Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-18T05:46:13.786Z Has data issue: false hasContentIssue false

Early Jurassic climate change and the radiation of organic-walled phytoplankton in the Tethys Ocean

Published online by Cambridge University Press:  08 April 2016

Bas van de Schootbrugge
Affiliation:
Institute for Marine and Coastal Sciences, Rutgers University, 71 Dudley Road, New Brunswick, New Jersey 08901. E-mail: vandesch@imcs.rutgers.edu
Trevor R. Bailey
Affiliation:
Institute for Marine and Coastal Sciences, Rutgers University, 71 Dudley Road, New Brunswick, New Jersey 08901. E-mail: vandesch@imcs.rutgers.edu
Yair Rosenthal
Affiliation:
Institute for Marine and Coastal Sciences, Rutgers University, 71 Dudley Road, New Brunswick, New Jersey 08901. E-mail: vandesch@imcs.rutgers.edu
Miriam E. Katz
Affiliation:
Department of Geological Sciences, Rutgers University, 610 Taylor Road, Piscataway, New Jersey 08854
James D. Wright
Affiliation:
Department of Geological Sciences, Rutgers University, 610 Taylor Road, Piscataway, New Jersey 08854
Kenneth G. Miller
Affiliation:
Department of Geological Sciences, Rutgers University, 610 Taylor Road, Piscataway, New Jersey 08854
Susanne Feist-Burkhardt
Affiliation:
Natural History Museum, Paleontology Department, Cromwell Road, London SW7 5BD, United Kingdom
Paul G. Falkowski
Affiliation:
Department of Geological Sciences, Rutgers University, 610 Taylor Road, Piscataway, New Jersey 08854

Abstract

During the Early Jurassic, cyst-forming dinoflagellates began a long-term radiation that would portend ecological importance of these taxa in the pelagic plankton community throughout the rest of the Mesozoic era. The factors that contributed to the evolutionary success of dinoflagellates are poorly understood. Here we examine the relationship between oceanographic and climatic conditions during the Hettangian–Toarcian interval in relation to the radiation of dinoflagellates and other organic-walled phytoplankton taxa in the Tethys Ocean. Our analysis is based on two data sets. The first includes δ13Ccarb, δ13Corg, total organic carbon (TOC), and quantitative palynological observations derived from the Mochras Core (Wales, U.K.), which spans the complete Early Jurassic. The second is a coupled Mg/Ca and δ18O record derived from analyses of belemnite calcite obtained from three sections in northern Spain, covering the upper Sinemurian to Toarcian. From these two data sets we reconstructed the influence of sea level, trophism, temperature, and salinity on dinoflagellate cyst abundance and diversity in northwest Europe. Our results suggest that organic-walled phytoplankton (acritarchs, prasinophytes, and dinoflagellates) diversity increased through the Early Jurassic. The radiation coincides with a long-term eustatic rise and overall increase in the areal extent of continental shelves, a factor critical to cyst germination. On shorter timescales, we observed short bursts of dinoflagellate diversification during the late Sinemurian and late Pliensbachian. The former diversification is consistent with the opening of the Hispanic Corridor during the late Sinemurian, which apparently allowed the pioneer dinoflagellate, Liasidium variabile, to invade the Tethys from the Paleo-Pacific. A true radiation pulse during the late Pliensbachian, with predominantly cold-water taxa, occurred during sea level fall, suggesting that climate change was critical to setting the evolutionary tempo. Our belemnite δ18O and Mg/Ca data indicate that late Pliensbachian water masses cooled (ΔT ≈ −6°C) and became more saline (ΔS ≈ +2 psu). Cooling episodes during generally warm and humid Early Jurassic climate conditions would have produced stronger winter monsoon northeast trade winds, resulting in hydrographic instability, increased vertical mixing, and ventilation of bottom waters. During the late Pliensbachian, dinoflagellates replaced green algae, including prasinophytes and acritarchs, as primary producers. By producing benthic resting cysts, dinoflagellates may have been better adapted to oxidized ocean regimes. This hypothesis is supported by palynological data from the early Toarcian ocean anoxic event, which was marked by highly stratified anoxic bottom water overlain by low-salinity, warm surface waters. These conditions were advantageous to green algae, while cyst-producing dinoflagellates temporarily disappeared. Our results suggest that the rise in dinoflagellate diversity later in the Jurassic appears to correspond to deep water ventilation as a result of the opening of the Atlantic seaway, conditions that appear to have simultaneously led to a loss of prasinophyte dominance in the global oceans.

Type
Articles
Copyright
Copyright © The Paleontological Society 

Access options

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

References

Literature Cited

Aberhan, M. 2001. Bivalve palaeobiogeography and the Hispanic Corridor: time of opening and effectiveness of a proto-Atlantic seaway. Palaeogeography, Palaeoclimatology, Palaeoecology 165:375394.Google Scholar
Aberhan, M. 2002. Opening of the Hispanic Corridor and Early Jurassic bivalve biodiversity. In Crame, J. A. and Owen, A.W., eds. Palaeobiogeography and biodiversity change: the Ordovician and Mesozoic-Cenozoic radiations. Geological Society of London Special Publication 194:127139.Google Scholar
Ainsworth, N. R., O'Neill, M., Rutherford, M. M., Clayton, G., Horton, N. F., and Penney, R. A. 1987. Biostratigraphy of the Lower Cretaceous, Jurassic and uppermost Triassic of the North Celtic Sea and Fastnet Basins. Pp. 611622in Brooks, J. and Glennie, K., eds. Petroleum geology of North West Europe. Graham and Trotman, London.Google Scholar
Anderson, T. F., and Arthur, M. A. 1983. Stable isotopes of oxygen and carbon and their application to sedimentologic and paleoenvironmental problems. SEPM Short Course 10:1.11.151.Google Scholar
Bailey, T. R., Rosenthal, Y., McArthur, J. M., van de Schootbrugge, B., and Thirlwall, M. F. 2003. Paleoceanographic changes of the late Pliensbachian–Early Toarcian interval: a possible link to the genesis of an Oceanic Anoxic Event. Earth and Planetary Science Letters 212:307320.Google Scholar
Beaudoin, F., Herbin, J. P., Bassoullet, J. P., Dercourt, J., Lachkar, G., Manivit, H., and Renard, M. 1990. Distribution of organic matter during the Toarcian in the Mediterranean Tethys and Middle East. In Huc, Y., ed. Deposition of organic facies. Association of American Petroleum Geologists, Studies in Geology 30:7392.Google Scholar
Below, R. 1987. Evolution und Systematik von Dinoflagellaten-Zysten aus der Ordnung Peridiniales. I. Allgemeine Grundlagen und Subfamilie Rhaetogonyaulacoideae (Familie Peridiniaceae). Palaeontographica, Abteilung B 205(1–6):1164.Google Scholar
Bjerrum, C. J., Surlyk, F., Callomon, J. H., and Slingerland, R. L. 2001. Numerical paleoceanographic study of the Early Jurassic transcontinental Laurasian Seaway. Paleoceanography 16:390404.Google Scholar
Blomeier, D. P. G., and Reijmer, J. G. 1999. Drowning of a Lower Jurassic carbonate platform: Jbel Bou Dahar, High Atlas, Morocco. Facies 41:81110.Google Scholar
Bonijoly, D., Perrin, J., Roure, F., Bergerat, F., Courel, L., Elmi, S., A. Mignot, and the GPF Team. 1996. The Ardeche palaeomargin of the South-East Basin of France: Mesozoic evolution of a part of the Tethyan continental margin (Géologie Profonde de la France programme). Marine and Petroleum Geology 13:607623.Google Scholar
Boomer, I. D. 1991. Lower Jurassic ostracod biozonation of the Mochras Borehole. Journal of Micropalaeontology 9:205218.Google Scholar
Boomer, I. D., and Whatley, R. 1992. Ostracoda and dysaerobia in the Lower Jurassic of Wales: the reconstruction of past oxygen levels. Palaeogeography, Palaeoclimatology, Palaeoecology 99:373379.Google Scholar
Boomer, I. D., Whatley, R., Bassi, D., Fugagnoli, A., and Loriga, C. 2001. An Early Jurassic oligohaline ostracod assemblage within the carbonate platform sequence of the Venetian Prealps, NE Italy. Palaeogeography, Palaeoclimatology, Palaeoecology 166:331344.Google Scholar
Borrego, A. G., Hagemann, H. W., Blanco, C. G., Valenzuela, M., and de Centi, C. Suarez 1996. The Pliensbachian (Early Jurassic) anoxic events in Asturias, northern Spain: Santa Mera Member, Rodiles Formation. Organic Geochemistry 25(5–7):295309.Google Scholar
Bown, P. R., Burnett, J. A., and Gallagher, L. T. 1992. Calcareous nannoplankton evolution. Memorie di Scienze Geologiche, Padova, XLIII:117.Google Scholar
Braga, J. C., Comas-Rengifo, M. J., Goy, A., Rivas, P., and Yebenes, A. 1988. El Lias inferior y medio en zona central de la cuenca Vasco-Cantabrica (Camino, Santander). Ciencias de la Tierra Geologia 11 (III Colloquio de estratigrafia y paleogeografia del Jurassico de España):1743.Google Scholar
Palliani, R. Bucefalo, and Riding, J. B. 1999. Relationships between the Early Toarcian anoxic event and organic-walled phytoplankton in central Italy. Marine Micropaleontology 37:101116.CrossRefGoogle Scholar
Palliani, R. Bucefalo, and Riding, J. B. 2000. A palynological investigation of the Lower and lowermost Middle Jurassic strata (Sinemurian to Aalenian) from North Yorkshire, UK. Proceedings of the Yorkshire Geological Society 53:116.Google Scholar
Palliani, R. Bucefalo, and Riding, J. B. 2003. Biostratigraphy, provincialism and evolution of European Early Jurassic (Pliensbachian to Early Toarcian) dinoflagellate cysts. Palynology 27:179214.Google Scholar
Palliani, R. Bucefalo, Mattioli, E., and Riding, J. B. 2002. The response of marine phytoplankton and sedimentary organic matter to the Early Toarcian (Lower Jurassic) oceanic anoxic event in northern England. Marine Micropaleontology 46:223245.Google Scholar
Butterfield, N. J., and Rainbird, R. H. 1998. Diverse organic-walled fossils, including possible dinoflagellates, from the Early Neoproterozoic of Arctic Canada. Geology 26:963966.Google Scholar
Cobianchi, M., and Picotti, V. 2001. Sedimentary and biological response to sea level and palaeoceanographic changes of a Lower-Middle Jurassic Tethyan platform margin (Southern Alps, Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 169:219244.Google Scholar
Comas-Rengifo, M. J., Goy, A., Rivas, P., and Yebenes, A. 1988. El Toarciense en Castillo Pedroso (Santander). Ciencias de la Tierra Geologia 11 (III Colloquio de estratigrafia y paleogeografia del Jurassico de España):6371.Google Scholar
Crevello, P. D. 1990. Stratigraphic evolution of Lower Jurassic carbonate platforms: Record of rift tectonics and eustasy, central and eastern High Atlas, Morocco. Ph.D. dissertation. Colorado School of Mines, Golden.Google Scholar
Dale, B. 1976. Cyst formation, sedimentation, and preservation: factors affecting dinoflagellate assemblages in Recent sediments from Trondheimsfjord, Norway. Review of Paleobotany and Palynology 22:3960.Google Scholar
Dale, B. 1983. Dinoflagellate resting cysts: ‘benthic plankton.’ Pp. 69136in Fryxell, G. A., ed. Survival strategies of the Algae. Cambridge University Press, Cambridge.Google Scholar
Dybkjaer, K. 1988. Palynological zonation and stratigraphy of the Jurassic section in the Gassum No.1-borehole, Denmark. Geological Survey of Denmark (DGU) A 21:573.Google Scholar
Espitalié, J., Deroo, G., and Marquis, F. 1985a. La pyrolyse Rock-Eval et ses applications. Partie 1. Revue de l'Institut Français du Pétrole 40:563579.Google Scholar
Espitalié, J., Deroo, G., and Marquis, F. 1985b. La pyrolyse Rock-Eval et ses applications. Partie 2. Revue de l'Institut Français du Pétrole 40:755784.Google Scholar
Espitalié, J., Deroo, G., and Marquis, F. 1986. La pyrolyse Rock-Eval et ses applications. Partie 3. Revue de l'Institut Français du Pétrole 41:7389.Google Scholar
Falkowski, P. G., Schofield, O., Katz, M. E., van de Schootbrugge, B. and Knoll, A. H.In press. Why is the land green and the ocean red? In Young, J. and Thierstein, H., eds. Coccolithophorids: from molecular processes to global impact. Springer.Google Scholar
Fauconnier, D. 1995. Jurassic palynology from a borehole in the Champagne area, France: correlation of the Lower Callovian–middle Oxfordian using sequence stratigraphy. Review of Paleobotany and Palynology 87:1526.Google Scholar
Feist-Burkhardt, S. 1992. Palynological investigations in the Lower and Middle Jurassic of Switzerland, France and Germany: palynofacies and type of organic matter, dinoflagellate cyst morphology and stratigraphy. . Université de Genève, Geneva.Google Scholar
Feist-Burkhardt, S. 1995a. Stratigraphic compilation of Below's data (1987a 1987b, and 1990) on Early and Middle Jurassic dinoflagellate cysts. Revue de Paléobiologie 13:313318.Google Scholar
Feist-Burkhardt, S. 1995b. Weiachia fenestrata gen. et sp. nov., a new Phallocystean dinoflagellate cyst from the Lower Jurassic of Switzerland. Palynology 19:211219.Google Scholar
Feist-Burkhardt, S. 1998. Palynostratigraphic characterization of the Sinemurian–Pliensbachian transition of the potential GSSP section at Aubach/Aselfingen, southwest Germany. Fifth international symposium on the Jurassic System, August, 12–25 1998, Vancouver, Abstracts and program, p. 29.Google Scholar
Fensome, R. A., Williams, G. L., Barss, M. S., Freeman, J. M., and Hill, J. M. 1990. Acritarchs and fossil prasinophytes: an index to genera, species and intraspecific taxa. American Association of Stratigraphic Palynologists Contributions Series No. 25.Google Scholar
Fensome, R. A., Taylor, F. J. R., Norris, G., Sarjeant, W. A. S., Wharton, D. I., and Williams, G. L. 1993. A classification of fossil and living dinoflagellates. Micropaleontology Press Special Publication 7.Google Scholar
Fensome, R. A., MacRae, R. A., Moldowan, J. M., Taylor, F. J. R., and Williams, G. L. 1996. The Early Mesozoic radiation of dinoflagellates. Paleobiology 22:329338.Google Scholar
Fensome, R. A., Saldarriaga, J. F., Taylor, F. J. R. 1999. Dinoflagellate phylogeny revisited: reconciling morphological and molecular based phylogenies. Grana 38:6680.Google Scholar
Gorin, G. E., and Feist-Burkhardt, S. 1990. Organic facies of Lower to Middle Jurassic sediments in the Jura Mountains, Switzerland. Review of Palaeobotany and Palynology 65:349355.Google Scholar
Gustomesov, V. A. 1978. The pre-Jurassic ancestry of Belemnitida and the evolutionary changes in the Belemnoidea at the boundary between the Triassic and the Jurassic. Palaeontological Journal 3:283292.Google Scholar
Hallam, A. 2001. A review of the broad pattern of Jurassic sea-level changes and their possible causes in the light of current knowledge. Palaeogeography, Palaeoclimatology, Palaeoecology 167:2337.Google Scholar
Haq, B. U., Hardenbol, J., and Vail, P. R. 1987. Chronology of fluctuating sea levels since the Triassic. Science 255:11561167.Google Scholar
Hesselbo, S. P., Gröcke, D. R., Jenkyns, H. C., Bjerrum, C. J., Farrimond, P., Bell, H. S. Morgans, and Green, O. R. 2000. Massive dissociation of gas hydrate during a Jurassic oceanic anoxic event. Nature 406:392395.Google Scholar
Hochüli, P., and Frank, S. M. 2000. Palynology (dinoflagellate cysts, spore-pollen) and stratigraphy of the Lower Carnian Raibl Group in the Eastern Swiss Alps. Eclogae Geologicae Helvetiae 93:429443.Google Scholar
Hollander, D. J., Bessereau, G., Belin, S., Houzay, J. P., and Huc, A. Y. 1991. Organic matter in the Early Toarcian shales, Paris Basin: a response to environmental change. Review de l'Institut Français du Petrole 46:543562.Google Scholar
Jenkyns, H. C. 1988. The Early Toarcian (Jurassic) anoxic event: stratigraphic, sedimentary, and geochemical evidence. American Journal of Science 288:101151.Google Scholar
Jenkyns, H. C., and Clayton, C. J. 1986. Black shales and carbon isotopes in pelagic sediments from the Tethyan Lower Jurassic. Sedimentology 33:87106.Google Scholar
Jenkyns, H. C., and Clayton, C. J. 1997. Lower Jurassic epicontinental carbonates and mudstones from England and Wales: chemostratigraphic signals and the Early Toarcian anoxic event. Sedimentology 44:687706.Google Scholar
Jenkyns, H. C., Gröcke, D. R., and Hesselbo, S. P. 2001. Nitrogen isotope evidence for water mass denitrification during the Early Toarcian (Jurassic) oceanic anoxic event. Palaeoceanography 16:111.Google Scholar
Jimenez, A. P., de Cisneros, C. Jimenez, Rivas, P., and Vera, J. A. 1996. The Early Toarcian anoxic event in the westernmost Tethys (Subbetic): paleogeographic and paleobiogeographic significance. Journal of Geology 104:399416.Google Scholar
Katz, M. E., Wright, J. D., Miller, K. G., Cramer, B. S., Fennel, K., and Falkowski, P. G.In press. Biological overprint of the geological carbon cycle. Marine Geology.Google Scholar
Kump, L. R., and Arthur, M. A. 1999. Interpreting carbon-isotope excursions: carbonates and organic matter. Chemical Geology 161:181198.Google Scholar
Kump, L. R., Arthur, M. A., Patzkowsky, M. E., Gibbs, M. T., Pinkus, D. S., and Sheehan, P. M. 1999. A weathering hypothesis for glaciation at high atmospheric pCO2 during the Late Ordovician. Palaeogeography, Palaeoclimatology, Palaeoecology 152:173187.Google Scholar
Küspert, W. 1982. Environmental changes during oil shale deposition as deduced from stable isotope ratios. Pp. 482501in Einsele, G. and Seilacher, A., eds. Cyclic and event stratification. Springer, Berlin.Google Scholar
Langford, F. F., and Blanc-Valleron, M. M. 1990. Interpreting Rock-Eval pyrolysis data using graphs of pyrolyzable hydrocarbons vs. total organic carbon. American Association Petroleum Geologists Bulletin 74:709806.Google Scholar
Lear, C. H., Rosenthal, Y., Slowey, N. 2002. Benthic foraminiferal Mg/Ca paleothermometry: a revised core-top calibration. Geochimica et Cosmochimica Acta 66:33753387.CrossRefGoogle Scholar
Little, C. T. S., and Benton, M. J. 1995. Early Jurassic mass extinction: a global long-term event. Geology 23:495498.Google Scholar
Lund, J. J. 2003. Rhaetian to Pliensbachian palynostratigraphy of the central part of the NW German Basin exemplified by the Eitzendorf 8 well. Courier Forschungsinstitut Senckenberg 241:6983.Google Scholar
Mallarino, G., Goldstein, R. H., and Di Stefano, P. 2002. New approach for qualifying water depth applied to the enigma of drowning carbonate platforms. Geological Society of America Bulletin 30:783786.Google Scholar
Martin, R. E. 1996. Secular increase in nutrient levels through the Phanerozoic: implications for productivity, biomass, diversity of the marine biosphere. Palaios 11:209219.Google Scholar
Martini, R., Zaninetti, L., Villeneuve, M., Cornee, J. J., Krystyn, L., Cirilli, S., De Wever, P., Dumitrica, P., and Harsolumakso, A. 2000. Triassic pelagic deposits of Timor: palaeogeographic and sea-level implications. Palaeogeography, Palaeoclimatology, Palaeoecology 160:123151.Google Scholar
McArthur, J. M., Donovan, D. T., Thirlwall, M. F., Fouke, B. W., and Mattey, D. 2000. Strontium isotope profile of the Early Toarcian (Jurassic) oceanic anoxic event, duration of ammonite biozones, and belemnite palaeotemperatures. Earth and Planetary Science Letters 179:269285.Google Scholar
Moldowan, J. M., and Talyzina, N. M. 1998. Biogeochemical evidence for dinoflagellate ancestors in the Early Cambrian. Science 281:11681170.Google Scholar
Morbey, S. J. 1978. Late Triassic and Early Jurassic subsurface palynostratigraphy in Northwestern Europe. Palinologia 1:355365.Google Scholar
Morettini, E., Santantonio, M., Bartolini, A., Cecca, F., Baumgartner, P. O., and Hunziker, J. C. 2002. Carbon isotope stratigraphy and carbonate production during the Early-Middle Jurassic: examples from the Umbria-Marche-Sabina Apennines (central Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 184:251273.Google Scholar
Müller, F. L. 1990, The paleoecology of the Liassic benthic foraminifera of Great Britain. Ph.D. dissertation. Rutgers University, New Brunswick, NJ.Google Scholar
Partington, M. A., Copestake, P., Mitchener, B. C., and Underhill, J. R. 1993. Biostratigraphic calibration of genetic stratigraphic sequences in the Jurassic-lowermost Cretaceous (Hettangian to Ryazanian) of the North Sea and adjacent areas. Pp. 371386in Parker, J. R. ed. Petroleum geology of north-west Europe: proceedings of the 4th conference. Geological Society of London.Google Scholar
Pfiester, L. A., and Anderson, D. M. 1987. Dinoflagellate reproduction. Pp. 611648in Taylor, F. J. R., ed. The biology of dinoflagellates. Botanical Monographs 21:611648.Google Scholar
Poulsen, N. E., and Riding, J. B. 2003. The Jurassic dinoflagellate cyst zonation of Subboreal north-west Europe. With a supplement by B. Buchardt: oxygen isotope palaeotemperatures from the Jurassic in Northwest Europe. In Ineson, J. and Surlyk, F., eds. The Jurassic of Denmark and Greenland. Geological Survey of Denmark and Greenland Bulletin 1:115144.Google Scholar
Prauss, M., and Riegel, W. 1989. Evidence of phytoplankton associations for causes of black shale formation in epicontinental seas. Neues Jahrbuch für Geologie und Paläontologie, Monatshefte 11:671682.Google Scholar
Prauss, M., Ligouis, B. and Luterbacher, H. 1991. Organic matter and palynomorphs in the ‘Posidonienschiefer’ (Toarcian, Lower Jurassic) of southern Germany. Pp. 335352in Tyson, R. V. and Pearson, T. H., eds. Modern and ancient continental shelf anoxia. Geological Society, London.Google Scholar
Price, G. D. 1999. The evidence and implications of polar ice during the Mesozoic. Earth-Science Reviews 48:183210.Google Scholar
Quesada, S., Dorronsoro, C., Robles, S., Chaler, R., and Grimalt, J. O. 1997. Geochemical correlation of oil from the Ayoluengo field to Liassic shale units in the southwestern Basque-Cantabrian Basin (northern Spain). Organic Geochemistry, 27(1–2):2540.Google Scholar
Quigg, A., Finkel, Z. V., Irwin, A., Rosenthal, Y., Ho, T. Y., Reinfelder, J. R., Schofield, O., Morel, F. M. M., and Falkowski, P. G. 2003. The evolutionary inheritance of elemental stoichiometry in marine phytoplankton. Nature 425:291294.Google Scholar
Rasmussen, E. S., Lomholt, S., Andersen, C., and Vejback, O. V. 1998. Effects of structural evolution of the Lusitanian Basin in Portugal and the shelf and slope area offshore Portugal. Tectonophysics 300:199225.Google Scholar
Rauscher, R., and Schmitt, J.-P. 1990. Recherches palynologiques dans le Jurassique d'Alsace (France). Review of Palaeobotany and Palynology 62:107156.Google Scholar
Riding, J. B. 1984. A palynological investigation of Toarcian and Early Aalenian strata from the Blea Wyke area, Ravenscar, North Yorkshire. Proceeding of the Yorkshire Geological Society 45:109122.Google Scholar
Riding, J. B. 1987. Dinoflagellate cyst stratigraphy of the Nettleton Bottom Borehole (Jurassic: Hettangian to Kimmeridgian), Lincolnshire, England. Proceeding of the Yorkshire Geological Society 46:231266.Google Scholar
Riding, J. B., and Hubbard, R. N. L. B. 1999. Jurassic (Toarcian-Kimmeridgian) dinoflagellate cysts and paleoclimates. Palynology 23:1530.Google Scholar
Riding, J. B., and Ioannides, N. S. 1996. A review of Jurassic dinoflagellate cyst biostratigraphy and global provincialism. Bulletin de la Société Géologique de France 167:314.Google Scholar
Röhl, H., Schmid-Röhl, A., Oschmann, W., Frimmel, A., and Schwark, L. 2001. The Posidonia Shale (lower Toarcian) of SW-Germany: an oxygen-depleted ecosystem controlled by sea level and palaeoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology, 165(1–2):2752.Google Scholar
Rosales, I., Quesada, S., and Robles, S. 2001. Primary and diagenetic isotopic signals in fossils and hemipelagic carbonates: the Lower Jurassic of northern Spain. Sedimentology 48:11491169.Google Scholar
Rosenthal, Y., Boyle, E. A., and Slowey, N. 1997. Temperature control on the incorporation of magnesium, strontium, fluorine, and cadmium into benthic foraminiferal shells from Little Bahama Bank: prospects for thermocline paleoceanography. Geochimica et Cosmochimica Acta 61:36333643.Google Scholar
Rosenthal, Y., Field, M. P., and Sherrell, R. M. 1999. Precise determination of element/calcium ratios in calcareous samples using sector field inductively coupled plasma mass spectrometry. Analytical Chemistry 71:32483253.Google Scholar
Sandy, M. R., and Stanley, G. D. Jr. 1993. Late Triassic brachiopods from the Luning Formation, Nevada, and their palaeobiogeographical significance. Palaeontology 36:439480.Google Scholar
Schouten, S., van Kaam-Peters, M. E., Rijpstra, I., Schoell, M., and Damste, J. S. Sinnighe 2000. Effects of an oceanic anoxic event on the stable carbon isotopic composition of Early Toarcian carbon. American Journal of Science 300:122.Google Scholar
Smith, P. L., and Tipper, H. W. 1986. Plate tectonics and paleobiogeography: Early Jurassic (Pliensbachian) endemism and diversity. Palaios 1:399412.Google Scholar
Stover, L. E., Brinkhuis, H., Damassa, S. P., De Verteuil, L., Helby, R. J., Monteil, E., Patridge, A. D., Powell, A. J., Riding, J. B., Smelror, M., and Williams, G. L. 1996. Mesozoic-Tertiary dinoflagellates, acritarchs and prasinophytes. Pp. 641750in Jansonius, J. and McGregor, D. C., eds. Palynology: principles and applications, Vol. 2. American Association of Stratigraphic Palynologists Foundation, College Station, Tex. Vega, L. C. Suarez1974. Estratigrafia del Jurassico en Asturias. Cuadernos Geologicos Iberica 3:1–368.Google Scholar
Taylor, K. G. 1998. Spatial and temporal variations in early diagenetic organic matter oxidation pathways in Lower Jurassic mudstones of eastern England. Chemical Geology 145:4760.Google Scholar
Tyson, R. V. 1995. Sedimentary organic matter. Chapman and Hall, London.Google Scholar
Vakhrameev, V. A. 1981. Pollen Classopollis: indicator of Jurassic and Cretaceous climate. Palaeobotanist 28–29:301307.Google Scholar
Van de Schootbrugge, B., Föllmi, K. B., Bulot, L. G., and Burns, S. J. 2000. Paleoceanographic changes during the Early Cretaceous (Valanginian-Hauterivian): evidence from oxygen and carbon stable isotopes. Earth and Planetary Science Letters 181:1531.Google Scholar
Wall, D. 1965. Microplankton, pollen and spores from the Lower Jurassic in Britain. Micropalaeontology 11:151–90.Google Scholar
Wall, D., Dale, B., Lohmann, G. P., and Smith, W. K. 1977. The environment and climatic distribution of dinoflagellate cysts in modern marine sediments from regions in the North and South Atlantic Oceans and adjacent seas. Marine Micropalaeontology 2:121200.Google Scholar
Weissert, H., Lini, A., Föllmi, K. B., and Kuhn, O. 1998. Correlation of Early Cretaceous carbon isotope stratigraphy and platform drowning events: a possible link? Palaeogeography, Palaeoclimatology, Palaeoecology 137:189203.Google Scholar
Weiß, M. 1986. Liasidium variabile, eine Dinoflagellate mit stratigraphischem Wert an der Grenze Unter-/Ober-Sinemurium. Neues Jahrbuch für Geologie und Paläontologie, Monatshefte 5:317320.Google Scholar
Wetzel, A., Allenbach, R., and Allia, V. 2003. Reactivated basement structures affecting the sedimentary facies in a tectonically “quiescent” epicontinental basin: an example from NW Switzerland. Sedimentary Geology 157:153172.Google Scholar
Wille, W. 1982a. Evolution and ecology of upper Liassic dinoflagellates from SW-Germany. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 164(1/2):7482.Google Scholar
Wille, W. 1982b. Palynology of upper Liassic bituminous shales. Pp. 505in Einsele, G. and Seilacher, A., eds. Cyclic and event stratification. Springer, Berlin.Google Scholar
Wille, W., and Gocht, H. 1979. Dinoflagellaten aus dem Lias Südwestdeutschlands, Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen221258.Google Scholar
Williams, G. L., Ascoli, P., Barss, M. S., Bujak, J. P., Davies, E. H., Fensome, R. A., and Williamson, M. A. 1990. Chapter 3: Biostratigraphy and related studies. In Keen, M. J. and Williams, G. L., eds. Geology of the continental margin of eastern Canada. Geology of Canada 2:86137. Geological Survey of Canada, Ottawa.Google Scholar
Williams, G. L., Lentin, J. K., and Fensome, R. A. 1998. The Lentin and Williams index of fossil dinoflagellates 1998 edition. American Association of Stratigraphic Palynologists, Contributions Series 34:817.Google Scholar
Woodland, A. W. 1971. The Llanbedr (Mochras Farm) Borehole. Report No. 71/18. Institute of Geological Sciences, London.Google Scholar