Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-24T22:58:05.266Z Has data issue: false hasContentIssue false

The expression of the Hangenberg Event (latest Devonian) in a relatively shallow-marine succession (Pomeranian Basin, Poland): the results of a multi-proxy investigation

Published online by Cambridge University Press:  20 August 2014

HANNA MATYJA*
Affiliation:
Polish Geological Institute – National Research Institute, Rakowiecka Str. 4, 00-975 Warszawa, Poland
KATARZYNA SOBIEŃ
Affiliation:
Polish Geological Institute – National Research Institute, Rakowiecka Str. 4, 00-975 Warszawa, Poland
LESZEK MARYNOWSKI
Affiliation:
Faculty of Earth Sciences, University of Silesia, Będzińska Str. 60, 41-200 Sosnowiec, Poland
MARZENA STEMPIEŃ-SAŁEK
Affiliation:
Faculty of Oceanography and Geography, University of Gdańsk, Marszałka Piłsudzkiego Str. 46, 81-378 Gdynia, Poland
KRZYSZTOF MAŁKOWSKI
Affiliation:
Institute of Palaeobiology, Polish Academy of Sciences, Twarda Str. 51/55, 00-818 Warszawa, Poland
*
Author for correspondence: hanna.matyja@pgi.gov.pl

Abstract

The uppermost Famennian – lowermost Tournaisian interval has been analysed in detail using biostratigraphy, sedimentology, magnetic susceptibility and geochemistry in a reference section of the relatively shallow carbonate ramp environment within the Pomeranian Basin. High-resolution biostratigraphic study, based on miospores, allows recognition of the standard western European lepidophyta–nitidus (LN) and verrucosus–incohatus (VI) zones, as well as the Convolutispora major Zone, a local Pomeranian equivalent of the European standard hibernicus–distinctus (HD) Zone. The sedimentary succession and specific phenomena recognized close to the Devonian/Carboniferous boundary, such as fluctuations in water column euxinia, wildfire evidence, relative sea-level changes and perturbations of the carbon cycle reflected by positive carbon excursions, display a pattern partly similar to that observed in many areas in Europe during the Hangenberg Event, although the Hangenberg Black Shale horizon is not developed here. These important microscale environmental perturbations were observed not only within the Famennian LN miospore Zone but in a wide interval between the LN and the lowermost local Convolutispora major miospore zones ( = lower part of HD standard miospore Zone). It is still uncertain whether the recognized event(s) were connected solely with the Hangenberg Event, which was possibly complex and multi-phased as is sometimes suggested, or whether they represent a succession of regionally limited, post-Hangenberg events. This question needs to be further investigated on broader stratigraphic and geographical scales.

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

Avchimovitch, V. I., Byvsheva, T. V., Higgs, K., Streel, M. & Umnova, V. T. 1988. Miospore systematics and stratigraphic correlation of Devonian-Carboniferous boundary deposits in the European part of the USSR and Western Europe. Courier Forschungsinstitut Senckenberg 100, 169–91.Google Scholar
Avchimovitch, V. I., Turnau, E. & Clayton, G. 1993. Correlation of uppermost Devonian and Lower Carboniferous miospore zonation in Byelorussia, Poland and western Europe. Annales de la Société Géologique de Belgique 115, 453–8.Google Scholar
Amler, M. R. 1993. Shallow marine bivalves at the Devonian/Carboniferous boundary from the Velbert Anticline (Rheinisches Schiefergebirge). Annales de la Société Géologique de Belgique 115, 4053–458.Google Scholar
Azmy, K., Poty, E. & Brand, U. 2009. High-resolution isotope stratigraphy of the Devonian-Carboniferous boundary in the Namur-Dinant Basin, Belgium. Sedimentary Geology 216, 117–24.CrossRefGoogle Scholar
Barnes, C. R., Zhang, Y., Jeppson, L., Fredholm, D., Varker, W. J., Swift, A., Merrill, G. K. & Dorning, K. J. 1987. Recent developments in rock disintegration techniques for the extraction of conodonts. In Conodonts: Investigative Techniques and Applications (ed. Austin, R. L.), pp. 3553. The British Micropalaeontological Society Series. Chichester: Ellis Horwood.Google Scholar
Becker, R. T. 1993. Analyses of ammonoid palaeogeography in relation to the global Hangenberg (terminal Devonian) and Lower Alum Shale (Middle Tournaisian) Events. Annales de la Société Géologique de Belgique 115, 459–73.Google Scholar
Bertrand, S., Charlet, F., Charlier, B., Renson, V. & Fagel, N. 2008. Climate variability of southern Chile since the last Glacial Maximum: a continuous sedimentological record from Lago Puyehue (40°S). Journal of Paleolimnology 39, 179–95.CrossRefGoogle Scholar
Bharadwaj, D. C. & Venkatachala, B. S. 1961. Spore assemblage out of a Lower Carboniferous shale from Spitsbergen. The Palaeobotanist 10, 1847.Google Scholar
Bless, M. J. M. 1993. Comparison between eustatic T-R cycles around the Devonian-Carboniferous boundary and the distribution of the ostracode of the taxon Pseudoleperditia gr. venulosa . Annales de la Société Géologique de Belgique 115, 475–81.Google Scholar
Bless, M. J. M., Becker, R. T., Higgs, K., Paproth, E. & Streel, M. 1993. Eustatic cycles around the Devonian-Carboniferous boundary and the sedimentary and fossil record in Sauerland (Federal Republic of Germany). Annales de la Société Géologique de Belgique 115, 689702.Google Scholar
Blumenstengel, H. 1993. Ostracods from the Devonian-Carboniferous boundary beds in Thuringia (Germany). Annales de la Société Géologique de Belgique 115, 483–9.Google Scholar
Bond, D., Wignall, P. B. & Racki, G. 2004. Extent and duration of marine anoxia during the Frasnian-Famennian (Late Devonian) mass extinction in Poland, Germany, Austria and France. Geological Magazine 141, 173–93.CrossRefGoogle Scholar
Bond, D. P. G., Zatoń, M., Wignall, P. G. & Marynowski, L. 2013. Evidence for shallow-water anoxic “Kellwasser Events” in the Frasnian-Famennian reefs of Alberta, Canada. Lethaia 46, 355–68.CrossRefGoogle Scholar
Brand, U., Legrand-Blain, M. & Streel, M. 2004. Biochemostratigraphy of the Devonian-Carboniferous boundary global stratotype section and point, Griotte Formation, La Serre, Montagne Noire, France. Palaeogeography, Palaeoclimatology, Palaeoecology 205, 337–57.CrossRefGoogle Scholar
Brezinski, D. K., Cecil, C. B. & Skema, V. W. 2010. Late Devonian glacigenic and associated facies from the central Appalachian Basin, eastern United States. Geological Society of America Bulletin 122, 265–81.CrossRefGoogle Scholar
Brezinski, D. K., Cecil, C. B., Skema, V. W. & Stamm, R. 2008. Late Devonian glacial deposits from the eastern United States signal an end of the mid-Paleozoic warm period. Palaeogeography, Palaeoclimatology, Palaeoecology 268, 143–51.CrossRefGoogle Scholar
Buggisch, W. & Joachimski, M. M. 2006. Carbon isotope stratigraphy of the Devonian of Central and Southern Europe. Palaeogeography, Palaeoclimatology, Palaeoecology 240, 6888.CrossRefGoogle Scholar
Byvsheva, T. V. 1985. Spory iz otlozhenij turnejskogo i vizenskogo jarusov Russkoj plity. In Atlas Spor i Pylcy Neftegazonosnykh Tolshtch Fanerozoja Russkoj i Turanskoj Plit (eds Menner, V. V. & Byvsheva, T. G.), pp. 80158. Trudy VNIGNI 253 (in Russian).Google Scholar
Calvert, S. E. & Pedersen, T. F. 1993. Geochemistry of recent oxic and anoxic marine sediments: implications for the geological record. Marine Geology 113, 6788.CrossRefGoogle Scholar
Caplan, M. L. & Bustin, M. R. 1999. Devonian-Carboniferous Hangenberg mass extinction event, widespread organic-rich mudrock and anoxia: causes and consequences. Palaeogeography, Palaeoclimatology, Palaeoecology 148, 197207.CrossRefGoogle Scholar
Caputo, M. V., Melo, J. H. G., Streel, M. & Isbell, J. L. 2008. Late Devonian and Early Carboniferous glacial records of South America. In Resolving the Late Paleozoic Ice in Time and Space (eds Fielding, C. R., Frank, T. D. & Isbell, J. L.), pp. 113. The Geological Society of America, Special Paper no. 441.Google Scholar
Clark, S., Day, J., Ellwood, B., Harry, R. & Tomkin, J. 2009. Astronomical tuning of integrated upper Famennian-early Carboniferous faunal, carbon isotope and high resolution magnetic susceptibility records: western Illinois basin. Subcommission on Devonian Stratigraphy, Newsletter 24, 2735.Google Scholar
Clayton, G. 1971. A Lower Carboniferous miospore assemblage from the Calciferous Sandstone Measures of the Cockburnspath region of eastern Scotland. Pollen et Spores 12, 577600.Google Scholar
Clayton, G., Coquel, R., Doubinger, J., Gueinn, K. J., Loboziak, S., Owens, B. & Streel, M. 1977. Carboniferous spores of Western Europe: illustration and zonation. Mededelingen, Rijks Geologische Dienst 29, 170.Google Scholar
Dadlez, R. 1997. Epicontinental basins in Poland: Devonian to Cretaceous – relationships between the crystalline basement and sedimentary infill. Geological Quarterly 41, 419–32.Google Scholar
Day, J., Witzke, B. J., Rowe, H. & Elwood, B. 2013. Magnetic susceptibility and carbon isotope stratigraphies through the Devonian-Carboniferous boundary interval in the western Illinois basin – central North America. In Geophysical and Geochemical Techniques: A Window on the Palaeozoic World, IGCP 580–596, 27 August – 1 September 2013, Programme with Abstracts (eds Whalen, M., Osadetz, K., Richards, B., Kabanov, P., Weissenberger, J., Potma, K., Koenigshof, P., Suttner, T., Kido, E. & Silva, A.-C. Da), pp. 26–7.Google Scholar
Dean, W. E. & Arthur, M. A. 1989. Iron-sulfur-carbon relationships in organic-carbon-rich sequences: I. Cretaceous Western Interior Seaway. American Journal of Science 289, 708–43.CrossRefGoogle Scholar
De Vleeschouwer, D., Rakociński, M., Racki, G., Bond, D. P. G., Sobień, K. & Claeys, P. 2013. The astronomical rhythm of Late-Devonian climate change (Kowala section, Holy Cross Mountains, Poland). Earth and Planetary Science Letters 365, 2537.CrossRefGoogle Scholar
Didyk, B. M., Simoneit, B. R. T., Brassel, S. C. & Eglinton, G. 1978. Organic geochemical indicators of palaeoenvironmental conditions of sedimentation. Nature 272, 216–22.CrossRefGoogle Scholar
Dolby, G. & Neves, R. 1970. Palynological evidence concerning the Devonian-Carboniferous boundary in the Mendips, England. 6-eme Congrés International de Stratigraphie et de Géologie du Carbonifère (Sheffield, 1967) . Comptes Rendus 2, 631–46.Google Scholar
Filipiak, P. 2004. Miospore stratigraphy of Upper Famennian and Lower Carboniferous deposits of the Holy Cross Mountains (central Poland). Review of Palaeobotany and Palynology 128, 291322.CrossRefGoogle Scholar
Filipiak, P. 2005. Late Devonian and Early Carboniferous acritarchs and prasinophytes from the Holy Cross Mountains (central Poland). Review of Palaeobotany and Palynology 134, 126.CrossRefGoogle Scholar
Filipiak, P. & Racki, G. 2010. Proliferation of abnormal palynoflora during the end-Devonian biotic crisis. Geological Quarterly 54, 114.Google Scholar
Finklestein, D. B., Pratt, L. M., Curtin, T. M. & Brassell, S. C. 2005. Wildfires and seasonal aridity recorded in Late Cretaceous strata from south-eastern Arizona, USA. Sedimentology 52, 587–99.CrossRefGoogle Scholar
Flajs, G. & Feist, R. 1988. Index conodonts, trilobites and environment of the Devonian-Carboniferous boundary beds at La Serre (Montagne Noire, France). Courier Forschungsinstitut Senckenberg 100, 53107.Google Scholar
Gross-Uffenorde, H., Lethiers, F. & Blumenstengel, H. 2000. Ostracodes and Devonian stratigraphy. Courier Forschungsinstitut Senckenberg 220, 99111.Google Scholar
Gross-Uffenorde, H. & Schindler, E. 1990. The effect of global events on entomozoacean Ostracoda. In Ostracoda and Global Events (eds Whatley, R. & Maybury, C.), pp. 101–12 (with Addendum to Fig. 4 by Gross-Uffenorde H., 1990. Present state of knowledge of the correlation of entomozoacean and conodont zonation within the Devonian). British Micropalaeontological Society Series. London: Chapman & Hall.CrossRefGoogle Scholar
Grotek, I., Matyja, H. & Skompski, S. 1998. Thermal maturity of organic matter in the Carboniferous deposits of the Radom-Lublin and Pomerania areas. In Sedimentary Basin Analysis of the Polish Lowland (ed. Narkiewicz, M.), pp. 245–53. Prace Państwowego Instytutu Geologicznego 165 (in Polish with English summary).Google Scholar
Hacquebard, P. A. 1957. Plant spores in coal from Horton Group (Mississippian) of Nova Scotia. Micropalaeontology 3, 301–24.CrossRefGoogle Scholar
Hartkopf-Fröder, C., Kloppisch, M., Mann, U., Neumann-Mahlkau, P., Schaefer, R. G. & Wilkes, H. 2007. The end-Frasnian mass extinction in the Eifel Mountains, Germany: new insights from organic matter composition and preservation. In Devonian Events and Correlations (eds Becker, R. T. & Kirchgasser, W. T.), pp. 173–96. Geological Society of London, Special Publication no. 278.Google Scholar
Heider, F., Bock, J. M., Hendy, J., Kennett, J. P., Matzka, J. & Schneider, J. 2001. Latest Quaternary rock magnetic record of climatic and ocean change, Tanner Basin, California borderland. Geological Society of America Bulletin 113, 346–59.2.0.CO;2>CrossRefGoogle Scholar
Higgs, K. T. 1975. Upper Devonian and Lower Carboniferous miospore assemblages from Hook Head, County Wexford, Ireland. Micropaleontology 21, 393419.CrossRefGoogle Scholar
Higgs, K. T., Clayton, G. & Keegan, J. B. 1988. Stratigraphic and systematic palynology of the Tournaisian rocks of Ireland. The Geological Survey of Ireland, Special Paper 1, 193.Google Scholar
Higgs, K. T., Streel, M., Korn, D. & Paproth, E. 1993. Palynological data from the Devonian-Carboniferous boundary beds in the New Stockum Trench II and the Hasselbachtal Borehole. Northern Rhenish Massif, Germany. Annales de la Société Géologique de Belgique 115, 551–7.Google Scholar
Hoffmeister, W. S., Staplin, F. L. & Malloy, R. E. 1955. Mississippian plant spores from the Hardinsburg Formation of Illinois and Kentucky. Journal of Paleontology 29, 372–99.Google Scholar
House, M. R. 1985. Correlations of mid-Palaeozoic ammonoid evolutionary events with global sedimentary perturbations. Nature 313, 1722.CrossRefGoogle Scholar
House, M. R. 2002. Strength, timing, setting and cause of mid-Palaeozoic extinctions. Palaeogeography, Palaeoclimatology, Palaeoecology 181, 525.CrossRefGoogle Scholar
Hrouda, F. & Kahan, Š. 1991. The magnetic fabric relationship between sedimentary and basement nappes in the High Tatra Mts. (N Slovakia). Journal of Structural Geology 13, 431–42.CrossRefGoogle Scholar
Isaacson, P. E., Diaz-Martinez, E., Grader, G. W., Kalvoda, J., Babek, O. & Devuyst, F. X. 2008. Late Devonian-earliest Mississippian glaciation in Gondwanaland and its biogeographic consequences. Palaeogeography, Palaeoclimatology, Palaeoecology 268, 126–42.CrossRefGoogle Scholar
Isaacson, P. E., Hladil, J., Shen, J.-W., Kalvoda, J. & Grader, G. 1999. Late Devonian (Famennian) Glaciation in South America and marine offlap on other continents. In North Gondwana: Mid-Paleozoic Terranes, Stratigraphy and Biota (eds Feist, R., Talent, J. A. & Daurer, A.), pp. 239–57. Abhandlungen der Geologischen Bundesanstalt 54.Google Scholar
Joachimski, M. M., Van Geldern, R., Breisig, S., Buggisch, W. & Day, J. 2004. Oxygen isotope evolution of biogenic calcite and apatite during the Middle and Late Devonian. International Journal of Earth Sciences 93, 542–53.CrossRefGoogle Scholar
Johnson, J. G., Klapper, G. & Sandberg, C. A. 1985. Devonian eustatic fluctuations in Euramerica. Geological Society of America Bulletin 96, 567–87.2.0.CO;2>CrossRefGoogle Scholar
Johnson, J. G., Klapper, G. & Sandberg, C. A. 1986. Late Devonian eustatic cycles around the margin of Old Red Continent. Annales de la Société Géologique de Belgique 109, 141–7.Google Scholar
Jones, B. & Manning, D. A. C. 1994. Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstone. Chemical Geology 111, 111–29.CrossRefGoogle Scholar
Kaiser, S. I., Becker, R. T., Spaletta, C. & Steuber, T. 2009. High-resolution conodont stratigraphy, biofacies, and extinctions around the Hangenberg Event in pelagic successions from Austria, Italy, and France. Palaeontographica Americana 63, 99143.Google Scholar
Kaiser, S. I., Becker, R. T., Steuber, T. & Aboussalam, S. Z. 2011. Climate-controlled mass extinctions, facies, and sea-level changes around the Devonian–Carboniferous boundary in the eastern Anti-Atlas (SE Morocco). Palaeogeography, Palaeoclimatology, Palaeoecology 310, 340–64.CrossRefGoogle Scholar
Kaiser, S. I., Steuber, T. & Becker, R. T. 2008. Environmental change during the Late Famennian and Early Tournaisian (Late Devonian-Early Carboniferous): implications from stable isotopes and conodont biofacies in southern Europe. Geological Journal 43, 241–60.CrossRefGoogle Scholar
Kaiser, S. I., Steuber, T., Becker, R. T. & Joachimski, M. M. 2006. Geochemical evidence for major environmental change at the Devonian-Carboniferous boundary in the Carnic Alps and the Rhenish Massif. Palaeogeography, Palaeoclimatology, Palaeoecology 240, 146–60.CrossRefGoogle Scholar
Kalvoda, J. & Kukal, Z. 1987. Devonian-Carboniferous boundary in the Moravian Karst at Lesni Lom Quarry, Brno-Lisen, Czechoslovakia. Courier Forschungsinstitut Senckenberg 98, 95117.Google Scholar
Kenig, F., Hudson, J. D., Sinninghe Damsté, J. S. & Popp, B. N. 2004. Intermittent euxinia: reconciliation of a Jurassic black shale with its biofacies. Geology 32, 421–4.CrossRefGoogle Scholar
Koopmans, M. P., Köster, J., Van Kaam-Peters, H. M. E., Kenig, F., Schouten, S., Hartgers, W. A., De Leeuw, J. W. & Sinninghe Damsté, J. S. 1996. Diagenetic and catagenetic products of isorenieratene: molecular indicators for photic zone anoxia. Geochimica et Cosmochimica Acta 60, 4467–96.CrossRefGoogle Scholar
Korn, D., Belka, Z., Fröhlich, S., Rücklin, M. & Wendt, J. 2004. The youngest African clymeniids (Ammonoidea, Late Devonian) – failed survivors of the Hangenberg Event. Lethaia 37, 307–15.CrossRefGoogle Scholar
Lane, H. R., Sandberg, C. A. & Ziegler, W. 1980. Taxonomy and phylogeny of some Lower Carboniferous conodonts and preliminary standard post-Siphonodella zonation. Geologica et Palaeontologica 14, 117–64.Google Scholar
Latta, D. K., Anastasio, D. J., Hinnov, L. A., Elrick, M. & Kodama, K. P. 2006. Magnetic record of Milankovitch rhythms in lithologically noncyclic marine carbonates. Geology 34, 2931.CrossRefGoogle Scholar
Marynowski, L. & Filipiak, P. 2007. Water column euxinia and wildfire evidence during deposition of the Upper Famennian Hangenberg event horizon from the Holy Cross Mountains (central Poland). Geological Magazine 144, 569–95.CrossRefGoogle Scholar
Marynowski, L., Filipiak, P. & Zatoń, M. 2010. Geochemical and palynological study of the Upper Famennian Dasberg event horizon from the Holy Cross Mountains (central Poland). Geological Magazine 147, 527–50.CrossRefGoogle Scholar
Marynowski, L., Kurkiewicz, S., Rakociński, M. & Simoneit, B. R. T. 2011. Effects of weathering on organic matter: I. changes in molecular composition of extractable organic compounds caused by paleoweathering of a Tournaisian marine black shale. Chemical Geology 285, 144–56.CrossRefGoogle Scholar
Marynowski, L., Narkiewicz, M. & Grelowski, C. 2000. Biomarkers as environmental indicators in a carbonate complex, examples from the Middle Devonian, the Holy Cross Mountains, Poland. Sedimentary Geology 137, 187212.CrossRefGoogle Scholar
Marynowski, L., Zatoń, M., Rakociński, M., Filipiak, P., Kurkiewicz, S. & Pearce, T. J. 2012. Deciphering the upper Famennian Hangenberg Black Shale depositional environments based on multi-proxy record. Palaeogeography, Palaeoclimatology, Palaeoecology 346/347, 68–66.Google Scholar
Matyja, H. 1993. Upper Devonian of Western Pomerania. Acta Geologica Polonica 43, 2794.Google Scholar
Matyja, H. 2006. Stratigraphy and facies development of Devonian and Carboniferous deposits in the Pomeranian basin and in western part of the Baltic basin, and palaeogeography of the northern TESZ during late Palaeozoic time. In Facies, Tectonic, and Thermal Evolution of the Pomeranian Sector of Trans-European Suture Zone and Adjacent Areas (eds Matyja, H. & Poprawa, P.), pp. 79–122. Prace Państwowego Instytutu Geologicznego 186 (in Polish with English summary).Google Scholar
Matyja, H. 2008. Pomeranian basin (NW Poland) and its sedimentary evolution during Mississippian times. In Carboniferous Platforms and Basins (eds Aretz, M., Herbig, H.-G. & Somerville, I. D.), pp. 123–50. Geological Journal 43.CrossRefGoogle Scholar
Matyja, H. 2009. Depositional history of the Devonian succession in the Pomeranian Basin, NW Poland. Geological Quarterly 53, 6392.Google Scholar
Matyja, H. & Stempień-Sałek, M. 1994. Devonian/Carboniferous boundary and the associated phenomena in Western Pomerania (NW Poland). Annales de la Société Géologique de Belgique 116, 249–63.Google Scholar
Matyja, H. & Turnau, E. 1989. Conodonts and spores from the Devonian/Carboniferous boundary beds in Poland. XI Congrès International de Stratigraphie et de Géologie du Carbonifère Beijing 1987, Compte Rendu 3, pp. 6172.Google Scholar
Merrill, G. K., Swift, A., Ryley, C. C., Barnes, C. R., O’Brien, F. H. C., Varker, W. J., Stone, J., Saunders, R., Fredholm, D. & Jeppson, L. 1987. Recent developments in conodont concentration techniques. In Conodonts: Investigative Techniques and Applications (ed. Austin, R. L.), pp. 5476. The British Micropalaeontological Society Series. Chichester: Ellis Horwood.Google Scholar
Meyer-Berthaud, B., Scheckler, S. E. & Wendt, J. 1999. Archaeopteris is the earliest known modern tree. Nature 398, 700–1.CrossRefGoogle Scholar
Myrow, P. M., Ramezani, J., Hanson, A., Bowring, S. A., Racki, G. & Rakociński, M. 2014. High-precision U-Pb age and duration of the Latest Devonian (Famennian) Hangenberg Event, and its implications. Terra Nova 26, 222–9.CrossRefGoogle Scholar
Nabbefeld, B., Grice, K., Summons, R. E., Hays, L. E. & Cao, C. 2010. Significance of polycyclic aromatic hydrocarbons (PAHs) in Permian/Triassic boundary sections. Applied Geochemistry 25, 1374–82.CrossRefGoogle Scholar
Narkiewicz, K., Grotek, I. & Matyja, H. 1998. Thermal maturity of organic matter in the Upper Devonian deposits of the Radom-Lublin and Pomerania area. In Sedimentary Basin Analysis of the Polish Lowland (ed. Narkiewicz, M.), pp. 235–44. Prace Państwowego Instytutu Geologicznego 165 (in Polish with English summary).Google Scholar
Naumova, S. N. 1953. Spore-pollen assemblages of the Russian platform and their stratigraphic significance. Trudy Instituta Geologicheskikh Nauk, Akademia Nauk USSR 143 (Geologicheskaya seriya 60), 1–204 (in Russian).Google Scholar
Olempska, E. 1997. Changes in benthic ostracod assemblages across the Devonian-Carboniferous boundary in the Holy Cross Mountains, Poland. Acta Palaeontologia Polonica 42, 291332.Google Scholar
Paproth, E., Feist, R. & Flajs, G. 1991. Decision on the Devonian-Carboniferous boundary stratotype. Episodes 14, 331–6.CrossRefGoogle Scholar
Peters, K. E., Walters, C. C. & Moldowan, J. M. 2005. The Biomarker Guide. Vol. 2. Cambridge University Press, 1155 pp.Google Scholar
Playford, G. 1964. Miospores from the Mississippian Horton Group, eastern Canada. Geological Survey of Canada Bulletin 107, 147.Google Scholar
Playford, G. 1971. Lower Carboniferous spores from the Bonaparte Gulf Basin, Western Australia and Northern Territory. Bulletin of the Bureau of Mineral Resources, Geology and Geophysics 115, 1105.Google Scholar
Playford, G. 1976. Plant microfossils from the Upper Devonian and Lower Carboniferous of the Canning Basin, Western Australia. Palaeontographica, Abt. B, 158, 171.Google Scholar
Playford, G. 1991. Australian Lower Carboniferous miospores relevant to extra-Gondwanic correlations: an evaluation. Courier Forschungsinstitut Senckenberg 130, 85125.Google Scholar
Potonié, R. 1966. Synopsis der Gattungen der Sporae dispersae. IV Teil: Nachträge zu allen Gruppen (Turmae). Beihefte zum Geologischen Jahrbuch 72, 3244.Google Scholar
Prestianni, C., Decombeix, A.-L., Thorez, J., Fokan, D. & Gerrienne, P. 2010. Famennian charcoal of Belgium. Palaeogeography, Palaeoclimatology, Palaeoecology 291, 6071.CrossRefGoogle Scholar
Racka, M., Marynowski, L., Filipiak, P., Sobstel, M., Pisarzowska, A. & Bond, D. P. J. 2010. Anoxic Annulata Events in the Late Famennian of the Holy Cross Mountains (Southern Poland): geochemical and palaeontological record. Palaeogeography, Palaeoclimatology, Palaeoecology 297, 549–75.CrossRefGoogle Scholar
Racki, G., Racka, M., Matyja, H. & Devleeschouwer, X. 2002. The Frasnian/Famennian boundary interval in the South Polish-Moravian shelf basins: integrated event-stratigraphical approach. In Late Devonian Biotic Crisis: Ecological, Depositional and Geochemical Records (eds Racki, G. & House, M. R.), pp. 251–97. Palaeogeography, Palaeoclimatology, Palaeoecology 181.Google Scholar
Radke, M., Welte, D. H. & Willsch, H. 1986. Maturity parameters based on aromatic hydrocarbons: influence of the organic matter type. In Advances on Organic Chemistry 1985 (eds Leythaenser, D. & Rullkötter, J.), pp. 51–63. Organic Geochemistry 10.Google Scholar
Radke, M. & Willsch, H. 1994. Extractable alkyldibenzothiophenes in Posidonia Shale (Toarcian) source rocks: relationship of yields to petroleum formation and expulsion. Geochimica and Cosmochimica Acta 58, 5223–44.CrossRefGoogle Scholar
Requejo, A. G., Allan, J., Creany, S., Gray, N. R. & Cole, K. S. 1992. Aryl isoprenoids and diaromatic carotenoids in Paleozoic source rocks and oils from the Western Canada and Williston basins. Organic Geochemistry 23, 205–22.CrossRefGoogle Scholar
Rimmer, S. M. 2004. Geochemical paleoredox indicators in Devonian-Mississippian black shales, Central Appalachian Basin (USA). Chemical Geology 206, 373–91.CrossRefGoogle Scholar
Rimmer, S. M., Thompson, J. A., Goodnight, S. A. & Robl, T. L. 2004. Multiple controls on the preservation of organic matter in Devonian-Mississippian marine black shales: geochemical and petrographic evidence. Palaeogeography, Palaeoclimatology, Palaeoecology 215, 125–54.CrossRefGoogle Scholar
Riquier, L., Averbuch, O., Devleeschouwer, X. & Tribovillard, N. 2010. Diagenetic versus detrital origin of the magnetic susceptibility variations in some carbonate Frasnian–Famennian boundary sections from Northern Africa and Western Europe: implications for paleoenvironmental reconstructions. International Journal of Earth Sciences (Geologische Rundschau) 99, 5773.CrossRefGoogle Scholar
Ross, C. A. & Ross, J. R. P. 1987. Late Paleozoic sea-levels and depositional sequences. In Timing and Depositional History of Eustatic Sequences: Constraints on Seismic Stratigraphy (eds Ross, C. A. & Haman, D.), pp. 137–49. Cushman Foundation for Foraminiferal Research, Special Publication 24.Google Scholar
Rowe, N. P. & Jones, T. P. 2000. Devonian charcoal. Palaeogeography, Palaeoclimatology, Palaeoecology, 164, 331–8.CrossRefGoogle Scholar
Sallan, L. C. & Coates, M. I. 2010. End-Devonian extinction and a bottleneck in the early evolution of modern jawed vertebrates. Proceedings of the National Academy of Sciences of the United States of America 107, 10131–5.Google Scholar
Sandberg, CH. A., Gutschick, R. C., Johnson, J. G., Poole, F. G. & Sando, W. J. 1986. Middle Devonian to late Mississippian event stratigraphy of overthrust belt region, western United States. Annales de la Société Géologique de Belgique 109, 205–7.Google Scholar
Sandberg, CH. A., Ziegler, W., Leuteritz, K. & Brill, S. M. 1978. Phylogeny, speciation, and zonation of Siphonodella (Conodonta, Upper Devonian and Lower Carboniferous). Newsletter on Stratigraphy 7, 102–20.CrossRefGoogle Scholar
Schneider, J., De Wall, H., Kontny, A., Bechstädt, T. 2004 a. Magnetic susceptibility variations in carbonates of the La Vid Group (Cantabrian Zone, NW-Spain) related to burial diagenesis. Sedimentary Geology 166, 7388.CrossRefGoogle Scholar
Schwark, L. & Frimmel, A. 2004. Chemostratigraphy of the Posidonia Black Shale, SW-Germany II. Assessment of extent and persistence of photic-zone anoxia using aryl isoprenoid distribution. Chemical Geology 206, 231–48.CrossRefGoogle Scholar
Scott, A. C. & Glasspool, I. J. 2006. The diversification of Paleozoic fire systems and fluctuations in atmospheric oxygen concentration. Proceedings of the National Academy of Sciences of the United States of America 103, 10861–5.Google Scholar
Sepkoski, J. J. 1996. Patterns of Phanerozoic extinction: perspective from global data bases. In Global Events and Event Stratigraphy (ed. Walliser, O. H.), pp. 3551. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Simakov, K. V., Bless, M. J. M., Bouckaert, J., Conil, R., Gagiev, M. H., Kolesov, Y. V., Onoprienko, Y. I., Poty, E., Razina, T. P., Shilo, N. A., Smirnova, L. V., Streel, M. & Swennen, R. 1983. Upper Famennian and Tournaisian deposits of the Omolon region (NE-USSR). Annales de la Société Géologique de Belgique 106, 335–99.Google Scholar
Staplin, F. L. & Jansonius, J. 1964. Elucidation of some Paleozoic densospores. Palaeontographica, B 114, 95117.Google Scholar
Steenwinkel, M. van. 1993. The Devonian-Carboniferous boundary: comparison between the Dinant Synclinorium and the northern border of the Rhenish Slate Mountains. Annales de la Société Géologique de Belgique 115, 665–81.Google Scholar
Stempień-Sałek, M. 2002. Miospore taxonomy and stratigraphy of Upper Devonian and lowermost Carboniferous in Western Pomerania (NW Poland). Annales Societatis Geologorum Poloniae 72, 163–90.Google Scholar
Stone, J. 1987. Review of investigative techniques used in the study of conodonts. In Conodonts: Investigative Techniques and Applications (ed. Austin, R. L.), pp. 1734. The British Micropalaeontological Society Series. Chichester: Ellis Horwood.Google Scholar
Streel, M. 1974. Similitudes des assemblages de spores d’Europe, d’Afrique du Nord et d’Amerique du Nord au Dévonien terminal. Bulletin Sciences Géologiques 27, 2537.CrossRefGoogle Scholar
Streel, M. 1999. Quantitative palynology of Famennian events in the Ardenne-Rhine regions. In North Gondwana: Mid-Paleozoic Terranes, Stratigraphy and Biota (eds Feist, R., Talent, J. A. & Daurer, A.), pp. 201–12. Abhandlungen der Geologischen Bundesanstalt 54.Google Scholar
Streel, M. 2009. Upper Devonian miospore and conodont zone correlation in western Europe. In Devonian Change: Case Studies in Palaeogeography and Palaeoecology (ed. Königshof, P.), pp. 163–76. Geological Society of London, Special Publication no. 314.Google Scholar
Streel, M., Caputo, M., Loboziak, S. & Melo, J. H. G. 2000. Late Frasnian-Famennian climates based on palynomorph analyses and the question of the Late Devonian glaciations. Earth-Science Reviews 52, 121–73.CrossRefGoogle Scholar
Streel, M., Higgs, K., Loboziak, S., Riegel, W. & Steemans, P. 1987. Spore stratigraphy and correlation with faunas and floras in the type marine Devonian of the Ardenne-Rhenish regions. Review of Palaeobotany and Palynology 50, 211–29.CrossRefGoogle Scholar
Streel, M. & Theron, J. N. 1999. The Devonian-Carboniferous boundary in South Africa and the age of the earliest episode of the Dwyka glaciation: new palynological result. Episodes 22, 41–4.CrossRefGoogle Scholar
Sullivan, H. 1964. Miospores from the Lower Limestone shales (Tournaisian) of the Forest of Dean Basin, Gloucestershire. Compte Rendu 5-eme Congrés International de Stratigraphie et de Géologie du Carbonifère, Paris (1963), 1249–59.Google Scholar
Sullivan, H. J. 1986. A Tournaisian spore flora from the Cementstone Group of Ayrshire, Scotland. Palaeontology 11, 116–31.Google Scholar
Summons, R. E. & Powell, T. G. 1987. Identification of aryl isoprenoids in source rocks and crude oils: biological markers for the green sulphur bacteria. Geochimica et Cosmochimica Acta 51, 557–66.CrossRefGoogle Scholar
Turnau, E. 1975. Correlations of Upper Devonian and Carboniferous deposits of Western Pomerania, based on miospore study. Rocznik Polskiego Towarzystwa Geologicznego (Annales de la Société Géologique de Pologne) 49, 231–69 [in Polish with English summary].Google Scholar
Turnau, E. 1978. Spore zonation of uppermost Devonian and lower Carboniferous of Western Pomerania. Mededelingen, Rijks Geologische Dienst 30, 134.Google Scholar
Turnau, E. 1979. Correlation of Upper Devonian and Carboniferous deposits of Western Pomerania, based on miospore study. Rocznik Polskiego Towarzystwa Geologicznego 49, 231–69 (in Polish with English summary).Google Scholar
Van Veen, P. M. 1981. Aspects of late Devonian and early Carboniferous palynology of southern Ireland; IV, morphological variation within Diducites a new formgenus to accommodate camerate spores with two-layered outer wall. Review of Palaeobotany and Palynology 31, 261–87.CrossRefGoogle Scholar
Walliser, O. H. 1996. Global events in the Devonian and Carboniferous. In Global Events and Event Stratigraphy in the Phanerozoic (ed. Walliser, O. H.), pp. 225–50. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Weissert, H., Joachimski, M. & Sarnthein, M. 2008. Chemostratigraphy. Newsletter on Stratigraphy 42, 145–79.CrossRefGoogle Scholar
Wignall, P. B. & Twitchett, R. J. 1996. Oceanic anoxia and the end Permian mass extinction. Science 272, 1155–8.CrossRefGoogle ScholarPubMed
Wood, G., Gabriel, A. M. & Lawson, J. C. 1996. Palynological techniques–processing and microscopy. In Palynology: Principles and Applications (eds Jansonius, J. & McGregor, D. C.), pp. 2950. American Association of Stratigraphic Palynologists Foundation, 1.Google Scholar
Żbikowska, B. 1992. Entomozoaceans (Ostracoda) from the Upper Devonian and Lower Carboniferous of Western Pomerania (in Polish). Przegląd Geologiczny 40, 612 (in Polish).Google Scholar
Zhang, S., Wang, X. & Hu, H. 2000. Magnetic susceptibility variations of carbonates controlled by sea-level changes – examples in Devonian to Carboniferous strata in southern Guizhou Province China. Science in China (Series D) 43, 266–76.CrossRefGoogle Scholar
Ziegler, P. 1990. Geological Atlas of Western and Central Europe. The Hague: Shell Internationale Petroleum Meatschappij B.V. Google Scholar
Ziegler, W. & Sandberg, C. A. 1984. Palmatolepis-based revision of upper part of standard Late Devonian conodont zonation. In Conodont Biofacies and Provincialism (ed. Clark, D. L.), pp. 179–94. Geological Society of America, Special Paper no. 196.CrossRefGoogle Scholar