Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-16T04:51:55.672Z Has data issue: false hasContentIssue false

Climatic evolution across oceanic anoxic event 1a derived from terrestrial palynology and clay minerals (Maestrat Basin, Spain)

Published online by Cambridge University Press:  30 October 2014

JEAN CORS
Affiliation:
Institute of Geology, Leibniz University Hannover, Callinstraße 30, 30167 Hannover, Germany
ULRICH HEIMHOFER*
Affiliation:
Institute of Geology, Leibniz University Hannover, Callinstraße 30, 30167 Hannover, Germany
THIERRY ADATTE
Affiliation:
Institute of Geology and Palaeontology, Université de Lausanne, CH-1015 Lausanne, Switzerland
PETER A. HOCHULI
Affiliation:
Palaeontological Institute and Museum, University of Zurich, Karl Schmid-Str. 4, Ch-8006 Zurich, Switzerland
STEFAN HUCK
Affiliation:
Institute of Geology, Leibniz University Hannover, Callinstraße 30, 30167 Hannover, Germany
TELM BOVER-ARNAL
Affiliation:
Departament de Geoquímica, Petrologia i Prospecció Geòlogica, Facultat de Geologia, Universitat de Barcelona, c/ de Martí i Franquès s/n, 08028 Barcelona, Spain
*
Author for correspondence: heimhofer@geowi.uni-hannover.de

Abstract

Studies dealing with the response of the continental biosphere to the environmental perturbations associated with Cretaceous oceanic anoxic events (OAEs) are comparatively rare. Here, a quantitative spore-pollen record combined with clay mineral data is presented, which covers the entire early Aptian OAE 1a interval (Forcall Formation, Maestrat basin, east Spain). The well-expressed OAE 1a carbon-isotope anomaly is paralleled by changes in the clay mineral assemblage and by a stepwise decline in the normalized frequency of Classopollis pollen (produced by xerophytic Cheirolepidiaceae) with lowest contents occurring during the positive δ13C shift. In contrast, Araucariacites and Inaperturopollenites pollen show a pronounced increase in relative abundance from low background values to become a significant component of the palynological assemblage during the Classopollis minimum. The observed changes in clay minerals and pollen distribution patterns are interpreted to reflect a major change in the composition of the hinterland vegetation of the Maestrat Basin, most probably due to short-lived but pronounced climatic cooling and changes in humidity. Temperature anomalies driven by organic carbon burial and associated CO2 decline have been postulated for all major Mesozoic OAEs. The palynomorph record from the Iberian Maestrat basins indicates that the climax of this cooling episode was significantly delayed in comparison to the end of organic carbon-rich deposition in the world oceans.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2014 

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

Abbink, O. A. 1998. Palynological investigations in the Jurassic of the North Sea region. Ph.D. thesis, University of Utrecht, Utrecht, Netherlands. Published thesis.Google Scholar
Abbink, O. A., Van Konijnenburg – Van Cittert, J. H. A. & Visscher, H. 2004. A sporomorph ecogroup model for the Northwest European Jurassic – Lower Cretaceous: concepts and framework. Geologie en Mijnbouw 83, 1731.Google Scholar
Adatte, T., Keller, G. & Stinnesbeck, W. 2002. Late Cretaceous to Early Paleogene climate and sea-level fluctuations: the Tunisian record. Palaeogeography, Palaeoclimatology, Palaeoecology 178, 165–96.CrossRefGoogle Scholar
Ando, A., Kaiho, K., Kawahata, H. & Kakegawa, T. 2008. Timing and magnitude of early Aptian extreme warming: unraveling primary δ18O variation in indurated pelagic carbonates at Deep Sea Drilling Project Site 463, central Pacific Ocean. Palaeogeography, Palaeoclimatology, Palaeoecology 269, 463–76.Google Scholar
Basset, K. N. & Kleinspehn, K. L. 1997. Early to middle Cretaceous paleogeography of north-central British Columbia: Stratigraphy and basin analysis of the Skeena Group Canadian Journal of Earth Sciences 34, 1644–69.CrossRefGoogle Scholar
Bellanca, A., Erba, E., Neri, R., Premoli, S. I., Sprovieri, M., Tremolada, F. & Verga, D. 2002. Palaeoceanographic significance of the Tethyan “Livello Selli” (Early Aptian) from the Hybla Formation, northwestern Sicily: biostratigraphy and high-resolution chemostratigraphic records. Palaeogeography, Palaeoclimatology, Palaeoecology 185, 175–96.CrossRefGoogle Scholar
Bonis, N. R. & Kürschner, W. M. 2012. Vegetation history, diversity patterns, and climate change across the Triassic/Jurassic boundary. Paleobiology 38, 240–64.CrossRefGoogle Scholar
Bonis, N. R., Kürschner, M. W. & Krystyn, L. 2009. A detailed palynological study of the Triassic-Jurassic transition in key sections of the Eiberg Basin (Northern Calcareous Alps, Austria). Review of Palaeobotany and Palynology 156, 376400.Google Scholar
Bottini, C., Cohen, A. S., Erba, E., Jenkyns, H. C. & Coe, A. L. 2012. Osmium-isotope evidence for volcanism, weathering, and ocean mixing during the early Aptian OAE 1a. Geology 40, 583–6.Google Scholar
Bover-Arnal, T., Moreno-Bedmar, J. A., Salas, R., Skelton, P. W., Bitzer, K. & Gili, E. 2010. Sedimentary evolution of an Aptian syn-rift carbonate system (Maestrat Basin, E Spain): effects of accommodation and environmental change. Geologica Acta 8, 249–80.Google Scholar
Bover-Arnal, T., Salas, R., Martin-Closas, C., Schlagintweit, F. & Moreno-Bedmar, J. A. 2011. Expression of an oceanic anoxic event in a neritic setting: Lower Aptian coral rubble deposits from the Western Maestrat Basin (Iberian Chain, Spain). Palaios 26, 1832.Google Scholar
Bover-Arnal, T., Salas, R., Moreno-Bedmar, J. A. & Bitzer, K. 2009. Sequence stratigraphy and architecture of a late Early-Middle Aptian carbonate platform succession from the western Maestrat Basin (Iberian Chain, Spain). Sedimentary Geology 219, 280301.Google Scholar
Bralower, T. J., Arthur, M. A., Leckie, R. M., Sliter, W. V., Allard, D. J. & Schlanger, S. O. 1994. Timing and paleoceanography of oceanic dysoxia/anoxia in the Late Barremian to Early Aptian (Early Cretaceous). Palaios 9, 335–69.Google Scholar
Bustin, R. M. & Smith, G. G. 1993. Coal deposits in the front ranges and foothills of the Canadian Rocky Mountains, southern Canadian Cordillera. International Journal of Coal Geology 23, 127.Google Scholar
Canérot, J., Cugny, P., Pardo, G., Salas, R. & Villena, J. 1982. Ibérica central-maestrazgo. In El Cretácico de España (ed. García, A.), pp. 273344. Universidad Complutense de Madrid, Madrid.Google Scholar
Chumakov, N. M., Zharkov, M. A., Herman, A. B., Doludenko, M. P., Kalandadze, N. M., Lebedev, E. L., Ponomarenko, A. G. & Rautian, A. S. 1995. Climatic belts of the mid-Cretaceous time. Stratigraphy and Geological Correlation 3, 241–60.Google Scholar
Diéguez, C., Peyrot, D. & Barrón, E. 2010. Floristic and vegetational changes in the Iberian Peninsula during Jurassic and Cretaceous. Review of Palaeobotany and Palynology 162, 325–40.Google Scholar
Doyle, J. A., Jardiné, S. & Doerenkamp, A. 1982. Afropollis, a new genus of early angiosperm pollen, with notes on the Cretaceous palynostratigraphy and palaeoenvironments of northern Gondwana. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine 6, 39117.Google Scholar
Dumitrescu, M., Brassel, S. C., Schouten, S., Hopmans, E. C. & Sinninghe Damsté, J. S. 2006. Instability in tropical Pacific sea-surface temperatures during the early Aptian. Geology 34, 833–6.Google Scholar
Dupont, L. 2011. Orbital scale vegetation change in Africa. Quaternary Science Reviews 30, 3589–602.CrossRefGoogle Scholar
Embry, J.-C., Vennin, E., Van Buchem, F. S. P., Schroeder, R., Pierre, C. & Aurell, M. 2010. Sequence stratigraphy and carbon isotope stratigraphy of an Aptian mixed carbonate-siliciclastic platform to basin transition (Galve sub-basin, NE Spain). In Mesozoic and Cenozoic Carbonate Systems of the Mediterranean and Middle East: Stratigraphic and Diagenetic Reference Models (eds Van Buchem, F. S. P., Gerdes, K. D. & Esteban, M.), pp. 113–43. Geological Society of London, Special Publication no. 329.Google Scholar
Erba, E., Bottini, C., Weissert, H. J. & Keller, C. E. 2010. Calcareous nannoplankton response to surface-water acidification around Oceanic Anoxic Event 1a. Science 329, 428–32.CrossRefGoogle ScholarPubMed
Erdenetsogt, B., Lee, I., Bat-Erdene, D. & Jargal, L. 2009. Mongolian coal-bearing basins: geological settings, coal characteristics, distribution, and resources. International Journal of Coal Geology 80, 87104.CrossRefGoogle Scholar
Föllmi, K. B., Godet, A., Bodin, S. & Linder, P. 2006. Interactions between environmental change and shallow water carbonate buildup along the northern Tethyan margin and their impact on the Early Cretaceous carbon isotope record. Paleoceanography 21, 116.CrossRefGoogle Scholar
Francis, J. E. 1983. The dominant conifer of the Jurassic Purbeck Formation, England. Palaeontology 26, 277–94.Google Scholar
Friis, E. M., Pedersen, K. R. & Crane, P. R. 2010. Cretaceous diversification of angiosperms in the western part of the Iberian Peninsula. Review of Palaeobotany and Palynology 162, 341–61.Google Scholar
Gaona-Narvaez, T., Maurrasse, F. J. M. R. & Moreno-Bedmar, J. A. 2013. Stable carbon-isotope stratigraphy and ammonite biochronology at Madotz, Navarra, northern Spain: implications for the timing and duration of oxygen depletion during OAE-1a. Cretaceous Research 40, 143–57.Google Scholar
Gorin, G. E. & Steffen, D. 1991. Organic facies as a tool for recording eustatic variations in marine fine-grained carbonates - example of the Berriasian stratotype at Berrias (Ardèche, SE France). Palaeogeography, Palaeoclimatology, Palaeoecology 85, 303–20.CrossRefGoogle Scholar
Gradstein, F. M., Ogg, J. G., Schmitz, M. & Ogg, G. 2012. The Geologic Time Scale 2012. Oxford, Amsterdam: Elsevier, 1176 pp.Google Scholar
Hay, W. W., DeConto, R. M., Wold, C. N., Wilson, K. M., Voigt, S., Schulz, M., Wold, A. R., Dullo, W. C., Ronov, A. B., Balukhovsky, A. N. & Soeding, E. 1999. Alternative global Cretaceous paleogeography. In Evolution of the Cretaceous Ocean–Climate System (eds Barrera, E. & Johnson, C. C.), pp. 147. Geological Society of America, Special Paper no. 332.Google Scholar
Hay, W. W. & Floegel, S. 2012. New thoughts about the Cretaceous climate and oceans. Earth-Science Reviews 115, 262–72.Google Scholar
Heimhofer, U., Adatte, T., Hochuli, P. A., Burla, S. & Weissert, H. 2008. Coastal sediments from the Algarve: low-latitude climate archive for the Aptian-Albian. International Journal of Earth Sciences 97, 785–97.CrossRefGoogle Scholar
Heimhofer, U., Hochuli, P. A., Herrle, J. O., Andersen, N. & Weissert, H. 2004. Absence of major vegetation and pCO2 changes associated with Oceanic Anoxic Event 1a (Early Aptian, SE France). Earth and Planetary Science Letters 223, 303–18.CrossRefGoogle Scholar
Heimhofer, U., Hochuli, P. A., Herrle, J. O. & Weissert, H. 2006. Contrasting origins of Early Cretaceous black shales in the Vocontian basin: evidence from palynological and calcareous nannofossil records. Palaeogeography, Palaeoclimatology, Palaeoecology 235, 93109.Google Scholar
Heldt, M., Bachmann, M. & Lehmann, J. 2008. Microfacies, biostratigraphy, and geochemistry of the hemipelagic Barremian-Aptian in north-central Tunisia: influence of the OAE 1a on the southern Tethys margin. Palaeogeography, Palaeoclimatology, Palaeoecology 261, 246–60.Google Scholar
Herman, A. B. & Spicer, R. A. 2010. Mid-Cretaceous floras and climate of the Russian high Arctic (Novosibirsk Islands, Northern Yakutiya). Palaeogeography, Palaeoclimatology, Palaeoecology 295, 409–22.Google Scholar
Hermann, E., Hochuli, P. A., Bucher, H., Brühwiler, T., Hautmann, M., Ware, D. & Roohi, G. 2011. Terrestrial ecosystems on North Gondwana following the end-Permian mass extinction. Gondwana Research 20, 630–7.Google Scholar
Herrle, J. O., Kössler, P., Friedrich, O., Erlenkeuser, H. & Hemleben, C. 2004. High-resolution carbon isotope records of the Aptian to Lower Albian from SE France and the Mazagan Plateau (DSPD site 545): a stratigraphic tool for paleocanographic and paleobiologic reconstruction. Earth and Planetary Science Letters 218, 149–61.Google Scholar
Heusser, L. & Balsam, W. L. 1977. Pollen distribution in the Northeast Pacific Ocean. Quaternary Research 7, 4562.Google Scholar
Hochuli, P. A., Menegatti, A. P., Weissert, H., Riva, A., Erba, E. & Premoli Silva, I. 1999. Episodes of high productivity and cooling in the early Aptian Alpine Tethys. Geology 27, 657–60.2.3.CO;2>CrossRefGoogle Scholar
Hochuli, P. A., Os Vigran, J., Hermann, E. & Bucher, H. 2010. Multiple climatic changes around the Permian-Triassic boundary event revealed by an expanded palynological record from mid-Norway. Geological Society of America Bulletin 122, 884–96.Google Scholar
Hu, X. M., Zhao, K. D., Yilmaz, I. O. & Li, Y. X. 2012. Stratigraphic transition and palaeoenvironmental changes from the Aptian oceanic anoxic event 1a (OAE1a) to the oceanic red bed 1 (ORB1) in the Yenicesihlar section, central Turkey. Cretaceous Research 38, 4051.Google Scholar
Huber, B. T., Hodell, D. A. & Hamilton, C. P. 1995. Middle-Late Cretaceous climate of the southern high latitudes: stable isotopic evidence for minimal equator-to-pole thermal gradients. Geological Society of America Bulletin 107, 1164–91.2.3.CO;2>CrossRefGoogle Scholar
Huck, S., Heimhofer, U. & Immenhauser, A. 2012. Early Aptian algal bloom in a neritic proto-North Atlantic setting: Harbinger of global change related to OAE 1a? Geological Society of America Bulletin 124, 1810–25.CrossRefGoogle Scholar
Immenhauser, A., Hillgärtner, H. & Van Betum, E. 2005. Microbial-foraminiferal episodes in the Early Aptian of the southern Tethyan margin: ecological significance and possible relation to oceanic anoxic event 1a. Sedimentology 52, 7799.Google Scholar
Jahren, A. H., Arens, N. C., Sarmiento, G., Guerrero, J. & Amundson, R. 2001. Terrestrial record of methane hydrate dissociation in the Early Cretaceous. Geology 29, 159–62.Google Scholar
Jaramillo, C., Ochoa, D., Contreras, L., Pagani, M., Carvajal-Ortiz, H., Pratt, L. M., Krishnan, S., Cardona, A., Romero, M., Quiroz, L., Rodriguez, G., Rueda, M. J., de la Parra, F., Moron, S., Green, W., Bayona, G., Montes, C., Quintero, O., Ramirez, R., Mora, G., Schouten, S., Bermudez, H., Navarrete, R., Parra, F., Alvaran, M., Osorno, J., Crowley, J. L., Valencia, V. & Vervoort, J. 2010. Effects of rapid global warming at the Paleocene-Eocene Boundary on Neotropical Vegetation. Science 330, 957–61.Google Scholar
Jenkyns, H. C. 2010. Geochemistry of oceanic anoxic events. Geochemistry, Geophysics, Geosystems 11, doi: 10.1029/2009GC002788.Google Scholar
Jenkyns, H. C., Schouten-Huibers, L., Schouten, S. & Sinninghe Damsté, J. S. 2012. Warm Middle Jurassic-Early Cretaceous high-latitude sea-surface temperatures from the Southern Ocean. Climate of the Past 8, 215–26.Google Scholar
Keller, C. E., Hochuli, P. A., Weissert, H., Bernasconi, S. M., Giorgioni, M. & Garcia, T. I. 2011. A volcanically induced climate warming and floral change preceded the onset of OAE1a (Early Cretaceous). Palaeogeography, Palaeoclimatology, Palaeoecology 305, 43–9.Google Scholar
Kidder, D. L. & Worsley, T. R. 2010. Phanerozoic Large Igneous Provinces (LIPs), HEATT (Haline Euxinic Acidic Thermal Transgression) episodes, and mass extinctions. Palaeogeography, Palaeoclimatology, Palaeoecology 295, 162–91.Google Scholar
Kirillova, G. L. 2003. Late Mesozoic – Cenozoic sedimentary basins of active continental margin of Southeast Russia: paleogeography, tectonics, and coal–oil–gas presence. Marine & Petroleum Geology 20, 385–97.Google Scholar
Kübler, B. 1987. Cristallinite de l’illite, methods normalisees de preparations, methods normalisees de measures. Cahiers de l’Institut de Géologie de Neuchâtel 1.1, 113.Google Scholar
Kuhnt, W., Holbourn, A. & Moullade, M. 2011. Transient global cooling at the onset of Early Aptian oceanic anoxic event (OAE) 1a. Geology 39, 323–6.Google Scholar
Kujau, A., Heimhofer, U., Ostertag-Henning, C., Greselle, B. & Mutterlose, J. 2012. No evidence for anoxia during the Valanginian carbon isotope event: An organic-geochemical study from the Vocontian Basin, SE France. Global and Planetary Change 92–93, 92104.CrossRefGoogle Scholar
Larson, R. L. & Erba, E. 1999. Onset of the mid-Cretaceous greenhouse in the Barremian–Aptian igneous events and the biological, sedimentary and geochemical responses. Paleoceanography 14, 663–78.Google Scholar
Li, Y.-X., Bralower, T. J., Montanez, I. P., Osleger, D. A., Arthur, M. A., Bice, D. M., Herbert, T. D., Erba, E. & Silva, I. P. 2008. Toward an orbital chronology for the early Aptian Oceanic Anoxic Event (OAE1a, ~120 Ma). Earth and Planetary Science Letters 271, 88100.CrossRefGoogle Scholar
Lirong, D., Dingsheng, C., Zhi, L., Zhiwei, Z. & Jingchun, W. 2013. Petroleum geology of the fula sub-basin, Muglad basin, Sudan. Journal of Petroleum Geology 36, 4359.Google Scholar
Looy, C. V., Brugman, W. A., Dilcher, D. L. & Visscher, H. 1999. The delayed resurgence of equatorial forests after the Permian-Triassic ecologic crisis. Proceedings of the National Academy of Sciences of the United States of America 96, 13857–62.Google Scholar
Malinverno, A., Erba, E. & Herbert, T. D. 2010. Orbital tuning as an inverse problem: chronology of the early Aptian oceanic anoxic event 1a (Selli Level) in the Cismon APTICORE. Paleoceanography 25, 116.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 (eds McCabe, P.J. & Parrish, J.T.), pp. 115. Geological Society of America, Special Paper no. 267.Google Scholar
McElwain, J. C., Beerling, D. J. & Woodward, F. I. 1999. Fossil plants and global warming at the Triassic–Jurassic Boundary. Science 285, 1386–90.CrossRefGoogle ScholarPubMed
Méhay, S., Keller, C. E., Bernasconi, S. M., Weissert, H., Erba, E., Bottini, C. & Hochuli, P. A. 2009. A volcanic CO2 pulse triggered the Cretaceous Oceanic Anoxic Event 1a and a biocalcification crisis. Geology 37, 819–22.Google Scholar
Mello, M. R. & Maxwell, J. R. 1990. Organic geochemical and biological marker characterization of source rocks and oils derived from lacustrine environments in the Brazilian continental margin. In Lacustrine Basin Exploration - Case Studies and Modern Analogs (eds Katz, B. J.), pp. 7797. American Association of Petroleum Geologists, Memoir no. 50.Google Scholar
Menegatti, A. P., Weissert, H., Brown, R. S., Tyson, R. V., Farrimond, P., Strasser, A. & Caron, M. 1998. High-resolution δ13C stratigraphy through the early Aptian “Livello Selli” of the Alpine Tethys. Paleoceanography 13, 530–45.Google Scholar
Millán, M. I., Weissert, H. J., Fernandez-Mendiola, P. A. & Garcia-Mondejar, J. 2009. Impact of Early Aptian carbon cycle perturbations on evolution of a marine shelf system in the Basque-Cantabrian Basin (Aralar, N Spain). Earth and Planetary Science Letters 287, 392401.Google Scholar
Moore, D. & Reynolds, R. 1997. X-Ray-Diffraction and the Identification and Analysis of Clay-Minerals. Oxford: Oxford University Press, 378 pp.Google Scholar
Moreno-Bedmar, J. A., Company, M., Bover-Arnal, T., Salas, R. & Delanoy, G. 2009. Biostratigraphic characterization by means of ammonoids of the lower Aptian Oceanic Anoxic Event (OAE 1a) in the eastern Iberian Chain (Maestrat Basin, eastern Spain). Cretaceous Research 30, 864–72.Google Scholar
Moreno-Bedmar, J. A., Company, M., Bover-Arnal, T., Salas, R., Delanoy, G., Maurrasse, F. J. M. R., Grauges, A. & Martinez, R. 2010. Lower Aptian ammonite biostratigraphy in the Maestrat Basin (Eastern Iberian Chain, Eastern Spain). A Tethyan transgressive record enhanced by synrift subsidence. Geologica Acta 8, 281–99.Google Scholar
Moreno-Bedmar, J. A., Company, M., Sandoval, J., Tavera, J. M., Bover-Arnal, T., Salas, R., Delanoy, G., Maurrasse, F. J. M. R. & Martinez, R. 2012. Lower Aptian ammonite and carbon isotope stratigraphy in the eastern Prebetic Domain (Betic Cordillera, southeastern Spain). Geologica Acta 10, 333–50.Google Scholar
Moss, P. T., Kershaw, A. P. & Grindrod, J. 2005. Pollen transport and deposition in riverine and marine environments within the humid tropics of northeastern Australia. Review of Palaeobotany and Palynology 134, 5569.Google Scholar
Moullade, M., Kuhnt, W., Bergen, J. A., Masse, J. P. & Tronchetti, G. 1998. Correlation of biostratigraphic and stable isotope events in the Aptian historical stratotype of La Bedoule (southeast France). Comptes Rendus de l’Académie des Sciences - Serie II A - Earth and Planetary Science 327, 693–8.Google Scholar
Najarro, M., Rosales, I., Moreno-Bedmar, J. A., de Gea, G. A., Barron, E., Company, M. & Delanoy, G. 2011. High-resolution chemo- and biostratigraphic records of the Early Aptian oceanic anoxic event in Cantabria (N Spain): Palaeoceanographic and palaeoclimatic implications. Palaeogeography, Palaeoclimatology, Palaeoecology 299, 137–58.Google Scholar
Ntamak-Nida, M.-J., Baudin, F., Schnyder, J., Makong, J.-C., Komguem, P. B. & Abolo, G. M. 2008. Depositional environments and characterisation of the organic matter of the Lower Mundeck Formation (Barremian ?-Aptian) of the Kribi-Campo sub-basin (South Cameroon): Implications for petroleum exploration. Journal of African Earth Sciences 51, 207–19.Google Scholar
Pancost, R. D., Crawford, N., Magness, S., Turner, A., Jenkyns, H. C. & Maxwell, J. R. 2004. Further evidence for the development of photic-zone euxinic conditions during Mesozoic oceanic anoxic events. Journal of the Geological Society, London 161, 353–64.Google Scholar
Pellaton, C. & Gorin, G. E. 2005. The Miocene New Jersey passive margin as a model for the distribution of sedimentary organic matter in siliciclastic deposits. Journal of Sedimentary Research 75, 1011–27.Google Scholar
Price, G. D. 1999. The evidence and implications of polar ice during the Mesozoic. Earth-Science Reviews 48, 183210.Google Scholar
Price, G. D. 2003. New constraints upon isotope variation during the early Cretaceous (Barremian–Cenomanian) from the Pacific Ocean. Geological Magazine 140, 513–22.Google Scholar
Quijano, L. M., Manuel Castro, J., Pancost, R. D., de Gea, G. A., Najarro, M., Aguado, R., Rosales, I. & Martin-Chivelet, J. 2012. Organic geochemistry, stable isotopes, and facies analysis of the Early Aptian OAE - New records from Spain (Western Tethys). Palaeogeography, Palaeoclimatology, Palaeoecology 365, 276–93.Google Scholar
Rameil, N., Götz, A. E. & Feist-Burkhardt, S. 2000. High-resolution sequence interpretation of epeiric shelf carbonates by means of palynofacies analysis: An example from the Germanic Triassic (Lower Muschelkalk, Anisian) of East Thuringia, Germany. Facies 43, 123–43.Google Scholar
Riding, J. B., Leng, M. J., Kender, S., Hesselbo, S. P. & Feist-Burkhardt, S. 2013. Isotopic and palynological evidence for a new Early Jurassic environmental perturbation. Palaeogeography, Palaeoclimatology, Palaeoecology 374, 1627.CrossRefGoogle Scholar
Robinson, S. A., Clarke, L. J., Nederbragt, A. & Wood, I. G. 2008. Mid-Cretaceous oceanic anoxic events in the Pacific Ocean revealed by carbon-isotope stratigraphy of the Calera Limestone, California, USA. Geological Society of America Bulletin 120, 1416–27.CrossRefGoogle Scholar
Ruffell, A. H. & Batten, D. J. 1990. The Barremian-Aptian arid phase in western Europe. Palaeogeography, Palaeoclimatology, Palaeoecology 80, 197212.Google Scholar
Salas, R., Guimera, J., Mas, R., Martin-Closas, C., Melendez, A. & Alonso, A. 2001. Evolution of the Mesozoic Central Iberian Rift System and its Cenozoic inversion (Iberian chain). In Peri-Tethys Memoir 6: Peri-Tethyan Rift/Wrench Basins and Passive Margins (eds Ziegler, P. A., Cavazza, W., Robertson, A. H. F. & CrasquinSoleau, S.), pp. 145–86. Memoires du Museum National d’Histoire Naturelle, Paris, 186.Google Scholar
Salas, R., Martín-Closas, C., Delclòs, X., Guimerà, J., Caja, M. A. & Mas, R. 2005. Factores principales de control de la sedimentación y los cambios bióticos durante el tránsito Jurásico-Cretácico en la Cadena Ibérica. Geogaceta 38, 15–8.Google Scholar
Schiøler, P., Crampton, J. S. & Laird, M. G. 2002. Palynofacies and sea-level changes in the Middle Coniacian-Late Campanian (Late Cretaceous) of the East Coast Basin, New Zealand. Palaeogeography, Palaeoclimatology, Palaeoecology 188, 101–25.Google Scholar
Schouten, S., Hopmans, E. C., Forster, A., van Breugel, Y., Kuypers, M. M. M. & Sinninghe Damsté, J. S. 2003. Extremely high sea-surface temperatures at low latitudes during the middle Cretaceous as revealed by archaeal membrane lipids. Geology 31, 1069–72.Google Scholar
Schrank, E. 2010. Pollen and spores from the Tendaguru Beds, Upper Jurassic and Lower Cretaceous of southeast Tanzania: palynostratigraphical and paleoecological implications. Palynology 34, 342.Google Scholar
Sewall, J. O., van de Wal, R. S. W., van der Zwan, K., van Oosterhout, C., Dijkstra, H. A. & Scotese, C. R. 2007. Climate model boundary conditions for four Cretaceous time slices. Climate of the Past 3, 647–57.Google Scholar
Sha, J., Hiromichi, H., Yao, X. & Pan, Y. 2008. Late Mesozoic transgressions of eastern Heilongjiang and their significance in tectonics, and coal and oil accumulation in northeast China. Palaeogeography, Palaeoclimatology, Palaeoecology 263, 119–30.Google Scholar
Sinninghe Damste, J. S., van Bentum, E. C., Reichart, G. J., Pross, J. & Schouten, S. 2010. A CO2 decrease-driven cooling and increased latitudinal temperature gradient during the mid-Cretaceous Oceanic Anoxic Event 2. Earth and Planetary Science Letters 293, 97103.Google Scholar
Skelton, P. W., Spicer, R. A., Kelley, S. P. & Gilmour, I. 2003. The Cretaceous World. Cambridge: Cambridge University Press, 360 pp.Google Scholar
Solé de Porta, N., Querol, X., Cabanes, R. & Salas, R. 1994. Nuevas aportaciones a la palinología y paleoclimatología de la Formación Esucha (Albiense inferior-medio) en las Cubetas de Utrillas y Oliete, Cordillera Ibérica Oriental. Cuadernos Geología Ibérica 18, 203–15.Google Scholar
Staplin, F. L. 1982. Determination of thermal alteration index from color of exinite (pollen, spores). In How to Assess Maturation and Paleotemperatures (ed. Staplin, F. L.), pp. 711. Society of Economic Paleontologists and Mineralogists, Short Course no. 7.Google Scholar
Steffen, D. & Gorin, G. 1993. Palynofacies of the Upper Tithonian – Berriasian deep-sea carbonates in the Vocontian Trough (SE France). Bulletin Des Centres De Recherches Exploration-Production Elf Aquitaine 17, 235–47.Google Scholar
Traverse, A. 2007. Paleopalynology. Dordrecht: Springer, 813 pp.Google Scholar
Tyson, R. V. 1995. Sedimentary Organic Matter. London: Chapman & Hall, 615 pp.Google Scholar
Vahrenkamp, V. C. 2010. Chemostratigraphy of the Lower Cretaceous Shu’aiba Formation: A d13C reference profile for the Aptian Stage from the Southern Neo-Tehtys Ocean. In Barremian-Aptian Stratigraphy and Hydrocarbon Habitat of the Eastern Arabian Plate (eds Van Buchem, F. S. P., Al-Husseini, M. I., Maurer, F. & Droste, H.), pp. 107–37. GeoArabia Special Publication no. 4.Google Scholar
Vakhrameyev, V. A. 1982. Classopollis pollen as an indicator of Jurassic and Cretaceous climate. International Geology Review 24, 1190–6.Google Scholar
Vakhrameyev, V. A. 1991. Jurassic and Cretaceous Floras and Climates of the Earth. Cambridge: Cambridge University Press, 318 pp.Google Scholar
van de Schootbrugge, B., Quan, T. M., Lindstrom, S., Puettmann, W., Heunisch, C., Pross, J., Fiebig, J., Petschick, R., Roehling, H. G., Richoz, S., Rosenthal, Y. & Falkowski, P. G. 2009. Floral changes across the Triassic/Jurassic boundary linked to flood basalt volcanism. Nature Geoscience 2, 589–94.Google Scholar
van der Kaars, S. 2001. Pollen distribution in marine sediments from the south-eastern Indonesian waters. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 341–61.Google Scholar
van Konijnenburg – van Cittert, J. 2002. Ecology of some Late Triassic to Early Cretaceous ferns in Eurasia. Review of Palaeobotany and Palynology 119, 113–24.Google Scholar
Vennin, E. & Aurell, M. 2001. Aptian paleoenvironmental evolution and sequence stratigraphy in the Galve sub-basin (Teruel, NE Spain). Bulletin de la Societe Geologique de France 172, 397410.Google Scholar
Weissert, H. & Erba, E. 2004. Volcanism, CO2 and palaeoclimate: a Late Jurassic–Early Cretaceous carbon and oxygen isotope record. Journal of the Geological Society, London 161, 695702.Google Scholar
Westermann, S., Föllmi, K. B., Adatte, T., Matera, V., Schnyder, J., Fleitmann, D., Fiet, N., Ploch, I. & Duchamp-Alphonse, S. 2010. The Valanginian δ13C excursion may not be an expression of a global anoxic event. Earth and Planetary Science Letters 290, 118–31.Google Scholar
Wing, S. L., Harrington, G. J., Smith, F. A., Bloch, J. I., Boyer, D. M. & Freeman, K. H. 2005. Transient floral change and rapid global warming at the Paleocene–Eocene boundary. Science 310, 993–6.Google Scholar
Wissler, L., Funk, H. & Weissert, H. 2003. Response of Early Cretaceous carbonate platforms to changes in atmospheric carbon dioxide levels. Palaeogeography, Palaeoclimatology, Palaeoecology 200, 187205.Google Scholar
Yamamoto, K., Ishibashi, M., Takayanagi, H., Asahara, Y., Sato, T., Nishi, H. & Iryu, Y. 2013. Early Aptian paleoenvironmental evolution of the Bab Basin at the southern Neo-Tethys margin: response to global carbon-cycle perturbations across Ocean Anoxic Event 1a. Geochemistry, Geophysics, Geosystems 14, doi: 10.1002/ggge.20083.Google Scholar
Ziegler, A. M. 1990. Geological Atlas of Western and Central Europe. London: Geological Society Publishing House, 239 pp.Google Scholar
Supplementary material: File

Cors Supplementary Material

Table S1

Download Cors Supplementary Material(File)
File 22.1 KB
Supplementary material: File

Cors Supplementary Material

Table S2

Download Cors Supplementary Material(File)
File 24.1 KB