., & Buckee, C. 2000. The structure and fluid mechanics of turbidity currents: a review of some recent studies and their geological implications. Sedimentology 47 (Supplement 1), 62–94.CrossRefGoogle Scholar
Kneller, B. C., Edwards, D., McCaffrey, W., & Moore, R. 1991. Oblique reflection of turbidity currents. Geology 19, 250–252.2.3.CO;2>CrossRefGoogle Scholar
Knight, J. B., Jaeger, H. M., & Nagel, S. R. 1993. Vibration-induced size separation in granular media: the convection connection. Physical Review Letters 70, 3728–3731.CrossRefGoogle ScholarPubMed
Knight, P. G. 2006. Glacier Science and Environmental Change. Oxford: Blackwells.CrossRefGoogle Scholar
Knight, R. J., & McLean, J. R. 1986. Shelf Sands and Sandstones. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Knighton, A. D. 1998. Fluvial Forms and Processes: A New Perspective. London: Arnold.Google Scholar
Knoll, A. H., & Carroll, S. B. 1999. Early animal evolution: emerging views from comparative biology and geology. Science 284, 2129–2137.CrossRefGoogle ScholarPubMed
Knox, J. C. 1983. Responses of river systems to Holocene climates. In Late Quaternary Environments of the United States, Volume 2, The Holocene (ed. Wright, H. E. & Porter, S. C.), pp. 26–41. Minneapolis, MN: University of Minnesota Press.Google Scholar
Knox, J. C. 1996. Fluvial systems since 20,000 years BP. In Continental Palaeohydrology (ed. Gregory, K. J., Starkel, L., & Baker, V. R.), pp. 87–108. Chichester: Wiley.Google Scholar
Kochel, R. C., & Baker, V. R. 1988. Palaeoflood analysis using slack water deposits. In Flood Geomorphology (ed. Baker, V. R., Kochel, R. C., & Patton, P. P.), pp. 357–376. New York: Wiley.Google Scholar
Kocurek, G. 1981. Significance of interdune deposits and bounding surfaces in eolian dune sands. Sedimentology 28, 753–780.CrossRefGoogle Scholar
Kocurek, G. 1988. First order and super bounding surfaces in eolian sequences – bounding surfaces revisited. Sedimentary Geology 56, 193–206.CrossRefGoogle Scholar
Kocurek, G. 1991. Interpretation of ancient eolian sand dunes. Annual Review of Earth and Planetary Sciences 19, 43–75.CrossRefGoogle Scholar
Kocurek, G.1996. Desert aeolian systems. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 125–155. Oxford: Blackwells.Google Scholar
Kocurek, G., & Havholm, K. G. 1993. Eolian sequence stratigraphy – a conceptual framework. In Recent Advances in and Application of Siliciclastic Sequence Stratigraphy (ed. Weimer, P. & Posamentier, H.), pp. 393–409. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Kocurek, G., & Lancaster, N. 1999. Aeolian sediment states: theory and Mojave Kelso Dunefield example. Sedimentology 46, 505–516.CrossRefGoogle Scholar
Kocurek, G., & Nielsen, J. 1986. Conditions favourable for the formation of warm-climate aeolian sand sheets. Sedimentology 33, 795–816.CrossRefGoogle Scholar
Kocurek, G., Townsley, M., Yeh, E., Havholm, K., & Sweet, M. L. 1992. Dune and dune-field development on Padre Island, Texas, with implications for interdune deposition and water-table controlled accumulation. Journal of Sedimentary Petrology 62, 622–635.Google Scholar
Kohout, F. A., Henry, H. R., & Banks, J. E. 1977. Hydrogeology related to geothermal conditions of the Florida Plateau. In The Geothermal Nature of the Florida Plateau (ed. Smith, D. L. & Griffin, G. E.), pp. 1–41. Tallahassee, FL: Florida Bureau of Geology.Google Scholar
Kolb, C. R., & Van Lopik, J. R. 1966. Depositional environments of the Mississippi River deltaic plain – southeastern Louisiana. In Deltas in Their Geologic Framework (ed. Shirley, M. L. & Ragsdale, J. A.), pp. 17–61. Houston, TX: Houston Geological Society.Google Scholar
Kolla, V., & Coumes, F. 1985. Indus fan, Indian Ocean. In Submarine Fans and Related Turbidite Systems (ed. Bouma, A. H., Normark, W. R., & Barnes, N. E.), pp. 129–136. New York: Springer-Verlag.CrossRefGoogle Scholar
Kolla, V., & Coumes, F. 1987. Morphology, internal structure, seismic stratigraphy, and sedimentation of Indus fan. American Association of Petroleum Geologists Bulletin 71, 650–677.Google Scholar
Kolla, V., & Perlmutter, M. A. 1993. Timing of turbidite sedimentation on the Mississippi Fan. American Association of Petroleum Geologists Bulletin 77, 1129–1141.Google Scholar
Koltermann, C. E., & Gorelick, S. M. 1992. Paleoclimatic signature in terrestrial flood deposits. Science 256, 1775–1782.CrossRefGoogle ScholarPubMed
Koltermann, C. E., & Gorelick, S. M. 1996. Heterogeneity in sedimentary deposits: a review of structure-imitating, process-imitating and descriptive approaches. Water Resources Research 32, 2617–2658.CrossRefGoogle Scholar
Komar, P. D. 1973. Computer models of delta growth due to sediment input from rivers and longshore transport. Geological Society of America Bulletin 84, 2,217–2,226.2.0.CO;2>CrossRefGoogle Scholar
Komar, P. D. 1974. Oscillatory ripple marks and the evaluation of ancient wave conditions and environments. Journal of Sedimentary Petrology 44, 169–180.Google Scholar
Komar, P. D. 1976. Beach Processes and Sedimentation. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Komar, P. D. 1977. Modeling of sand transport on beaches and the resulting shoreline evolution. In The Sea (ed. Goldberg, E.et al.) Vol. 6, pp. 499–513. New York: Wiley.Google Scholar
Komar, P. D.1996. Entrainment of sediments from deposits of mixed grain sizes and densities. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 127–181. Chichester: Wiley.Google Scholar
Komar, P. D. 1998. Beach Processes and Sedimentation, 2nd edn. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Komar, P. D., & Miller, M. C. 1973. The threshold of sediment motion under oscillatory water waves. Journal of Sedimentary Petrology 43, 1101–1110.Google Scholar
Komar, P. D., & Miller, M. C. 1975a. On the comparison between the threshold of sediment motion under waves and unidirectional currents with a discussion of the practical evaluation of the threshold. Journal of Sedimentary Petrology 45, 362–367.Google Scholar
Komar, P. D., & Miller, M. C. 1975b. The initiation of oscillatory ripple marks and the development of plane-bed at high shear stresses under waves. Journal of Sedimentary Petrology 45, 697–703.Google Scholar
Kooi, H., & Beaumont, C. 1994. Escarpment evolution on high-elevation rifted margins: insights derived from a surface processes model that combines diffusion, advection, and reaction. Journal of Geophysical Research 99, 12191–12209.CrossRefGoogle Scholar
Kooi, H., & Beaumont, C. 1996. Large-scale geomorphology: classical concepts reconciled and integrated with contemporary ideas using a surface processes model. Journal of Geophysical Research 101, 3361–3386.CrossRefGoogle Scholar
Kor, P. S. G., Shaw, J., & Sharpe, D. R. 1991. Erosion of bedrock by subglacial meltwater, Georgian Bay, Ontario: a regional view. Canadian Journal of Earth Sciences 28, 623–642.CrossRefGoogle Scholar
Kosters, E. C. 1989. Organic–clastic facies relationships and chronostratigraphy of the Barataria interlobe basin, Mississippi delta plain. Journal of Sedimentary Petrology 59, 98–113.Google Scholar
Kosters, E. C., & Suter, J. R. 1993. Facies relationships and systems tracts in the late Holocene Mississippi delta plain. Journal of Sedimentary Petrology 63, 727–733.Google Scholar
Kraft, J. C. 1971. Sedimentary facies patterns and geologic history of a Holocene marine transgression. Geological Society of America Bulletin 82, 2131–2158.CrossRefGoogle Scholar
Kraft, J. C., Chrzastowski, M. J., Belknap, D. F., Toscano, M. A., & Fletcher, C. H. 1987. The transgressive barrier–lagoon coast of Delaware: morphostratigraphy, sedimentary sequences and responses to relative rises in sea level. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 129–144. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Kraus, M. J. 1987. Integration of channel and floodplain suites, II. Vertical relations of alluvial paleosols. Journal of Sedimentary Petrology 56, 602–612.Google Scholar
Kraus, M. J. 1996. Avulsion deposits in lower Eocene alluvial rocks, Bighorn Basin, Wyoming. Journal of Sedimentary Research 66B, 354–363.Google Scholar
Kraus, M. J. 1999. Paleosols in clastic sedimentary rocks: their geologic applications. Earth-Science Reviews 47, 41–70.CrossRefGoogle Scholar
Kraus, M. J., & Aslan, A. 1999. Palaeosol sequences in floodplain environments: a hierarchical approach. In Palaeoweathering, Palaeosurfaces and Related Continental Deposits (ed. Thiry, M. & Simon-Coicon, R.), pp. 303–321. Oxford: Blackwells.Google Scholar
Kraus, M. J., & Bown, T. M. 1986. Palaeosols and time resolution in alluvial stratigraphy. In Paleosols, Their Recognition and Interpretation (ed. Wright, V. P.), pp. 180–207. Princeton, NJ: Princeton University Press.Google Scholar
Kraus, M. J., & Gwinn, B. M. 1997. Controls on the development of early Eocene avulsion deposits and floodplain paleosols, Willwood Formation, Bighorn Basin. Sedimentary Geology 114, 33–54.CrossRefGoogle Scholar
Kraus, M. J., & Hasiotis, S. T. 2006. Significance of different modes of rhizolith preservation to interpreting paleoenvironmental and paleohydologic settings: examples from Paleogene paleosols, Bighorn Basin, Wyoming, U.S.A. Journal of Sedimentary Research 76, 633–646.CrossRefGoogle Scholar
Kraus, M. J., & Wells, T. M. 1999. Recognizing avulsion deposits in the ancient stratigraphic record. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 251–268. Oxford: Blackwells.CrossRefGoogle Scholar
Krause, D. C., White, W. C., Piper, D. J. W., & Heezen, B. C. 1970. Turbidity currents and cable breaks in the western New Britain Trench. Geological Society of America Bulletin 81, 2153–2160.CrossRefGoogle Scholar
Krynine, P. D. 1942. Differential sedimentation and its products during one complete geosynclinal cycle. Proceedings of the 1st Pan American Congress on Mining Engineering Geology, Part 1, Vol. 2, pp. 537–560. Santiago: Chilean Institute of Mining Engineering.Google Scholar
Krynine, P. D. 1950. Petrology, stratigraphy and origin of the Triassic sedimentary rocks of Connecticut. Connecticut State Geological and Natural History Survey Bulletin 73, 1–247.Google Scholar
Kuenen, P. H. 1965. Value of experiments in geology. Geologie en Mijnbouw 44, 22–36.Google Scholar
Kuenen, P. H., & Migliorini, C. I. 1950. Turbidity currents as a cause of graded bedding. Journal of Geology 58, 91–127.CrossRefGoogle Scholar
Kuijper, C., Cornelisse, J. M., & Winterwerp, J. C. 1989. Research on erosive properties of cohesive sediments. Journal of Geophysical Research 95, 14,341–14,350.CrossRefGoogle Scholar
Kumar, N., & Sanders, J. E. 1974. Inlet sequences: a vertical succession of sedimentary structures and textures created by the lateral migration of tidal inlets. Sedimentology 21, 491–532.CrossRefGoogle Scholar
Kummel, B., & Raup, D. 1965. Handbook of Paleontological Techniques. San Francisco, CA: W. H. Freeman.Google Scholar
Kump, L. R., Brantley, S. L., & Arthur, M. A. 2000. Chemical weathering, atmospheric CO2, and climate. Annual Review of Earth and Planetary Sciences 28, 611–667.CrossRefGoogle Scholar
Kump, L. R., Kasting, J. F., & Crane, R. G. 2004. The Earth System, 2nd edn. Upper Saddle River, NJ: Prentice-Hall.Google Scholar
Lachenbruch, A. H. 1962. Mechanics of thermal contraction cracks and ice-wedge polygons. Geological Society of America Special Paper 70.
Lal, R. 2004. Soil carbon sequestration impacts on global climate change and food security. Science 304, 1623–1627.CrossRefGoogle ScholarPubMed
Lancaster, N. 1982. Linear dunes. Progress in Physical Geography 6, 475–504.CrossRefGoogle Scholar
Lancaster, N. 1989. Star dunes. Progress in Physical Geography 13, 67–91.CrossRefGoogle Scholar
Lancaster, N. 1995. Geomorphology of Desert Dunes. London: Routledge.CrossRefGoogle Scholar
Lancaster, N.2005. Aeolian processes. In Encyclopedia of Geology 4 (ed. Selley, R. C., Cocks, R. L. M., & Plimer, I. R.), pp. 612–627. Oxford: Elsevier.Google Scholar
Lancaster, N.2007. Low-latitude dune fields. In Encyclopedia of Quaternary Sciences (ed. Elias, S. A.). Amsterdam: Elsevier.Google Scholar
Lancaster, N., Kocurek, G., Singhvi, A.et al. 2002. Late Pleistocene and Holocene dune activity and wind regimes in the western Sahara Desert of Mauritania. Geology 30, 991–994.2.0.CO;2>CrossRefGoogle Scholar
Land, L. S., Behrens, E. W., & Frishman, S. A. 1979. The ooids of Baffin Bay, Texas. Journal of Sedimentary Petrology 49, 1269–1278.Google Scholar
Langbein, W. B. 1964. Geometry of river channels. Journal of the Hydraulics Division, ASCE 90, 301–312.Google Scholar
Langbein, W. B., & Leopold, L. B. 1964. Quasi-equilibrium states in channel morphology. American Journal of Science 262, 782–794.CrossRefGoogle Scholar
Langbein, W. B., & Leopold, L. B. 1966. River meanders – theory of minimum variance. United States Geological Survey Professional Paper 422H.
Langford, R. P., & Chan, M. A. 1988. Flood surfaces and deflation surfaces within the Cutler Formation and Cedar Mesa Sandstone (Permian), southeastern Utah. Geological Society of America Bulletin 100, 1541–1549.2.3.CO;2>CrossRefGoogle Scholar
Langford, R. P., & Chan, M. A.1993. Downwind changes within an ancient dune sea, Permian Cedar Mesa Sandstone, southeast Utah. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 109–126. Oxford: Blackwells.CrossRefGoogle Scholar
Larson, G. J., Lawson, D. E., Evenson, E. B.et al. 2006. Glaciohydraulic supercooling in former ice sheets?Geomorphology 75, 20–32.CrossRefGoogle Scholar
Larson, M., & Kraus, N. C. 1991. Numerical model of longshore current for bar and trough beaches. Journal of the Waterways, Harbors and Coastal Engineering Division, ASCE 117, 326–347.CrossRefGoogle Scholar
Lasemi, Z., Boardman, M. R., & Sandberg, P. A. 1989. Cement origin of supratidal dolomite, Andros Island, Bahamas, Journal of Sedimentary Petrology 59, 249–257.Google Scholar
Lauff, G. H. 1967. Estuaries. Washington, D.C.: American Association for the Advancement of Science.Google Scholar
Laval, B., Cady, S. L, Pollack, J. C.et al. 2000. Modern freshwater microbialite analogues for ancient dendritic reef structures. Nature 407, 626–629.Google ScholarPubMed
Lawson, D. E. 1982. Mobilisation, movement and deposition of active subaerial sediment flows, Matanuska Glacier, Alaska. Journal of Geology 90, 279–300.CrossRefGoogle Scholar
Lawson, D. E., Strasser, J. C., Evenson, E. B.et al. 1998. Glaciohydraulic supercooling: a mechanism to create stratified, debris-rich basal ice.1. Field evidence and conceptual model. Journal of Glaciology 44, 547–562.CrossRefGoogle Scholar
Lay, T., Kanamori, H., Ammon, C. J.et al. 2005. The Great Sumatra–Andaman earthquake of 26 December 2004. Science 308, 1127–1133.CrossRefGoogle ScholarPubMed
Leatherman, S. P. 1979. Barrier Islands. New York: Academic Press.Google Scholar
Leatherman, S. P., Williams, A. T., & Fisher, J. S. 1977. Overwash sedimentation associated with a large-scale northeaster. Marine Geology 24, 109–121.CrossRefGoogle Scholar
Leckie, D. A., & Krystinik, L. F. 1989. Is there evidence for geostrophic currents preserved in the sedimentary record of inner to middle-shelf deposits?Journal of Sedimentary Petrology 59, 862–870.Google Scholar
Leclair, S. F. 2002. Preservation of cross-strata due to the migration of subaqueous dunes: an experimental investigation. Sedimentology 49, 1157–1180.CrossRefGoogle Scholar
Leclair, S. F., & Bridge, J. S. 2001. Quantitative interpretation of sedimentary structures formed by river dunes. Journal of Sedimentary Research 71, 713–716.CrossRefGoogle Scholar
Leclair, S. F., Bridge, J. S., & Wang, F. 1997. Preservation of cross-strata due to migration of subaqueous dunes over aggrading and non-aggrading beds: comparison of experimental data with theory. Geoscience Canada 24, 55–66.Google Scholar
Lee, M., Aronson, J. L, & Savin, S. M. 1989. Timing and conditions of Permian Rotliegende sandstone diagenesis, southern North Sea; K/Ar and oxygen isotope data. American Association of Petroleum Geologists Bulletin 73, 195–215.Google Scholar
Leeder, M. R. 1975. Pedogenic carbonates and flood sediment accretion rates: a quantitative method for alluvial arid-zone lithofacies. Geological Magazine 112, 257–270.CrossRefGoogle Scholar
Leeder, M. R.1978. A quantitative stratigraphic model for alluvium with special reference to channel deposit density and interconnectedness. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 587–596. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Leeder, M. R. 1980. On the stability of lower stage plane beds and the absence of current ripples in coarse sands. Journal of the Geological Society of London 137, 423–430.CrossRefGoogle Scholar
Leeder, M. R. 1982. Sedimentology: Process and Product. London: George Allen and Unwin.Google Scholar
Leeder, M. R.1993. Tectonic controls upon drainage basin development, river channel migration and alluvial architecture: implications for hydrocarbon reservoir development and characterization. In Characterization of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 7–22. London: Geological Society of London.Google Scholar
Leeder, M. R. 1999. Sedimentology and Sedimentary Basins: From Turbulence to Tectonics. Oxford: Blackwells.Google Scholar
Leeder, M. R., & Alexander, J. A. 1987. The origin and tectonic significance of asymmetrical meander belts. Sedimentology 34, 217–226.CrossRefGoogle Scholar
Leeder, M. R., Mack, G. H., Peakall, J., & Salyards, S. L. 1996. First quantitative test of alluvial stratigraphy models: southern Rio Grande rift, New Mexico. Geology 24, 87–90.2.3.CO;2>CrossRefGoogle Scholar
Leeder, M. R., & Stewart, M. 1996. Fluvial incision and sequence stratigraphy: alluvial responses to relative base level fall and their detection in the geological record. In Sequence Stratigraphy in British Geology (ed. Hesselbo, S. P. & Parkinson, D. N.), pp. 47–61. London: Geological Society of London.Google Scholar
Lees, A. 1975. Possible influences of salinity and temperature on modern shelf carbonate sedimentation. Marine Geology 19, 159–198.CrossRefGoogle Scholar
Legros, F. 2002. Can dispersive pressure cause inverse grading in grain flows?Journal of Sedimentary Research 72, 166–170.CrossRefGoogle Scholar
Leopold, L. B., & Langbein, W. B. 1962. The concept of entropy in landscape evolution. United States Geological Survey Professional Paper 500A.
Leopold, L. B., & Langbein, W. B. 1966. River meanders. Scientific American 214, 60–70.CrossRefGoogle Scholar
Leopold, L. B., & Wolman, M. G. 1957. River channel patterns: braided, meandering and straight. United States Geological Survey Professional Paper 282-B.
Leopold, L. B., Wolman, M. G., & Miller, J. P. 1964. Fluvial Processes in Geomorphology. San Francisco, CA: W. H. Freeman.Google Scholar
Lerman, A. 1978. Lakes: Chemistry Geology and Physics. New York: Springer-Verlag.CrossRefGoogle Scholar
Lerman, A., Imboden, D., & Gat, J. 1995. Physics and Chemistry of Lakes, 2nd edn. New York: Springer-Verlag.CrossRefGoogle Scholar
Levey, R. A. 1978. Bedform distribution and internal stratification of coarse-grained point bars, Upper Congaree River, South Carolina. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 105–127. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Levinton, J. S. 1982. Marine Ecology. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Li, M. Z., & Amos, C. L. 1999. Sheet flow and large wave ripples under combined waves and currents: field observations, model predictions and effects on boundary layer dynamics. Continental Shelf Research 19, 637–663.CrossRefGoogle Scholar
Lippman, F. 1973. Sedimentary Carbonate Minerals. New York: Springer-Verlag.CrossRefGoogle Scholar
Livingston, D. A. 1963. Chemical composition of rivers and lakes. United States Geological Survey Professional Paper 440-G.
Livingstone, I., & Warren, A. 1996. Aeolian Geomorphology: An Introduction. Harlow: Longman.Google Scholar
Lizitsin, A. P. 1996. Oceanic Sedimentation. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Lliboutry, L. 1979. Local friction laws for glaciers: a critical review and new openings. Journal of Glaciology 23, 67–95.CrossRefGoogle Scholar
Lockley, M. G. 1991. Tracking Dinosaurs: A New Look at an Ancient World. Cambridge: Cambridge University Press.Google Scholar
Lockley, M. G., & Hunt, A. P. 1995. Dinosaur Tracks and Other Fossil Footprints of the Western United States. New York: Columbia University Press.Google Scholar
Logan, B. W. 1987. The MacLeod Evaporite Basin Western Australia. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., Davies, G. R., Read, J. F., & Cebulski, D. E. 1970. Carbonate Sedimentation and Environments, Shark Bay, Western Australia. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., Hoffman, P., & Gebelein, C. D. 1974. Algal mats, crypt-algal fabrics and structures, Hamelin Pool, Western Australia. In Evolution and Diagenesis of Quaternary Carbonate Sequences, Shark Bay, Western Australia (ed. Logan, B. W., Read, J. F., Hagan, G. M.et al.), pp. 140–193. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., Read, J. F., Hagan, G. M.et al. 1974. Evolution and Diagenesis of Quaternary Carbonate Sequences, Shark Bay, Western Australia. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., & Semeniuk, V. 1976. Dynamic Metamorphism: Processes and Products in Devonian Carbonate Rocks, Canning Basin, Western Australia. Sydney: Geological Society of Australia.Google Scholar
Longman, M. W. 1980. Carbonate diagenetic textures from near surface diagenetic environments. American Association of Petroleum Geologists Bulletin 64, 461–487.Google Scholar
Longuet-Higgins, M. S. 1953. Mass transport in water waves. Philosophical Transactions of the Royal Society245A.
Longuet-Higgins, M. S. 1970a. Longshore currents generated by obliquely incident waves, 1. Journal of Geophysical Research 75, 6778–6789.CrossRefGoogle Scholar
Longuet-Higgins, M. S. 1970b. Longshore currents generated by obliquely incident waves, 2. Journal of Geophysical Research 75, 6790–6801.CrossRefGoogle Scholar
Longuet-Higgins, M. S. 1981. Oscillating flow over steep sand ripples. Journal of Fluid Mechanics 107, 1–35.CrossRefGoogle Scholar
Loope, D. B. 1984. Origin of extensive bedding planes in aeolian sandstones: a defence of Stokes' hypothesis. Sedimentology 31, 123–132.CrossRefGoogle Scholar
Loope, D. B. 1985. Episodic deposition and preservation of eolian sands – a Late Paleozoic example from southeastern Utah. Geology 13, 73–76.2.0.CO;2>CrossRefGoogle Scholar
Loucks, R. G., & Sarg, J. F. 1993. Carbonate Sequence Stratigraphy: Recent Developments and Applications. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Lowe, D. R. 1975. Water escape structures in coarse-grained sediments. Sedimentology 22, 157–204.CrossRefGoogle Scholar
Lowe, D. R. 1976a. Subaqueous liquefied and fluidized sediment flows and their deposits. Sedimentology 23, 285–308.CrossRefGoogle Scholar
Lowe, D. R. 1976b. Grain flow and grain flow deposits. Journal of Sedimentary Petrology 46, 188–199.Google Scholar
Lowe, D. R.1979. Sediment gravity flows: their classification and some problems of application to natural flows and deposits. In Geology of Continental Slopes (ed. Doyle, L. J. & Pilkey, O. H.), pp. 75–82. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Lowe, D. R. 1982. Sediment gravity flows: 2. Depositional models with special reference to the deposits of high-density turbidity currents. Journal of Sedimentary Petrology 52, 279–297.Google Scholar
Lowe, D. R., Anderson, K. S., & Braunstein, D. 2001. The zonation and structuring of siliceous sinter around hot springs, Yellowstone National Park, and the role of thermophilic bacteria in its deposition. In Thermophiles: Biodiversity, Ecology, and Evolution (ed. Reysenbach, A.-L., Voytek, M., & Mancinelli, R.), pp. 143–166. New York: Kluwer Academic.CrossRefGoogle Scholar
Lowe, D. R., & Braunstein, D. 2003. Microstructure of high-temperature (> 73 ℃) siliceous sinter deposited around hot springs and geysers, Yellowstone National Park: the role of biological and abiological processes in sedimentation. Canadian Journal of Earth Sciences 40, 1611–1642.CrossRefGoogle Scholar
Lowe, D. R., & Lopiccolo, R. D. 1974. The characteristics and origin of dish and pillar structures. Journal of Sedimentary Petrology 44, 484–501.Google Scholar
Lowenstam, H. A. 1954. Factors affecting the aragonite : calcite ratios in carbonate-secreting marine organisms. Journal of Geology 62, 284–322.CrossRefGoogle Scholar
Lowenstam, H. A., & Weiner, S. 1989. On Biomineralization. Oxford: Oxford University Press.Google Scholar
Lowenstein, T. K. 2002. Pleistocene lakes and paleoclimates (0 to 200 ka) in Death Valley, California. Smithsonian Contributions to the Earth Sciences 33, 109–120.Google Scholar
Lowenstein, T. K., & Hardie, L. A. 1985. Criteria for the recognition of salt-pan evaporites. Sedimentology 32, 627–644.CrossRefGoogle Scholar
Lowenstein, T. K., Hardie, L. A., Timofeeff, M. N., & Demicco, R. V. 2003. Secular variation in seawater chemistry and the origin of calcium chloride basinal brines. Geology 31, 857–860.CrossRefGoogle Scholar
Lowenstein, T. K., Li., J., Brown, C.et al. 1999. 200 k.y. paleoclimate record from Death Valley salt core. Geology 27, 3–6.2.3.CO;2>CrossRefGoogle Scholar
Lowenstein, T. K., Timofeeff, M. N., Brennan, S. T., Hardie, L. A., & Demicco, R. V. 2001. Oscillations in Phanerozoic seawater chemistry: Evidence from fluid inclusions. Science 294, 1086–1088.CrossRefGoogle ScholarPubMed
Lundqvist, J. 1989. Rogen (ribbed) moraine – identification and possible origin. Sedimentary Geology 62, 281–292.CrossRefGoogle Scholar
Lundqvist, J. 1997. Rogen moraine – an example of two-step formation of glacial landscapes. Sedimentary Geology 111, 27–40.CrossRefGoogle Scholar
Lunt, I. A., & Bridge, J. S. 2006. Formation and preservation of open-framework gravel strata in unidirectional water flows. Sedimentology 53, 1–17.Google Scholar
Lyell, C. 1830. The Principles of Geology. London: John Murray.Google Scholar
MacEachern, J. A., Bann, K. L., Bhattacharya, J. P., & Howell, C. D. Jr. 2005. Ichnology of deltas: organism responses to the dynamic interplay of rivers, waves, storms, and tides. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J. P.), pp. 49–85. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Macintyre, I. G. 1985. Submarine cements – the peloidal question. In Carbonate Cements (ed. Schneidermann, N. & Harris, P. M.), pp. 109–116. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Mack, G. H., & James, W. C. 1994. Paleoclimate and global distribution of paleosols. Journal of Geology 102, 360–362.CrossRefGoogle Scholar
Mack, G. H., James, W. C., & Monger, H. C. 1993. Classification of paleosols. Geological Society of America Bulletin 105, 129–136.2.3.CO;2>CrossRefGoogle Scholar
Mack, G. H., & Leeder, M. R. 1998. Channel shifting of the Rio Grande, southern Rio Grande rift: implications for alluvial stratigraphy models. Sedimentary Geology 177, 207–219.CrossRefGoogle Scholar
Mackenzie, F. T., & Pigott, J. D. 1981. Tectonic controls of Phanerozoic sedimentary rock cycling. Journal of the Geological Society of London 138, 183–196.CrossRefGoogle Scholar
Mackey, S. D., & Bridge, J. S. 1995. Three dimensional model of alluvial stratigraphy: theory and application. Journal of Sedimentary Research B65, 7–31.CrossRefGoogle Scholar
Mader, C. L. 1974. Numerical simulation of tsunamis. Journal of Physical Oceanography 4, 74–82.2.0.CO;2>CrossRefGoogle Scholar
Mader, C. L. 1988. Numerical Modeling of Water Waves. Berkeley, CA: University of California Press.Google Scholar
Madsen, O. S. 1977. A realistic model of the wind-induced Ekman boundary layer. Journal of Physical Oceanagraphy 7, 248–255.2.0.CO;2>CrossRefGoogle Scholar
Maizels, J. K. 1989. Sedimentology, paleoflow dynamics and flood history of jokulhlaup deposits: paleohydrology of Holocene sediment sequences in southern Iceland sadur deposits. Journal of Sedimentary Petrology 59, 204–223.Google Scholar
Maizels, J. K.2002. Sediments and landforms of modern proglacial terrestrial environments. In Modern and Past Glacial Environments (ed. Menzies, J.), pp. 279–333. Oxford: Butterworth-Heineman.CrossRefGoogle Scholar
Major, J. J. 1997. Depositional processes in large-scale debris-flow experiments. Journal of Geology 105, 345–366.CrossRefGoogle Scholar
Makaske, B. 2001. Anastomosing rivers: a review of their classification, origin and sedimentary products. Earth-Science Reviews 53, 149–196.CrossRefGoogle Scholar
Makse, H. A., Havlin, S., King, P. R., & Stanley, H. E. 1997. Spontaneous stratification in granular mixtures. Nature 386, 379–382.CrossRefGoogle Scholar
Maliva, R. G., Knoll, A. H., & Simonson, B. M. 2005. Secular change in the Precambrian silica cycle: insights from chert petrology. Geological Society of America Bulletin 117, 835–845.CrossRefGoogle Scholar
Maliva, R. G., & Siever, R. 1989. Nodular chert formation in carbonate rocks. Journal of Geology 97, 421–433.CrossRefGoogle Scholar
Mancini, E. A, Llinas, J. C., Parcell, W. C.et al. 2004. Upper Jurassic thrombolite reservoir play, northeastern Gulf of Mexico. American Association of Petroleum Geologists Bulletin 88, 1573–1602.CrossRefGoogle Scholar
Mann, K. H. 2000. Ecology of Coastal Waters: With Implications for Management. Massachusetts, PA: Blackwell Science.Google Scholar
Mann, S. 2001. Biomineralization: Principles and Concepts in Bioinorganic Materials Chemistry. Oxford: Oxford University Press.Google Scholar
Marani, M., Lanzoni, S., Zandolin, D, Seminara, G., & Rinaldo, A. 2002. Tidal meanders. Water Resources Research 38, 7–1 to 7–11.CrossRefGoogle Scholar
Marr, J. G., Swenson, J. B., Paola, C., & Voller, V. R. 2000. A two-diffusion model of fluvial stratigraphy in closed depositional basins. Basin Research 12, 381–398.CrossRefGoogle Scholar
Marshall, J. F., & Davies, P. J. 1975. High magnesium calcite ooids from the Great Barrier Reef. Journal of Sedimentary Petrology 45, 285–291.Google Scholar
Martini, I. P., Baker, V. R., & Garzon, G. 2002. Flood and Megaflood Processes and Deposits. Oxford: Blackwells.CrossRefGoogle Scholar
Martini, I. P., Brookfield, M. E., & Sadura, S. 2001. Glacial Geomorphology and Geology. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Mastbergen, D. R., & Berg, J. H. 2003. Breaching in fine sands and the generation of sustained turbidity currents in submarine canyons. Sedimentology 50, 625–637.CrossRefGoogle Scholar
Matter, A., & Tucker, M. E. 1978. Modern and Ancient Lake Sediments. Oxford: Blackwells.CrossRefGoogle Scholar
McCabe, M., & Eyles, N. 1988. Sedimentology of an ice-contact glaciomarine delta, Carey Valley, Northern Ireland. Sedimentary Geology 59, 1–14.CrossRefGoogle Scholar
McCabe, P. J. 1984. Depositional environments of coal and coal-bearing strata. In Sedimentology of Coal and Coal-bearing Sequences (ed. Rahmani, R. A. & Flores, R. M.), pp. 13–42. Oxford: Blackwells.Google Scholar
McCabe, P. J., & Parrish, J. T. 1992. Tectonic and climatic controls on the distribution and quality of Cretaceous coals. In Controls on the Distribution and Quality of Cretaceous Coals (ed. McCabe, P. J. & Parrish, J. T.), pp. 1–15. Boulder, CO: Geological Society of America.Google Scholar
McCaffrey, W. D., Kneller, B. C., & Peakall, J. 2001. Particulate Gravity Currents. Oxford: Blackwells.CrossRefGoogle Scholar
McCarthy, P. J., & Plint, A. G. 1998. Recognition of interfluve sequence boundaries: integrating paleopedology and sequence stratigraphy. Geology 26, 387–390.2.3.CO;2>CrossRefGoogle Scholar
McCave, I. N. 1968. Shallow and marginal marine sediments associated with the Catskill complex in the Middle Devonian of New York. In Late Paleozoic and Mesozoic Continental Sedimentation, Northeastern North America (ed. Klein, G. deV.), pp. 75–108. Boulder, CO: Geological Society of America.Google Scholar
McCave, I. N. 1969. Correlation of marine and nonmarine strata with example from Devonian of New York State. American Association of Petroleum Geologists Bulletin 53, 155–162.Google Scholar
McCave, I. N. 1970. Deposition of fine-grained suspended sediment from tidal currents. Journal of Geophysical Research 75, 4151–4159.CrossRefGoogle Scholar
McCave, I. N.1985. Recent shelf clastic sediment. In Sedimentology: Recent Developments and Applied Aspects (ed. Brenchley, P. J. & Williams, B. P. J.), pp. 49–65. London: Geological Society of London.Google Scholar
McCubbin, D. G. 1982. Barrier island and strand plain facies. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D. R.), pp. 247–279. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
McHargue, T. R. 1991. Seismic facies, processes and evolution of Miocene inner fan channels, Indus submarine fan. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 403–414. New York: Springer-Verlag.CrossRefGoogle Scholar
McIlreath, I. A. 1977. Accumulation of a Middle Cambrian deep water, basinal limestone adjacent to a vertical, submarine carbonate escarpment, southern Rocky Mountains. In Deep Water Carbonate Environments (ed. Cook, H. E. & Enos, P.), pp. 113–124. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
McIlreath, I. A., & James, N. P. 1984. Carbonate slopes. In Facies Models, 2nd edn. (ed. Walker, R. G.), pp. 245–257. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
McIlreath, I. A., & Morrow, D. W. 1990. Diagenesis. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
McKee, E. D. 1966. Structure of dunes at White Sands National Monument, New Mexico. Sedimentology 7, 3–69.CrossRefGoogle Scholar
McKee, E. D.1979. A study of global sand seas. United States Geological Survey Professional Paper 1052.
McKee, E. D.1982. Sedimentary structures in dunes of the Namib Desert, South West Africa. Geological Society of America Special Paper 188.CrossRef
McKee, E. D., & Gutschick, R. C. 1969. History of the Redwall Limestone of Northern Arizona. Boulder, CO: Geological Society of America.CrossRefGoogle Scholar
McKee, E. D., & Tibbitts, G. C. Jr. 1964. Primary structures of a seif dune and associated deposits in Libya. Journal of Sedimentary Petrology 34, 5–17.Google Scholar
McKee, E. D., & Weir, G. W. 1953. Terminology for stratification and cross-stratification in sedimentary rocks. Geological Society of America Bulletin 64, 381–390.CrossRefGoogle Scholar
McKenzie, J. A., Hsü, K. J., & Schneider, J. F. 1980. Movement of subsurface waters under the sabkha, Abu Dhabi, UAE, and its relation to evaporative dolomite genesis. In Concepts and Models of Dolomitization (ed. Zenger, D. H., Dunham, J. B., & Ethington, R. L.), pp. 11–30. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
McLean, S. R. 1990. The stability of ripples and dunes. Earth-Science Reviews 29, 131–144.CrossRefGoogle Scholar
McNeill, J. R., & Winiwarter, V. 2004. Breaking the sod: humankind, history, and soil. Science 304, 1627–1629.CrossRefGoogle Scholar
McPhee, J. 1989. Control of Nature. New York: Farrar, Straus and Giroux.Google Scholar
Mehta, A. J. 1989. On estuarine cohesive sediment suspension behaviour. Journal of Geophysical Research 94, 14,303–14,314.CrossRefGoogle Scholar
Melim, L. A., & Swart, P. K. 2001. Meteoric and marine-burial diagenesis in the subsurface of Great Bahama Bank. In Subsurface Geology of a Prograding Carbonate Platform Margin, Great Bahama Bank: Results of the Bahamas Drilling Project (ed. Ginsburg, R. N.), pp. 137–161. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Melosh, H. J. 1987. The mechanics of large rock avalanches. In Debris Flows/Avalanches: Processes, Recognition, and Mitigation (ed. Costa, J. E. & Wieczorek, G. F.), pp. 41–49. Boulder, CO: Geological Society of America.Google Scholar
Menzies, J. 1995. Modern Glacial Environments – Processes, Dynamics and Sediments. Oxford: Butterworth-Heinemann.Google Scholar
Menzies, J. 1996. Past Glacial Environments – Sediments, Forms and Techniques. Oxford: Butterworth-Heinemann.Google Scholar
Menzies, J. 2002. Modern and Past Glacial Environments. Oxford: Butterworth-Heinemann.CrossRefGoogle Scholar
Menzies, J., & Rose, J. 1989. Subglacial bedforms – drumlins, Rogen moraine and associated subglacial bedforms. Sedimentary Geology 62, 117–407.CrossRefGoogle Scholar
Meybeck, M. 1995. Global distribution of lakes. In Physics and Chemistry of Lakes, 2nd edn. (ed. Lerman, A., Imboden, D., & Gat, J.), pp. 1–35. New York: Springer-Verlag.CrossRefGoogle Scholar
Meyer, G. A., Wells, S. G., Balling, R. C., & Jull, A. J. T. 1992. Response of alluvial systems to fire and climate change in Yellowstone National Park. Nature 357, 147–150.CrossRefGoogle Scholar
Meyers, W. J. 1974. Carbonate cement stratigraphy of the Lake Valley Formation (Mississippian), Sacramento Mountains, New Mexico. Journal of Sedimentary Petrology 44, 837–861.Google Scholar
Meyers, W. J.1991. Calcite cement stratigraphy: an overview. In Luminescence Microscopy: Quantitative and Qualitative Analysis (ed. Barker, C. E. & Kipp, O. C.), pp. 133–148. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Miall, A. D. 1986. Eustatic sea level changes interpreted from seismic stratigraphy: a critique of the methodology with particular reference to the North Sea Jurassic record. American Association of Petroleum Geologists Bulletin 70, 131–137.Google Scholar
Miall, A. D. 1991. Stratigraphic sequences and their chronostratigraphic correlation. Journal of Sedimentary Petrology 61, 497–505.Google Scholar
Miall, A. D. 1996. The Geology of Fluvial Deposits. New York: Springer-Verlag.Google Scholar
Miall, A. D. 1997. The Geology of Stratigraphic Sequences. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Miall, A. D., & Smith, N. D. 1989. Rivers and Their Deposits. Slide Set No. 4. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Mickelson, D. M., & Attig, J. W. 1999. Glacial processes past and present. Geological Society of America Special Paper 337.
Middleton, G. V. 1966a. Experiments on density and turbidity currents. 1: motion of the head. Canadian Journal of Earth Sciences 3, 523–546.Google Scholar
Middleton, G. V. 1966b. Experiments on density and turbidity currents. 2: uniform flow of turbidity currents. Canadian Journal of Earth Sciences 3, 627–637.CrossRefGoogle Scholar
Middleton, G. V. 1966c. Experiments on density and turbidity currents. 3: deposition of sediment. Canadian Journal of Earth Sciences 4, 475–505.CrossRefGoogle Scholar
Middleton, G. V. 1993. Sediment deposition from turbidity currents. Annual Review of Earth and Planetary Sciences 21, 89–114.CrossRefGoogle Scholar
Middleton, G. V. 2003. Encyclopedia of Sediments and Sedimentary Rocks. Dordrecht: Kluwer Academic Publishers.Google Scholar
Middleton, G. V., & Hampton, M. A. 1973. Sediment gravity flows: mechanics of flow and deposition. In Turbidites and Deep-Water Sedimentation (ed. Middleton, G. V. & Bouma, A. H.), pp. 1–38. Los Angeles, CA: SEPM (Society for Sedimentary Geology) Pacific Section.Google Scholar
Middleton, G. V., & Hampton, M. A.1976. Subaqueous sediment transport and deposition by sediment gravity flows. In Marine Sediment Transport and Environmental Management (ed. Stanley, D. J. & Swift, D. J. P.), pp. 197–218. New York: Wiley.Google Scholar
Middleton, G. V., & Southard, J. B. 1984. Mechanics of Sediment Movement. Tulsa, OK: SEPM.CrossRefGoogle Scholar
Middleton, G. V., & Wilcock, P. R. 1994. Mechanics in the Earth and Environmental Sciences. Cambridge: Cambridge University Press.Google Scholar
Miller, J. M. G. 1996. Glacial sediments. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 454–484. Oxford: Blackwells.Google Scholar
Miller, M. C., & Komar, P. D. 1980. Oscillation ripples generated by laboratory experiments. Journal of Sedimentary Petrology 50, 173–182.Google Scholar
Milliman, J. D., & Barretto, H. T. 1975. Relic magnesian calcite oolite and subsidence of the Amazon shelf. Sedimentology 22, 137–145.CrossRefGoogle Scholar
Mitchell, R. W. III. 1985. Comparative sedimentology of shelf carbonates of the Middle Ordovician St. Paul Group of the central Appalachians. Sedimentary Geology 43, 1–41.CrossRefGoogle Scholar
Mohrig, D., Heller, P. L., Paola, C., & Lyons, W. J. 2000. Interpreting avulsion process from ancient alluvial sequences: Guadalope–Matarranya system (northern Spain) and Wasatch Formation (western Colorado). Geological Society of America Bulletin 112, 1787–1803.2.0.CO;2>CrossRefGoogle Scholar
Mojzsis, S. J., Arrhenius, G., McKeegan, K. D.et al. 1996. Evidence for life on Earth before 3,800 million years ago. Nature 384, 55–59.CrossRefGoogle ScholarPubMed
Moller, P. 2006. Rogen moraine: an example of glacial reshaping of pre-existing landforms. Quaternary Science Reviews 25, 362–389.CrossRefGoogle Scholar
Molnia, B. F. 1983. Glacial–Marine Sedimentation. New York: Plenum.CrossRefGoogle Scholar
Montañez, I. P., Gregg, J. M., & Shelton, K. L. 1997. Basin-wide Diagenetic Patterns: Integrated Petrologic, Geochemical, and Hydrologic Considerations. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Montgomery, D. R., & Dietrich, W. E. 1989. Source areas, drainage density, and channel initiation. Water Resources Research 25, 1907–1918.CrossRefGoogle Scholar
Montgomery, D. R., & Dietrich, W. E.1994. Landscape dissection and drainage area-slope thresholds. In Process Models and Theoretical Geomorphology (ed. Kirkby, M. J.), pp. 221–246. Chichester: Wiley.Google Scholar
Mora, C. I., & Driese, S. G. 1993. A steep, mid- to late Paleozoic decline in atmospheric CO2: evidence from soil carbonate CO2 barometer. Chemical Geology 107, 217–219.CrossRefGoogle Scholar
Mora, C. I., & Driese, S. G.1999. Palaeoenvironment, palaeoclimate and stable carbon isotopes of Palaeozoic red-bed palaeosols, Appalachian Basin, USA and Canada. In Palaeoweathering, Palaeosurfaces and Related Continental Deposits (ed. Thiry, M. & Simon-Coicon, R.), pp. 61–84. Oxford: Blackwells.Google Scholar
Mora, C. I., Driese, S. G., & Colarusso, L. A. 1996. Middle to late Paleozoic atmospheric CO2 levels from soil carbonate and organic matter. Science 271, 1105–1107.CrossRefGoogle Scholar
Morad, S., Ketzer, J. M., & Ros, L. F. 2000. Spatial and temporal distribution of diagenetic alterations in siliciclastic rocks: implications for mass transfer in sedimentary basins. Sedimentology 47, 95–120.CrossRefGoogle Scholar
Morgan, J. P. 1970. Deltaic Sedimentation: Modern and Ancient. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Morgan, J. P., Coleman, J. M., & Gagliano, S. M. 1968. Mudlumps: diapiric structures in Mississippi delta sediments. In Diapirism and Diapirs (ed. Braunstein, J. & O'Brien, G. D.), pp. 145–161. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Moore, A., Nishimura, Y., Gelfenbaum, G., Kamataki, T., & Triyono, R. 2006. Sedimentary deposits of the 26 December 2004 tsunami on the northwest coast of Aceh, Indonesia. Earth Planets Space 58, 253–258.CrossRefGoogle Scholar
Morozova, G. S., & Smith, N. D. 1999. Holocene avulsion history of the lower Saskatchewan fluvial system, Cumberland Marshes, Saskatchewan–Manitoba, Canada. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 231–249. Oxford: Blackwells.CrossRefGoogle Scholar
Morris, P. E., & Williams, D. J. 1997. Exponential longitudinal profiles of streams. Earth Surface Processes and Landforms 22, 143–163.3.0.CO;2-Z>CrossRefGoogle Scholar
Morris, P. E., & Williams, D. J. 1999a. A worldwide correlation for exponential bed particle size variations in subaerial aqueous flows. Earth Surface Processes and Landforms 24, 835–847.3.0.CO;2-G>CrossRefGoogle Scholar
Morris, P. E., & Williams, D. J. 1999b. Worldwide correlations for subaerial aqueous flows with exponential longitudinal profiles. Earth Surface Processes and Landforms 24, 867–879.3.0.CO;2-L>CrossRefGoogle Scholar
Morse, J. W., Millero, F. J., Thurmond, V., Brown, E., & Ostlund, H. G. 1984. The carbonate chemistry of Grand Bahama Bank waters: after 18 years another look. Journal of Geophysical Research 89, 3604–3614.CrossRefGoogle Scholar
Morse, J. W., Wang, Q., & Tzio, M. Y. 1997. Influences of temperature and Mg : Ca ratio on CaCO3 precipitates from seawater. Geology 25, 85–87.2.3.CO;2>CrossRefGoogle Scholar
Morton, R. A. 1981. Formation of storm deposits by wind-forced currents in the Gulf of Mexico and the North Sea. In Holocene Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 385–396. Oxford: Blackwells.CrossRefGoogle Scholar
Moslow, T. F., & Heron, S. D. 1978. Relict inlets: preservation and occurrence in the Holocene stratigraphy of southern Core Banks, North Carolina. Journal of Sedimentary Petrology 48, 1275–1286.Google Scholar
Moslow, T. F., & Tye, R. S. 1985. Recognition and characterization of Holocene tidal inlet sequences. Marine Geology 63, 129–151.CrossRefGoogle Scholar
Mountney, N. P. 2006. Periodic accumulation and destruction of aeolian erg sequences in the Permian Cedar Mesa Sandstone, White Canyon, southern Utah, USA. Sedimentology 53, 789–823.CrossRefGoogle Scholar
Mountney, N. P., & Jagger, A. 2004. Stratigraphic evolution of an aeolian erg margin system: the Permian Cedar Mesa sandstone, SE Utah, USA. Sedimentology 51, 713–743.CrossRefGoogle Scholar
Mulder, T., & Syvitski, J. P. M. 1995. Turbidity currents generated at river mouths during exceptional discharges to the world's oceans. Journal of Geology 103, 285–299.CrossRefGoogle Scholar
Mullins, H. T. 1983. Modern carbonate slopes and basins of the Bahamas. In Platform Margin and Deep Water Carbonates (ed. Cook, H. E., Hine, A. C., & Mullins, H. T.), pp. 4.1–4.138. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Multer, H. G. 1975. Field Guide to Some Carbonate Rock Environments, Florida Keys and Western Bahamas. Madison, NJ: Farleigh Dickinson University.Google Scholar
Multer, H. G., & Hoffmeister, J. E. 1975. Petrology and significance of the Key Largo (Pleistocene) Limestone, Florida Keys. In Field Guide to some Carbonate Rock Environments; Florida Keys and Western Bahamas (ed. Multer, H. G.) pp. 111–112. Madison, NJ: Farleigh Dickinson University.Google Scholar
Murray, A. B., & Paola, C. 1994. A cellular model of braided rivers. Nature 371, 54–57.CrossRefGoogle Scholar
Murray, A. B., & Paola, C. 1997. Properties of a cellular braided stream model. Earth Surface Processes and Landforms 22, 1001–1025.3.0.CO;2-O>CrossRefGoogle Scholar
Murray, P. B., Davies, A. G., & Soulsby, R. L. 1991. Sediment pick-up in wave and current flows. In Sand Transport in Rivers, Estuaries and the Sea (ed. Soulsby, R. L. & Bettess, R.), pp. 37–44. Rotterdam: Balkema.Google Scholar
Murray, S. P. 1970. Bottom currents near the coast during hurricane Camille. Journal of Geophysical Research 75, 4579–4582.CrossRefGoogle Scholar
Murton, J. B., Peterson, R., & Ozouf, J.-C. 2006. Bedrock fracture by ice segregation in cold regions. Science 314, 1127–1129.CrossRefGoogle ScholarPubMed
Mutti, E. 1985. Turbidite systems and their relations to depositional sequences. In Provenance of Arenites (ed. Zuffa, G. G.), pp. 65–93. Amsterdam: Reidel.CrossRefGoogle Scholar
Mutti, E., & Normark, W. R. 1987. Comparing examples of modern and ancient turbidite systems: problems and concepts. In Marine Clastic Sedimentology: Concepts and Case Studies (ed. Legget, J. K. & Zuffa, G. G.), pp. 1–38. London: Graham and Trotman.CrossRefGoogle Scholar
Mutti, E., & Normark, W. R.1991. An integrated approach to the study of turbidite systems. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 75–106. New York: Springer-Verlag.CrossRefGoogle Scholar
Mutti, E., & Lucci, Ricci F. 1972. Le torbiditi dell'Appennino Settentrionale: introduzione all'analisi di facies. Memorie della Società Geologica Italiana 11, 161–199.Google Scholar
Myrow, P. M., & Southard, J. B. 1991. Combined-flow model for vertical stratification sequences in shallow marine storm-deposited beds. Journal of Sedimentary Petrology 61, 202–210.Google Scholar
Nardin, T. R., Hein, F. J., Gorsline, D. S., & Edwards, B. D. 1979. A review of mass movement processes, sediment and acoustic characteristics, and contrasts in slope and base-of-slope versus canyon–fan–basin floor systems. In Geology of Continental Slopes (ed. Doyle, L. J. & Pilkey, O. H.), pp. 61–73. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Nealson, K. H. 1997. Sedimentary bacteria: who's there, what are they doing, and what's new?Annual Review of Earth and Planetary Sciences 25, 403–434.CrossRefGoogle ScholarPubMed
Necker, F., Hartel, C., Kleiser, L., & Meiburg, E. 2005. Mixing and dissipation in particle-driven gravity currents. Journal of Fluid Mechanics 545, 339–372.CrossRefGoogle Scholar
Neev, D., & Emery, K. O. 1967. The Dead Sea: depositional processes and environments of evaporites. Israel Geological Survey Bulletin 41.Google Scholar
Nemec, W., & Steel, R. J. 1988. Fan Deltas: Sedimentology and Tectonic Settings. London: Blackie.Google Scholar
Nezu, I., & Nakagawa, H. 1993. Turbulence in Open-Channel Flows. Rotterdam: Balkema.Google Scholar
Nguyen, Q. D., & Boger, D. V. 1992. Measuring the flow properties of yield stress fluids. Annual Review of Fluid Mechanics 24, 47–88.CrossRefGoogle Scholar
Nichols, G. 1999. Sedimentology and Stratigraphy. Oxford: Blackwells.Google Scholar
Nichols, M. J. 1989. Sediment accumulation rates and relative sea level rise in lagoons. Marine Geology 88, 201–220.CrossRefGoogle Scholar
Nichols, R. J., Sparks, R. S. J., & Wilson, C. J. N. 1994. Experimental studies of the fluidization of layered sediments and the formation of fluid escape structures. Sedimentology 41, 233–253.CrossRefGoogle Scholar
Nickling, W. G. 1994. Aeolian sediment transport and deposition. In Sediment Transport and Depositional Processes (ed. Pye, K.), pp. 293–350. Oxford: Blackwells.Google Scholar
Nielsen, P. 1992. Coastal Bottom Boundary Layers and Sediment Transport. Singapore: World Scientific.CrossRefGoogle Scholar
Nielson, J., & Kocurek, G. 1986. Climbing zibars in the Algodones. Sedimentary Geology 48, 1–15.CrossRefGoogle Scholar
Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E. 1981. Holocene Sedimentation in the North Sea Basin. Oxford: Blackwells.CrossRefGoogle Scholar
Nio, S. D., Seigenthaler, C., & Yang, C. S. 1983. Megaripple cross-bedding as a tool for the reconstruction of the palaeohydraulics in a Holocene subtidal environment, S.W. Netherlands. Geologie en Mijnbouw 62, 499–510.Google Scholar
Nio, S. D., & Yang, C. S. 1991. Diagnostic attributes of clastic tidal deposits: a review. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 3–28. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Nittrouer, C. A., & Wright, L. D. 1994. Transport of particles across continental shelves. Reviews of Geophysics 32, 85–113.CrossRefGoogle Scholar
Noffke, N., Gerdes, G., & Klenke, T. 2005. Benthic cyanobacteria and their influence on the sedimentary dynamics of peritidal depositional systems (siliciclastic, evaporitic salty, and evaporitic carbonatic). Earth-Science Reviews 62, 163–176.CrossRefGoogle Scholar
Normark, W. R. 1970. Growth patterns of deep-sea fans. American Association of Petroleum Geologists Bulletin 54, 2170–2195.Google Scholar
Normark, W. R. 1978. Fan valleys, channels, and depositional lobes on modern submarine fans: characters for recognition of sandy turbidite environments. American Association of Petroleum Geologists Bulletin 62, 912–931.Google Scholar
Normark, W. R., & Piper, D. J. W. 1985. Navy Fan, Pacific Ocean. In Submarine Fans and Related Turbidite Systems (ed. Bouma, A. H., Normark, W. R., & Barnes, N. E.), pp. 87–94. New York: Springer-Verlag.CrossRefGoogle Scholar
Normark, W. R., Piper, D. J. W., & Hess, G. R. 1979. Distributary channels, sand lobes, and mesotopography of Navy submarine fan, California Borderland, with applications to ancient fan sediments. Sedimentology 26, 749–774.CrossRefGoogle Scholar
North, C. P. 1996. The prediction and modelling of subsurface fluvial stratigraphy. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 395–508. Chichester: Wiley.Google Scholar
North, C. P., & Prosser, D. J. 1993. Characterisation of Fluvial and Aeolian Reservoirs. London: Geological Society of London.Google Scholar
North, C. P., & Taylor, K. S. 1996. Ephemeral–fluvial deposits: integrated outcrop and simulation studies reveal complexity. American Association of Petroleum Geologists Bulletin 80, 811–830.Google Scholar
Notholt, A. J. G., Sheldon, R. P., & Davidson, D. F. 1989. Phosphate Deposits of the World, Volume 2, Phosphate Rock Resources. Cambridge: Cambridge University Press.Google Scholar
Nottvedt, A., & Kreisa, R. D. 1987. Model for the combined flow origin of hummocky cross-stratification. Geology 15, 357–361.2.0.CO;2>CrossRefGoogle Scholar
Nummedal, D., & Penland, S. 1981. Sediment dispersal in Norderneyer Seegate, West Germany. In Holocene Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 187–210. Oxford: Blackwells.CrossRefGoogle Scholar
Nummedal, D., Pilkey, O. H., & Howard, J. D. 1987. Sea-Level Fluctuations and Coastal Evolution. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Nummedal, D., & Swift, D. J. P. 1987. Transgressive stratigraphy at sequence-bounding unconformities: some principles derived from Holocene and Cretaceous examples. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 241–260. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Nye, J. F. 1957. The distribution of stress and velocity in glaciers and ice sheets. Proceedings of the Royal Society of London 239A, 113–133.CrossRefGoogle Scholar
Nye, J. F. 1965. The flow of a glacier in a channel of rectangular, elliptic or parabolic cross-section. Journal of Glaciology 5, 661–690.CrossRefGoogle Scholar
Oertel, G. F. 1979. Barrier island development during the Holocene recession, SE United States. In Barrier Islands (ed. Leatherman, S. P.), pp. 273–290. New York: Academic Press.Google Scholar
Oertel, G. F. 1985. The barrier island system. Marine Geology 63, 1–18.CrossRefGoogle Scholar
Oertel, G. F.1988. Processes of sediment exchange between tidal inlets, ebb deltas and barrier islands. In Hydrodynamics and Sediment Dynamics of Tidal Inlets (ed. Aubrey, D. G. & Weishar, L.), pp. 297–318. Berlin: Springer-Verlag.Google Scholar
Oertel, G. F., Kearney, M. S., Leatherman, S. P., & Woo, J. 1989. Anatomy of a barrier platform: outer barrier lagoon, southern Delmarva Peninsula, Virginia. Marine Geology 88, 303–318.CrossRefGoogle Scholar
Ohfuji, H., & Rickard, D. 2005. Experimental synthesis of framboids – a review. Earth-Science Reviews 71, 147–170.CrossRefGoogle Scholar
Ohmoto, H., & Skinner, B. J. 1983. The Kuroko and Related Volcanogenic Massive Sulfide Deposits. New Haven, CT: The Economic Geology Publishing Company.Google Scholar
Olcott, A. N., Sessions, A. L., Corsetti, F. A., Kaufman, A. J., & Oliviera, T. F. 2005. Biomarker evidence for photosynthesis during Neoproterozoic glaciation. Science 310, 471–473.CrossRefGoogle ScholarPubMed
Oliver, J. 1986. Fluids expelled tectonically from orogenic belts: their role in hydrocarbon migration and other geological phenomena. Geology 14, 99–102.2.0.CO;2>CrossRefGoogle Scholar
Olsen, H. 1990. Astronomical forcing of meandering river behaviour: Milankovitch cycles in Devonian of East Greenland. Palaeogeography, Palaeoclimatology, Palaeoecology 79, 99–115.CrossRefGoogle Scholar
Olsen, H.1994. Orbital forcing on continental depositional systems – lacustrine and fluvial cyclicity in the Devonian of East Greenland. In Orbital Forcing and Cyclic Sequences (ed. Boer, P. L. & Smith, D. G.), pp. 429–438. Oxford: Blackwells.CrossRefGoogle Scholar
Olsen, P. E. 1986. A 40-million-year lake record of Early Mesozoic orbital climatic forcing. Science 234, 842–848.CrossRefGoogle ScholarPubMed
Oltman-Shay, J., & Guza, R. T. 1987. Infragravity edge wave observations on two California beaches. Journal of Physical Oceanography 17, 644–663.2.0.CO;2>CrossRefGoogle Scholar
Open University 1989. Ocean Circulation. Oxford: Pergamon Press.
Orton, G. J. 1996. Volcanic environments. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 485–567. Oxford: Blackwells.Google Scholar
Orton, G. J., & Reading, H. G. 1993. Variability of deltaic processes in terms of sediment supply, with particular emphasis on grain size. Sedimentology 40, 475–512.CrossRefGoogle Scholar
Osborne, P. D., & Greenwood, B. 1993. Sediment suspension under waves and currents: time scales and vertical structure. Sedimentology 40, 599–622.CrossRefGoogle Scholar
Osgood, R. G. Jr. 1987. Trace fossils. In Fossil Invertebrates (ed. Boardman, R. S., Cheetham, A. H., & Rowell, A. J.), pp. 663–674. Oxford: Blackwell Scientific.Google Scholar
Otvos, E. G., & Price, W. A. 1979. Problems of chenier genesis and terminology – an overview. Marine Geology 31, 251–263.CrossRefGoogle Scholar
Ouchi, S. 1985. Response of alluvial rivers to slow active tectonic movement. Geological Society of America Bulletin, 96, 504–515.2.0.CO;2>CrossRefGoogle Scholar
Overeem, I., Syvitski, J. P. M., & Hutton, E. W. H. 2005. Three-dimensional numerical modeling of deltas. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J. P.), pp. 13–30. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Owen, G. 1996. Experimental soft-sediment deformation: structures formed by the liquefaction of unconsolidated sands and some ancient examples. Sedimentology 43, 279–293.CrossRefGoogle Scholar
Paola, C. 2000. Quantitative models of sedimentary basin filling. Sedimentology 47 (Supplement 1), 121–178.CrossRefGoogle Scholar
Paola, C., & Borgman, L. 1991. Reconstructing random topography from preserved stratification. Sedimentology 38, 553–565.CrossRefGoogle Scholar
Paola, C., Heller, P. L., & Angevine, C. L. 1992. The large-scale dynamics of grain-size variation in alluvial basins. 1 – theory. Basin Research 4, 73–90.CrossRefGoogle Scholar
Paola, C., Parker, G., Mohrig, D. C., & Whipple, K. X. 1999. The influence of transport fluctuations on spatially averaged topography on a sandy, braided fluvial plain. In Numerical Experiments in Stratigraphy: Recent Advances in Stratigraphic and Sedimentologic Computer Simulations (ed. Harbaugh, J. W., Watney, W. L., Rankey, E. C.et al.), pp. 211–218. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Pantin, H. M. 1979. Interaction between velocity and effective density in turbidity flow: phase plane analysis, with criteria for autosuspension. Marine Geology 31, 59–99.CrossRefGoogle Scholar
Pantin, H. M.2001. Experimental evidence for autosuspension. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 189–205. Oxford: Blackwells.CrossRefGoogle Scholar
Pantin, H. M., & Leeder, M. R. 1987. Reverse flow in turbidity currents: the role of internal solitons. Sedimentology 34, 1143–1155.CrossRefGoogle Scholar
Parker, G. 1978a. Self-formed straight rivers with equilibrium banks and mobile bed. 1. The sand–silt river. Journal of Fluid Mechanics 89, 109–126.CrossRefGoogle Scholar
Parker, G. 1978b. Self-formed straight rivers with equilibrium banks and mobile bed. 2. The gravel river. Journal of Fluid Mechanics 89, 127–146.CrossRefGoogle Scholar
Parker, G. 1979. Hydraulic geometry of active gravel rivers. Journal of the Hydraulic Division, ASCE 105, 1185–1201.Google Scholar
Parker, G., Fukushima, Y., & Pantin, H. M. 1986. Self-accelerating turbidity currents. Journal of Fluid Mechanics 171, 145–181.CrossRefGoogle Scholar
Parker, R. S. 1977. Experimental study of drainage basin evolution and its hydrologic implications. Colorado State University, Fort Collins, Colorado, Hydrology Paper 90.
Parrish, J. T. 1993. Climate of the supercontinent Pangea. Journal of Geology 101, 215–233.CrossRefGoogle Scholar
Parrish, J. T. 1998. Interpreting Pre-Quaternary Climate from the Geological Record. New York: Columbia University Press.Google Scholar
Parrish, J. T., & Barron, E. J. 1986. Paleoclimates and Economic Geology. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Paterson, W. S. B. 1994. The Physics of Glaciers, 3rd edn. Oxford: Pergamon.Google Scholar
Patterson, R. J., & Kinsman, D. J. J. 1982. Formation of diagenetic dolomite in coastal sabkha along Arabian (Persian) Gulf. American Association of Petroleum Geologists Bulletin 66, 28–43.Google Scholar
Paul, E. A., & Clark, F. E. 1989. Soil Microbiology and Biochemistry. San Diego, CA: Academic Press.Google Scholar
Paul, M. A., & Eyles, N. 1990. Constraints on the preservation of diamict facies (melt-out tills) at the margins of stagnant glaciers. Quaternary Science Reviews 9, 51–69.CrossRefGoogle Scholar
Paxton, S. T., Szabo, J. O., Ajdukiewicz, J. M., & Klimentidis, R. E. 2002. Construction of an intergranular volume compaction curve for evaluating and predicting compaction and porosity loss in rigid-grain sandstone reservoirs. American Association of Petroleum Geologists Bulletin 86, 2047–2067.Google Scholar
Peakall, J. 1998. Axial river evolution in response to half-graben faulting: Carson River, Nevada. Journal of Sedimentary Research 68, 788–799.CrossRefGoogle Scholar
Peakall, J., Felix, M., McCaffrey, B., & Kneller, B. 2001. Particulate gravity currents: perspectives. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 1–8. Oxford: Blackwells.CrossRefGoogle Scholar
Peakall, J., Leeder, M., Best, J., & Ashworth, P. 2000a. River response to lateral ground tiltng: a synthesis and some implications for the modelling of alluvial architecture in extensional basins. Basin Research 12, 413–424.CrossRefGoogle Scholar
Peakall, J., McCaffrey, B., & Kneller, B. 2000b. A process model for the evolution, morphology, and architecture of sinuous submarine channels. Journal of Sedimentary Research 70, 434–448.CrossRefGoogle Scholar
Pemberton, S. G., MacEachern, J. A., & Frey, R. W. 1992. Trace fossil facies models: environmental and allostratigraphic significance. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 47–72. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Penland, S., Boyd, R., & Suter, J. R. 1988. Transgressive depositional systems of the Mississippi delta plain: a model for barrier shoreline and shelf sand development. Journal of Sedimentary Petrology 58, 932–949.Google Scholar
Pennisi, E. 2004. The secret life of fungi. Science 304, 1620–1622.CrossRefGoogle ScholarPubMed
Pérez-Arlucea, M., & Smith, N. D. 1999. Depositional patterns following the 1870's avulsion of the Saskatchewan River (Cumberland Marshes, Saskatchewan). Journal of Sedimentary Research 69, 62–73.CrossRefGoogle Scholar
Perillo, G. M. E. 1995. Geomorphology and Sedimentology of Estuaries. Amsterdam: Elsevier.Google Scholar
Pernetta, J. 1994. Atlas of the Oceans. London: Mitchell Beazly.Google Scholar
Person, M., & Garven, G. 1992. Hydrologic constraints on petroleum generation within continental rift basins: theory and application to the Rhine Graben. American Association of Petroleum Geologists Bulletin 76, 468–488.Google Scholar
Person, M., Raffensperger, J. P., Ge, S., & Garven, G. 1996. Basin-scale hydrogeologic modeling. Reviews of Geophysics 34, 61–87.CrossRefGoogle Scholar
Peryt, T. 1983. Coated Grains. New York: Springer-Verlag.CrossRefGoogle Scholar
Peterson, M. N. A., & der Borch, C. C. 1965. Chert: modern inorganic deposition in a carbonate-precipitating locality. Science 149, 1501–1503.CrossRefGoogle Scholar
Petit, J. R., Jouzel, J., Raynaud, D.et al. 1999. Climate and atmospheric history of the past 420,000 years from the Vostock ice core, Antarctica. Nature 399, 429–436.CrossRefGoogle Scholar
Pettijohn, F. J. 1975. Sedimentary Rocks, 3rd edn. New York: Harper and Row.Google Scholar
Pettijohn, F. J., & Potter, P. E. 1964. Atlas and Glossary of Primary Sedimentary Structures. New York: Springer-Verlag.CrossRefGoogle Scholar
Pettijohn, F. J., Potter, P. E., & Siever, R. 1972. Sand and Sandstone. New York: Springer-Verlag.Google Scholar
Petts, G., & Calow, P. 1996. River Flows and Channel Forms. Oxford: Blackwell Science Limited.Google Scholar
Phillips, O. M. 1991. Flow and Reactions in Permeable Rocks. Cambridge: Cambridge University Press.Google Scholar
Pickard, G. L., & Emery, W. J. 1990. Descriptive Physical Oceanography, 5th edn. Oxford: Pergamon Press.Google Scholar
Pickering, K. T., & Hiscott, R. N. 1985. Contained (reflected) turbidity currents from the Middle Ordovician Cloridorme Formation, Quebec, Canada: an alternative to the antidune hypothesis. Sedimentology 32, 373–394.CrossRefGoogle Scholar
Pickering, K. T., Hiscott, R. N., & Hein, F. J. 1989. Deep Marine Environments: Clastic Sedimentation and Tectonics. London: Unwin Hyman.Google Scholar
Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, Ricci F., & Smith, R. D. A. 1995. Atlas of Deep Water Environments: Architectural Style in Turbidite Systems. London: Chapman and Hall.CrossRefGoogle Scholar
Pickering, K. T., Soh, W., & Tiara, A. 1991. Scale of tsunami-generated sedimentary structures in deep water. Journal of the Geological Society of London 148, 211–214.CrossRefGoogle Scholar
Pickering, K. T., Stow, D. A. V., Watson, M. P., & Hiscott, R. N. 1986. Deep-water facies, processes and models: a review and classification scheme for modern and ancient sediments. Earth-Science Reviews 23, 75–174.CrossRefGoogle Scholar
Pierson, T. C. 1981. Dominant particle support mechanisms in debris flows at Mt Thomas, New Zealand, and implications for flow mobility. Sedimentology 28, 49–60.CrossRefGoogle Scholar
Pierson, T. C. 1995. Flow characteristics of large eruption-triggered debris flows at snow-clad volcanoes: constraints for debris-flow models. Journal of Volcanology and Geothermal Research 66, 283–294.CrossRefGoogle Scholar
Pierson, T. C., & Scott, K. M. 1985. Downstream dilution of a lahar: transition from debris flow to hyperconcentrated streamflow. Water Resources Research 21, 1511–1524.CrossRefGoogle Scholar
Pinet, P. R. 2006. Invitation to Oceanography, 4th edn. Boston: Jones and Bartlett.Google Scholar
Piper, D. J. W., Cochonat, P., & Morrison, M. L. 1999. The sequence of events around the epicenter of the 1929 Grand Banks earthquake: initiation of debris flow and turbidity current inferred from sidescan sonar. Sedimentology 46, 79–97.CrossRefGoogle Scholar
Piper, D. J. W., & Normark, W. R. 1983. Turbidite depositional patterns and flow characteristics, Navy submarine fan, California Borderland. Sedimentology 30, 681–694.CrossRefGoogle Scholar
Pitzer, K. 1973. Thermodynamics of electrolytes. I. Theoretical basis and general equations. Journal of Physical Chemistry 77, 268–277.CrossRefGoogle Scholar
Playford, P. E. 1984. Platform–margin and marginal–slope relationships in the Devonian reef complexes of the Canning Basin. In The Canning Basin, Western Australia (ed. Purcell, P. G.), pp. 189–214. Perth: Geological Society of Australia and Petroleum Exploration Society of Australia.Google Scholar
Playford, P. E., & Cockbain, A. E. 1976. Modern algal stromatolites at Hamelin Pool, a hypersaline barred basin, Western Australia. In Stromatolites (ed. Walter, M. R.), pp. 389–411. Amsterdam: Elsevier.Google Scholar
Plummer, L. N., Parkhurst, D. L., Fleming, G. W., & Dunkle, S. A. 1988. A computer program incorporating Pitzer's equations for calculation of geochemical reactions in brines. United States Geological Survey Water-resources Investigations Report 88–4153.
Pond, S., & Pickard, G. L. 1983. Introductory Dynamical Oceanography, 2nd edn. London: Pergamon.Google Scholar
Pope, M. C., & Grotzinger, J. P. 2000. Controls on fabric development and morphology of tufas and stromatolites, Uppermost Pethei Group (1.8 Ga), Great Slave Lake, Northwest Canada. In Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World (ed. Grotzinger, J. P. & James, N. P.), pp. 103–121. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Portman, C. P., Andrews, J. E., Rowe, P. J., Leeder, M. R., & Hoodewerff, J. 2005. Submarine-spring controlled calcification and growth of large Rivularia bioherms, Late Pleistocene (MIS 5e), Gulf of Corinth, Greece. Sedimentology 52, 441–465.CrossRefGoogle Scholar
Posamentier, H. W., & Allen, G. P. 1999. Siliciclastic Sequence Stratigraphy – Concepts and Applications. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Posamentier, H. W., Allen, G. P., James, D. P., & Tesson, M. 1992. Forced regressions in a sequence stratigraphic framework: concepts, examples, and exploration significance. American Association of Petroleum Geologists Bulletin 76, 1687–1709.Google Scholar
Posamentier, H. W., Erskine, R. D., & Mitchum, R. M. 1991. Models for submarine fan deposition within a sequence stratigraphic framework. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 127–136. New York: Springer-Verlag.CrossRefGoogle Scholar
Posamentier, H. W., Jervey, M. T., & Vail, P. R. 1988. Eustatic controls on clastic deposition I – conceptual framework. In Sea Level Changes; An Integrated Approach (ed. Wilgus, C. K., Hastings, B. S., St., C. G.Kendall, C.et al.), pp. 109–124. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Posamentier, H. W., & Kolla, V. 2003. Seismic geomorphology and stratigraphy of depositional elements in deep-water settings. Journal of Sedimentary Research 73, 367–388.CrossRefGoogle Scholar
Posamentier, H. W., Summerhayes, C. P., Haq, B. U., & Allen, G. P. 1993. Sequence Stratigraphy and Facies Associations. Oxford: Blackwells.CrossRefGoogle Scholar
Posamentier, H. W., & Walker, R. G. 2006. Deep-water turbidites and submarine fans. In Facies Models Revisited (ed. Posamentier, H. W. & Walker, R. G.), pp. 397–520. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Post, A., & LaChapelle, E. R. 2000. Glacier Ice. Seattle, WA: University of Washington Press.Google Scholar
Postma, G. 1990. Depositional architecture and facies of river and fan deltas: a synthesis. In Coarse-Grained Deltas (ed. Colella, A. & Prior, D. B.), pp. 13–27. Oxford: Blackwells.CrossRefGoogle Scholar
Postma, G., Nemec, W., & Keinspehn, K. L. 1988. Large floating clasts in turbidites: a mechanism for their emplacement. Sedimentary Geology 58, 47–61.CrossRefGoogle Scholar
Potter, P. E., Maynard, J. B., & Depetris, P. J. 2005. Mud and Mudstone. Berlin: Springer-Verlag.Google Scholar
Powell, R. D. 1990. Glacimarine processes at grounding-line fans and their growth to ice-contact deltas. In Glacimarine Environments (ed. Dowdeswell, J. A. & Scourse, J. D.), pp. 53–73. London: Geological Society of London.Google Scholar
Powell, R. D., & Molnia, B. F. 1989. Glacimarine sedimentary processes, facies and morphology of the south-southeast Alaska shelf and fiords. Marine Geology 85, 359–390.CrossRefGoogle Scholar
Pratt, B. R., & James, N. P. 1982. Crypt-algal–metazoan bioherms of early Ordovician age in the St. George Group, western Newfoundland. Sedimentology 29, 313–343.CrossRefGoogle Scholar
Pratt, B. R., James, N. P., & Cowan, C. A. 1992. Peritidal carbonates. In Facies Models, Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 303–323. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Press, F., & Siever, R. 1978. Earth, 2nd edn. New York: W. H. Freeman and Company.Google Scholar
Press, F., Siever, R., Grotzinger, J., & Jordan, T. H. 2004. Understanding Earth, 4th edn. New York: W. H. Freeman and Company.Google Scholar
Prior, D. B., Bornhold, B. D., Wiseman, W. J., & Lowe, D. R. 1987. Turbidity current activity in a British Columbia fjord. Science 237, 1330–1333.CrossRefGoogle Scholar
Prior, D. B., & Coleman, J. M. 1979. Submarine landslides – geometry and nomenclature. Zeitschrift für Geomorphologie 23, 415–426.Google Scholar
Prothero, D. R. 2004. Bringing Fossils to Life: An Introduction to Paleobiology, 2nd edn. Boston, MA: McGraw-Hill.Google Scholar
Pryor, W. A. 1975. Biogenic sedimentation and alteration of argillaceous sediments in shallow marine environments. Geological Society of America Bulletin 86, 1244–1254.2.0.CO;2>CrossRefGoogle Scholar
Purser, B. H. 1973. The Persian Gulf: Holocene Carbonate Sedimentation and Diagenesis in a Shallow Epicontinental Sea. New York: Springer-Verlag.CrossRefGoogle Scholar
Purser, B. H., & Evans, G. 1973. Regional sedimentation along the Trucial Coast, SE Persian Gulf. In The Persian Gulf: Holocene Carbonate Sedimentation and Diagenesis in a Shallow Epicontinental Sea (ed. Purser, B. H.), pp. 211–231. New York: Springer-Verlag.CrossRefGoogle Scholar
Putnis, A. 2002. Mineral replacement reactions: from macroscopic observations to microscopic mechanisms. Mineralogical Magazine 66, 689–708.CrossRefGoogle Scholar
Pye, K. 1987. Aeolian Dust and Dust Deposits. London: Academic Press.Google Scholar
Pye, K.1993a. Late Quaternary development of coastal parabolic megadune complexes in northeastern Australia. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 23–44. Oxford: Blackwells.CrossRefGoogle Scholar
Pye, K. 1993b. The Dynamics and Environmental Context of Aeolian Sedimentary Systems. London: Geological Society of London.Google Scholar
Pye, K., & Lancaster, N. 1993. Aeolian Sediments: Ancient and Modern. Oxford: Blackwells.CrossRefGoogle Scholar
Pye, K., & Tsoar, H. 1990. Aeolian Sand and Sand Dunes. London: Chapman and Hall.CrossRefGoogle Scholar
Quinlan, G. M., & Beaumont, C. 1984. Appalachian thrusting, lithospheric flexure, and the Paleozoic stratigraphy of the eastern interior of North America. Canadian Journal of Earth Sciences 21, 973–996.CrossRefGoogle Scholar
Rahmani, R. A., & Flores, R. M. 1984. Sedimentology of Coal and Coal-bearing Sequences. Oxford: Blackwells.Google Scholar
Railsback, L. B. 2003. Stylolites. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 690–692. Dordrecht: Kluwer Academic Publishers.Google Scholar
Rampino, M. R., & Sanders, J. E. 1980. Holocene transgression in south-central Long Island, New York. Journal of Sedimentary Petrology 50, 1063–1080.Google Scholar
Rasmussen, B., Bengston, S., Fletcher, I. R., & McNaughton, N. J. 2002. Discoidal impressions and trace-like fossils more than 1200 million years old. Science 296, 1112–1115.CrossRefGoogle ScholarPubMed
Raymo, M. E. & Ruddiman, W. F. 1992. Tectonic forcing of late Cenozoic climate. Nature 359, 117–122.CrossRefGoogle Scholar
Raymond, C. F. 1987. How do glaciers surge? A review. Journal of Geophysical Research 92, 9121–9134.CrossRefGoogle Scholar
Read, J. F. 1985. Carbonate platform facies models. Bulletin of the American Association of Petroleum Geologists 66, 860–878.Google Scholar
Read, J. F. 1973. Carbonate cycles, Pillara Formation (Devonian), Canning Basin, Western Australia. Bulletin of Canadian Petroleum Geology 16, 649–653.Google Scholar
Read, J. F.1974. Calcrete deposits and Quaternary sediments, Edel Province, Shark Bay, Western Australia. In Evolution and Diagenesis of Quaternary Carbonate Sequences, Shark Bay, Western Australia (ed. Logan, B. W., Read, J. F., Hagan, G. M.et al.), pp. 250–282. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Read, W. A. 1994. High-frequency, glacial–eustatic sequences in early Namurian coal-bearing fluviodeltaic deposits, central Scotland. In Orbital Forcing and Cyclic Sequences (ed. Boer, P. L. & Smith, D. G.), pp. 413–428. Oxford: Blackwells.CrossRefGoogle Scholar
Reading, H. G. (ed.) 1986. Sedimentary Environments and Facies, 2nd edn. Oxford: Blackwells.Google Scholar
Reading, H. G. 1996. Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. Oxford: Blackwells.Google Scholar
Reading, H. G., & Collinson, J. D. 1996. Clastic coasts. In Sedimentary Environments: Processes, Facies and Stratigraphy (ed. Reading, H. G.), pp. 154–231. Oxford: Blackwells.Google Scholar
Reading, H. G., & Richards, M. 1994. Turbidite systems in deep-water basin margins classified by grain size and feeder system. American Association of Petroleum Geologists Bulletin 78, 792–822.Google Scholar
Reeder, R. J. 1983. Carbonates: Mineralogy and Chemistry. Chelsea, MI: Mineralogical Society of America.Google Scholar
Reesink, A. J. H., & Bridge, J. S. 2007. Influence of superimposed bedforms and flow unsteadiness on formation of cross strata in dunes and unit bars. Sedimentary Geology doi:10.1016/j.sedgeo.2007.02.005.CrossRef
Reeves, C. C. 1977. Caliche: Origin, Classification, Morphology and Uses. Lubbock, TX: Escado Books.Google Scholar
Reid, R. P., MacIntyre, I. G., & James, N. P. 1990. Internal precipitation of microcrystalline carbonate: a fundamental problem for sedimentologists. Sedimentary Geology 68, 163–170.CrossRefGoogle Scholar
Reineck, H.-E., & Singh, I. B. 1980. Depositional Sedimentary Environments: With Special Reference to Terrigenous Clastics. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Reineck, H.-E., & Wunderlich, F. 1968a. Zur Unterscheidung von asymmetrischen Oszillationripplen und Stromungsripplen. Senckenbergiana Lethaia 49, 321–345.Google Scholar
Reineck, H.-E., & Wunderlich, F. 1968b. Classification and origin of flaser and lenticular bedding. Sedimentology 11, 99–104.CrossRefGoogle Scholar
Reinharz, E., Nilsen, K. J., Boesch, D. F., Bertelsen, R., & O'Connell, A. E. 1982. A Radiographic Examination of Physical and Biogenic Sedimentary Structures in the Chesapeake Bay. Baltimore, MD: Maryland Geological Survey.Google Scholar
Reinson, G. E. 1992. Transgressive barrier island and estuarine systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 179–194. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Reitner, J. 1993. Modern cryptic microbialite/metazoan facies from Lizard Island (Great Barrier Reef, Australia): formation and concepts. Facies 29, 3–39.CrossRefGoogle Scholar
Reitner, J., & Neuweiler, F. 1995. Mud mounds: a polygenetic spectrum of fine-grained carbonate buildups. Facies 32, 1–70.Google Scholar
Renard, F., & Dysthe, D. 2003. Pressure solution. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 542–544. Dordrecht: Kluwer Academic Publishers.Google Scholar
Retallack, G. J. 1997. A Color Guide to Paleosols. Chichester: Wiley.Google Scholar
Retallack, G. J. 2001. Soils of the Past, 2nd edn. Oxford: Blackwells.CrossRefGoogle Scholar
Retallack, G. J.2003. Weathering, soils, and paleosols. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 770–776. Dordrecht: Kluwer Academic Publishers.Google Scholar
Retallack, G. J. 2005. Pedogenic carbonate proxies for amount and seasonality of precipitation in paleosols. Geology 33, 333–336.CrossRefGoogle Scholar
Reynaud, J. Y., Tessier, B., Proust, J. N.et al. 1999. Eustatic and hydrodynamic controls on the architecture of a deep shelf sand bank (Celtic Sea). Sedimentology 46, 703–721.CrossRefGoogle Scholar
Rhodes, B., Tuttle, M., Horton, B.et al. 2006. Paleotsunami research. EOS 87, 205.CrossRefGoogle Scholar
Rhodes, E. G., & Moslow, T. F. 1993. Marine Clastic Reservoirs: Examples and Analogues. New York: Springer-Verlag.CrossRefGoogle Scholar
Rice, S. P., & Church, M. 2001. Longitudinal profiles in simple alluvial systems. Water Resources Research 37, 417–426.CrossRefGoogle Scholar
Lucci, Ricci F. 1995. Sedimentographica: Photographic Atlas of Sedimentary Structures, 2nd edn. New York: Columbia University Press.Google Scholar
Richardson, K., & Carling, P. A. 2005. A typology of sculpted forms in open bedrock channels. Geological Society of America Special Paper 392.
Ries, J. R. 2004. Effect of ambient Mg/Ca ratio on Mg fractionation in calcareous marine invertebrates: a record of the oceanic Mg/Ca ratio over the Phanerozoic. Geology 32, 981–984.CrossRefGoogle Scholar
Riding, R. 1979. Origin and diagenesis of lacustrine algal bioherms at the margin of the Ries Crater, Upper Miocene, southern Germany. Sedimentology 26, 645–680.CrossRefGoogle Scholar
Riding, R. 2000. Microbial carbonates: the geologic record of calcified bacterial–algal mats and biofilms. Sedimentology 47, 179–214.CrossRefGoogle Scholar
Rinaldo, A., Rodriguez-Iturbe, I., & Rigon, R. 1998. Channel networks. Annual Review of Earth and Planetary Sciences 26, 289–327.CrossRefGoogle Scholar
Rinaldo, A., Rodriguez-Iturbe, I., Rigon, R.et al. 1992. Minimum energy and fractal structures of drainage networks. Water Resources Research 28, 2183–2195.CrossRefGoogle Scholar
Riggs, S. D. 1984. Paleoceanographic model of Neogene phosphorite deposition, U.S. Atlantic continental margin. Science 223, 123–131.CrossRefGoogle ScholarPubMed
Rine, J. M., & Ginsburg, R. N. 1985. Depositional facies of a mud shoreface in Suriname, South America. A mud analogue to sandy, shallow-marine deposits. Journal of Sedimentary Petrology 55, 633–652.Google Scholar
Ritter, D. F., Kochel, R. C., & Miller, J. R. 2002. Process Geomorphology, 4th edn. New York: McGraw-Hill.Google Scholar
Roberson, H. E., & Lahann, R. W. 1981. Smectite to illite conversion rates: effects of solution chemistry. Clay and Clay Minerals 29, 129–135.CrossRefGoogle Scholar
Roberts, H. H., Adams, R. D., & Cunningham, R. H. W. 1980. Evolution of the sand-dominant subaerial phase, Atchafalaya Delta, Louisiana. American Association of Petroleum Geologists Bulletin 64, 264–279.Google Scholar
Roberts, J. A., Bennett, P. C., González, L. A., Macpherson, G. L., & Milliken, K. L. 2004. Microbial precipitation of dolomite in methanogenic groundwater. Geology 32, 277–280.CrossRefGoogle Scholar
Robie, R. A., & Hemingway, B. S. 1995. Thermodynamic Properties of Minerals and Related Substances at 298.15 K and 1 bar (105 Pascals) Pressure and Higher Temperatures. Washington, D.C.: United States Geological Survey.Google Scholar
Robinson, R. L., & Slingerland, R. L. 1998a. Origin of fluvial grain-size trends in a foreland basin: the Pocono Formation of the central Appalachian basin. Journal of Sedimentary Research A68, 473–486.CrossRefGoogle Scholar
Robinson, R. L., & Slingerland, R. L. 1998b. Grain-size trends and basin subsidence in the Campanian Castlegate Sandstone and equivalent conglomerates of central Utah. Basin Research 10, 109–127.CrossRefGoogle Scholar
Robinson, R. A. J., Slingerland, R. L., & Walsh, J. M. 2001. Predicting fluvial–deltaic aggradation in Lake Roxburgh, New Zealand: test of a water and sediment routing model. In Geologic Modeling and Simulation: Sedimentary Systems (ed. Merriam, D. F. & Davis, J. C.), pp. 119–132. New York: Kluwer Academic/Plenum.CrossRefGoogle Scholar
Robock, A. 2002. The climatic aftermath. Science 295, 1242–1244.CrossRefGoogle ScholarPubMed
Rodriguez, A. B., Hamilton, M. D., & Anderson, J. B. 2000. Facies and evolution of the modern Brazos delta, Texas; wave versus flood influence. Journal of Sedimentary Research 70, 283–295.CrossRefGoogle Scholar
Rodriguez-Iturbe, I., & Rinaldo, A. 1997. Fractal River Basins: Chance and Self-organization. Cambridge: Cambridge University Press.Google Scholar
Rodriguez-Iturbe, I., Rinaldo, A., Rigon, R.et al. 1992. Energy dissipation, runoff production, and the three-dimensional structure of river basins. Water Resources Research 28, 1095–1103.CrossRefGoogle Scholar
Roscoe, R. 1953. Suspensions. In Flow Properties of Disperse Systems (ed. Hermans, J. J.), pp. 1–38. New York: Wiley Interscience.Google Scholar
Rosgen, D. L. 1994. A classification of natural rivers. Catena 22, 169–199.CrossRefGoogle Scholar
Rosgen, D. L. 1996. Applied River Morphology. Fort Collins, CO: Wildland Hydrology.Google Scholar
Ross, D. A., & Degens, E. T. 1969. Shipboard collection and preservation of sediment samples collected during CHAIN 61 from the Red Sea. In Hot Brines and Recent Heavy Metal Deposits in the Red Sea: A Geochemical and Geophysical Account (ed. Degens, E. T. & Ross, D. A.), pp. 363–367. New York: Springer-Verlag.CrossRefGoogle Scholar
Rothlisberger, H., & Lang, H. 1987. Glacial hydrology. In Glacio-Fluvial Sediment Transfer – An Alpine Perspective (ed. Gurnell, A. M. & Clark, M. J.), pp. 207–284. Chichester: Wiley.Google Scholar
Rowley, D. B. 2002. Rate of plate creation and destruction: 180 Ma to present. Geological Society of America Bulletin 114, 927–933.2.0.CO;2>CrossRefGoogle Scholar
Rubin, D. M. 1987. Cross-Bedding, Bedforms and Paleocurrents. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Rubin, D. M. 1990. Lateral migration of linear dunes in the Strzelecki Desert, Australia. Earth Surface Processes and Landforms 15, 1–14.CrossRefGoogle Scholar
Rubin, D. M., & Hunter, R. E. 1982. Bedform climbing in theory and nature. Sedimentology 29, 121–138.CrossRefGoogle Scholar
Rubin, D. M., & Hunter, R. E. 1985. Why deposits of longitudinal dunes are rarely recognized in the rock record. Sedimentology 32, 147–157.CrossRefGoogle Scholar
Rubin, D. M., & Hunter, R. E. 1987. Bedform alignment in directionally varying flows. Science 237, 276–278.CrossRefGoogle ScholarPubMed
Rust, B. R. 1978. A classification of alluvial channel systems. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 187–198. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Rust, B. R., & Romanelli, R. 1975. Late Quaternary subaqueous outwash deposits near Ottawa, Canada. In Glaciofluvial and Glaciolacustrine Sediments (ed. Jopling, A. V. & McDonald, B. C.), pp. 177–192. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ryan, W. B. F., & Hsü, K. J. 1973. Initial Reports of the Deep Sea Drilling Project, Vol. 13. Washington, D.C.: U.S. Government Printing Office.CrossRefGoogle Scholar
Sallenger, A. 1979. Inverse grading and hydraulic equivalence in grain-flow deposits. Journal of Sedimentary Petrology 49, 553–562.Google Scholar
Salter, T. 1993. Fluvial scour and incision: models for their influence on the development of realistic reservoir geometries. In Characterization of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 33–51. London: Geological Society of London.Google Scholar
Sandberg, P. A. 1975. New interpretations of Great Salt Lake ooids and of ancient non-skeletal carbonate mineralogy. Sedimentology 22, 497–538.CrossRefGoogle Scholar
Sandberg, P. A. 1983. An oscillating trend in Phanerozoic non-skeletal carbonate mineralogy. Nature 305, 19–22.CrossRefGoogle Scholar
Sarg, J. F. 1988. Carbonate sequence stratigraphy. In Sea Level Changes: An Integrated Approach (ed. Wilgus, C. K., Hastings, B. S., C. Kendall, C. G. St.et al.), pp. 155–182. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Sarna-Wojcicki, A. M., & Davis, J. O. 1991. Quaternary tephrochronology. In Quaternary Nonglacial Geology, Conterminous U.S. (ed. Morrison, R. B.), pp. 93–116. Boulder, CO: Geological Society of America.Google Scholar
Satake, K. 2005. Tsunamis: Case Studies and Recent Developments. Dordrecht: Springer-Verlag.CrossRefGoogle Scholar
Saucier, R. T. 1994. Geomorphology and Quaternary Geologic History of the Lower Mississippi Valley. Vicksburg, VA: Mississippi River Commission.Google Scholar
Saunders, I., & Young, A. 1983. Rates of surface processes on slopes, slope retreat and denudation. Earth Surface Processes and Landforms 8, 473–501.CrossRefGoogle Scholar
Saunderson, H. C., & Lockett, F. P. 1983. Flume experiments on bedforms and structures at the dune-plane bed transition. In Modern and Ancient Fluvial Systems (ed. Collinson, J. D. & Lewin, J.), pp. 49–58. Oxford: Blackwells.CrossRefGoogle Scholar
Savage, S. B. 1979. Gravity flow of cohesionless granular materials in chutes and channels. Journal of Fluid Mechanics 92, 53–96.CrossRefGoogle Scholar
Schenk, C. J., Gautier, D. L., Olhoeft, G. R., & Lucius, J. E. 1993. Internal structure of an aeolian dune using ground-penetrating radar. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 61–70. Oxford: Blackwells.CrossRefGoogle Scholar
Schlager, W. 2005. Carbonate Sedimentology and Sequence Stratigraphy. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Schlager, W., & Bolz, H. 1977. Clastic accumulation of sulphate evaporites in deep water. Journal of Sedimentary Petrology 47, 600–609.Google Scholar
Schlager, W., & Ginsburg, R. N. 1981. Bahamian carbonate platforms – the deep and the past. Marine Geology 44, 1–24.CrossRefGoogle Scholar
Schlichting, H. 1979. Boundary Layer Theory. New York: McGraw-Hill.Google Scholar
Schminke, H.-U. 2004. Volcanism. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Schneidermann, N., & Harris, P. M. 1985. Carbonate Cements. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Scholle, P. A. 1978. A Color Illustrated Guide to Carbonate Rock Constituents, Textures, Cements, and Porosities. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A. 1979. A Color Guide to Constituents, Textures, Cements, and Porosities of Sandstones and Associated Rocks. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A., Arthur, M. A, & Ekdale, A. A. 1983. Pelagic environments. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 620–691. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A., & Halley, R. B. 1985. Burial diagenesis: out of sight, out of mind! In Carbonate Cements (ed. Schneidermann, N. & Harris, P. M.), pp. 309–334. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Scholle, P. A., & Spearing, D. R. 1982. Sandstone Depositional Environments. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A., & Ulmer-Scholle, D. S. 2003. A Color Guide to the Petrography of Carbonate Rocks: Grains, Textures, Porosity, Diagenesis. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Schopf, J. W. 1983. Earth's Earliest Biosphere. Princeton, NJ: Princeton University Press.Google Scholar
Schubel, K. A., & Simonson, B. M. 1990. Petrography and diagenesis of cherts from Lake Magadi, Kenya. Journal of Sedimentary Petrology 60, 761–776.Google Scholar
Schulz, H. D., & Zabel, M. 2000. Marine Geochemistry. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Schwartz, R. K. 1982. Bedform and stratification characteristics of some modern small-scale washover sand bodies. Sedimentology 29, 835–849.CrossRefGoogle Scholar
Schumm, S. A. 1977. The Fluvial System. New York: Wiley.Google Scholar
Schumm, S. A. 1993. River response to baselevel change: implications for sequence stratigraphy. Journal of Geology 101, 279–294.CrossRefGoogle Scholar
Schumm, S. A., Dumont, J. F., & Holbrook, J. M. 2000. Active Tectonics and Alluvial Rivers. Cambridge: Cambridge University Press.Google Scholar
Schumm, S. A., & Lichty, R. W. 1963. Channel widening and floodplain construction, Cimarron River, Kansas. U.S. Geological Survey Professional Paper352-D, pp. 71–88.
Schumm, S. A., Mosley, M. P., & Weaver, W. E. 1987. Experimental Fluvial Geomorphology. New York: Wiley.Google Scholar
Scruton, P. C. 1960. Delta building and the deltaic sequence. In Recent Sediments, Northwest Gulf of Mexico (ed. Shepard, F. P., Phleger, F. B., & Andel, T. H.), pp. 82–102. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Seifert, D., & Jensen, J. L. 1999. Using sequential indicator simulation as a tool in reservoir description: issues and uncertainties. Mathematical Geology 31, 527–550.CrossRefGoogle Scholar
Seifert, D., & Jensen, J. L. 2000. Object and pixel-based reservoir modeling of a braided fluvial reservoir. Mathematical Geology 32, 581–603.CrossRefGoogle Scholar
Seilacher, A, Bose, P. K., & Pflüger, F. 1998. Triploblastic animals more than 1 billion years ago: trace fossil evidence from India. Science 282, 80–83.CrossRefGoogle ScholarPubMed
Selby, M. J. 1993. Hillslope Materials and Processes, 2nd edn. Oxford: Oxford University Press.Google Scholar
Semikhatov, M. A., Gebelein, C. D., Cloud, P., Awramik, S. M., & Benmore, W. C. 1979. Stromatolite morphogenesis – progress and problems. Canadian Journal of Earth Sciences 16, 992–1015.CrossRefGoogle Scholar
Sepkoski, J. J. Jr. 1981. A factor analytic description of the Phanerozoic marine record. Paleobiology 7, 36–53.CrossRefGoogle Scholar
Sha, L. P., & de Boer, P. L. 1991. Ebb-tidal delta deposits along the west Friesian islands (The Netherlands): processes, facies architecture and preservation. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 199–218. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Shanley, K. W., & McCabe, P. J. 1993. Alluvial architecture in a sequence stratigraphic framework: a case history from the Upper Cretaceous of southern Utah, USA. In The Geological Modeling of Hydrocarbon Reservoirs and Outcrop Analogues (ed. Flint, S. & Bryant, I. D.), pp. 21–56. Oxford: Blackwells.Google Scholar
Shanley, K. W., & McCabe, P. J. 1994. Perspectives on the sequence stratigraphy of continental strata. American Association of Petroleum Geologists Bulletin 78, 544–568.Google Scholar
Shanmugan, G. 1996. High-density turbidity currents: are they sandy debris flows?Journal of Sedimentary Research 66, 2–10.CrossRefGoogle Scholar
Shanmugan, G. 1997. The Bouma sequence and the turbidite mind set. Earth-Science Reviews 42, 201–229.CrossRefGoogle Scholar
Shanmugan, G., & Moiola, R. J. 1988. Submarine fans: characteristics, models, classification and reservoir potential. Earth-Science Reviews 24, 383–428.CrossRefGoogle Scholar
Sharp, M. 1988a. Surging glaciers: behaviour and mechanisms. Progress in Physical Geography 12, 349–370.CrossRefGoogle Scholar
Sharp, M. 1988b. Surging glaciers: geomorphic effects. Progress in Physical Geography 12, 533–559.CrossRefGoogle Scholar
Sharp, M., Jouzel, J., Hubbard, B., & Lawson, W. 1994. The character, structure and origin of the basal ice layer of a surge-type glacier. Journal of Glaciology 40, 327–340.CrossRefGoogle Scholar
Sharp, M., Richards, K. S., & Tranter, M. 1998. Glacier Hydrology and Hydrochemistry. Chichester: Wiley.Google Scholar
Sharp, R. P. 1963. Wind ripples. Journal of Geology 71, 617–636.CrossRefGoogle Scholar
Sharpe, D. R., & Shaw, J. 1989. Erosion of bedrock by subglacial meltwater, Cantley, Quebec. Geological Society of America Bulletin 101, 1011–1020.2.3.CO;2>CrossRefGoogle Scholar
Shaw, J. 2006. A glimpse at meltwater effects associated with continental ice sheets. In Glacier Science and Environmental Change (ed. Knight, P. G.), pp. 25–32. Oxford: Blackwells.CrossRefGoogle Scholar
Shaw, J., Kvill, D., & Rains, B. 1989. Drumlins and catastrophic subglacial floods. Sedimentary Geology 62, 177–202.CrossRefGoogle Scholar
Shaw, J., & Sharpe, D. R. 1987. Drumlin formation by subglacial meltwater erosion. Canadian Journal of Earth Sciences 24, 2316–2322.CrossRefGoogle Scholar
Shear, W. A., & Selden, P. A. 2001. Rustling in the undergrowth: animals in early terrestrial ecosystems. In Plants Invade the Land (ed. Gensel, P. G. & Edwards, D.), pp. 29–51. New York: Columbia University Press.CrossRefGoogle Scholar
Shearman, D. J. 1963. Recent anhydrite, gypsum, dolomite and halite from the coastal flats of Arabian shore of the Persian Gulf. Proceedings of the Geological Society of London 1607, 63–65.Google Scholar
Shearman, D. J., McCugan, A., Stein, C., & Smith, A. J. 1989. Ikaite, CaCO3⋅6H2O, precursor of the thinolites in the Quaternary tufas and tufa mounds of the Lahontin and Mono Lake basins, western United States. Geological Society of America Bulletin 101, 913–917.2.3.CO;2>CrossRefGoogle Scholar
Sheehan, P. M., & Harris, M. T. 2004. Microbialite resurgence after the Late Ordovician extinction. Nature 430, 75–78.CrossRefGoogle ScholarPubMed
Shepard, F. P. 1932. Sediments on continental shelves. Geological Society of America Bulletin 43, 1017–1034.CrossRefGoogle Scholar
Shepard, F. P. 1973. Submarine Geology. New York: Harper and Row.Google Scholar
Shepard, F. P., & Inman, D. L. 1950. Nearshore circulation related to bottom topography and wave refraction. Transactions of the American Geophysical Union 31, 555–565.CrossRefGoogle Scholar
Sheridan, R. E. 1974. Altantic continental margin of North America. In The Geology of Continental Margins (ed. Burke, C. A. & Drake, C. L.), pp. 391–407. New York: Springer-Verlag.CrossRefGoogle Scholar
Shields, A. 1936. Anwendung der Ähnlichkeitsmechanik und der Turbulenzforschung auf die Geschiebebewegung. Mitteilungen der Preuβischen Versuchsanstalt für Wasserbau und Schiffbau 26, 26 pp.Google Scholar
Shiki, T., Cita, M. B., & Gorsline, D. S. 2000. Seismoturbidites, seismites and tsunamiites. Sedimentary Geology (special issue) 135, 1–322.Google Scholar
Shinn, E. A. 1968a. Selective dolomitization of recent sedimentary structures. Journal of Sedimentary Petrology 38, 612–616.Google Scholar
Shinn, E. A. 1968b. Practical significance of birdseye structures in carbonate rocks. Journal of Sedimentary Petrology 38, 611–616.CrossRefGoogle Scholar
Shinn, E. A. 1969. Submarine lithification of Holocene carbonate sediments in the Persian Gulf. Sedimentology 12, 109–144.CrossRefGoogle Scholar
Shinn, E. A. 1983a. Birdseyes, fenestrae, shrinkage pores, and Loferites: a reevaluation. Journal of Sedimentary Petrology 53, 619–628.Google Scholar
Shinn, E. A.1983b. Tidal flats. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 171–210. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Shinn, E. A., & Lidz, B. H. 1988. Blackened limestone pebbles: fire at subaerial unconformities. In Paleokarst (ed. James, N. P. & Choquette, P. W.), pp. 117–131. New York: Springer-Verlag.CrossRefGoogle Scholar
Shinn, E. A., Lloyd, R. M., & Ginsburg, R. N. 1969. Anatomy of a modern carbonate tidal flat, Andros Island, Bahamas. Journal of Sedimentary Petrology 39, 1202–1228.Google Scholar
Shinn, E. A., & Robbin, D. M. 1983. Mechanical and chemical compaction in fine-grained shallow-water limestones. Journal of Sedimentary Petrology 53, 596–618.Google Scholar
Shinn, E. A., Steinen, R. P., Lidz, B. H., & Swart, R. K. 1989. Whitings, a sedimentological dilemma. Journal of Sedimentary Petrology 59, 147–161.CrossRefGoogle Scholar
Shreve, R. L. 1966. Statistical law of stream numbers. Journal of Geology 74, 17–37.CrossRefGoogle Scholar
Shvidchenko, A. B., & Pender, G. 2001. Macroturbulent structure of open-channel flow over gravel beds. Water Resources Research 37, 709–719.CrossRefGoogle Scholar
Sibley, D. F., & Gregg, J. M. 1987. Classification of dolomite rock texture. Journal of Sedimentary Petrology 57, 967–975.Google Scholar
Sidi, F. H., Nummedal, D., Imbert, P., Darman, H., & Posamentier, H. 2003. Tropical Deltas of Southeast Asia – Sedimentology Stratigraphy, and Petroleum Geology. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Siegert, M. J. 2000. Antarctic subglacial lakes. Earth-Science Reviews 50, 29–50.CrossRefGoogle Scholar
Simms, M. A. 1984. Dolomitization by groundwater-flow systems in carbonate platforms. Transactions of the Gulf Coast Association of Geological Societies 34, 411–420.Google Scholar
Simone, L. 1980. Ooids: a review. Earth-Science Reviews 16, 319–335.CrossRefGoogle Scholar
Simonson, B. M. 2003a. Ironstones and iron formations. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 379–385. Dordrecht: Kluwer Academic Publishers.Google Scholar
Simonson, B. M. 2003b. Origin and evolution of large Precambrian iron formations. Geological Society of America Special Paper 370, pp. 231–244.Google Scholar
Simonson, B. M., & Carney, K. E. 1999. Roll-up structures: evidence of in situ microbial mats in Late Archean deep shelf environments. Palaios 14, 13–24.CrossRefGoogle Scholar
Simpson, J. E. 1982. Gravity currents in the laboratory, atmosphere and ocean. Annual Review of Fluid Mechanics 14, 213–234.CrossRefGoogle Scholar
Simpson, J. E. 1997. Gravity Currents in the Environment and the Laboratory, 2nd edn. New York: Cambridge University Press.Google Scholar
Singh, U. 1987. Ooids and cements from the Late Precambrian of the Flinders Range, South Australia. Journal of Sedimentary Petrology 57, 117–127.CrossRefGoogle Scholar
Sinha, S. K., & Parker, G. 1996. Causes of concavity in longitudinal profiles of rivers. Water Resources Research 32, 1417–1428.CrossRefGoogle Scholar
Siringan, F. P., & Anderson, J. B. 1993. Seismic facies architecture and evolution of the Bolivar Roads tidal inlet/delta complex, east Texas Gulf Coast. Journal of Sedimentary Petrology 63, 794–808.Google Scholar
Siringan, F. P., & Anderson, J. B. 1994. Modern shoreface and inner-shelf storm deposits off the East Texas coast, Gulf of Mexico. Journal of Sedimentary Petrology 64, 99–110.Google Scholar
Sleath, J. F. A. 1984. Sea Bed Mechanics. New York: Wiley.Google Scholar
Slingerland, R. L. 1986. Numerical computation of co-oscillating paleotides in the Catskill epeiric sea of eastern North America. Sedimentology 33, 817–829.CrossRefGoogle Scholar
Slingerland, R., Harbaugh, J. W., & Furlong, K. P. 1994. Simulating Clastic Sedimentary Basins. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Slingerland, R. L., & Smith, N. D. 1998. Necessary conditions for a meandering-river avulsion. Geology 26, 435–438.2.3.CO;2>CrossRefGoogle Scholar
Slingerland, R. L., & Smith, N. D. 2004. River avulsions and their deposits. Annual Review of Earth and Planetary Sciences 32, 255–283.CrossRefGoogle Scholar
Sloss, L. L. 1963. Sequences in the cratonic interior of North America. Geological Society of America Bulletin 74, 93–114.CrossRefGoogle Scholar
Smith, C. R. 1996. Coherent flow structures in smooth-wall turbulent boundary layers: facts, mechanisms and speculation. In Coherent Flow Structures in Open Channels (ed. Ashworth, P. J., Bennett, S. J., Best, J. L., & McLelland, S. J.), pp. 1–39. Chichester: Wiley.Google Scholar
Smith, D. G. 1988. Modern point bar deposits analogous to the Athabasca Oil Sands, Alberta, Canada. In Tide-Influenced Sedimentary Environments and Facies (ed. DeBoer, P. L., Gelder, A., & Nio, S. D.), pp. 417–432. Boston, MA: D. Reidel Publishing Company.CrossRefGoogle Scholar
Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A. 1991. Clastic Tidal Sedimentology. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Smith, D. G., & Smith, N. D. 1980. Sedimentation in anastomosed river systems: examples from alluvial valleys near Banff, Alberta. Journal of Sedimentary Petrology 50, 157–164.CrossRefGoogle Scholar
Smith, L. C., Sheng, Y., Magilligan, F. J.et al. 2006. Geomorphic impact and rapid subsequent recovery from the 1996 Skeiðarásandur jokulhlaup, Iceland, measured with multi-year airborne lidar. Geomorphology 75, 65–75.CrossRefGoogle Scholar
Smith, N. D. 1974. Sedimentology and bar formation in the upper Kicking Horse River, a braided outwash stream. Journal of Geology 81, 205–223.CrossRefGoogle Scholar
Smith, N. D., Cross, T. A., Dufficy, J. P., & Clough, S. R. 1989. Anatomy of an avulsion. Sedimentology 36, 1–23.CrossRefGoogle Scholar
Smith, N. D., & Rogers, J. 1999. Fluvial Sedimentology VI. Oxford: Blackwells.CrossRefGoogle Scholar
Smith, N. D., Slingerland, R. L., Pérez-Arlucea, M., & Morozova, G. S. 1998. The 1870s avulsion of the Saskatchewan River. Canadian Journal of Earth Sciences 35, 453–466.CrossRefGoogle Scholar
Smith, S. V., & Kinsey, D. W. 1976. Calcium carbonate production, coral reef growth, and sea level change. Science 194, 937–939.CrossRefGoogle ScholarPubMed
Smoot, J. P. 1983. Depositional subenvironments in an arid closed basin: the Wilkens Peak Member of the Green River Formation (Eocene), Wyoming. Sedimentology 30, 801–827.CrossRefGoogle Scholar
Smoot, J. P., & Lowenstein, T. K. 1991. Depositional environments of non-marine evaporites. In Evaporites, Petroleum and Mineral Resources (ed. Melvin, J. L.), pp. 189–347. Amsterdam: Elsevier.Google Scholar
Snedden, J. W., & Nummedal, D. 1991. Origin and geometry of storm-deposited sand beds in modern sediments of the Texas continental shelf. In Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (ed. Swift, D. J. P., Oertel, G. F., Tillman, R. W., & Thorne, J. A.), pp. 283–308. Oxford: Blackwells.Google Scholar
Sonnenfeld, P. 1984. Brines and Evaporites. Orlando, FL: Academic Press.Google Scholar
Sorby, H. C. 1879. The structure and origin of limestones. Proceedings of the Geological Society of London 35, 56–95.Google Scholar
Soulsby, R. L., & Bettess, R. 1991. Sand Transport in Rivers, Estuaries and the Sea. Rotterdam: Balkema.Google Scholar
Southard, J. B., & Boguchwal, L. A. 1990. Bed configurations in steady unidirectional flows. Part 2. Synthesis of flume data. Journal of Sedimentary Petrology 60, 658–679.CrossRefGoogle Scholar
Southard, J. B., Lambie, J. M., Federico, D. C., Pile, H. T., & Weidman, C. R. 1990. Experiments on bed configurations in fine sands under bidirectional purely oscillatory flow, and the origin of hummocky cross-stratification. Journal of Sedimentary Petrology 60, 1–17.Google Scholar
Sparks, R. S. J. 1976. Grain-size variations in ignimbrites and implications for the transport of pyroclastic flows. Sedimentology 23, 147–188.CrossRefGoogle Scholar
Sparks, R. S. J., Bonnecaze, R. T., Huppert, H. E.et al. 1993. Sediment-laden gravity currents with reversing buoyancy. Earth and Planetary Science Letters 114, 249–288.CrossRefGoogle Scholar
Sparks, R. S. J., & Gilbert, J. S. 2002. Physics of Explosive Volcanic Eruptions. London: Geological Society of London.Google Scholar
Spencer, R. J., & Lowenstein, T. K. 1990. Evaporites. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 141–163. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Spencer, R. J., Möller, N., & Weare, J. H. 1990. The prediction of mineral solubilities in natural waters: a chemical equilibrium model for the Na–K–Ca–Mg–Cl–SO4–H2O system at temperatures below 25 ℃. Geochimica et Cosmochimica Acta 54, 575–590.CrossRefGoogle Scholar
Sposito, G. 1989. The Chemistry of Soils. New York: Oxford University Press.Google Scholar
Srivastava, R. M. 1994. An overview of stochastic methods for reservoir characterization. In Stochastic Modeling and Geostatistics (ed. Yarus, J. M. & Chambers, R. L.), pp. 3–16. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Stallard, R. F., & Edmond, J. M. 1981. Geochemistry of the Amazon 1. Precipitation chemistry and the marine contribution to the dissolved load at the time of peak discharge. Journal of Geophysical Research 86, 9844–9858.CrossRefGoogle Scholar
Stallard, R. F., & Edmond, J. M. 1983. Geochemistry of the Amazon 2. The influence of geology and weathering environment on the dissolved load. Journal of Geophysical Research 88, 9671–9688.CrossRefGoogle Scholar
Stallard, R. F., & Edmond, J. M. 1987. Geochemistry of the Amazon 3. Weathering chemistry and the limits to dissolved inputs. Journal of Geophysical Research 92, 8293–8302.CrossRefGoogle Scholar
Stanley, S. M. 1966. Paleoecology and diagenesis of Key Largo Limestone, Florida. American Association of Petroleum Geologists Bulletin 50, 1927–1947.Google Scholar
Stanley, S. M. 2005. Earth System History, 2nd edn. New York: W. H. Freeman and Company.Google Scholar
Stanley, S. M., & Hardie, L. A. 1998. Secular oscillations in the carbonate mineralogy of reef-building and sediment-producing organisms driven by tectonically forced shifts in seawater chemistry. Palaeogeography, Palaeoclimatology, Palaeoecology 144, 3–19.CrossRefGoogle Scholar
Stanley, S. M., Ries, J. B., & Hardie, L. A. 2002. Low-magnesium calcite produced by coralline algae in seawater of Late Cretaceous composition. Proceedings of the National Academy of Sciences of the USA 99, 15323–15326.CrossRefGoogle ScholarPubMed
Starkel, L., Gregory, K. J., & Thornes, J. B. 1991. Temperate Palaeohydrology. Chichester: Wiley.Google Scholar
Steel, R. J., Maehle, S., Nilsen, H., Roe, S. L., & Spinnanger, A. 1977. Coarsening-upward cycles in the alluvium of the Hornelen basin (Devonian, Norway): sedimentary response to tectonic events. Geological Society of America Bulletin 88, 1124–1134.2.0.CO;2>CrossRefGoogle Scholar
Steel, R. J., & Ryseth, A. 1990. The Triassic–Early Jurassic succession in the northern North Sea: megasequence stratigraphy and intra-Triassica tectonics. In Tectonic Events Responsible for Britain's Oil and Gas Reserves (ed. Hardman, R. F. P. & Brooks, J.), pp. 139–168. London: Geological Society of London.Google Scholar
Stern, C. W., Scoffin, T. P., & Martindale, W. 1977. Calcium carbonate budget of a fringing reef on the west coast of Barbados, pt. 1, zonation and productivity. Bulletin of Marine Science 27, 779–810.Google Scholar
Stewart, F. H. 1963. Marine evaporites. United States Geological Survey Professional Paper 440-Y.
Stockman, K. W., Ginsburg, R. N., & Shinn, E. A. 1967. The production of lime mud by algae in South Florida. Journal of Sedimentary Petrology 37, 633–648.Google Scholar
Stokes, C. R., & Clark, C. D. 2001. Palaeo-ice streams. Quaternary Science Reviews 20, 1437–1457.CrossRefGoogle Scholar
Stokes, C. R., & Clark, C. D. 2002. Ice stream shear margin moraines. Earth Surface Processes and Landforms 27, 547–558.CrossRefGoogle Scholar
Stokes, C. R., & Clark, C. D. 2003. Laurentide ice streaming on the Canadian Shield: a conflict with the soft-bedded ice stream paradigm. Geology 31, 347–350.2.0.CO;2>CrossRefGoogle Scholar
Stokes, W. L. 1968. Multiple parallel-truncation bedding planes – a feature of wind deposited sandstone formations. Journal of Sedimentary Petrology 38, 510–515.Google Scholar
Stokstad, E. 2004. Defrosting the carbon freezer of the north. Science 304, 1618–1620.CrossRefGoogle ScholarPubMed
Stone, G. W., & Orford, J. D. 2004. Storms and their significance in coastal morpho-sedimentary dynamics. Marine Geology 210, 1–368.CrossRefGoogle Scholar
Stonecipher, S. A. 2000. Applied Sandstone Diagenesis – Practical Petrographic Solutions for a Variety of Common Exploration, Development, and Production Problems. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Storms, J. E. A. 2003. Event-based stratigraphic simulation of wave-dominated shallow-marine environments. Marine Geology 199, 83–100.CrossRefGoogle Scholar
Storms, J. E. A., & Swift, D. J. P. 2003. Shallow-marine sequences as the building blocks of stratigraphy: insights from numerical modeling. Basin Research 15, 287–303.CrossRefGoogle Scholar
Stouthamer, E., & Berendsen, H. J. A. 2000. Factors controlling the Holocene avulsion history of the Rhine–Meuse delta (The Netherlands). Journal of Sedimentary Research 70, 1051–1064.CrossRefGoogle Scholar
Stouthamer, E., & Berendsen, H. J. A. 2001. Avulsion frequency, avulsion duration and interavulsion period of Holocene channel belts in the Rhine–Meuse delta, The Netherlands. Journal of Sedimentary Research 71, 589–598.CrossRefGoogle Scholar
Stouthamer, E. 2005. Reoccupation of channel belts and its influence on alluvial architecture in the Holocene Rhine-Meuse delta, the Netherlands. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J.), pp. 319–339. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Stow, D. A. V. 1986. Deep clastic seas. In Sedimentary Environments and Facies, 2nd edn. (ed. Reading, H. G.), pp. 399–444. Oxford: Blackwells.Google Scholar
Stow, D. A. V.1994. Deep sea processes of sediment transport and deposition. In Sediment Transport and Depositional Processes (ed. Pye, K.), pp. 257–291. Oxford: Blackwells.Google Scholar
Stow, D. A. V., & Faugeres, J. C. (eds.) 1993. Contourites and Bottom Currents. Sedimentary Geology (special issue) 82.
Stow, D. A. V., & Faugeres, J. C.1998. Contourites, Turbidites and Process Interaction. Sedimentary Geology (special issue) 115.
Stow, D. A. V., & Lovell, J. P. B. 1979. Contourites: their recognition in modern and ancient sediments. Earth-Science Reviews 14, 251–291.CrossRefGoogle Scholar
Stow, D. A. V., & Piper, D. J. W. 1984. Fine-Grained Sediments: Deep-Water Processes and Facies. London: Geological Society of London.Google Scholar
Stow, D. A. V., Pudsey, C. J., Howe, J. A., Faugeres, J. -C., & Vianna, A. R. 2002. Deep-water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics. London: Geological Society of London.Google Scholar
Stow, D. A. V., Reading, H. G., & Collinson, J. D. 1996. Deep seas. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 395–453. Oxford: Blackwells.Google Scholar
Stow, D. A. V., & Shanmugan, G. 1980. Sequences of structures in fine-grained turbidites: comparison of recent deep sea and ancient flysch sediments. Sedimentary Geology 25, 23–42.CrossRefGoogle Scholar
Stow, D. A. V., & Tabrez, A. 1998. Hemipelagites: Facies, Processes, and Models, pp. 317–338. London: Geological Society of London.Google Scholar
Stow, D. A. V., & Wetzel, A. 1990. Hemiturbidite: a new type of deep-water sediment. Proceedings of the Ocean Drilling Program, Scientific Results 116, 25–34.Google Scholar
Strahler, A. N. 1957. Quantitative analysis of watershed geomorphology. Transactions of the American Geophysical Union 38, 913–920.CrossRefGoogle Scholar
Straub, S. 2001. Bagnold revisited: implications for the rapid motion of high-concentration sediment flows. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 91–109. Oxford: Blackwells.CrossRefGoogle Scholar
Stride, A. H. 1982. Offshore Tidal Sands. London: Chapman and Hall.CrossRefGoogle Scholar
Stumm, W., & Morgan, J. J. 1981. Aquatic Chemistry: An Introduction Emphazing Chemical Equilibria in Natural Waters. New York: John Wiley and Sons.Google Scholar
Sugden, D. E., & John, B. S. 1976. Glaciers and Landscape. London: Edward Arnold.Google Scholar
Summerhayes, C. P., & Thorpe, S. A. 1996. Oceanography: An Illustrated Guide. London: Manson.Google Scholar
Summerfield, M. A. 1983. Petrography and diagenesis of silcrete from the Kalahari Basin and Cape coastal zone. Journal of Sedimentary Petrology 53, 895–909.Google Scholar
Sumner, D. Y., & Grotzinger, J. P. 2000. Late Archean aragonite precipitation: petrography, facies associations, and environmental significance. In Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World (ed. Grotzinger, J. P. & James, N. P.), pp. 123–144. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Sun, T., Meakin, P., & Jossang, T. 2001a. Meander migration and the lateral tilting of floodplains. Water Resources Research 37, 1485–1502.CrossRefGoogle Scholar
Sun, T., Meakin, P., & Jossang, T. 2001b. A computer model for meandering rivers with multiple bedload sediment sizes 1. Theory. Water Resources Research 37, 2227–2241.CrossRefGoogle Scholar
Sun, T., Meakin, P., & Jossang, T. 2001c. A computer model for meandering rivers with multiple bedload sediment sizes 2. Computer simulations. Water Resources Research 37, 2243–2258.CrossRefGoogle Scholar
Sun, T., Paola, C., Parker, G., & Meakin, P. 2002. Fluvial fan deltas: linking channel processes with large-scale morphodynamics. Water Resources Research 38, 1151.CrossRefGoogle Scholar
Swift, D. J. P. 1975. Barrier island genesis: evidence from the Middle Atlantic Shelf of North America. Sedimentary Geology 14, 1–43.CrossRefGoogle Scholar
Swift, D. J. P., & Field, M. E. 1981. Evolution of a classic sand ridge field; Maryland sector, North American inner shelf. Sedimentology 28, 461–482.CrossRefGoogle Scholar
Swift, D. J. P., Figueiredo, A. R., Freeland, G. L., & Oertal, G. F. 1983. Hummocky cross stratification and megaripples: a geological double standard?Journal of Sedimentary Petrology 53, 1295–1317.Google Scholar
Swift, D. J. P., Han, G., & Vincent, C. E. 1986. Fluid processes and sea-floor response on a modern storm-dominated shelf: middle Atlantic shelf of North America. Part I: the storm-current regime. In Shelf Sands and Sandstones (ed. Knight, R. J. & McLean, J. R.), pp. 99–119. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Swift, D. J. P., Oertel, G. F., Tillman, R. W., & Thorne, J. A. 1991. Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy. Oxford: Blackwells.Google Scholar
Swift, D. J. P., & Thorne, J. A. 1991. Sedimentation on continental margins, I: a general model for shelf sedimentation. In Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (ed. Swift, D. J. P., Oertel, G. F., Tillman, R. W., & Thorne, J. A.), pp. 3–31. Oxford: Blackwells.Google Scholar
Syvitski, J. P. M., Burrell, D. C., & Skei, J. M. 1987. Fjords: Processes and Products. New York: Springer-Verlag.CrossRefGoogle Scholar
Takahashi, T. 1981. Debris flow. Annual Review of Fluid Mechanics 13, 57–77.CrossRefGoogle Scholar
Takahashi, T.2001. Mechanics and simulation of snow avalanches, pyroclastic flows and debris flows. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 11–43. Oxford: Blackwells.CrossRefGoogle Scholar
Talbot, M. R. 1985. Major bounding surfaces in aeolian sandstones – a climatic model. Sedimentology 32, 257–265.CrossRefGoogle Scholar
Talbot, M. R., & Allen, P. A. 1996. Lakes. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 83–124. Oxford: Blackwell Science.Google Scholar
Taylor, A., Goldring, R., & Gowland, S. 2003. Analysis and application of ichnofabrics. Earth-Science Reviews 60, 227–259.CrossRefGoogle Scholar
Taylor, K. G., Gawthorpe, R. L., Curtis, C. D., Marshall, J. D., & Anwiller, D. A. 2000. Carbonate cementation in a sequence-stratigraphic framework: Upper Cretaceous sandstones, Book Cliffs, Utah–Colorado. Journal of Sedimentary Research 70, 360–372.CrossRefGoogle Scholar
Tennant, C. B., & Berger, R. W. 1957. X-ray determination of the dolomite–calcite ratios of a carbonate rock. American Mineralogist 42, 23–29.Google Scholar
Terwindt, J. H. J. 1981. Origin and sequences of sedimentary structures in inshore mesotidal deposits of the North Sea. In Holocene Marine Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 51–64. Oxford: Blackwells.CrossRefGoogle Scholar
Terwindt, J. H. J.1988. Palaeo-tidal reconstructions of inshore tidal depositional environments. In Tide-influenced Sedimentary Environments and Facies (ed. DeBoer, P. L., Gelder, A., & Nio, S. D.), pp. 233–263. Boston, MA: D. Reidel Publishing Company.CrossRefGoogle Scholar
Tetzlaff, D. M., & Harbaugh, J. W. 1989. Simulating Clastic Sedimentation. New York: Van Nostrand Reinhold.CrossRefGoogle Scholar
Tetzlaff, D. M., & Priddy, G. 2001. Sedimentary process modeling: from academia to industry. In Geologic Modeling and Simulation: Sedimentary Systems (ed. Merriam, D. F. & Davis, J. C.), pp. 45–69. New York: Kluwer Academic/Plenum Publishers.CrossRefGoogle Scholar
Thomas, D. 1997. Arid Zone Geomorphology. London: Bellhaven/Hallsted Press.Google Scholar
Thomas, R. G., Smith, D. G., Wood, J. M.et al. 1987. Inclined heterolithic stratification – terminology, description, interpretation and significance. Sedimentary Geology 53, 123–179.CrossRefGoogle Scholar
Thrailkill, J. 1976. Speleothems. In Stromatolites (ed. Walter, M. R.), pp. 73–86. Amsterdam: Elsevier.Google Scholar
Thurman, H. V., & Trujilo, A. P. 2004. Introductory oceanography, 10th edn. Englewood Cliffs, NJ, Prentice-Hall.Google Scholar
Tillman, R. W., & Martinsen, R. S. 1984. The Shannon shelf-ridge sandstone complex, Salt Creek anticline area, Powder River Basin, Wyoming. In Siliciclastic Shelf Sediments (ed. Tillman, R. W. & Seimers, C. T.), pp. 85–142. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Tillman, R. W., & Martinsen, R. S. 1987. Sedimentological model and production characteristics of Hartzog Draw Field, Wyoming. In Reservoir Sedimentology (ed. Tillman, R. W. & Weber, K. J.), pp. 15–112. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Tinkler, K. J., & Wohl, E. E. 1998. Rivers Over Rock: Fluvial Processes in Bedrock Channels. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Thorne, C. R., Russell, A. P. G., & Alam, M. K. 1993. Planform pattern and channel evolution of Brahmaputra River, Bangladesh. In Braided Rivers (ed. Best, J. L. & Bristow, C. S.), pp. 257–276. London: Geological Society of London.Google Scholar
Törnqvist, T. E. 1993. Holocene alternation of meandering and anastomosing fluvial systems in the Rhine–Meuse delta (central Netherlands) controlled by sea-level rise and subsoil erodibility. Journal of Sedimentary Petrology 63, 683–693.Google Scholar
Törnqvist, T. E. 1994. Middle and late Holocene avulsion history of the River Rhine (Rhine–Meuse delta, Netherlands). Geology 22, 711–714.2.3.CO;2>CrossRefGoogle Scholar
Törnqvist, T. E., & Bridge, J. S. 2002. Spatial variation of overbank aggradation rate and its influence on avulsion frequency. Sedimentology 49, 891–905.CrossRefGoogle Scholar
Törnqvist, T. E., Kidder, T. R., Autin, W. J.et al. 1996. A revised chronology for Mississippi River subdeltas. Science 273, 1693–1696.CrossRefGoogle Scholar
Törnqvist, T. E., Wallinga, J., Murray, A. S.et al. 2000. Response of the Rhine–Meuse system (west-central Netherlands) to the last Quaternary glacio-eustatic cycles. Global and Planetary Change 27, 89–111.CrossRefGoogle Scholar
Traverse, A., & Ginsburg, R. N. 1966. Palynology of the surface sediments of Great Bahama Bank, as related to water movements and sedimentation. Marine Geology 4, 417–459.CrossRefGoogle Scholar
Trendall, A. F., & Morris, R. C. 1983. Iron-Formation: Facts and Problems. Amsterdam: Elsevier.Google Scholar
Trendall, A. F., & Blockley, J. G. 1970. The iron formations of the Precambrian Hamersley Group, Western Australia. Geological Survey of Western Australia Bulletin 119.
Trewin, N. H. 1994. Depositional environment and preservation of biota in the Lower Devonian hot-springs of Rhynie, Aberdeenshire, Scotland. Transactions of the Royal Society of Edinburgh: Earth Sciences 84, 433–442.CrossRefGoogle Scholar
Trewin, N. H., & Rice, C. M. 1992. Stratigraphy and sedimentology of the Devonian Rhynie Chert locality. Scottish Journal of Geology 28, 37–47.CrossRefGoogle Scholar
Tripati, A., & Elderfield, H. 2005. Deep-sea temperature and circulation changes at the Paleocene–Eocene Thermal Maximum. Science 308, 1894–1898.CrossRefGoogle ScholarPubMed
Tsoar, H. 1982. Internal structure and surface geometry of longitudinal (seif) dunes. Journal of Sedimentary Petrology 52, 823–831.Google Scholar
Tsoar, H. 1983. Dynamic processes acting on a longitudinal (seif) dune. Sedimentology 30, 567–578.CrossRefGoogle Scholar
Tsoar, H. 1984. The formation of seif dunes from barchans – a discussion. Zeitschrift für Geomorphologie 28, 99–103.Google Scholar
Tsoar, H. 1989. Linear dunes – forms and formation. Progress in Physical Geography 13, 507–528.CrossRefGoogle Scholar
Tsoar, H., & Pye, K. 1987. Dust transport and the question of desert loess formation. Sedimentology 34, 139–153.CrossRefGoogle Scholar
Tucker, G. E., & Slingerland, R. L. 1996. Predicting sediment flux from fold and thrust belts. Basin Research 8, 329–349.CrossRefGoogle Scholar
Tucker, G. E., & Slingerland, R. L. 1997. Drainage basin responses to climate change. Water Resources Research 33, 2031–2047.CrossRefGoogle Scholar
Tucker, G. E., Lancaster, S. T., Gasparini, N. M., & Bras, R. L. 2002. The Channel-Hillslope Integrated Landscape Development Model (CHILD). In Landscape Erosion and Evolution Modeling (ed. Harmon, R. S. & Doe, W. W. III), pp. 349–388. New York: Kluwer Academic Publishing.Google Scholar
Tucker, M. E. 2001. Sedimentary Petrology: An Introduction to the Origin of Sedimentary Rocks. Oxford: Blackwell Scientific.Google Scholar
Tucker, M. E., & Wright, V. P. 1990. Carbonate Sedimentology. Oxford: Blackwells.CrossRefGoogle Scholar
Turcotte, D. L., & Schubert, G. 2002. Geodynamics, 2nd edn. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Turner, E. C., Narbonne, G. M., & James, N. P. 1993. Neoproterozoic reef microstructures from the Little Dal Group, northwestern Canada. Geology 21, 259–262.2.3.CO;2>CrossRefGoogle Scholar
Turner, E. C., Narbonne, G. M., & James, N. P.2000a. Framework composition of Early Neoproterozoic calcimicrobial reefs and associated microbialites, Mackenzie Mountains, N.W.T., Canada. In Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World (ed. Grotzinger, J. P. & James, N. P.), pp. 179–205. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Turner, E. C., Narbonne, G. M., & James, N. P. 2000b. Taphonomic control on microstructures in Early Neoproterozoic reefal stromatolites and thrombolites. Palaios 15, 87–111.2.0.CO;2>CrossRefGoogle Scholar
Tuttle, M. P., Ruffman, A., Anderson, T., & Jeter, H. 2004. Distinguishing tsunami deposits from storm deposits along the coast of northeastern North America: lessons learned from the 1929 Grand Banks tsunami and the 1991 Halloween storm. Seismological Research Letters 75, 117–131.CrossRefGoogle Scholar
Twichell, D. C., Kenyon, N. H., Parson, L. M., & McGregor, B. A. 1991. Depositional patterns of the Mississippi Fan surface: evidence from GLORIA II and high resolution seismic profiles. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 349–364. New York: Springer-Verlag.CrossRefGoogle Scholar
Tye, R. S. 1984. Geomorphic evolution and stratigraphy of Price and Capers Inlets, South Carolina. Sedimentology 31, 655–674.CrossRefGoogle Scholar
Tye, R. S. 2004. Geomorphology: an approach to determining subsurface reservoir dimensions. American Association of Petroleum Geologists Bulletin 88, 1123–1147.CrossRefGoogle Scholar
Tye, R. S., & Coleman, J. M. 1989a. Depositional processes and stratigraphy of fluvially dominated lacustrine deltas: Mississippi Delta Plain. Journal of Sedimentary Petrology 59, 973–996.Google Scholar
Tye, R. S., & Coleman, J. M. 1989b. Evolution of Atchafalaya lacustrine deltas, south-central Louisiana. Sedimentary Geology 65, 95–112.CrossRefGoogle Scholar
Tye, R. S., & Moslow, T. F. 1993. Tidal inlet reservoirs: insights from modern examples. In Marine Clastic Reservoirs: Examples and Analogues (ed. Rhodes, E. G. & Moslow, T. F.), pp. 77–99. New York: Springer-Verlag.CrossRefGoogle Scholar
Tyler, K., Henriquez, A., & Svanes, T. 1994. Modeling heterogeneities in fluvial domains: a review of the influence on production profiles. In Stochastic Modeling and Geostatistics (ed. Yarus, J. M. & Chambers, R. L.), pp. 77–89. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
US Department of Agriculture Soil Survey Staff 1993. Soil Survey Manual. Washington, D.C.: United States Department of Agriculture.
US Department of Agriculture Soil Survey Staff2003. Keys to Soil Taxonomy. Washington, D.C.: United States Department of Agriculture.
Usiglio, J. 1894. Analyse de l'eau de la Mediterranée sur la côte de France. Annales de Chimie et de Physique 27, 172–191.Google Scholar
Vail, P. R., Mitchum, R. M. Jr, Todd, R. G. et al. 1977. Seismic stratigraphy and global changes in sea-level. In Seismic Stratigraphy – Applications to Hydrocarbon Exploration (ed. Payton, C. E.), pp. 49–212. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Vallance, J. W., & Scott, K. M. 1997. The Osceola Mudflow from Mount Rainier: sedimentology and hazard implications of a huge clay-rich debris flow. Geological Society of America Bulletin 109, 143–163.2.3.CO;2>CrossRefGoogle Scholar
Andel, , , T. H., & Komar, P. D. 1969. Ponded sediments of the Mid-Atlantic Ridge between 22° and 23° north latitude. Geological Society of America Bulletin 80, 1163–1190.CrossRefGoogle Scholar
Berg, J. H. 1982. Migration of large-scale bedforms and preservation of crossbedded sets in highly accretional parts of tidal channels in the Oosterschelde, S.W. Netherlands. Geologie en Mijnbouw 61, 253–263.Google Scholar
Berg, J. H. 1995. Prediction of alluvial channel pattern of perennial rivers. Geomorphology 12, 259–279.CrossRefGoogle Scholar
Van Den Berg, J. H., & Van Gelder, A. 1993. A new bedform stability diagram, with emphasis on the transition of ripples to plane bed in flows over fine sand and silt. In Alluvial Sedimentation (ed. Marzo, M. & Puidefabregas, C.), pp. 11–21. Oxford: Blackwells.CrossRefGoogle Scholar
Berg, J. H., Gelder, A., & Mastbergen, D. R. 2002. The importance of breaching as a mechanism of subaqueous slope failure in fine sand. Sedimentology 49, 81–95.CrossRefGoogle Scholar
Van der Wateren, F. M. 2002. Processes of glaciotectonism. In Modern and Past Glacial Environments (ed. Menzies, J.), pp. 417–443. Oxford: Butterworth-Heinemann.CrossRefGoogle Scholar
Dyke, M. 1982. An Album of Fluid Motion. Stanford, CA: Parabolic Press.Google Scholar
Heerden, I. L., & Roberts, H. H. 1988. Facies development of Atchafalaya Delta, Louisiana: a modern bayhead delta. American Association of Petroleum Geologists Bulletin 72, 439–453.Google Scholar
Heijst, M. W. I. M., & Postma, G. 2001. Fluvial response to sea-level changes: a quantitative analogue, experimental approach. Basin Research 13, 269–292.CrossRefGoogle Scholar
Houten, F. B. 1964. Cyclic lacustrine sedimentation, Upper Triassic Lockatong Formation, central New Jersey and adjacent Pennsylvania. Kansas Geological Survey Bulletin 169, 497–532.Google Scholar
Houten, F. B. 1985. Oolitic ironstones and contrasting Ordovician and Jurassic paleogeography. Geology 13, 722–724.2.0.CO;2>CrossRefGoogle Scholar
Houten, F. B. & Bhattacharyya, P. D. 1982. Phanerozoic oolitic ironstones – geologic record and facies model. Annual Review of Earth and Planetary Sciences 10, 441–457.CrossRefGoogle Scholar
Lith, Y., Warthmann, R., Vasconcelos, C., & McKenzie, J. A. 2003. Microbial fossilization in carbonate sediments: a result of the bacterial surface involvement in dolomite precipitation. Sedimentology 50, 237–245.CrossRefGoogle Scholar
Lith, Y., Warthmann, R., Vasconcelos, C., & McKenzie, J. A. 2004. Sulfphate-reducing bacteria induce low-temperature Ca-dolomite and high Mg-calcite formation. Geobiology 1, 71–79.CrossRefGoogle Scholar
Niekerk, A.Vogel, A. K., Slingerland, R., & Bridge, J. S. 1992. Routing heterogeneous size–density sediments over a moveable bed: model development. Journal of Hydraulic Engineering, ASCE 118, 246–262.CrossRefGoogle Scholar
Rijn, L. C. 1990. Principles of Fluid Flow and Surface Waves in Rivers, Estuaries, Seas, and Oceans. Amsterdam: Aqua Publications.Google Scholar
Rijn, L. C. 1993. Principles of Sediment Transport in Rivers, Estuaries and Coastal Seas. Amsterdam: Aqua Publications.Google Scholar
Van Tassell, J. 1994. Cyclic deposition of the Devonian Catskill Delta of the Appalachian, U.S.A. In Orbital Forcing and Cyclic Sequences (ed. Boer, P. L. & Smith, D. G.), pp. 395–411. Oxford: Blackwells.CrossRefGoogle Scholar
Wagoner, J. C., & Bertram, G. T. 1995. Sequence Stratigraphy of Foreland Basin Deposits. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wagoner, J. C., Mitchum, R. M. Jr., Campion, K. M., & Rahmanian, V. D. 1990. Siliciclastic Sequence Stratigraphy in Well Logs, Cores and Outcrop: Concepts for High Resolution Correlation of Time and Facies. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Veizer, J., Godderis, Y., & Francois, L. M. 2000. Evidence for decoupling of atmospheric CO2 and global climate during the Phanerozoic eon. Nature 408, 698–701.CrossRefGoogle ScholarPubMed
Verway, J. 1952. On the ecology of distribution of cockle and mussel in the Dutch Waddensee, their role in sedimentation and the source of their food supply. Archives Néerlandaises de Zoologie 10, 171–240.CrossRefGoogle Scholar
Vigilar, G. G. Jr., & Diplas, P. 1997. Stable channels with mobile bed: formulation and numerical solution. Journal of Hydraulic Engineering, ASCE 123, 189–199.CrossRefGoogle Scholar
Vigilar, G. G. Jr., & Diplas, P. 1998. Stable channels with mobile bed: model verification and graphical solution. Journal of Hydraulic Engineering, ASCE 124, 1097–1108.CrossRefGoogle Scholar
Vincent, C. E., Young, R. A., & Swift, D. J. P. 1982. On the relationship between bedload and suspended sand transport on the inner shelf, Long Island, New York. Journal of Geophysical Research 87, 4163–4170.CrossRefGoogle Scholar
Vinopal, R. J., & Coogan, A. H. 1978. Effect of particle shape on the packing of carbonate sands and gravels. Journal of Sedimentary Petrology 48, 7–24.Google Scholar
Visscher, P. T., & Stolz, J. F. 2005. Microbial mats as bioreactors: populations, processes, and products. Palaeogeography, Palaeoclimatology, Palaeoecology 219, 87–100.CrossRefGoogle Scholar
Visser, M. J. 1980. Neap–spring cycles reflected in Holocene subtidal large-scale bedform deposits: a preliminary note. Geology 8, 543–546.2.0.CO;2>CrossRefGoogle Scholar
Vogel, K., Niekerk, A., Slingerland, R., & Bridge, J. S. 1992. Routing of heterogeneous size–density sediments over a moveable bed: model verification and testing. Journal of Hydraulic Engineering, ASCE 118, 263–279.CrossRefGoogle Scholar
Vose, R. S., Schmoyer, R. L., Steurer, P. M., Peterson, T. C., Heim, R., Karl, T. R., & Eischeid, J. K. 1992. The Global Historical Climatology Network: Long-term Monthly Temperature, Precipitation, Sea Level Pressure, and Station Pressure Data. Oak Ridge, TN: Oak Ridge National Laboratory Environmental Sciences Division.Google Scholar
Walgreen, M., Calvete, D., & Swart, H. E. 2002. Growth of large-scale bedforms due to storm-driven and tidal currents: a model approach. Continental Shelf Research 22, 2777–2793.CrossRefGoogle Scholar
Walker, R. G. 1965. The origin and significance of the internal structures of turbidites. Proceedings of the Yorkshire Geological Society 35, 1–32.CrossRefGoogle Scholar
Walker, R. G. 1978. Deep water sandstone facies and ancient submarine fans: models for exploration for stratigraphic traps. American Association of Petroleum Geologists Bulletin 62, 932–966.Google Scholar
Walker, R. G.1979. Turbidites and associated coarse clastic deposits. In Facies Models (ed. Walker, R. G.), pp. 91–103. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G.1984. Shelf and shallow marine sands. In Facies Models, 2nd edn. (ed. Walker, R. G.), pp. 141–170. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G.1992. Turbidites and submarine fans. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 239–263. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G., & James, N. P. 1992. Facies Models: Response to Sea Level Change. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G., & Mutti, E. 1973. Turbidite facies and facies associations. In Turbidites and Deep-Water Sedimentation (ed. Middleton, G. V. & Bouma, A. H.), pp. 19–157. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Walker, R. G., & Plint, A. G. 1992. Wave- and storm-dominated shallow marine systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 219–238. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walter, M. R. 1976. Stromatolites. Amsterdam: Elsevier.Google Scholar
Walther, J. V. 2005. Essentials of Geochemistry. Sudbury, MA: A. Jones and Bartlett Publishers.Google Scholar
Wanless, H. R. 1979. Limestone response to stress: pressure solution and dolomitization. Journal of Sedimentary Petrology 49, 437–462.Google Scholar
Ward, A. W., & Greeley, R. 1984. The yardangs at Rogers Lake, California. Geological Society of America Bulletin 95, 829–837.2.0.CO;2>CrossRefGoogle Scholar
Ward, W. B. 1996. Evolution and diagenesis of Frasnian carbonate platforms, Devonian Reef Complexes, Napier Range, Canning Basin, Western Australia. Unpublished Ph.D. dissertation, State University of New York at Stony Brook, New York.
Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setata, H., Putten, W. H., & Wall, D. H. 2004. Ecological linkages between aboveground and belowground biota. Science 304, 1629–1633.CrossRefGoogle ScholarPubMed
Warren, J. K. 1982. The hydrological significance of Holocene tepees, stromatolites, and boxwork limestones in coastal Salinas in south Australia. Journal of Sedimentary Petrology 52, 1171–1201.Google Scholar
Warren, J. K. 2006. Evaporites: Sediments, Resources and Hydrocarbons. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Warren, W. P., & Ashley, G. M. 1994. Origins of the ice-contact ridges (eskers) of Ireland. Journal of Sedimentary Research 64, 433–449.Google Scholar
Watabe, N., & Wilbur, K. M. 1976. The Mechanisms of Mineralization in the Invertebrates and Plants. Columbia, SC: University of South Carolina Press.Google Scholar
Waters, B. B., Spencer, R. J., & Demicco, R. V. 1989. Three-dimensional architecture of shallowing-upward carbonate cycles: Middle and Upper Cambrian Waterfowl Formation, southern Canadian Rocky Mountains. Bulletin of Canadian Petroleum Geology 37, 198–209.Google Scholar
Watts, N. L. 1977. Pseudo-anticlines and other structures in some calcretes of Botswana and South Africa. Earth Surface Processes and Landforms 2, 63–74.CrossRefGoogle Scholar
Watts, N. L. 1980. Quaternary pedogenic calcretes from the Kalahari (southern Africa); mineralogy, genesis and diagenesis. Sedimentology 27, 661–686.CrossRefGoogle Scholar
Weaver, P. P. E., & Thomson, J. 1987. Geology and Geochemistry of Abyssal Plains. London: Geological Society of London.Google Scholar
Weertman, J. 1979. The unsolved general glacier sliding problem. Journal of Glaciology 23, 97–115.CrossRefGoogle Scholar
Weertman, J. 1986. Basal water and high-pressure basal ice. Journal of Glaciology 32, 455–463.CrossRefGoogle Scholar
Weimer, P. 1990. Sequence stratigraphy, facies geometries, and depositional history of the Mississippi Fan, Gulf of Mexico. American Association of Petroleum Geologists Bulletin 74, 425–453.Google Scholar
Weimer, P.1995. Sequence stratigraphy of the Mississippi Fan (late Miocene–Pleistocene), northern deep Gulf of Mexico. In Atlas of Deep Water Environments; Architectural Style in Turbidite Systems (ed. Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, F. Ricci, & Smith, R. D. A.), pp. 94–99. London: Chapman and Hall.CrossRefGoogle Scholar
Weimer, P., Bouma, A. H., & Perkins, B. F. 1994. Submarine Fans and Turbidite Systems. Austin, TX: Society of Economic Paleontologists and Mineralogists Gulf Coast Section.Google Scholar
Weimer, P., & Link, M. H. 1991. Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems. New York: Springer-Verlag.CrossRefGoogle Scholar
Weimer, P., & Posamentier, H. 1993. Siliciclastic Sequence Stratigraphy. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Weimer, R. J., Howard, J. D., & Lindsay, D. R. 1982. Tidal flats. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D. A.), pp. 191–245. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wellner, R. W., Ashley, G. M., & Sheridan, R. E. 1993. Seismic stratigraphic evidence for a submerged middle Wisconsin barrier: implications for sea-level history. Geology 21, 109–112.2.3.CO;2>CrossRefGoogle Scholar
Werner, B. T. 1995. Eolian dunes: computer simulations and attractor interpretation. Geology 23, 1107–1110.2.3.CO;2>CrossRefGoogle Scholar
Wescott, W. A. 1993. Geomorphic thresholds and complex response of fluvial systems – some implications for sequence stratigraphy. American Association of Petroleum Geologists Bulletin 77, 1208–1218.Google Scholar
Westbroek, P., & Jong, E. W. 1983. Biomineralization and Biological Metal Accumulation. Dordrecht: D. Reidel Publishing Company.CrossRefGoogle Scholar
Whateley, M. K. G., & Pickering, K. T. 1989. Deltas: Sites and Traps for Fossil Fuels. London: Geological Society of London.Google Scholar
Whipple, K. L. 1997. Open-channel flow of Bingham fluids: applications in debris-flow research. Journal of Geology 105, 243–262.CrossRefGoogle Scholar
Whipple, K. X., Hancock, G. S., & Anderson, R. S. 2000. River incision into bedrock: mechanics and relative efficacy of plucking, abrasion, and cavitation. Geological Society of America Bulletin 112, 490–503.2.0.CO;2>CrossRefGoogle Scholar
White, W. A. 1961. Colloid phenomena in sedimentation of argillaceous rocks. Journal of Sedimentary Petrology 31, 560–570.Google Scholar
Wignall, P. B. 1994. Black Shales. Oxford: Oxford University Press.Google Scholar
Wilgus, C. K., Hastings, B. S., Kendall, C. G. St. , C.et al. 1988. Sea-Level Changes: An Integrated Approach. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Wilken, R. T. 2003. Sulfide minerals in sediments. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 701–703. Dordrecht: Kluwer Academic Publishers.Google Scholar
Wilkinson, B. H. 1979. Biomineralization, paleoceanography and the evolution of calcareous marine organisms. Geology 7, 524–527.2.0.CO;2>CrossRefGoogle Scholar
Willgoose, G. R., Bras, R. L., & Rodriguez-Iturbe, I. 1991a. A physically-based coupled network growth and hillslope evolution model: 1. Theory. Water Resource Research 27, 1,671–1,684.CrossRefGoogle Scholar
Willgoose, G. R., Bras, R. L., & Rodriguez-Iturbe, I. 1991b. A physically-based coupled network growth and hillslope evolution model: 2. Applications. Water Resource Research 27, 1,685–1,696.CrossRefGoogle Scholar
Williams, G. P. 1988. Paleofluvial estimates from dimensions of former channels and meanders. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Baker, V. R., Kochel, R. C., & Patton, P. P.), pp. 321–334. Chichester: Wiley.Google Scholar
Willis, B. J. 1989. Paleochannel reconstructions from point bar deposits: a three-dimensional perspective. Sedimentology 36, 757–766.CrossRefGoogle Scholar
Willis, B. J. 1993a. Ancient river systems in the Himalayan foredeep, Chinji village area, northern Pakistan. Sedimentary Geology 88, 1–76.CrossRefGoogle Scholar
Willis, B. J.1993b. Interpretation of bedding geometry within ancient point-bar deposits In Alluvial Sedimentation (ed. Marzo, M. & Puigdefabregas, C.), pp. 101–114. Oxford: Blackwells.CrossRefGoogle Scholar
Willis, B. J. 1997. Architecture of fluvial-dominated valley-fill deposits in the Cretaceous Fall River Formation. Sedimentology 44, 735–757.CrossRefGoogle Scholar
Willis, B. J.2005. Deposits of tide-influenced river deltas. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J. P.), pp. 87–129. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Willis, B. J., & Behrensmeyer, A. K. 1994. Architecture of Miocene overbank deposits in Northern Pakistan. Journal of Sedimentary Research B64, 60–67.Google Scholar
Wilson, A. M., Sanford, W., Whitaker, F., & Smart, P. 2001. Spatial patterns of diagenesis during geothermal circulation in carbonate platforms. American Journal of Science 301, 727–752.CrossRefGoogle Scholar
Wilson, E. N., Hardie, L. A., & Phillips, O. M. 1990. Dolomitization front geometry, fluid flow patterns, and the origin of massive dolomite: the Triassic Latemar Buildup, northern Italy. American Journal of Science 290, 741–796.CrossRefGoogle Scholar
Wilson, I. G. 1972a. Aeolian bedforms – their development and origins. Sedimentology 19, 173–210.CrossRefGoogle Scholar
Wilson, I. G. 1972b. Universal discontinuities in bedforms produced by wind. Journal of Sedimentary Petrology 42, 667–669.Google Scholar
Wilson, J. L. 1975. Carbonate Facies in Geologic History. New York: Springer-Verlag.CrossRefGoogle Scholar
Wilson, J. L., & Jordan, C. 1983. Middle shelf. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 297–343. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wolf, K. H., Easton, A. J., & Warme, S. 1967. Techniques of examining and analyzing carbonate skeletons, minerals and rocks. In Carbonate Rocks (ed. Chilingar, G. V., Bissell, H. J., & Fairbridge, R. W.), pp. 253–341. Amsterdam: Elsevier.Google Scholar
Wollast, R. 1976. Kinetics of the alteration of K-feldspar in buffered solutions at low temperature. Geochimica et Cosmochimica Acta 31, 635–648.CrossRefGoogle Scholar
Wolman, M. G., & Miller, J. P. 1960. Magnitude and frequency of forces in geomorphic processes. Journal of Geology 68, 54–74.CrossRefGoogle Scholar
Wood, L. J., Ethridge, F. G., & Schumm, S. A. 1993. An experimental study of the influence of subaqueous shelf angles on coastal plain and shelf deposits. In Siliciclastic Sequence Stratigraphy (ed. Weimer, P. & Posamentier, H. W.), pp. 381–391. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wood, R. A., Grotzinger, J. P., & Dickson, J. A. D. 2002. Proterozoic modular biomineralized metazoan from the Nama Group, Namibia. Science 296, 2383–2386.CrossRefGoogle ScholarPubMed
Woodroffe, C. D. 2003. Coasts; Form, Process and Evolution. Cambridge: Cambridge University Press.Google Scholar
Woodroffe, C. D., Chappell, J., Thom, B. G., & Wallensky, E. 1989. Depositional model of a macrotidal estuary and floodplain, South Alligator River, Northern Australia. Sedimentology 36, 737–756.CrossRefGoogle Scholar
Wright, H. E. Jr., Kutzbach, J. E., Webb, T. IIIet al. 1993. Global Climates since the Last Glacial Maximum. Minneapolis, MN: University of Minnesota Press.Google Scholar
Wright, L. D. 1977. Sediment transport and deposition at river mouths: a synthesis. Geological Society of America Bulletin 88, 857–868.2.0.CO;2>CrossRefGoogle Scholar
Wright, L. D., & Coleman, J. M. 1973. Variations in morphology of major river deltas as functions of ocean wave and river discharge regimes. American Association of Petroleum Geologists Bulletin 57, 370–398.Google Scholar
Wright, L. D., & Coleman, J. M. 1974. Mississippi River mouth processes: effluent dynamics and morphological development. Journal of Geology 82, 751–778.CrossRefGoogle Scholar
Wright, L. D., Coleman, J. M., & Thom, B. G. 1975. Sediment transport and deposition in a macrotidal river channel, Ord River, Western Australia. In Estuarine Research II (ed. Cronin, L. E.), pp. 309–322. New York: Academic Press.Google Scholar
Wright, L. D., & Short, A. D. 1984. Morphodynamic variability of surf zones and beaches: a synthesis. Marine Geology 56, 93–118.CrossRefGoogle Scholar
Wright, V. P. 1986. Paleosols: Their Recognition and Interpretation. Oxford: Blackwell Scientific.Google Scholar
Wright, V. P.1999. Assessing flood duration gradients and fine-scale environmental change on ancient floodplains. In Floodplains: Interdisciplinary Approaches (ed. Marriott, S. B. & Alexander, J.), pp. 279–287. London: Geological Society of London.Google Scholar
Wright, V. P., & Burchette, T. P. 1996. Shallow-water carbonate environments. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 325–394. Oxford: Blackwells.Google Scholar
Wright, V. P., & Marriott, S. B. 1993. The sequence stratigraphy of fluvial depositional systems: the role of floodplain sediment storage. Sedimentary Geology 86, 203–210.CrossRefGoogle Scholar
Yalin, M. S. 1977. Mechanics of Sediment Transport, 2nd edn. Oxford: Pergamon Press.Google Scholar
Yalin, M. S. 1992. River Mechanics. Oxford: Pergamon Press.Google Scholar
Yao, Q., & Demicco, R. V. 1995. Paleoflow patterns of dolomitizing fluids and paleohydrogeology of southern Canadian Rocky Mountains: evidence from dolomite geometry and numerical modeling. Geology 23, 791–794.2.3.CO;2>CrossRefGoogle Scholar
Yao, Q., & Demicco, R. V. 1997. Dolomitization of the Cambrian carbonate platform, southern Canadian Rocky Mountains: dolomite front geometry, fluid inclusion geochemistry, isotopic signature, and hydrogeologic modeling studies. American Journal of Science 297, 892–938.CrossRefGoogle Scholar
Yarnold, Y. C. 1993. Rock-avalanche characteristics in dry climates and the effects of flow into lakes: insights from mid-Tertiary sedimentary breccias near Artillery Peak, Arizona. Geological Society of America Bulletin 105, 345–360.2.3.CO;2>CrossRefGoogle Scholar
Young, A. 1972. Slopes. Edinburgh: Oliver and Boyd.Google Scholar
Young, I. M., & Crawford, J. W. 2004. Interactions and self-organization in the soil-microbe complex. Science 304, 1634–1637.CrossRefGoogle ScholarPubMed
Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K. 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686–693.CrossRefGoogle ScholarPubMed
Zachos, J. C., Röhl, U., Schellenberg, S. A.et al. 2005. Rapid acidification of the ocean during the Paleocene–Eocene Thermal Maximum. Science 308, 1611–1615.CrossRefGoogle ScholarPubMed
Zajac, I. S. 1974. The Stratigraphy and Mineralogy of the Sokoman Formation in the Knob Lake Area, Quebec and Newfoundland. Ottawa: Geological Survey of Canada.Google Scholar
Zeng, J., & Lowe, D. R. 1997. Numerical simulations of turbidity current flow and sedimentation: I. Theory. Sedimentology 44, 67–84.CrossRefGoogle Scholar
Zharkov, M. A. 1984. Paleozoic Salt Bearing Formations of the World. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Zaitlin, B. A., Dalrymple, R. W., & Boyd, R. 1994. The stratigraphic organization of incised-valley systems associated with relative sea-level change. In Incised Valley Systems: Origin and Sedimentary Sequences (ed. Dalrymple, R. W., Boyd, R., & Zaitlin, B. A.), pp. 45–60. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Aber, J. S., Croot, D. G., & Fenton, M. M. 1989. Glaciotectonic Landforms and Structures. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Abrahams, A. D. 1986. Hillslope Processes. Boston: Allen and Unwin.Google Scholar
Abrahams, A. D., & Parsons, A. J. 1994. Geomorphology of Desert Environments. London: Chapman and Hall.CrossRefGoogle Scholar
Adams, A. E., & MacKenzie, W. S. 1998. A Colour Atlas of Carbonate Sediments and Rocks under the Microscope. London: Manson Publishing.Google Scholar
Adams, A. E., MacKenzie, W. S., & Guilford, C. 1984. Atlas of Sedimentary Rocks under the Microscope. New York: John Wiley and Sons.Google Scholar
Adams, J. E., & Rhodes, M. L. 1960. Dolomitization by seepage refluxion. American Association of Petroleum Geologists Bulletin 44, 1912–1921.Google Scholar
Ahlbrandt, T. S., & Fryberger, S. G. 1980. Eolian deposits in the Nebraska Sand Hills. United States Geological Survey Professional Paper1120A.
Ahlbrandt, T. S., & Fryberger, S. G.1981. Sedimentary features and significance of interdune deposits. In Recent and Ancient Nonmarine Depositional Environments: Models for Exploration (ed. Ethridge, F. G. and Flores, R. M.), pp. 293–314. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Aigner, T., & Reineck, H. -E. 1982. Proximality trends in modern storm sands from the Helgoland Bight (North Sea) and their implications for basin analysis. Senckenbergiana Maritima 14, 183–215.Google Scholar
Aitken, J. F., & Flint, S. S. 1995. The application of high-resolution sequence stratigraphy to fluvial systems: a case study from the Upper Carboniferous Breathitt Group, eastern Kentucky, USA. Sedimentology 42, 3–30.CrossRefGoogle Scholar
Alexander, C. R, Davis, R. A., & Henry, V. J. 1998. Tidalites: Processes and Products. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Alexander, C. R., Nittrouer, C. A., DeMaster, D. J., Park, Y. A., & Park, S. C. 1991. Macrotidal mudflats of the southwestern Korean coast: a model for interpretation of intertidal deposits. Journal of Sedimentary Petrology 61, 805–824.Google Scholar
Alexander, J., Bridge, J. S., Cheel, R. J., & Leclair, S. F. 2001. Bed forms and associated sedimentary structures formed under supercritical water flows over aggrading sand beds. Sedimentology 48, 133–152.CrossRefGoogle Scholar
Alexander, J., Bridge, J. S., Leeder, M. R., Collier, R. E. L., & Gawthorpe, R. L. 1994. Holocene meander belt evolution in an active extensional basin, southwestern Montana. Journal of Sedimentary Research B64, 542–559.Google Scholar
Alexander, J., & Leeder, M. R. 1987. Active tectonic control on alluvial architecture. In Recent Developments in Fluvial Sedimentology (ed. Ethridge, F. G., Flores, R. M., & Harvey, M. D.), pp. 243–252. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Alexander, J., & Morris, S. 1994. Observations on the experimental, nonchannelised, high-concentration turbidity currents and variations in deposits around obstacles. Journal of Sedimentary Research 64A, 899–909.Google Scholar
Algeo, T. J., Scheckler, S. E., & Maynard, J. B. 2001. Effects of the Middle to Late Devonian spread of vascular plants on weathering regimes, marine biotas, and global climate. In Plants Invade the Land (ed. Gensel, P. G. & Edwards, D.), pp. 216–236. New York: Columbia University Press.CrossRefGoogle Scholar
Algeo, T. J., & Wilkinson, B. H. 1988. Periodicity of mesoscale Phanerozoic sedimentary cycles and the role of Milankovitch orbital modulation. Journal of Geology 96, 313–322.CrossRefGoogle Scholar
Allen, G. P. 1991. Sedimentary processes and facies in the Gironde estuary: a recent model for macrotidal estuarine systems. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 29–40. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Allen, J. R. L. 1965. A review of the origin and characteristics of recent alluvial sediments. Sedimentology 5, 89–191.CrossRefGoogle Scholar
Allen, J. R. L. 1966. On bedforms and paleocurrents. Sedimentology 6, 153–190.CrossRefGoogle Scholar
Allen, J. R. L. 1970. Physical Processes of Sedimentation. London: George Allen and Unwin.Google Scholar
Allen, J. R. L. 1973. Features of cross-stratified units due to random and other changes in bed forms. Sedimentology 20, 189–202.CrossRefGoogle 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, 181–208.CrossRefGoogle Scholar
Allen, J. R. L. 1978. Studies in fluviatile sedimentation: an exploratory quantitative model for architecture of avulsion-controlled alluvial suites. Sedimentary Geology 21, 129–147.CrossRefGoogle Scholar
Allen, J. R. L. 1979a. A model for the interpretation of wave-ripple marks using their wavelength, textural composition and shape. Journal of the Geological Society of London 136, 673–682.CrossRefGoogle Scholar
Allen, J. R. L. 1979b. Studies in fluviatile sedimentation: an elementary geometrical model for the connectedness of avulsion-related channel sand bodies. Sedimentary Geology 24, 253–267.CrossRefGoogle Scholar
Allen, J. R. L. 1980. Sand waves: a model of origin and internal structure. Marine Geology 26, 281–328.Google Scholar
Allen, J. R. L. 1982a. Sedimentary Structures: Their Character and Physical Basis. Amsterdam: Elsevier Science Publishers.Google Scholar
Allen, J. R. L. 1982b. Mud drapes in sand-wave deposits: a physical model with application to the Folkestone beds (early Cretaceous, southeast England). Philosophical Transactions of the Royal Society of London 306A, 291–345.CrossRefGoogle Scholar
Allen, J. R. L. 1985. Principles of Physical Sedimentology. London: George Allen and Unwin.Google Scholar
Allen, J. R. L. 1991. The Bouma division A and the possible duration of turbidity currents. Journal of Sedimentary Petrology 61, 291–295.Google Scholar
Allen, J. R. L., & Banks, N. L. 1972. An interpretation and analysis of recumbent-folded deformed cross bedding. Sedimentology 19, 257–283.CrossRefGoogle Scholar
Allen, J. R. L., & Pye, K. 1992. Saltmarshes. Cambridge: Cambridge University Press.Google Scholar
Allen, P. A. 1981. Some guidelines in reconstructing ancient sea conditions from wave ripple marks. Marine Geology 43, M59–67.CrossRefGoogle Scholar
Allen, P. A. 1985. Hummocky cross-stratification is not produced purely under progressive gravity waves. Nature 313, 562–564.CrossRefGoogle Scholar
Allen, P. A. 1997. Earth Surface Processes. Oxford: Blackwells.CrossRefGoogle Scholar
Allen, P. A., & Allen, J. R. 2005. Basin analysis, 2nd edn. Oxford: Blackwells.Google Scholar
Allen, P. A., & Densmore, A. L. 2000. Sediment flux from an uplifting fault block. Basin Research 12, 367–380.CrossRefGoogle Scholar
Allen, P. A., & Homewood, P. 1984. Evolution and mechanics of a Miocene tidal sand wave. Sedimentology 31, 63–81.CrossRefGoogle Scholar
Alley, R. B. 1991. Deforming-bed origin for southern Laurentide till sheets. Journal of Glaciology 37, 67–76.CrossRefGoogle Scholar
Alley, R. B., Lawson, D. E., Evenson, E. B., Larson, G. J., & Baker, G. S. 2003. Stabilizing feedbacks in glacier bed erosion. Nature 424, 758–760.CrossRefGoogle ScholarPubMed
Alley, R. B., Lawson, D. E., Evenson, E. B., Strasser, J. C., & Larson, G. J. 1998. Glaciohydraulic supercooling: a freeze-on mechanism to create stratified, debris-rich basal ice. 2. Theory. Journal of Glaciology 44, 563–569.CrossRefGoogle Scholar
Alsharhan, A. S., & Kendall, C. G.St, C. 2003. Holocene coastal carbonates and evaporites of the southern Arabian Gulf and their ancient analogues. Earth-Science Reviews 61, 191–243.CrossRefGoogle Scholar
American Society of Civil Engineers Task Force on Bed Forms in Alluvial Channels 1966. Nomenclature for bedforms in alluvial channels. Journal of the Hydraulics Division ASCE, 92, 51–64.
Amos, C. L., Li, M. Z., & Choung, K.-S. 1996. Storm-generated, hummocky stratification on the outer-Scotian shelf. Geomarine Letters 16, 85–94.Google Scholar
Anadón, P., Cabrera, L., & Kelts, K. 1991. Lacustrine Facies Analysis. Oxford: Blackwells.CrossRefGoogle Scholar
Anbar, A. D., & Knoll, A. H. 2002. Proterozoic ocean chemistry and evolution: a bioinorganic bridge?Science 297, 1137–1192.CrossRefGoogle ScholarPubMed
Anderson, J. B., & Ashley, G. M. 1991. Glacial Marine Sedimentation: Palaeoclimatic Significance. Boulder, CO: Geological Society of America.Google Scholar
Anderson, M. G., Walling, D. E., & Bates, P. D. 1996. Floodplain Processes. Chichester: Wiley.Google Scholar
Anderson, M. P. 1997. Characterization of geological heterogeneity. In Subsurface Flow and Transport: A Stochastic Approach (ed. Dagan, G. & Neuman, S. P.), pp. 23–43. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Anderson, R. S. 1987. A theoretical model of aeolian impact ripples. Sedimentology 34, 943–956.CrossRefGoogle Scholar
Anderson, R. S., & Bunas, K. L. 1993. Grain size segregation and stratigraphy in aeolian ripples modeled with a cellular automaton. Nature 365, 740–743.CrossRefGoogle Scholar
Archer, A. W. 1998. Hierarchy of controls on cyclic rhythmite deposition: Carboniferous basins of eastern and mid-continental USA. In Tidalites: Processes and Products (ed. Alexander, C. R., Davis, R. A., & Henry, V. J.), pp. 59–68. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Arnott, R. W. C., & Hand, B. M. 1989. Bedforms, primary structures and grain fabric in the presence of sediment rain. Journal of Sedimentary Petrology 59, 1062–1069.Google Scholar
Arnott, R. W., & Southard, J. B. 1990. Experimental study of combined-flow bed configurations in fine sands, and some implications for stratification. Journal of Sedimentary Petrology 60, 211–219.Google Scholar
Aronson, J. L., & Hower, J. 1976. Mechanism of burial metamorphism of argillaceous sediment: 2, radiogenic argon evidence. Geological Society of America Bulletin 87, 738–744.2.0.CO;2>CrossRefGoogle Scholar
Arthur, M. A., Dean, W. E., & Stow, D. A. V. 1984. Models for the deposition of Mesozoic–Cenozoic fine-grained organic-carbon-rich sediment in the deep sea. In Fine-Grained Sediments: Deep-Water Processes and Facies (ed. Stow, D. A. V. & Piper, D. J. W.), pp. 527–560. London: Geological Society of London.Google Scholar
Arthur, M. A., & Sageman, B. B. 1994. Marine black shales: depositional mechanisms and environments of ancient deposits. Annual Review of Earth and Planetary Sciences 22, 499–551.CrossRefGoogle Scholar
Ashley, G. M. 1988. The hydrodynamics and sedimentology of a back-barrier lagoon–salt marsh system, Great Sound, New Jersey. Marine Geology 82, 1–132.Google Scholar
Ashley, G. M. 1990. Classification of large-scale subaqueous bed-forms: a new look at an old problem. Journal of Sedimentary Petrology 60, 160–172.Google Scholar
Ashley, G. M.2002. Glaciolacustrine environments. In Modern and Past Glacial Environments (ed. Menzies, J.), pp. 335–359. Oxford: Butterworth-Heinemann.CrossRef
Ashley, G. M., Shaw, J., & Smith, N. D. 1985. Glacial Sedimentary Environments. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ashley, G. M., Wellner, R. W., Esker, D., & Sheridan, R. E. 1991. Clastic sequences developed during late Quaternary glacio-eustatic sea-level fluctuations on a passive margin: example from the inner continental shelf near Barnegat Inlet, New Jersey. Geological Society of America Bulletin 103, 1607–1621.2.3.CO;2>CrossRefGoogle Scholar
Ashworth, P. J., Best, J. L., & Jones, M. 2004. Relationship between sediment supply and avulsion frequency in braided rivers. Geology 32, 21–24.CrossRefGoogle Scholar
Aslan, A., & Blum, M. D. 1999. Contrasting styles of Holocene avulsion, Texas Gulf Coastal Plain, USA. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 193–209 J. Oxford: Blackwells.CrossRefGoogle Scholar
Assereto, R., & Folk, R. L. 1980. Diagenetic fabrics of aragonite, calcite, and dolomite in an ancient peritidal-spelean environment: Triassic Calcare Rosso, Lombardia, Italy. Journal of Sedimentary Petrology 50, 371–394.Google Scholar
Atwater, B. F., & Hemphill-Haley, E. 1997. Recurrence intervals for great earthquakes of the past 3,500 years at northeastern Willapa Bay, Washington. United States Geological Survey Professional Paper 1576.
Atwater, B. F., Musumi-Rokaku, S., Satake, K. et al. 2005. The Orphan Tsunami of 1700: Japanese clues to a parent earthquake in North America. United States Geological Survey Professional Paper 1707.
Augustinius, P. G. E. F. 1989. Cheniers and chenier plains: a general introduction. Marine Geology 90, 219–229.CrossRefGoogle Scholar
Augustinius, P. G. E. F., Hazelhoff, L., & Kroon, A. 1989. The chenier coast of Suriname; modern and geological development. Marine Geology 90, 269–281.CrossRefGoogle Scholar
Ausich, W. I., & Bottjer, D. J. 1982. Tiering in suspension-feeding communities on soft substrata throughout the Phanerozoic. Science 216, 173–174.CrossRefGoogle ScholarPubMed
Ausich, W. I., & Bottjer, D. J.1985. Phanerozoic tiering in suspension-feeding communities on soft substrata: implications for diversity. In Phanerozoic Diversity Patterns (ed. Valentine, J. W.), pp. 255–274. New Jersey: Princeton University Press.Google Scholar
Autin, W. J., Burns, S. F., Miller, R. T., Saucier, R. T., & Snead, J. I. 1991. Quaternary geology of the lower Mississippi valley. In Quaternary Nonglacial Geology: Conterminous U.S. The Geology of North America, pp. 547–582. Boulder, CO: Geological Society of America.Google Scholar
Baas, J. H., Kesteren, W., & Postma, G. 2004. Deposits of depletive high-density turbidity currents: a flume analogue of bed geometry, structure and texture. Sedimentology 51, 1053–1088.CrossRefGoogle Scholar
Badiozamani, K. 1973. The dorag dolomitization model – application to the Middle Ordovician of Wisconsin. Journal of Sedimentary Petrology 43, 965–984.Google Scholar
Baeuerlein, E. 2004. Biomineralization: Progress in Biology, Molecular Biology and Application. Weinheim: Wiley-VCH.CrossRefGoogle Scholar
Bagnold, R. A. 1941. The Physics of Blown Sand and Desert Dunes. London: Methuen.Google Scholar
Bagnold, R. A. 1946. Motion of waves in shallow water. Interaction between waves and sand bottoms. Proceedings of the Royal Society of London 187A, 1–16.Google Scholar
Bagnold, R. A. 1954. Experiments on the gravity-free dispersion of large solid spheres in a Newtonian fluid under shear. Proceedings of the Royal Society of London 187A, 49–63.CrossRefGoogle Scholar
Bagnold, R. A. 1956. The flow of cohesionless grains in fluids. Philosophical Transactions of the Royal Society 249A, 235–297.CrossRefGoogle Scholar
Bagnold, R. A. 1960. Some aspects of the shape of river meanders. United States Geological Survey Professional Paper 282E, pp. 135–144.Google Scholar
Bagnold, R. A. 1962. Auto-suspension of transported sediment: turbidity currents. Philosophical Transactions of the Royal Society 265A, 315–319.Google Scholar
Bagnold, R. A.1963. Mechanics of marine sedimentation. In The Sea: Volume 3 (ed. Hill, M. N.), pp. 507–528. New York: Wiley.Google Scholar
Bagnold, R. A. 1966. An approach to the sediment transport problem from general physics. United States Geological Survey Professional Paper 442-I.
Bagnold, R. A. 1973. The nature of saltation and of “bed-load” transport in water. Philosophical Transactions of the Royal Society 332A, 473–504.Google Scholar
Baker, V. R. 1973. Palaeohydrology and sedimentology of Lake Missoula flooding in eastern Washington. Geological Society of America Special Paper 144.
Baker, V. R. 1990. Geological fluvial geomorphology. Geological Society of America Bulletin 100, 1157–1167.2.3.CO;2>CrossRefGoogle Scholar
Baker, V. R., Kochel, R. C., & Patton, P. C. 1988. Flood Geomorphology. New York: Wiley.Google Scholar
Baker, V. R., & Nummedal, D. 1978. The Channeled Scabland. Washington, D.C.: NASA Planetary Geology Program.Google Scholar
Baldwin, B., & Butler, C. O. 1985. Compaction curves. American Association of Petroleum Geologists Bulletin 69, 622–626.Google Scholar
Ball, M. M. 1967. Carbonate sand bodies of Florida and the Bahamas. Journal of Sedimentary Petrology 37, 556–591.Google Scholar
Banerjee, I., & McDonald, B. C. 1975. Nature of esker sedimentation. In Glaciofluvial and Glaciolacustrine Sediments (ed. Jopling, A. V. & McDonald, B. C.), pp. 132–154. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Barnes, M. A., Barnes, W. C., & Bustin, R. M. 1990. Chemistry and diagenesis of organic matter in sediments and fossil fuels. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 189–204. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Barnes, R. S. K., & Hughes, R. N. 1982. An Introduction to Marine Ecology. Oxford: Blackwells.Google Scholar
Barry, J. M. 1997. Rising Tide: The Great Mississippi Flood of 1927 and How It Changed America. New York: Simon and Schuster.Google Scholar
Barwis, J. H. 1978. Sedimentology of some South Carolina tidal creek point bars, and a comparison with their fluvial counterparts. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 129–160. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Bates, C. C. 1953. Rational theory of delta formation. American Association of Petroleum Geologists Bulletin 37, 2119–2162.Google Scholar
Bathurst, R. G. C. 1975. Carbonate Sediments and Their Diagenesis, 2nd edn. Amsterdam: Elsevier.Google Scholar
Bathurst, R. G. C. 1980. Lithification of carbonate sediments. Science Progress Oxford 66, 451–471.Google Scholar
Bathurst, R. G. C. 1987. Diagenetically enhanced bedding in argillaceous platform limestones: stratified cementation and selective compaction. Sedimentology 34, 749–778.CrossRefGoogle Scholar
Batten, K. L., Narbonne, G. M., & James, N. P. 2004. Paleoenvironments and growth of early Neoproterozoic calcimicrobial reefs: platformal Little Dal Group, northwestern Canada. Precambrian Research 133, 249–269.CrossRefGoogle Scholar
Bear, J. 1972. Dynamics of Fluids in Porous Media. New York: Dover Publications.Google Scholar
Beard, B. L., Johnson, C. M., Cox, L.et al. 1999. Iron isotope biosignatures. Science 285, 1889–1892.CrossRefGoogle ScholarPubMed
Beaumont, C. 1981. Foreland basins. Geophysical Journal of the Royal Astronomical Society 65, 291–329.CrossRefGoogle Scholar
Beaumont, C., Fullsack, P., & Hamilton, J. 1992. Erosional control of active compressional orogens. In Thrust Tectonics (ed. McClay, K. R.), pp. 1–18. London: Chapman and Hall.CrossRefGoogle Scholar
Beaumont, C., Quinlan, G., & Hamilton, J. 1988. Orogeny and stratigraphy: numerical models of the Paleozoic in the eastern interior of North America. Tectonics 7, 389–416.CrossRefGoogle Scholar
Behrensmeyer, A. K. 1987. Miocene fluvial facies and vertebrate taphonomy in northern Pakistan. In Recent Developments in Fluvial Sedimentology (ed. Ethridge, F. G., Flores, R. M., & Harvey, M. D.), pp. 169–176. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H. -D., & Wing, S. L. 1992. Terrestrial Ecosystems Through Time. Chicago: University of Chicago Press.Google Scholar
Behrensmeyer, A. K., & Hook, R. W. 1992. Paleoenvironmental contexts and taphonomic modes. In Terrestrial Ecosystems Through Time (ed. Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A.et al.), pp. 15–136. Chicago: University of Chicago Press.Google Scholar
Bekker, A., Holland, H. D., Wang., P. -L., Rumble, D. III. et al. 2004. Dating the rise of atmospheric oxygen. Nature 427, 117–120.CrossRefGoogle ScholarPubMed
Belderson, R. H., Johnson, M. A., & Kenyon, N. H. 1982. Bedforms. In Offshore Tidal Sands (ed. Stride, A. H.), pp. 25–57. London: Chapman and Hall.CrossRefGoogle Scholar
Bengtson, S. 1999. Early Life on Earth: Nobel Symposium No. 84. New York: Columbia University Press.Google Scholar
Benito, G., Baker, V. R., & Gregory, K. J. 1998. Palaeohydrology and Environmental Change. Chichester: Wiley.Google Scholar
Benjamin, T. B. 1968. Gravity currents and related phenomena. Journal of Fluid Mechanics 31, 209–248.CrossRefGoogle Scholar
Benn, D. I. 1994. Fluted moraine and till genesis below a temperate valley glacier: Slettmarkbreen, Jotunheimen, southern Norway. Sedimentology 41, 279–292.CrossRefGoogle Scholar
Benn, D. I., & Evans, D. J. A. 1998. Glaciers and Glaciation. London: Arnold.Google Scholar
Benn, D. I., & Evans, D. J. A.2006. Subglacial megafoods: outrageous hypothesis or just outrageous? In Glacier Science and Environmental Change (ed. Knight, P. G.), pp. 42–50. Oxford: Blackwells.CrossRefGoogle Scholar
Bennett, R. H., Bryant, W. R., & Hulbert, M. H. 1991. Microstructure of Fine-Grained Sediments: From Mud to Shale. New York: Springer-Verlag.CrossRefGoogle Scholar
Bennett, S. J., & Best, J. L. 1995. Mean flow and turbulence structure over fixed, two-dimensional dunes: implications for sediment transport and dune stability. Sedimentology 42, 491–514.CrossRefGoogle Scholar
Bennett, S. J., & Best, J. L.1996. Mean flow and turbulence structure over fixed ripples and the ripple–dune transition. In Coherent Flow Structures in Open Channels (ed. Ashworth, P. J., Bennett, S. J., Best, J. L., & McLelland, S. J.), pp. 281–304. Chichester: Wiley.Google Scholar
Bennett, S. J., & Bridge, J. S. 1995. An experimental study of flow, bedload transport and bed topography under conditions of erosion and deposition and comparison with theoretical models. Sedimentology 42, 117–146.CrossRefGoogle Scholar
Bennett, S. J., Bridge, J. S., & Best, J. L. 1998. The fluid and sediment dynamics of upper-stage plane beds. Journal of Geophysical Research 103, 1239–1274.CrossRefGoogle Scholar
Bennett, S. J., & Simon, A. 2004. Riparian Vegetation and Fluvial Geomorphology. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Berelson, W. M., & Heron, S. D. 1985. Correlations between Holocene flood tidal delta and barrier island inlet sequences: Back Sound–Shackleford Banks, North Carolina. Sedimentology 32, 215–222.CrossRefGoogle Scholar
Berendsen, H. J. A., & Stouthamer, E. 2001. Paleogeographic Development of the Rhine–Meuse Delta, the Netherlands. Assen: Koninklijke Van Gorcum.Google Scholar
Berger, A., & Loutre, M.-F. 1991. Insolation values for the climate of the last 10 million years. Quaternary Science Reviews 10, 297–317.CrossRefGoogle Scholar
Bergman, K. M., & Snedden, J. W. 1999. Isolated Shallow Marine Sand Bodies: Sequence Stratigraphic Analysis and Sedimentological Interpretation. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Bernard, H. A., LeBlanc, R. J., & Major, C. F. Jr. 1962. Recent and Pleistocene geology of southwest Texas, field excursion 3. Geology of the Gulf Coast and Central Texas and Guidebook of Excursion, pp. 175–225. Houston: Houston Geological Society.Google Scholar
Berné, S., Auffret, J. P., & Walker, P. 1988. Internal structures of subtidal sandwaves revealed by high-resolution seismic reflection. Sedimentology 35, 5–20.CrossRefGoogle Scholar
Berné, S., Castaing, P., LeDrezen, E., & Lericolais, G. 1993. Morphology, internal structure and reversal of asymmetry of large subtidal dunes in the entrance to Gironde estuary (France). Journal of Sedimentary Petrology 63, 780–793.Google Scholar
Berné, S., Lericolais, G., Marsset, T., Bourillet, J., & Batist, M. 1998. Erosional offshore sand ridges and lowstand shorelines: examples from tide- and wave-dominated environments of France. Journal of Sedimentary Research 68, 540–555.CrossRefGoogle Scholar
Berner, R. A. 1980. Early Diagenesis: A Theoretical Approach. Princeton, New Jersey: Princeton University Press.Google Scholar
Berner, R. A. 1991. A model for atmospheric CO2 over Phanerozoic time. American Journal of Science 291, 339–376.CrossRefGoogle Scholar
Berner, R. A. 1994. GEOCARB II: A revised model of atmospheric CO2 over Phanerozoic time. American Journal of Science 294, 56–91.CrossRefGoogle Scholar
Berner, R. A.2001. The effect of the rise of land plants on atmospheric CO2 during the Paleozoic. In Plants Invade the Land (ed. Gensel, P. G. & Edwards, D.), pp. 173–178. New York: Columbia University Press.CrossRefGoogle Scholar
Berner, E. K., & Berner, R. A. 1996. Global Environment: Water, Air, and Geochemical Cycles. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Berner, R. A., & Holdren, G. R. Jr. 1977. Mechanisms of feldspar weathering: some observational evidence. Geology 5, 369–372.2.0.CO;2>CrossRefGoogle Scholar
Berner, R. A., & Holdren, G. R. Jr. 1979. Mechanism of feldspar weathering: II. Observations of feldspars from soils. Geochimica et Cosmochimica Acta 43, 1173–1186.CrossRefGoogle Scholar
Berner, R. A., Holdern, G. R. Jr., & Schott, J. 1985. Surface layers on dissolving silicates. Geochimica et Cosmochimica Acta 49, 1657–1658.CrossRefGoogle Scholar
Berner, R. A., & Kothavala, Z. 2001. GEOCARB III: a revised model of atmospheric CO2 over Phanerozoic time: American Journal of Science 301, 182–204.CrossRefGoogle Scholar
Best, J. L. 1992a. On the entrainment of sediment and initiation of bed defects: insights from recent developments within turbulent boundary layer research. Sedimentology 39, 797–811.CrossRefGoogle Scholar
Best, J. L. 1992b. Sedimentology and event timing of a catastrophic volcaniclastic mass flow, Volcan Hudson, Southern Chile. Bulletin of Volcanology 54, 299–318.CrossRefGoogle Scholar
Best, J. L.1993. On the interactions between turbulent flow structure, sediment transport and bedform development: some considerations from recent experimental research. In Turbulence: Perspectives on Flow and Sediment Transport (ed. Clifford, N. J., French, J. R., & Hardisty, J.), pp. 61–92. Chichester: Wiley.Google Scholar
Best, J. L.1996. The fluid dynamics of small-scale alluvial bedforms. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 67–125. Chichester: Wiley.Google Scholar
Best, J. L. 2005. The fluid dynamics of river dunes: a review and some future research directions, Journal of Geophysical Research, 110, F04S02, doi:10.1029/2004JF000218.CrossRefGoogle Scholar
Best, J. L., & Ashworth, P. J. 1997. Scour in large braided rivers and the recognition of sequence stratigraphic boundaries. Nature 387, 275–277.CrossRefGoogle Scholar
Bhattacharya, J. P., & Giosan, L. 2003. Wave-influenced deltas: geomorphological implications for facies reconstruction. Sedimentology 50, 187–210.CrossRefGoogle Scholar
Bhattacharya, J. P., & Walker, R. G. 1992. Deltas. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 157–177. Ottawa: Geological Association of Canada.Google Scholar
Birkeland, C. 1997. Life and Death of Coral Reefs. New York: Chapman and Hall.CrossRefGoogle Scholar
Birkeland, P. W. 1999. Soils and Geomorphology, 3rd edn. New York: Oxford University Press.Google Scholar
Bierkens, M. F. P., & Weerts, H. J. T. 1994. Application of indicator simulation to modeling the lithological properties of a complex confining layer. Geodema 62, 265–284.CrossRefGoogle Scholar
Bjørlykke, K. 2003. Compaction (consolidation) of sediments. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 161–168. Dordrecht: Kluwer Academic Publishers.Google Scholar
Bjørkum, P. A. 1996. How important is pressure in causing dissolution of quartz in sandstones?Journal of Sedimentary Petrology 66, 147–154.Google Scholar
Bjornsson, H. 1992. Jokulhlaups in Iceland: prediction, characteristics, and simulation. Annals of Glaciology 16, 95–106.CrossRefGoogle Scholar
Black, K. S., Paterson, D. M., & Cramp, A. 1998. Sedimentary Processes in the Intertidal Zone. London: Geological Society of London.Google Scholar
Black, K. S., Tolhurst, T. J., Paterson, D. M., & Hagerthey, S. E. 2002. Working with natural cohesive sediments. Journal of Hydraulic Engineering ASCE 128, 2–8.CrossRefGoogle Scholar
Black, M. 1933. The algal sediments of Andros Island, Bahamas. Philosophical Transactions of the Royal Society 122B, 165–191.Google Scholar
Blakey, R. C., Havholm, K. G., & Jones, L. S. 1996. Stratigraphic analysis of eolian interactions with marine and fluvial deposits, Middle Jurassic Page Sandstone and Carmel Formation, Colorado Plateau, USA. Journal of Sedimentary Research 66, 324–342.Google Scholar
Blatt, H. 1982. Sedimentary Petrology. San Francisco: W. H. Freeman.Google Scholar
Blatt, H., Middleton, G., & Murray, R. 1980. Origin of Sedimentary Rocks, 2nd edn. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Blum, M. D., & Price, D. M. 1998. Quaternary alluvial plain construction in response to glacio-eustatic and climatic controls, Texas Gulf Coastal Plain. In Relative Role of Eustasy, Climate and Tectonism on Continental Rocks (ed. Shanley, K. J. & McCabe, P. J.), pp. 31–48. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Blum, M. D., & Törnqvist, T. E. 2000. Fluvial responses to climate and sea-level change: a review and look forward. Sedimentology 47 (Supplement 1), 2–48.CrossRefGoogle Scholar
Boardman, R. S., Cheetham, A. H., & Rowell, A. J. 1987. Fossil Invertebrates. Oxford: Blackwell Scientific.Google Scholar
Boersma, J. R. 1970. Distinguishing features of wave-ripple cross-stratification and morphology. Unpublished Ph.D. thesis, Utrecht University, the Netherlands.
Boersma, J. R., & Terwindt, J. H. J. 1981. Neap–spring tide sequences of intertidal shoal deposits in a mesotidal estuary. Sedimentology 28, 151–170.CrossRefGoogle Scholar
Bohacs, K., & Suter, J. 1997. Sequence stratigraphic distribution of coaly rocks: fundamental controls and paralic examples. American Association of Petroleum Geologists Bulletin 81, 1612–1639.Google Scholar
Boles, J. R., & Franks, S. G. 1979. Clay diagenesis in Wilcox sandstones of southwest Texas: implications of smectite diagenesis on sandstone cementation. Journal of Sedimentary Petrology 49, 55–70.Google Scholar
Bond, G. C., & Lotti, R. 1995. Iceberg discharges into the North Atlantic on millennial time scales during the last glaciation. Science 267, 1005–1010.CrossRefGoogle ScholarPubMed
Bondevik, S., Mangerud, J., Dawson, S., Dawson, A., & Lohne, O. 2005. Evidence for three North Sea tsunamis at the Shetland Islands between 8000 and 1500 years ago. Quaternary Science Reviews 24, 1757–1775.CrossRefGoogle Scholar
Boothroyd, J. C. 1985. Tidal inlets and tidal deltas. In Coastal Sedimentary Environments (ed. Davis, R. A.), pp. 445–533. New York: Springer-Verlag.CrossRefGoogle Scholar
Boothroyd, J. C., Friedrich, N. E., & McGinn, S. R. 1985. Geology of microtidal coastal lagoons: Rhode Island. Marine Geology 63, 35–76.CrossRefGoogle Scholar
Borchert, H., & Muir, R. O. 1964. Salt Deposits; The Origin, Metamorphism and Deformation of Evaporites. Princeton, NJ: Van Nostrand.Google Scholar
Bosence, D. W. J. 1989. Biogenic carbonate production in Florida Bay. Bulletin of Marine Science 44, 419–433.Google Scholar
Bosence, D. W. J., & Allison, P. A. 1995. Marine Paleoenvironmental Analysis from Fossils. London: Geological Society of London.Google Scholar
Bosence, D. W. J., Rowlands, R. J., & Quine, M. L. 1985. Sedimentology and budget of a Recent carbonate mound, Florida Keys. Sedimentology 32, 3137–343.CrossRefGoogle Scholar
Bosscher, H., & Schlager, W. 1992. Computer simulation of reef growth. Sedimentology 39, 503–512.CrossRefGoogle Scholar
Bossellini, A. 1984. Progradation geometries of carbonate platforms: examples from the Triassic of the Dolomites, northern Italy. Sedimentology 31, 1–24.CrossRefGoogle Scholar
Boucot, A. J., & Gray, J., 2001. A critique of Phanerozoic climatic models involving changes in the CO2 content of the atmosphere. Earth-Science Reviews 56, 1–159.CrossRefGoogle Scholar
Boulton, G. S. 1972. Modern Arctic glaciers as depositional models for former ice sheets. Quarterly Journal of the Geological Society of London 128, 361–393.CrossRefGoogle Scholar
Boulton, G. S.1974. Processes and patterns of glacial erosion. In Glacial Geomorphology (ed. Coates, D. R.), pp. 41–87. Binghamton, NY: Publications in Geomorphology.Google Scholar
Boulton, G. S. 1976. The origin of glacially fluted surfaces: observations and theory. Journal of Glaciology 17, 287–309.CrossRefGoogle Scholar
Boulton, G. S. 1979. Processes of glacier erosion on different substrata. Journal of Glaciology 23, 15–37.CrossRefGoogle Scholar
Boulton, G. S.1987. A theory of drumlin formation by subglacial sediment deformation. In Drumlin Symposium (ed. Menzies, J. & Rose, J.), pp. 25–80. Rotterdam: Balkema.Google Scholar
Boulton, G. S.1990. Sedimentary and sea level changes during glacial cycles and their control on glacimarine facies architecture. In Glacimarine Environments (ed. Dowdeswell, J. A. & Scourse, J. D.), pp. 15–52. London: Geological Society of London.Google Scholar
Boulton, G. S. 1996a. Theory of glacial erosion, transport and deposition as a consequence of subglacial sediment deformation. Journal of Glaciology 42, 43–62.CrossRefGoogle Scholar
Boulton, G. S. 1996b. The origin of till sequences by subglacial sediment deformation beneath mid-latitude ice sheets. Annals of Glaciology 22, 75–84.CrossRefGoogle Scholar
Boulton, G. S.2006. Glaciers and their coupling with hydraulic and sedimentary processes. In Glacier Science and Environmental Change (ed. Knight, P. G.), pp. 3–22. Oxford: Blackwells.CrossRefGoogle Scholar
Boulton, G. S., & Hindmarsh, R. C. A. 1987. Sediment deformation beneath glaciers: rheology and geological consequences. Journal of Geophysical Research 92B, 9059–9082.CrossRefGoogle Scholar
Boulton, G. S., Dobbie, K. E., & Zatsepin, S. 2001. Sediment deformation beneath glaciers and its coupling to the subglacial hydraulic system. Quaternary International 86, 3–28.CrossRefGoogle Scholar
Bouma, A. H. 1962. Sedimentology of Some Flysch Deposits. Amsterdam: Elsevier.Google Scholar
Bouma, A. H., Coleman, J. R., Stetling, C. E., & Kohl, B. 1989. Influence of relative sea level on the construction of the Mississippi Fan. Geo-Marine Letters 9, 161–170.CrossRefGoogle Scholar
Bouma, A. H., Normark, W. R., & Barnes, N. E. 1985a. Submarine Fans and Related Turbidite Systems. New York: Springer-Verlag.CrossRefGoogle Scholar
Bouma, A. H., Stetling, C. E., & Coleman, J. M. 1985b. Mississippi Fan, Gulf of Mexico. In Submarine Fans and Related Turbidite Systems (ed. Bouma, A. H., Normark, W. R., & Barnes, N. E.), pp. 143–150. New York: Springer-Verlag.CrossRefGoogle Scholar
Bourgeois, J., Hansen, T. A., Wiberg, P. L., & Kauffman, E. G. 1988. A tsunami deposit at the Cretaceous–Tertiary boundary in Texas. Science 241, 567–570.CrossRefGoogle ScholarPubMed
Bowden, K. F. 1983. Physical Oceanography of Coastal Waters. Chichester: Ellis Horwood.Google Scholar
Bowen, A. J. 1969. Rip currents, 1: theoretical investigations. Journal of Geophysical Research 74, 5467–5478.CrossRefGoogle Scholar
Bowen, A. J., & Inman, D. L. 1969. Rip currents, 2: laboratory and field observations. Journal of Geophysical Research 74, 5479–5490.CrossRefGoogle Scholar
Bowen, A. J., Normark, W. R., & Piper, D. J. W. 1984. Modelling of turbidity currents of Navy Submarine Fan, California Continental Borderland. Sedimentology 31, 169–185.CrossRefGoogle Scholar
Boyce, J. I., & Eyles, N. 1991. Drumlins carved by deforming till streams below the Laurentide Ice Sheet. Geology 19, 787–790.2.3.CO;2>CrossRefGoogle Scholar
Boyd, R., Dalrymple, R. W., & Zaitlin, B. A. 1992. Classification of clastic coastal depositional environments. Sedimentary Geology 80, 139–150.CrossRefGoogle Scholar
Boyd, R., Suter, J., & Penland, S. 1989. Sequence stratigraphy of the Mississippi delta. Transactions of the Gulf Coast Association of Geological Societies 39, 331–340.Google Scholar
Bradley, R. S. 1999. Paleoclimatology. San Diego, CA: Academic Press.Google Scholar
Brady, N. C., & Weil, R. R. 2002. The Nature and Properties of Soils, 13th edn. Upper Saddle River, NJ: Prentice-Hall.Google Scholar
Braitsch, O. 1971. Salt Deposits; Their Origin and Composition. New York: Springer-Verlag.CrossRefGoogle Scholar
Branney, M., & Kokelaar, P. 2003. Pyroclastic Density Currents and the Sedimentation of Ignimbrites. London: Geological Society of London.Google Scholar
Brayshaw, A. C., Davies, G. W., & Corbett, P. W. M. 1996. Depositional controls on primary permeability and porosity at the bedform scale in fluvial reservoir sandstones. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 374–394. Chichester: Wiley.Google Scholar
Brett, C. E., & Baird, G. C. 1985. Carbonate-shale cycles in the Middle Devonian of New York: an evaluation of models for the origin of limestones in terrigenous shelf sequences. Geology 13, 324–327.2.0.CO;2>CrossRefGoogle Scholar
Brewer, R. 1976. Fabric and Mineral Analysis of Soils, 2nd edn. New York: Krieger.Google Scholar
Bricker, O. P. 1971. Carbonate Cements. Baltimore, MA: The Johns Hopkins University Press.Google Scholar
Bridge, J. S. 1978. Paleohydraulic interpretation using mathematical models of contemporary flow and sedimentation in meandering channels. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 723–742. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Bridge, J. S. 1984. Large-scale facies sequences in alluvial overbank environments. Journal of Sedimentary Petrology 54, 583–588.Google Scholar
Bridge, J. S. 1985. Paleochannel patterns inferred from alluvial deposits: a critical evaluation. Journal of Sedimentary Petrology 55, 579–589.Google Scholar
Bridge, J. S. 1992. A revised model for water flow, sediment transport, bed topography and grain size sorting in natural river bends. Water Resources Research 28, 999–1023.CrossRefGoogle Scholar
Bridge, J. S.1993a. The interaction between channel geometry, water flow, sediment transport and deposition in braided rivers. In Braided Rivers (ed. Best, J. L. & Bristow, C. S.), pp. 13–72. London: Geological Society of London.Google Scholar
Bridge, J. S. 1993b. Description and interpretation of fluvial deposits: a critical perspective. Sedimentology 40, 801–810.CrossRefGoogle Scholar
Bridge, J. S. 1997. Thickness of sets of cross strata and planar strata as a function of formative bed-wave geometry and migration, and aggradation rate. Geology 25, 971–974.2.3.CO;2>CrossRefGoogle Scholar
Bridge, J. S.1999. Alluvial architecture of the Mississippi Valley: predictions using a 3D simulation model. In Floodplains: Interdisciplinary Approaches (ed. Marriott, S. B. & Alexander, J.), pp. 269–278. London: Geological Society of London.Google Scholar
Bridge, J. S.2000. The geometry, flow and sedimentary processes of Catskill rivers and coasts. In New Perspectives on the Old Red Sandstone (ed. Friend, P. F. & Williams, B. P. J.), pp. 85–108. London: Geological Society of London.Google Scholar
Bridge, J. S. 2003. Rivers and Floodplains: Forms, Processes, and Sedimentary Record. Oxford: Blackwells.Google Scholar
Bridge, J. S. 2006. Fluvial Facies Models. In Facies Models Revisited (ed. Posamentier, H. W. and Walker, R. G.), pp. 85–170. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Bridge, J. S.2007. Numerical modelling of alluvial deposits: recent developments. In Analogue and Numerical Forward Modelling of Sedimentary Systems; from Understanding to Prediction (ed. Boer, P. L., Postima, G., Kukla, P., Zwan, K. van des, & Burgess, P.). Oxford: Blackwells.Google Scholar
Bridge, J. S., & Bennett, S. J. 1992. A model for the entrainment and transport of sediment grains of mixed sizes, shapes and densities. Water Resource Research 28, 337–363.CrossRefGoogle Scholar
Bridge, J. S., & Best, J. L. 1988. Flow, sediment transport and bedform dynamics over the transition from upper-stage plane beds: implications for the formation of planar laminae. Sedimentology 35, 753–763.CrossRefGoogle Scholar
Bridge, J. S., & Gabel, S. L. 1992. Flow and sediment dynamics in a low sinuosity, braided river: Calamus River, Nebraska Sandhills. Sedimentology, 39, 125–142.CrossRefGoogle Scholar
Bridge, J. S. & Karssenberg, D. 2005. Simulation of flow and sedimentary processes, including channel bifurcation and avulsion, on alluvial fans. 8th International Conference on Fluvial Sedimentology, Delft. The Netherlands, August 7–12, 2005.Google Scholar
Bridge, J. S., & Leeder, M. R. 1979. A simulation model of alluvial stratigraphy. Sedimentology 26, 617–644.CrossRefGoogle Scholar
Bridge, J. S., & Lunt, I. A. 2006. Depositional models of braided rivers. In Braided Rivers II (ed. Smith, G. H. Sambrook, Best, J. L., Bristow, C. S., & Petts, G.), pp. 11–50. Oxford: Blackwells.CrossRefGoogle Scholar
Bridge, J. S., & Mackey, S. D. 1993a. A theoretical study of fluvial sandstone body dimensions. In Geological Modeling of Hydrocarbon Reservoirs (ed. Flint, S. S. & Bryant, I. D.), pp. 213–236. Oxford: Blackwells.Google Scholar
Bridge, J. S., & Mackey, S. D.1993b. A revised alluvial stratigraphy model. In Alluvial Sedimentation (ed. Marzo, M. & Puidefabregas, C.), pp. 319–337. Oxford: Blackwells.CrossRefGoogle Scholar
Bridge, J. S., & Tye, R. S. 2000. Interpreting the dimensions of ancient fluvial channel bars, channels, and channel belts from wireline-logs and cores. American Association of Petroleum Geologists Bulletin 84, 1205–1228.Google Scholar
Bridge, J. S., & Willis, B. J. 1994. Marine transgressions and regressions recorded in Middle Devonian shore-zone deposits of the Catskill clastic wedge. Geological Society of America Bulletin 106, 1440–1458.2.3.CO;2>CrossRefGoogle Scholar
Bridges, P. H., & Leeder, M. R. 1976. Sedimentary model for intertidal mudflat channels with examples from the Solway Firth, Scotland. Sedimentology 23, 533–552.CrossRefGoogle Scholar
Brierley, G. J., Ferguson, R. J., & Woolfe, K. J. 1997. What is a fluvial levee?Sedimentary Geology 114, 1–9.CrossRefGoogle Scholar
Bristow, C. S., Bailey, S. D., & Lancaster, N. 2000. The sedimentary structures of linear sand dunes. Nature 406, 56–59.CrossRefGoogle ScholarPubMed
Bristow, C. S., Lancaster, N., & Duller, G. A. T. 2005. Combining ground penetrating radar surveys and optical dating to determine dune migration in Namibia. Journal of the Geological Society of London 162, 315–321.CrossRefGoogle Scholar
Bristow, C. S., Pugh, J., & Goodall, T. 1996. Internal structure of aeolian dunes in Abu Dhabi determined using ground-penetrating radar. Sedimentology 43, 995–1003.CrossRefGoogle Scholar
Bristow, C. S., Skelly, R. L., & Ethridge, F. G. 1999. Crevasse splays from the rapidly aggrading, sand-bed, braided Niobrara River, Nebraska: effect of base-level rise. Sedimentology, 46, 1029–1047.CrossRefGoogle Scholar
Brodzikowski, K., & Loon, A. J. 1991. Glacigenic Sediments. Amsterdam: Elsevier.Google Scholar
Broecker, W. A., & Denton, G. H. 1989. The role of ocean–atmosphere reorganizations in glacial cycles. Geochimica et Cosmochimica Acta 53, 2465–2501.CrossRefGoogle Scholar
Broecker, W. A., & Takahashi, T. 1966. Calcium carbonate production on the Bahama Bank. Journal of Geophysical Research 71, 1575–1602.CrossRefGoogle Scholar
Bromley, R. G. 1996. Trace Fossils: Biology, Taphonomy and Applications, 2nd edn. London: Chapman and Hall.CrossRefGoogle Scholar
Brookfield, M. E. 1977. The origin of bounding surfaces in ancient aeolian sandstones. Sedimentology 24, 303–332.CrossRefGoogle Scholar
Brookfield, M. E.1992. Eolian Systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 143–156. St. John's Newfoundland: Geological Association of Canada.Google Scholar
Brookfield, M. E., & Ahlbrandt, T. S. 1983. Eolian Sediments and Processes. Amsterdam: Elsevier.Google Scholar
Brown, T. M., & Kraus, M. J. 1987. Integration of channel floodplain suite, I. Developmental sequence and lateral relations of alluvial paleosols. Journal of Sedimentary Petrology 57, 587–601.Google Scholar
Brown, T. M., & Ratcliffe, B. C. 1988. The origin of Chubutolithes lhering, ichnofossils for the Eocene and Oligocene of Chubut Province, Argentine. Journal of Paleontology 62, 163–167.Google Scholar
Brunsden, D., & Prior, D. B. 1984. Slope Instability. New York: Wiley.Google Scholar
Bryant, E. 2001. Tsunami: The Underrated Hazard. Cambridge: Cambridge University Press.Google Scholar
Bryant, I. D., & Flint, S. 1993. Quantitative clastic reservoir geological modeling: problems and perspectives. In Geological Modeling of Hydrocarbon Reservoirs (ed. Flint, S. & Bryant, I. D.), pp. 3–20. Oxford: Blackwells.Google Scholar
Buatois, L. A., Mangano, M. G., Genise, J. F., & Taylor, T. N. 1998. The ichnologic record of the continental invertebrate invasion: evolutionary trends in environmental expansion, ecospace utilization, and behavioral complexity. Palaios 13, 217–240.CrossRefGoogle Scholar
Bull, W. B. 1991. Geomorphic Responses to Climate Change. New York: Oxford University Press.Google Scholar
Bullock, P., Fedoroff, N., Jungerius, A., Stoops, G., & Tursina, T. 1988. Handbook of Soil Thin Section Description. Albrighton: Waine Research.Google Scholar
Burbank, D. W. 1992. Causes of recent Himalayan uplift deduced from deposited patterns in the Ganges basin. Nature 357, 680–683.CrossRefGoogle Scholar
Burbank, D. W., & Anderson, R. S. 2001. Tectonic Geomorphology. Oxford: Blackwells.Google Scholar
Burley, S. D. 2003. Cathodoluminescence (applied to the study of sedimentary rocks). In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 102–106. Dordrecht: Kluwer Academic Publishers.Google Scholar
Burley, S. D., Kantorowicz, J. D., & Waugh, B. 1985. Clastic diagenesis. In Sedimentology: Recent Developments and Applied Aspects (ed. Brenchley, P. J. & Williams, B. P. J.), pp. 189–228. London, Blackwell Scientific Publications.Google Scholar
Burley, S. D., & Worden, R. H. 2003. Sandstone Diagenesis: Recent and Ancient. Malden, MA: Blackwells.CrossRefGoogle Scholar
Burnett, W., & Riggs, S. R. 1990. Phosphate Deposits of the World, Volume 3, Neogene to Modern Phosphorites. Cambridge: Cambridge University Press.Google Scholar
Burst, J. F. 1965. Subaqueously formed shrinkage cracks in clay. Journal of Sedimentary Petrology 35, 348–353.CrossRefGoogle Scholar
Busby, C., & Ingersoll, R. V. 1995. Tectonics of Sedimentary Basins. Oxford: Blackwells.Google Scholar
Bustillo, M. A. 2003. Silcrete. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 659–660. Dordrecht: Kluwer Academic Publishers.Google Scholar
Butler, G. P. 1969. Modern evaporite deposition and geochemistry of coexisting brines, the sabkha, Trucial Coast, Arabian Gulf. Journal of Sedimentary Petrology 39, 70–89.Google Scholar
Campbell, C. S. 1990. Rapid granular flows. Annual Review of Fluid Mechanics 22, 57–92.CrossRefGoogle Scholar
Campbell, C. S. 2001. Granular flows in the elastic limit. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 83–89. Oxford: Blackwells.CrossRefGoogle Scholar
Campbell, N. A., Reece, J. B., & Mitchell, L. G. 1999. Biology, 5th edn. Park, Menlo, CA: Addison Wesley Longman.Google Scholar
Cao, S., & Knight, D. W. 1998. Design of hydraulic geometry of alluvial channels. Journal of Hydraulic Engineering, ASCE 124, 484–492.CrossRefGoogle Scholar
Carballo, J. D., Land, L. S., & Miser, D. E. 1987. Holocene dolomitization of supratidal sediments by active tidal pumping, Sugarloaf Key, Florida. Journal of Sedimentary Petrology 57, 153–165.Google Scholar
Carle, S. F., Labolle, E. M., Weissmann, G. S. et al. 1998. Conditional simulation of hydrofacies architecture: a transition probability/Markov approach. In Hydrogeologic Models of Sedimentary Aquifers (ed. Fraser, G. S. & Davis, J. M.), pp. 147–170. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Carling, P. A. 1981. Sediment transport by tidal currents and waves: observations from a sandy intertidal zone (Burry Inlet, South Wales). In Holocene Marine Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 65–80. Oxford: Blackwells.CrossRefGoogle Scholar
Carling, P. A. 1999. Subaqueous gravel dunes. Journal of Sedimentary Research 69, 534–545.CrossRefGoogle Scholar
Carling, P. A., & Dawson, M. R. 1996. Advances in Fluvial Dynamics and Stratigraphy. Chichester: Wiley.Google Scholar
Carling, P. A. & Shvidchenko, A. B. 2002. A consideration of the dune : antidune transition in fine gravel. Sedimentology, 49, 1,269–1,282.CrossRefGoogle Scholar
Carpenter, A. B. 1978. Origin and chemical evolution of brines in sedimentary basins. In Thirteenth Annual Forum on the Geology of Industrial Minerals (ed. Johnson, K. S. & Russell, J. A.), pp. 60–77. Tulsa, OK: Oklahoma Geological Survey.Google Scholar
Carson, M. A., & Kirkby, M. J. 1972. Hillslope Form and Process. Cambridge: Cambridge University Press.Google Scholar
Carter, D. J. T. 1982. Prediction of wave height and period for a constant wind velocity using the JONSWAP results. Ocean Engineering 9, 17–33.CrossRefGoogle Scholar
Carter, L., & McCave, I. N. 1997. The sedimentary regime beneath the deep Western Boundary Current inflow to the southwest Pacific Ocean. Journal of Sedimentary Research 67, 1005–1017.Google Scholar
Carter, R. M. 1975. A discussion and classification of subaqueous mass-transport with particular application to grain flow, slurry-flow, and fluxoturbidites. Earth-Science Reviews 11, 145–177.CrossRefGoogle Scholar
Carter, R. W. G., & Woodroffe, C. D. 1994. Coastal Evolution: Late Quaternary Shoreline Morphodynamics. Cambridge: Cambridge University Press.Google Scholar
Cas, R. A. F., & Wright, J. V. 1987. Volcanic Successions: Modern and Ancient. London: Allen & Unwin.CrossRefGoogle Scholar
Cathles, L. M. 1993. A discussion of flow mechanisms responsible for alteration and mineralization in the Cambrian aquifers of the Ouachita–Arkoma Basin–Ozark system. In Diagenesis and Basin Development (ed. Horbury, A. D. & Robinson, A. G.), pp. 99–112. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Cayeux, L. 1935. Les roches sédimentaires de France: roches carbonatées. Paris: Masson. [English translation by Carozzi, A. V. 1970. Sedimentary Rocks of France: Carbonate Rocks. Darien, CN: Hafner Publishing.]Google Scholar
Cerling, T. E. 1999. Stable carbon isotopes in paleosol carbonates. In Palaeoweathering, Palaeosurfaces and Related Continental Deposits (ed. Thiry, M. & Simon-Coicon, R.), pp. 43–60. Oxford: Blackwells.Google Scholar
Cerling, T. E., & Craig, H. 1994. Geomorphology and in-situ cosmogenic isotopes. Annual Review of Earth and Planetary Sciences 22, 273–317.CrossRefGoogle Scholar
Chafetz, H. S. 1986. Marine pisoids: a product of bacterially induced precipitation. Journal of Sedimentary Petrology 56, 812–817.Google Scholar
Chafetz, H. S., & Folk, R. L. 1984. Travertines, depositional morphology and the bacterially constructed constituents. Journal of Sedimentary Petrology 54, 289–316.Google Scholar
Chang, H. H. 1988. Fluvial Processes in River Engineering. New York: Wiley.Google Scholar
Chave, K. E. 1954. Aspects of the biogeochemistry of magnesium 1. Calcareous marine organisms. Journal of Geology 62, 266–283.CrossRefGoogle Scholar
Chave, K. E., Smith, S. V., & Roy, R. J. 1972. Carbonate production by coral reefs. Marine Geology 12, 123–140.CrossRefGoogle Scholar
Cheel, R. J., & Leckie, D. A. 1992. Coarse-grained storm beds of the Upper cretaceous Chungo Member (Wapiabi Formation), southern Alberta, Canada. Journal of Sedimentary Petrology 62, 933–945.Google Scholar
Chester, R. 2003. Marine Geochemistry, 2nd edn. Oxford: Blackwells.Google Scholar
Chilingarian, G. V., & Wolf, K. H. 1988. Diagenesis. Amsterdam: Elsevier.Google Scholar
Choquette, P. W., & Pray, L. C. 1970. Geological nomenclature and classification of porosity in sedimentary carbonates. American Association of Petroleum Geologists Bulletin 54, 207–250.Google Scholar
Christie-Blick, N., Sohl, L. E., & Kennedy, M. J. 1999. Considering a Neoproterozoic snowball Earth. Science 284, 1087.CrossRefGoogle Scholar
Chuhan, F. A., Bjørlykke, K., & Lowrey, C. J. 2000. The role of provenance in illitization of deeply buried reservoir sandstones from Haltenbanken and north Viking Graben, offshore Norway. Marine and Petroleum Geology 17, 673–689.CrossRefGoogle Scholar
Chuhan, F. A., Bjørlykke, K., & Lowrey, C. J. 2001. Closed-system burial diagenesis in reservoir sandstones: examples from the Garn Formation at Haltenbanken Area, offshore mid-Norway. Journal of Sedimentary Petrology 71, 15–26.CrossRefGoogle Scholar
Church, M. 1996. Channel morphology and typology. In River Flows and Channel Forms (ed. Petts, G. & Calow, P.), pp. 185–202. Oxford: Blackwell Science Limited.Google Scholar
Clark, C. D., & Meehan, R. T. 2001. Subglacial bedform morphology of the Irish Ice Sheet reveals major configuration changes during growth and decay. Journal of Quaternary Science 16, 483–496.CrossRefGoogle Scholar
Clark, J. D., & Pickering, K. T. 1996. Submarine Channels: Processes and Architecture. London: Vallis Press.Google Scholar
Clark, J. D., Kenyon, N. H., & Pickering, K. T. 1992. Quantitative analysis of the geometry of submarine channels: implications for the classification of submarine fans. Geology 20, 633–636.2.3.CO;2>CrossRefGoogle Scholar
Clarke, G. K. C., Collins, S. G., & Thompson, D. E. 1984. Flow, thermal structure, and subglacial conditions of a surge-type glacier. Canadian Journal of Earth Sciences 21, 232–240.CrossRefGoogle Scholar
Clarkson, E. N. K. 1998. Invertebrate Palaeontology and Evolution, 4th edn. Oxford: Blackwell Scientific.Google Scholar
Clemmensen, L. B. 1989. Preservation of interdraa and draa plinth deposits by the lateral migration of large linear draas (Lower Permian Yellow Sands, northeast England). Sedimentary Geology 65, 139–151.CrossRefGoogle Scholar
Clemmensen, L. B., Olsen, H., & Blakey, R. C. 1989. Erg-margin deposits in the Lower Jurassic Moenave Formation and Wingate Sandstone, southern Utah. Geological Society of America Bulletin 101, 759–773.2.3.CO;2>CrossRefGoogle Scholar
Clevis, Q., Boer, P., & Nijman, W. 2004a. Differentiating the effect of episodic tectonism and eustatic sea-level fluctuations in foreland basins filled by alluvial fans and axial deltaic systems: insights from a three-dimensional stratigraphic forward model. Sedimentology 51, 809–835.CrossRefGoogle Scholar
Clevis, Q., Boer, P., & Wachter, M. 2003. Numerical modeling of drainage basin evolution and three-dimensional alluvial fan stratigraphy. Sedimentary Geology 163, 85–110.CrossRefGoogle Scholar
Clevis, Q., Jager, G., Nijman, W., & Boer, P. L. 2004b. Stratigraphic signatures of translation of thrust-sheet top basins over low-angle detachment faults. Basin Research 16, 145–163.CrossRefGoogle Scholar
Clifton, H. E. 1969. Beach lamination: nature and origin. Marine Geology 7, 553–559.CrossRefGoogle Scholar
Clifton, H. E., & Dingler, J. R. 1984. Wave-formed structures and palaeoenvironmental reconstruction. Marine Geology 60, 165–198.CrossRefGoogle Scholar
Clifton, H. E., Hunter, R. E., & Phillips, R. L. 1971. Depositional structures and processes in the non-barred, high energy nearshore. Journal of Sedimentary Petrology 41, 651–670.Google Scholar
Clout, J. M. F., & Simonson, B. M. 2005. Precambrian iron formations and iron formation-hosted iron ore deposits. Economic Geology 100th Anniversary Issue, 643–679.Google Scholar
Cohen, A. S., 2003. Paleolimnology: The History and Evolution of Lake Systems. Oxford: Oxford University Press.Google Scholar
Cole, J. M., Rasbury, E. T., Montañez, I. P.et al. 2005. Petrographic and trace element analysis of uranium-rich tufa calcite, middle Miocene Barstow Formation, California, USA. Sedimentology 51, 433–453.CrossRefGoogle Scholar
Coles, D. F. 1956. The law of the wake in turbulent boundary layer. Journal of Fluid Mechanics 1, 191–226.CrossRefGoogle Scholar
Colella, A., & Prior, D. B. 1990. Coarse-Grained Deltas. Oxford: Blackwells.CrossRefGoogle Scholar
Coleman, J. M. 1969. Brahmaputra River: channel processes and sedimentation. Sedimentary Geology 3, 129–239.CrossRefGoogle Scholar
Coleman, J. M. 1988. Dynamic changes and processes in the Mississippi River delta. Geological Society of America Bulletin 100, 999–1015.2.3.CO;2>CrossRefGoogle Scholar
Coleman, J. M., & Gagliano, S. M. 1964. Cyclic sedimentation in the Mississippi River deltaic plain. Transactions of the Gulf Coast Association of Geological Societies 14, 67–80.Google Scholar
Coleman, J. M., & Gagliano, S. M.1965. Sedimentary structures: Mississippi River deltaic plain. In Primary Sedimentary Structures and Their Hydrodynamic Interpretation (ed. Middleton, G. V.), pp. 133–148. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Coleman, J. M., Gagliano, S. M., & Webb, J. E. 1964. Minor sedimentary structures in a prograding distributary. Marine Geology 1, 240–258.CrossRefGoogle Scholar
Coleman, J. M., & Prior, D. B. 1982. Deltaic environments. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D. R.), pp. 139–178. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Coleman, J. M., Prior, D. B., & Lindsay, J. F. 1983. Deltaic influences on shelfedge instability processes. In The Shelfbreak: Critical Interface on Continental Margins (ed. Stanley, D. J. & Moore, G. T.), pp. 121–137. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Coleman, J. M., & Wright, L. D. 1975. Modern river deltas: variability of processes and sand bodies. In Deltas, Models for Exploration (ed. Broussard, M. L.), pp. 99–149. Houston, TX: Houston Geological Society.Google Scholar
Collins, M. B., Amos, C. L., & Evans, G. 1981. Observations of some sediment transport processes over intertidal flats, The Wash, UK. In Holocene Marine Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 81–98. Oxford: Blackwells.CrossRefGoogle Scholar
Coniglio, M., & Dix, G. R. 1992. Carbonate slopes. In Facies Models, Response to Sea Level Changes (ed. Walker, R. G. & James, N. P.), pp. 349–373. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Cook, H. E., Hine, A. C., & Mullins, H. T. 1983. Platform Margin and Deep Water Carbonates. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Cook, H. E., & Mullins, H. T. 1983. Basin margin environments. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 540–617. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Cook, P. J., & Shergold, J. H. 1986. Phosphate Deposits of the World, Volume 1, Proterozoic and Cambrian Phosphorites. Cambridge: Cambridge University Press.Google Scholar
Cooke, R. U., & Warren, A. 1973. Geomorphology in Deserts. London: Batsford.Google Scholar
Cooke, R. U., Warren, A., & Goudie, A. 1993. Desert Geomorphology. London: UCL Press.Google Scholar
Corney, R. K. T., Peakall, J., Parsons, D. R.et al. 2006. The orientation of helical flow in curved channels. Sedimentology 53, 249–257.CrossRefGoogle Scholar
Costa, J. E., & Wieczorek, G. F. 1987. Debris Flows/Avalanches: Process, Recognition, and Mitigation. Boulder, CO: Geological Society of America.Google Scholar
Costa, J. E., & Williams, G. P. 1984. Debris Flow Dynamics. United States Geological Survey Open File Report 84–606 (VHS Videotape). Reston, VA: United States Geological Survey.Google Scholar
Cotter, E., & Link, J. E. 1993. Deposition and diagenesis of Clinton ironstones (Silurian) in the Appalachian Foreland basin of Pennsylvania. Geological Society of America Bulletin 105, 911–922.2.3.CO;2>CrossRefGoogle Scholar
Coulthard, T. J., Macklin, M. G., & Kirkby, M. J. 2002. A cellular model of Holocene upland river basin and alluvial fan evolution. Earth Surface Processes and Landforms 27, 269–288.CrossRefGoogle Scholar
Crabaugh, M., & Kocurek, G. 1993. Entrada sandstone: an example of a wet aeolian system. In The Dynamics and Environmental Context of Aeolian Sedimentary Systems (ed. Pye, K.), pp. 103–126. London: Geological Society of London.Google Scholar
Craft, J. H., & Bridge, J. S. 1987. Shallow-marine sedimentary processes in the Late Devonian Catskill Sea, New York State. Geological Society of America Bulletin 98, 338–355.2.0.CO;2>CrossRefGoogle Scholar
Crawford, D. A., & Mader, C. L. 1998. Modeling asteroid impact and tsunami. Science of Tsunami Hazards 16, 21–30.Google Scholar
Crevello, P. D., & Schlager, W. 1980. Carbonate debris sheets and turbidites, Exuma Sound, Bahamas. Journal of Sedimentary Petrology 50, 1121–1148.Google Scholar
Crick, R. E. 1986. Origin, Evolution, and Modern Aspects of Biomineralization in Plants and Animals. New York: Plenum Press.Google Scholar
Cross, T. A. 1990. Quantitative Dynamic Stratigraphy. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Crowley, T. J., & Berner, R. A. 2001. CO2 and climate change. Science 289, 270–277.CrossRefGoogle Scholar
Crowley, T. J., & North, G. R. 1991. Paleoclimatology. New York: Oxford University Press.Google Scholar
Csanady, G. T. 1978. Water circulation and dispersal mechanisms. In Lakes: Chemistry Geology and Physics (ed. Lerman, A.), pp. 21–64. New York: Springer-Verlag.CrossRefGoogle Scholar
Curray, J. R., Emmel, F. J., & Crampton, P. J. S. 1969. Holocene history of a strand plain, lagoonal coast, Narayit, Mexico. In Coastal Lagoons – A Symposium (ed. Castanares, A. A. & Phleger, F. B.), pp. 63–100. Mexico City: Universidad Nacional Autónoma.Google Scholar
Dade, W. B., & Huppert, H. E. 1994. Predicting the geometry of channelized deep-sea turbidites. Geology 22, 645–648.2.3.CO;2>CrossRefGoogle Scholar
Dade, W. B., Nowell, A. R. M., & Jumars, P. A. 1992. Predicting the erosion resistance of muds. Marine Geology 105, 285–297.CrossRefGoogle Scholar
Dalrymple, R. W. 1984. Morphology and internal structure of sand waves in the Bay of Fundy. Sedimentology 31, 365–382.CrossRefGoogle Scholar
Dalrymple, R. W.1992. Tidal depositional systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 195–218. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Dalrymple, R. W., Baker, E. K., Hughes, M., & Harris, P. T. 2003. Geomorphology and sedimentology of the muddy, tide-domonated, Fly River delta, Papua New Guinea. In Tropical Deltas of South-East Asia – Sedimentology, Stratigraphy, and Petroleum Geology (ed. Sidi, F. H., Nummedal, D., Imbert, P., Darman, H., & Posamentier, H.), pp. 147–174. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Dalrymple, R. W., Boyd, R., & Zaitlin, B. A. 1994. Incised Valley Systems: Origin and Sedimentary Sequences. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Dalrymple, R. W., Knight, R. J., Zaitlin, B. A., & Middleton, G. V. 1990. Dynamics and facies model of a macrotidal sand-bar complex, Cobequid Bay–Salmon River Estuary (Bay of Fundy). Sedimentology 37, 577–612.CrossRefGoogle Scholar
Dalrymple, R. W., Makino, Y., & Zaitlin, B. A. 1991. Temporal and spatial patterns of rhythmite deposition on mudflats in the macrotidal Cobequid Bay–Salmon River Estuary, Bay of Fundy. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 137–160. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Dalrymple, R. W., & Rhodes, R. N. 1995. Estuarine dunes and barforms. In Geomorphology and Sedimentology of Estuaries (ed. Perillo, G. M.), pp. 359–422. Amsterdam: Elsevier.Google Scholar
Dalrymple, R. W., Zaitlin, B. A., & Boyd, R. 1992. Estuarine facies models: Conceptual basis and stratigraphic implications. Journal of Sedimentary Petrology 62, 1130–1146.CrossRefGoogle Scholar
Damuth, J. E., & Embley, R. W. 1981. Mass-transport processes on Amazon cone: western equatorial Atlantic. American Association of Petroleum Geologists Bulletin 65, 629–643.Google Scholar
Damuth, J. E., Flood, R. D., Kowsmann, R. O., Belderson, R. H., & Gorini, M. A. 1988. Anatomy and growth pattern of Amazon deep-sea fans as revealed by long-range side-scan sonar (GLORIA) and high-resolution seismic studies. American Association of Petroleum Geologists Bulletin 72, 885–911.Google Scholar
Damuth, J. E., Flood, R. D., Pirmez, C., & Manley, P. L. 1995. Architectural elements and depositional processes of Amazon deep-sea fan imaged by sidescal sonar (GLORIA), bathymetric swath mapping (SeaBeam), high-resolution seismic, and piston-core data. In Atlas of Deep Water Environments: Architectural Style in Turbidite Systems (ed. Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, F. Ricci, & Smith, R. D. A.), pp. 105–121. London: Chapman and Hall.CrossRefGoogle Scholar
Damuth, J. E., Jacobi, R. D., & Hayes, D. E. 1983a. Sedimentary processes in Northwest Pacific Basin revealed by echo-character mapping studies. Geological Society of America Bulletin 94, 381–395.2.0.CO;2>CrossRefGoogle Scholar
Damuth, J. E., Kolla, V., Flood, R. D.et al. 1983b. Distributary channel meandering and bifurcation patterns on the Amazon deep-sea fan as revealed by long-range side-scan sonar (GLORIA). Geology 11, 94–98.2.0.CO;2>CrossRefGoogle Scholar
Damuth, J. E., Kowsmann, R. O., Flood, R. D., Belderson, R. H., & Gorini, M. A. 1983c. Age relationships of distributary channels on Amazon Deep-Sea fan: implications for fan growth pattern. Geology 11, 470–473.2.0.CO;2>CrossRefGoogle Scholar
Davidson, J. F., Harrison, D., & Carvalho, Guedes J. R. F. 1977. On the liquidlike behaviour of fluidized beds. Annual Review of Fluid Mechanics 9, 55–86.CrossRefGoogle Scholar
Davidson-Arnott, R. G. D., & Greenwood, B. 1974. Bedforms and structures associated with bar topography in the shallow water wave environment, Kouchibouguac Bay, New Brunswick, Canada. Journal of Sedimentary Petrology 44, 698–704.Google Scholar
Davies, G. R. 1970. Algal-laminated sediments, Gladstone Embayment, Shark Bay, Western Australia. In Carbonate Sedimentation and Environments, Shark Bay, Western Australia (ed. Logan, B. W., Davies, G. R., Read, J. F., & Cebulski, D. E.), pp. 169–205. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Davies, P. J., & Till, R. 1968. Stained dry cellulose peels of ancient and recent impregnated carbonate sediments. Journal of Sedimentary Petrology 38, 234–237.CrossRefGoogle Scholar
Davies, S., Hampson, G., Flint, S., & Elliott, T. 1999. Continent-scale sequence stratigraphy of the Namurian, Upper Carboniferous and its applications to reservoir prediction. In Petroleum Geology of Northwest Europe: Proceedings 5th Conference (ed. Fleet, A. J. & Boldy, A. A. R.), pp. 757–770. London: The Geological Society.Google Scholar
Davis, R. A. 1985. Coastal Sedimentary Environments. New York: Springer-Verlag.CrossRefGoogle Scholar
Davis, R. A., & Clifton, H. E. 1987. Sea-level change and the preservation of wave-dominated and tide-dominated coastal sequences. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 167–178. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Davis, R. A., & Fitzgerald, D. M. 2004. Beaches and Coasts. Oxford: Blackwells.Google Scholar
Davis, R. A., & Flemming, B. W. 1995. Stratigraphy of a combined wave- and tide-dominated intertidal sand body: Martens Plate, East Frisian Wadden Sea, Germany. In Tidal Signatures in Modern and Ancient Sediments (ed. Flemming, B. W. & Bartholoma, A.), pp. 121–132. Oxford: Blackwells Publication.CrossRefGoogle Scholar
Dawson, A. G. 1992. Ice Age Earth: Late Quaternary Geology and Climate. London: Routledge.Google Scholar
Dawson, A. G. 1994. Geomorphological effects of tsunami run-up and backwash. Geomorphology 10, 83–94.CrossRefGoogle Scholar
Dawson, A. G. 1999. Linking tsunami deposits, submarine slides and offshore earthquakes. Quaternary International 60, 119–126.CrossRefGoogle Scholar
Dawson, A. G., Foster, I. K. L., Shi, S., Smith, D. E., & Long, D. 1991. The identification of tsunami deposits in coastal sediment sequences. Science of Tsunami Hazards 9, 73–82.Google Scholar
Dean, W. E., & Fouch, T. D. 1983. Lacustrine. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 97–130. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Boer, P. L. 1979. Convolute lamination in modern sands of the estuary of the Oosterschelde, the Netherlands, formed as the result of entrapped air. Sedimentology 26, 283–294.CrossRefGoogle Scholar
Boer, P. L., & Smith, D. G. 1994. Orbital Forcing and Cyclic Sequences. Oxford: Blackwells.CrossRefGoogle Scholar
Boer, P. L., Gelder, A., & Nio, S. D. 1988. Tide-influenced Sedimentary Environments and Facies. Boston: D. Reidel Publishing Company.CrossRefGoogle Scholar
Decho, A. W., Visscher, P. T., & Reid, R. P. 2005. Production and cycling of natural microbial exopolymers (EPS) within a marine stromatolite. Palaeogeography, Palaeoclimatology, Palaeoecology, 219 71–86.CrossRefGoogle Scholar
Deer, W. A., Howie, R. A., & Zussman, J. 1962. Rock Forming Minerals, Volume 3: Sheet Silicates. New York: Wiley.Google Scholar
Defant, A. 1958. Ebb and Flow. Ann Arbor, MI: University of Michigan Press.Google Scholar
Deffeyes, K. S., Lucia, F. J., & Weyl, P. K. 1965. Dolomitization of Recent and Plio-Pleistocene sediments by marine evaporite waters on Bonaire, Netherlands Antilles. In Dolomitization and Limestone Diagenesis, A Symposium (ed. Pray, L. C. & Murray, R. C.), pp. 71–80. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Degens, E. T., & Ross, D. A. 1969. Hot Brines and Recent Heavy Metal Deposits in the Red Sea: A Geochemical and Geophysical Account. New York: Springer-Verlag.CrossRefGoogle Scholar
Demarest, J. M. II, & Kraft, J. C. 1987. Stratigraphic record of Quaternary sea levels: implications for more ancient strata. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 223–239. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Demicco, R. V. 1983. Wavy and lenticular-bedded carbonate ribbon rocks of the Upper Cambrian Conococheague Limestone, western Maryland. Journal of Sedimentary Petrology 53, 1121–1132.Google Scholar
Demicco, R. V. 1998. CYCOPATH 2D – a two-dimensional model of cyclic sedimentation on carbonate platforms. Computers & Geosciences 24, 405–423.CrossRefGoogle Scholar
Demicco, R. V., & Gierlowski-Kordesch, E. 1986. Facies sequences of a semi-arid closed basin: the Lower Jurassic East Berlin Formation of the Hartford Basin, New England, U.S.A. Sedimentology, 33, 107–118.CrossRefGoogle Scholar
Demicco, R. V., & Hardie, L. A. 1994. Sedimentary Structures and Early Diagentic Features of Shallow Marine Carbonate Deposits. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Demicco, R. V., & Hardie, L. A. 2002. The “carbonate factory” revisited: a reexamination of sediment production functions used to model deposition on carbonate platforms. Journal of Sedimentary Research 72, 849–857.CrossRefGoogle Scholar
Demicco, R. V., Lowenstein, T. K., Hardie, L. A., & Spencer, R. J. 2005. Model of seawater composition for the Phanerozoic. Geology 33, 877–880.CrossRefGoogle Scholar
Demicco, R. V., & Mitchell, R. W. III 1982. Facies of the Great American Bank in the central Appalachians. In Central Appalachian Geology, Field Trip Guidebook (ed. Lyttle, P. T.), pp. 171–266. Falls Church, VA: American Geological Institute.Google Scholar
Mowbray, T. 1983. The genesis of lateral accretion deposits in recent intertidal mudflat channels, Solway Firth, Scotland. Sedimentology 30, 425–435.CrossRefGoogle Scholar
Mowbray, T., & Visser, M. J. 1984. Reactivation surfaces in subtidal channel deposits, Oosterschelde, southwest Netherlands. Journal of Sedimentary Petrology 54, 811–824.Google Scholar
Raaf, J. F. M., Boersma, J. R., & Gelder, A. 1977. Wave-generated structures and sequences from a shallow marine succession, Lower Carboniferous, County Cork, Ireland. Sedimentology 24, 451–483.CrossRefGoogle Scholar
Deutsch, C. V. 2002. Geostatistical Reservoir Modeling. New York: Oxford University Press.Google Scholar
Deutsch, C., & Cockerham, P. 1994. Practical considerations in the application of simulated annealing to stochastic simulation. Mathematical Geology 26, 67–82.CrossRefGoogle Scholar
Deutsch, C. V., & Wang, L. 1996. Hierarchical object-based stochastic modeling of fluvial reservoirs. Mathematical Geology 28, 857–880.CrossRefGoogle Scholar
D'Hondt, S., Rutherford, S., & Spivak, A. J. 2002. Metabolic activity of subsurface life in deep-sea sediments. Science 295, 2067–2070.CrossRefGoogle ScholarPubMed
D'Hondt, S., Jørgensen, B. B., Miller, D. J.et al. 2004. Distributions of microbial activities in deep subseafloor sediments. Science 306, 2216–2221.CrossRefGoogle ScholarPubMed
Dickinson, W. R. 1976. Plate Tectonic Evolution of Sedimentary Basins. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Dickinson, W. R.1985. Interpreting provenance relations from detrital modes of sandstones. In Provenance of Arenites (ed. Zuffa, G. G.), pp. 333–361. Dordrecht: Reidel.CrossRefGoogle Scholar
Dickinson, W. R. 1995. Forearc basins. In Tectonics of Sedimentary Basins (ed. Busby, C. & Ingersoll, R.), pp. 221–262. Oxford: Blackwell Science.Google Scholar
Dickson, J. A. D. 1966. Carbonate identification and genesis as revealed by staining. Journal of Sedimentary Petrology 36, 491–505.Google Scholar
Dickson, J. A. D. 2002. Fossil echinoderms as monitor of the Mg/Ca ratio of Phanerozoic oceans. Science 298, 1222–1224.CrossRefGoogle ScholarPubMed
Diem, B. 1984. Analytical method for estimating paleowave climate and water depth from wave ripple marks. Sedimentology 32, 705–720.CrossRefGoogle Scholar
Dietrich, W. E., & Dunne, T. 1993. The channel head. In Channel Network Hydrology (ed. Beven, K. & Kirkby, M. J.), pp. 175–219. Chichester: Wiley.Google Scholar
Dill, R. F., Kendall, C. G.St., C., & Shinn, E. A. 1989. Giant Subtidal Stromatolites and Related Sedimentary Features, Lee Stocking Island, Exumas, Bahamas. IGC Field TripT 373, Washington, D.C.: American Geophysical Union.Google Scholar
Dill, R. F., Shinn, E. A., Jones, A. T., Kelly, K., & Steinen, R. P. 1986. Giant subtidal stromatolites forming in normal salinity water. Nature 324, 55–58.CrossRefGoogle Scholar
Dingler, J. R. 1979. The threshold of grain motion under oscillatory flow in a laboratory wave channel. Journal of Sedimentary Petrology 49, 287–294.Google Scholar
Domack, E. W., & Ishman, S. 1993. Oceanographic and physiographic controls on modern sedimentation within Antarctic fjords. Geological Society of America Bulletin 105, 1175–1189.2.3.CO;2>CrossRefGoogle Scholar
Dominguez, J. M. L. 1996. The São Francisco strandplain: a paradigm for wave-dominated deltas? In Geology of Siliciclastic Shelf Seas (ed. Baptist, M. & Jacobs, P.), pp. 217–231. London: Geological Society of London.Google Scholar
Dominguez, J. M. L., Martin, L., & Bittencourt, A. C. S. P. 1987. Sea-level history and Quaternary evolution of river mouth-associated beach-ridge plains along the east-southeast Brazilian coast: a summary. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 115–127. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRef
Donovan, R. N., & Foster, R. J. 1972. Subaqueous shrinkage cracks from the Caithness Flagstone series (Middle Devonian) of NE Scotland. Journal of Sedimentary Petrology 42, 309–317.Google Scholar
Donovan, S. K. 1994. Insects and other arthropods as trace-makers in non-marine environments and paleoenvironments. In The Paleobiology of Trace Fossils (ed. Donovan, S. K.), pp. 200–220. Baltimore, MA: The Johns Hopkins University Press.Google Scholar
Dott, R. H. Jr. 1992. Eustasy: the ups and downs of a major geological concept. Boulder, CO: Geological Society of America.
Dott, R. H., & Bourgeois, J. 1982. Hummocky stratification: significance of its variable bedding sequences. Geological Society of America Bulletin 93, 663–680.2.0.CO;2>CrossRefGoogle Scholar
Dove, P. M., DeYoreo, J. J., & Weiner, S. 2003. Biomineralization. Chantilly, VA: The Mineralogical Society of America.Google Scholar
Doveton, J. H. 1994. Theory and applications of vertical variability measures from Markov chain analysis. In Stochastic Modeling and Geostatistics (ed. Yarus, J. M. & Chambers, R. L.), pp. 55–64. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Dowdeswell, J. A., Elverhoi, A., Andrews, J. T., & Hebbeln, D. 1999. Asynchronous deposition of ice-rafted layers in the Nordic seas and North Atlantic Ocean. Nature 400, 348–351.CrossRefGoogle Scholar
Dowdeswell, J. A., Maslin, M. A., Andrews, J. T., & McCave, I. N. 1995. Iceberg production, debris rafting, and the extent and thickness of Heinrich layers (H-1, H-2) in North Atlantic sediments. Geology 23, 301–304.2.3.CO;2>CrossRefGoogle Scholar
Dowdeswell, J. A., & Cofaigh, O C. 2002. Glacier-influenced Sedimentation on High-latitude Continental Margins. London: Geological Society of London.Google Scholar
Dowdeswell, J. A., O Cofaigh, C., Taylor, J. et al. 2002. On the architecture of high-latitude continental margins: the influence of ice-sheet and sea-ice processes in the Polar North Atlantic. In Glacier-influenced Sedimentation on High-latitude Continental Margins (ed. Dowdeswell, J. A. & Cofaigh, C. O), pp. 33–54. London: Geological Society of London.Google Scholar
Dowdeswell, J. A., & Scourse, J. D. 1990. Glacimarine Environments. London: Geological Society of London.Google Scholar
Dowdeswell, J. A., Whittington, R. J., & Marienfeld, P. 1994. The origin of massive diamicton facies by iceberg rafting and scouring, Scoresby Sund, East Greenland. Sedimentology 41, 21–35.CrossRefGoogle Scholar
Dravis, J. J. 1979. Rapid and widespread generation of Recent oolitic hardgrounds on a high energy Bahamian Platform, Eleuthera Bank, Bahamas. Journal of Sedimentary Petrology 49, 195–208.Google Scholar
Dravis, J. J. 1983. Hardened subtidal stromatolites. Science 219, 385–386.CrossRefGoogle ScholarPubMed
Drever, J. J. 1994. The effect of land plants on weathering rates of silicate minerals. Geochimica et Cosmochimica Acta 58, 2325–2332.CrossRefGoogle Scholar
Drever, J. J. 1997. The Geochemistry of Natural Waters, 3rd edn. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Drever, J. J., & Stillings, L. L. 1997. The role of organic acids in mineral weathering. Colloids and Surfaces A: Physicochemical and Engineering Aspects 120, 167–181.CrossRefGoogle Scholar
Driemanis, A. 1988. Tills: their genetic terminology and classification. In Genetic Classification of Glacigenic Deposits (ed. Goldthwait, R. P. & Matsch, C. L.), pp. 17–83. Rotterdam: Balkema.Google Scholar
Driese, S. G., & Mora, C. I. 2001. Diversification of Siluro-Devonian plant traces in paleosols and influence on estimates of paleoatmospheric CO2 levels. In Plants Invade the Land (ed. Gensel, P G. & Edwards, D.), pp. 237–253. New York: Columbia University Press.CrossRefGoogle Scholar
Driese, S. G., Mora, C. I., & Elrick, J. M. 2000. The paleosol record of increasing plant diversity and depth of rooting and changes in atmospheric pCO2 in the Siluro-Devonian. In Phanerozoic Terrestrial Ecosystems: A Short Course (ed. Gastaldo, R. A. & DiMichele, W. A.), pp. 47–61. Lawrence, KA: The Paleontological Society.Google Scholar
Droser, M. L., & Bottjer, D. L. 1986. A semiquantitative field classification of ichnofabric. Journal of Sedimentary Petrology 56, 558–559.CrossRefGoogle Scholar
Droser, M. L., & Bottjer, D. L. 1989. Ichnofabric of sandstones deposited in high energy nearshore environments: measurement and utilization. Palaios 4, 598–604.CrossRefGoogle Scholar
Droz, L., & Bellaiche, G. 1985, Rhone deep-sea fan: morphostructure and growth pattern. American Association of Petroleum Geologists Bulletin 69, 460–479.Google Scholar
Droz, L., & Bellaiche, G.1991. Seismic facies and geological evolution of the central portion of the Indus fan. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 383–402. New York: Springer-Verlag.CrossRefGoogle Scholar
Duc, A. W., & Tye, R. S. 1987. Evolution and stratigraphy of a regressive/backbarrier complex: Kaiwah Island, South Carolina. Sedimentology 34, 237–251.CrossRefGoogle Scholar
Duke, W. L. 1985. Hummocky cross-stratification, tropical hurricanes, and intense winter storms. Sedimentology 32, 167–194.CrossRefGoogle Scholar
Duke, W. L. 1990. Geostrophic circulation or shallow marine turbidity currents? The dilemma of paleoflow patterns in storm-influenced prograding shoreline systems. Journal of Sedimentary Petrology 60, 870–883.CrossRefGoogle Scholar
Duke, W. L., Arnott, R. W., & Cheel, R. J. 1991. Shelf sandstones and hummocky cross-stratification: new insights on a stormy debate. Geology 19, 625–628.2.3.CO;2>CrossRefGoogle Scholar
Dumas, S., Arnott, R. W. C., & Southard, J. B. 2005. Experiments on oscillatory-flow and combined-flow bed forms: implications for interpreting parts of the shallow-marine sedimentary record. Journal of Sedimentary Research 75, 501–513.CrossRefGoogle Scholar
Dunham, R. J. 1969a. Early vadose silt in Townsend mound (reef), New Mexico. In Depositional Environments in Carbonate Rocky, a Symposium (ed. Friedman, G. M.), pp. 139–182. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Dunham, R. J.1969b. Vadose pisolite in the Capitan reef (Permian), New Mexico and Texas. In Depositional Environments in Carbonate Rocky, a Symposium (ed. Friedman, G. M.), pp. 182–190. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Dunlop, P., & Clarke, C. D. 2006. The morphological characteristics of ribbed moraine. Quaternary Science Reviews 25, 1668–1691.CrossRefGoogle Scholar
Dunne, T. 1980. Formation and controls of channel networks. Progress in Physical Geography 4, 211–239.CrossRefGoogle Scholar
Dury, G. H. 1985. Attainable standards of accuracy in the retrodiction of palaeodischarge from channel dimensions. Earth Surface Processes and Landforms 10, 205–213.CrossRefGoogle Scholar
Dutton, S. P., Flanders, W. A., & Barton, M. D. 2003. Reservoir characterization of a Permian deep-water sandstone, East Ford field, Delaware basin, Texas. American Association of Petroleum Geologists Bulletin 87, 609–627.CrossRefGoogle Scholar
Dutton, S. P., White, C. D., Willis, B. J., & Novakovic, D. 2002. Calcite cement distribution and its effect on fluid flow in a deltaic sandstone, Frontier Formation, Wyoming. American Association of Petroleum Geologists Bulletin 86, 2007–2021.Google Scholar
Dyer, K. R. 1989. Sediment processes in estuaries: future research requirements. Journal of Geophysical Research 94C, 14,327–14,339.CrossRefGoogle Scholar
Dyer, K. R., & Huntley, D. A. 1999. The origin, classification and modeling of sand banks and ridges. Continental Shelf Research 19, 1285–1330.CrossRefGoogle Scholar
Dzulyński, S., & Kotlarczyk, J. 1962. On load-casted ripples. Annales de la Société Géologique de Pologne 32, 148–159.Google Scholar
Dzulyński, S., & Walton, E. K. 1965. Sedimentary Features of Flysch and Graywackes. Amsterdam: Elsevier.Google Scholar
Easterbrook, D. J. 1999. Surface Processes and Landforms, 2nd edn. New York: Macmillan Publishing Company.Google Scholar
Edmonds, D. A., & Slingerland, R. 2007. Mechanics of middle ground bar formation: implications for the formation of delta distributary channel networks. Journal of Geophysical Research – Earth Surface 112, F02034, doi:10.1029/2006 JF000574.Google Scholar
Edwards, D. E., Leeder, M. R., Best, J. L., & Pantin, H. M. 1994. An experimental study of reflected density currents and the interpretation of certain turbidites. Sedimentology 41, 437–461.CrossRefGoogle Scholar
Ehrenberg, S. N., & Nadeau, P. H. 2005. Sandstone vs. carbonate petroleum reservoirs: a global perspective on porosity–depth and porosity–permeability relationships. American Association of Petroleum Geologists Bulletin 89, 435–445.CrossRefGoogle Scholar
Einsele, G. 1992. Sedimentary Basins, Evolution, Facies and Sediment Budget. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Einsele, G., Ricken, W., & Seilacher, A. 1991. Cycles and Events in Stratigraphy. Berlin: Springer-Verlag.Google Scholar
Eisma, D. 1997. Intertidal Deposits. Boca Raton, FL: CRC Press.Google Scholar
Ekdale, A. A., Bromley, R. G., & Pemberton, S. G. 1984. Ichnology: The Use of Trace Fossils in Sedimentology and Stratigraphy. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ekdale, A. A., & Pollard, J. 1991. Ichnofabric and ichnofacies. Palios 6, 197–343.CrossRefGoogle Scholar
Elfeki, A., & Dekking, M. 2001. A Markov chain model for subsurface characterization: theory and application. Mathematical Geology 33, 569–589.CrossRefGoogle Scholar
Elliott, T. 1974. Interdistributary bay sequences and their genesis. Sedimentology 21, 611–622.CrossRefGoogle Scholar
Elrick, M., & Snider, A. C. 2002. Deep-water stratigraphic cyclicity and carbonate mud mound development in the Middle Cambrian Marjum Formation, House Range, Utah. Sedimentology 49, 1021–1047.CrossRefGoogle Scholar
Embley, R. W. 1976. New evidence for occurrence of debris flow deposits in the deep sea. Geology 4, 371–374.2.0.CO;2>CrossRefGoogle Scholar
Embley, R. W. 1980. The role of mass transport in the distribution and character of deep-ocean sediments with special reference to the North Atlantic. Marine Geology 38, 23–50.CrossRefGoogle Scholar
Embry, A. F., & Klovan, J. E. 1971. A Late Devonian reef tract on north-eastern Banks Island, N.W.T. Bulletin of Canadian Petroleum Geology 19, 730–781.Google Scholar
Emery, D., & Myers, K. J. (eds.) 1996. Sequence Stratigraphy. Oxford: Blackwells.CrossRefGoogle Scholar
Emery, K. O. 1952. Continental shelf sediments off southern California. Geological Society of America Bulletin 63, 1105–1108.CrossRefGoogle Scholar
Enos, P. 1974. Surface Sediment Facies Map of the Florida–Bahamas Plateau. Boulder, CO: Geological Society of America.Google Scholar
Enos, P.1983. Shelf. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 267–295. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Enos, P. 1991. Sedimentary parameters for computer modeling. Kansas Geological Survey Bulletin 233, 63–99.Google Scholar
Enos, P., & Moore, C. H. 1983. Fore-reef slope. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 507–538. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Enos, P., & Perkins, R. D. 1977. Quaternary Sedimentation in South Florida. Boulder, CO: Geological Society of America.Google Scholar
Enos, P., & Perkins, R. D. 1979. Evolution of Florida Bay from island stratigraphy. Geological Society of America Bulletin 90, 59–83.2.0.CO;2>CrossRefGoogle Scholar
Enos, P., & Sawatsky, L. H. 1981. Pore networks in Holocene carbonate sediments. Journal of Sedimentary Petrology 51, 961–985.Google Scholar
Erickson, M. C., Masson, D. S., Slingerland, R., & Swetland, D. W. 1990. Numerical simulation of circulation and sediment transport in the Late Devonian Catskill Sea. In Quantitative Dynamic Stratigraphy (ed. Cross, T. A.), pp. 293–305. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Ernst, W. G. 2000. Earth Systems: Processes and Issues. Cambridge: Cambridge University Press.Google Scholar
Esteban, M. 1974. Caliche textures and “Microcodium.”Bollettino della Società Geologica Italiana 92, 105–125.Google Scholar
Esteban, M., & Klappa, C. F. 1983. Subaerial exposure. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 1–54. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Esteban, M., & Pray, L. C. 1983. Pisoids and pisolitic facies (Permian), Guadalupe Mountains, New Mexico and West Texas. In Coated Grains (ed. Peryt, T. M.), pp. 503–537. New York: Springer-Verlag.CrossRefGoogle Scholar
Ethridge, F. G., & Schumm, S. A. 1978. Reconstructing paleochannel morphologic and flow characteristics: methodology, limitations and assessment. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 703–721. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Ethridge, F. G., Skelly, R. L., & Bristow, C. S. 1999. Avulsion and crevassing in the sandy, braided Niobrara River: complex response to base-level rise and aggradation. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 179–191. Oxford: Blackwells.CrossRefGoogle Scholar
Eugster, H. P. 1969. Inorganic bedded cherts from the Magadi area, Kenya. Contributions to Mineralogy and Petrology 22, 1–31.CrossRefGoogle Scholar
Eugster, H. P., & Hardie, L. A. 1978. Saline lakes. In Lakes: Chemistry Geology Physics (ed. Lerman, A.), pp. 237–293. New York: Springer-Verlag.CrossRefGoogle Scholar
Evamy, B. D. 1962. The application of a chemical staining technique to a study of dedolomitization. Sedimentology 2, 164–170.CrossRefGoogle Scholar
Evans, D. J. A. 2003. Glacial Landsystems. London: Arnold.Google Scholar
Evans, D. J. A., Phillips, E. R., Hiemstra, J. F., & Auton, C. A. 2006a. Subglacial till: formation, sedimentary characteristics and classification. Earth-Science Reviews 78, 115–176.CrossRefGoogle Scholar
Evans, D. J. A., Rea, B. R., Hiemstra, J. F., & Cofaigh, O C. 2006b. A critical assessment of subglacial mega-floods: a case study of glacial sediments and landforms in south-central Alberta, Canada. Quaternary Science Reviews 25, 1638–1667.CrossRefGoogle Scholar
Evans, G. 1965. Intertidal flat sediments and their environments of deposition in the Wash. Quarterly Journal of the Geology Society of London 121, 209–245.CrossRefGoogle Scholar
Ewers, W. E., & Morris, R. C. 1981. Studies of the Dales Gorge Member of the Brockman Iron Formation, Western Australia. Economic Geology 76, 1929–1953.CrossRefGoogle Scholar
Eyles, C. H., Eyles, N., & Miall, A. D. 1985. Models of glaciomarine sedimentation and their application to the ancient glacial record. Palaeogeography, Palaeoclimatology, Palaeoecology 51, 15–84.CrossRefGoogle Scholar
Eyles, N., & Eyles, C. H. 1992. Glacial depositional systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 73–100. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Eyles, N., Sladen, J., & Gilroy, S. 1982. A depositional model for stratigraphic complexes and facies superimposition in lodgement tills. Boreas 11, 317–333.CrossRefGoogle Scholar
Faugeres, J.-C., Mezerais, M.-L., & Stow, D. A. V. 1993. Contourite drift types and their distribution in the North and South Atlantic ocean basins. Sedimentary Geology 82, 189–206.CrossRefGoogle Scholar
Farquhar, J., Bao, H., & Thiemens, M. 2000. Atmospheric influence of Earth's earliest sulfur cycle. Science 289, 756–758.CrossRefGoogle ScholarPubMed
Farrell, K. M. 1987. Sedimentology and facies architecture of overbank deposits of the Mississippi River, False River Region, Louisiana. In Recent Developments in Fluvial Sedimentology (ed. Ethridge, F. G., Flores, R. M., & Harvey, M. D.), pp. 111–120. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Farrell, K. M. 2001. Geomorphology, facies architecture, and high resolution, non-marine sequence stratigraphy in avulsion deposits, Cumberland marshes, Saskatchewan. Sedimentary Geology 139, 93–150.CrossRefGoogle Scholar
Faure, G. 1991. Principles and Applications of Inorganic Geochemistry, 2nd edn. New York: MacMillan Publishing Company.Google Scholar
Fawthrop, N. P. 1996. Modelling hydrological processes for river management. In River Flows and Channel Forms (ed. Petts, G. & Calow, P.), pp. 51–76. Oxford: Blackwell Science Ltd.Google Scholar
Felix, M. 2001. A two-dimensional numerical model for a turbidity current. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 71–81. Oxford: Blackwells.CrossRefGoogle Scholar
Field, M. E. 1980. Sand bodies on coastal plain shelves: Holocene record of the U.S. Atlantic shelf off Maryland. Journal of Sedimentary Petrology 50, 505–528.Google Scholar
Figueredo, A. G. J., Sanders, J. E., & Swift, D. J. P. 1982. Storm graded layers on inner continental shelves: examples from south Brazil and the Atlantic coast of the central United States. Sedimentary Geology 31, 171–190.CrossRefGoogle Scholar
Fischer, A. G. 1964. The Lofer cyclothems of the Alpine Triassic. Kansas State Geological Survey Bulletin 169, 7–149.Google Scholar
Fischer, A. G. 1982. Long-term climatic oscillations recorded in stratigraphy. In Climate in Earth History (ed. Geophysics Study Committee), pp. 97–104. Washington, D.C.: National Academy Press.Google Scholar
Fisher, R. V., & Schminke, H. U. 1984. Pyroclastic Rocks. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Fisher, R. V., & Smith, G. A. 1991. Sedimentation in Volcanic Settings. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Fisher, W. L., Brown, L. F., Scott, A. J., & McGowen, J. H. 1969. Delta Systems in the Exploration for Oil and Gas. Austin, TX: Bureau of Economic Geology, University of Texas.Google Scholar
Fisk, H. N. 1961. Bar finger sands of the Mississippi delta. In Geometry of Sandstone Bodies – A Symposium (ed. Peterson, J. A. & Osmond, J. C.), pp. 29–52. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Fisk, H. N., McFarlan, E. Jr., Kolb, C. R., & Wilbert, L. J. Jr. 1954. Sedimentary framework of the modern Mississippi delta. Journal of Sedimentary Petrology 24, 76–99.CrossRefGoogle Scholar
Fitzgerald, D. M. 1988. Shoreline erosional–depositional processes associated with tidal inlets. In Hydrodynamics and Sediment Dynamics of Tidal Inlets (ed. Aubrey, D. G. & Weishar, L.), pp. 186–225. Berlin: Springer-Verlag.Google Scholar
Fitzgerald, D. M.1993. Origin and stability of tidal inlets in Massachusetts. In Formation and Evolution of Multiple Tidal Inlets (ed. Aubrey, D. G. & Geise, F. J.), pp. 1–61. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Fitzgerald, M. G., Mitchum, R. M. Jr., Uliana, M. A., & Biddle, K. T. 1990. Evolution of San Jorge Basin. American Association of Petroleum Geologists Bulletin 74, 879–920.Google Scholar
Fitzpatrick, E. A. 1984. Micromorphology of Soils. London: Chapman and Hall.CrossRefGoogle Scholar
Flemming, B. W. 1980. Sand transport and bedform patterns on the continental shelf between Durban and Port Elizabeth, southeast African continental margin. Sedimentary Geology 26, 179–205.CrossRefGoogle Scholar
Flemming, B. W., & Bartholoma, A. 1995. Tidal Signatures in Modern and Ancient Sediments. Oxford: Blackwells.CrossRefGoogle Scholar
Flemings, P. B., & Jordan, T. E. 1989. A synthetic stratigraphic model of foreland basin development. Journal of Geophysical Research 94, 3851–3866.CrossRefGoogle Scholar
Flemings, P. B., & Jordan, T. E. 1990. Stratigraphic modeling of foreland basins: interpreting thrust deformation and lithospheric rheology. Geology 18, 430–434.2.3.CO;2>CrossRefGoogle Scholar
Flint, R. F. 1971. Glacial and Quaternary Geology. New York: Wiley.Google Scholar
Flint, S. S., Aitken, J., & Hampson, G. 1995. Application of sequence stratigraphy to coal-bearing coastal plain successions: implications for the UK coal measures. In European Coal Geology (ed. Whateley, M. K. G. & Spears, D. A.), pp. 1–16. London: Geological Society of London.Google Scholar
Flood, R. D., & Damuth, J. E. 1987. Quantitative characteristics of sinuous distributary channels on the Amazon deep-sea fan. Geological Society of America Bulletin 98, 728–738.2.0.CO;2>CrossRefGoogle Scholar
Flood, R. D., Manley, P. C., Kowsman, R. O., Appi, C. J., & Pirmez, C. 1991. Seismic facies and Late Quaternary growth of Amazon submarine fan. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 415–433. New York: Springer-Verlag.CrossRefGoogle Scholar
Flügel, E. 2004. Microfacies of Carbonate Rocks: Analysis, Interpretation and Application. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Folk, R. L. 1959. Practical petrographic classification of limestones. American Association of Petroleum Geologists Bulletin 43, 1–38.Google Scholar
Folk, R. L. 1965a. Petrology of Sedimentary Rocks. Austin, TX: Hemphil's.Google Scholar
Folk, R. L.1965b. Some aspects of recrystalization in ancient limestones. In Dolomitization and Limestone Diagenesis; A Symposium (ed. Pray, L. C. & Murray, R. C.), pp. 14–48. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Folk, R. L., & McBride, E. F. 1976. The Caballos Novaculite revisited. Part I: origin of the novaculite members. Journal of Sedimentary Petrology 46, 659–669.Google Scholar
Forristall, G. S., Hamilton, R. C., & Cardone, V. J. 1977. Continental shelf currents in Tropical Storm Delia: observation and theory. Journal of Physical Oceanography 7, 532–546.2.0.CO;2>CrossRefGoogle Scholar
Fortin, D., & Langley, S. 2005. Formation and occurrence of biogenic iron-rich minerals. Earth-Science Reviews 72, 1–19.CrossRefGoogle Scholar
Fouch, T. D., & Dean, W. E., 1982. Lacustrine environments. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D.), pp. 87–114. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Frakes, L. A. 1979. Climates Throughout Geologic History. Amsterdam: Elsevier.Google Scholar
Frakes, L. A., Francis, J. E., & Syktus, J. I. 1992. Climate Modes of the Phanerozoic. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Francis, J. E., & Smith, M. P. 2002. Paleoclimate reconstruction using fossils and lithologic indicators. Palaeogeography, Palaeoclimatology, Palaeoecology 182, 1–143.CrossRefGoogle Scholar
Francis, P. W. 1993. Volcanoes: A Planetary Perspective. Oxford: Oxford University Press.Google Scholar
Franseen, E. K., Watney, W. L., Kendall, C. G., C.St., & Ross, W. 1991. Sedimentary Modeling: Computer Simulations and Methods for Improved Parameter Definition. Lawrence, KA: Kansas Geological Survey.Google Scholar
Fraser, G. S., & DeCelles, P. G. 1992. Geomorphic controls on sediment accumulation at margins of foreland basins. Basin Research 4, 233–252.CrossRefGoogle Scholar
Frazier, D. E. 1967. Recent deltaic deposits of the Mississippi delta: their development and chronology. Transactions of the Gulf Coast Association of Geological Societies 17, 287–315.Google Scholar
Fredsoe, J., & Deigaard, R. 1992. Mechanics of Coastal Sediment Transport. Singapore: World Scientific.CrossRefGoogle Scholar
Freeze, R. A., & Cherry, J. A. 1979. Groundwater. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Frey, R. W., Howard, J. D., Han, S. J., & Park, B. K. 1989. Sediments and sedimentary sequences on a modern macrotidal flat, Inchon, Korea. Journal of Sedimentary Petrology 59, 28–44.Google Scholar
Friedman, G. M. 1959. Identification of carbonate minerals by staining methods. Journal of Sedimentary Petrology 29, 87–97.Google Scholar
Friedman, G. M., & Sanders, J. E. 1967. Origin and occurrence of dolostones. In Carbonate Rocks, Part A: Origin, Occurrence, and Classification (ed. Chilingar, G. V., Bissell, H. J., & Fairbridge, R. W.), pp. 167–348. Amsterdam: Elsevier.Google Scholar
Friedman, S. J. 1998. Rock-avalanche elements of the Shadow Valley Basin, Eastern Mojave Desert, California: processes and problems. Journal of Sedimentary Research 67, 792–804.Google Scholar
Frostick, L. E., & Reid, I. 1987. Desert Sediment: Ancient and Modern. London: Geological Society of London.Google Scholar
Fryberger, S. G. 1993. A review of aeolian bounding surfaces, with examples from the Permian Minnelusa Formation, USA. In Characterisation of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 167–197. London: Geological Society of London.Google Scholar
Fryberger, S. G., Ahlbrandt, T. S., & Andrews, S. 1979. Origin, sedimentary features and significance of low-angle eolian “sand sheet” deposits, Great Sand Dunes National Monument and vicinity, Colorado. Journal of Sedimentary Petrology 49, 733–746.CrossRefGoogle Scholar
Füchtbauer, H., & Hardie, L. A. 1976. Experimentally determined homogeneous distribution coefficients for precipitated magnesium calcites: application to marine carbonate cements. Geological Society of America Abstracts with Program 8, 877.Google Scholar
Füchtbauer, H., & Hardie, L. A. 1980. Comparison of experimental and natural magnesium calcites. IAS International Meeting, Bochum. Bochum: International Association of Sedimentologists. pp. 167–169.Google Scholar
Galloway, W. E. 1975. Process framework for describing the morphologic and stratigraphic evolution of deltaic depositional systems. In Deltas, Models for Exploration (ed. Broussard, M. L.), pp. 87–98. Houston, TX: Houston Geological Society.Google Scholar
Galloway, W. E.1981. Depositional architecture of Cenozoic Gulf coastal plain fluvial systems. In Recent and Ancient Nonmarine Depositional Environments: Models for Exploration (ed. Ethridge, F. G. & Flores, R. M.), pp. 127–155. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Galloway, W. E. 1989a. Genetic stratigraphic sequences in basin analysis I: architecture and genesis of flooding-suface bounded depositional units. American Association of Petroleum Geologists Bulletin 73, 125–142.Google Scholar
Galloway, W. E. 1989b. Genetic stratigraphic sequences in basin analysis II: application to northwest Gulf of Mexico Cenozoic basin. American Association of Petroleum Geologists Bulletin 73, 143–154.Google Scholar
Garcia-Castellanos, D. 2002. Interplay between lithospheric flexure and river transport in foreland basins. Basin Research 14, 89–104.CrossRefGoogle Scholar
Garrels, R. M., & MacKenzie, F. T. 1971. Evolution of Sedimentary Rocks. New York: W. W. Norton.Google Scholar
Garret, P. 1977. Biological communities and their sedimentary record. In Sedimentation on the Modern Carbonate Tidal Flats of Northwest Andros Island (ed. Hardie, L. A.), pp. 124–158. Baltimore, MA: The Johns Hopkins University Press.
Garven, G., Ge, S., Person, M. A., & Sverjensky, D. A. 1993. Genesis of stratabound ore deposits in the mid-continent of North America. 1. The role of regional groundwater flow. American Journal of Science 293, 497–568.CrossRefGoogle Scholar
Gawthorpe, R. L., & Leeder, M. R. 2000. Tectono-sedimentary evolution of active extensional basins. Basin Research 12, 195–218.CrossRefGoogle Scholar
Gaynor, G. C., & Swift, D. J. P. 1988. Shannon Sandstone depositional model; sand ridge formation on the Campanian western interior shelf. Journal of Sedimentary Petrology 58, 868–880.Google Scholar
Gebelein, C. D. 1974. Guidebook for Modern Bahamian Platform Environments. St. George, Bermuda: Geological Society of America.Google Scholar
Gebelein, C. D., Steinen, R. P., Garrett, P. et al. 1980. Subsurface dolomitization beneath the tidal flats of Central West Andros Island, Bahamas. In Concepts and Models of Dolomitization (ed. Zenger, D. H., Dunham, J. B., & Ethington, R. L.), pp. 31–49. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Genise, J. F., & Brown, T. M. 1994a. New Miocene scarabeid and hymenopterous nests and early Miocene (Santacrucian) paleoenvironments. Ichnos 3, 107–117.CrossRefGoogle Scholar
Genise, J. F., & Brown, T. M. 1994b. New trace fossils of termites (insecta: isoptera) from the late Eocene–early Miocene of Egypt, and the reconstruction of ancient isopteran behavior. Ichnos 3, 155–183.CrossRefGoogle Scholar
Genise, J. F., Mangano, M. G., Buatois, L. A., Laza, J. H., & Verde, M. 2000. Insect trace fossil associations in paleosols: the Coprinisphaera ichnofacies. Palaios 15, 49–64.2.0.CO;2>CrossRefGoogle Scholar
Gensel, P. G., & Edwards, D. 2001. Plants Invade the Land. New York: Columbia University Press.CrossRefGoogle Scholar
George, G., & Berry, J. K. 1993. A new lithostratigraphy and depositional model for the Upper Rotleigend of the UK sector of the southern North Sea. In Characterisation of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 291–320. London: Geological Society of London.Google Scholar
Ghibaudo, G. 1992. Subaqueous sediment gravity flow deposits: practical criteria for their field description and classification. Sedimentology 39, 423–454.CrossRefGoogle Scholar
Ghosh, B., & Lowe, D. R. 1993. The architecture of deep-water channel complexes, Cretaceous Venado Sandstone Member, Sacramento Valley, California. In Advances in the Sedimentary Geology of the Great Valley Group, Sacramento Valley, California (ed. Graham, S. A. & Lowe, D. R.), pp. 51–65. Los Angeles, CA: Pacific Section, Society of Economic Paleontologists and Mineralogists.Google Scholar
Gibling, M. R., & Bird, D. J. 1994. Late Carboniferous cyclothems and alluvial paleovalleys in the Sydney Basin, Nova Scotia. Geological Society of America Bulletin 106, 105–117.2.3.CO;2>CrossRefGoogle Scholar
Gilbert, G. K. 1885. The Topographic Features of Lake Shores. Washington, D.C.: United States Geological Survey.Google Scholar
Giles, M. R. 1997. Diagenesis: A Quantitative Perspective, Implications for Basin Modelling and Rock Property Prediction. Dordrecht: Kluwer Academic Publishers.Google Scholar
Giles, M. R., & Boer, R. B. 1990. Origin and significance of redistributional secondary porosity. Marine and Petroleum Geology 7, 378–397.CrossRefGoogle Scholar
Gill, W. D., & Kuenen, P. H. 1958. Sand volcanoes on slumps in the Carboniferous of County Clare, Ireland. Quarterly Journal of the Geology Society of London 113, 441–460.CrossRefGoogle Scholar
Ginsburg, R. N. 1967. Stromatolites. Science 157, 339–340.CrossRefGoogle ScholarPubMed
Ginsburg, R. N. 2001. Subsurface Geology of a Prograding Carbonate Platform Margin, Great Bahama Bank: Results of the Bahamas Drilling Project. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ginsburg, R. N., Isham, L. B., Bein, S. J., & Kuperburg, J. 1954. Laminated Algal Sediments of South Florida and Their Recognition in the Fossil Record. Coral Gables, FL: University of Miami Marine Laboratory.Google Scholar
Ginsburg, R. N., & James, N. P. 1974. Holocene carbonate sediments of continental margins. In The Geology of Continental Margins (ed. Burke, C. A. & Drake, C. L.), pp. 137–155. New York: Springer-Verlag.CrossRefGoogle Scholar
Giosan, L., & Bhattacharya, J. P. 2005. River Deltas – Concepts, Models, and Examples. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Glen, J. W. 1952. Experiments on the deformation of ice. Journal of Glaciology 2, 111–114.CrossRefGoogle Scholar
Glen, J. W. 1955. The creep of polycrystalline ice. Proceedings of the Royal Society of London 228A, 519–538.CrossRefGoogle Scholar
Glenn, C. R., & Garrison, R. E. 2003. Phosphorites. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 257–263. Dordrecht: Kluwer Academic Publisher.Google Scholar
Glennie, K. W. 1970. Desert Sedimentary Environments. Amsterdam: Elsevier.Google Scholar
Glennie, K. W.1986. Early Permian Rotleigend. In Introduction to the Petroleum Geology of the North Sea (ed. Glennie, K. W.), pp. 63–86. Oxford: Blackwells.Google Scholar
Goldhammer, R. K. 1997. Compaction and decompaction algorithms for sedimentary carbonates. Journal of Sedimentary Petrology 67, 26–35.CrossRefGoogle Scholar
Goldhammer, R. K., Dunn, P. A., & Hardie, L. A. 1987. High frequency glacio-eustatic sea level oscillations with Milankovitch characteristics recorded in Middle Triassic platform carbonates in northern Italy. American Journal of Science 287, 853–892.CrossRefGoogle Scholar
Goldhammer, R. K., Dunn, P. A., & Hardie, L. A. 1990. Depositional cycles, composite sea-level changes, cycle stacking patterns, and the hierarchy of stratigraphic forcing: examples from Alpine Triassic platform carbonates. Geological Society of America Bulletin 102, 535–562.2.3.CO;2>CrossRefGoogle Scholar
Goldhammer, R. K., & Harris, M. T. 1989. Eustatic controls on the stratigraphy and geometry of the Latemar buildup (Middle Triassic): the Dolomites of northern Italy. In Controls on Carbonate Platform and Basin Development (ed. Crevello, P. D., Wilson, J. L., Sarg, J. F., & Read, J. F.), pp. 323–338. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Goldrich, S. S. 1938. A study in rock weathering. Journal of Geology 46, 17–58.CrossRefGoogle Scholar
Goldsmith, J. R., & Graf, D. L. 1958. Relation between lattice constants and composition of the Ca–Mg carbonates. American Mineralogist 43, 84–101.Google Scholar
Goldsmith, J. R., Graf, D. L., & Joensuu, O. I. 1955. The occurrence of magnesium calcites in nature. Geochimica et Cosmochimica Acta 7, 212–230.CrossRefGoogle Scholar
Gosse, J. C., & Phillips, F. M. 2001. Terrestrial in situ cosmogenic nuclides: theory and applications. Quaternary Sciences Reviews 20, 1475–1560.CrossRefGoogle Scholar
Goudie, A. S. 1973. Duricrusts in Tropical and Subtropical Landscapes. Oxford: Clarendon Press.Google Scholar
Goudie, A. S.1999. Wind erosional landforms: yardangs and pans. In Aeolian Environments, Sediments and Landforms (ed. Goudie, A. S., Livingstone, I., & Stokes, S.), pp. 167–180. Chichester: Wiley.Google Scholar
Goudie, A. S., Livingstone, I., & Stokes, S. 1999. Aeolian Environments, Sediments and Landforms. Chichester: Wiley.Google Scholar
Gould, H. R. 1970. The Mississippi delta complex. In Deltaic Sedimentation Modern and Ancient (ed. Morgan, J. P. & Shaver, R. H.), pp. 3–30. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Graf, D. L., & Goldsmith, J. R. 1956. Some hydrothermal syntheses of dolomite and protodolomite. Journal of Geology 64, 173–186.CrossRefGoogle Scholar
Grant, W. D., & Madsen, O. S. 1979. Combined wave and current interaction with a rough bottom. Journal of Geophysical Research 84, 1797–1808.CrossRefGoogle Scholar
Grass, A. J. 1970. Initial instability of fine sand beds. Journal of the Hydraulic Division of the ASCE 96, 619–632.Google Scholar
Greeley, R., & Iverson, J. D. 1985. Wind as a Geological Process: On Earth, Mars, Venus and Titan. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Green, M. O., Rees, J. M., & Pearson, N. D. 1990. Evidence for the influence of wave–current interactions in a tidal boundary layer. Journal of Geophysical Research 95C, 9629–9644.CrossRefGoogle Scholar
Greenwood, B., & Mittler, P. R. 1985. Vertical sequence and lateral transitions in the facies of a barred nearshore environment. Journal of Sedimentary Petrology 55, 366–375.Google Scholar
Greenwood, B., & Sherman, D. J. 1986. Hummocky cross-stratification in the surf zone: flow parameters and bedding genesis. Sedimentology 33, 33–45.CrossRefGoogle Scholar
Greenwood, B., & Sherman, D. J. 1993. Waves, currents, sediment flux and morphological response in a barred nearshore system. Marine Geology 60, 31–61.CrossRefGoogle Scholar
Gregg, J. M., & Sibley, D. F. 1984. Epigenetic dolomitization and the origin of xenotopic dolomite textures. Journal of Sedimentary Petrology 54, 908–931.Google Scholar
Gregory, K. J. 1983. Background to Palaeohydrology. Chichester: Wiley.Google Scholar
Gregory, K. J., Lewin, J., & Thornes, J. B. 1987. Palaeohydrology in Practice. Chichester: Wiley.Google Scholar
Gregory, K. J., Starkel, L., & Baker, V. R. 1996. Continental Palaeohydrology. Chichester: Wiley.Google Scholar
Griffing, D. H., Bridge, J. S. & Hotton, C. L. 2000. Coastal–fluvial paleoenvironments and plant ecology of the early Devonian (Emsian), Gaspe Bay, Canada. In New Perspectives on the Old Red Sandstone (ed. Friend, P. F., & Williams, B. P. J.), pp. 61–84. London: Geological Society of London.Google Scholar
Griffiths, C. M., Dyt, C., Paraschivoiu, E., & Liu, K. 2001. SEDSIM in hydrocarbon exploration. In Geologic Modeling and Simulation: Sedimentary Systems (ed. Merriam, D. F. & Davis, J. C.), pp. 71–97. New York: Kluwer Academic/Plenum Publishers.CrossRefGoogle Scholar
Grotzinger, J. P., & James, N. P. 2000. Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Grotzinger, J. P., Jordan, T. H., Press, F., & Siever, R. 2007. Understanding Earth, 5th edn. New York: W. H. Freeman and Company.Google Scholar
Grotzinger, J. P., & Knoll, A. H. 1999. Stromatolites in Precambrian Carbonates: evolutionary mileposts or environmental dipsticks. Annual Review of Earth and Planetary Sciences 27, 313–358.CrossRefGoogle ScholarPubMed
Grover, G. Jr., & Read, J. F. 1978. Fenestral and associated vadose diagenetic fabrics of tidal flat carbonates, Middle Ordovician New Market Limestone, southwestern Virginia. Journal of Sedimentary Petrology 48, 453–473.Google Scholar
Grunau, H. R. 1965. Radiolarian cherts and associated rocks in space and time. Eclogae Geologicae Helvetiae 58, 157–208.Google Scholar
Gruszczynski, M., Rudowski, S., Semil, J., Slominski, J., & Zrobek, J. 1993. Rip currents as a geological tool. Sedimentology 40, 217–236.CrossRefGoogle Scholar
Gupta, S. 1997. Himalayan drainage patterns and the origin of fluvial megafans in the Ganges foreland basin. Geology 25, 11–14.2.3.CO;2>CrossRefGoogle Scholar
Gupta, S., & Cowie, P. 2000. Processes and controls in the stratigraphic development of extensional basins. Basin Research 12, 185–241.CrossRefGoogle Scholar
Gustard, A. 1996. Analysis of river regimes. In River Flows and Channel Forms (ed. Petts, G. & Calow, P.), pp. 32–50. Oxford: Blackwell Science Ltd.Google Scholar
Guza, R. T., & Bowen, A. J. 1975. The resonant instabilities of long waves obliquely incident on a beach. Journal of Geophysical Research 80, 4529–4534.CrossRefGoogle Scholar
Guza, R. T., & Davis, R. E. 1974. Excitation of edge waves by waves incident on a beach. Journal of Geophysical Research 79, 1285–1291.CrossRefGoogle Scholar
Haff, P. K. 1983. Grain flow as a fluid-mechanical phenomenon. Journal of Fluid Mechanics 134, 401–430.CrossRefGoogle Scholar
Hallet, B. 1996. Glacial quarrying: a simple theoretical model. Annals of Glaciology 22, 1–8.CrossRefGoogle Scholar
Halley, R. B., & Evans, C. C. 1983. The Miami Limestone. A Guide to Selected Outcrops and Their Interpretation (with a Discussion of Diagenesis in the Formation). Miami, FL: Miami Geological Society.Google Scholar
Halley, R. B., Harris, P. M., & Hine, A. C., 1983. Bank margin. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 463–506. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Hambrey, M. 1994. Glacial Environments. London: UCL Press.Google Scholar
Hambrey, M., & Harland, W. B. 1981. Earth's Pre-Pleistocene Glacial Record. Cambridge: Cambridge University Press.Google Scholar
Hampson, G. J., Elliott, T., & Davies, S. J. 1997. The application of sequence stratigraphy to Upper Carboniferous fluvio-deltaic strata of the onshore UK and Ireland: implications for the southern North Sea. Journal of the Geological Society of London 154, 719–733.CrossRefGoogle Scholar
Hampson, G. J., Davies, S. J., Elliott, T., Flint, S. S., & Stollhofen, H. 1999. Incised valley fill sandstone bodies in Upper Carboniferous fluvio-deltaic strata: recognition and reservoir characterization of Southern North Sea analogues. In Petroleum Geology of Northwest Europe (ed. Fleet, A. J. & Boldy, S. A. R.), pp. 771–788. London: The Geological Society.Google Scholar
Hampton, M. A. 1972. The role of subaqueous debris flow in generating turbidity currents. Journal of Sedimentary Petrology 42, 775–793.Google Scholar
Hampton, M. A. 1975. Competence of fine-grained debris flows. Journal of Sedimentary Petrology 45, 834–844.Google Scholar
Handford, C. F., Kendall, A. C., Prezbindowski, D. R., Dunham, J. B., & Logan, B. W. 1984. Salina-margin tepees, pisolites, and aragonite cements, Lake MacLeod, Western Australia: their significance in interpreting ancient analogs. Geology 12, 523–527.2.0.CO;2>CrossRefGoogle Scholar
Handford, C. R., Loucks, R. G., & Davies, G. R. 1985. Depositional and Diagenetic Spectra of Evaporites – A Core Workshop. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Hanna, S. R. 1969. The formation of longitudinal sand dunes by large helical eddies in the atmosphere. Journal of Applied Meteorology 8, 874–883.2.0.CO;2>CrossRefGoogle Scholar
Hannington, M. D., Jonasson, I. R., Herzig, P. M., & Petersen, S. 1995. Physical and chemical processes of seafloor mineralization at mid-ocean ridges. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions (ed. Humphris, S. E., Zierenberg, R. A., Mullineaux, L. S., & Thomson, R. E.), pp. 115–157. Washington: D.C.: American Geophysical Union.CrossRefGoogle Scholar
Hanor, J. S. 1994. Origin of saline fluids in sedimentary basins. In Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins (ed. Parnell, J.), pp. 151–174. London: Geological Society of London.Google Scholar
Hanor, J. S., & McIntosh, J. C. 2006. Are secular variations in seawater chemistry reflected in the compositions of basinal brines?Journal of Geochemical Exploration 89, 153–156.CrossRefGoogle Scholar
Häntschel, W. 1975. Trace Fossils and Problematica. Treatise on Invertebrate Paleontology, Part W Supplement 1. Boulder, CO: Geological Society of America; and Lawrence, KA: The University of Kansas.Google Scholar
Harbaugh, J. W., Watney, W. L., Rankey, E. C.et al. 1999. Numerical Experiments in Stratigraphy: Recent Advances in Stratigraphic and Sedimentologic Computer Simulations. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Harbor, J. M. 1992. Numerical modeling of the development of U-shaped valleys by glacial erosion. Geological Society of America Bulletin 104, 1364–1375.2.3.CO;2>CrossRefGoogle Scholar
Hardie, L. A. 1977. Sedimentation on the Modern Carbonate Tidal Flats of Northwest Andros Island, Bahamas. Baltimore, MA: The Johns Hopkins University Press.Google Scholar
Hardie, L. A. 1984. Evaporites: marine or nonmarine?American Journal of Science 284, 193–240.CrossRefGoogle Scholar
Hardie, L. A. 1987. Dolomitization: a critical view of some current views. Journal of Sedimentary Petrology 57, 166–183.CrossRefGoogle Scholar
Hardie, L. A. 1990. The roles of rifting and hydrothermal CaCl2 brines in the origin of potash evaporites: an hypothesis. American Journal of Science 290, 43–106.CrossRefGoogle Scholar
Hardie, L. A. 1991. On the significance of evaporites. Annual Review of Earth and Planetary Sciences 19, 131–168.CrossRefGoogle Scholar
Hardie, L. A. 1996. Secular variation in seawater chemistry: an explanation for the coupled secular variation in the mineralogies of marine limestones and potash evaporites over the past 600 m.y. Geology 34, 279–283.2.3.CO;2>CrossRefGoogle Scholar
Hardie, L. A.2003a Evaporites. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 257–263. Dordrecht: Kluwer Academic Publishing.Google Scholar
Hardie, L. A.2003b Sabkha, salt flat, salina. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 584–585. Dordrecht: Kluwer Academic Publishing.Google Scholar
Hardie, L. A., & Eugster, H. P. 1970. The evolution of closed basin brines. Mineralogical Society of America Special Paper 3, 273–290.Google Scholar
Hardie, L. A., & Lowenstein, T. K. 2004. Did the Mediterranean Sea dry out during the Miocene? A reassessment of the evaporite evidence from DSDP Legs 13 and 42 A cores. Journal of Sedimentary Research 74, 453–461.CrossRefGoogle Scholar
Hardie, L. A., Lowenstein, T. K., & Spencer, R. J. 1985. The problem of distinguishing primary from secondary evaporites. In Proceedings of the Sixth International Symposium on Salt (ed. Schreiber, B. C. & Harner, H. L.), pp. 11–39. Alexandria, VA: The Salt Institute.Google Scholar
Hardie, L. A., & Shinn, E. A. 1986. Carbonate depositional environments modern and ancient, part 3: tidal flats. Colorado School of Mines Quarterly 81, 1–74.Google Scholar
Hardie, L. A., Smoot, J. P., & Eugster, H. P. 1978. Saline lakes and their deposits, a sedimentological approach. In Modern and Ancient Lake Sediments (ed. Matter, A. & Tucker, M. E.), pp. 7–41. Oxford: Blackwells.CrossRefGoogle Scholar
Harms, J. C., Southard, J. B., & Walker, R. G. 1982. Structures and Sequences in Clastic Rocks. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Harris, N. B. 1989. Diagenetic quartzarenite and destruction of secondary porosity: an example from the Middle Jurassic Brent sandstone of northwest Europe. Geology 17, 361–364.2.3.CO;2>CrossRefGoogle Scholar
Harris, P. M. 1979. Facies Anatomy and Diagenesis of a Bahamian Ooid Shoal. Sedimenta 7. Miami, FL: Comparative Sedimentology Laboratory, University of Miami.Google Scholar
Harris, P. T., Pattiaratchi, C. B., Cole, A. R., & Keene, J. B. 1992. Evolution of subtidal sandbanks in Moreton bay, eastern Australia. Marine Geology 103, 225–247.CrossRefGoogle Scholar
Harris, P. T., Pattiaratchi, C. B., Collins, M. B., & Dalrymple, R. W. 1995. What is a bedload parting? In Tidal Signatures in Modern and Ancient Sediments (ed. Flemming, B. W. & Bartholoma, A.), pp. 3–18. Oxford: Blackwells.CrossRefGoogle Scholar
Harris, P. T., Pattiaratchi, C. B., Keene, J. B.et al. 1996. Late Quaternary deltaic and carbonate sedimentation in the Gulf of Papua foreland basin: response to sea-level change. Journal of Sedimentary Research B66, 801–819.Google Scholar
Harrison, W. J., & Summa, L. L. 1991. Paleohydrology of the Gulf of Mexico Basin. American Journal of Science 291, 109–176.CrossRefGoogle Scholar
Hart, B. S., & Plint, A. G. 1989. Gravelly shoreface deposits: a comparison of modern and ancient facies sequences. Sedimentology 36, 551–557.CrossRefGoogle Scholar
Hartley, A. J., & Prosser, D. J. 1995. Characterization of Deep Marine Clastic Systems. London: Geological Society of London.Google Scholar
Hartmann, D. L. 1994. Global Physical Climatology. San Diego, CA: Academic Press.Google Scholar
Harvie, C. E., Möller, N., & Weare, J. H. 1984. The prediction of mineral solubilities in natural waters: the Na–K–Mg–Ca–H–Cl–SO4–OH–HCO3–CO3–CO2–H2O system to high ionic strengths at 25 °C. Geochimica et Cosmochimica Acta 48, 723–752.CrossRefGoogle Scholar
Harvie, C. E., Weare, J. H., Hardie, L. A., & Eugster, H. P. 1980. Evaporation of sea water: calculated mineral sequences. Science 208, 498–500.Google Scholar
Hasiotis, S. T. 2002. Continental Trace Fossils. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google 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, 259–320.CrossRefGoogle Scholar
Hasiotis, S. T. 2004. Reconnaissance of Upper Jurassic Morrison Formation ichnofossils, Rocky Mountain Region, USA: paleoenvironmental, stratigraphic, and paleoclimatic significance of terrestrial and freshwater ichnocoenoses. Sedimentary Geology 167, 177–268.CrossRefGoogle Scholar
Hättestrand, C., & Kleman, J. 1999. Ribbed moraine formation. Quaternary Science Reviews 18, 43–61.CrossRefGoogle Scholar
Havholm, K. G., Blakey, R. C., Capps, M. et al. 1993. Aeolian genetic stratigraphy: an example from the Middle Jurassic Page Sandstone, Colorado Plateau. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 87–107. Oxford: Blackwells.CrossRefGoogle Scholar
Hawley, N. 1981. Flume experiments on the origin of flaser bedding. Sedimentology 28, 699–712.CrossRefGoogle Scholar
Hayes, M. O. 1967. Hurricanes as Geological Agents: Case Studies of Hurricanes Carla, 1961, and Cindy, 1963. Austin, TX: Bureau of Economic Geology.Google Scholar
Hayes, M. O.1975. Morphology of sand accumulation in estuaries: an introduction to the symposium. In Estuarine Research: Volume II Geology and Engineering (ed. Cronin, L E.), pp. 3–22. London: Academic Press.Google Scholar
Hayes, M. O.1979. Barrier island morphology as a function of tidal and wave regime. In Barrier Islands – From the Gulf of St Lawrence to the Gulf of Mexico (ed. Leatherman, S. P.), pp. 1–27. London: Academic Press.Google Scholar
Heckel, P. H., Gibling, M. R., & King, N. R. 1998. Stratigraphic model for glacial–eustatic Pennsylvanian cyclotherms in highstand nearshore detrital regimes. Journal of Geology 106, 373–383.CrossRefGoogle Scholar
Heezen, B. C., Ericson, D. B., & Ewing, M. 1954. Further evidence for a turbidity current following the 1929 Grand banks earthquake. Deep-Sea Research 1, 193–202.CrossRefGoogle Scholar
Heiken, G., & Wohletz, K. 1985. Volcanic Ash. Berkeley: University of California Press.Google Scholar
Heikoop, J. M., Tsujita, C. J., Risk, M. J., Tomascik, T., & Mah, A. J. 1996. Modern iron ooids from a shallow-marine volcanic setting: Mahengetang, Indonesia. Geology 24, 759–762.2.3.CO;2>CrossRefGoogle Scholar
Heller, P. L., & Paola, C. 1992. The large scale dynamics of grain-size variation in alluvial basins, 2. Application to syntectonic conglomerate. Basin Research 4, 91–102.CrossRefGoogle Scholar
Heller, P. L., & Paola, C. 1996. Downstream changes in alluvial architecture: an exploration of controls on channel-stacking patterns. Journal of Sedimentary Research B66, 297–306.Google Scholar
Hequette, A., & Hill, P. R. 1993. Storm-generated currents and offshore sediment transport on a sandy shoreface, Tibjak Beach, Canadian Beaufort Sea. Marine Geology 113, 283–304.CrossRefGoogle Scholar
Hesp, P., Hyde, R., Hesp, V., & Zengyu, Q. 1989. Longitudinal dunes can move sideways. Earth Surface Processes and Landforms 14, 447–451.CrossRefGoogle Scholar
Hesse, R. 1990a. Origin of chert: diagenesis of biogenic siliceous sediments. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 227–251. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Hesse, R.1990b. Silica diagenesis: origin of inorganic and replacement cherts. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 253–275. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Hills, J. G., & Mader, C. L. 1997. Tsunami produced by the impacts of small asteroids. Annals of New York Academy of Sciences 822, 381–394.CrossRefGoogle Scholar
Hindmarsh, R. C. A. 1998. Drumlinization and drumlin-forming instabilities: viscous till mechanisms. Journal of Glaciology 44, 293–314.CrossRefGoogle Scholar
Hindmarsh, R. C. A, Dunlop, P., Clark, C. D. 2003. Modeling the geomorphological effects of till redistribution; assessing a dynamic theory for Rogen moraine formation and drumlin formation. In 16th Congress of the International Union for Quaternary Research.Reno, NV: INQUA.Google Scholar
Hine, A. C. 1975. Bedform distribution and migration patterns on tidal deltas in the Chatham Harbor estuary, Cape Cod, Massachusetts. In Estuarine Research: Volume II, Geology and Engineering (ed. Cronin, L. E.), pp. 235–252. London: Academic Press.Google Scholar
Hine, A. C. 1977. Lily Bank, Bahamas: history of an active oolite sand shoal. Journal of Sedimentary Petrology 47, 1554–1581.Google Scholar
Hine, A. C., & Neumann, A. C. 1977. Shallow carbonate bank margin growth and structure, Little Bahama Bank, Bahamas. American Association of Petroleum Geologists Bulletin 61, 376–406.Google Scholar
Hinnov, L. 2000. New perspectives on orbitally forced stratigraphy. Annual Review of Earth and Planetary Sciences 28, 419–475.CrossRefGoogle Scholar
Hoffman, P. 1973. Recent and ancient algal stromatolites: seventy years of pedagogic cross-pollination. In Evolving Concepts in Sedimentology (ed. Ginsburg, R. N.), pp. 178–191. Baltimore, MA: The Johns Hopkins University Press.
Hoffman, P. 1974. Shallow and deepwater stromatolites in Lower Proterozoic platform-to-basin facies change, Great Slave Lake, Canada. American Association of Petroleum Geologists Bulletin 58, 856–867.Google Scholar
Hoffman, P. 1976. Stromatolite morphogenesis in Shark Bay, Western Australia. In Stromatolites (ed. Walter, M. R.), pp. 261–271. Amsterdam: Elsevier.CrossRef
Hoffman, P. F., Kaufman, A. J., Halverson, G. P., & Schrag, D. P. 1998. A Neoproterozoic snowball Earth. Science 281, 1342–1346.CrossRefGoogle ScholarPubMed
Hoffman, P. F., & Schrag, D. P. 2002. The snowball Earth hypothesis: testing the limits of global change. Terra Nova 14, 129–155.CrossRefGoogle Scholar
Hogg, A. J., & Huppert, H. E. 2001. Two-dimensional and axisymmetric models for compositional and particle-driven gravity currents in uniform ambient flows. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 121–134. Oxford: Blackwells.CrossRefGoogle Scholar
Holland, H. D. 1972. The geologic history of seawater – an attempt to solve the problem. Geochimica et Cosmochimica Acta 36, 637–651.CrossRefGoogle Scholar
Holland, H. D.1994. Early Proterozoic atmospheric change. In Early Life on Earth, Nobel Symposium No. 84 (ed. Bengston, S.), pp. 237–244. New York: Columbia University Press.Google Scholar
Holland, H. D. 2005. Sea level, sediments and the composition of sea water. American Journal of Science 305, 220–239.CrossRefGoogle Scholar
Hollister, C. D. 1993. The concept of deep-sea contourites. Sedimentary Geology 82, 5–15.CrossRefGoogle Scholar
Hollister, C. D., & McCave, I. N. 1984. Sedimentation under deep-sea storms. Nature 309, 220–225.CrossRefGoogle Scholar
Holman, R. A. 1983. Edge waves and the configuration of the shoreline. In CRC Handbook of Coastal Processes and Erosion (ed. Komar, P. D.), pp. 21–34. Boca Raton, FL: CRC Press.Google Scholar
Holman, R. A., & Sallenger, A. H. 1993. Sand bar generation: a discussion of the Duck experimental series. Journal of Coastal Research, Special Issue 15, 76–92.Google Scholar
Holmlund, P. 1988. Internal geometry and evolution of moulins, Storglaciaren, Sweden. Journal of Glaciology 34, 242–248.CrossRefGoogle Scholar
Homewood, P., & Allen, P. A. 1981. Wave-, tide-, and current-controlled sandbodies of Miocene molasse, western Switzerland. American Association of Petroleum Geologists Bulletin 65, 2534–2545.Google Scholar
Honji, H., Kaneko, A., & Matsunaga, N. 1980. Flows above oscillatory ripples. Sedimentology 27, 225–229.CrossRefGoogle Scholar
Hooke, R. LeB. 1998. Principles of Glacier Mechanics. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Hooke, R. L., & Jennings, C. E. 2006. On the formation of the tunnel valleys of the southern Laurentide ice sheet. Quaternary Science Reviews 25, 1364–1372.CrossRefGoogle Scholar
Horbury, A. D., & Robinson, A. G. 1993. Diagenesis and Basin Development. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Hori, K., Saito, Y., Zhang, Q., & Wang, P. 2002. Architecture and evolution of the tide-dominated Changjiang (Yangtze) River delta, China. Sedimentary Geology 146, 249–264.CrossRefGoogle Scholar
Horita, J., Zimmermann, H., & Holland, H. D. 2002. Chemical evolution of seawater during the Phanerozoic: implications from the record of marine evaporites. Geochimica et Cosmochimica Acta 66, 3733–3756.CrossRefGoogle Scholar
Horowitz, D. H. 1982. Geometry and origin of large-scale deformation structures in some ancient windblown sand deposits. Sedimentology 29, 155–180.CrossRefGoogle Scholar
Horton, R. E. 1945. Erosional development of streams and their drainage basins: hydrophysical approach to quantitative morphology. Geological Society of America Bulletin 56, 275–370.CrossRefGoogle Scholar
House, M. R., & Gale, A. S. 1995. Orbital Forcing Timescales and Cyclostratigraphy. London: Geological Society of London.Google Scholar
Howard, A. D. 1967. Drainage analysis in geologic interpretation: a summary. American Association of Petroleum Geologists Bulletin 51, 2246–2259.Google Scholar
Howard, A. D. 1987. Modeling fluvial systems: rock-, gravel- and sand-bed channels. In River Channels: Environment and Process (ed. Richards, K. S.), pp. 69–94. Oxford: Blackwells.Google Scholar
Howard, J. D., & Frey, R. W. 1975. Estuaries of the Georgia coast, U.S.A.: sedimentology and biology, II. Regional animal-sediment characteristics of Georgia Estuaries. Senckenbergiana Maritima 7, 33–103.Google Scholar
Howard, J. D., & Reineck, H.-E. 1972. Georgia coastal region, Sapelo Island, U.S.A.: sedimentology and biology, IV. Physical and biogenic sedimentary structures of the nearshore shelf. Senckenbergiana Maritima 4, 47–79.Google Scholar
Howard, J. D., & Reineck, H.-E. 1981. Depositional facies of high-energy beach-to-offshore sequence, comparison with low energy sequence. American Association of Petroleum Geologists Bulletin 65, 807–830.Google Scholar
Hower, J., Eslinger, E. V., Hower, M. E., & Perry, E. A. 1976. Mechanism of burial metamorphism of argillaceous sediment: 1. Mineralogical and chemical evidence. Geological Society of America Bulletin 87, 725–737.2.0.CO;2>CrossRefGoogle Scholar
Hoyt, J. H., & Henry, V. J. 1967. Influence of island migration on barrier island sedimentation. Geological Society of America Bulletin 78, 77–86.CrossRefGoogle Scholar
Hoyt, J. H., & Henry, V. J. 1971. Origin of capes and shoals along the southeastern coast of the United States. Geological Society of America Bulletin 82, 59–66.CrossRefGoogle Scholar
Hsü, K. J. 1972a. Origin of saline giants: a critical review after the discovery of the Mediterranean evaporite. Earth-Science Reviews 8, 371–396.CrossRefGoogle Scholar
Hsü, K. J. 1972b. When the Mediterranean dried up. Scientific American 227, 26–36.CrossRefGoogle Scholar
Hsü, K. J., & Jenkyns, H. C. 1974. Pelagic Sediments: On Land and Under the Sea. Oxford: Blackwells.Google Scholar
Hubbard, D. K., Oertel, G., & Nummedal, D. 1979. The role of waves and tidal currents in the development of tidal-inlet sedimentary structures and sand body geometry: examples from North Carolina, South Carolina and Georgia. Journal of Sedimentary Petrology 49, 1073–1092.Google Scholar
Hulbe, C. L., MacAyeal, D. R., Denton, G. H., Kleman, J., & Lowell, T. V. 2004. Catastrophic ice shelf breakup as the source of Heinrich event icebergs. Paleoceanography19, PA1004, doi: 10.1029/2003PA000890.CrossRef
Hulscher, S. 1996. Tide-induced large-scale regular bedform patterns in three-dimensional shallow water model. Journal of Geophysical Research 101, 20,727–20,744.CrossRefGoogle Scholar
Hunt, D., & Gawthorpe, R. L. 2000. Sedimentary Responses to Forced Regressions. London: Geological Society of London.Google Scholar
Hunter, R. E. 1977. Basic types of stratification in small eolian dunes. Sedimentology 24, 361–387.CrossRefGoogle Scholar
Hunter, R. E. 1985. Subaqueous sand flow cross strata. Journal of Sedimentary Petrology 55, 886–894.Google Scholar
Hunter, R. E., & Clifton, H. E. 1982. Cyclic deposits and hummocky cross-stratification of probable storm origin in the Upper Cretaceous rocks of the Cape Sebastian area, south-western Oregon. Journal of Sedimentary Petrology 52, 127–143.Google Scholar
Hunter, R. E., Clifton, H. E., & Phillips, R. L. 1979. Depositional processes, sedimentary structures and predicted vertical sequences in barred nearshore systems, southern Oregon coast. Journal of Sedimentary Petrology 49, 711–726.Google Scholar
Hunter, R. E., & Kocurek, G. 1986. An experimental study of subaqueous slipface deposition. Journal of Sedimentary Petrology 56, 387–394.CrossRefGoogle Scholar
Huntley, D. A., & Bowen, A. J. 1973. Field observations of edge waves. Nature 243, 160–161.CrossRefGoogle Scholar
Huntley, D. A., Guza, R. T., & Thornton, D. B. 1981. Field observations of surf beat, 1: progressive edge waves. Journal of Geophysical Research 83, 1913–1920.Google Scholar
Huppert, H. E. 1998. Quantitative modeling of granular suspension flows. Philosophical Transactions of the Royal Society of London 356A, 2471–2496.Google Scholar
Hutcheon, I. E. 1990. Aspects of diagenesis in coarse-grained siliciclastic rocks. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 165–176. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Huthnance, J. M. 1982a. On one mechanism forming linear sand banks. Estuarine, Coastal and Shelf Science 14, 79–99.CrossRefGoogle Scholar
Huthnance, J. M. 1982b. On the formation of sand banks of finite extent. Estuarine, Coastal and Shelf Science 15, 277–299.CrossRefGoogle Scholar
Hyndman, D., & Hyndman, D. 2006. Natural Hazards and Disasters. Belmont, CA: Thomson Brooks/Cole.Google Scholar
Iida, K., & Iwasaki, T. 1983. Tsunamis: Their Science and Engineering. Dordrecht: Reidel.CrossRefGoogle Scholar
Ikeda, S., & Parker, G. 1989. River Meandering. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Illing, L. V., Wells, A. J., & Taylor, J. C. M. 1965. Penecontemporary dolomite in the Persian Gulf. In Dolomitization and Limestone Diagenesis, A Symposium (ed. Pray, L. C. & Murray, R. C.), pp. 89–111. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Imboden, D. M., & Wüest, A. 1995. Mixing mechanisms in lakes. In Physics and Chemistry of Lakes, 2nd edn. (ed. Lerman, A., Imboden, D., & Gat, J.), pp. 83–138. New York: Springer-Verlag.CrossRefGoogle Scholar
Imbrie, J., Hays, J. D., Martinson, D. G. et al. 1984. The orbital theory of Pleistocene climate: support from a revised chronology of the marine δ18O record. In Milankovitch and Climate: Understanding the Response to Astronomical Forcing (ed. Berger, A., Imbrie, J., Hays, J., & Kukla, G.), pp. 269–305. Dordrecht: Reidel.CrossRefGoogle Scholar
Imbrie, J., & Imbrie, K. P. 1979. Ice Ages: Solving the Mystery. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Imbrie, J., Mix, A. C., & Martinson, D. G. 1993. Milankovitch theory viewed from Devils Hole. Nature 363, 531–533.CrossRefGoogle Scholar
Imperato, D. P., Sexton, W. J., & Hayes, M. O. 1988. Stratigraphy and sediment characteristics of a mesotidal ebb-tidal delta, North Edisto Inlet, South Carolina. Journal of Sedimentary Petrology 58, 950–958.Google Scholar
Immenhauser, A., Hillgartner, H., & Benthum, E. 2005. Microbial–foraminiferal episodes in the Early Aptian of the southern Tethyan margin: ecological significance and possible relation to oceanic anoxia. Sedimentology 52, 77–99.CrossRefGoogle Scholar
Inden, R. F., & Moore, C. H. 1983. Beach. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 211–265. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Iverson, N. R. 1991. Potential effects of subglacial water-pressure fluctuations on quarrying. Journal of Glaciology 37, 27–36.CrossRefGoogle Scholar
Iverson, R. M. 1997. The physics of debris flows. Reviews of Geophysics 35, 245–296.CrossRefGoogle Scholar
Jackson, R. G. 1975. Hierarchical attributes and a unifying model of bedforms composed of cohesionless sediment and produced by shearing flow. Geological Society of America Bulletin 86, 1,523–1,533.2.0.CO;2>CrossRefGoogle Scholar
Jackson, R. G. 1976. Sedimentological and fluid-dynamic implications of the turbulent bursting phenomenon in geophysical flows. Journal of Fluid Mechanics 77, 531–560.CrossRefGoogle Scholar
Jaeger, H. M., Nagel, S. R., & Behringer, R. P. 1996. The physics of granular materials. Physics Today 49, 32–36.CrossRefGoogle Scholar
Jahren, A. H. 2004. Factors of soil formation: biota. In Encyclopedia of Soils in the Environment (ed. Hillel, D., Rosenzweig, C., Powlson, D.et al.), pp. 507–512. New York: Academic Press.Google Scholar
James, N. P., & Choquette, P. W. 1983. Diagenesis 6. Limestones – the sea floor diagenetic environment. Geoscience Canada 11, 161–194.Google Scholar
James, N. P. & MacIntyre, I. G. 1985. Carbonate depositional environments, modern and ancient. Part 1 – reefs, zonation, depositional facies and diagenesis. Quarterly Journal of the Colorado School of Mines 80, 70pp.Google Scholar
James, H. L. 1954. Sedimentary facies of iron formations. Economic Geology 49, 235–293.CrossRefGoogle Scholar
James, N. P. 1981. Megablocks of calcified algae in the Cow Head Breccia, western Newfoundland: vestiges of a Cambro-Ordovician margin. Geological Society of America Bulletin 42, 799–811.2.0.CO;2>CrossRefGoogle Scholar
James, N. P.1983. Reefs. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 345–462. Tulsa, OK: American Association of Petroleum.Google Scholar
James, N. P., Bone, Y., Collins, L. B., & Kyser, T. K. 2001. Surficial sediments of the Great Australian Bight: facies dynamics and oceanography on a vast cool-water carbonate shelf. Journal of Sedimentary Research 71, 5549–567.CrossRefGoogle Scholar
James, N. P., & Bourque, P. A. 1992. Reefs and mounds. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 323–345. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Choquette, P. W. 1988. Paleokarst. New York: Springer-Verlag.CrossRefGoogle Scholar
James, N. P., & Choquette, P. W.1990a. Limestones – introduction. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 9–12. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Choquette, P. W.1990b. Limestones – the sea floor diagenetic environment. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 13–34. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Choquette, P. W.1990c. Limestones – the meteoric diagenetic environment. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 35–74. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
James, N. P., & Clarke, A. D. 1997. Cool Water Carbonates. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
James, N. P., & Ginsburg, R. N. 1979. The Seaward Margin of the Belize Reef. Oxford: Blackwells.Google Scholar
James, N. P., & Kendall, A. C. 1992. Introduction to carbonate and evaporite facies models. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 265–275. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Jenkins, J. G., & Savage, S. B. 1983. A theory for the rapid flow of identical, smooth, nearly elastic spherical particles. Journal of Fluid Mechanics 130, 187–202.CrossRefGoogle Scholar
Jenkyns, H. C., Forster, A., Schouten, S., & Damste, J. S. S. 2004. High temperatures in the Late Cretaceous Arctic Ocean. Nature 432, 888–892.CrossRefGoogle ScholarPubMed
Johnson, A. M. 1970. Physical Processes in Geology. San Francisco: W. H. Freeman.Google Scholar
Johnson, D. D., & Beaumont, C. 1995. Preliminary results from a planform kinematic model of orogen evolution, surface processes and the development of clastic foreland basin stratigraphy. In Stratigraphic Evolution of Foreland Basins (ed. Dorobek, S. L. & Ross, G. M.), pp. 3–24. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Johnson, M. A., Kenyon, N. H., Belderson, R. H., & Stride, A. H. 1982. Sand transport. In Offshore Tidal Sands (ed. Stride, A. H.), pp. 58–94. London: Chapman and Hall.CrossRefGoogle Scholar
Johnson, N. M., & Driess, S. J. 1989. Hydrostratigraphic interpretation using indicator geostatistics. Water Resources Research 25, 2501–2510.CrossRefGoogle Scholar
Jones, B., & Desrochers, A. 1992. Shallow platform carbonates. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 277–301. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Jones, B., & Luth, R. W. 2002. Dolostones from Grand Cayman, British West Indes. Journal of Sedimentary Research 72, 559–569.CrossRefGoogle Scholar
Jones, L. S., & Schumm, S. A. 1999. Causes of avulsion: an overview. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 171–178. Oxford: Blackwells.CrossRefGoogle Scholar
Jopling, A. V., & McDonald, B. C. 1975. Glaciofluvial and Glaciolacustrine Sediments. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Jordan, T. E. 1981. Thrust loads and foreland basin evolution, Cretaceous western United States. American Association of Petroleum Geologists Bulletin 65, 2506–2620.Google Scholar
Jordan, T. E., & Flemings, P. B. 1990. From geodynamical models to basin fill – a stratigraphic perspective. In Quantitative Dynamic Stratigraphy (ed. Cross, T. A.), pp. 149–163. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Jordan, T. E., & Flemings, P. B. 1991. Large-scale stratigraphic architecture, eustatic variation, and unsteady tectonism: a theoretical evaluation. Journal of Geophysical Research 96, 6681–6699.CrossRefGoogle Scholar
Jorgenson, B. B. 2000. Bacteria and marine biogeochemistry. In Marine Geochemistry (ed. Schulz, H. D. & Zabel, M.), pp. 173–207. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Julien, P. Y. 2002. River Mechanics. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jullien, R., Meakin, P., & Pavlovitch, A. 2002. Three-dimensional model for particle-size segregation by shaking. Physical Review Letters 69, 640–643.CrossRefGoogle Scholar
Kaiser, J. 2004. Wounding Earth's fragile skin. Science 304, 1616–1618.CrossRefGoogle ScholarPubMed
Kamb, B. 1987. Glacier surge mechanism based on linked cavity configuration of the basal water conduit system. Journal of Geophysical Research 92, 9083–9100.CrossRefGoogle Scholar
Kamb, B. 1991. Rheological nonlinearity and flow instability in the deforming bed mechanism of ice stream motion. Journal of Geophysical Research 96B, 16,585–16,595.CrossRefGoogle Scholar
Kamb, B., Raymond, C. F., Harrison, W. D.et al. 1985. Glacier surge mechanism: 1982–1983 surge of Variegated Glacier, Alaska. Science 227, 469–479.CrossRefGoogle ScholarPubMed
Kanfoush, S. L., Hodell, D. A., Charles, C. D.et al. 2000. Millennial-scale instability of the Antarctic Ice sheet during the last glaciation. Science 288, 1815–1819.CrossRefGoogle ScholarPubMed
Karssenberg, D., Törnqvist, T. E., & Bridge, J. S. 2001. Conditioning a process-based model of sedimentary architecture to well data. Journal of Sedimentary Research 71, 868–879.CrossRefGoogle Scholar
Karssenberg, D., Dalman, R., Weltje, G. J., Postma, G., & Bridge, J. S. 2004. Numerical modelling of delta evolution by nesting high and low resolution process-based models of sedimentary basin filling. In Joint EURODELTA/EUROSTRATAFORM Meeting, Venice (Italy), pp. 1–37.
Katz, M. E., Finkel, Z. V., Grzebyk, D., Knoll, A. H., & Falkowski, P. G. 2004. Evolutionary trajectories and biogeochemical impacts of marine eukaryotic phytoplankton. Annual Review of Ecology, Evolution, and Systematics 35, 523–556.CrossRefGoogle Scholar
Keen, T. R., & Glenn, S. M. 1994. A coupled hydrodynamic–bottom boundary layer model of Ekman flow on stratified continental shelves. Journal of Physical Oceanography 24, 1732–1749.2.0.CO;2>CrossRefGoogle Scholar
Kelly, D. S., Karson, J. A., Früh-Green, G. L.et al. 2005. A serpentine-hosted ecosystem: the Lost City hydrothermal field. Science 307, 1428–1434.CrossRefGoogle Scholar
Kemp, P. H., & Simons, R. R. 1982. The interaction between waves and a turbulent current: waves propagating with the current. Journal of Fluid Mechanics 116, 227–250.CrossRefGoogle Scholar
Kendall, A. C. 1992. Evaporites. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 375–409. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Kendall, A. C., & Harwood, G. M. 1996. Marine evaporites: arid shorelines and basins. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 281–324. Oxford: Blackwell Science.Google Scholar
Kendall, C. G. St. C., & Skipwith, P. A. D'E. 1968. Recent algal mats of a Persian Gulf lagoon. Journal of Sedimentary Petrology 38, 1040–1058.Google Scholar
Kendall, C. G. St. C., & Skipwith, P. A. D'E. 1969. Geomorphology of a Recent shallow-water carbonate province: Khor al Bazam, Trucial Coast, southwestern Persian Gulf. Geological Society of America Bulletin 80, 865–892.CrossRefGoogle Scholar
Kendall, C. G. St. C., & Warren, J. 1987. A review of the origin and setting of tepees and their associated fabrics. Sedimentology 34, 1007–1027.CrossRefGoogle Scholar
Kennard, J. M., & James, N. P. 1986. Thrombolites and stromatolites: two distinct types of microbial structures. Palios 1, 492–503.CrossRefGoogle Scholar
Kennedy, J. F. 1963. The mechanics of dunes and antidunes in erodible-bed channels. Journal of Fluid Mechanics 16, 521–544.CrossRefGoogle Scholar
Kenyon, N. H., Amir, A., & Cramp, A. 1995. Geometry of the younger sediment bodies on the Indus Fan. In Atlas of Deep Water Environments: Architectural Style in Turbidite Systems (ed. Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, F. Ricci, & Smith, R. D. A.), pp. 89–93. London: Chapman and Hall.CrossRefGoogle Scholar
Kiessling, W., Flügel, E., & Golonka, J. 2002. Phanerozoic Reef Patterns. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Kim, J., Dong, H., Seabaugh, J., Newell, S. W., & Eberl, D. D. 2004. Role of microbes in the smectite-to-illite reaction. Science 303, 830–832.CrossRefGoogle ScholarPubMed
King, L. H. 1993. Till in the marine environment. Journal of Quaternary Science 8, 347–358.CrossRefGoogle Scholar
Kinsman, D. J. 1966. Gypsum and anhydrite of Recent age, Trucial Coast, Persian Gulf. In Proceedings of the Second Salt Symposium (ed. Rau, J. L.), Vol. 1, pp. 302–326. Cleveland, OH: Northern Ohio Geological Society.Google Scholar
Kinsman, D. J. J., & Park, R. K. 1976. Algal belt and coastal sabkha evolution, Trucial Coast, Persian Gulf. In Stromatolites (ed. Walter, M. R.), pp. 421–433. Amsterdam: Elsevier.Google Scholar
Kirkby, M. J. 1994. Thresholds and instability in stream head hollows: a model of magnitude and frequency for wash processes. In Process Models and Theoretical Geomorphology (ed. Kirkby, M. J.), pp. 294–314. Chichester: Wiley.Google Scholar
Kirschvink, J. L. 1992. Late Proterozoic low latitude glaciation: the snowball Earth. In The Proterozoic Biosphere (ed. Schopf, J. W. & Klein, C.), pp. 51–52. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Klappa, C. F. 1978. Biolithogenesis of Microcodium; elucidation. Sedimentology 25, 489–522.CrossRefGoogle Scholar
Klein, C., & Beukes, N. J. 1992. Proterozoic iron formations. In Proterozoic Crustal Evolution (ed. Condie, K. C.), pp. 383–418. Amsterdam: Elsevier.Google Scholar
Klein, G. deV. 1970. Depositional and dispersal dynamics of intertidal sand bars. Journal of Sedimentary Petrology 40, 1095–1127.Google Scholar
Kleinhans, M. G. 2004. Sorting in grains flows at the lee side of dunes. Earth-Science Reviews 65, 75–102.CrossRefGoogle Scholar
Kleinhans, M. G. 2005. Grain-size sorting in grainflows at the lee side of deltas. Sedimentology 52, 291–311.CrossRefGoogle Scholar
Kleypas, J. A., Buddemeier, R. W., Archer, D.et al. 1999. Geochemical consequences of increased atmospheric carbon dioxide on coral reefs. Science 284, 118–120.CrossRefGoogle ScholarPubMed
Knauth, L. P. 2003. Siliceous sediments. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 660–666. Dordrecht: Kluwer Academic Publishers.Google Scholar
Kneller, B. 1996. Beyond the turbidite paradigm: physical models for deposition of turbidites and their implication for reservoir prediction. In Characterisation of Deep Marine Clastic Systems (ed. Hartley, A. & Prosser, D. J.), pp. 29–46. London: Geological Society of London.Google Scholar
Kneller, B. C., & Branney, M. J. 1995. Sustained high density turbidity currents and the deposition of massive sands. Sedimentology 42, 607–616.CrossRefGoogle Scholar
Kneller, B. C., & Buckee, C. 2000. The structure and fluid mechanics of turbidity currents: a review of some recent studies and their geological implications. Sedimentology 47 (Supplement 1), 62–94.CrossRefGoogle Scholar
Kneller, B. C., Edwards, D., McCaffrey, W., & Moore, R. 1991. Oblique reflection of turbidity currents. Geology 19, 250–252.2.3.CO;2>CrossRefGoogle Scholar
Knight, J. B., Jaeger, H. M., & Nagel, S. R. 1993. Vibration-induced size separation in granular media: the convection connection. Physical Review Letters 70, 3728–3731.CrossRefGoogle ScholarPubMed
Knight, P. G. 2006. Glacier Science and Environmental Change. Oxford: Blackwells.CrossRefGoogle Scholar
Knight, R. J., & McLean, J. R. 1986. Shelf Sands and Sandstones. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Knighton, A. D. 1998. Fluvial Forms and Processes: A New Perspective. London: Arnold.Google Scholar
Knoll, A. H., & Carroll, S. B. 1999. Early animal evolution: emerging views from comparative biology and geology. Science 284, 2129–2137.CrossRefGoogle ScholarPubMed
Knox, J. C. 1983. Responses of river systems to Holocene climates. In Late Quaternary Environments of the United States, Volume 2, The Holocene (ed. Wright, H. E. & Porter, S. C.), pp. 26–41. Minneapolis, MN: University of Minnesota Press.Google Scholar
Knox, J. C. 1996. Fluvial systems since 20,000 years BP. In Continental Palaeohydrology (ed. Gregory, K. J., Starkel, L., & Baker, V. R.), pp. 87–108. Chichester: Wiley.Google Scholar
Kochel, R. C., & Baker, V. R. 1988. Palaeoflood analysis using slack water deposits. In Flood Geomorphology (ed. Baker, V. R., Kochel, R. C., & Patton, P. P.), pp. 357–376. New York: Wiley.Google Scholar
Kocurek, G. 1981. Significance of interdune deposits and bounding surfaces in eolian dune sands. Sedimentology 28, 753–780.CrossRefGoogle Scholar
Kocurek, G. 1988. First order and super bounding surfaces in eolian sequences – bounding surfaces revisited. Sedimentary Geology 56, 193–206.CrossRefGoogle Scholar
Kocurek, G. 1991. Interpretation of ancient eolian sand dunes. Annual Review of Earth and Planetary Sciences 19, 43–75.CrossRefGoogle Scholar
Kocurek, G.1996. Desert aeolian systems. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 125–155. Oxford: Blackwells.Google Scholar
Kocurek, G., & Havholm, K. G. 1993. Eolian sequence stratigraphy – a conceptual framework. In Recent Advances in and Application of Siliciclastic Sequence Stratigraphy (ed. Weimer, P. & Posamentier, H.), pp. 393–409. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Kocurek, G., & Lancaster, N. 1999. Aeolian sediment states: theory and Mojave Kelso Dunefield example. Sedimentology 46, 505–516.CrossRefGoogle Scholar
Kocurek, G., & Nielsen, J. 1986. Conditions favourable for the formation of warm-climate aeolian sand sheets. Sedimentology 33, 795–816.CrossRefGoogle Scholar
Kocurek, G., Townsley, M., Yeh, E., Havholm, K., & Sweet, M. L. 1992. Dune and dune-field development on Padre Island, Texas, with implications for interdune deposition and water-table controlled accumulation. Journal of Sedimentary Petrology 62, 622–635.Google Scholar
Kohout, F. A., Henry, H. R., & Banks, J. E. 1977. Hydrogeology related to geothermal conditions of the Florida Plateau. In The Geothermal Nature of the Florida Plateau (ed. Smith, D. L. & Griffin, G. E.), pp. 1–41. Tallahassee, FL: Florida Bureau of Geology.Google Scholar
Kolb, C. R., & Van Lopik, J. R. 1966. Depositional environments of the Mississippi River deltaic plain – southeastern Louisiana. In Deltas in Their Geologic Framework (ed. Shirley, M. L. & Ragsdale, J. A.), pp. 17–61. Houston, TX: Houston Geological Society.Google Scholar
Kolla, V., & Coumes, F. 1985. Indus fan, Indian Ocean. In Submarine Fans and Related Turbidite Systems (ed. Bouma, A. H., Normark, W. R., & Barnes, N. E.), pp. 129–136. New York: Springer-Verlag.CrossRefGoogle Scholar
Kolla, V., & Coumes, F. 1987. Morphology, internal structure, seismic stratigraphy, and sedimentation of Indus fan. American Association of Petroleum Geologists Bulletin 71, 650–677.Google Scholar
Kolla, V., & Perlmutter, M. A. 1993. Timing of turbidite sedimentation on the Mississippi Fan. American Association of Petroleum Geologists Bulletin 77, 1129–1141.Google Scholar
Koltermann, C. E., & Gorelick, S. M. 1992. Paleoclimatic signature in terrestrial flood deposits. Science 256, 1775–1782.CrossRefGoogle ScholarPubMed
Koltermann, C. E., & Gorelick, S. M. 1996. Heterogeneity in sedimentary deposits: a review of structure-imitating, process-imitating and descriptive approaches. Water Resources Research 32, 2617–2658.CrossRefGoogle Scholar
Komar, P. D. 1973. Computer models of delta growth due to sediment input from rivers and longshore transport. Geological Society of America Bulletin 84, 2,217–2,226.2.0.CO;2>CrossRefGoogle Scholar
Komar, P. D. 1974. Oscillatory ripple marks and the evaluation of ancient wave conditions and environments. Journal of Sedimentary Petrology 44, 169–180.Google Scholar
Komar, P. D. 1976. Beach Processes and Sedimentation. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Komar, P. D. 1977. Modeling of sand transport on beaches and the resulting shoreline evolution. In The Sea (ed. Goldberg, E.et al.) Vol. 6, pp. 499–513. New York: Wiley.Google Scholar
Komar, P. D.1996. Entrainment of sediments from deposits of mixed grain sizes and densities. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 127–181. Chichester: Wiley.Google Scholar
Komar, P. D. 1998. Beach Processes and Sedimentation, 2nd edn. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Komar, P. D., & Miller, M. C. 1973. The threshold of sediment motion under oscillatory water waves. Journal of Sedimentary Petrology 43, 1101–1110.Google Scholar
Komar, P. D., & Miller, M. C. 1975a. On the comparison between the threshold of sediment motion under waves and unidirectional currents with a discussion of the practical evaluation of the threshold. Journal of Sedimentary Petrology 45, 362–367.Google Scholar
Komar, P. D., & Miller, M. C. 1975b. The initiation of oscillatory ripple marks and the development of plane-bed at high shear stresses under waves. Journal of Sedimentary Petrology 45, 697–703.Google Scholar
Kooi, H., & Beaumont, C. 1994. Escarpment evolution on high-elevation rifted margins: insights derived from a surface processes model that combines diffusion, advection, and reaction. Journal of Geophysical Research 99, 12191–12209.CrossRefGoogle Scholar
Kooi, H., & Beaumont, C. 1996. Large-scale geomorphology: classical concepts reconciled and integrated with contemporary ideas using a surface processes model. Journal of Geophysical Research 101, 3361–3386.CrossRefGoogle Scholar
Kor, P. S. G., Shaw, J., & Sharpe, D. R. 1991. Erosion of bedrock by subglacial meltwater, Georgian Bay, Ontario: a regional view. Canadian Journal of Earth Sciences 28, 623–642.CrossRefGoogle Scholar
Kosters, E. C. 1989. Organic–clastic facies relationships and chronostratigraphy of the Barataria interlobe basin, Mississippi delta plain. Journal of Sedimentary Petrology 59, 98–113.Google Scholar
Kosters, E. C., & Suter, J. R. 1993. Facies relationships and systems tracts in the late Holocene Mississippi delta plain. Journal of Sedimentary Petrology 63, 727–733.Google Scholar
Kraft, J. C. 1971. Sedimentary facies patterns and geologic history of a Holocene marine transgression. Geological Society of America Bulletin 82, 2131–2158.CrossRefGoogle Scholar
Kraft, J. C., Chrzastowski, M. J., Belknap, D. F., Toscano, M. A., & Fletcher, C. H. 1987. The transgressive barrier–lagoon coast of Delaware: morphostratigraphy, sedimentary sequences and responses to relative rises in sea level. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 129–144. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Kraus, M. J. 1987. Integration of channel and floodplain suites, II. Vertical relations of alluvial paleosols. Journal of Sedimentary Petrology 56, 602–612.Google Scholar
Kraus, M. J. 1996. Avulsion deposits in lower Eocene alluvial rocks, Bighorn Basin, Wyoming. Journal of Sedimentary Research 66B, 354–363.Google Scholar
Kraus, M. J. 1999. Paleosols in clastic sedimentary rocks: their geologic applications. Earth-Science Reviews 47, 41–70.CrossRefGoogle Scholar
Kraus, M. J., & Aslan, A. 1999. Palaeosol sequences in floodplain environments: a hierarchical approach. In Palaeoweathering, Palaeosurfaces and Related Continental Deposits (ed. Thiry, M. & Simon-Coicon, R.), pp. 303–321. Oxford: Blackwells.Google Scholar
Kraus, M. J., & Bown, T. M. 1986. Palaeosols and time resolution in alluvial stratigraphy. In Paleosols, Their Recognition and Interpretation (ed. Wright, V. P.), pp. 180–207. Princeton, NJ: Princeton University Press.Google Scholar
Kraus, M. J., & Gwinn, B. M. 1997. Controls on the development of early Eocene avulsion deposits and floodplain paleosols, Willwood Formation, Bighorn Basin. Sedimentary Geology 114, 33–54.CrossRefGoogle Scholar
Kraus, M. J., & Hasiotis, S. T. 2006. Significance of different modes of rhizolith preservation to interpreting paleoenvironmental and paleohydologic settings: examples from Paleogene paleosols, Bighorn Basin, Wyoming, U.S.A. Journal of Sedimentary Research 76, 633–646.CrossRefGoogle Scholar
Kraus, M. J., & Wells, T. M. 1999. Recognizing avulsion deposits in the ancient stratigraphic record. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 251–268. Oxford: Blackwells.CrossRefGoogle Scholar
Krause, D. C., White, W. C., Piper, D. J. W., & Heezen, B. C. 1970. Turbidity currents and cable breaks in the western New Britain Trench. Geological Society of America Bulletin 81, 2153–2160.CrossRefGoogle Scholar
Krynine, P. D. 1942. Differential sedimentation and its products during one complete geosynclinal cycle. Proceedings of the 1st Pan American Congress on Mining Engineering Geology, Part 1, Vol. 2, pp. 537–560. Santiago: Chilean Institute of Mining Engineering.Google Scholar
Krynine, P. D. 1950. Petrology, stratigraphy and origin of the Triassic sedimentary rocks of Connecticut. Connecticut State Geological and Natural History Survey Bulletin 73, 1–247.Google Scholar
Kuenen, P. H. 1965. Value of experiments in geology. Geologie en Mijnbouw 44, 22–36.Google Scholar
Kuenen, P. H., & Migliorini, C. I. 1950. Turbidity currents as a cause of graded bedding. Journal of Geology 58, 91–127.CrossRefGoogle Scholar
Kuijper, C., Cornelisse, J. M., & Winterwerp, J. C. 1989. Research on erosive properties of cohesive sediments. Journal of Geophysical Research 95, 14,341–14,350.CrossRefGoogle Scholar
Kumar, N., & Sanders, J. E. 1974. Inlet sequences: a vertical succession of sedimentary structures and textures created by the lateral migration of tidal inlets. Sedimentology 21, 491–532.CrossRefGoogle Scholar
Kummel, B., & Raup, D. 1965. Handbook of Paleontological Techniques. San Francisco, CA: W. H. Freeman.Google Scholar
Kump, L. R., Brantley, S. L., & Arthur, M. A. 2000. Chemical weathering, atmospheric CO2, and climate. Annual Review of Earth and Planetary Sciences 28, 611–667.CrossRefGoogle Scholar
Kump, L. R., Kasting, J. F., & Crane, R. G. 2004. The Earth System, 2nd edn. Upper Saddle River, NJ: Prentice-Hall.Google Scholar
Lachenbruch, A. H. 1962. Mechanics of thermal contraction cracks and ice-wedge polygons. Geological Society of America Special Paper 70.
Lal, R. 2004. Soil carbon sequestration impacts on global climate change and food security. Science 304, 1623–1627.CrossRefGoogle ScholarPubMed
Lancaster, N. 1982. Linear dunes. Progress in Physical Geography 6, 475–504.CrossRefGoogle Scholar
Lancaster, N. 1989. Star dunes. Progress in Physical Geography 13, 67–91.CrossRefGoogle Scholar
Lancaster, N. 1995. Geomorphology of Desert Dunes. London: Routledge.CrossRefGoogle Scholar
Lancaster, N.2005. Aeolian processes. In Encyclopedia of Geology 4 (ed. Selley, R. C., Cocks, R. L. M., & Plimer, I. R.), pp. 612–627. Oxford: Elsevier.Google Scholar
Lancaster, N.2007. Low-latitude dune fields. In Encyclopedia of Quaternary Sciences (ed. Elias, S. A.). Amsterdam: Elsevier.Google Scholar
Lancaster, N., Kocurek, G., Singhvi, A.et al. 2002. Late Pleistocene and Holocene dune activity and wind regimes in the western Sahara Desert of Mauritania. Geology 30, 991–994.2.0.CO;2>CrossRefGoogle Scholar
Land, L. S., Behrens, E. W., & Frishman, S. A. 1979. The ooids of Baffin Bay, Texas. Journal of Sedimentary Petrology 49, 1269–1278.Google Scholar
Langbein, W. B. 1964. Geometry of river channels. Journal of the Hydraulics Division, ASCE 90, 301–312.Google Scholar
Langbein, W. B., & Leopold, L. B. 1964. Quasi-equilibrium states in channel morphology. American Journal of Science 262, 782–794.CrossRefGoogle Scholar
Langbein, W. B., & Leopold, L. B. 1966. River meanders – theory of minimum variance. United States Geological Survey Professional Paper 422H.
Langford, R. P., & Chan, M. A. 1988. Flood surfaces and deflation surfaces within the Cutler Formation and Cedar Mesa Sandstone (Permian), southeastern Utah. Geological Society of America Bulletin 100, 1541–1549.2.3.CO;2>CrossRefGoogle Scholar
Langford, R. P., & Chan, M. A.1993. Downwind changes within an ancient dune sea, Permian Cedar Mesa Sandstone, southeast Utah. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 109–126. Oxford: Blackwells.CrossRefGoogle Scholar
Larson, G. J., Lawson, D. E., Evenson, E. B.et al. 2006. Glaciohydraulic supercooling in former ice sheets?Geomorphology 75, 20–32.CrossRefGoogle Scholar
Larson, M., & Kraus, N. C. 1991. Numerical model of longshore current for bar and trough beaches. Journal of the Waterways, Harbors and Coastal Engineering Division, ASCE 117, 326–347.CrossRefGoogle Scholar
Lasemi, Z., Boardman, M. R., & Sandberg, P. A. 1989. Cement origin of supratidal dolomite, Andros Island, Bahamas, Journal of Sedimentary Petrology 59, 249–257.Google Scholar
Lauff, G. H. 1967. Estuaries. Washington, D.C.: American Association for the Advancement of Science.Google Scholar
Laval, B., Cady, S. L, Pollack, J. C.et al. 2000. Modern freshwater microbialite analogues for ancient dendritic reef structures. Nature 407, 626–629.Google ScholarPubMed
Lawson, D. E. 1982. Mobilisation, movement and deposition of active subaerial sediment flows, Matanuska Glacier, Alaska. Journal of Geology 90, 279–300.CrossRefGoogle Scholar
Lawson, D. E., Strasser, J. C., Evenson, E. B.et al. 1998. Glaciohydraulic supercooling: a mechanism to create stratified, debris-rich basal ice.1. Field evidence and conceptual model. Journal of Glaciology 44, 547–562.CrossRefGoogle Scholar
Lay, T., Kanamori, H., Ammon, C. J.et al. 2005. The Great Sumatra–Andaman earthquake of 26 December 2004. Science 308, 1127–1133.CrossRefGoogle ScholarPubMed
Leatherman, S. P. 1979. Barrier Islands. New York: Academic Press.Google Scholar
Leatherman, S. P., Williams, A. T., & Fisher, J. S. 1977. Overwash sedimentation associated with a large-scale northeaster. Marine Geology 24, 109–121.CrossRefGoogle Scholar
Leckie, D. A., & Krystinik, L. F. 1989. Is there evidence for geostrophic currents preserved in the sedimentary record of inner to middle-shelf deposits?Journal of Sedimentary Petrology 59, 862–870.Google Scholar
Leclair, S. F. 2002. Preservation of cross-strata due to the migration of subaqueous dunes: an experimental investigation. Sedimentology 49, 1157–1180.CrossRefGoogle Scholar
Leclair, S. F., & Bridge, J. S. 2001. Quantitative interpretation of sedimentary structures formed by river dunes. Journal of Sedimentary Research 71, 713–716.CrossRefGoogle Scholar
Leclair, S. F., Bridge, J. S., & Wang, F. 1997. Preservation of cross-strata due to migration of subaqueous dunes over aggrading and non-aggrading beds: comparison of experimental data with theory. Geoscience Canada 24, 55–66.Google Scholar
Lee, M., Aronson, J. L, & Savin, S. M. 1989. Timing and conditions of Permian Rotliegende sandstone diagenesis, southern North Sea; K/Ar and oxygen isotope data. American Association of Petroleum Geologists Bulletin 73, 195–215.Google Scholar
Leeder, M. R. 1975. Pedogenic carbonates and flood sediment accretion rates: a quantitative method for alluvial arid-zone lithofacies. Geological Magazine 112, 257–270.CrossRefGoogle Scholar
Leeder, M. R.1978. A quantitative stratigraphic model for alluvium with special reference to channel deposit density and interconnectedness. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 587–596. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Leeder, M. R. 1980. On the stability of lower stage plane beds and the absence of current ripples in coarse sands. Journal of the Geological Society of London 137, 423–430.CrossRefGoogle Scholar
Leeder, M. R. 1982. Sedimentology: Process and Product. London: George Allen and Unwin.Google Scholar
Leeder, M. R.1993. Tectonic controls upon drainage basin development, river channel migration and alluvial architecture: implications for hydrocarbon reservoir development and characterization. In Characterization of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 7–22. London: Geological Society of London.Google Scholar
Leeder, M. R. 1999. Sedimentology and Sedimentary Basins: From Turbulence to Tectonics. Oxford: Blackwells.Google Scholar
Leeder, M. R., & Alexander, J. A. 1987. The origin and tectonic significance of asymmetrical meander belts. Sedimentology 34, 217–226.CrossRefGoogle Scholar
Leeder, M. R., Mack, G. H., Peakall, J., & Salyards, S. L. 1996. First quantitative test of alluvial stratigraphy models: southern Rio Grande rift, New Mexico. Geology 24, 87–90.2.3.CO;2>CrossRefGoogle Scholar
Leeder, M. R., & Stewart, M. 1996. Fluvial incision and sequence stratigraphy: alluvial responses to relative base level fall and their detection in the geological record. In Sequence Stratigraphy in British Geology (ed. Hesselbo, S. P. & Parkinson, D. N.), pp. 47–61. London: Geological Society of London.Google Scholar
Lees, A. 1975. Possible influences of salinity and temperature on modern shelf carbonate sedimentation. Marine Geology 19, 159–198.CrossRefGoogle Scholar
Legros, F. 2002. Can dispersive pressure cause inverse grading in grain flows?Journal of Sedimentary Research 72, 166–170.CrossRefGoogle Scholar
Leopold, L. B., & Langbein, W. B. 1962. The concept of entropy in landscape evolution. United States Geological Survey Professional Paper 500A.
Leopold, L. B., & Langbein, W. B. 1966. River meanders. Scientific American 214, 60–70.CrossRefGoogle Scholar
Leopold, L. B., & Wolman, M. G. 1957. River channel patterns: braided, meandering and straight. United States Geological Survey Professional Paper 282-B.
Leopold, L. B., Wolman, M. G., & Miller, J. P. 1964. Fluvial Processes in Geomorphology. San Francisco, CA: W. H. Freeman.Google Scholar
Lerman, A. 1978. Lakes: Chemistry Geology and Physics. New York: Springer-Verlag.CrossRefGoogle Scholar
Lerman, A., Imboden, D., & Gat, J. 1995. Physics and Chemistry of Lakes, 2nd edn. New York: Springer-Verlag.CrossRefGoogle Scholar
Levey, R. A. 1978. Bedform distribution and internal stratification of coarse-grained point bars, Upper Congaree River, South Carolina. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 105–127. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Levinton, J. S. 1982. Marine Ecology. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Li, M. Z., & Amos, C. L. 1999. Sheet flow and large wave ripples under combined waves and currents: field observations, model predictions and effects on boundary layer dynamics. Continental Shelf Research 19, 637–663.CrossRefGoogle Scholar
Lippman, F. 1973. Sedimentary Carbonate Minerals. New York: Springer-Verlag.CrossRefGoogle Scholar
Livingston, D. A. 1963. Chemical composition of rivers and lakes. United States Geological Survey Professional Paper 440-G.
Livingstone, I., & Warren, A. 1996. Aeolian Geomorphology: An Introduction. Harlow: Longman.Google Scholar
Lizitsin, A. P. 1996. Oceanic Sedimentation. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Lliboutry, L. 1979. Local friction laws for glaciers: a critical review and new openings. Journal of Glaciology 23, 67–95.CrossRefGoogle Scholar
Lockley, M. G. 1991. Tracking Dinosaurs: A New Look at an Ancient World. Cambridge: Cambridge University Press.Google Scholar
Lockley, M. G., & Hunt, A. P. 1995. Dinosaur Tracks and Other Fossil Footprints of the Western United States. New York: Columbia University Press.Google Scholar
Logan, B. W. 1987. The MacLeod Evaporite Basin Western Australia. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., Davies, G. R., Read, J. F., & Cebulski, D. E. 1970. Carbonate Sedimentation and Environments, Shark Bay, Western Australia. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., Hoffman, P., & Gebelein, C. D. 1974. Algal mats, crypt-algal fabrics and structures, Hamelin Pool, Western Australia. In Evolution and Diagenesis of Quaternary Carbonate Sequences, Shark Bay, Western Australia (ed. Logan, B. W., Read, J. F., Hagan, G. M.et al.), pp. 140–193. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., Read, J. F., Hagan, G. M.et al. 1974. Evolution and Diagenesis of Quaternary Carbonate Sequences, Shark Bay, Western Australia. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Logan, B. W., & Semeniuk, V. 1976. Dynamic Metamorphism: Processes and Products in Devonian Carbonate Rocks, Canning Basin, Western Australia. Sydney: Geological Society of Australia.Google Scholar
Longman, M. W. 1980. Carbonate diagenetic textures from near surface diagenetic environments. American Association of Petroleum Geologists Bulletin 64, 461–487.Google Scholar
Longuet-Higgins, M. S. 1953. Mass transport in water waves. Philosophical Transactions of the Royal Society245A.
Longuet-Higgins, M. S. 1970a. Longshore currents generated by obliquely incident waves, 1. Journal of Geophysical Research 75, 6778–6789.CrossRefGoogle Scholar
Longuet-Higgins, M. S. 1970b. Longshore currents generated by obliquely incident waves, 2. Journal of Geophysical Research 75, 6790–6801.CrossRefGoogle Scholar
Longuet-Higgins, M. S. 1981. Oscillating flow over steep sand ripples. Journal of Fluid Mechanics 107, 1–35.CrossRefGoogle Scholar
Loope, D. B. 1984. Origin of extensive bedding planes in aeolian sandstones: a defence of Stokes' hypothesis. Sedimentology 31, 123–132.CrossRefGoogle Scholar
Loope, D. B. 1985. Episodic deposition and preservation of eolian sands – a Late Paleozoic example from southeastern Utah. Geology 13, 73–76.2.0.CO;2>CrossRefGoogle Scholar
Loucks, R. G., & Sarg, J. F. 1993. Carbonate Sequence Stratigraphy: Recent Developments and Applications. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Lowe, D. R. 1975. Water escape structures in coarse-grained sediments. Sedimentology 22, 157–204.CrossRefGoogle Scholar
Lowe, D. R. 1976a. Subaqueous liquefied and fluidized sediment flows and their deposits. Sedimentology 23, 285–308.CrossRefGoogle Scholar
Lowe, D. R. 1976b. Grain flow and grain flow deposits. Journal of Sedimentary Petrology 46, 188–199.Google Scholar
Lowe, D. R.1979. Sediment gravity flows: their classification and some problems of application to natural flows and deposits. In Geology of Continental Slopes (ed. Doyle, L. J. & Pilkey, O. H.), pp. 75–82. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Lowe, D. R. 1982. Sediment gravity flows: 2. Depositional models with special reference to the deposits of high-density turbidity currents. Journal of Sedimentary Petrology 52, 279–297.Google Scholar
Lowe, D. R., Anderson, K. S., & Braunstein, D. 2001. The zonation and structuring of siliceous sinter around hot springs, Yellowstone National Park, and the role of thermophilic bacteria in its deposition. In Thermophiles: Biodiversity, Ecology, and Evolution (ed. Reysenbach, A.-L., Voytek, M., & Mancinelli, R.), pp. 143–166. New York: Kluwer Academic.CrossRefGoogle Scholar
Lowe, D. R., & Braunstein, D. 2003. Microstructure of high-temperature (> 73 ℃) siliceous sinter deposited around hot springs and geysers, Yellowstone National Park: the role of biological and abiological processes in sedimentation. Canadian Journal of Earth Sciences 40, 1611–1642.CrossRefGoogle Scholar
Lowe, D. R., & Lopiccolo, R. D. 1974. The characteristics and origin of dish and pillar structures. Journal of Sedimentary Petrology 44, 484–501.Google Scholar
Lowenstam, H. A. 1954. Factors affecting the aragonite : calcite ratios in carbonate-secreting marine organisms. Journal of Geology 62, 284–322.CrossRefGoogle Scholar
Lowenstam, H. A., & Weiner, S. 1989. On Biomineralization. Oxford: Oxford University Press.Google Scholar
Lowenstein, T. K. 2002. Pleistocene lakes and paleoclimates (0 to 200 ka) in Death Valley, California. Smithsonian Contributions to the Earth Sciences 33, 109–120.Google Scholar
Lowenstein, T. K., & Hardie, L. A. 1985. Criteria for the recognition of salt-pan evaporites. Sedimentology 32, 627–644.CrossRefGoogle Scholar
Lowenstein, T. K., Hardie, L. A., Timofeeff, M. N., & Demicco, R. V. 2003. Secular variation in seawater chemistry and the origin of calcium chloride basinal brines. Geology 31, 857–860.CrossRefGoogle Scholar
Lowenstein, T. K., Li., J., Brown, C.et al. 1999. 200 k.y. paleoclimate record from Death Valley salt core. Geology 27, 3–6.2.3.CO;2>CrossRefGoogle Scholar
Lowenstein, T. K., Timofeeff, M. N., Brennan, S. T., Hardie, L. A., & Demicco, R. V. 2001. Oscillations in Phanerozoic seawater chemistry: Evidence from fluid inclusions. Science 294, 1086–1088.CrossRefGoogle ScholarPubMed
Lundqvist, J. 1989. Rogen (ribbed) moraine – identification and possible origin. Sedimentary Geology 62, 281–292.CrossRefGoogle Scholar
Lundqvist, J. 1997. Rogen moraine – an example of two-step formation of glacial landscapes. Sedimentary Geology 111, 27–40.CrossRefGoogle Scholar
Lunt, I. A., & Bridge, J. S. 2006. Formation and preservation of open-framework gravel strata in unidirectional water flows. Sedimentology 53, 1–17.Google Scholar
Lyell, C. 1830. The Principles of Geology. London: John Murray.Google Scholar
MacEachern, J. A., Bann, K. L., Bhattacharya, J. P., & Howell, C. D. Jr. 2005. Ichnology of deltas: organism responses to the dynamic interplay of rivers, waves, storms, and tides. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J. P.), pp. 49–85. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Macintyre, I. G. 1985. Submarine cements – the peloidal question. In Carbonate Cements (ed. Schneidermann, N. & Harris, P. M.), pp. 109–116. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Mack, G. H., & James, W. C. 1994. Paleoclimate and global distribution of paleosols. Journal of Geology 102, 360–362.CrossRefGoogle Scholar
Mack, G. H., James, W. C., & Monger, H. C. 1993. Classification of paleosols. Geological Society of America Bulletin 105, 129–136.2.3.CO;2>CrossRefGoogle Scholar
Mack, G. H., & Leeder, M. R. 1998. Channel shifting of the Rio Grande, southern Rio Grande rift: implications for alluvial stratigraphy models. Sedimentary Geology 177, 207–219.CrossRefGoogle Scholar
Mackenzie, F. T., & Pigott, J. D. 1981. Tectonic controls of Phanerozoic sedimentary rock cycling. Journal of the Geological Society of London 138, 183–196.CrossRefGoogle Scholar
Mackey, S. D., & Bridge, J. S. 1995. Three dimensional model of alluvial stratigraphy: theory and application. Journal of Sedimentary Research B65, 7–31.CrossRefGoogle Scholar
Mader, C. L. 1974. Numerical simulation of tsunamis. Journal of Physical Oceanography 4, 74–82.2.0.CO;2>CrossRefGoogle Scholar
Mader, C. L. 1988. Numerical Modeling of Water Waves. Berkeley, CA: University of California Press.Google Scholar
Madsen, O. S. 1977. A realistic model of the wind-induced Ekman boundary layer. Journal of Physical Oceanagraphy 7, 248–255.2.0.CO;2>CrossRefGoogle Scholar
Maizels, J. K. 1989. Sedimentology, paleoflow dynamics and flood history of jokulhlaup deposits: paleohydrology of Holocene sediment sequences in southern Iceland sadur deposits. Journal of Sedimentary Petrology 59, 204–223.Google Scholar
Maizels, J. K.2002. Sediments and landforms of modern proglacial terrestrial environments. In Modern and Past Glacial Environments (ed. Menzies, J.), pp. 279–333. Oxford: Butterworth-Heineman.CrossRefGoogle Scholar
Major, J. J. 1997. Depositional processes in large-scale debris-flow experiments. Journal of Geology 105, 345–366.CrossRefGoogle Scholar
Makaske, B. 2001. Anastomosing rivers: a review of their classification, origin and sedimentary products. Earth-Science Reviews 53, 149–196.CrossRefGoogle Scholar
Makse, H. A., Havlin, S., King, P. R., & Stanley, H. E. 1997. Spontaneous stratification in granular mixtures. Nature 386, 379–382.CrossRefGoogle Scholar
Maliva, R. G., Knoll, A. H., & Simonson, B. M. 2005. Secular change in the Precambrian silica cycle: insights from chert petrology. Geological Society of America Bulletin 117, 835–845.CrossRefGoogle Scholar
Maliva, R. G., & Siever, R. 1989. Nodular chert formation in carbonate rocks. Journal of Geology 97, 421–433.CrossRefGoogle Scholar
Mancini, E. A, Llinas, J. C., Parcell, W. C.et al. 2004. Upper Jurassic thrombolite reservoir play, northeastern Gulf of Mexico. American Association of Petroleum Geologists Bulletin 88, 1573–1602.CrossRefGoogle Scholar
Mann, K. H. 2000. Ecology of Coastal Waters: With Implications for Management. Massachusetts, PA: Blackwell Science.Google Scholar
Mann, S. 2001. Biomineralization: Principles and Concepts in Bioinorganic Materials Chemistry. Oxford: Oxford University Press.Google Scholar
Marani, M., Lanzoni, S., Zandolin, D, Seminara, G., & Rinaldo, A. 2002. Tidal meanders. Water Resources Research 38, 7–1 to 7–11.CrossRefGoogle Scholar
Marr, J. G., Swenson, J. B., Paola, C., & Voller, V. R. 2000. A two-diffusion model of fluvial stratigraphy in closed depositional basins. Basin Research 12, 381–398.CrossRefGoogle Scholar
Marshall, J. F., & Davies, P. J. 1975. High magnesium calcite ooids from the Great Barrier Reef. Journal of Sedimentary Petrology 45, 285–291.Google Scholar
Martini, I. P., Baker, V. R., & Garzon, G. 2002. Flood and Megaflood Processes and Deposits. Oxford: Blackwells.CrossRefGoogle Scholar
Martini, I. P., Brookfield, M. E., & Sadura, S. 2001. Glacial Geomorphology and Geology. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Mastbergen, D. R., & Berg, J. H. 2003. Breaching in fine sands and the generation of sustained turbidity currents in submarine canyons. Sedimentology 50, 625–637.CrossRefGoogle Scholar
Matter, A., & Tucker, M. E. 1978. Modern and Ancient Lake Sediments. Oxford: Blackwells.CrossRefGoogle Scholar
McCabe, M., & Eyles, N. 1988. Sedimentology of an ice-contact glaciomarine delta, Carey Valley, Northern Ireland. Sedimentary Geology 59, 1–14.CrossRefGoogle Scholar
McCabe, P. J. 1984. Depositional environments of coal and coal-bearing strata. In Sedimentology of Coal and Coal-bearing Sequences (ed. Rahmani, R. A. & Flores, R. M.), pp. 13–42. Oxford: Blackwells.Google Scholar
McCabe, P. J., & Parrish, J. T. 1992. Tectonic and climatic controls on the distribution and quality of Cretaceous coals. In Controls on the Distribution and Quality of Cretaceous Coals (ed. McCabe, P. J. & Parrish, J. T.), pp. 1–15. Boulder, CO: Geological Society of America.Google Scholar
McCaffrey, W. D., Kneller, B. C., & Peakall, J. 2001. Particulate Gravity Currents. Oxford: Blackwells.CrossRefGoogle Scholar
McCarthy, P. J., & Plint, A. G. 1998. Recognition of interfluve sequence boundaries: integrating paleopedology and sequence stratigraphy. Geology 26, 387–390.2.3.CO;2>CrossRefGoogle Scholar
McCave, I. N. 1968. Shallow and marginal marine sediments associated with the Catskill complex in the Middle Devonian of New York. In Late Paleozoic and Mesozoic Continental Sedimentation, Northeastern North America (ed. Klein, G. deV.), pp. 75–108. Boulder, CO: Geological Society of America.Google Scholar
McCave, I. N. 1969. Correlation of marine and nonmarine strata with example from Devonian of New York State. American Association of Petroleum Geologists Bulletin 53, 155–162.Google Scholar
McCave, I. N. 1970. Deposition of fine-grained suspended sediment from tidal currents. Journal of Geophysical Research 75, 4151–4159.CrossRefGoogle Scholar
McCave, I. N.1985. Recent shelf clastic sediment. In Sedimentology: Recent Developments and Applied Aspects (ed. Brenchley, P. J. & Williams, B. P. J.), pp. 49–65. London: Geological Society of London.Google Scholar
McCubbin, D. G. 1982. Barrier island and strand plain facies. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D. R.), pp. 247–279. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
McHargue, T. R. 1991. Seismic facies, processes and evolution of Miocene inner fan channels, Indus submarine fan. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 403–414. New York: Springer-Verlag.CrossRefGoogle Scholar
McIlreath, I. A. 1977. Accumulation of a Middle Cambrian deep water, basinal limestone adjacent to a vertical, submarine carbonate escarpment, southern Rocky Mountains. In Deep Water Carbonate Environments (ed. Cook, H. E. & Enos, P.), pp. 113–124. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
McIlreath, I. A., & James, N. P. 1984. Carbonate slopes. In Facies Models, 2nd edn. (ed. Walker, R. G.), pp. 245–257. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
McIlreath, I. A., & Morrow, D. W. 1990. Diagenesis. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
McKee, E. D. 1966. Structure of dunes at White Sands National Monument, New Mexico. Sedimentology 7, 3–69.CrossRefGoogle Scholar
McKee, E. D.1979. A study of global sand seas. United States Geological Survey Professional Paper 1052.
McKee, E. D.1982. Sedimentary structures in dunes of the Namib Desert, South West Africa. Geological Society of America Special Paper 188.CrossRef
McKee, E. D., & Gutschick, R. C. 1969. History of the Redwall Limestone of Northern Arizona. Boulder, CO: Geological Society of America.CrossRefGoogle Scholar
McKee, E. D., & Tibbitts, G. C. Jr. 1964. Primary structures of a seif dune and associated deposits in Libya. Journal of Sedimentary Petrology 34, 5–17.Google Scholar
McKee, E. D., & Weir, G. W. 1953. Terminology for stratification and cross-stratification in sedimentary rocks. Geological Society of America Bulletin 64, 381–390.CrossRefGoogle Scholar
McKenzie, J. A., Hsü, K. J., & Schneider, J. F. 1980. Movement of subsurface waters under the sabkha, Abu Dhabi, UAE, and its relation to evaporative dolomite genesis. In Concepts and Models of Dolomitization (ed. Zenger, D. H., Dunham, J. B., & Ethington, R. L.), pp. 11–30. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
McLean, S. R. 1990. The stability of ripples and dunes. Earth-Science Reviews 29, 131–144.CrossRefGoogle Scholar
McNeill, J. R., & Winiwarter, V. 2004. Breaking the sod: humankind, history, and soil. Science 304, 1627–1629.CrossRefGoogle Scholar
McPhee, J. 1989. Control of Nature. New York: Farrar, Straus and Giroux.Google Scholar
Mehta, A. J. 1989. On estuarine cohesive sediment suspension behaviour. Journal of Geophysical Research 94, 14,303–14,314.CrossRefGoogle Scholar
Melim, L. A., & Swart, P. K. 2001. Meteoric and marine-burial diagenesis in the subsurface of Great Bahama Bank. In Subsurface Geology of a Prograding Carbonate Platform Margin, Great Bahama Bank: Results of the Bahamas Drilling Project (ed. Ginsburg, R. N.), pp. 137–161. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Melosh, H. J. 1987. The mechanics of large rock avalanches. In Debris Flows/Avalanches: Processes, Recognition, and Mitigation (ed. Costa, J. E. & Wieczorek, G. F.), pp. 41–49. Boulder, CO: Geological Society of America.Google Scholar
Menzies, J. 1995. Modern Glacial Environments – Processes, Dynamics and Sediments. Oxford: Butterworth-Heinemann.Google Scholar
Menzies, J. 1996. Past Glacial Environments – Sediments, Forms and Techniques. Oxford: Butterworth-Heinemann.Google Scholar
Menzies, J. 2002. Modern and Past Glacial Environments. Oxford: Butterworth-Heinemann.CrossRefGoogle Scholar
Menzies, J., & Rose, J. 1989. Subglacial bedforms – drumlins, Rogen moraine and associated subglacial bedforms. Sedimentary Geology 62, 117–407.CrossRefGoogle Scholar
Meybeck, M. 1995. Global distribution of lakes. In Physics and Chemistry of Lakes, 2nd edn. (ed. Lerman, A., Imboden, D., & Gat, J.), pp. 1–35. New York: Springer-Verlag.CrossRefGoogle Scholar
Meyer, G. A., Wells, S. G., Balling, R. C., & Jull, A. J. T. 1992. Response of alluvial systems to fire and climate change in Yellowstone National Park. Nature 357, 147–150.CrossRefGoogle Scholar
Meyers, W. J. 1974. Carbonate cement stratigraphy of the Lake Valley Formation (Mississippian), Sacramento Mountains, New Mexico. Journal of Sedimentary Petrology 44, 837–861.Google Scholar
Meyers, W. J.1991. Calcite cement stratigraphy: an overview. In Luminescence Microscopy: Quantitative and Qualitative Analysis (ed. Barker, C. E. & Kipp, O. C.), pp. 133–148. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Miall, A. D. 1986. Eustatic sea level changes interpreted from seismic stratigraphy: a critique of the methodology with particular reference to the North Sea Jurassic record. American Association of Petroleum Geologists Bulletin 70, 131–137.Google Scholar
Miall, A. D. 1991. Stratigraphic sequences and their chronostratigraphic correlation. Journal of Sedimentary Petrology 61, 497–505.Google Scholar
Miall, A. D. 1996. The Geology of Fluvial Deposits. New York: Springer-Verlag.Google Scholar
Miall, A. D. 1997. The Geology of Stratigraphic Sequences. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Miall, A. D., & Smith, N. D. 1989. Rivers and Their Deposits. Slide Set No. 4. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Mickelson, D. M., & Attig, J. W. 1999. Glacial processes past and present. Geological Society of America Special Paper 337.
Middleton, G. V. 1966a. Experiments on density and turbidity currents. 1: motion of the head. Canadian Journal of Earth Sciences 3, 523–546.Google Scholar
Middleton, G. V. 1966b. Experiments on density and turbidity currents. 2: uniform flow of turbidity currents. Canadian Journal of Earth Sciences 3, 627–637.CrossRefGoogle Scholar
Middleton, G. V. 1966c. Experiments on density and turbidity currents. 3: deposition of sediment. Canadian Journal of Earth Sciences 4, 475–505.CrossRefGoogle Scholar
Middleton, G. V. 1993. Sediment deposition from turbidity currents. Annual Review of Earth and Planetary Sciences 21, 89–114.CrossRefGoogle Scholar
Middleton, G. V. 2003. Encyclopedia of Sediments and Sedimentary Rocks. Dordrecht: Kluwer Academic Publishers.Google Scholar
Middleton, G. V., & Hampton, M. A. 1973. Sediment gravity flows: mechanics of flow and deposition. In Turbidites and Deep-Water Sedimentation (ed. Middleton, G. V. & Bouma, A. H.), pp. 1–38. Los Angeles, CA: SEPM (Society for Sedimentary Geology) Pacific Section.Google Scholar
Middleton, G. V., & Hampton, M. A.1976. Subaqueous sediment transport and deposition by sediment gravity flows. In Marine Sediment Transport and Environmental Management (ed. Stanley, D. J. & Swift, D. J. P.), pp. 197–218. New York: Wiley.Google Scholar
Middleton, G. V., & Southard, J. B. 1984. Mechanics of Sediment Movement. Tulsa, OK: SEPM.CrossRefGoogle Scholar
Middleton, G. V., & Wilcock, P. R. 1994. Mechanics in the Earth and Environmental Sciences. Cambridge: Cambridge University Press.Google Scholar
Miller, J. M. G. 1996. Glacial sediments. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 454–484. Oxford: Blackwells.Google Scholar
Miller, M. C., & Komar, P. D. 1980. Oscillation ripples generated by laboratory experiments. Journal of Sedimentary Petrology 50, 173–182.Google Scholar
Milliman, J. D., & Barretto, H. T. 1975. Relic magnesian calcite oolite and subsidence of the Amazon shelf. Sedimentology 22, 137–145.CrossRefGoogle Scholar
Mitchell, R. W. III. 1985. Comparative sedimentology of shelf carbonates of the Middle Ordovician St. Paul Group of the central Appalachians. Sedimentary Geology 43, 1–41.CrossRefGoogle Scholar
Mohrig, D., Heller, P. L., Paola, C., & Lyons, W. J. 2000. Interpreting avulsion process from ancient alluvial sequences: Guadalope–Matarranya system (northern Spain) and Wasatch Formation (western Colorado). Geological Society of America Bulletin 112, 1787–1803.2.0.CO;2>CrossRefGoogle Scholar
Mojzsis, S. J., Arrhenius, G., McKeegan, K. D.et al. 1996. Evidence for life on Earth before 3,800 million years ago. Nature 384, 55–59.CrossRefGoogle ScholarPubMed
Moller, P. 2006. Rogen moraine: an example of glacial reshaping of pre-existing landforms. Quaternary Science Reviews 25, 362–389.CrossRefGoogle Scholar
Molnia, B. F. 1983. Glacial–Marine Sedimentation. New York: Plenum.CrossRefGoogle Scholar
Montañez, I. P., Gregg, J. M., & Shelton, K. L. 1997. Basin-wide Diagenetic Patterns: Integrated Petrologic, Geochemical, and Hydrologic Considerations. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Montgomery, D. R., & Dietrich, W. E. 1989. Source areas, drainage density, and channel initiation. Water Resources Research 25, 1907–1918.CrossRefGoogle Scholar
Montgomery, D. R., & Dietrich, W. E.1994. Landscape dissection and drainage area-slope thresholds. In Process Models and Theoretical Geomorphology (ed. Kirkby, M. J.), pp. 221–246. Chichester: Wiley.Google Scholar
Mora, C. I., & Driese, S. G. 1993. A steep, mid- to late Paleozoic decline in atmospheric CO2: evidence from soil carbonate CO2 barometer. Chemical Geology 107, 217–219.CrossRefGoogle Scholar
Mora, C. I., & Driese, S. G.1999. Palaeoenvironment, palaeoclimate and stable carbon isotopes of Palaeozoic red-bed palaeosols, Appalachian Basin, USA and Canada. In Palaeoweathering, Palaeosurfaces and Related Continental Deposits (ed. Thiry, M. & Simon-Coicon, R.), pp. 61–84. Oxford: Blackwells.Google Scholar
Mora, C. I., Driese, S. G., & Colarusso, L. A. 1996. Middle to late Paleozoic atmospheric CO2 levels from soil carbonate and organic matter. Science 271, 1105–1107.CrossRefGoogle Scholar
Morad, S., Ketzer, J. M., & Ros, L. F. 2000. Spatial and temporal distribution of diagenetic alterations in siliciclastic rocks: implications for mass transfer in sedimentary basins. Sedimentology 47, 95–120.CrossRefGoogle Scholar
Morgan, J. P. 1970. Deltaic Sedimentation: Modern and Ancient. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Morgan, J. P., Coleman, J. M., & Gagliano, S. M. 1968. Mudlumps: diapiric structures in Mississippi delta sediments. In Diapirism and Diapirs (ed. Braunstein, J. & O'Brien, G. D.), pp. 145–161. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Moore, A., Nishimura, Y., Gelfenbaum, G., Kamataki, T., & Triyono, R. 2006. Sedimentary deposits of the 26 December 2004 tsunami on the northwest coast of Aceh, Indonesia. Earth Planets Space 58, 253–258.CrossRefGoogle Scholar
Morozova, G. S., & Smith, N. D. 1999. Holocene avulsion history of the lower Saskatchewan fluvial system, Cumberland Marshes, Saskatchewan–Manitoba, Canada. In Fluvial Sedimentology VI (ed. Smith, N. D. & Rogers, J.), pp. 231–249. Oxford: Blackwells.CrossRefGoogle Scholar
Morris, P. E., & Williams, D. J. 1997. Exponential longitudinal profiles of streams. Earth Surface Processes and Landforms 22, 143–163.3.0.CO;2-Z>CrossRefGoogle Scholar
Morris, P. E., & Williams, D. J. 1999a. A worldwide correlation for exponential bed particle size variations in subaerial aqueous flows. Earth Surface Processes and Landforms 24, 835–847.3.0.CO;2-G>CrossRefGoogle Scholar
Morris, P. E., & Williams, D. J. 1999b. Worldwide correlations for subaerial aqueous flows with exponential longitudinal profiles. Earth Surface Processes and Landforms 24, 867–879.3.0.CO;2-L>CrossRefGoogle Scholar
Morse, J. W., Millero, F. J., Thurmond, V., Brown, E., & Ostlund, H. G. 1984. The carbonate chemistry of Grand Bahama Bank waters: after 18 years another look. Journal of Geophysical Research 89, 3604–3614.CrossRefGoogle Scholar
Morse, J. W., Wang, Q., & Tzio, M. Y. 1997. Influences of temperature and Mg : Ca ratio on CaCO3 precipitates from seawater. Geology 25, 85–87.2.3.CO;2>CrossRefGoogle Scholar
Morton, R. A. 1981. Formation of storm deposits by wind-forced currents in the Gulf of Mexico and the North Sea. In Holocene Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 385–396. Oxford: Blackwells.CrossRefGoogle Scholar
Moslow, T. F., & Heron, S. D. 1978. Relict inlets: preservation and occurrence in the Holocene stratigraphy of southern Core Banks, North Carolina. Journal of Sedimentary Petrology 48, 1275–1286.Google Scholar
Moslow, T. F., & Tye, R. S. 1985. Recognition and characterization of Holocene tidal inlet sequences. Marine Geology 63, 129–151.CrossRefGoogle Scholar
Mountney, N. P. 2006. Periodic accumulation and destruction of aeolian erg sequences in the Permian Cedar Mesa Sandstone, White Canyon, southern Utah, USA. Sedimentology 53, 789–823.CrossRefGoogle Scholar
Mountney, N. P., & Jagger, A. 2004. Stratigraphic evolution of an aeolian erg margin system: the Permian Cedar Mesa sandstone, SE Utah, USA. Sedimentology 51, 713–743.CrossRefGoogle Scholar
Mulder, T., & Syvitski, J. P. M. 1995. Turbidity currents generated at river mouths during exceptional discharges to the world's oceans. Journal of Geology 103, 285–299.CrossRefGoogle Scholar
Mullins, H. T. 1983. Modern carbonate slopes and basins of the Bahamas. In Platform Margin and Deep Water Carbonates (ed. Cook, H. E., Hine, A. C., & Mullins, H. T.), pp. 4.1–4.138. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Multer, H. G. 1975. Field Guide to Some Carbonate Rock Environments, Florida Keys and Western Bahamas. Madison, NJ: Farleigh Dickinson University.Google Scholar
Multer, H. G., & Hoffmeister, J. E. 1975. Petrology and significance of the Key Largo (Pleistocene) Limestone, Florida Keys. In Field Guide to some Carbonate Rock Environments; Florida Keys and Western Bahamas (ed. Multer, H. G.) pp. 111–112. Madison, NJ: Farleigh Dickinson University.Google Scholar
Murray, A. B., & Paola, C. 1994. A cellular model of braided rivers. Nature 371, 54–57.CrossRefGoogle Scholar
Murray, A. B., & Paola, C. 1997. Properties of a cellular braided stream model. Earth Surface Processes and Landforms 22, 1001–1025.3.0.CO;2-O>CrossRefGoogle Scholar
Murray, P. B., Davies, A. G., & Soulsby, R. L. 1991. Sediment pick-up in wave and current flows. In Sand Transport in Rivers, Estuaries and the Sea (ed. Soulsby, R. L. & Bettess, R.), pp. 37–44. Rotterdam: Balkema.Google Scholar
Murray, S. P. 1970. Bottom currents near the coast during hurricane Camille. Journal of Geophysical Research 75, 4579–4582.CrossRefGoogle Scholar
Murton, J. B., Peterson, R., & Ozouf, J.-C. 2006. Bedrock fracture by ice segregation in cold regions. Science 314, 1127–1129.CrossRefGoogle ScholarPubMed
Mutti, E. 1985. Turbidite systems and their relations to depositional sequences. In Provenance of Arenites (ed. Zuffa, G. G.), pp. 65–93. Amsterdam: Reidel.CrossRefGoogle Scholar
Mutti, E., & Normark, W. R. 1987. Comparing examples of modern and ancient turbidite systems: problems and concepts. In Marine Clastic Sedimentology: Concepts and Case Studies (ed. Legget, J. K. & Zuffa, G. G.), pp. 1–38. London: Graham and Trotman.CrossRefGoogle Scholar
Mutti, E., & Normark, W. R.1991. An integrated approach to the study of turbidite systems. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 75–106. New York: Springer-Verlag.CrossRefGoogle Scholar
Mutti, E., & Lucci, Ricci F. 1972. Le torbiditi dell'Appennino Settentrionale: introduzione all'analisi di facies. Memorie della Società Geologica Italiana 11, 161–199.Google Scholar
Myrow, P. M., & Southard, J. B. 1991. Combined-flow model for vertical stratification sequences in shallow marine storm-deposited beds. Journal of Sedimentary Petrology 61, 202–210.Google Scholar
Nardin, T. R., Hein, F. J., Gorsline, D. S., & Edwards, B. D. 1979. A review of mass movement processes, sediment and acoustic characteristics, and contrasts in slope and base-of-slope versus canyon–fan–basin floor systems. In Geology of Continental Slopes (ed. Doyle, L. J. & Pilkey, O. H.), pp. 61–73. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Nealson, K. H. 1997. Sedimentary bacteria: who's there, what are they doing, and what's new?Annual Review of Earth and Planetary Sciences 25, 403–434.CrossRefGoogle ScholarPubMed
Necker, F., Hartel, C., Kleiser, L., & Meiburg, E. 2005. Mixing and dissipation in particle-driven gravity currents. Journal of Fluid Mechanics 545, 339–372.CrossRefGoogle Scholar
Neev, D., & Emery, K. O. 1967. The Dead Sea: depositional processes and environments of evaporites. Israel Geological Survey Bulletin 41.Google Scholar
Nemec, W., & Steel, R. J. 1988. Fan Deltas: Sedimentology and Tectonic Settings. London: Blackie.Google Scholar
Nezu, I., & Nakagawa, H. 1993. Turbulence in Open-Channel Flows. Rotterdam: Balkema.Google Scholar
Nguyen, Q. D., & Boger, D. V. 1992. Measuring the flow properties of yield stress fluids. Annual Review of Fluid Mechanics 24, 47–88.CrossRefGoogle Scholar
Nichols, G. 1999. Sedimentology and Stratigraphy. Oxford: Blackwells.Google Scholar
Nichols, M. J. 1989. Sediment accumulation rates and relative sea level rise in lagoons. Marine Geology 88, 201–220.CrossRefGoogle Scholar
Nichols, R. J., Sparks, R. S. J., & Wilson, C. J. N. 1994. Experimental studies of the fluidization of layered sediments and the formation of fluid escape structures. Sedimentology 41, 233–253.CrossRefGoogle Scholar
Nickling, W. G. 1994. Aeolian sediment transport and deposition. In Sediment Transport and Depositional Processes (ed. Pye, K.), pp. 293–350. Oxford: Blackwells.Google Scholar
Nielsen, P. 1992. Coastal Bottom Boundary Layers and Sediment Transport. Singapore: World Scientific.CrossRefGoogle Scholar
Nielson, J., & Kocurek, G. 1986. Climbing zibars in the Algodones. Sedimentary Geology 48, 1–15.CrossRefGoogle Scholar
Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E. 1981. Holocene Sedimentation in the North Sea Basin. Oxford: Blackwells.CrossRefGoogle Scholar
Nio, S. D., Seigenthaler, C., & Yang, C. S. 1983. Megaripple cross-bedding as a tool for the reconstruction of the palaeohydraulics in a Holocene subtidal environment, S.W. Netherlands. Geologie en Mijnbouw 62, 499–510.Google Scholar
Nio, S. D., & Yang, C. S. 1991. Diagnostic attributes of clastic tidal deposits: a review. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 3–28. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Nittrouer, C. A., & Wright, L. D. 1994. Transport of particles across continental shelves. Reviews of Geophysics 32, 85–113.CrossRefGoogle Scholar
Noffke, N., Gerdes, G., & Klenke, T. 2005. Benthic cyanobacteria and their influence on the sedimentary dynamics of peritidal depositional systems (siliciclastic, evaporitic salty, and evaporitic carbonatic). Earth-Science Reviews 62, 163–176.CrossRefGoogle Scholar
Normark, W. R. 1970. Growth patterns of deep-sea fans. American Association of Petroleum Geologists Bulletin 54, 2170–2195.Google Scholar
Normark, W. R. 1978. Fan valleys, channels, and depositional lobes on modern submarine fans: characters for recognition of sandy turbidite environments. American Association of Petroleum Geologists Bulletin 62, 912–931.Google Scholar
Normark, W. R., & Piper, D. J. W. 1985. Navy Fan, Pacific Ocean. In Submarine Fans and Related Turbidite Systems (ed. Bouma, A. H., Normark, W. R., & Barnes, N. E.), pp. 87–94. New York: Springer-Verlag.CrossRefGoogle Scholar
Normark, W. R., Piper, D. J. W., & Hess, G. R. 1979. Distributary channels, sand lobes, and mesotopography of Navy submarine fan, California Borderland, with applications to ancient fan sediments. Sedimentology 26, 749–774.CrossRefGoogle Scholar
North, C. P. 1996. The prediction and modelling of subsurface fluvial stratigraphy. In Advances in Fluvial Dynamics and Stratigraphy (ed. Carling, P. A. & Dawson, M. R.), pp. 395–508. Chichester: Wiley.Google Scholar
North, C. P., & Prosser, D. J. 1993. Characterisation of Fluvial and Aeolian Reservoirs. London: Geological Society of London.Google Scholar
North, C. P., & Taylor, K. S. 1996. Ephemeral–fluvial deposits: integrated outcrop and simulation studies reveal complexity. American Association of Petroleum Geologists Bulletin 80, 811–830.Google Scholar
Notholt, A. J. G., Sheldon, R. P., & Davidson, D. F. 1989. Phosphate Deposits of the World, Volume 2, Phosphate Rock Resources. Cambridge: Cambridge University Press.Google Scholar
Nottvedt, A., & Kreisa, R. D. 1987. Model for the combined flow origin of hummocky cross-stratification. Geology 15, 357–361.2.0.CO;2>CrossRefGoogle Scholar
Nummedal, D., & Penland, S. 1981. Sediment dispersal in Norderneyer Seegate, West Germany. In Holocene Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 187–210. Oxford: Blackwells.CrossRefGoogle Scholar
Nummedal, D., Pilkey, O. H., & Howard, J. D. 1987. Sea-Level Fluctuations and Coastal Evolution. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Nummedal, D., & Swift, D. J. P. 1987. Transgressive stratigraphy at sequence-bounding unconformities: some principles derived from Holocene and Cretaceous examples. In Sea-level Fluctuations and Coastal Evolution (ed. Nummedal, D., Pilkey, O. H., & Howard, J. D.), pp. 241–260. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Nye, J. F. 1957. The distribution of stress and velocity in glaciers and ice sheets. Proceedings of the Royal Society of London 239A, 113–133.CrossRefGoogle Scholar
Nye, J. F. 1965. The flow of a glacier in a channel of rectangular, elliptic or parabolic cross-section. Journal of Glaciology 5, 661–690.CrossRefGoogle Scholar
Oertel, G. F. 1979. Barrier island development during the Holocene recession, SE United States. In Barrier Islands (ed. Leatherman, S. P.), pp. 273–290. New York: Academic Press.Google Scholar
Oertel, G. F. 1985. The barrier island system. Marine Geology 63, 1–18.CrossRefGoogle Scholar
Oertel, G. F.1988. Processes of sediment exchange between tidal inlets, ebb deltas and barrier islands. In Hydrodynamics and Sediment Dynamics of Tidal Inlets (ed. Aubrey, D. G. & Weishar, L.), pp. 297–318. Berlin: Springer-Verlag.Google Scholar
Oertel, G. F., Kearney, M. S., Leatherman, S. P., & Woo, J. 1989. Anatomy of a barrier platform: outer barrier lagoon, southern Delmarva Peninsula, Virginia. Marine Geology 88, 303–318.CrossRefGoogle Scholar
Ohfuji, H., & Rickard, D. 2005. Experimental synthesis of framboids – a review. Earth-Science Reviews 71, 147–170.CrossRefGoogle Scholar
Ohmoto, H., & Skinner, B. J. 1983. The Kuroko and Related Volcanogenic Massive Sulfide Deposits. New Haven, CT: The Economic Geology Publishing Company.Google Scholar
Olcott, A. N., Sessions, A. L., Corsetti, F. A., Kaufman, A. J., & Oliviera, T. F. 2005. Biomarker evidence for photosynthesis during Neoproterozoic glaciation. Science 310, 471–473.CrossRefGoogle ScholarPubMed
Oliver, J. 1986. Fluids expelled tectonically from orogenic belts: their role in hydrocarbon migration and other geological phenomena. Geology 14, 99–102.2.0.CO;2>CrossRefGoogle Scholar
Olsen, H. 1990. Astronomical forcing of meandering river behaviour: Milankovitch cycles in Devonian of East Greenland. Palaeogeography, Palaeoclimatology, Palaeoecology 79, 99–115.CrossRefGoogle Scholar
Olsen, H.1994. Orbital forcing on continental depositional systems – lacustrine and fluvial cyclicity in the Devonian of East Greenland. In Orbital Forcing and Cyclic Sequences (ed. Boer, P. L. & Smith, D. G.), pp. 429–438. Oxford: Blackwells.CrossRefGoogle Scholar
Olsen, P. E. 1986. A 40-million-year lake record of Early Mesozoic orbital climatic forcing. Science 234, 842–848.CrossRefGoogle ScholarPubMed
Oltman-Shay, J., & Guza, R. T. 1987. Infragravity edge wave observations on two California beaches. Journal of Physical Oceanography 17, 644–663.2.0.CO;2>CrossRefGoogle Scholar
Open University 1989. Ocean Circulation. Oxford: Pergamon Press.
Orton, G. J. 1996. Volcanic environments. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 485–567. Oxford: Blackwells.Google Scholar
Orton, G. J., & Reading, H. G. 1993. Variability of deltaic processes in terms of sediment supply, with particular emphasis on grain size. Sedimentology 40, 475–512.CrossRefGoogle Scholar
Osborne, P. D., & Greenwood, B. 1993. Sediment suspension under waves and currents: time scales and vertical structure. Sedimentology 40, 599–622.CrossRefGoogle Scholar
Osgood, R. G. Jr. 1987. Trace fossils. In Fossil Invertebrates (ed. Boardman, R. S., Cheetham, A. H., & Rowell, A. J.), pp. 663–674. Oxford: Blackwell Scientific.Google Scholar
Otvos, E. G., & Price, W. A. 1979. Problems of chenier genesis and terminology – an overview. Marine Geology 31, 251–263.CrossRefGoogle Scholar
Ouchi, S. 1985. Response of alluvial rivers to slow active tectonic movement. Geological Society of America Bulletin, 96, 504–515.2.0.CO;2>CrossRefGoogle Scholar
Overeem, I., Syvitski, J. P. M., & Hutton, E. W. H. 2005. Three-dimensional numerical modeling of deltas. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J. P.), pp. 13–30. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Owen, G. 1996. Experimental soft-sediment deformation: structures formed by the liquefaction of unconsolidated sands and some ancient examples. Sedimentology 43, 279–293.CrossRefGoogle Scholar
Paola, C. 2000. Quantitative models of sedimentary basin filling. Sedimentology 47 (Supplement 1), 121–178.CrossRefGoogle Scholar
Paola, C., & Borgman, L. 1991. Reconstructing random topography from preserved stratification. Sedimentology 38, 553–565.CrossRefGoogle Scholar
Paola, C., Heller, P. L., & Angevine, C. L. 1992. The large-scale dynamics of grain-size variation in alluvial basins. 1 – theory. Basin Research 4, 73–90.CrossRefGoogle Scholar
Paola, C., Parker, G., Mohrig, D. C., & Whipple, K. X. 1999. The influence of transport fluctuations on spatially averaged topography on a sandy, braided fluvial plain. In Numerical Experiments in Stratigraphy: Recent Advances in Stratigraphic and Sedimentologic Computer Simulations (ed. Harbaugh, J. W., Watney, W. L., Rankey, E. C.et al.), pp. 211–218. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Pantin, H. M. 1979. Interaction between velocity and effective density in turbidity flow: phase plane analysis, with criteria for autosuspension. Marine Geology 31, 59–99.CrossRefGoogle Scholar
Pantin, H. M.2001. Experimental evidence for autosuspension. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 189–205. Oxford: Blackwells.CrossRefGoogle Scholar
Pantin, H. M., & Leeder, M. R. 1987. Reverse flow in turbidity currents: the role of internal solitons. Sedimentology 34, 1143–1155.CrossRefGoogle Scholar
Parker, G. 1978a. Self-formed straight rivers with equilibrium banks and mobile bed. 1. The sand–silt river. Journal of Fluid Mechanics 89, 109–126.CrossRefGoogle Scholar
Parker, G. 1978b. Self-formed straight rivers with equilibrium banks and mobile bed. 2. The gravel river. Journal of Fluid Mechanics 89, 127–146.CrossRefGoogle Scholar
Parker, G. 1979. Hydraulic geometry of active gravel rivers. Journal of the Hydraulic Division, ASCE 105, 1185–1201.Google Scholar
Parker, G., Fukushima, Y., & Pantin, H. M. 1986. Self-accelerating turbidity currents. Journal of Fluid Mechanics 171, 145–181.CrossRefGoogle Scholar
Parker, R. S. 1977. Experimental study of drainage basin evolution and its hydrologic implications. Colorado State University, Fort Collins, Colorado, Hydrology Paper 90.
Parrish, J. T. 1993. Climate of the supercontinent Pangea. Journal of Geology 101, 215–233.CrossRefGoogle Scholar
Parrish, J. T. 1998. Interpreting Pre-Quaternary Climate from the Geological Record. New York: Columbia University Press.Google Scholar
Parrish, J. T., & Barron, E. J. 1986. Paleoclimates and Economic Geology. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Paterson, W. S. B. 1994. The Physics of Glaciers, 3rd edn. Oxford: Pergamon.Google Scholar
Patterson, R. J., & Kinsman, D. J. J. 1982. Formation of diagenetic dolomite in coastal sabkha along Arabian (Persian) Gulf. American Association of Petroleum Geologists Bulletin 66, 28–43.Google Scholar
Paul, E. A., & Clark, F. E. 1989. Soil Microbiology and Biochemistry. San Diego, CA: Academic Press.Google Scholar
Paul, M. A., & Eyles, N. 1990. Constraints on the preservation of diamict facies (melt-out tills) at the margins of stagnant glaciers. Quaternary Science Reviews 9, 51–69.CrossRefGoogle Scholar
Paxton, S. T., Szabo, J. O., Ajdukiewicz, J. M., & Klimentidis, R. E. 2002. Construction of an intergranular volume compaction curve for evaluating and predicting compaction and porosity loss in rigid-grain sandstone reservoirs. American Association of Petroleum Geologists Bulletin 86, 2047–2067.Google Scholar
Peakall, J. 1998. Axial river evolution in response to half-graben faulting: Carson River, Nevada. Journal of Sedimentary Research 68, 788–799.CrossRefGoogle Scholar
Peakall, J., Felix, M., McCaffrey, B., & Kneller, B. 2001. Particulate gravity currents: perspectives. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 1–8. Oxford: Blackwells.CrossRefGoogle Scholar
Peakall, J., Leeder, M., Best, J., & Ashworth, P. 2000a. River response to lateral ground tiltng: a synthesis and some implications for the modelling of alluvial architecture in extensional basins. Basin Research 12, 413–424.CrossRefGoogle Scholar
Peakall, J., McCaffrey, B., & Kneller, B. 2000b. A process model for the evolution, morphology, and architecture of sinuous submarine channels. Journal of Sedimentary Research 70, 434–448.CrossRefGoogle Scholar
Pemberton, S. G., MacEachern, J. A., & Frey, R. W. 1992. Trace fossil facies models: environmental and allostratigraphic significance. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 47–72. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Penland, S., Boyd, R., & Suter, J. R. 1988. Transgressive depositional systems of the Mississippi delta plain: a model for barrier shoreline and shelf sand development. Journal of Sedimentary Petrology 58, 932–949.Google Scholar
Pennisi, E. 2004. The secret life of fungi. Science 304, 1620–1622.CrossRefGoogle ScholarPubMed
Pérez-Arlucea, M., & Smith, N. D. 1999. Depositional patterns following the 1870's avulsion of the Saskatchewan River (Cumberland Marshes, Saskatchewan). Journal of Sedimentary Research 69, 62–73.CrossRefGoogle Scholar
Perillo, G. M. E. 1995. Geomorphology and Sedimentology of Estuaries. Amsterdam: Elsevier.Google Scholar
Pernetta, J. 1994. Atlas of the Oceans. London: Mitchell Beazly.Google Scholar
Person, M., & Garven, G. 1992. Hydrologic constraints on petroleum generation within continental rift basins: theory and application to the Rhine Graben. American Association of Petroleum Geologists Bulletin 76, 468–488.Google Scholar
Person, M., Raffensperger, J. P., Ge, S., & Garven, G. 1996. Basin-scale hydrogeologic modeling. Reviews of Geophysics 34, 61–87.CrossRefGoogle Scholar
Peryt, T. 1983. Coated Grains. New York: Springer-Verlag.CrossRefGoogle Scholar
Peterson, M. N. A., & der Borch, C. C. 1965. Chert: modern inorganic deposition in a carbonate-precipitating locality. Science 149, 1501–1503.CrossRefGoogle Scholar
Petit, J. R., Jouzel, J., Raynaud, D.et al. 1999. Climate and atmospheric history of the past 420,000 years from the Vostock ice core, Antarctica. Nature 399, 429–436.CrossRefGoogle Scholar
Pettijohn, F. J. 1975. Sedimentary Rocks, 3rd edn. New York: Harper and Row.Google Scholar
Pettijohn, F. J., & Potter, P. E. 1964. Atlas and Glossary of Primary Sedimentary Structures. New York: Springer-Verlag.CrossRefGoogle Scholar
Pettijohn, F. J., Potter, P. E., & Siever, R. 1972. Sand and Sandstone. New York: Springer-Verlag.Google Scholar
Petts, G., & Calow, P. 1996. River Flows and Channel Forms. Oxford: Blackwell Science Limited.Google Scholar
Phillips, O. M. 1991. Flow and Reactions in Permeable Rocks. Cambridge: Cambridge University Press.Google Scholar
Pickard, G. L., & Emery, W. J. 1990. Descriptive Physical Oceanography, 5th edn. Oxford: Pergamon Press.Google Scholar
Pickering, K. T., & Hiscott, R. N. 1985. Contained (reflected) turbidity currents from the Middle Ordovician Cloridorme Formation, Quebec, Canada: an alternative to the antidune hypothesis. Sedimentology 32, 373–394.CrossRefGoogle Scholar
Pickering, K. T., Hiscott, R. N., & Hein, F. J. 1989. Deep Marine Environments: Clastic Sedimentation and Tectonics. London: Unwin Hyman.Google Scholar
Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, Ricci F., & Smith, R. D. A. 1995. Atlas of Deep Water Environments: Architectural Style in Turbidite Systems. London: Chapman and Hall.CrossRefGoogle Scholar
Pickering, K. T., Soh, W., & Tiara, A. 1991. Scale of tsunami-generated sedimentary structures in deep water. Journal of the Geological Society of London 148, 211–214.CrossRefGoogle Scholar
Pickering, K. T., Stow, D. A. V., Watson, M. P., & Hiscott, R. N. 1986. Deep-water facies, processes and models: a review and classification scheme for modern and ancient sediments. Earth-Science Reviews 23, 75–174.CrossRefGoogle Scholar
Pierson, T. C. 1981. Dominant particle support mechanisms in debris flows at Mt Thomas, New Zealand, and implications for flow mobility. Sedimentology 28, 49–60.CrossRefGoogle Scholar
Pierson, T. C. 1995. Flow characteristics of large eruption-triggered debris flows at snow-clad volcanoes: constraints for debris-flow models. Journal of Volcanology and Geothermal Research 66, 283–294.CrossRefGoogle Scholar
Pierson, T. C., & Scott, K. M. 1985. Downstream dilution of a lahar: transition from debris flow to hyperconcentrated streamflow. Water Resources Research 21, 1511–1524.CrossRefGoogle Scholar
Pinet, P. R. 2006. Invitation to Oceanography, 4th edn. Boston: Jones and Bartlett.Google Scholar
Piper, D. J. W., Cochonat, P., & Morrison, M. L. 1999. The sequence of events around the epicenter of the 1929 Grand Banks earthquake: initiation of debris flow and turbidity current inferred from sidescan sonar. Sedimentology 46, 79–97.CrossRefGoogle Scholar
Piper, D. J. W., & Normark, W. R. 1983. Turbidite depositional patterns and flow characteristics, Navy submarine fan, California Borderland. Sedimentology 30, 681–694.CrossRefGoogle Scholar
Pitzer, K. 1973. Thermodynamics of electrolytes. I. Theoretical basis and general equations. Journal of Physical Chemistry 77, 268–277.CrossRefGoogle Scholar
Playford, P. E. 1984. Platform–margin and marginal–slope relationships in the Devonian reef complexes of the Canning Basin. In The Canning Basin, Western Australia (ed. Purcell, P. G.), pp. 189–214. Perth: Geological Society of Australia and Petroleum Exploration Society of Australia.Google Scholar
Playford, P. E., & Cockbain, A. E. 1976. Modern algal stromatolites at Hamelin Pool, a hypersaline barred basin, Western Australia. In Stromatolites (ed. Walter, M. R.), pp. 389–411. Amsterdam: Elsevier.Google Scholar
Plummer, L. N., Parkhurst, D. L., Fleming, G. W., & Dunkle, S. A. 1988. A computer program incorporating Pitzer's equations for calculation of geochemical reactions in brines. United States Geological Survey Water-resources Investigations Report 88–4153.
Pond, S., & Pickard, G. L. 1983. Introductory Dynamical Oceanography, 2nd edn. London: Pergamon.Google Scholar
Pope, M. C., & Grotzinger, J. P. 2000. Controls on fabric development and morphology of tufas and stromatolites, Uppermost Pethei Group (1.8 Ga), Great Slave Lake, Northwest Canada. In Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World (ed. Grotzinger, J. P. & James, N. P.), pp. 103–121. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Portman, C. P., Andrews, J. E., Rowe, P. J., Leeder, M. R., & Hoodewerff, J. 2005. Submarine-spring controlled calcification and growth of large Rivularia bioherms, Late Pleistocene (MIS 5e), Gulf of Corinth, Greece. Sedimentology 52, 441–465.CrossRefGoogle Scholar
Posamentier, H. W., & Allen, G. P. 1999. Siliciclastic Sequence Stratigraphy – Concepts and Applications. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Posamentier, H. W., Allen, G. P., James, D. P., & Tesson, M. 1992. Forced regressions in a sequence stratigraphic framework: concepts, examples, and exploration significance. American Association of Petroleum Geologists Bulletin 76, 1687–1709.Google Scholar
Posamentier, H. W., Erskine, R. D., & Mitchum, R. M. 1991. Models for submarine fan deposition within a sequence stratigraphic framework. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 127–136. New York: Springer-Verlag.CrossRefGoogle Scholar
Posamentier, H. W., Jervey, M. T., & Vail, P. R. 1988. Eustatic controls on clastic deposition I – conceptual framework. In Sea Level Changes; An Integrated Approach (ed. Wilgus, C. K., Hastings, B. S., St., C. G.Kendall, C.et al.), pp. 109–124. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Posamentier, H. W., & Kolla, V. 2003. Seismic geomorphology and stratigraphy of depositional elements in deep-water settings. Journal of Sedimentary Research 73, 367–388.CrossRefGoogle Scholar
Posamentier, H. W., Summerhayes, C. P., Haq, B. U., & Allen, G. P. 1993. Sequence Stratigraphy and Facies Associations. Oxford: Blackwells.CrossRefGoogle Scholar
Posamentier, H. W., & Walker, R. G. 2006. Deep-water turbidites and submarine fans. In Facies Models Revisited (ed. Posamentier, H. W. & Walker, R. G.), pp. 397–520. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Post, A., & LaChapelle, E. R. 2000. Glacier Ice. Seattle, WA: University of Washington Press.Google Scholar
Postma, G. 1990. Depositional architecture and facies of river and fan deltas: a synthesis. In Coarse-Grained Deltas (ed. Colella, A. & Prior, D. B.), pp. 13–27. Oxford: Blackwells.CrossRefGoogle Scholar
Postma, G., Nemec, W., & Keinspehn, K. L. 1988. Large floating clasts in turbidites: a mechanism for their emplacement. Sedimentary Geology 58, 47–61.CrossRefGoogle Scholar
Potter, P. E., Maynard, J. B., & Depetris, P. J. 2005. Mud and Mudstone. Berlin: Springer-Verlag.Google Scholar
Powell, R. D. 1990. Glacimarine processes at grounding-line fans and their growth to ice-contact deltas. In Glacimarine Environments (ed. Dowdeswell, J. A. & Scourse, J. D.), pp. 53–73. London: Geological Society of London.Google Scholar
Powell, R. D., & Molnia, B. F. 1989. Glacimarine sedimentary processes, facies and morphology of the south-southeast Alaska shelf and fiords. Marine Geology 85, 359–390.CrossRefGoogle Scholar
Pratt, B. R., & James, N. P. 1982. Crypt-algal–metazoan bioherms of early Ordovician age in the St. George Group, western Newfoundland. Sedimentology 29, 313–343.CrossRefGoogle Scholar
Pratt, B. R., James, N. P., & Cowan, C. A. 1992. Peritidal carbonates. In Facies Models, Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 303–323. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Press, F., & Siever, R. 1978. Earth, 2nd edn. New York: W. H. Freeman and Company.Google Scholar
Press, F., Siever, R., Grotzinger, J., & Jordan, T. H. 2004. Understanding Earth, 4th edn. New York: W. H. Freeman and Company.Google Scholar
Prior, D. B., Bornhold, B. D., Wiseman, W. J., & Lowe, D. R. 1987. Turbidity current activity in a British Columbia fjord. Science 237, 1330–1333.CrossRefGoogle Scholar
Prior, D. B., & Coleman, J. M. 1979. Submarine landslides – geometry and nomenclature. Zeitschrift für Geomorphologie 23, 415–426.Google Scholar
Prothero, D. R. 2004. Bringing Fossils to Life: An Introduction to Paleobiology, 2nd edn. Boston, MA: McGraw-Hill.Google Scholar
Pryor, W. A. 1975. Biogenic sedimentation and alteration of argillaceous sediments in shallow marine environments. Geological Society of America Bulletin 86, 1244–1254.2.0.CO;2>CrossRefGoogle Scholar
Purser, B. H. 1973. The Persian Gulf: Holocene Carbonate Sedimentation and Diagenesis in a Shallow Epicontinental Sea. New York: Springer-Verlag.CrossRefGoogle Scholar
Purser, B. H., & Evans, G. 1973. Regional sedimentation along the Trucial Coast, SE Persian Gulf. In The Persian Gulf: Holocene Carbonate Sedimentation and Diagenesis in a Shallow Epicontinental Sea (ed. Purser, B. H.), pp. 211–231. New York: Springer-Verlag.CrossRefGoogle Scholar
Putnis, A. 2002. Mineral replacement reactions: from macroscopic observations to microscopic mechanisms. Mineralogical Magazine 66, 689–708.CrossRefGoogle Scholar
Pye, K. 1987. Aeolian Dust and Dust Deposits. London: Academic Press.Google Scholar
Pye, K.1993a. Late Quaternary development of coastal parabolic megadune complexes in northeastern Australia. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 23–44. Oxford: Blackwells.CrossRefGoogle Scholar
Pye, K. 1993b. The Dynamics and Environmental Context of Aeolian Sedimentary Systems. London: Geological Society of London.Google Scholar
Pye, K., & Lancaster, N. 1993. Aeolian Sediments: Ancient and Modern. Oxford: Blackwells.CrossRefGoogle Scholar
Pye, K., & Tsoar, H. 1990. Aeolian Sand and Sand Dunes. London: Chapman and Hall.CrossRefGoogle Scholar
Quinlan, G. M., & Beaumont, C. 1984. Appalachian thrusting, lithospheric flexure, and the Paleozoic stratigraphy of the eastern interior of North America. Canadian Journal of Earth Sciences 21, 973–996.CrossRefGoogle Scholar
Rahmani, R. A., & Flores, R. M. 1984. Sedimentology of Coal and Coal-bearing Sequences. Oxford: Blackwells.Google Scholar
Railsback, L. B. 2003. Stylolites. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 690–692. Dordrecht: Kluwer Academic Publishers.Google Scholar
Rampino, M. R., & Sanders, J. E. 1980. Holocene transgression in south-central Long Island, New York. Journal of Sedimentary Petrology 50, 1063–1080.Google Scholar
Rasmussen, B., Bengston, S., Fletcher, I. R., & McNaughton, N. J. 2002. Discoidal impressions and trace-like fossils more than 1200 million years old. Science 296, 1112–1115.CrossRefGoogle ScholarPubMed
Raymo, M. E. & Ruddiman, W. F. 1992. Tectonic forcing of late Cenozoic climate. Nature 359, 117–122.CrossRefGoogle Scholar
Raymond, C. F. 1987. How do glaciers surge? A review. Journal of Geophysical Research 92, 9121–9134.CrossRefGoogle Scholar
Read, J. F. 1985. Carbonate platform facies models. Bulletin of the American Association of Petroleum Geologists 66, 860–878.Google Scholar
Read, J. F. 1973. Carbonate cycles, Pillara Formation (Devonian), Canning Basin, Western Australia. Bulletin of Canadian Petroleum Geology 16, 649–653.Google Scholar
Read, J. F.1974. Calcrete deposits and Quaternary sediments, Edel Province, Shark Bay, Western Australia. In Evolution and Diagenesis of Quaternary Carbonate Sequences, Shark Bay, Western Australia (ed. Logan, B. W., Read, J. F., Hagan, G. M.et al.), pp. 250–282. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Read, W. A. 1994. High-frequency, glacial–eustatic sequences in early Namurian coal-bearing fluviodeltaic deposits, central Scotland. In Orbital Forcing and Cyclic Sequences (ed. Boer, P. L. & Smith, D. G.), pp. 413–428. Oxford: Blackwells.CrossRefGoogle Scholar
Reading, H. G. (ed.) 1986. Sedimentary Environments and Facies, 2nd edn. Oxford: Blackwells.Google Scholar
Reading, H. G. 1996. Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. Oxford: Blackwells.Google Scholar
Reading, H. G., & Collinson, J. D. 1996. Clastic coasts. In Sedimentary Environments: Processes, Facies and Stratigraphy (ed. Reading, H. G.), pp. 154–231. Oxford: Blackwells.Google Scholar
Reading, H. G., & Richards, M. 1994. Turbidite systems in deep-water basin margins classified by grain size and feeder system. American Association of Petroleum Geologists Bulletin 78, 792–822.Google Scholar
Reeder, R. J. 1983. Carbonates: Mineralogy and Chemistry. Chelsea, MI: Mineralogical Society of America.Google Scholar
Reesink, A. J. H., & Bridge, J. S. 2007. Influence of superimposed bedforms and flow unsteadiness on formation of cross strata in dunes and unit bars. Sedimentary Geology doi:10.1016/j.sedgeo.2007.02.005.CrossRef
Reeves, C. C. 1977. Caliche: Origin, Classification, Morphology and Uses. Lubbock, TX: Escado Books.Google Scholar
Reid, R. P., MacIntyre, I. G., & James, N. P. 1990. Internal precipitation of microcrystalline carbonate: a fundamental problem for sedimentologists. Sedimentary Geology 68, 163–170.CrossRefGoogle Scholar
Reineck, H.-E., & Singh, I. B. 1980. Depositional Sedimentary Environments: With Special Reference to Terrigenous Clastics. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Reineck, H.-E., & Wunderlich, F. 1968a. Zur Unterscheidung von asymmetrischen Oszillationripplen und Stromungsripplen. Senckenbergiana Lethaia 49, 321–345.Google Scholar
Reineck, H.-E., & Wunderlich, F. 1968b. Classification and origin of flaser and lenticular bedding. Sedimentology 11, 99–104.CrossRefGoogle Scholar
Reinharz, E., Nilsen, K. J., Boesch, D. F., Bertelsen, R., & O'Connell, A. E. 1982. A Radiographic Examination of Physical and Biogenic Sedimentary Structures in the Chesapeake Bay. Baltimore, MD: Maryland Geological Survey.Google Scholar
Reinson, G. E. 1992. Transgressive barrier island and estuarine systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 179–194. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Reitner, J. 1993. Modern cryptic microbialite/metazoan facies from Lizard Island (Great Barrier Reef, Australia): formation and concepts. Facies 29, 3–39.CrossRefGoogle Scholar
Reitner, J., & Neuweiler, F. 1995. Mud mounds: a polygenetic spectrum of fine-grained carbonate buildups. Facies 32, 1–70.Google Scholar
Renard, F., & Dysthe, D. 2003. Pressure solution. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 542–544. Dordrecht: Kluwer Academic Publishers.Google Scholar
Retallack, G. J. 1997. A Color Guide to Paleosols. Chichester: Wiley.Google Scholar
Retallack, G. J. 2001. Soils of the Past, 2nd edn. Oxford: Blackwells.CrossRefGoogle Scholar
Retallack, G. J.2003. Weathering, soils, and paleosols. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 770–776. Dordrecht: Kluwer Academic Publishers.Google Scholar
Retallack, G. J. 2005. Pedogenic carbonate proxies for amount and seasonality of precipitation in paleosols. Geology 33, 333–336.CrossRefGoogle Scholar
Reynaud, J. Y., Tessier, B., Proust, J. N.et al. 1999. Eustatic and hydrodynamic controls on the architecture of a deep shelf sand bank (Celtic Sea). Sedimentology 46, 703–721.CrossRefGoogle Scholar
Rhodes, B., Tuttle, M., Horton, B.et al. 2006. Paleotsunami research. EOS 87, 205.CrossRefGoogle Scholar
Rhodes, E. G., & Moslow, T. F. 1993. Marine Clastic Reservoirs: Examples and Analogues. New York: Springer-Verlag.CrossRefGoogle Scholar
Rice, S. P., & Church, M. 2001. Longitudinal profiles in simple alluvial systems. Water Resources Research 37, 417–426.CrossRefGoogle Scholar
Lucci, Ricci F. 1995. Sedimentographica: Photographic Atlas of Sedimentary Structures, 2nd edn. New York: Columbia University Press.Google Scholar
Richardson, K., & Carling, P. A. 2005. A typology of sculpted forms in open bedrock channels. Geological Society of America Special Paper 392.
Ries, J. R. 2004. Effect of ambient Mg/Ca ratio on Mg fractionation in calcareous marine invertebrates: a record of the oceanic Mg/Ca ratio over the Phanerozoic. Geology 32, 981–984.CrossRefGoogle Scholar
Riding, R. 1979. Origin and diagenesis of lacustrine algal bioherms at the margin of the Ries Crater, Upper Miocene, southern Germany. Sedimentology 26, 645–680.CrossRefGoogle Scholar
Riding, R. 2000. Microbial carbonates: the geologic record of calcified bacterial–algal mats and biofilms. Sedimentology 47, 179–214.CrossRefGoogle Scholar
Rinaldo, A., Rodriguez-Iturbe, I., & Rigon, R. 1998. Channel networks. Annual Review of Earth and Planetary Sciences 26, 289–327.CrossRefGoogle Scholar
Rinaldo, A., Rodriguez-Iturbe, I., Rigon, R.et al. 1992. Minimum energy and fractal structures of drainage networks. Water Resources Research 28, 2183–2195.CrossRefGoogle Scholar
Riggs, S. D. 1984. Paleoceanographic model of Neogene phosphorite deposition, U.S. Atlantic continental margin. Science 223, 123–131.CrossRefGoogle ScholarPubMed
Rine, J. M., & Ginsburg, R. N. 1985. Depositional facies of a mud shoreface in Suriname, South America. A mud analogue to sandy, shallow-marine deposits. Journal of Sedimentary Petrology 55, 633–652.Google Scholar
Ritter, D. F., Kochel, R. C., & Miller, J. R. 2002. Process Geomorphology, 4th edn. New York: McGraw-Hill.Google Scholar
Roberson, H. E., & Lahann, R. W. 1981. Smectite to illite conversion rates: effects of solution chemistry. Clay and Clay Minerals 29, 129–135.CrossRefGoogle Scholar
Roberts, H. H., Adams, R. D., & Cunningham, R. H. W. 1980. Evolution of the sand-dominant subaerial phase, Atchafalaya Delta, Louisiana. American Association of Petroleum Geologists Bulletin 64, 264–279.Google Scholar
Roberts, J. A., Bennett, P. C., González, L. A., Macpherson, G. L., & Milliken, K. L. 2004. Microbial precipitation of dolomite in methanogenic groundwater. Geology 32, 277–280.CrossRefGoogle Scholar
Robie, R. A., & Hemingway, B. S. 1995. Thermodynamic Properties of Minerals and Related Substances at 298.15 K and 1 bar (105 Pascals) Pressure and Higher Temperatures. Washington, D.C.: United States Geological Survey.Google Scholar
Robinson, R. L., & Slingerland, R. L. 1998a. Origin of fluvial grain-size trends in a foreland basin: the Pocono Formation of the central Appalachian basin. Journal of Sedimentary Research A68, 473–486.CrossRefGoogle Scholar
Robinson, R. L., & Slingerland, R. L. 1998b. Grain-size trends and basin subsidence in the Campanian Castlegate Sandstone and equivalent conglomerates of central Utah. Basin Research 10, 109–127.CrossRefGoogle Scholar
Robinson, R. A. J., Slingerland, R. L., & Walsh, J. M. 2001. Predicting fluvial–deltaic aggradation in Lake Roxburgh, New Zealand: test of a water and sediment routing model. In Geologic Modeling and Simulation: Sedimentary Systems (ed. Merriam, D. F. & Davis, J. C.), pp. 119–132. New York: Kluwer Academic/Plenum.CrossRefGoogle Scholar
Robock, A. 2002. The climatic aftermath. Science 295, 1242–1244.CrossRefGoogle ScholarPubMed
Rodriguez, A. B., Hamilton, M. D., & Anderson, J. B. 2000. Facies and evolution of the modern Brazos delta, Texas; wave versus flood influence. Journal of Sedimentary Research 70, 283–295.CrossRefGoogle Scholar
Rodriguez-Iturbe, I., & Rinaldo, A. 1997. Fractal River Basins: Chance and Self-organization. Cambridge: Cambridge University Press.Google Scholar
Rodriguez-Iturbe, I., Rinaldo, A., Rigon, R.et al. 1992. Energy dissipation, runoff production, and the three-dimensional structure of river basins. Water Resources Research 28, 1095–1103.CrossRefGoogle Scholar
Roscoe, R. 1953. Suspensions. In Flow Properties of Disperse Systems (ed. Hermans, J. J.), pp. 1–38. New York: Wiley Interscience.Google Scholar
Rosgen, D. L. 1994. A classification of natural rivers. Catena 22, 169–199.CrossRefGoogle Scholar
Rosgen, D. L. 1996. Applied River Morphology. Fort Collins, CO: Wildland Hydrology.Google Scholar
Ross, D. A., & Degens, E. T. 1969. Shipboard collection and preservation of sediment samples collected during CHAIN 61 from the Red Sea. In Hot Brines and Recent Heavy Metal Deposits in the Red Sea: A Geochemical and Geophysical Account (ed. Degens, E. T. & Ross, D. A.), pp. 363–367. New York: Springer-Verlag.CrossRefGoogle Scholar
Rothlisberger, H., & Lang, H. 1987. Glacial hydrology. In Glacio-Fluvial Sediment Transfer – An Alpine Perspective (ed. Gurnell, A. M. & Clark, M. J.), pp. 207–284. Chichester: Wiley.Google Scholar
Rowley, D. B. 2002. Rate of plate creation and destruction: 180 Ma to present. Geological Society of America Bulletin 114, 927–933.2.0.CO;2>CrossRefGoogle Scholar
Rubin, D. M. 1987. Cross-Bedding, Bedforms and Paleocurrents. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Rubin, D. M. 1990. Lateral migration of linear dunes in the Strzelecki Desert, Australia. Earth Surface Processes and Landforms 15, 1–14.CrossRefGoogle Scholar
Rubin, D. M., & Hunter, R. E. 1982. Bedform climbing in theory and nature. Sedimentology 29, 121–138.CrossRefGoogle Scholar
Rubin, D. M., & Hunter, R. E. 1985. Why deposits of longitudinal dunes are rarely recognized in the rock record. Sedimentology 32, 147–157.CrossRefGoogle Scholar
Rubin, D. M., & Hunter, R. E. 1987. Bedform alignment in directionally varying flows. Science 237, 276–278.CrossRefGoogle ScholarPubMed
Rust, B. R. 1978. A classification of alluvial channel systems. In Fluvial Sedimentology (ed. Miall, A. D.), pp. 187–198. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Rust, B. R., & Romanelli, R. 1975. Late Quaternary subaqueous outwash deposits near Ottawa, Canada. In Glaciofluvial and Glaciolacustrine Sediments (ed. Jopling, A. V. & McDonald, B. C.), pp. 177–192. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Ryan, W. B. F., & Hsü, K. J. 1973. Initial Reports of the Deep Sea Drilling Project, Vol. 13. Washington, D.C.: U.S. Government Printing Office.CrossRefGoogle Scholar
Sallenger, A. 1979. Inverse grading and hydraulic equivalence in grain-flow deposits. Journal of Sedimentary Petrology 49, 553–562.Google Scholar
Salter, T. 1993. Fluvial scour and incision: models for their influence on the development of realistic reservoir geometries. In Characterization of Fluvial and Aeolian Reservoirs (ed. North, C. P. & Prosser, D. J.), pp. 33–51. London: Geological Society of London.Google Scholar
Sandberg, P. A. 1975. New interpretations of Great Salt Lake ooids and of ancient non-skeletal carbonate mineralogy. Sedimentology 22, 497–538.CrossRefGoogle Scholar
Sandberg, P. A. 1983. An oscillating trend in Phanerozoic non-skeletal carbonate mineralogy. Nature 305, 19–22.CrossRefGoogle Scholar
Sarg, J. F. 1988. Carbonate sequence stratigraphy. In Sea Level Changes: An Integrated Approach (ed. Wilgus, C. K., Hastings, B. S., C. Kendall, C. G. St.et al.), pp. 155–182. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Sarna-Wojcicki, A. M., & Davis, J. O. 1991. Quaternary tephrochronology. In Quaternary Nonglacial Geology, Conterminous U.S. (ed. Morrison, R. B.), pp. 93–116. Boulder, CO: Geological Society of America.Google Scholar
Satake, K. 2005. Tsunamis: Case Studies and Recent Developments. Dordrecht: Springer-Verlag.CrossRefGoogle Scholar
Saucier, R. T. 1994. Geomorphology and Quaternary Geologic History of the Lower Mississippi Valley. Vicksburg, VA: Mississippi River Commission.Google Scholar
Saunders, I., & Young, A. 1983. Rates of surface processes on slopes, slope retreat and denudation. Earth Surface Processes and Landforms 8, 473–501.CrossRefGoogle Scholar
Saunderson, H. C., & Lockett, F. P. 1983. Flume experiments on bedforms and structures at the dune-plane bed transition. In Modern and Ancient Fluvial Systems (ed. Collinson, J. D. & Lewin, J.), pp. 49–58. Oxford: Blackwells.CrossRefGoogle Scholar
Savage, S. B. 1979. Gravity flow of cohesionless granular materials in chutes and channels. Journal of Fluid Mechanics 92, 53–96.CrossRefGoogle Scholar
Schenk, C. J., Gautier, D. L., Olhoeft, G. R., & Lucius, J. E. 1993. Internal structure of an aeolian dune using ground-penetrating radar. In Aeolian Sediments: Ancient and Modern (ed. Pye, K. & Lancaster, N.), pp. 61–70. Oxford: Blackwells.CrossRefGoogle Scholar
Schlager, W. 2005. Carbonate Sedimentology and Sequence Stratigraphy. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Schlager, W., & Bolz, H. 1977. Clastic accumulation of sulphate evaporites in deep water. Journal of Sedimentary Petrology 47, 600–609.Google Scholar
Schlager, W., & Ginsburg, R. N. 1981. Bahamian carbonate platforms – the deep and the past. Marine Geology 44, 1–24.CrossRefGoogle Scholar
Schlichting, H. 1979. Boundary Layer Theory. New York: McGraw-Hill.Google Scholar
Schminke, H.-U. 2004. Volcanism. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Schneidermann, N., & Harris, P. M. 1985. Carbonate Cements. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Scholle, P. A. 1978. A Color Illustrated Guide to Carbonate Rock Constituents, Textures, Cements, and Porosities. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A. 1979. A Color Guide to Constituents, Textures, Cements, and Porosities of Sandstones and Associated Rocks. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A., Arthur, M. A, & Ekdale, A. A. 1983. Pelagic environments. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 620–691. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A., & Halley, R. B. 1985. Burial diagenesis: out of sight, out of mind! In Carbonate Cements (ed. Schneidermann, N. & Harris, P. M.), pp. 309–334. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Scholle, P. A., & Spearing, D. R. 1982. Sandstone Depositional Environments. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Scholle, P. A., & Ulmer-Scholle, D. S. 2003. A Color Guide to the Petrography of Carbonate Rocks: Grains, Textures, Porosity, Diagenesis. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Schopf, J. W. 1983. Earth's Earliest Biosphere. Princeton, NJ: Princeton University Press.Google Scholar
Schubel, K. A., & Simonson, B. M. 1990. Petrography and diagenesis of cherts from Lake Magadi, Kenya. Journal of Sedimentary Petrology 60, 761–776.Google Scholar
Schulz, H. D., & Zabel, M. 2000. Marine Geochemistry. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Schwartz, R. K. 1982. Bedform and stratification characteristics of some modern small-scale washover sand bodies. Sedimentology 29, 835–849.CrossRefGoogle Scholar
Schumm, S. A. 1977. The Fluvial System. New York: Wiley.Google Scholar
Schumm, S. A. 1993. River response to baselevel change: implications for sequence stratigraphy. Journal of Geology 101, 279–294.CrossRefGoogle Scholar
Schumm, S. A., Dumont, J. F., & Holbrook, J. M. 2000. Active Tectonics and Alluvial Rivers. Cambridge: Cambridge University Press.Google Scholar
Schumm, S. A., & Lichty, R. W. 1963. Channel widening and floodplain construction, Cimarron River, Kansas. U.S. Geological Survey Professional Paper352-D, pp. 71–88.
Schumm, S. A., Mosley, M. P., & Weaver, W. E. 1987. Experimental Fluvial Geomorphology. New York: Wiley.Google Scholar
Scruton, P. C. 1960. Delta building and the deltaic sequence. In Recent Sediments, Northwest Gulf of Mexico (ed. Shepard, F. P., Phleger, F. B., & Andel, T. H.), pp. 82–102. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Seifert, D., & Jensen, J. L. 1999. Using sequential indicator simulation as a tool in reservoir description: issues and uncertainties. Mathematical Geology 31, 527–550.CrossRefGoogle Scholar
Seifert, D., & Jensen, J. L. 2000. Object and pixel-based reservoir modeling of a braided fluvial reservoir. Mathematical Geology 32, 581–603.CrossRefGoogle Scholar
Seilacher, A, Bose, P. K., & Pflüger, F. 1998. Triploblastic animals more than 1 billion years ago: trace fossil evidence from India. Science 282, 80–83.CrossRefGoogle ScholarPubMed
Selby, M. J. 1993. Hillslope Materials and Processes, 2nd edn. Oxford: Oxford University Press.Google Scholar
Semikhatov, M. A., Gebelein, C. D., Cloud, P., Awramik, S. M., & Benmore, W. C. 1979. Stromatolite morphogenesis – progress and problems. Canadian Journal of Earth Sciences 16, 992–1015.CrossRefGoogle Scholar
Sepkoski, J. J. Jr. 1981. A factor analytic description of the Phanerozoic marine record. Paleobiology 7, 36–53.CrossRefGoogle Scholar
Sha, L. P., & de Boer, P. L. 1991. Ebb-tidal delta deposits along the west Friesian islands (The Netherlands): processes, facies architecture and preservation. In Clastic Tidal Sedimentology (ed. Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A.), pp. 199–218. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Shanley, K. W., & McCabe, P. J. 1993. Alluvial architecture in a sequence stratigraphic framework: a case history from the Upper Cretaceous of southern Utah, USA. In The Geological Modeling of Hydrocarbon Reservoirs and Outcrop Analogues (ed. Flint, S. & Bryant, I. D.), pp. 21–56. Oxford: Blackwells.Google Scholar
Shanley, K. W., & McCabe, P. J. 1994. Perspectives on the sequence stratigraphy of continental strata. American Association of Petroleum Geologists Bulletin 78, 544–568.Google Scholar
Shanmugan, G. 1996. High-density turbidity currents: are they sandy debris flows?Journal of Sedimentary Research 66, 2–10.CrossRefGoogle Scholar
Shanmugan, G. 1997. The Bouma sequence and the turbidite mind set. Earth-Science Reviews 42, 201–229.CrossRefGoogle Scholar
Shanmugan, G., & Moiola, R. J. 1988. Submarine fans: characteristics, models, classification and reservoir potential. Earth-Science Reviews 24, 383–428.CrossRefGoogle Scholar
Sharp, M. 1988a. Surging glaciers: behaviour and mechanisms. Progress in Physical Geography 12, 349–370.CrossRefGoogle Scholar
Sharp, M. 1988b. Surging glaciers: geomorphic effects. Progress in Physical Geography 12, 533–559.CrossRefGoogle Scholar
Sharp, M., Jouzel, J., Hubbard, B., & Lawson, W. 1994. The character, structure and origin of the basal ice layer of a surge-type glacier. Journal of Glaciology 40, 327–340.CrossRefGoogle Scholar
Sharp, M., Richards, K. S., & Tranter, M. 1998. Glacier Hydrology and Hydrochemistry. Chichester: Wiley.Google Scholar
Sharp, R. P. 1963. Wind ripples. Journal of Geology 71, 617–636.CrossRefGoogle Scholar
Sharpe, D. R., & Shaw, J. 1989. Erosion of bedrock by subglacial meltwater, Cantley, Quebec. Geological Society of America Bulletin 101, 1011–1020.2.3.CO;2>CrossRefGoogle Scholar
Shaw, J. 2006. A glimpse at meltwater effects associated with continental ice sheets. In Glacier Science and Environmental Change (ed. Knight, P. G.), pp. 25–32. Oxford: Blackwells.CrossRefGoogle Scholar
Shaw, J., Kvill, D., & Rains, B. 1989. Drumlins and catastrophic subglacial floods. Sedimentary Geology 62, 177–202.CrossRefGoogle Scholar
Shaw, J., & Sharpe, D. R. 1987. Drumlin formation by subglacial meltwater erosion. Canadian Journal of Earth Sciences 24, 2316–2322.CrossRefGoogle Scholar
Shear, W. A., & Selden, P. A. 2001. Rustling in the undergrowth: animals in early terrestrial ecosystems. In Plants Invade the Land (ed. Gensel, P. G. & Edwards, D.), pp. 29–51. New York: Columbia University Press.CrossRefGoogle Scholar
Shearman, D. J. 1963. Recent anhydrite, gypsum, dolomite and halite from the coastal flats of Arabian shore of the Persian Gulf. Proceedings of the Geological Society of London 1607, 63–65.Google Scholar
Shearman, D. J., McCugan, A., Stein, C., & Smith, A. J. 1989. Ikaite, CaCO3⋅6H2O, precursor of the thinolites in the Quaternary tufas and tufa mounds of the Lahontin and Mono Lake basins, western United States. Geological Society of America Bulletin 101, 913–917.2.3.CO;2>CrossRefGoogle Scholar
Sheehan, P. M., & Harris, M. T. 2004. Microbialite resurgence after the Late Ordovician extinction. Nature 430, 75–78.CrossRefGoogle ScholarPubMed
Shepard, F. P. 1932. Sediments on continental shelves. Geological Society of America Bulletin 43, 1017–1034.CrossRefGoogle Scholar
Shepard, F. P. 1973. Submarine Geology. New York: Harper and Row.Google Scholar
Shepard, F. P., & Inman, D. L. 1950. Nearshore circulation related to bottom topography and wave refraction. Transactions of the American Geophysical Union 31, 555–565.CrossRefGoogle Scholar
Sheridan, R. E. 1974. Altantic continental margin of North America. In The Geology of Continental Margins (ed. Burke, C. A. & Drake, C. L.), pp. 391–407. New York: Springer-Verlag.CrossRefGoogle Scholar
Shields, A. 1936. Anwendung der Ähnlichkeitsmechanik und der Turbulenzforschung auf die Geschiebebewegung. Mitteilungen der Preuβischen Versuchsanstalt für Wasserbau und Schiffbau 26, 26 pp.Google Scholar
Shiki, T., Cita, M. B., & Gorsline, D. S. 2000. Seismoturbidites, seismites and tsunamiites. Sedimentary Geology (special issue) 135, 1–322.Google Scholar
Shinn, E. A. 1968a. Selective dolomitization of recent sedimentary structures. Journal of Sedimentary Petrology 38, 612–616.Google Scholar
Shinn, E. A. 1968b. Practical significance of birdseye structures in carbonate rocks. Journal of Sedimentary Petrology 38, 611–616.CrossRefGoogle Scholar
Shinn, E. A. 1969. Submarine lithification of Holocene carbonate sediments in the Persian Gulf. Sedimentology 12, 109–144.CrossRefGoogle Scholar
Shinn, E. A. 1983a. Birdseyes, fenestrae, shrinkage pores, and Loferites: a reevaluation. Journal of Sedimentary Petrology 53, 619–628.Google Scholar
Shinn, E. A.1983b. Tidal flats. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 171–210. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Shinn, E. A., & Lidz, B. H. 1988. Blackened limestone pebbles: fire at subaerial unconformities. In Paleokarst (ed. James, N. P. & Choquette, P. W.), pp. 117–131. New York: Springer-Verlag.CrossRefGoogle Scholar
Shinn, E. A., Lloyd, R. M., & Ginsburg, R. N. 1969. Anatomy of a modern carbonate tidal flat, Andros Island, Bahamas. Journal of Sedimentary Petrology 39, 1202–1228.Google Scholar
Shinn, E. A., & Robbin, D. M. 1983. Mechanical and chemical compaction in fine-grained shallow-water limestones. Journal of Sedimentary Petrology 53, 596–618.Google Scholar
Shinn, E. A., Steinen, R. P., Lidz, B. H., & Swart, R. K. 1989. Whitings, a sedimentological dilemma. Journal of Sedimentary Petrology 59, 147–161.CrossRefGoogle Scholar
Shreve, R. L. 1966. Statistical law of stream numbers. Journal of Geology 74, 17–37.CrossRefGoogle Scholar
Shvidchenko, A. B., & Pender, G. 2001. Macroturbulent structure of open-channel flow over gravel beds. Water Resources Research 37, 709–719.CrossRefGoogle Scholar
Sibley, D. F., & Gregg, J. M. 1987. Classification of dolomite rock texture. Journal of Sedimentary Petrology 57, 967–975.Google Scholar
Sidi, F. H., Nummedal, D., Imbert, P., Darman, H., & Posamentier, H. 2003. Tropical Deltas of Southeast Asia – Sedimentology Stratigraphy, and Petroleum Geology. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Siegert, M. J. 2000. Antarctic subglacial lakes. Earth-Science Reviews 50, 29–50.CrossRefGoogle Scholar
Simms, M. A. 1984. Dolomitization by groundwater-flow systems in carbonate platforms. Transactions of the Gulf Coast Association of Geological Societies 34, 411–420.Google Scholar
Simone, L. 1980. Ooids: a review. Earth-Science Reviews 16, 319–335.CrossRefGoogle Scholar
Simonson, B. M. 2003a. Ironstones and iron formations. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 379–385. Dordrecht: Kluwer Academic Publishers.Google Scholar
Simonson, B. M. 2003b. Origin and evolution of large Precambrian iron formations. Geological Society of America Special Paper 370, pp. 231–244.Google Scholar
Simonson, B. M., & Carney, K. E. 1999. Roll-up structures: evidence of in situ microbial mats in Late Archean deep shelf environments. Palaios 14, 13–24.CrossRefGoogle Scholar
Simpson, J. E. 1982. Gravity currents in the laboratory, atmosphere and ocean. Annual Review of Fluid Mechanics 14, 213–234.CrossRefGoogle Scholar
Simpson, J. E. 1997. Gravity Currents in the Environment and the Laboratory, 2nd edn. New York: Cambridge University Press.Google Scholar
Singh, U. 1987. Ooids and cements from the Late Precambrian of the Flinders Range, South Australia. Journal of Sedimentary Petrology 57, 117–127.CrossRefGoogle Scholar
Sinha, S. K., & Parker, G. 1996. Causes of concavity in longitudinal profiles of rivers. Water Resources Research 32, 1417–1428.CrossRefGoogle Scholar
Siringan, F. P., & Anderson, J. B. 1993. Seismic facies architecture and evolution of the Bolivar Roads tidal inlet/delta complex, east Texas Gulf Coast. Journal of Sedimentary Petrology 63, 794–808.Google Scholar
Siringan, F. P., & Anderson, J. B. 1994. Modern shoreface and inner-shelf storm deposits off the East Texas coast, Gulf of Mexico. Journal of Sedimentary Petrology 64, 99–110.Google Scholar
Sleath, J. F. A. 1984. Sea Bed Mechanics. New York: Wiley.Google Scholar
Slingerland, R. L. 1986. Numerical computation of co-oscillating paleotides in the Catskill epeiric sea of eastern North America. Sedimentology 33, 817–829.CrossRefGoogle Scholar
Slingerland, R., Harbaugh, J. W., & Furlong, K. P. 1994. Simulating Clastic Sedimentary Basins. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Slingerland, R. L., & Smith, N. D. 1998. Necessary conditions for a meandering-river avulsion. Geology 26, 435–438.2.3.CO;2>CrossRefGoogle Scholar
Slingerland, R. L., & Smith, N. D. 2004. River avulsions and their deposits. Annual Review of Earth and Planetary Sciences 32, 255–283.CrossRefGoogle Scholar
Sloss, L. L. 1963. Sequences in the cratonic interior of North America. Geological Society of America Bulletin 74, 93–114.CrossRefGoogle Scholar
Smith, C. R. 1996. Coherent flow structures in smooth-wall turbulent boundary layers: facts, mechanisms and speculation. In Coherent Flow Structures in Open Channels (ed. Ashworth, P. J., Bennett, S. J., Best, J. L., & McLelland, S. J.), pp. 1–39. Chichester: Wiley.Google Scholar
Smith, D. G. 1988. Modern point bar deposits analogous to the Athabasca Oil Sands, Alberta, Canada. In Tide-Influenced Sedimentary Environments and Facies (ed. DeBoer, P. L., Gelder, A., & Nio, S. D.), pp. 417–432. Boston, MA: D. Reidel Publishing Company.CrossRefGoogle Scholar
Smith, D. G., Reinson, G. E., Zaitlin, B. A., & Rahmani, R. A. 1991. Clastic Tidal Sedimentology. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Smith, D. G., & Smith, N. D. 1980. Sedimentation in anastomosed river systems: examples from alluvial valleys near Banff, Alberta. Journal of Sedimentary Petrology 50, 157–164.CrossRefGoogle Scholar
Smith, L. C., Sheng, Y., Magilligan, F. J.et al. 2006. Geomorphic impact and rapid subsequent recovery from the 1996 Skeiðarásandur jokulhlaup, Iceland, measured with multi-year airborne lidar. Geomorphology 75, 65–75.CrossRefGoogle Scholar
Smith, N. D. 1974. Sedimentology and bar formation in the upper Kicking Horse River, a braided outwash stream. Journal of Geology 81, 205–223.CrossRefGoogle Scholar
Smith, N. D., Cross, T. A., Dufficy, J. P., & Clough, S. R. 1989. Anatomy of an avulsion. Sedimentology 36, 1–23.CrossRefGoogle Scholar
Smith, N. D., & Rogers, J. 1999. Fluvial Sedimentology VI. Oxford: Blackwells.CrossRefGoogle Scholar
Smith, N. D., Slingerland, R. L., Pérez-Arlucea, M., & Morozova, G. S. 1998. The 1870s avulsion of the Saskatchewan River. Canadian Journal of Earth Sciences 35, 453–466.CrossRefGoogle Scholar
Smith, S. V., & Kinsey, D. W. 1976. Calcium carbonate production, coral reef growth, and sea level change. Science 194, 937–939.CrossRefGoogle ScholarPubMed
Smoot, J. P. 1983. Depositional subenvironments in an arid closed basin: the Wilkens Peak Member of the Green River Formation (Eocene), Wyoming. Sedimentology 30, 801–827.CrossRefGoogle Scholar
Smoot, J. P., & Lowenstein, T. K. 1991. Depositional environments of non-marine evaporites. In Evaporites, Petroleum and Mineral Resources (ed. Melvin, J. L.), pp. 189–347. Amsterdam: Elsevier.Google Scholar
Snedden, J. W., & Nummedal, D. 1991. Origin and geometry of storm-deposited sand beds in modern sediments of the Texas continental shelf. In Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (ed. Swift, D. J. P., Oertel, G. F., Tillman, R. W., & Thorne, J. A.), pp. 283–308. Oxford: Blackwells.Google Scholar
Sonnenfeld, P. 1984. Brines and Evaporites. Orlando, FL: Academic Press.Google Scholar
Sorby, H. C. 1879. The structure and origin of limestones. Proceedings of the Geological Society of London 35, 56–95.Google Scholar
Soulsby, R. L., & Bettess, R. 1991. Sand Transport in Rivers, Estuaries and the Sea. Rotterdam: Balkema.Google Scholar
Southard, J. B., & Boguchwal, L. A. 1990. Bed configurations in steady unidirectional flows. Part 2. Synthesis of flume data. Journal of Sedimentary Petrology 60, 658–679.CrossRefGoogle Scholar
Southard, J. B., Lambie, J. M., Federico, D. C., Pile, H. T., & Weidman, C. R. 1990. Experiments on bed configurations in fine sands under bidirectional purely oscillatory flow, and the origin of hummocky cross-stratification. Journal of Sedimentary Petrology 60, 1–17.Google Scholar
Sparks, R. S. J. 1976. Grain-size variations in ignimbrites and implications for the transport of pyroclastic flows. Sedimentology 23, 147–188.CrossRefGoogle Scholar
Sparks, R. S. J., Bonnecaze, R. T., Huppert, H. E.et al. 1993. Sediment-laden gravity currents with reversing buoyancy. Earth and Planetary Science Letters 114, 249–288.CrossRefGoogle Scholar
Sparks, R. S. J., & Gilbert, J. S. 2002. Physics of Explosive Volcanic Eruptions. London: Geological Society of London.Google Scholar
Spencer, R. J., & Lowenstein, T. K. 1990. Evaporites. In Diagenesis (ed. McIlreath, I. A. & Morrow, D. W.), pp. 141–163. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Spencer, R. J., Möller, N., & Weare, J. H. 1990. The prediction of mineral solubilities in natural waters: a chemical equilibrium model for the Na–K–Ca–Mg–Cl–SO4–H2O system at temperatures below 25 ℃. Geochimica et Cosmochimica Acta 54, 575–590.CrossRefGoogle Scholar
Sposito, G. 1989. The Chemistry of Soils. New York: Oxford University Press.Google Scholar
Srivastava, R. M. 1994. An overview of stochastic methods for reservoir characterization. In Stochastic Modeling and Geostatistics (ed. Yarus, J. M. & Chambers, R. L.), pp. 3–16. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Stallard, R. F., & Edmond, J. M. 1981. Geochemistry of the Amazon 1. Precipitation chemistry and the marine contribution to the dissolved load at the time of peak discharge. Journal of Geophysical Research 86, 9844–9858.CrossRefGoogle Scholar
Stallard, R. F., & Edmond, J. M. 1983. Geochemistry of the Amazon 2. The influence of geology and weathering environment on the dissolved load. Journal of Geophysical Research 88, 9671–9688.CrossRefGoogle Scholar
Stallard, R. F., & Edmond, J. M. 1987. Geochemistry of the Amazon 3. Weathering chemistry and the limits to dissolved inputs. Journal of Geophysical Research 92, 8293–8302.CrossRefGoogle Scholar
Stanley, S. M. 1966. Paleoecology and diagenesis of Key Largo Limestone, Florida. American Association of Petroleum Geologists Bulletin 50, 1927–1947.Google Scholar
Stanley, S. M. 2005. Earth System History, 2nd edn. New York: W. H. Freeman and Company.Google Scholar
Stanley, S. M., & Hardie, L. A. 1998. Secular oscillations in the carbonate mineralogy of reef-building and sediment-producing organisms driven by tectonically forced shifts in seawater chemistry. Palaeogeography, Palaeoclimatology, Palaeoecology 144, 3–19.CrossRefGoogle Scholar
Stanley, S. M., Ries, J. B., & Hardie, L. A. 2002. Low-magnesium calcite produced by coralline algae in seawater of Late Cretaceous composition. Proceedings of the National Academy of Sciences of the USA 99, 15323–15326.CrossRefGoogle ScholarPubMed
Starkel, L., Gregory, K. J., & Thornes, J. B. 1991. Temperate Palaeohydrology. Chichester: Wiley.Google Scholar
Steel, R. J., Maehle, S., Nilsen, H., Roe, S. L., & Spinnanger, A. 1977. Coarsening-upward cycles in the alluvium of the Hornelen basin (Devonian, Norway): sedimentary response to tectonic events. Geological Society of America Bulletin 88, 1124–1134.2.0.CO;2>CrossRefGoogle Scholar
Steel, R. J., & Ryseth, A. 1990. The Triassic–Early Jurassic succession in the northern North Sea: megasequence stratigraphy and intra-Triassica tectonics. In Tectonic Events Responsible for Britain's Oil and Gas Reserves (ed. Hardman, R. F. P. & Brooks, J.), pp. 139–168. London: Geological Society of London.Google Scholar
Stern, C. W., Scoffin, T. P., & Martindale, W. 1977. Calcium carbonate budget of a fringing reef on the west coast of Barbados, pt. 1, zonation and productivity. Bulletin of Marine Science 27, 779–810.Google Scholar
Stewart, F. H. 1963. Marine evaporites. United States Geological Survey Professional Paper 440-Y.
Stockman, K. W., Ginsburg, R. N., & Shinn, E. A. 1967. The production of lime mud by algae in South Florida. Journal of Sedimentary Petrology 37, 633–648.Google Scholar
Stokes, C. R., & Clark, C. D. 2001. Palaeo-ice streams. Quaternary Science Reviews 20, 1437–1457.CrossRefGoogle Scholar
Stokes, C. R., & Clark, C. D. 2002. Ice stream shear margin moraines. Earth Surface Processes and Landforms 27, 547–558.CrossRefGoogle Scholar
Stokes, C. R., & Clark, C. D. 2003. Laurentide ice streaming on the Canadian Shield: a conflict with the soft-bedded ice stream paradigm. Geology 31, 347–350.2.0.CO;2>CrossRefGoogle Scholar
Stokes, W. L. 1968. Multiple parallel-truncation bedding planes – a feature of wind deposited sandstone formations. Journal of Sedimentary Petrology 38, 510–515.Google Scholar
Stokstad, E. 2004. Defrosting the carbon freezer of the north. Science 304, 1618–1620.CrossRefGoogle ScholarPubMed
Stone, G. W., & Orford, J. D. 2004. Storms and their significance in coastal morpho-sedimentary dynamics. Marine Geology 210, 1–368.CrossRefGoogle Scholar
Stonecipher, S. A. 2000. Applied Sandstone Diagenesis – Practical Petrographic Solutions for a Variety of Common Exploration, Development, and Production Problems. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Storms, J. E. A. 2003. Event-based stratigraphic simulation of wave-dominated shallow-marine environments. Marine Geology 199, 83–100.CrossRefGoogle Scholar
Storms, J. E. A., & Swift, D. J. P. 2003. Shallow-marine sequences as the building blocks of stratigraphy: insights from numerical modeling. Basin Research 15, 287–303.CrossRefGoogle Scholar
Stouthamer, E., & Berendsen, H. J. A. 2000. Factors controlling the Holocene avulsion history of the Rhine–Meuse delta (The Netherlands). Journal of Sedimentary Research 70, 1051–1064.CrossRefGoogle Scholar
Stouthamer, E., & Berendsen, H. J. A. 2001. Avulsion frequency, avulsion duration and interavulsion period of Holocene channel belts in the Rhine–Meuse delta, The Netherlands. Journal of Sedimentary Research 71, 589–598.CrossRefGoogle Scholar
Stouthamer, E. 2005. Reoccupation of channel belts and its influence on alluvial architecture in the Holocene Rhine-Meuse delta, the Netherlands. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J.), pp. 319–339. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Stow, D. A. V. 1986. Deep clastic seas. In Sedimentary Environments and Facies, 2nd edn. (ed. Reading, H. G.), pp. 399–444. Oxford: Blackwells.Google Scholar
Stow, D. A. V.1994. Deep sea processes of sediment transport and deposition. In Sediment Transport and Depositional Processes (ed. Pye, K.), pp. 257–291. Oxford: Blackwells.Google Scholar
Stow, D. A. V., & Faugeres, J. C. (eds.) 1993. Contourites and Bottom Currents. Sedimentary Geology (special issue) 82.
Stow, D. A. V., & Faugeres, J. C.1998. Contourites, Turbidites and Process Interaction. Sedimentary Geology (special issue) 115.
Stow, D. A. V., & Lovell, J. P. B. 1979. Contourites: their recognition in modern and ancient sediments. Earth-Science Reviews 14, 251–291.CrossRefGoogle Scholar
Stow, D. A. V., & Piper, D. J. W. 1984. Fine-Grained Sediments: Deep-Water Processes and Facies. London: Geological Society of London.Google Scholar
Stow, D. A. V., Pudsey, C. J., Howe, J. A., Faugeres, J. -C., & Vianna, A. R. 2002. Deep-water Contourite Systems: Modern Drifts and Ancient Series, Seismic and Sedimentary Characteristics. London: Geological Society of London.Google Scholar
Stow, D. A. V., Reading, H. G., & Collinson, J. D. 1996. Deep seas. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 395–453. Oxford: Blackwells.Google Scholar
Stow, D. A. V., & Shanmugan, G. 1980. Sequences of structures in fine-grained turbidites: comparison of recent deep sea and ancient flysch sediments. Sedimentary Geology 25, 23–42.CrossRefGoogle Scholar
Stow, D. A. V., & Tabrez, A. 1998. Hemipelagites: Facies, Processes, and Models, pp. 317–338. London: Geological Society of London.Google Scholar
Stow, D. A. V., & Wetzel, A. 1990. Hemiturbidite: a new type of deep-water sediment. Proceedings of the Ocean Drilling Program, Scientific Results 116, 25–34.Google Scholar
Strahler, A. N. 1957. Quantitative analysis of watershed geomorphology. Transactions of the American Geophysical Union 38, 913–920.CrossRefGoogle Scholar
Straub, S. 2001. Bagnold revisited: implications for the rapid motion of high-concentration sediment flows. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 91–109. Oxford: Blackwells.CrossRefGoogle Scholar
Stride, A. H. 1982. Offshore Tidal Sands. London: Chapman and Hall.CrossRefGoogle Scholar
Stumm, W., & Morgan, J. J. 1981. Aquatic Chemistry: An Introduction Emphazing Chemical Equilibria in Natural Waters. New York: John Wiley and Sons.Google Scholar
Sugden, D. E., & John, B. S. 1976. Glaciers and Landscape. London: Edward Arnold.Google Scholar
Summerhayes, C. P., & Thorpe, S. A. 1996. Oceanography: An Illustrated Guide. London: Manson.Google Scholar
Summerfield, M. A. 1983. Petrography and diagenesis of silcrete from the Kalahari Basin and Cape coastal zone. Journal of Sedimentary Petrology 53, 895–909.Google Scholar
Sumner, D. Y., & Grotzinger, J. P. 2000. Late Archean aragonite precipitation: petrography, facies associations, and environmental significance. In Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World (ed. Grotzinger, J. P. & James, N. P.), pp. 123–144. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Sun, T., Meakin, P., & Jossang, T. 2001a. Meander migration and the lateral tilting of floodplains. Water Resources Research 37, 1485–1502.CrossRefGoogle Scholar
Sun, T., Meakin, P., & Jossang, T. 2001b. A computer model for meandering rivers with multiple bedload sediment sizes 1. Theory. Water Resources Research 37, 2227–2241.CrossRefGoogle Scholar
Sun, T., Meakin, P., & Jossang, T. 2001c. A computer model for meandering rivers with multiple bedload sediment sizes 2. Computer simulations. Water Resources Research 37, 2243–2258.CrossRefGoogle Scholar
Sun, T., Paola, C., Parker, G., & Meakin, P. 2002. Fluvial fan deltas: linking channel processes with large-scale morphodynamics. Water Resources Research 38, 1151.CrossRefGoogle Scholar
Swift, D. J. P. 1975. Barrier island genesis: evidence from the Middle Atlantic Shelf of North America. Sedimentary Geology 14, 1–43.CrossRefGoogle Scholar
Swift, D. J. P., & Field, M. E. 1981. Evolution of a classic sand ridge field; Maryland sector, North American inner shelf. Sedimentology 28, 461–482.CrossRefGoogle Scholar
Swift, D. J. P., Figueiredo, A. R., Freeland, G. L., & Oertal, G. F. 1983. Hummocky cross stratification and megaripples: a geological double standard?Journal of Sedimentary Petrology 53, 1295–1317.Google Scholar
Swift, D. J. P., Han, G., & Vincent, C. E. 1986. Fluid processes and sea-floor response on a modern storm-dominated shelf: middle Atlantic shelf of North America. Part I: the storm-current regime. In Shelf Sands and Sandstones (ed. Knight, R. J. & McLean, J. R.), pp. 99–119. Calgary, Alberta: Canadian Society of Petroleum Geologists.Google Scholar
Swift, D. J. P., Oertel, G. F., Tillman, R. W., & Thorne, J. A. 1991. Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy. Oxford: Blackwells.Google Scholar
Swift, D. J. P., & Thorne, J. A. 1991. Sedimentation on continental margins, I: a general model for shelf sedimentation. In Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (ed. Swift, D. J. P., Oertel, G. F., Tillman, R. W., & Thorne, J. A.), pp. 3–31. Oxford: Blackwells.Google Scholar
Syvitski, J. P. M., Burrell, D. C., & Skei, J. M. 1987. Fjords: Processes and Products. New York: Springer-Verlag.CrossRefGoogle Scholar
Takahashi, T. 1981. Debris flow. Annual Review of Fluid Mechanics 13, 57–77.CrossRefGoogle Scholar
Takahashi, T.2001. Mechanics and simulation of snow avalanches, pyroclastic flows and debris flows. In Particulate Gravity Currents (ed. McCaffrey, W. D., Kneller, B. C., & Peakall, J.), pp. 11–43. Oxford: Blackwells.CrossRefGoogle Scholar
Talbot, M. R. 1985. Major bounding surfaces in aeolian sandstones – a climatic model. Sedimentology 32, 257–265.CrossRefGoogle Scholar
Talbot, M. R., & Allen, P. A. 1996. Lakes. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 83–124. Oxford: Blackwell Science.Google Scholar
Taylor, A., Goldring, R., & Gowland, S. 2003. Analysis and application of ichnofabrics. Earth-Science Reviews 60, 227–259.CrossRefGoogle Scholar
Taylor, K. G., Gawthorpe, R. L., Curtis, C. D., Marshall, J. D., & Anwiller, D. A. 2000. Carbonate cementation in a sequence-stratigraphic framework: Upper Cretaceous sandstones, Book Cliffs, Utah–Colorado. Journal of Sedimentary Research 70, 360–372.CrossRefGoogle Scholar
Tennant, C. B., & Berger, R. W. 1957. X-ray determination of the dolomite–calcite ratios of a carbonate rock. American Mineralogist 42, 23–29.Google Scholar
Terwindt, J. H. J. 1981. Origin and sequences of sedimentary structures in inshore mesotidal deposits of the North Sea. In Holocene Marine Sedimentation in the North Sea Basin (ed. Nio, S. D., Schuttenhelm, R. T. E., & Weering, T. C. E.), pp. 51–64. Oxford: Blackwells.CrossRefGoogle Scholar
Terwindt, J. H. J.1988. Palaeo-tidal reconstructions of inshore tidal depositional environments. In Tide-influenced Sedimentary Environments and Facies (ed. DeBoer, P. L., Gelder, A., & Nio, S. D.), pp. 233–263. Boston, MA: D. Reidel Publishing Company.CrossRefGoogle Scholar
Tetzlaff, D. M., & Harbaugh, J. W. 1989. Simulating Clastic Sedimentation. New York: Van Nostrand Reinhold.CrossRefGoogle Scholar
Tetzlaff, D. M., & Priddy, G. 2001. Sedimentary process modeling: from academia to industry. In Geologic Modeling and Simulation: Sedimentary Systems (ed. Merriam, D. F. & Davis, J. C.), pp. 45–69. New York: Kluwer Academic/Plenum Publishers.CrossRefGoogle Scholar
Thomas, D. 1997. Arid Zone Geomorphology. London: Bellhaven/Hallsted Press.Google Scholar
Thomas, R. G., Smith, D. G., Wood, J. M.et al. 1987. Inclined heterolithic stratification – terminology, description, interpretation and significance. Sedimentary Geology 53, 123–179.CrossRefGoogle Scholar
Thrailkill, J. 1976. Speleothems. In Stromatolites (ed. Walter, M. R.), pp. 73–86. Amsterdam: Elsevier.Google Scholar
Thurman, H. V., & Trujilo, A. P. 2004. Introductory oceanography, 10th edn. Englewood Cliffs, NJ, Prentice-Hall.Google Scholar
Tillman, R. W., & Martinsen, R. S. 1984. The Shannon shelf-ridge sandstone complex, Salt Creek anticline area, Powder River Basin, Wyoming. In Siliciclastic Shelf Sediments (ed. Tillman, R. W. & Seimers, C. T.), pp. 85–142. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Tillman, R. W., & Martinsen, R. S. 1987. Sedimentological model and production characteristics of Hartzog Draw Field, Wyoming. In Reservoir Sedimentology (ed. Tillman, R. W. & Weber, K. J.), pp. 15–112. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Tinkler, K. J., & Wohl, E. E. 1998. Rivers Over Rock: Fluvial Processes in Bedrock Channels. Washington, D.C.: American Geophysical Union.CrossRefGoogle Scholar
Thorne, C. R., Russell, A. P. G., & Alam, M. K. 1993. Planform pattern and channel evolution of Brahmaputra River, Bangladesh. In Braided Rivers (ed. Best, J. L. & Bristow, C. S.), pp. 257–276. London: Geological Society of London.Google Scholar
Törnqvist, T. E. 1993. Holocene alternation of meandering and anastomosing fluvial systems in the Rhine–Meuse delta (central Netherlands) controlled by sea-level rise and subsoil erodibility. Journal of Sedimentary Petrology 63, 683–693.Google Scholar
Törnqvist, T. E. 1994. Middle and late Holocene avulsion history of the River Rhine (Rhine–Meuse delta, Netherlands). Geology 22, 711–714.2.3.CO;2>CrossRefGoogle Scholar
Törnqvist, T. E., & Bridge, J. S. 2002. Spatial variation of overbank aggradation rate and its influence on avulsion frequency. Sedimentology 49, 891–905.CrossRefGoogle Scholar
Törnqvist, T. E., Kidder, T. R., Autin, W. J.et al. 1996. A revised chronology for Mississippi River subdeltas. Science 273, 1693–1696.CrossRefGoogle Scholar
Törnqvist, T. E., Wallinga, J., Murray, A. S.et al. 2000. Response of the Rhine–Meuse system (west-central Netherlands) to the last Quaternary glacio-eustatic cycles. Global and Planetary Change 27, 89–111.CrossRefGoogle Scholar
Traverse, A., & Ginsburg, R. N. 1966. Palynology of the surface sediments of Great Bahama Bank, as related to water movements and sedimentation. Marine Geology 4, 417–459.CrossRefGoogle Scholar
Trendall, A. F., & Morris, R. C. 1983. Iron-Formation: Facts and Problems. Amsterdam: Elsevier.Google Scholar
Trendall, A. F., & Blockley, J. G. 1970. The iron formations of the Precambrian Hamersley Group, Western Australia. Geological Survey of Western Australia Bulletin 119.
Trewin, N. H. 1994. Depositional environment and preservation of biota in the Lower Devonian hot-springs of Rhynie, Aberdeenshire, Scotland. Transactions of the Royal Society of Edinburgh: Earth Sciences 84, 433–442.CrossRefGoogle Scholar
Trewin, N. H., & Rice, C. M. 1992. Stratigraphy and sedimentology of the Devonian Rhynie Chert locality. Scottish Journal of Geology 28, 37–47.CrossRefGoogle Scholar
Tripati, A., & Elderfield, H. 2005. Deep-sea temperature and circulation changes at the Paleocene–Eocene Thermal Maximum. Science 308, 1894–1898.CrossRefGoogle ScholarPubMed
Tsoar, H. 1982. Internal structure and surface geometry of longitudinal (seif) dunes. Journal of Sedimentary Petrology 52, 823–831.Google Scholar
Tsoar, H. 1983. Dynamic processes acting on a longitudinal (seif) dune. Sedimentology 30, 567–578.CrossRefGoogle Scholar
Tsoar, H. 1984. The formation of seif dunes from barchans – a discussion. Zeitschrift für Geomorphologie 28, 99–103.Google Scholar
Tsoar, H. 1989. Linear dunes – forms and formation. Progress in Physical Geography 13, 507–528.CrossRefGoogle Scholar
Tsoar, H., & Pye, K. 1987. Dust transport and the question of desert loess formation. Sedimentology 34, 139–153.CrossRefGoogle Scholar
Tucker, G. E., & Slingerland, R. L. 1996. Predicting sediment flux from fold and thrust belts. Basin Research 8, 329–349.CrossRefGoogle Scholar
Tucker, G. E., & Slingerland, R. L. 1997. Drainage basin responses to climate change. Water Resources Research 33, 2031–2047.CrossRefGoogle Scholar
Tucker, G. E., Lancaster, S. T., Gasparini, N. M., & Bras, R. L. 2002. The Channel-Hillslope Integrated Landscape Development Model (CHILD). In Landscape Erosion and Evolution Modeling (ed. Harmon, R. S. & Doe, W. W. III), pp. 349–388. New York: Kluwer Academic Publishing.Google Scholar
Tucker, M. E. 2001. Sedimentary Petrology: An Introduction to the Origin of Sedimentary Rocks. Oxford: Blackwell Scientific.Google Scholar
Tucker, M. E., & Wright, V. P. 1990. Carbonate Sedimentology. Oxford: Blackwells.CrossRefGoogle Scholar
Turcotte, D. L., & Schubert, G. 2002. Geodynamics, 2nd edn. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Turner, E. C., Narbonne, G. M., & James, N. P. 1993. Neoproterozoic reef microstructures from the Little Dal Group, northwestern Canada. Geology 21, 259–262.2.3.CO;2>CrossRefGoogle Scholar
Turner, E. C., Narbonne, G. M., & James, N. P.2000a. Framework composition of Early Neoproterozoic calcimicrobial reefs and associated microbialites, Mackenzie Mountains, N.W.T., Canada. In Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World (ed. Grotzinger, J. P. & James, N. P.), pp. 179–205. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Turner, E. C., Narbonne, G. M., & James, N. P. 2000b. Taphonomic control on microstructures in Early Neoproterozoic reefal stromatolites and thrombolites. Palaios 15, 87–111.2.0.CO;2>CrossRefGoogle Scholar
Tuttle, M. P., Ruffman, A., Anderson, T., & Jeter, H. 2004. Distinguishing tsunami deposits from storm deposits along the coast of northeastern North America: lessons learned from the 1929 Grand Banks tsunami and the 1991 Halloween storm. Seismological Research Letters 75, 117–131.CrossRefGoogle Scholar
Twichell, D. C., Kenyon, N. H., Parson, L. M., & McGregor, B. A. 1991. Depositional patterns of the Mississippi Fan surface: evidence from GLORIA II and high resolution seismic profiles. In Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems (ed. Weimer, P. & Link, M. H.), pp. 349–364. New York: Springer-Verlag.CrossRefGoogle Scholar
Tye, R. S. 1984. Geomorphic evolution and stratigraphy of Price and Capers Inlets, South Carolina. Sedimentology 31, 655–674.CrossRefGoogle Scholar
Tye, R. S. 2004. Geomorphology: an approach to determining subsurface reservoir dimensions. American Association of Petroleum Geologists Bulletin 88, 1123–1147.CrossRefGoogle Scholar
Tye, R. S., & Coleman, J. M. 1989a. Depositional processes and stratigraphy of fluvially dominated lacustrine deltas: Mississippi Delta Plain. Journal of Sedimentary Petrology 59, 973–996.Google Scholar
Tye, R. S., & Coleman, J. M. 1989b. Evolution of Atchafalaya lacustrine deltas, south-central Louisiana. Sedimentary Geology 65, 95–112.CrossRefGoogle Scholar
Tye, R. S., & Moslow, T. F. 1993. Tidal inlet reservoirs: insights from modern examples. In Marine Clastic Reservoirs: Examples and Analogues (ed. Rhodes, E. G. & Moslow, T. F.), pp. 77–99. New York: Springer-Verlag.CrossRefGoogle Scholar
Tyler, K., Henriquez, A., & Svanes, T. 1994. Modeling heterogeneities in fluvial domains: a review of the influence on production profiles. In Stochastic Modeling and Geostatistics (ed. Yarus, J. M. & Chambers, R. L.), pp. 77–89. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
US Department of Agriculture Soil Survey Staff 1993. Soil Survey Manual. Washington, D.C.: United States Department of Agriculture.
US Department of Agriculture Soil Survey Staff2003. Keys to Soil Taxonomy. Washington, D.C.: United States Department of Agriculture.
Usiglio, J. 1894. Analyse de l'eau de la Mediterranée sur la côte de France. Annales de Chimie et de Physique 27, 172–191.Google Scholar
Vail, P. R., Mitchum, R. M. Jr, Todd, R. G. et al. 1977. Seismic stratigraphy and global changes in sea-level. In Seismic Stratigraphy – Applications to Hydrocarbon Exploration (ed. Payton, C. E.), pp. 49–212. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Vallance, J. W., & Scott, K. M. 1997. The Osceola Mudflow from Mount Rainier: sedimentology and hazard implications of a huge clay-rich debris flow. Geological Society of America Bulletin 109, 143–163.2.3.CO;2>CrossRefGoogle Scholar
Andel, , , T. H., & Komar, P. D. 1969. Ponded sediments of the Mid-Atlantic Ridge between 22° and 23° north latitude. Geological Society of America Bulletin 80, 1163–1190.CrossRefGoogle Scholar
Berg, J. H. 1982. Migration of large-scale bedforms and preservation of crossbedded sets in highly accretional parts of tidal channels in the Oosterschelde, S.W. Netherlands. Geologie en Mijnbouw 61, 253–263.Google Scholar
Berg, J. H. 1995. Prediction of alluvial channel pattern of perennial rivers. Geomorphology 12, 259–279.CrossRefGoogle Scholar
Van Den Berg, J. H., & Van Gelder, A. 1993. A new bedform stability diagram, with emphasis on the transition of ripples to plane bed in flows over fine sand and silt. In Alluvial Sedimentation (ed. Marzo, M. & Puidefabregas, C.), pp. 11–21. Oxford: Blackwells.CrossRefGoogle Scholar
Berg, J. H., Gelder, A., & Mastbergen, D. R. 2002. The importance of breaching as a mechanism of subaqueous slope failure in fine sand. Sedimentology 49, 81–95.CrossRefGoogle Scholar
Van der Wateren, F. M. 2002. Processes of glaciotectonism. In Modern and Past Glacial Environments (ed. Menzies, J.), pp. 417–443. Oxford: Butterworth-Heinemann.CrossRefGoogle Scholar
Dyke, M. 1982. An Album of Fluid Motion. Stanford, CA: Parabolic Press.Google Scholar
Heerden, I. L., & Roberts, H. H. 1988. Facies development of Atchafalaya Delta, Louisiana: a modern bayhead delta. American Association of Petroleum Geologists Bulletin 72, 439–453.Google Scholar
Heijst, M. W. I. M., & Postma, G. 2001. Fluvial response to sea-level changes: a quantitative analogue, experimental approach. Basin Research 13, 269–292.CrossRefGoogle Scholar
Houten, F. B. 1964. Cyclic lacustrine sedimentation, Upper Triassic Lockatong Formation, central New Jersey and adjacent Pennsylvania. Kansas Geological Survey Bulletin 169, 497–532.Google Scholar
Houten, F. B. 1985. Oolitic ironstones and contrasting Ordovician and Jurassic paleogeography. Geology 13, 722–724.2.0.CO;2>CrossRefGoogle Scholar
Houten, F. B. & Bhattacharyya, P. D. 1982. Phanerozoic oolitic ironstones – geologic record and facies model. Annual Review of Earth and Planetary Sciences 10, 441–457.CrossRefGoogle Scholar
Lith, Y., Warthmann, R., Vasconcelos, C., & McKenzie, J. A. 2003. Microbial fossilization in carbonate sediments: a result of the bacterial surface involvement in dolomite precipitation. Sedimentology 50, 237–245.CrossRefGoogle Scholar
Lith, Y., Warthmann, R., Vasconcelos, C., & McKenzie, J. A. 2004. Sulfphate-reducing bacteria induce low-temperature Ca-dolomite and high Mg-calcite formation. Geobiology 1, 71–79.CrossRefGoogle Scholar
Niekerk, A.Vogel, A. K., Slingerland, R., & Bridge, J. S. 1992. Routing heterogeneous size–density sediments over a moveable bed: model development. Journal of Hydraulic Engineering, ASCE 118, 246–262.CrossRefGoogle Scholar
Rijn, L. C. 1990. Principles of Fluid Flow and Surface Waves in Rivers, Estuaries, Seas, and Oceans. Amsterdam: Aqua Publications.Google Scholar
Rijn, L. C. 1993. Principles of Sediment Transport in Rivers, Estuaries and Coastal Seas. Amsterdam: Aqua Publications.Google Scholar
Van Tassell, J. 1994. Cyclic deposition of the Devonian Catskill Delta of the Appalachian, U.S.A. In Orbital Forcing and Cyclic Sequences (ed. Boer, P. L. & Smith, D. G.), pp. 395–411. Oxford: Blackwells.CrossRefGoogle Scholar
Wagoner, J. C., & Bertram, G. T. 1995. Sequence Stratigraphy of Foreland Basin Deposits. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wagoner, J. C., Mitchum, R. M. Jr., Campion, K. M., & Rahmanian, V. D. 1990. Siliciclastic Sequence Stratigraphy in Well Logs, Cores and Outcrop: Concepts for High Resolution Correlation of Time and Facies. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Veizer, J., Godderis, Y., & Francois, L. M. 2000. Evidence for decoupling of atmospheric CO2 and global climate during the Phanerozoic eon. Nature 408, 698–701.CrossRefGoogle ScholarPubMed
Verway, J. 1952. On the ecology of distribution of cockle and mussel in the Dutch Waddensee, their role in sedimentation and the source of their food supply. Archives Néerlandaises de Zoologie 10, 171–240.CrossRefGoogle Scholar
Vigilar, G. G. Jr., & Diplas, P. 1997. Stable channels with mobile bed: formulation and numerical solution. Journal of Hydraulic Engineering, ASCE 123, 189–199.CrossRefGoogle Scholar
Vigilar, G. G. Jr., & Diplas, P. 1998. Stable channels with mobile bed: model verification and graphical solution. Journal of Hydraulic Engineering, ASCE 124, 1097–1108.CrossRefGoogle Scholar
Vincent, C. E., Young, R. A., & Swift, D. J. P. 1982. On the relationship between bedload and suspended sand transport on the inner shelf, Long Island, New York. Journal of Geophysical Research 87, 4163–4170.CrossRefGoogle Scholar
Vinopal, R. J., & Coogan, A. H. 1978. Effect of particle shape on the packing of carbonate sands and gravels. Journal of Sedimentary Petrology 48, 7–24.Google Scholar
Visscher, P. T., & Stolz, J. F. 2005. Microbial mats as bioreactors: populations, processes, and products. Palaeogeography, Palaeoclimatology, Palaeoecology 219, 87–100.CrossRefGoogle Scholar
Visser, M. J. 1980. Neap–spring cycles reflected in Holocene subtidal large-scale bedform deposits: a preliminary note. Geology 8, 543–546.2.0.CO;2>CrossRefGoogle Scholar
Vogel, K., Niekerk, A., Slingerland, R., & Bridge, J. S. 1992. Routing of heterogeneous size–density sediments over a moveable bed: model verification and testing. Journal of Hydraulic Engineering, ASCE 118, 263–279.CrossRefGoogle Scholar
Vose, R. S., Schmoyer, R. L., Steurer, P. M., Peterson, T. C., Heim, R., Karl, T. R., & Eischeid, J. K. 1992. The Global Historical Climatology Network: Long-term Monthly Temperature, Precipitation, Sea Level Pressure, and Station Pressure Data. Oak Ridge, TN: Oak Ridge National Laboratory Environmental Sciences Division.Google Scholar
Walgreen, M., Calvete, D., & Swart, H. E. 2002. Growth of large-scale bedforms due to storm-driven and tidal currents: a model approach. Continental Shelf Research 22, 2777–2793.CrossRefGoogle Scholar
Walker, R. G. 1965. The origin and significance of the internal structures of turbidites. Proceedings of the Yorkshire Geological Society 35, 1–32.CrossRefGoogle Scholar
Walker, R. G. 1978. Deep water sandstone facies and ancient submarine fans: models for exploration for stratigraphic traps. American Association of Petroleum Geologists Bulletin 62, 932–966.Google Scholar
Walker, R. G.1979. Turbidites and associated coarse clastic deposits. In Facies Models (ed. Walker, R. G.), pp. 91–103. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G.1984. Shelf and shallow marine sands. In Facies Models, 2nd edn. (ed. Walker, R. G.), pp. 141–170. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G.1992. Turbidites and submarine fans. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 239–263. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G., & James, N. P. 1992. Facies Models: Response to Sea Level Change. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walker, R. G., & Mutti, E. 1973. Turbidite facies and facies associations. In Turbidites and Deep-Water Sedimentation (ed. Middleton, G. V. & Bouma, A. H.), pp. 19–157. Tulsa, OK: SEPM (Society for Sedimentary Geology).Google Scholar
Walker, R. G., & Plint, A. G. 1992. Wave- and storm-dominated shallow marine systems. In Facies Models: Response to Sea Level Change (ed. Walker, R. G. & James, N. P.), pp. 219–238. St. John's, Newfoundland: Geological Association of Canada.Google Scholar
Walter, M. R. 1976. Stromatolites. Amsterdam: Elsevier.Google Scholar
Walther, J. V. 2005. Essentials of Geochemistry. Sudbury, MA: A. Jones and Bartlett Publishers.Google Scholar
Wanless, H. R. 1979. Limestone response to stress: pressure solution and dolomitization. Journal of Sedimentary Petrology 49, 437–462.Google Scholar
Ward, A. W., & Greeley, R. 1984. The yardangs at Rogers Lake, California. Geological Society of America Bulletin 95, 829–837.2.0.CO;2>CrossRefGoogle Scholar
Ward, W. B. 1996. Evolution and diagenesis of Frasnian carbonate platforms, Devonian Reef Complexes, Napier Range, Canning Basin, Western Australia. Unpublished Ph.D. dissertation, State University of New York at Stony Brook, New York.
Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setata, H., Putten, W. H., & Wall, D. H. 2004. Ecological linkages between aboveground and belowground biota. Science 304, 1629–1633.CrossRefGoogle ScholarPubMed
Warren, J. K. 1982. The hydrological significance of Holocene tepees, stromatolites, and boxwork limestones in coastal Salinas in south Australia. Journal of Sedimentary Petrology 52, 1171–1201.Google Scholar
Warren, J. K. 2006. Evaporites: Sediments, Resources and Hydrocarbons. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Warren, W. P., & Ashley, G. M. 1994. Origins of the ice-contact ridges (eskers) of Ireland. Journal of Sedimentary Research 64, 433–449.Google Scholar
Watabe, N., & Wilbur, K. M. 1976. The Mechanisms of Mineralization in the Invertebrates and Plants. Columbia, SC: University of South Carolina Press.Google Scholar
Waters, B. B., Spencer, R. J., & Demicco, R. V. 1989. Three-dimensional architecture of shallowing-upward carbonate cycles: Middle and Upper Cambrian Waterfowl Formation, southern Canadian Rocky Mountains. Bulletin of Canadian Petroleum Geology 37, 198–209.Google Scholar
Watts, N. L. 1977. Pseudo-anticlines and other structures in some calcretes of Botswana and South Africa. Earth Surface Processes and Landforms 2, 63–74.CrossRefGoogle Scholar
Watts, N. L. 1980. Quaternary pedogenic calcretes from the Kalahari (southern Africa); mineralogy, genesis and diagenesis. Sedimentology 27, 661–686.CrossRefGoogle Scholar
Weaver, P. P. E., & Thomson, J. 1987. Geology and Geochemistry of Abyssal Plains. London: Geological Society of London.Google Scholar
Weertman, J. 1979. The unsolved general glacier sliding problem. Journal of Glaciology 23, 97–115.CrossRefGoogle Scholar
Weertman, J. 1986. Basal water and high-pressure basal ice. Journal of Glaciology 32, 455–463.CrossRefGoogle Scholar
Weimer, P. 1990. Sequence stratigraphy, facies geometries, and depositional history of the Mississippi Fan, Gulf of Mexico. American Association of Petroleum Geologists Bulletin 74, 425–453.Google Scholar
Weimer, P.1995. Sequence stratigraphy of the Mississippi Fan (late Miocene–Pleistocene), northern deep Gulf of Mexico. In Atlas of Deep Water Environments; Architectural Style in Turbidite Systems (ed. Pickering, K. T., Hiscott, R. N., Kenyon, N. H., Lucci, F. Ricci, & Smith, R. D. A.), pp. 94–99. London: Chapman and Hall.CrossRefGoogle Scholar
Weimer, P., Bouma, A. H., & Perkins, B. F. 1994. Submarine Fans and Turbidite Systems. Austin, TX: Society of Economic Paleontologists and Mineralogists Gulf Coast Section.Google Scholar
Weimer, P., & Link, M. H. 1991. Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems. New York: Springer-Verlag.CrossRefGoogle Scholar
Weimer, P., & Posamentier, H. 1993. Siliciclastic Sequence Stratigraphy. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Weimer, R. J., Howard, J. D., & Lindsay, D. R. 1982. Tidal flats. In Sandstone Depositional Environments (ed. Scholle, P. A. & Spearing, D. A.), pp. 191–245. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wellner, R. W., Ashley, G. M., & Sheridan, R. E. 1993. Seismic stratigraphic evidence for a submerged middle Wisconsin barrier: implications for sea-level history. Geology 21, 109–112.2.3.CO;2>CrossRefGoogle Scholar
Werner, B. T. 1995. Eolian dunes: computer simulations and attractor interpretation. Geology 23, 1107–1110.2.3.CO;2>CrossRefGoogle Scholar
Wescott, W. A. 1993. Geomorphic thresholds and complex response of fluvial systems – some implications for sequence stratigraphy. American Association of Petroleum Geologists Bulletin 77, 1208–1218.Google Scholar
Westbroek, P., & Jong, E. W. 1983. Biomineralization and Biological Metal Accumulation. Dordrecht: D. Reidel Publishing Company.CrossRefGoogle Scholar
Whateley, M. K. G., & Pickering, K. T. 1989. Deltas: Sites and Traps for Fossil Fuels. London: Geological Society of London.Google Scholar
Whipple, K. L. 1997. Open-channel flow of Bingham fluids: applications in debris-flow research. Journal of Geology 105, 243–262.CrossRefGoogle Scholar
Whipple, K. X., Hancock, G. S., & Anderson, R. S. 2000. River incision into bedrock: mechanics and relative efficacy of plucking, abrasion, and cavitation. Geological Society of America Bulletin 112, 490–503.2.0.CO;2>CrossRefGoogle Scholar
White, W. A. 1961. Colloid phenomena in sedimentation of argillaceous rocks. Journal of Sedimentary Petrology 31, 560–570.Google Scholar
Wignall, P. B. 1994. Black Shales. Oxford: Oxford University Press.Google Scholar
Wilgus, C. K., Hastings, B. S., Kendall, C. G. St. , C.et al. 1988. Sea-Level Changes: An Integrated Approach. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Wilken, R. T. 2003. Sulfide minerals in sediments. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Middleton, G. V.), pp. 701–703. Dordrecht: Kluwer Academic Publishers.Google Scholar
Wilkinson, B. H. 1979. Biomineralization, paleoceanography and the evolution of calcareous marine organisms. Geology 7, 524–527.2.0.CO;2>CrossRefGoogle Scholar
Willgoose, G. R., Bras, R. L., & Rodriguez-Iturbe, I. 1991a. A physically-based coupled network growth and hillslope evolution model: 1. Theory. Water Resource Research 27, 1,671–1,684.CrossRefGoogle Scholar
Willgoose, G. R., Bras, R. L., & Rodriguez-Iturbe, I. 1991b. A physically-based coupled network growth and hillslope evolution model: 2. Applications. Water Resource Research 27, 1,685–1,696.CrossRefGoogle Scholar
Williams, G. P. 1988. Paleofluvial estimates from dimensions of former channels and meanders. In Encyclopedia of Sediments and Sedimentary Rocks (ed. Baker, V. R., Kochel, R. C., & Patton, P. P.), pp. 321–334. Chichester: Wiley.Google Scholar
Willis, B. J. 1989. Paleochannel reconstructions from point bar deposits: a three-dimensional perspective. Sedimentology 36, 757–766.CrossRefGoogle Scholar
Willis, B. J. 1993a. Ancient river systems in the Himalayan foredeep, Chinji village area, northern Pakistan. Sedimentary Geology 88, 1–76.CrossRefGoogle Scholar
Willis, B. J.1993b. Interpretation of bedding geometry within ancient point-bar deposits In Alluvial Sedimentation (ed. Marzo, M. & Puigdefabregas, C.), pp. 101–114. Oxford: Blackwells.CrossRefGoogle Scholar
Willis, B. J. 1997. Architecture of fluvial-dominated valley-fill deposits in the Cretaceous Fall River Formation. Sedimentology 44, 735–757.CrossRefGoogle Scholar
Willis, B. J.2005. Deposits of tide-influenced river deltas. In River Deltas – Concepts, Models, and Examples (ed. Giosan, L. & Bhattacharya, J. P.), pp. 87–129. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar
Willis, B. J., & Behrensmeyer, A. K. 1994. Architecture of Miocene overbank deposits in Northern Pakistan. Journal of Sedimentary Research B64, 60–67.Google Scholar
Wilson, A. M., Sanford, W., Whitaker, F., & Smart, P. 2001. Spatial patterns of diagenesis during geothermal circulation in carbonate platforms. American Journal of Science 301, 727–752.CrossRefGoogle Scholar
Wilson, E. N., Hardie, L. A., & Phillips, O. M. 1990. Dolomitization front geometry, fluid flow patterns, and the origin of massive dolomite: the Triassic Latemar Buildup, northern Italy. American Journal of Science 290, 741–796.CrossRefGoogle Scholar
Wilson, I. G. 1972a. Aeolian bedforms – their development and origins. Sedimentology 19, 173–210.CrossRefGoogle Scholar
Wilson, I. G. 1972b. Universal discontinuities in bedforms produced by wind. Journal of Sedimentary Petrology 42, 667–669.Google Scholar
Wilson, J. L. 1975. Carbonate Facies in Geologic History. New York: Springer-Verlag.CrossRefGoogle Scholar
Wilson, J. L., & Jordan, C. 1983. Middle shelf. In Carbonate Depositional Environments (ed. Scholle, P. A., Bebout, D. G., & Moore, C. H.), pp. 297–343. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wolf, K. H., Easton, A. J., & Warme, S. 1967. Techniques of examining and analyzing carbonate skeletons, minerals and rocks. In Carbonate Rocks (ed. Chilingar, G. V., Bissell, H. J., & Fairbridge, R. W.), pp. 253–341. Amsterdam: Elsevier.Google Scholar
Wollast, R. 1976. Kinetics of the alteration of K-feldspar in buffered solutions at low temperature. Geochimica et Cosmochimica Acta 31, 635–648.CrossRefGoogle Scholar
Wolman, M. G., & Miller, J. P. 1960. Magnitude and frequency of forces in geomorphic processes. Journal of Geology 68, 54–74.CrossRefGoogle Scholar
Wood, L. J., Ethridge, F. G., & Schumm, S. A. 1993. An experimental study of the influence of subaqueous shelf angles on coastal plain and shelf deposits. In Siliciclastic Sequence Stratigraphy (ed. Weimer, P. & Posamentier, H. W.), pp. 381–391. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Wood, R. A., Grotzinger, J. P., & Dickson, J. A. D. 2002. Proterozoic modular biomineralized metazoan from the Nama Group, Namibia. Science 296, 2383–2386.CrossRefGoogle ScholarPubMed
Woodroffe, C. D. 2003. Coasts; Form, Process and Evolution. Cambridge: Cambridge University Press.Google Scholar
Woodroffe, C. D., Chappell, J., Thom, B. G., & Wallensky, E. 1989. Depositional model of a macrotidal estuary and floodplain, South Alligator River, Northern Australia. Sedimentology 36, 737–756.CrossRefGoogle Scholar
Wright, H. E. Jr., Kutzbach, J. E., Webb, T. IIIet al. 1993. Global Climates since the Last Glacial Maximum. Minneapolis, MN: University of Minnesota Press.Google Scholar
Wright, L. D. 1977. Sediment transport and deposition at river mouths: a synthesis. Geological Society of America Bulletin 88, 857–868.2.0.CO;2>CrossRefGoogle Scholar
Wright, L. D., & Coleman, J. M. 1973. Variations in morphology of major river deltas as functions of ocean wave and river discharge regimes. American Association of Petroleum Geologists Bulletin 57, 370–398.Google Scholar
Wright, L. D., & Coleman, J. M. 1974. Mississippi River mouth processes: effluent dynamics and morphological development. Journal of Geology 82, 751–778.CrossRefGoogle Scholar
Wright, L. D., Coleman, J. M., & Thom, B. G. 1975. Sediment transport and deposition in a macrotidal river channel, Ord River, Western Australia. In Estuarine Research II (ed. Cronin, L. E.), pp. 309–322. New York: Academic Press.Google Scholar
Wright, L. D., & Short, A. D. 1984. Morphodynamic variability of surf zones and beaches: a synthesis. Marine Geology 56, 93–118.CrossRefGoogle Scholar
Wright, V. P. 1986. Paleosols: Their Recognition and Interpretation. Oxford: Blackwell Scientific.Google Scholar
Wright, V. P.1999. Assessing flood duration gradients and fine-scale environmental change on ancient floodplains. In Floodplains: Interdisciplinary Approaches (ed. Marriott, S. B. & Alexander, J.), pp. 279–287. London: Geological Society of London.Google Scholar
Wright, V. P., & Burchette, T. P. 1996. Shallow-water carbonate environments. In Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd edn. (ed. Reading, H. G.), pp. 325–394. Oxford: Blackwells.Google Scholar
Wright, V. P., & Marriott, S. B. 1993. The sequence stratigraphy of fluvial depositional systems: the role of floodplain sediment storage. Sedimentary Geology 86, 203–210.CrossRefGoogle Scholar
Yalin, M. S. 1977. Mechanics of Sediment Transport, 2nd edn. Oxford: Pergamon Press.Google Scholar
Yalin, M. S. 1992. River Mechanics. Oxford: Pergamon Press.Google Scholar
Yao, Q., & Demicco, R. V. 1995. Paleoflow patterns of dolomitizing fluids and paleohydrogeology of southern Canadian Rocky Mountains: evidence from dolomite geometry and numerical modeling. Geology 23, 791–794.2.3.CO;2>CrossRefGoogle Scholar
Yao, Q., & Demicco, R. V. 1997. Dolomitization of the Cambrian carbonate platform, southern Canadian Rocky Mountains: dolomite front geometry, fluid inclusion geochemistry, isotopic signature, and hydrogeologic modeling studies. American Journal of Science 297, 892–938.CrossRefGoogle Scholar
Yarnold, Y. C. 1993. Rock-avalanche characteristics in dry climates and the effects of flow into lakes: insights from mid-Tertiary sedimentary breccias near Artillery Peak, Arizona. Geological Society of America Bulletin 105, 345–360.2.3.CO;2>CrossRefGoogle Scholar
Young, A. 1972. Slopes. Edinburgh: Oliver and Boyd.Google Scholar
Young, I. M., & Crawford, J. W. 2004. Interactions and self-organization in the soil-microbe complex. Science 304, 1634–1637.CrossRefGoogle ScholarPubMed
Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K. 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686–693.CrossRefGoogle ScholarPubMed
Zachos, J. C., Röhl, U., Schellenberg, S. A.et al. 2005. Rapid acidification of the ocean during the Paleocene–Eocene Thermal Maximum. Science 308, 1611–1615.CrossRefGoogle ScholarPubMed
Zajac, I. S. 1974. The Stratigraphy and Mineralogy of the Sokoman Formation in the Knob Lake Area, Quebec and Newfoundland. Ottawa: Geological Survey of Canada.Google Scholar
Zeng, J., & Lowe, D. R. 1997. Numerical simulations of turbidity current flow and sedimentation: I. Theory. Sedimentology 44, 67–84.CrossRefGoogle Scholar
Zharkov, M. A. 1984. Paleozoic Salt Bearing Formations of the World. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Zaitlin, B. A., Dalrymple, R. W., & Boyd, R. 1994. The stratigraphic organization of incised-valley systems associated with relative sea-level change. In Incised Valley Systems: Origin and Sedimentary Sequences (ed. Dalrymple, R. W., Boyd, R., & Zaitlin, B. A.), pp. 45–60. Tulsa, OK: SEPM (Society for Sedimentary Geology).CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • John Bridge, State University of New York, Binghamton, Robert Demicco, State University of New York, Binghamton
  • Book: Earth Surface Processes, Landforms and Sediment Deposits
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511805516.022
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • John Bridge, State University of New York, Binghamton, Robert Demicco, State University of New York, Binghamton
  • Book: Earth Surface Processes, Landforms and Sediment Deposits
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511805516.022
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • John Bridge, State University of New York, Binghamton, Robert Demicco, State University of New York, Binghamton
  • Book: Earth Surface Processes, Landforms and Sediment Deposits
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511805516.022
Available formats
×