Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-27T17:34:13.179Z Has data issue: false hasContentIssue false

Geochemical and palynological study of the Upper Famennian Dasberg event horizon from the Holy Cross Mountains (central Poland)

Published online by Cambridge University Press:  15 January 2010

LESZEK MARYNOWSKI
Affiliation:
University of Silesia, Faculty of Earth Sciences, Będzińska Str. 60, 41-200 Sosnowiec, Poland
PAWEŁ FILIPIAK*
Affiliation:
University of Silesia, Faculty of Earth Sciences, Będzińska Str. 60, 41-200 Sosnowiec, Poland
MICHAŁ ZATOŃ
Affiliation:
University of Silesia, Faculty of Earth Sciences, Będzińska Str. 60, 41-200 Sosnowiec, Poland
*
*Author for correspondence: filipiak@us.edu.pl

Abstract

Integrated palynological, organic and inorganic geochemical and petrographical methods have been used for deciphering the depositional redox conditions and character of organic matter of the Famennian Dasberg event horizon from the deep-shelf Kowala succession of the Holy Cross Mountains. The ages of the investigated samples have been established, using miospore data, as VF (Diducites versabilis–Grandispora famenensis) and LV (Retispora lepidophyta–Apiculiretusispora verrucosa) miospore Zones of the Middle/Upper Famennian. In the standard conodont zonation, this corresponds to the uppermost postera to lowermost praesulcata Zones. The presence of green sulphur bacteria biomarkers and dominance of small-sized framboids together with the presence of large framboids and low values of the U/Th ratio may indicate that during sedimentation of the lower Dasberg shale, intermittent anoxia occurred in the water column, or the anoxic conditions prevailed in the upper part of the water column, while the bottom waters were oxygenated, at least briefly. Deposition of the upper Dasberg shale was characterized by both bottom water and water column anoxia. The lack of acritarcha taxa from these intervals could have been due to anoxia in the photic zone. Moreover, organic content is high in those samples. There is no geochemical evidence for anoxia during sedimentation of the deposits sandwiched between the lower and upper Dasberg shales, or in the deposits which underlie and overlie both Dasberg shale horizons. The two discrete anoxic events are interpreted to be the result of major transgressions and the blooming of primary producers. Above the Dasberg shales, small fragments of charcoal and raised concentrations of polycyclic aromatic hydrocarbons are detected. This supports the presence of wildfires during deposition of shales just above the boundary of VF/LV palynological zones. Temperatures calculated from the fusinite reflectance values suggest that the charcoal was formed in low-temperature ground and/or surface fires. The typical marine character of sedimentation combined with the high proportion of charcoals suggests that wildfires were large-scale, and that there was intensive transport of terrestrial material. The main causes of intensive wildfires were a significant rise of O2 in the atmosphere and important progress in the land plant diversity during Late Devonian times. Palynofacies studies suggest that the transgression corresponds to the part IIf of the Late Devonian sea-level curve.

Type
Original Article
Copyright
Copyright © Cambridge University Press 2010

Access options

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

References

Algeo, T. J. & Scheckler, S. E. 1998. Terrestrial-marine teleconnections in the Devonian: links between the evolution of land plants, weathering processes, and marine anoxic events. Philosophical Transactions of the Royal Society London B 353, 113–30.CrossRefGoogle Scholar
Algeo, T. J. & Ingall, E. 2007. Sedimentary Corg:P ratios, paleocean ventilation, and Phanerozoic atmospheric pO2. Palaeogeography, Palaeoclimatology, Palaeoecology 256, 130–55.Google Scholar
Arinobu, T., Ishiwatari, R., Kaiho, K. & Lamolda, M. A. 1999. Spike of pyrosynthetic polycyclic aromatic hydrocarbons associated with an abrupt decrease in δ13C of a terrestrial biomarker at the Cretaceous–Tertiary boundary at Caravaca, Spain. Geology 27, 723–6.Google Scholar
Armstroff, A., Wilkes, H., Schwarzbauer, J., Littke, R. & Horsfield, B. 2006. Aromatic hydrocarbon biomarkers in terrestrial organic matter of Devonian to Permian age. Palaeogeography, Palaeoclimatology, Palaeoecology 240, 253–74.Google Scholar
Avkhimovitch, V. I. 1993. Zonation and spore complexes of Devonian and Carboniferous boundary deposits of the Pripyat depression (Byelorussia). Annales de la Société Géologique de Belgique 115, 425–52.Google Scholar
Avkhimovitch, 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 USRR and western Europe. Courier Forschungsinstitut Senckenberg 100, 169–91.Google Scholar
Avkhimovitch, V. I., Tchibrikova, E. V., Obukhovskaya, T. G., Nazarenko, A. M., Umnova, V. T., Raskatova, L. G., Mantsourova, V. N., Loboziak, S. & Streel, M. 1993. Middle and Upper Devonian miospore zonation of Eastern Europe. Bulletin Des Centres de Recherches Exploration–Production Elf-Aquitaine 17, 79147.Google Scholar
Bakr, M. M. Y. & Wilkes, H. 2002. The influence of facies and depositional environment on the occurrence and distribution of carbazoles and benzocarbazoles in crude oils: a case study from the Gulf of Suez, Egypt. Organic Geochemistry 33, 561–80.CrossRefGoogle Scholar
Bastow, T. P., Alexander, R., Fisher, S. J., Singh, R. K., Van Aarssen, B. G. K. & Kagi, R. I. 2000. Geosynthesis of organic compounds. Part V – methylation of alkylnaphthalenes. Organic Geochemistry 31, 523–34.Google Scholar
Batten, D. J. 1996. Chapter 26A. Palynofacies and palaeoenvironmental interpretation. In Palynology: principles and applications, Volume 3 (eds Jansonius, J. & McGregor, D. C.), pp. 1011–64. American Association of Stratigraphic Palynologists Foundation.Google Scholar
Becker, G., Bless, M. J. M., Streel, M. & Thorez, J. 1974. Palynology and ostracod distribution in the Upper Devonian and basal Dinantian of Belgium and their dependence on sedimentary facies. Mededelingen, Rijks Geologische Dienst 25, 999.Google Scholar
Becker, R. T. 1993. Anoxia, eustatic changes, and Upper Devonian to lowermost Carboniferous global ammonoid diversity. In The Ammonoidea: Environment, Ecology, and Evolution Change (ed. House, M. R.), pp. 115–63. Systematics Association Special Volume 47. Oxford: Clarendon Press.Google Scholar
Bergman, N. M., Lenton, T. M. & Watson, A. J. 2004. COPSE: a new model of biogeochemical cycling over Phanerozoic time. American Journal of Science 304, 397437.Google Scholar
Berkowski, B. 2002. Famennian rugosa and heterocorallia from southern Poland. Palaeontologica Polonica 61, 187.Google Scholar
Behrens, A., Wilkes, H., Schaeffer, P., Clegg, H. & Albrecht, P. 1998. Molecular characterization of organic matter in sediments from the Keg River formation (Elk Point group), western Canada sedimentary basin. Organic Geochemistry 29, 1905–20.CrossRefGoogle Scholar
Belcher, C. M. & McElwain, J. C. 2008. Limits for combustion in low O2 redefine paleoatmospheric predictions for the Mesozoic. Science 321, 11971200.Google Scholar
Bond, D. & Zatoń, M. 2003. Gamma-ray spectrometry across the Upper Devonian basin succession at Kowala in the Holy Cross Mountains (Poland). Acta Geololgica Pololonica 53, 93–9.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.Google Scholar
Bond, D. P. G. & Wignall, P. B. 2008. The role of sea-level change and marine anoxia in the Frasnian–Famennian (Late Devonian) mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology 263, 107–18.Google Scholar
Brown, T. C. & Kenig, F. 2004. Water column structure during deposition of Middle Devonian – Lower Mississippian black and green/gray shales of the Illinois and Michigan Basin: a biomarker approach. Palaeogeography, Palaeoclimatology, Palaeoecology 215, 5985.Google Scholar
Byvsheva, T. V. 1985. Spores from deposits of the Tournaisian and Visean stages of the Russian Plate. In Atlas of Spores and Pollen of Phanerozoic Oil and Gas-Bearing Strata of the Russian and Turanian Plates (eds Menner, V. V. & Byvsheva, T. V.), pp. 80158. Trudy Vsezoiuznogo Nauchno-Issledovatel'skogo Goelogorazvedochnogo Neftianogo Instituta (VNIGRI) 253 (in Russian).Google 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, 187207.Google Scholar
Chow, N., Wendte, J. & Stasiuk, L. D. 1995. Productivity versus preservation controls on two organic-rich carbonate facies in the Devonian of Alberta: sedimentological and organic petrological evidence. Bulletin of Canadian Petroleum Geology 43, 433–60.Google Scholar
Clifford, D. J., Clayton, J. L. & Sinninghe Damsté, J. S. 1998. 2,3,6-/3,4,5-Trimethyl substituted diaryl carotenoid derivatives (Chlorobiaceae) in petroleums of the Belarussian Pripyat River Basin. Organic Geochemistry 29, 1253–67.Google Scholar
Collinson, M. E., Steart, D. C., Scott, A. C., Glasspool, I. J. & Hooker, J. J. 2007. Episodic fire, runoff and deposition at the Palaeocene–Eocene boundary. Journal of the Geological Society, London 164, 8797.Google Scholar
Courtinat, B. 1998. New genera and new species of scolecodonts (fossil annelids) with paleoenvironmental and evolutionary considerations. Micropaleontology 44, 435–40.Google Scholar
Cressler, W. L. 2001. Evidence of earliest known wildfires. Palaios 16, 171–4.Google Scholar
Dąbrowska, K. & Filipiak, P. 2006. The new findings of macroflora from the Famennian of the Holy Cross Mountains (English summary). Przegląd Geologiczny 54, 720–3.Google Scholar
Deflandre, G. 1945. Microfossiles des calcaires siluriens de la Montagne Noire. Annales de Paleontologie 31, 3875.Google Scholar
Disnar, J. R. & Harouna, M. 1994. Biological origin of tetracyclic diterpanes, n-alkanes and other biomarkers found in Lower Carboniferous Gondwana coals (Niger). Organic Geochemistry 21, 143–52.Google Scholar
Dzik, J. 1997. Emergence and succession of Carboniferous conodont and ammonoid communities in the Polish part of the Variscan sea. Acta Palaeontologica Polonica 42, 57170.Google Scholar
Dzik, J. 2006. The Famennian ‘Golden Age’ of conodonts and ammonoids in the Polish part of the Variscan sea. Palaeontologica Polonica 63, 1360.Google Scholar
Elie, M., Faure, P., Michels, R., Landais, P. & Griffault, L. 2000. Natural and laboratory oxidation of low-organic-carbon-content sediments: comparison of chemical changes in hydrocarbons. Energy & Fuels 14, 854–61.CrossRefGoogle Scholar
Fairon-Demaret, M. & Hartkopf-Fröder, C. 2004. Late Famennian plant mesofossils from the Refrath 1 Borehole (Bergisch Gladbach-Pffrath Syncline; Ardennes-Rhenish Massif, Germany). Courier Forschungsinstitut Senckenberg 251, 89121.Google Scholar
Filipiak, P. 1996. The miospore horizons from the Devonian–Carboniferous boundary beds in the Bolechowice IG 1 borehole (Holy Cross Mountains). Geological Quarterly 40 (2), 169–84.Google Scholar
Filipiak, P. 2002. Palynofacies around the Frasnian/Famennian boundary in the Holy Cross Mountains, southern Poland. Palaeogeography, Palaeoclimatology, Palaeoecology 181, 313–24.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.Google Scholar
Filipiak, P. 2005. Late Devonian and Early Carboniferous acritarchs and prasinophyta from the Holy Cross Mountains (central Poland). Review of Palaeobotany and Palynology 134, 126.Google Scholar
Filipiak, P. & Racki, G. 2005. Unikatowy zapis zdarzeń beztlenowych w profile kamieniołomu Kowala k. Kielc (English summary). Przegląd Geologiczny 53, 846–7.Google Scholar
Finkelstein, 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
Gonzalez, F. 2009. Reappraisal of the organic-walled microphytoplankton genus Maranhites: morphology, excystment, and speciation. Review of Palaeobotany and Palynology 154, 621.Google Scholar
Gough, M. A., Rhead, M. M. & Rowland, S. J. 1992. Biodegradation studies of unresolved complex mixtures of hydrocarbons: model UCM hydrocarbons and the aliphatic UCM. Organic Geochemistry 18, 1722.CrossRefGoogle Scholar
Grice, K., Cao, C., Love, G. D., Böttcher, M. E., Twitchett, R. J., Grosjean, E., Summons, R. E., Turgeon, S. C., Dunning, W. & Jin, Y. 2005. Photic zone euxinia during the Permian–Triassic superanoxic event. Science 307, 706–9.CrossRefGoogle ScholarPubMed
Grice, K., Nabbefeld, B. & Maslen, E. 2007. Source and significance of selected polycyclic aromatic hydrocarbons in sediments (Hovea-3 well, Perth Basin, Western Australia) spanning the Permian–Triassic boundary. Organic Geochemistry 38, 17951803.Google Scholar
Guy-Ohlson, D. 1996. Chapter 7B. Prasinophycean algae. In Palynology: principles and applications, Volume 1 (eds Jansonius, J. & McGregor, D. C.), pp. 181–9. American Association of Stratigraphic Palynologists Foundation.Google Scholar
Habib, D. & Knapp, S. D. 1982. Stratigraphic utility of Cretaceous small acritarchs. Micropaleontology 28, 335–71.CrossRefGoogle Scholar
Hallam, A. & Wignall, P. B. 1999. Mass extinctions and sea-level changes. Earth-Science Reviews 48, 217–50.Google Scholar
Hartenfels, S. & Becker, R. T. In press. Timing of the global Dasberg Event – implications for Famennian eustasy and chronostratigraphy. Palaeontographica Americana.Google Scholar
Hartgers, W. A., Sinninghe Damsté, J. S., Koopmans, M. P. & De Leeuw, J. W. 1993. Sedimentary evidence for a diaromatic carotenoid with an unprecedented aromatic-substitution pattern. Journal of Chemical Society, Chemical Communications 23, 1715–16.Google Scholar
Hartgers, W. A., Sinninghe Damsté, J. S., Requejo, A. G., Allan, J., Hayes, J. M., Ling, Yue, Xie, Tian-Min, Primack, J. & De Leeuw, J. W. 1994. A molecular and carbon isotopic study towards the origin and diagenetic fate of diaromatic carotenoids. In Advances in Organic Geochemistry 1993 (eds Telnæs, N. et al. ), pp. 703–25. Organic Geochemistry 22.Google Scholar
Hartkopf-Fröder, C. 2004. Palynostratigraphy of upper Famennian sediment from the Refrath 1 Borehole (Bergish Gladbach-Paffrath Syncline; Ardennes-Rhenish Massif, Germany). Courier Forschungsinstitut Senckenberg 251, 7787.Google 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
Hays, L. E., Beatty, T., Henderson, C. M., Love, G. D. & Summons, R. E. 2007. Evidence for photic zone euxinia through the end-Permian mass extinction in the Panthalassic Ocean (Peace River Basin, Western Canada). Palaeoworld 16, 3950.CrossRefGoogle Scholar
Heimhofer, U., Hesselbo, S. P., Pancost, R. D., Martill, D. M., Hochuli, P. A. & Guzzo, J. V. P. 2008. Evidence for photic-zone euxinia in the Early Albian Santana Formation (Araripe Basin, NE Brazil). Terra Nova 20, 347–54.Google Scholar
House, M. R. 2002. Strength timing, setting and cause of mid-Palaeozoic extinctions. Palaeogeography, Palaeoclimatology, Palaeoecology 181, 525.Google Scholar
Jiang, C., Alexander, R., Kagi, R. I. & Murray, A. P. 1998. Polycyclic aromatic hydrocarbons in ancient sediments and their relationship to palaeoclimate. Organic Geochemistry 29, 1721–35.Google Scholar
Joachimski, M. M., Ostertag-Henning, C., Pancost, R. D., Strauss, H., Freeman, K. H., Littke, R., Sinninghe Damsté, J. S. & Racki, G. 2001. Water column anoxia, enhanced productivity and concomitant changes in δ13C and δ34S across the Frasnian–Famennian boundary (Kowala – Holy Cross Mountains/Poland). Chemical Geology 175, 109–31.Google Scholar
Johnson, J. G., Klapper, G. & Sandberg, C. A. 1986. Late Devonian eustatic cycles around margin of Old Red Continent. Annales de la Société géologique de Belgique 109, 141–7.Google Scholar
Jones, T. P. & Lim, B. 2000. Extraterrestrial impacts and wildfires. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 5766.Google Scholar
Kawka, O. E. & Simoneit, B. R. T. 1990. Polycyclic aromatic hydrocarbons in hydrothermal petroleums from the Guaymas Basin spreading center. Applied Geochemistry 5, 1727.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.Google Scholar
Killops, S. D. & Massoud, M. S. 1992. Polycyclic aromatic hydrocarbons of pyrolytic origin in ancient sediments: evidence for Jurassic vegetation fires. Organic Geochemistry 18, 17.Google 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 isorenieratane: Molecular indicators for photic zone anoxia. Geochimica et Cosmochimica Acta 60, 4467–96.Google Scholar
Korn, D. 2004. The mid-Famennian ammonoid succession in the Rhenish Mountains: the “annulata Event” reconsidered. Geological Quarterly 48, 245–52.Google Scholar
Köster, J., Rospondek, M., Schouten, S., Kotarba, M., Zubrzycki, A. & Sinninghe Damsté, J. S. 1998. Biomarker geochemistry of a foreland basin: Oligocene Menilite Formation in the Flysch Carpatians of Southeast Poland. In Advances in Organic Geochemistry 1997 (eds Horsfield, B., Radke, M., Schaefer, R. G. & Wilkes, H.), pp. 649–69. Organic Geochemistry 29.Google Scholar
Kruge, M. A., Stankiewicz, B. A., Crelling, J. C., Montanari, A. & Bensley, D. F. 1994. Fossil charcoal in Cretaceous–Tertiary boundary strata: evidence for catastrophic firestorm and megawave. Geochimica et Cosmochimica Acta 58, 1393–7.Google Scholar
Marynowski, L., Czechowski, F. & Simoneit, B. R. T. 2001. Phenylnaphthalenes and polyphenyls in Palaeozoic source rocks of the Holy Cross Mountains, Poland. Organic Geochemistry 32, 6985.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.Google Scholar
Marynowski, L., Narkiewicz, M. & Grelowski, C. 2000. Biomarkers as environmental indicators in a carbonate complex, example from the Middle to Upper Devonian, Holy Cross Mts., Poland. Sedimentary Geology 137, 187212.Google Scholar
Marynowski, L., Rakociński, M. & Zatoń, M. 2007. Middle Famennian (Late Devonian) interval with pyritized fauna from the Holy Cross Mountains (Poland): organic geochemistry and pyrite framboid diameter study. Geochemical Journal 41, 187200.Google Scholar
Marynowski, L., Filipiak, P. & Pisarzowska, A. 2008. Organic geochemistry and palynofacies of the Early–Middle Frasnian transition (Late Devonian) of the Holy Cross Mts, southern Poland. Palaeogeography, Palaeoclimatology, Palaeoecology 269, 152–65.Google Scholar
Maziane, N., Higgs, K. T. & Streel, M. 1999. Revision of late Famennian miospore zonation scheme in eastern Belgium. Journal of Micropaleontology 18, 1725.Google Scholar
Meyer, K. J. & Kump, L. R. 2008. Oceanic euxinia in Earth history: causes and consequences. Annual Review of Earth and Planetary Sciences 36, 251–88.Google Scholar
Moreau-Benoit, A. 1980. Les spores du Devonien de Libye. Cahiers de micropaléontologie 1, 353.Google Scholar
Murphy, A. E., Sageman, B. B. & Hollander, D. J. 2000. Eutrophication by decoupling of marine biogeochemical cycles of C, N, and P: a mechanism for the Late Gevonian mass extinction. Geology 28, 427–30.Google Scholar
Pancost, R. D., Freeman, K. H., Patzkowsky, M. E., Wavrek, D. A. & Collister, J. W. 1998. Molecular indicators of redox and marine photoautotroph composition in the late Middle Ordovician of Iowa, U.S.A. Organic Geochemistry 29, 1649–62.Google Scholar
Peters, K. E., Walters, C. C. & Moldowan, J. M. 2005. The Biomarker Guide. Vol. 2. Cambridge University Press, 1155 pp.Google Scholar
Pisarzowska, A., Sobstel, M. & Racki, G. 2006. Conodont-based event stratigraphy of the Early–Middle Frasnian transition on South Polish carbonate shelf. Acta Palaeontologica Polonica 51, 609–46.Google Scholar
Playford, G. 1976. Plant microfossils from the Upper Devonian and Lower Carboniferous of the Canning Basin, Western Australia. Palaeontographica B 158B, 171.Google Scholar
Playford, G. 1977. Lower to Middle Devonian Acritarchs of the Moose River Basin, Ontario. Geological Survey of Canada, Bulletin 279, 187.Google Scholar
Playford, G. & Dring, R. S. 1981. Late Devonian acritarchs from the Carnarvon Basin, Western Australia. Special Papers in Palaeontology 27, 178.Google Scholar
Prestianni, C., Decombeix, A.-L., Thorez, J., Fokan, D. & Gerrienne, P. 2009. Famennian charcoal of Belgium. Palaeogeography, Palaeoclimatology, Palaeoecology doi:10.1016/j.palaeo.2009.10.008, in press.Google Scholar
Racka, M. & Marynowski, L. 2008. Geochemical proxies of the late Famennian Annulata event from the Kowala quarry, Holy Cross Mountains, Poland (in Polish). Pierwszy Polski Kongres Geologiczny, June 26–28, 2008. Abstract, Polskie Towarzystwo Geologiczne, Kraków 95.Google Scholar
Racki, G. 1998. Frasnian–Famennian biotic crisis: undervalued tectonic control? Palaeogeography, Palaeoclimatology, Palaeoecology 141, 177–98.Google Scholar
Racki, G. 2005. Toward understanding Late Devonian global events: few answers, many questions. In Understanding Late Devonian and Permian–Triassic Biotic and Climatic Events: Towards an Integrated Approach (eds Over, D. J., Morrow, J. R. & Wignall, P. B.), pp. 536. Amsterdam: Elsevier.Google 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. Palaeogeography, Palaeoclimatology, Palaeoecology 181, 251–97.Google Scholar
Radke, M., Vriend, S. P. & Ramanampisoa, L. R. 2000. Alkyldibenzofurans in terrestrial rocks: Influence of organic facies and maturation. Geochimica et Cosmochimica Acta 64, 275–86.Google Scholar
Rauscher, R. 1969. Présence d'une forme nouvelle d'Acritarches dans le Dévonien de Normandie. Comptes Rendus des Séances de l'Académie Sciences Paris, série D 268, 34–6.Google 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.Google Scholar
Rimmer, S. M. & Scott, A. C. 2006. Charcoal (inertinite) in Late Devonian marine black shales: Implications for terrestrial and marine systems and for paleo-atmospheric composition. European Geosciences Union General Assembly, April 2–7, 2006, Vienna, Austria, Geophysical Research Abstracts 8, 07972.Google Scholar
Rimmer, S. M., Scott, A. C. & Cressler, W. L. 2006, Terrestrial-marine linkages: Records of Devonian charcoal, fire, and paleoatmospheric oxygen level. 2006 Annual Meeting Geological Society of America, October 22–25, 2006, Philadelphia, PA, Program with Abstracts 38, 341.Google Scholar
Rowe, N. P. & Jones, T. P. 2000. Devonian charcoal. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 331–8.Google Scholar
Sageman, B. B., Murphy, A. E., Werne, J. P., Ver Straeten, Ch. A., Hollander, D. J. & Lyons, T. W. 2002. A tale of shales: the relative roles of production, decomposition, and dilution in the accumulation of organic-rich strata, Middle–Upper Devonian, Appalachian basin. Chemical Geology 195, 229–73.Google Scholar
Schieber, J. 2009. Discovery of agglutinated benthic foraminifera in Devonian black shales and their relevance for the redox state of ancient seas. Palaeogeography, Palaeoclimatology, Palaeoecology 271, 292300.Google Scholar
Schwab, V. F. & Spangenberg, J. E. 2007. Molecular and isotopic characterization of biomarkers in the Frick Swiss Jura sediments: A palaeoenvironmental reconstruction on the northern Tethys margin. Organic Geochemistry 38, 419–39.Google 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.Google Scholar
Scott, A. C. 2000. The Pre-Quaternary history of fire. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 281329.Google Scholar
Scott, A. C. & Glasspool, I. J. 2006. The diversification of Paleozoic fire systems and fluctuations in atmospheric oxygen concentration. PNAS 103, 10861–5.Google Scholar
Sephton, A., Looy, C. V., Brinkhuist, H., Wignall, P. B., De Leeuw, J. W. & Visscher, H. 2005. Catastrophic soil erosion during the end-Permian biotic crisis. Geology 33, 941–4.Google Scholar
Simoneit, B. R. T. & Fetzer, J. C. 1996. High molecular weight polycyclic aromatic hydrocarbons in hydrotermal petroleums from the Gulf of California and Northeast Pacific Ocean. Organic Geochemistry 24, 1065–77.Google Scholar
Sinninghe Damsté, J. S. & Schouten, S. 2005. Biological markers for anoxia in the photic zone of the water column. Handbook of Environmental Chemistry 2, 137.Google Scholar
Sinninghe Damsté, J. S. & Hopmans, E. C. 2008. Does fossil pigment and DNA data from Mediterranean sediments invalidate the use of green sulfur bacterial pigments and their diagenetic derivatives as proxies for the assessment of past photic zone euxinia? Environmental Microbiology 10, 1392–9.CrossRefGoogle ScholarPubMed
Sinninghe Damsté, J. S., Kenig, F., Koopmans, M. P., Köster, J., Schouten, S., Hayes, J. M. & De Leeuw, J. W. 1995. Evidence for gammacerane as an indicator of water column stratification. Geochimica et Cosmochimica Acta 59, 18951900.Google Scholar
Staplin, F. L. 1961. Reef-controlled distribution of Devonian microplancton in Alberta. Paleontology 4 (3), 392424.Google Scholar
Streel, M., Higgs, K., Loboziak, S., Riegel, W. & Steemans, P. 1987. Spore stratigraphy and correlation with faunas and floras in type marine Devonian of the Ardenne-Rhenish region. Review of Palaeobotany and Palynology 50, 211–29.Google Scholar
Summerhayes, C. P. 1987. Organic-rich Cretaceous sediments from the North Atlantic. In Marine petroleum source rocks (eds Brooks, J. & Fleet, A. J.), pp. 301–16. Geological Society of London, Special Publication no. 26.Google Scholar
Summons, R. E. & Powell, T. G. 1986. Chlorobiaceae in Paleozoic seas revealed by biological markers, isotopes and geology. Nature 319, 763–5.Google Scholar
Szulczewski, M. 1971. Upper Devonian conodonts, stratigraphy and facial development in the Holy Cross Mts. Acta Geologica Pololonica 21, 1129.Google Scholar
Tappan, H. 1980. The palaeobiology of plant protists. San Francisco: W. H. Freeman and Co., 1028 pp.Google Scholar
Tappan, H. 1982. Extinction or survival: Selectivity and causes of Phanerozoic crises. Geological Society of America, Special Paper 190, 265–76.Google Scholar
Tappan, H. 1986. Phytoplankton: Below the salt at the global table. Journal of Palaeontology 60, 545–54.Google Scholar
Tinner, W., Hofstetter, S., Zeugin, F., Conedera, M., Wohlgemuth, T., Zimmerman, L. & Zweifel, R. 2006. Long-distance transport of macroscopic charcoal by an intensive crown fire in the Swiss Alps – implications for fire history reconstruction. The Holocene 16, 287–92.Google Scholar
Trela, W. & Malec, J. 2007. Carbon isotope record in sediments of the Devonian–Carboniferous boundary in the southern Holy Cross Mountains. (English summary). Przegląd Geologiczny 55, 411–15.Google Scholar
Turnau, E. 1975. Micloflora of the Famenian and Tournaisian deposits from boreholes of northern Poland. Acta Geologica Polonica 25, 505–28.Google Scholar
Turnau, E. 1990. Spore zones of Famennian and Tournaisian deposits from the Kowla 1 borehole. (English summary). Geological Quarterly 34, 291304.Google Scholar
Tyson, R. V. 1993. Palynofacies analysis. In Applied micropaleontology (ed. Jenkins, D. G.), pp. 153–91. Kluwer Academic Publishers.Google Scholar
Tyson, R. V. 1995. Sedimentary Organic Matter. Organic Facies and Palynofacies. London: Chapman and Hall, 615 pp.Google Scholar
Van Veen, P. M. 1981. Aspects of Late Devonian Palynology of Southern Ireland, IV: Morphological variation within Diducites a new form genus to accommodate camerate spores with a two layered outer wall. Review of Palaeobotany and Palynology 31, 261–87.Google Scholar
Venkatesan, M. I. & Dahl, J. 1989. Organic geochemical evidence for global fires at the Cretaceous/Tertiary boundary. Nature 338, 5760.Google Scholar
Wang, Ch. & Visscher, H. 2007. Abundance anomalies of aromatic biomarkers in the Permian–Triassic boundary section at Meishan, China – Evidence of end-Permian terrestrial ecosystem collapse. Palaeogeography, Palaeoclimatology, Palaeoecology 252, 291303.Google Scholar
Watson, J. S., Sephton, M. A., Looy, C. V. & Gilmour, I. 2005. Oxygen-containing aromatic compounds in a Late Permian sediment. Organic Geochemistry 36, 371–84.Google Scholar
Wicander, R. 1974. Upper Devonian–Lower Mississippian acritarchs and prasinophycean algae from Ohio, USA. Palaeontographica Abteilung B 148, 943.Google Scholar
Wignall, P. B. 1994. Black shales. Oxford: Clarendon Press, 127 pp.Google Scholar
Wignall, P. B. & Newton, R. 1998. Pyrite framboid diameter as a measure of oxygen deficiency in ancient mudrocks. American Journal of Science 298, 537–52.Google Scholar
Wilkin, R. T., Barnes, H. L. & Brantley, S. L. 1996. The size distribution of framboidal pyrite in modern sediments: an indicator of redox conditions. Geochimica et Cosmochimica Acta 60, 3897–912.Google Scholar
Yans, J., Corfield, R. M., Racki, G. & Préat, A. 2007. Evidence for major perturbation of carbon cycle in the middle Frasnian punctata conodont Zone. Geological Magazine 144, 263–70.Google Scholar
Ziegler, W. & Sandberg, C. A. 1990. The Late Devonian Standard Conodont Zonation. Courier Forschungsinstitut Senckenberg 121, 1115.Google Scholar
Żakowa, H., Nehring-Lefeld, M. & Malec, J. 1985. Devonian–Carboniferous boundary in the borehole Kowla 1 (Southern Holy Cross Mts, Poland). Macro- and microfauna. Geological Quarterly 33, 8795.Google Scholar
Supplementary material: File

Marynowski supplementary material

Appendix.doc

Download Marynowski supplementary material(File)
File 84.5 KB
Supplementary material: Image

Marynowski supplementary material

Figure.jpg

Download Marynowski supplementary material(Image)
Image 402.7 KB