Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-04-30T22:47:07.171Z Has data issue: false hasContentIssue false

Late Jurassic – earliest Cretaceous prolonged shelf dysoxic–anoxic event and its possible causes

Published online by Cambridge University Press:  19 August 2020

MA Rogov*
Affiliation:
Geological Institute of RAS, Pyzhevski lane 7/1, Moscow 119017, Russia Saint-Petersburg State University, Universitetskaya nab. 7/9, Saint Petersburg 199034, Russia
EV Shchepetova
Affiliation:
Geological Institute of RAS, Pyzhevski lane 7/1, Moscow 119017, Russia
VA Zakharov
Affiliation:
Geological Institute of RAS, Pyzhevski lane 7/1, Moscow 119017, Russia
*
Author for correspondence: MA Rogov, Email: russianjurassic@gmail.com

Abstract

The Late Jurassic – earliest Cretaceous time interval was characterized by a widespread distribution of dysoxiс–anoxiс environments in temperate- and high-latitude epicontinental seas, which could be defined as a shelf dysoxic–anoxic event (SDAE). In contrast to black shales related to oceanic anoxic events, deposits generated by the SDAE were especially common in shelf sites in the Northern Hemisphere. The onset and termination of the SDAE was strongly diachronous across different regions. The SDAE was not associated with significant disturbances of the carbon cycle. Deposition of organic-carbon-rich sediment and the existence of dysoxic–anoxic conditions during the SDAE lasted up to c. 20 Ma, but this event did not cause any remarkable biotic extinction. Temperate- and high-latitude black shale occurrences across the Jurassic–Cretaceous boundary have been reviewed. Two patterns of black shale deposition during the SDAE are recognized: (1) Subboreal type, with numerous thin black shale beds, bounded by sediments with very low total organic carbon (TOC) values; and (2) Boreal type, distinguished by predominantly thick black shale successions showing high TOC values and prolonged anoxic–dysoxic conditions. These types appear to be unrelated to differences in accommodation space, and can be clearly recognized irrespective of the thickness of shale-bearing units. Black shales in high-latitude areas in the Southern Hemisphere strongly resemble Boreal types of black shale by their mode of occurrence. The causes of this SDAE are linked to long-term warming and changes in oceanic circulation. Additionally, the long-term disturbance of planktonic communities may have triggered overall increased productivity in anoxia-prone environments.

Type
Original Article
Copyright
© The Author(s), 2020. Published by Cambridge University Press

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

Afanasenkov, AP, Petrov, AL and Grayzer, EM (2018) Geochemical description and oil-and-gas generation potential of Mesozoic formations within the Gydan and Yenisei-Khatanga oil and gas bearing regions. Oil and Gas Geology 6, 109–27 (in Russian). doi: 10.31087/0016-7894-2018-6-109-127Google Scholar
Afanasieva, MS and Mikhailova, MV (2001) The Domanik Formation of the Timan-Pechora basin: Radiolarians, biostratigraphy, and sedimentation conditions. Stratigraphy and Geological Correlation 9, 419–40.Google Scholar
Alsen, P and Piasecki, S (2018) Biostratigraphy of the Hareelv Formation (Upper Jurassic) in the Blokelv-1 core, Jameson Land, central East Greenland. Geological Survey of Denmark and Greenland Bulletin 42, 15–37. doi: 10.34194/geusb.v42.4308CrossRefGoogle Scholar
Alsgaard, PC, Felt, VL, Vosgerau, H and Surlyk, F (2003) The Jurassic of Kuhn Ø, North-East Greenland. Geological Survey of Denmark and Greenland Bulletin 1, 865–92.10.34194/geusb.v1.4691CrossRefGoogle Scholar
Arora, A, Banerjee, S and Dutta, S (2015) Black shale in late Jurassic Jhuran Formation of Kutch: possible indicator of oceanic anoxic event? Journal of the Geological Society of India 85, 265–78. doi: 10.1007/s12594-015-0215-6CrossRefGoogle Scholar
Arthur, MA and Sageman, BB (1994) Marine black shales: depositional mechanisms and environments of ancient deposits. Annual Review in Earth and Planetary Sciences 22, 499551. doi: 10.1146/annurev.ea.22.050194.002435CrossRefGoogle Scholar
Badics, B, Avu, A and Mackie, S (2015) Assessing source rock distribution in Heather and Draupne Formations of the Norwegian North Sea: a workflow using organic geochemical, petrophysical, and seismic character. Interpretation 3, SV4568. doi: 10.1190/INT-2014-0242.1CrossRefGoogle Scholar
Baraboshkin, EYu (2004) Boreal-Tethyan correlation of Lower Cretaceous ammonite scales Moscow University Geology Bulletin 59, 920.Google Scholar
Basov, VA, Nikitenko, BL and Kupriyanova, NV (2009) Lower–Middle Jurassic foraminiferal and ostracode biostratigraphy of the Barents Sea shelf. Russian Geology and Geophysics 50, 396416. doi: 10.1016/j.rgg.2008.08.006CrossRefGoogle Scholar
Baudin, F, Tribovillard, N, Laggoun-Défarge, F, Lichtfouse, E, Monod, O and Gardin, S (1999) Depositional environment of a Kimmeridgian carbonate ‘black band’(Akkuyu Formation, south-western Turkey). Sedimentology 46, 589602. doi: 10.1046/j.1365-3091.1999.00226.xCrossRefGoogle Scholar
Bayliss, G and Magoon, LB (1988) Organic facies and thermal maturity of sedimentary rocks in the National Petroleum Reserve in Alaska. United States Geological Survey Professional Paper 1399, 489518.Google Scholar
Bird, KJ and Molenaar, CM (1987) Stratigraphy. In Petroleum Geology of the Arctic National Wildlife Refuge (eds Bird, KJ and Magoon, LB), pp. 3759. Northeastern Alaska: United States Geological Survey, Bulletin 1778.Google Scholar
Bojesen-Koefoed, JA, Bjerager, M, Nytoft, HP, Petersen, HI, Piasecki, S and Pilgaard, A (2018) Petroleum potential of the Upper Jurassic Hareelv Formation, Jameson Land, East Greenland. Geological Survey of Denmark and Greenland Bulletin 42, 85113. doi: 10.34194/geusb.v42.4314CrossRefGoogle Scholar
Bown, PR, Lees, JA and Young, JR (2004) Calcareous nannoplankton evolution and diversity through time. In Coccolithophores. From Molecular Processes to Global Impact (eds Thierstein, HR and Young, JR), pp. 481508. Berlin, Heidelberg: Springer.Google Scholar
Braduchan, YuV, Gurari, FG, Zakharov, VA, Bulynnikova, SP, Vyachkileva, NP, Golbert, AV, Klimova, IG, Kozlova, GE, Lebedev, AI, Mesezhnikov, MS, Nalnyaeva, TI and Turbina, AS (1986) Bazhenovo horizon of West Siberia (stratigraphy, paleogeography, ecosystem, oil-and-gas content). Siberian Branch of the Russian Academy of Science, Transactions of the Institute of Geology and Geophysics, 649, 1216 (in Russian).Google Scholar
Braduchan, YuV, Zakharov, VA and Mesezhnikov, MS (1989) Stratigraphy and depositional conditions of bituminous deposits of Upper Jurassic - Neocomian of the European part of USSR and Western Siberia. In Sedimentary Cover of the Earth in Space and Time. Stratigraphy and Paleontology (ed. Sokolov, BS), pp. 108–15. Мoscow: Nauka (in Russian).Google Scholar
Bragin, VYu, Dzyuba, OS, Kazansky, AYu and Shurygin, BN (2013) New data on the magnetostratigraphy of the Jurassic–Cretaceous boundary interval, Nordvik Peninsula (northern East Siberia). Russian Geology and Geophysics 54, 335–48. doi: 10.1016/j.rgg.2013.02.008CrossRefGoogle Scholar
Bushnev, DA, Shchepetova, EV and Lyyurov, SV (2006) Organic geochemistry of Oxfordian carbon-rich sedimentary rocks of the Russian Plate. Lithology and Mineral Resources 41, 423–34. doi: 10.1134/s0024490206050038CrossRefGoogle Scholar
Callomon, JH and Cope, JCW (1971) The stratigraphy and ammonite succession of the Oxford and Kimmeridge clays in the Warlingham Borehole. Bulletin of the Geological Survey of Great Britain 36, 147–76.Google Scholar
Cariou, E, Enay, R, Bassoullet, J-P and Colchen, M (1994) Biochronologie du Jurassique moyen de la Thakkhola (Népal central) et biogéographie du domaine himalayen. Comptes Rendus de l’Académie des Sciences, Paris, Séries II 318, 93–9.Google Scholar
Carmeille, M, Bourillot, R, Pellenard, P, Dupias, V, Schnyder, J, Riquier, L, Mathieu, O, Brunet, MF, Enay, R, Grossi, V and Gaborieau, C (2020) Formation of microbial organic carbonates during the late Jurassic from the Northern Tethys (Amu Darya Basin, Uzbekistan): implications for Jurassic anoxic events. Global and Planetary Change 186, 103127. doi: 10.1016/j.gloplacha.2020.103127CrossRefGoogle Scholar
Cecca, F (1999) Palaeobiogeography of Tethyan ammonites during the Tithonian (latest Jurassic). Palaeogeography, Palaeoclimatology, Palaeoecology 147, 137. doi: 10.1016/s0031-0182(98)00149-7CrossRefGoogle Scholar
Chen, L, Tien-Shun Lin, A, Da, X, Yi, H, Tsai, LL-Y and Xu, G (2012) Sea-level changes recorded by Cerium anomalies in the Late Jurassic (Tithonian) black rock series of Qiangtang Basin, North-Central Tibet. Oil Shale 29, 1835. doi: 10.3176/oil.2012.1.03CrossRefGoogle Scholar
Clark, DN, Riley, LA and Ainsworth, NR (1993) Stratigraphic, structural and depositional history of the Jurassic in the Fisher Bank Basin, UK North Sea. Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference (ed. by JR Parker), pp. 415–24. Geological Society of London.CrossRefGoogle Scholar
Cope, JCW (1967) The palaeontology and stratigraphy of the lower part of the Upper Kimmeridge Clay of Dorset. Bulletin of the British Museum (Natural History) Geology 15, 179.Google Scholar
Cope, JCW (1978) The ammonite fauna and stratigraphy of the upper part of the Upper Kimmeridge Clay of Dorset. Palaeontology 21, 469533.Google Scholar
Cornford, C (1994) Mandal-Ekofisk petroleum system in the Central Graben of the North Sea. In The Petroleum System from Source to Trap (eds Magoon, LB and Dow, WG), pp. 523–39. American Association of Petroleum Geologists, Memoir no. 60. doi: 10.1306/m60585c33Google Scholar
Cornford, C (1998) Source rocks and hydrocarbons of the North Sea. In Petroleum Geology of the North Sea, 4th ed. (ed. Glennie, KW), pp. 376462. London: Blackwell Science Ltd.CrossRefGoogle Scholar
Dalland, A, Worsley, D and Ofstad, K (1988) A lithostratigraphic scheme for the Mesozoic and Cenozoic succession off shore mid- and northern Norway. Norwegian Petroleum Directorate Bulletin 4, 165.Google Scholar
Davison, I (2005) Central Atlantic margin basins of North West Africa: geology and hydrocarbon potential (Morocco to Guinea). Journal of African Earth Sciences 43, 254–74. doi: 10.1016/j.jafrearsci.2005.07.018CrossRefGoogle Scholar
Delsett, LL, Novis, LK, Roberts, AJ, Koevoets, MJ, Hammer, Ø, Druckenmiller, PS and Hurum, JH (2016) The Slottsmoya marine reptile Lagerstatte: depositional environments, taphonomy and diagenesis. In Dinosaurs and Other Extinct Saurians: A Historical Perspective (eds Moody, RTJ, Buffetaut, E, Naish, D and Martill, DM), pp. 165–88. Geological Society of London, Special Publication no. 434. doi: 10.1144/SP434.2Google Scholar
Delvaux, D, Martin, H, Leplat, P and Paulet, J (1990) Geochemical characterization of sedimentary organic matter by means of pyrolysis kinetic parameters. Organic Geochemistry 16, 175–87. doi: 10.1016/0146-6380(90)90038-2CrossRefGoogle Scholar
Deroo, G, Herbin, JP and Roucaché, J (1983) Organic geochemistry of Upper Jurassic–Cretaceous sediments from Site 511, Leg 71, Western South Atlantic. Initial Reports of the Deep Sea Drilling Project 71, 1001–13. doi: 10.2973/dsdp.proc.71.137.1983Google Scholar
Doyle, P and Whitham, AG (1991) Palaeoenvironments of the Nordenskjöld Formation: an Antarctic Late Jurassic-Early Cretaceous black shale-tuff sequence. In Modern and Ancient Continental Shelf Anoxia (eds Tyson, RV and PEarson, TH), pp. 397414. Geological Society of London, Special Publication no. 58. doi: 10.1144/gsl.sp.1991.058.01.25Google Scholar
Dypvik, H (1985) Jurassic and Cretaceous black shales of the Janusfjellet Formation, Svalbard, Norway. Sedimentary Geology 41, 235–48. doi: 10.1016/0037-0738(84)90064-2CrossRefGoogle Scholar
Dzyuba, OS, Pestchevitskaya, EB, Urman, OS, Shurygin, BN, Alifirov, AS, Igolnikov, AE and Kosenko, IN (2018) The Maurynya section, West Siberia: a key section of the Jurassic-Cretaceous boundary deposits of shallow marine genesis. Russian Geology and Geophysics 59, 864–90. doi: 10.1016/j.rgg.2018.07.010CrossRefGoogle Scholar
Eichenseer, K, Balthasar, U, Smart, CW, Stander, J, Haaga, KA and Kiessling, W (2019) Jurassic shift from abiotic to biotic control on marine ecological success. Nature Geoscience 12, 638–42. doi: 10.1038/s41561-019-0392-9CrossRefGoogle Scholar
Enay, R (1972) Paléobiogéographie des ammonites du Jurassique terminal (Tithonique/Volgien/Portlandien sl) et mobilité continentale. Geobios 5, 355407. doi: 10.1016/s0016-6995(72)80013-5CrossRefGoogle Scholar
Enay, R (2009) Les faunes d’ammonites de l’Oxfordien au Tithonien et la biostratigraphie des Spiti-shales (Callovien supérieur-Tithonien) de Thakkhola, Népal central. Documents des Laboratoires de Géologie, Lyon 166, 1350.Google Scholar
Etches, S and Clarke, J (1999) Steve Etches Kimmeridge Collection Illustrated Catalogue. Privately printed, Chandler’s Ford, Hants (with additional pages added in 2000 and 2001), 131 pp.Google Scholar
Gallois, RW (2004) The Kimmeridge Clay: the most intensively studied formation in Britain. Open University Geological Society Journal 25, 3338.Google Scholar
Gallois, RW (2005) Correlation of the Kimmeridgian succession of the Normandy coast, northern France with that of the Dorset-type area, southern England. Comptes rendus Geosciences 337, 347–55. doi: 10.1016/j.crte.2004.12.001CrossRefGoogle Scholar
Gallois, RW (2011) A revised description of the lithostratigraphy of the Kimmeridgian-Tithonian and Kimmeridgian-Volgian boundary beds at Kimmeridge, Dorset, UK. Geoscience in South-West England 12, 288–94.Google Scholar
Galloway, J, Vickers, M, Price, G, Poulton, T, Grasby, S, Hadlari, T, Beauchamp, B and Sulphur, K (2019) Finding the VOICE: organic carbon isotope chemostratigraphy of the Late Jurassic-Early Cretaceous of Arctic Canada. Geological Magazine, published online 20 December 2019, 115. doi: 10.1017/s0016756819001316Google Scholar
Gautier, DL (2005) Kimmeridgian Shales Total Petroleum System of the North Sea Graben Province. U.S. Geological Survey Bulletin 2204-C, 124. doi: 10.3133/b2204cGoogle Scholar
Gavrilov, YuO, Shchepetova, EV and Shcherbinina, EA (2014) Sedimentological and geochemical conditions of carbonaceous units occurrences in Mesozoic palaeobasins of the European part of Russia. Geresources, Geoenergetics, Geopolitics 1, 130 (in Russian).Google Scholar
Gentzis, T, Goodarzi, F and Embry, AF (1996) Thermal maturation, potential source rocks and hydrocarbon generation in Mesozoic rocks, Lougheed Island area, Central Canadian Arctic archipelago. Marine and Petroleum Geology 13, 879905. doi: 10.1016/s0264-8172(96)00028-1CrossRefGoogle Scholar
Georgiev, SV, Stein, HJ, Hannah, JL, Xu, G, Bingen, B and Weiss, HM (2017) Timing, duration, and causes for Late Jurassic–Early Cretaceous anoxia in the Barents Sea. Earth and Planetary Science Letters 461, 151–62. doi: 10.1016/j.epsl.2016.12.035CrossRefGoogle Scholar
Geyssant, JR, Vidier, J-P, Herbin, J-P, Proust, J-N and Deconinck, J-F (1993) Biostratigraphie et paléoenvironnement des couches de passage Kimméridgien/Tithonien du Boulonnais (Pas-de-Calais): nouvelles données paléontologiques (ammonites), organisation séquentielle et contenu en matière organique. Géologie de la France 4, 1124.Google Scholar
Głowniak, E, Kiselev, DN, Rogov, M, Wierzbowski, A and Wright, J (2010) The Middle Oxfordian to lowermost Kimmeridgian ammonite succession at Mikhalenino (Kostroma District) of Russian Platform, and its stratigraphical and palaeogeographical importance. Volumina Jurassica VIII, 845.Google Scholar
Goldhammer, RK and Johnson, CA (2002) Middle Jurassic-Upper Cretaceous paleogeographic evolution and sequence-stratigraphic framework of the northwest Gulf of Mexico rim. In The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems (eds Bartolini, C, Buffler, RT, and Cantú-Chapa, A), 4581. American Association of Petroleum Geologists, Memoir no. 75. doi: 10.1306/m75768c3Google Scholar
Goncharov, IV, Fadeeva, SV, Samoilenko, VV, Oblasov, NV and Bakhtina, ES (2014) Generation potential of organic matter Bazhenov Formation rocks in the south-east of Western Siberia (Tomsk region). Neftyanoe khozyaystvo = Oil Industry 12, 1216 (in Russian).Google Scholar
Grishkevich, VF (2018) Neocomian paleogeography, gas hydrate cementation of sediments, and abnormal sequences of the Bazhenov Formation (West Siberia). Russian Geology and Geophysics 59, 157–67. doi: 10.1016/j.rgg.2018.01.013CrossRefGoogle Scholar
Håkansson, E, Birkelund, T, Piasecki, S and Zakharov, V (1981) Jurassic–Cretaceous boundary strata of the extreme Arctic (Peare Land, North Greenland). Bulletin of the Geological Society of Denmark 30, 1142.CrossRefGoogle Scholar
Hakimi, MH and Ahmed, AF (2016) Petroleum generation modeling of the organic-rich shales of Late Jurassic–Early Cretaceous succession from Mintaq-01 well in the Wadi Hajar sub-basin, Yemen. Canadian Journal of Earth Sciences 53, 1053–72. doi: 10.1139/cjes-2015-0224CrossRefGoogle Scholar
Hallam, A (1969) Faunas realms and facies in the Jurassic. Palaeontology 12, 118.Google Scholar
Hammer, Ø, Collignon, M and Nakrem, HA (2012) Organic carbon isotope chemostratigraphy and cyclostratigraphy in the Volgian of Svalbard. Norwegian Journal of Geology 92, 103–12.Google Scholar
Hammer, Ø, Hryniewicz, K, Hurum, JH, Høyberget, M, Knutsen, EM and Nakrem, HA (2013) Large onychites (cephalopod hooks) from the Upper Jurassic of the Boreal Realm. Acta Palaeontologica Polonica 58, 827–35. doi: 10.4202/app.2012.0020Google Scholar
Hammes, U, Hamlin, HS and Ewing, TE (2011) Geologic analysis of the Upper Jurassic Haynesville Shale in east Texas and west Louisiana. American Association of Petroleum Geologists Bulletin 95, 1643–66. doi: 10.1306/02141110128CrossRefGoogle Scholar
Hantzpergue, P, Baudin, F, Mitta, V, Olferiev, A and Zakharov, V (1998) The Upper Jurassic of the Volga basin: ammonite biostratigraphy and occurrence of organic-carbon rich facies. Correlations between boreal-subboreal and submediterranean provinces. Mémoires du Muséum national d’histoire naturelle 179, 933.Google Scholar
Hatem, E, Tribovillard, N, Averbuch, O, Bout-Roumazeilles, V, Trentesaux, A, Deconinck, J-F, Baudin, F and Adatte, T (2018) Small-scaled lateral variations of an organic-rich formation in a ramp-type depositional environment (the Late Jurassic of the Boulonnais, France): impact of the clastic supply. Bulletin de la Société Géologique de France, 188, 117. doi: 10.1051/bsgf/2017193Google Scholar
Helleren, S (2019) Lateral compositional variations in the Upper Jurassic source rock in the southwestern Barents Sea – an organic or inorganic disclosure. M.Sc. thesis, University of Stavanger, Norway. Published thesis.CrossRefGoogle Scholar
Herbin, JP, Fernandez-Martinez, JL, Geyssant, JR, Albani, AEl, Deconinck, JF, Proust, JN, Colbeaux, JP and Vidier, JP (1995) Sequence stratigraphy of source rocks applied to the study of the Kimmeridgian/Tithonian in the north-west European shelf (Dorset/UK, Yorkshire/UK and Boulonnais/France). Marine and Petroleum Geology 12, 177–94. doi: 10.1016/0264-8172(95)92838-nCrossRefGoogle Scholar
Houša, V, Pruner, P, Zakharov, VA, Kostak, M, Chadima, M, Rogov, MA, Šlechta, S and Mazuch, M (2007) Boreal–Tethyan correlation of the Jurassic–Cretaceous boundary interval by magneto- and biostratigraphy. Stratigraphy and Geological Correlation 15, 297309. doi: 10.1134/s0869593807030057CrossRefGoogle Scholar
Hovikoski, J, Pedersen, GK, Alsen, P, Lauridsen, BW, Svennevig, K, Nøhr-Hansen, H, Sheldon, E, Dybkjær, K, Bojesen-Koefoed, J, Piasecki, S, Bjerager, M and Ineson, J (2018) The Jurassic–Cretaceous lithostratigraphy of Kilen, Kronprins Christian Land, eastern North Greenland. Bulletin of the Geological Society of Denmark 66, 61112.CrossRefGoogle Scholar
Ilyasov, VS, Staroverov, VN and Vorobyeva, EV (2018) Geochemical characteristics of organic matter in the Upper Jurassic organic-rich shales. Volga and Pricaspian Region Resources 93, 2636 (in Russian).Google Scholar
Ilyina, VI (1985) Jurassic Palynology of Siberia. Moscow: Nauka. Siberian Branch of the Russian Academy of Sciences, Transactions of the Institute of Geology and Geophysics 638, 1237 (in Russian).Google Scholar
Imlay, RW (1981) Late Jurassic ammonites from Alaska. United States Geological Survey Professional Paper 1190, 140. doi: doi.org/10.3133/pp1190Google Scholar
Ineson, JR, Bojesen-Koefoed, JA, Dybkjær, K and Nielsen, LH (2003) Volgian–Ryazanian ‘hot shales’ of the Bo Member (Farsund Formation) in the Danish Central Graben, North Sea: stratigraphy, facies and geochemistry Geological Survey of Denmark and Greenland Bulletin 1, 403–36.CrossRefGoogle Scholar
Jeletzky, JA (1984) Jurassic-Cretaceous Boundary Beds of western and Arctic Canada and the problem of the Tithonian-Berriasian stages in the Boreal Realms. Geological Association of Canada Special Paper 27, 175255.Google Scholar
Jenkyns, HC (1999) Mesozoic anoxic events and palaeoclimate. Zentralblatt für Geologie und Paläontologie 1997, 943–9.Google Scholar
Jenkyns, HC (2010) Geochemistry of oceanic anoxic events. Geochemistry, Geophysics, Geosystems 11, Q03004, doi: 10.1029/2009GC002788CrossRefGoogle Scholar
Kashirtsev, VA, Nikitenko, BL, Peshchevitskaya, EB and Fursenko, EA (2018) Biogeochemistry and microfossils of the Upper Jurassic and Lower Cretaceous, Anabar Bay, Laptev Sea. Russian Geology and Geophysics 59, 386404. doi: 10.1016/j.rgg.2017.09.004CrossRefGoogle Scholar
Kelly, SR, Gregory, FJ, Braham, W, Strogen, DP and Whitham, AG (2015) Towards an integrated Jurassic biostratigraphy for eastern Greenland. Volumina Jurassica 13, 4364.Google Scholar
Khotylev, OV, Balushkina, NA, Vishnevskaya, VS, Korobova, NI, Kalmykov, GA and Roslyakova, AS (2019) A model of the accumulation of radiolarite layers in the Bazhenov Formation of West Siberia. Moscow University Geology Bulletin 74, 206–11. doi: 10.3103/s0145875219020054CrossRefGoogle Scholar
Kietzmann, DA, Ambrosio, AL, Suriano, J, Alonso, MS, Tomassini, FG, Depine, G and Repol, D (2016) The Vaca Muerta–Quintuco system (Tithonian–Valanginian) in the Neuquén Basin, Argentina: a view from the outcrops in the Chos Malal fold and thrust belt. American Association of Petroleum Geologists Bulletin 100, 743–71. doi: 10.1306/02101615121CrossRefGoogle Scholar
Koevoets, MJ, Abay, TB, Hammer, Ø and Olaussen, S (2016) High-resolution organic carbon-isotope stratigraphy of the Middle Jurassic-Lower Cretaceous Agardhfjellet Formation of central Spitsbergen, Svalbard. Palaeogeography, Palaeoclimatology, Palaeoecology 449, 266–74. doi: 10.1016/j.palaeo.2016.02.029CrossRefGoogle Scholar
Koevoets, MJ, Hammer, Ø and Little, CTS (2019) Palaeoecology and palaeoenvironments of the Middle Jurassic to lowermost Cretaceous Agardhfjellet Formation (Bathonian–Ryazanian), Spitsbergen, Svalbard. Norwegian Journal of Geology 99, 124. doi: 10.17850/njg99-1-02Google Scholar
Koevoets, MJ, Hammer, Ø, Olaussen, S, Senger, K and Smelror, M (2018a) Integrating subsurface and outcrop data of the Middle Jurassic to Lower Cretaceous Agardhfjellet Formation in central Spitsbergen. Norwegian Journal of Geology 98, 134. doi: 10.17850/njg98-4-01Google Scholar
Koevoets, MJ, Hurum, HH and Hammer, Ø (2018b) New Late Jurassic teleost remains from the Agardhfjellet Formation, Spitsbergen, Svalbard. Norwegian Journal of Geology 98, 112. doi: 10.17850/njg98-2-01Google Scholar
Kosteva, NN (2005) Stratigraphy of the Jurassic-Cretaceous deposits of Franz-Josef Land archipelago. Arctic and Antarctic 4, 1632 (in Russian).Google Scholar
Kozlova, EV, Fadeeva, NP, Kalmykov, GA, Balushkina, NS, Pronina, NV, Poludetkina, EN, Kostenko, OV, Yurchenko, AY, Borisov, RS, Bychkov, AY and Kalmykov, AG (2015) Geochemical technique of organic matter research in deposits enrich in kerogene (the Bazhenov Formation, West Siberia). Moscow University Geology Bulletin 70, 409–18. doi: 10.3103/s0145875215050075CrossRefGoogle Scholar
Kulyova, GV, Yanochkina, ZA, Bukina, TF, Ivanov, AV, Baryshnikova, VN, Troitskaya, EA and Eryomin, VN (2004) Upper Jurassic shale-bearing section of the Volga Basin (Dorsoplanites panderi zone). Transactions of the Geological Scientific Institute of Saratov State University XVII, 1110 (in Russian).Google Scholar
Langrock, U and Stein, R (2004) Origin of marine petroleum source rocks from the Late Jurassic to Early Cretaceous Norwegian Greenland Seaway - evidence for stagnation and upwelling. Marine and Petroleum Geology 21, 157176. doi: 10.1016/j.marpetgeo.2003.11.011CrossRefGoogle Scholar
Langrock, U, Stein, R, Lipinski, M and Brumsack, HJ (2003) Paleoenvironment and sea-level change in the early Cretaceous Barents Sea - implications from near-shore marine sapropels. Geo-Marine Letters 23, 3442. doi: 10.1007/s00367-003-0122-5CrossRefGoogle Scholar
Leith, TL, Weiss, HM, Mørk, A, Elvebakk, G, Embry, AF, Brooks, PW, Stewart, KR, Pchelina, TM, Bro, EG, Verba, ML and Danyushevskaya, A (1992) Mesozoic hydrocarbon source-rocks of the Arctic region. Norwegian Petroleum Society Special Publications 2, 125. doi: 10.1016/b978-0-444-88943-0.50006-xGoogle Scholar
Lopatin, NV and Yemets, TP (1987) Pyrolysis in the Oil and Gas Geochemistry. Moscow: Nauka, 144 pp.Google Scholar
Lott, GK, Thomas, JE, Riding, JB, Davey, RJ and Butler, N (1989) Late Ryazanian black shales in the southern North Sea basin and their lithostratigraphical significance. Proceedings of the Yorkshire Geological Society 47, 321–24. doi: 10.1144/pygs.47.4.321CrossRefGoogle Scholar
Magoon, LB, Woodward, PV, Banet, AC Jr, Griscom, SB and Daws, TA (1987) Thermal maturity, richness, and type of organic matter of source-rock units. In Petroleum Geology of the Arctic National Wildlife Refuge (eds Bird, KJ and Magoon, LB), pp. 127–79. Northeastern Alaska: United States Geological Survey, Bulletin no. 1778.Google Scholar
Malone, TC (1971) The relative importance of nannoplankton and netplankton as primary producers in tropical oceanic and neritic phytoplankton communities. Limnology and Oceanography 16, 633–39. doi: 10.4319/lo.1971.16.4.0633CrossRefGoogle Scholar
Martinez, M and Dera, G (2015) Orbital pacing of carbon fluxes by a ~9-My eccentricity cycle during the Mesozoic. Proceedings of the National Academy of Sciences 112, 12604–9. doi: 10.1073/pnas.1419946112CrossRefGoogle ScholarPubMed
Miller, RG (1990) A paleoceanographic approach to the Kimmeridge clay formation. In Deposition of Organic Facies (ed. AY Huс), pp. 13–26. Tulsa: American Association of Petroleum Geologists, Studies in Geology no. 30.CrossRefGoogle Scholar
Mohammedyasin, MS, Wudie, G, Anteneh, ZL and Bawoke, GT (2019) Paleoredox conditions of the Middle-Upper Jurassic black shales in the Blue Nile Basin, Ethiopia. Journal of African Earth Sciences 151, 136–45. doi: 10.1016/j.jafrearsci.2018.12.009CrossRefGoogle Scholar
Morales, C, Rogov, M, Wierzbowski, H, Ershova, V, Suan, G, Adatte, T, Föllmi, KB, Tegelaar, E, Reichart, GJ, de Lange, GJ and Middelburg, JJ (2017) Glendonites track methane seepage in Mesozoic polar seas. Geology 45, 503–6. doi: 10.1130/g38967.1CrossRefGoogle Scholar
Morgans-Bell, HS, Coe, AL, Hesselbo, SP, Jenkyns, HC, Weedon, GP, Marshall, JE, Tyson, RV and Williams, CJ (2001) Integrated stratigraphy of the Kimmeridge Clay Formation (Upper Jurassic) based on exposures and boreholes in south Dorset, UK. Geological Magazine 138, 511–39. doi: 10.1017/s0016756801005738CrossRefGoogle Scholar
Mutterlose, J, Bornemann, A and Herrle, JO (2005) Mesozoic calcareous nannofossils - state of the art. Paläontologische Zeitschrift 79, 113–33. doi: 10.1007/BF03021757CrossRefGoogle Scholar
Mutterlose, J, Brumsack, H, Flögel, S, Hay, W, Klein, C, Langrock, U, Lipinski, M, Ricken, W, Söding, E, Stein, R and Swientek, O (2003) The Greenland–Norwegian Seaway: a key area for understanding Late Jurassic to Early Cretaceous paleoenvironments. Paleoceanography and Paleoclimatology 18. doi: 10.1029/2001PA000625Google Scholar
Mutterlose, J and Ruffell, A (1999) Milankovitch-scale palaeoclimate changes in pale–dark bedding rhythms from the Early Cretaceous (Hauterivian and Barremian) of eastern England and northern Germany. Palaeogeography, Palaeoclimatology, Palaeoecology 154, 133–60. doi: 10.1016/s0031-0182(99)00107-8CrossRefGoogle Scholar
Nagy, J, Løfaldli, M and Bäckström, SA (1988) Aspects of foraminiferal distribution and depositional conditions in Middle Jurassic to Early Cretaceous shales in eastern Spitsbergen. Abhandlungen der Geologischen Bundesanstalt 30, 297300.Google Scholar
Nikitenko, BL, Knyazev, VG, Peshchevitskaya, EB and Glinskikh, LA (2015) The upper Jurassic of the Laptev Sea: interregional correlations and paleoenvironments. Russian Geology and Geophysics 56, 1173–93. doi: 10.1016/j.rgg.2015.07.008CrossRefGoogle Scholar
Nikitenko, BL, Pestchevitskaya, EB, Lebedeva, NK and Ilyina, VI (2008) Micropalaeontological and palynological analyses across the Jurassic-Cretaceous boundary on Nordvik Peninsula, Northeast Siberia. Newsletters on Stratigraphy 42, 181222. doi: 10.1127/0078-0421/2008/0042-0181CrossRefGoogle Scholar
Nozaki, T, Kato, Y and Suzuki, K (2013) Late Jurassic ocean anoxic event: evidence from voluminous sulphide deposition and preservation in the Panthalassa. Scientific Reports 3, Article no. 1889. doi: 10.1038/srep01889CrossRefGoogle ScholarPubMed
Nunn, EV, Price, GD, Hart, MB, Page, KN and Leng, MJ (2009) Isotopic signals from Callovian–Kimmeridgian (Middle–Upper Jurassic) belemnites and bulk organic carbon, Staffin Bay, Isle of Skye, Scotland. Journal of the Geological Society 166, 633–41. doi: 10.1144/0016-76492008-067CrossRefGoogle Scholar
Oschmann, W (1988) Upper Kimmeridgian and Portlandian marine macrobenthic associations from southern England and northern France. Facies 18, 4982. doi: 10.1007/bf02536795CrossRefGoogle Scholar
Oschmann, W (1994) Der Kimmeridge Clay von Yorkshire als Beispiel eines fossilen Sauerstoff-kontrollierten Milieus. Berengeria 9, 3153.Google Scholar
Panchenko, IV, Balushkina, NS, Baraboshkin, EYu, Vishnevskaya, VS, Kalmikov, GA and Shurekova, OV (2015) Complexes of paleobiota in Abalak-Bazhenov deposits in the central part of Western Siberia. Neftegazovaya Geologiya: Teoriya i Praktika 10. doi: 353/2070-5379/24_2015Google Scholar
Panchenko, IV, Nemova, VD, Smirnova, ME and Ilyina, MV (2016) Stratification and detailed correlation of Bazhenov horizon in the central part of the Western Siberia according to lithological and paleontological core analysis and well logging. Oil and Gas Geology 6, 2234 (in Russian).Google Scholar
Pauly, S, Mutterlose, J and Alsen, P (2013) Depositional environments of Lower Cretaceous (Ryazanian–Barremian) sediments from Wollaston Forland and Kuhn Ø, North-East Greenland. Bulletin of the Geological Society of Denmark 61, 1936.CrossRefGoogle Scholar
Peters, KE, Kontorovich, AE, Moldowan, JM, Andrusevich, VE, Huizinga, BJ, Demaison, GJ and Stasova, OF (1993) Geochemistry of selected oils and rocks from the central portion of the West Siberian basin, Russia. American Association of Petroleum Geologists Bulletin 77, 863–87. doi: 10.1306/bdff8d80-1718-11d7-8645000102c1865dGoogle Scholar
Ponomareva, EV, Burshtein, LM, Kontorovich, AE and Kostyreva, EA (2018) Organic carbon distribution in the Bazhenov Horizon rocks of the Western Siberian megabasin. Doklady Earth Sciences 481, 918–21. doi: 10.1134/s1028334x18070176CrossRefGoogle Scholar
Ponsaing, L, Mathiesen, A, Petersen, HI, Bojesen-Koefoed, JA, Schovsbo, NH, Nytoft, HP and Stemmerik, L (2020) Organofacies composition of the Upper Jurassic–Lowermost Cretaceous source rocks, Danish central graben, and insight into the correlation to oils in the Valdemar field. Marine and Petroleum Geology 114, 104239. doi: 10.1016/j.marpetgeo.2020.104239CrossRefGoogle Scholar
Price, GD and Rogov, MA (2009) An isotopic appraisal of the Late Jurassic greenhouse phase in the Russian Platform. Palaeogeography, Palaeoclimatology, Palaeoecology 273, 41–9. doi: 10.1016/j.palaeo.2008.11.011CrossRefGoogle Scholar
Proust, JN, Deconinck, JF, Geyssant, JR, Herbin, JP and Vidier, JP (1995) Sequence analytical approach to the Upper Kimmeridgian-Lower Tithonian storm-dominated ramp deposits of the Boulonnais (Northern France). A landward time-equivalent to offshore marine source rocks. Geologische Rundschau 84, 255–71. doi: 10.1007/bf00260439CrossRefGoogle Scholar
Rakociński, M, Zatoń, M, Marynowski, L, Gedl, P and Lehmann, J (2018) Redox conditions, productivity, and volcanic input during deposition of uppermost Jurassic and Lower Cretaceous organic-rich siltstones in Spitsbergen, Norway. Cretaceous Research 89, 126–47. doi: 10.1016/j.cretres.2018.10.014CrossRefGoogle Scholar
Rawson, PF (1973) Lower Cretaceous (Ryazanian-Barremian) marine connections and cephalopod migrations between the Tethyan and Boreal Realms. Geological Journal, Special Issue, 5, 131–44.Google Scholar
Rees, PM, Ziegler, AM and Valdes, PJ (2000) Jurassic phytogeography and climates: new data and model comparisons. In Warm Climates in Earth History (eds Huber, BT, MacLeod, KG and Wing, SL), pp. 297318. Cambridge: Cambridge University Press.Google Scholar
Remane, J (1991) The Jurassic-Cretaceous boundary: problems of definition and procedure. Cretaceous Research 12, 447–53. doi: 10.1016/0195-6671(91)90001-sCrossRefGoogle Scholar
Remírez, MN and Algeo, TJ (2020) Paleosalinity determination in ancient epicontinental seas: a case study of the T-OAE in the Cleveland Basin (UK). Earth-Science Reviews 201, 103072. doi: 10.1016/j.earscirev.2019.103072CrossRefGoogle Scholar
Riboulleau, A, Derenne, S, Largeau, C and Baudin, F (2001) Origin of contrasting features and preservation pathways in kerogens from the Kashpir oil shales (Upper Jurassic, Russian Platform). Organic Geochemistry 32, 647–65. doi: 10.1016/s0146-6380(01)00017-1CrossRefGoogle Scholar
Rogov, MA (2004) The Russian Platform as a key region for Volgian/Tithonian correlation: a review of the Mediterranean faunal elements and ammonite biostratigraphy of the Volgian stage. Rivista Italiana di Paleontologia e Stratigrafia 110, 321–8.Google Scholar
Rogov, MA (2010) A precise ammonite biostratigraphy through the Kimmeridgian-Volgian boundary beds in the Gorodischi section (Middle Volga area, Russia), and the base of the Volgian Stage in its type area. Volumina Jurassica VIII, 103–30.Google Scholar
Rogov, MA (2013) Ammonites and infrazonal subdivision of the Dorsoplanites panderi Zone (Volgian Stage, Upper Jurassic) of the European part of Russia. Doklady Earth Sciences 451, 803–8. doi: 10.1134/s1028334x13080059CrossRefGoogle Scholar
Rogov, MA (2014) Infrazonal subdivision of the Volgian Stage in its type area using ammonites and correlation of the Volgian and Tithonian Stages. In STRATI 2013. First International Congress on Stratigraphy. At the Cutting Edge of Stratigraphy, pp. 577–580. Springer Geology. doi: 10.1007/978-3-319-04364-7_111CrossRefGoogle Scholar
Rogov, MA, Baraboshkin, EY, Guzhikov, AY, Efimov, VM, Kiselev, DN, Morov, VP and Gusev, VV (2015) The Jurassic-Cretaceous boundary in the Middle Volga region. Field guide to the International meeting on the Jurassic/Cretaceous boundary. September 7-13, 2015, Samara (Russia). Samara: Samara State Technical University, 130 pp.Google Scholar
Rogov, M and Bizikov, V (2006) New data on Middle Jurassic-Lower Cretaceous Belemnoteuthidae from Russia. What can shell tell about the animal and its mode of life. Acta Universitatis Carolinae Geologica 49, 149–63.Google Scholar
Rogov, MA and Ustinova, MA (2018) High-latitude Late Jurassic nannofossils and their implication for climate and palaeogeography. Norwegian Journal of Geology 98, 1723. doi: 10.17850/njg98-1-02Google Scholar
Rogov, M and Zakharov, V (2009) Ammonite- and bivalve-based biostratigraphy and Panboreal correlation of the Volgian Stage. Science in China Series D, Earth Sciences 52, 1890–909. doi: 10.1007/s11430-009-0182-0CrossRefGoogle Scholar
Rogov, MA, Zakharov, VA and Ershova, VB (2011) Detailed stratigraphy of the Jurassic–Cretaceous boundary beds of the Lena River lower reaches based on ammonites and buchiids. Stratigraphy and Geological Correlation 19, 641–62. doi: 10.1134/s0869593811060050CrossRefGoogle Scholar
Rogov, MA, Zverkov, NG, Zakharov, VA and Arkhangelsky, MS (2019) Marine reptiles and climates of the Jurassic and Cretaceous of Siberia. Stratigraphy and Geological Correlation 27, 398423. doi: 10.31857/s0869-592x27413-39CrossRefGoogle Scholar
Rost, B and Riebesell, U (2004) Coccolithophores and the biological pump: responses to environmental changes. In Coccolithophores: From Molecular Processes to Global Impact (eds Thierstein, HR and Young, JR), pp. 99125. Berlin, Heidelberg: Springer.CrossRefGoogle Scholar
Ruffell, AH, Price, GD, Mutterlose, J, Kessels, K, Baraboshkin, E and Gröcke, DR (2002) Palaeoclimate indicators (clay minerals, calcareous nannofossils, stable isotopes) compared from two successions in the late Jurassic of the Volga Basin (SE Russia). Geological Journal 37, 1733. doi: 10.1002/gj.903CrossRefGoogle Scholar
Ryzhkova, SV, Burshtein, LM, Ershov, SV, Kazanenkov, VA, Kontorovich, AE, Kontorovich, VA, Nekhaev, AY, Nikitenko, BL, Fomin, MA, Shurygin, BN, Beizel, AL, Borisov, EV, Zolotova, OV, Kalinina, LM and Ponomareva, EV (2018) The Bazhenov Horizon of West Siberia: structure, correlation, and thickness. Russian Geology and Geophysics 59, 846–63. doi: 10.1016/j.rgg.2018.07.009CrossRefGoogle Scholar
Sælen, G, Tyson, RV, Telnæs, N and Talbot, MR (2000) Contrasting watermass conditions during deposition of the Whitby Mudstone (Lower Jurassic) and Kimmeridge Clay (Upper Jurassic) formations, UK. Palaeogeography, Palaeoclimatology, Palaeoecology 163, 163–96. doi: 10.1016/s0031-0182(00)00150-4CrossRefGoogle Scholar
Samson, Y, Lepage, G, Hantzpergue, P, Guyader, J, Saint-Germès, M, Baudin, F and Bignot, G (1996) Révision lithostratigraphique et biostratigraphique du Kimméridgien de la région havraise (Normandie). Géologie de la France 3, 319.Google Scholar
Savrda, CE and Bottjer, DJ (1986) Trace fossil model for reconstruction of paleo-oxygenation in bottom water. Geology 14, 36. doi: 10.1130/0091-7613(1986)14<3:tmfrop>2.0.co;22.0.CO;2>CrossRefGoogle Scholar
Scotchman, IC (1991) Kerogen facies and maturity of the Kimmeridge Clay Formation in southern and eastern England. Marine and Petroleum Geology 8, 278–95. doi: 10.1016/0264-8172(91)90082-cCrossRefGoogle Scholar
Shchepetova, EV (2009) Sedimentology of Volgian oil shale formation (Upper Jurassic, Panderi Zone) in North Russian Plate. Bulletin of Moscow Society of Naturalists, Geological Series 84, 7489 (in Russian).Google Scholar
Shchepetova, EV and Rogov, MA (2013) Organic carbon-rich horizons in the Upper Kimmeridgian of the northern part of Ulyanovsk-Saratov trough (Russian Platform): biostratigraphy, sedimentology, geochemistry. In Jurassic System of Russia: Problems of Stratigraphy and Рaleogeography, Fifth All-Russian Meeting (eds VA Zakharov, MA Rogov and BN Shurygin), pp. 249–252. 23–27 September 2013, Tyumen. Scientific Materials. Yekaterinburg, ID IzdatNaukaServis LLC (in Russian).Google Scholar
Shchepetova, EV and Rogov, MA (2016) Organic carbon-rich shales within coarse-grained lithofacies of Jurassic–Cretaceous transition at the Russian Platform. In Field Trip Guide and Abstracts Book, XIIth Jurassica Conference. IGCP 632 and ICS Berriasian Workshop, pp. 95–96. 19–23 April 2016, Smolenice, Slovakia. Bratislava, Earth Science Institute, Slovak Academy of Sciences.Google Scholar
Shurygin, BN and Dzyuba, OS (2015) The Jurassic/Cretaceous boundary in northern Siberia and Boreal-Tethyan correlation of the boundary beds. Russian Geology and Geophysics 56, 652–62. doi: 10.1016/j.rgg.2015.03.013CrossRefGoogle Scholar
Smelror, M, Mørk, A, Monteil, E, Rutledge, D and Leereveld, H (1998) The Klippfisk Formation - a new lithostratigraphic unit of Lower Cretaceous platform carbonates on the Western Barents Shelf. Polar Research 17, 181202. doi: 10.1111/j.1751-8369.1998.tb00271.xGoogle Scholar
Smelror, M, Mørk, A, Mørk, MBE, Weiss, HM and Løseth, H (2001) Middle Jurassic-Lower Cretaceous transgressive-regressive sequences and facies distribution off northern Nordland and Troms, Norway. Norwegian Petroleum Society Special Publications 10, 211–32. doi: 10.1016/s0928-8937(01)80015-1CrossRefGoogle Scholar
Socha, K and Makos, M (2016) Revealing what has been overlooked – petroleum potential of the Jurassic deposits in Central Poland. In Field Trip Guide and Abstracts Book, XIIth Jurassica Conference. IGCP 632 and ICS Berriasian Workshop, pp. 84–85. 19–23 April 2016, Smolenice, Slovakia. Bratislava, Earth Science Institute, Slovak Academy of Sciences.Google Scholar
Stemmerik, L, Dam, G, Noe-Nygaard, N, Piasecki, S and Surlyk, F (1998) Sequence stratigraphy of source and reservoir rocks in the Upper Permian and Jurassic of Jameson Land, East Greenland. Geology of Greenland Survey Bulletin 180, 4354.CrossRefGoogle Scholar
Strachoff, NM (1934) Brennschiefer der Zone Perisphinctes Panderi d’Orb. (Lithologische Übersicht). Bulletin of Moscow Society of Naturalists, Geological Series XII, 200–50 (in Russian, with abstract in German).Google Scholar
Suchéras-Marx, B, Mattioli, E, Allemand, P, Giraud, F, Pittet, B, Plancq, J and Escarguel, G (2019) The colonization of the oceans by calcifying pelagic algae. Biogeosciences 16, 2501–10. doi: 10.5194/bg-2018-493CrossRefGoogle Scholar
Tobia, FH, Al-Jaleel, HS and Ahmad, IN (2019) Provenance and depositional environment of the Middle-Late Jurassic shales, northern Iraq. Geosciences Journal 23, 747–65. doi: 10.1007/s12303-018-0072-6CrossRefGoogle Scholar
Trabucho-Alexandre, J, Hay, WW and De Boer, PL (2012) Phanerozoic environments of black shale deposition and the Wilson Cycle. Solid Earth 3, 2942. doi: 10.5194/se-3-29-2012CrossRefGoogle Scholar
Tribovillard, N, Bialkowski, A, Tyson, RV, Lallier-Vergès, E, and Deconinck, JF (2001) Organic facies variation in the late Kimmeridgian of the Boulonnais area (northernmost France). Marine and Petroleum Geology 18, 371–89. doi: 10.1016/s0264-8172(01)00006-xCrossRefGoogle Scholar
Tribovillard, N, Ramdani, A and Trentesaux, A (2005) Controls on organic accumulation in Upper Jurassic shales of northwestern Europe as inferred from trace-metal geochemistry. In The Deposition of Organic-Carbon-Rich Sediments: Models, Mechanisms, and Consequences (ed. N Harris), pp. 145–64. Tulsa: Society for Sedimentary Geology, Special Publication no. 82.Google Scholar
Turner, H, Batenburg, SJ, Gale, A and Gradstein, F (2019) The Kimmeridge Clay Formation (Upper Jurassic–Lower Cretaceous) of the Norwegian Continental Shelf and Dorset, UK: a chemostratigraphic correlation. Newsletters on Stratigraphy 52, 132. doi: 10.1127/nos/2018/0436CrossRefGoogle Scholar
Turov, AV (2000) On environmental conditions of the Upper Jurassic oil shales in the Russian Plate. Proceedings of Higher Educational Establishments: Geology and Exploration 3, 920 (in Russian).Google Scholar
Tyson, RV (1987) The genesis and palynofacies characteristics of marine petroleum source rocks. In Marine Petroleum Source Rocks (eds Brooks, J and Fleet, AJ), pp. 4768. Geological Society of London, Special Publication no. 26. doi: 10.1144/gsl.sp.1987.026.01.03Google Scholar
Tyson, RV (1995) Sedimentary Organic Matter: Organic Facies and Paliynofacies. London: Chapman and Hall, 615 pp.CrossRefGoogle Scholar
Ulmishek, GF (1993) Petroleum Geology and Resources of the West Siberian Basin, Russia. US Geological Survey Bulletin 2201-G, 153. doi: 10.3133/b2201gGoogle Scholar
Underhill, JR (1998) Jurassic. In Petroleum Geology of the North Sea, 4th edition (ed. Glennie, KW), pp. 245–93. London: Blackwell Science Ltd.CrossRefGoogle Scholar
Vakhrameyev, VA (1982) Classopollis pollen as an indicator of Jurassic and Cretaceous climate. International Geology Review, 24, 1190–6. doi: 10.1080/00206818209451058CrossRefGoogle Scholar
Vishnevskaya, VS, de Wever, P, Baraboshkin, EY, Bogdanov, NA, Bragin, NY, Bragina, LG, Kostyuchenko, AS, Lambert, E, Malinovsky, YM, Sedaeva, KM and Zukova, GA (1999) New stratigraphic and palaeogeographic data on Upper Jurassic to Cretaceous deposits from the eastern periphery of the Russian Platform (Russia). Geodiversitas 21, 347–63.Google Scholar
Vollset, J and Doré, AG (1984) A revised Triassic and Jurassic lithostratigraphic nomenclature for the Norwegian North Sea. Norwegian Petroleum Directorate Bulletin 3, 153.Google Scholar
Więcław, D (2016) Habitat and hydrocarbon potential of the Kimmeridgian strata in the centralpart of the Polish Lowlands. Geological Quarterly 60, 192210. doi: 10.7306/gq.1260CrossRefGoogle Scholar
Wierzbowski, A, Hryniewicz, K, Hammer, Ø, Nakrem, HA and Little, CTS (2011) Ammonites from hydrocarbon seep carbonate bodies from the uppermost Jurassic – lowermost Cretaceous of Spitsbergen and their biostratigraphical importance. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 262, 267–88. doi: 10.1127/0077-7749/2011/0198CrossRefGoogle Scholar
Wierzbowski, A and Smelror, M (2020) The Bajocian to Kimmeridgian (Middle to Upper Jurassic) ammonite succession at Sentralbanken High (core 7533/3-U-1), Barents Sea, and its stratigraphical and palaeobiogeographical significance. Volumina Jurassica XVIII, 122.CrossRefGoogle Scholar
Wierzbowski, A, Smelror, M and Mørk, A (2002) Ammonites and dinoflagellate cysts in the Upper Oxfordian and Kimmeridgian of the northeastern Norwegian Sea (Nordland VII offshore area): biostratigraphical and biogeographical significance. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 226, 145–64. doi: 10.1127/njgpa/226/2002/145CrossRefGoogle Scholar
Wierzbowski, A and Wierzbowski, H (2019) Ammonite stratigraphy and organic matter of the Pałuki Fm. (Upper Kimmeridgian-Lower Tithonian) from the central-eastern part of the Łódź Synclinorium (Central Poland). Volumina Jurassica XVII, 4980. doi: 10.7306/VJ.17.4Google Scholar
Wierzbowski, H, Bajnai, D, Wacker, U, Rogov, MA, Fiebig, J and Tesakova, EM (2018) Clumped isotope record of salinity variations in the Subboreal Province at the Middle–Late Jurassic transition. Global and Planetary Change 167, 172–89. doi: 10.1016/j.gloplacha.2018.05.014CrossRefGoogle Scholar
Wignall, PB (1990) Benthic palaeoecology of the Late Jurassic Kimmeridge Clay of England. Special Papers in Palaeontology 43, 174.Google Scholar
Wimbledon, WA (2008) The Jurassic-Cretaceous boundary: an age-old correlative enigma. Episodes 31, 423–8. doi: 10.18814/epiiugs/2008/v31i4/008CrossRefGoogle Scholar
Yang, R, Cao, J, Hu, G, Bian, L, Hu, K and Fu, X (2017) Marine to brackish depositional environments of the Jurassic–Cretaceous Suowa Formation, Qiangtang Basin (Tibet), China. Palaeogeography, Palaeoclimatology, Palaeoecology 473, 4156. doi: 10.1016/j.palaeo.2017.02.031CrossRefGoogle Scholar
Zakharov, VA (2006) Tithonian–Berriasian deposition environments of the Bazhenov Formation bituminous shale in West Siberia, from paleoecological data. In Evolution of Biosphere and Biodiversity, on the 70th Anniversary of Yu.A. Rosanov (eds Leonova, TB, Lopatin, AV, Rozhnov, SV, Ushatinskaya, GT and Shevyrev, AA). Moscow: KMK Society of Scientific Publications, pp. 552568 (in Russian).Google Scholar
Zakharov, VA, Baudin, F, Dzyuba, OS, Daux, V, Zverev, KV and Renard, M (2005) Isotopic and faunal record of high paleotemperatures in the Kimmeridgian of Subpolar Urals. Russian Geology and Geophysics 46, 320.Google Scholar
Zakharov, VA, Nalnyaeva, TI and Shulgina, NI (1983) New data on the biostratigraphy of the Upper Jurassic and Lower Cretaceous deposits on Paksa peninsula, Anabar embayment (north of the Middle Siberia). Moscow: Nauka. Siberian Branch of the Russian Academy of Sciences, Transactions of the Institute of Geology and Geophysics, 528, 5699 (in Russian).Google Scholar
Zakharov, VA and Rogov, MA (2003) Boreal-Tethyan mollusk migrations at the Jurassic-Cretaceous boundary time and biogeographic ecotone position in the Northern Hemisphere. Stratigraphy and Geological Correlation 11, 152–71.Google Scholar
Zakharov, VA, Rogov, MA, Dzyuba, OS, Žák, K, Košt’ák, M, Pruner, P, Skupien, P, Chadima, M, Mazuch, M and Nikitenko, BL (2014) Palaeoenvironments and palaeoceanography changes across the Jurassic/Cretaceous boundary in the Arctic realm: case study of the Nordvik section (north Siberia, Russia). Polar Research 33, 19714. doi: 10.3402/polar.v33.19714CrossRefGoogle Scholar
Zakharov, VA, Rogov, MA and Shchepetova, ЕV (2017) Black shale events in the Late Jurassic – earliest Cretaceous of Central Russia. In Jurassic System of Russia: Problems of Stratigraphy and Paleogeography (eds Zakharov, VA, Rogov, MA and Shchepetova, EV), pp. 5763. Proceedings of 7th All-Russian Conference, 18–22 September 2017. Scientific Materials. Moscow: Siberian Branch of the Russian Academy of Sciences, Transactions of the Institute of Geology and Geophysics (in Russian).Google Scholar
Zakharov, VA and Yudovny, EG (1974) Conditions of sediment deposition and fauna existence in the Early Cretaceous sea of Khatanga depression. Novosibirsk: Nauka. Siberian Branch of the Russian Academy of Sciences, Transactions of the Institute of Geology and Geophysics 80, 127–74 (in Russian).Google Scholar
Zanin, YN, Zamirailova, AG and Eder, VG (2012) Some calcareous nannofossils from the Upper Jurassic-Lower Cretaceous Bazhenov Formation of the West Siberian marine basin, Russia. The Open Geology Journal 6, 2531. doi: 10.2174/1874262901206010025CrossRefGoogle Scholar
Ziegs, V, Horsfield, B, Skeie, JE and Rinna, J (2017) Petroleum retention in the Mandal Formation, Central Graben, Norway. Marine and Petroleum Geology 83, 195214. doi: 10.1016/j.marpetgeo.2017.03.005CrossRefGoogle Scholar
Zverkov, NG and Efimov, VM (2019) Revision of Undorosaurus, a mysterious Late Jurassic ichthyosaur of the Boreal Realm. Journal of Systematic Palaeontology 17, 963–93. doi: 10.1080/14772019.2018.1515793CrossRefGoogle Scholar
Supplementary material: PDF

Rogov et al. supplementary material

Rogov et al. supplementary material

Download Rogov et al. supplementary material(PDF)
PDF 158.1 KB