Hostname: page-component-7c8c6479df-xxrs7 Total loading time: 0 Render date: 2024-03-17T20:46:47.475Z Has data issue: false hasContentIssue false

Ordovician reef and mound evolution: the Baltoscandian picture

Published online by Cambridge University Press:  08 June 2016

BJÖRN KRÖGER*
Affiliation:
Finnish Museum of Natural History, PO Box 44, Fi-00014 University of Helsinki, Finland
LINDA HINTS
Affiliation:
Institute of Geology at Tallinn University of Technology, Ehitajate tee 5, 19086, Tallinn, Estonia
OLIVER LEHNERT
Affiliation:
Institute of Geology at Tallinn University of Technology, Ehitajate tee 5, 19086, Tallinn, Estonia Geo-Center of Northern Bavaria, Lithosphere Dynamics, Friedrich-Alexander University of Erlangen-Nürnberg, Schlossgarten 5, D-91054 Erlangen, Germany Department of Geology, Lund University, Sölvegatan 12, SE-223 62 Lund, Sweden
*
Author for correspondence: bjorn.kroger@helsinki.fi

Abstract

The widespread growth of reefs formed by a framework of biogenic constructors and frame-lacking carbonate mounds began on Baltica during Ordovician time. Previously, Ordovician reef and mound development on Baltica was considered to be sporadic and local. A review of all known bioherm localities across the Baltic Basin reveals a more consistent pattern. Ordovician bioherms grew in a wide E–W-aligned belt across the Baltic Basin and occur in several places in Norway. Substantial reef development began simultaneously across the region during the late Sandbian – early Katian interval and climaxed during the late Katian Pirgu age. The current spatiotemporal distribution of bioherms is a result of interdependent factors that involve original drivers of reef development such as relative sea level, climate during the time of deposition and effects of post-depositional erosion. Oceanographic conditions were likely more favourable during times of cooler global climates, low sea level and glacial episodes. At the same time, the likelihood that bioherms are preserved from long-term erosion is higher when deposited during low sea level in deeper parts of the basin. A main factor controlling the timing of the reef and mound evolution was Baltica's shift toward palaeotropical latitudes during Late Ordovician time. The time equivalence between initial reef growth and the Guttenberg isotope carbon excursion (GICE) suggests that global climatic conditions were important.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2016 

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

Ainsaar, L., Kaljo, D., Martma, T., Meidla, T., Männik, P., Nõlvak, J. & Tinn, O. 2010. Middle and Upper Ordovician carbon isotope chemostratigraphy in Baltoscandia: a correlation standard and clues to environmental history. Palaeogeography, Palaeoclimatology, Palaeoecology 294, 189201.CrossRefGoogle Scholar
Ainsaar, L., Meidla, T. & Martma, T. 1999. Evidence for a widespread carbon isotopic event associated with late Middle Ordovician sedimentological and faunal changes in Estonia. Geological Magazine 136, 4962.Google Scholar
Ainsaar, L., Meidla, T. & Martma, T. 2004. The Middle Caradoc Facies and Faunal Turnover in the Late Ordovician Baltoscandian palaeobasin. Palaeogeography, Palaeoclimatology, Palaeoecology 210, 119–33.CrossRefGoogle Scholar
Alberstad, L. P. & Walker, K. R. 1976. A receptaculitid–echinoderm pioneer community in a Middle Ordovician reef. Lethaia 9, 261–72.Google Scholar
Alikhova, T. N. 1953. Rukovodyashchaya fauna brakhiopod Ordovikskikh otloshenij Severo-Zapadnoj chasti Russkoj Platformy. Trudy Vsesoyuznogo Nauchno-Issledovatélskogo Geologicheskogo Instituta (VSEGEI) Ministerstva Geologii i Okhrany 1953, 1127.Google Scholar
Antoshkina, A. I. 1996. Ordovician reefs of the Ural Mountains, Russia: A review. Facies 35, 17.CrossRefGoogle Scholar
Antoshkina, A. I. 1998. Organic buildups and reefs on the Palaeozoic carbonate platform margin, Pechora Urals, Russia. Sedimentary Geology 118, 187211.CrossRefGoogle Scholar
Artyushkov, E. V., Tesakov, Y. & Chekhovich, P. A. 2008. Ordovician sea-level change and rapid change in crustal subsidence rates in East Siberia and Baltoscandia. Russian Geology and Geophysics 49, 633–47.Google Scholar
Axberg, S. 1980. Seismic stratigraphy and bedrock geology of the Bothnian Sea, Northern Baltic. Stockholm Contributions in Geology 36, 153213.Google Scholar
Bathurst, R. G. C. 1982. Genesis of stromatactis cavities between submarine crusts in Palaeozoic carbonate mud buildups. Journal of the Geological Society 139, 165–81.Google Scholar
Bauert, H., Isozaki, Y., Holmer, L. E., Aoki, K., Sakata, S. & Hirata, T. 2014. New U–Pb zircon ages of the Sandbian (Upper Ordovician) ‘Big K-bentonite’ in Baltoscandia (Estonia and Sweden) by LA-ICPMS. GFF 136, 14.Google Scholar
Bergström, S. M., Calner, M., Lehnert, O. & Noor, A. 2011a. A new upper Middle Ordovician–Lower Silurian drillcore standard succession from Borenshult in Östergötland, southern Sweden: 1. Stratigraphical review with regional comparisons. GFF 133, 149–71.CrossRefGoogle Scholar
Bergström, S.M., Chen, X., Gutiérrez-Marco, J.C. & Dronov, A. 2009. The new chronostratigraphic classification of the Ordovician System and its relation to major regional series and stages and to ∂13C chemostratigraphy. Lethaia 42, 97107.CrossRefGoogle Scholar
Bergström, S. M., Huff, W. D., Kolata, D. R., Yost, D. A. & Hart, C. 1997. A unique middle Ordovician K-bentonite bed succession at Röstånga, S. Sweden. GFF 119, 231–44.Google Scholar
Bergström, S. M., Lehnert, O., Calner, M. & Joachimski, M. M. 2012. A new upper Middle Ordovician–Lower Silurian drillcore standard succession from Borenshult in Östergötland, southern Sweden: 2. Significance of δ13C chemostratigraphy. GFF 134, 3963.Google Scholar
Bergström, S. M., Löfgren, A. & Grahn, Y. 2004. The stratigraphy of the Upper Ordovician carbonate mounds in the subsurface of Gotland. GFF 126, 289–96.Google Scholar
Bergström, S. M., Saltzman, M. R., Leslie, S. A., Ferretti, A. & Young, S. A. 2015. Trans-Atlantic application of the Baltic Middle and Upper Ordovician carbon isotope zonation. Estonian Journal of Earth Sciences 64, 812.Google Scholar
Bergström, S. M., Schmitz, B., Saltzman, M. R. & Huff, W. D. 2010a. The Upper Ordovician Guttenberg δ13C excursion (GICE) in North America and Baltoscandia: Occurrence, chronostratigraphic significance, and paleoenvironmental relationships. Geological Society of America Special Papers 466, 3767.Google Scholar
Bergström, S. M., Schmitz, B., Young, S. A. & Bruton, D. L. 2010b. The δ13C chemostratigraphy of the Upper Ordovician Mjøsa Formation at Furuberget near Hamar, southeastern Norway: Baltic, trans-Atlantic, and Chinese relations. Norsk Geologisk Tidsskrift 90, 6578.Google Scholar
Bergström, S. M., Schmitz, B., Young, S. A. & Bruton, D. L. 2011b. Lower Katian (Upper Ordovician) δ13C chemostratigraphy, global correlation and sea-level changes in Baltoscandia. GFF 133, 3147.Google Scholar
Bergström, S. M., Young, S. & Schmitz, B. 2010. Katian (Upper Ordovician) δ13C chemostratigraphy and sequence stratigraphy in the United States and Baltoscandia: A regional comparison. Palaeogeography, Palaeoclimatology, Palaeoecology 296, 217–34.CrossRefGoogle Scholar
Bockelie, J. F. 1978. The Oslo region during the early Palaeozoic. In Tectonics and Geophysics of Continntal Rifts (eds Ramberg, I. B. & Neumann, E. R.), pp. 195202. Dordrecht, Holland: Reidel Publishing.Google Scholar
Braithwaite, C. J. R., Owen, A. W. & Heath, R. A. 1995. Sedimentological changes across the Ordovician–Silurian boundary in Hadeland and their implications for regional patterns of deposition in the Oslo Region. Norsk Geologisk Tidsskrift 75, 199218.Google Scholar
Calner, M., Jeppsson, L. & Munnecke, A. 2004. The Silurian of Gotland–Part I: Review of the stratigraphic framework, event stratigraphy, and stable carbon and oxygen isotope development. Erlanger Geologische Abhandlungen, Sonderband 5, 113–31.Google Scholar
Calner, M., Lehnert, O. & Joachimski, M. 2010. Carbonate mud mounds, conglomerates, and sea-level history in the Katian (Upper Ordovician) of central Sweden. Facies 56, 157–72.Google Scholar
Calner, M., Lehnert, O. & Nõlvak, J. 2010. Palaeokarst evidence for widespread regression and subaerial exposure in the middle Katian (Upper Ordovician) of Baltoscandia: significance for global climate. Palaeogeography, Palaeoclimatology, Palaeoecology 296, 235–47.CrossRefGoogle Scholar
Cocks, R. L. M. & Torsvik, T. H. 2005. Baltica from the late Precambrian to mid-Palaeozoic times: the gain and loss of a terrane's identity. Earth-Science Reviews 72, 3966.Google Scholar
Cooper, R. A., Sadler, P. M., Hammer, O. & Gradstein, F. M. 2012. Chapter 20: The Ordovician Period. In The Geologic Time Scale 2012 (eds Gradstein, F. M., Schmitz, J. G. O. D. & Ogg, G. M.), pp. 489523. Boston: Elsevier.CrossRefGoogle Scholar
Cuffey, R. J. 2006. Bryozoan-built reef mounds: the overview from integrating recent studies with previous investigations. Courier Forschungsinstitut Senckenberg 257, 3547.Google Scholar
Cuffey, R. J., Chuantao, X., Zhu, Z., Spjeldnaes, N. & Hu, Z.-X. 2013. The world's oldest-known Bryozoan reefs: late Tremadocian, mid-Early Ordovician; Yichang, Central China. In Bryozoan Studies 2010 (eds Ernst, A., Schäfer, P., & Scholz, J.), pp. 1327. New York: Springer.Google Scholar
Dronov, A. 2005. Aspects of sedimentation. In Guidebook of the Pre-Conference Field Trip (eds Dronov, A., Tolmacheva, T., Raevskaya, E. & Nestell, M.), pp. 912. 6th Baltic Stratigraphical Conference, IGCP 503 Meeting, August 23–25, 2005, St Petersburg.Google Scholar
Dronov, A., Ainsaar, L., Kaljo, D., Meidla, T., Saadre, T. & Einasto, R. 2011. Ordovician of Baltoscandia: facies, sequences and sea level changes. Ordovician of the World: 11th International Symposium on the Ordovician System, Cuadernos del Museo Geominero 14, 143–50.Google Scholar
Dronov, A. & Dolgov, O. 2005. Stop 13 ‘Kegel’ dolomites of the Elisavetino quarry. In Guidebook of the Pre-Conference Field Trip (eds Dronov, A., Tolmacheva, T., Raevskaya, E. & Nestell, M.), pp. 58–9. 6th Baltic Stratigraphical Conference, Cambrian and Ordovician of St Petersburg Region, IGCP 503 Meeting, August 23–25, 2005, St Petersburg.Google Scholar
Dronov, A. & Rozhnov, S. V. 2007. Climatic changes in the Baltoscandian basin during the Ordovician: sedimentological and palaeontological aspects. Acta Palaeontologica Sinica 46, 108–13.Google Scholar
Ebbestad, J. O. R. & Högström, A. E. S. 2007. Ordovician of the Siljan District, Sweden. WOGOGOB 2007, 9th meeting of the Working Group on Ordovician Geology of Baltoscandia, Field Guide and Abstracts, Rapporter och Meddelanden 128, 52–8.Google Scholar
Ebbestad, J. O. R., Högström, A. E. S., Frisk, Å. M., Martma, T., Kaljo, D., Kröger, B. & Pärnaste, H. 2015. Terminal Ordovician stratigraphy of the Siljan district, Sweden. GFF 137, 3556.Google Scholar
Eriksson, M. E. & Hints, O. 2009. Vagrant benthos (Annelida; Polychaeta) associated with Upper Ordovician carbonate mud-mounds of subsurface Gotland, Sweden. Geological Magazine 146, 451–62.Google Scholar
Ernst, G., Niebuhr, B., Wiese, F. & Wilmsen, M. 1996. Facies development, basin dynamics, event correlation and sedimentary cycles in theUpper Cretaceous of selected areas of Germany and Spain. In Global and Regional Controls on Biogenic Sedimentation. II. Cretaceous Sedimentation (eds Reitner, J., Neuweiler, F. & Gunkel, F.), pp. 87100. Göttingen: University of Göttingen.Google Scholar
Ettensohn, F. R. 2010. Origin of Late Ordovician (mid-Mohawkian) temperate-water conditions on southeastern Laurentia: Glacial or tectonic? Geological Society of America Special Papers 466, 163–75.Google Scholar
Fedorov, P. V. 2003. Lower Ordovician mud mounds from the St. Petersburg region, northwestern Russia. Bulletin of the Geological Society of Denmark 50, 125–37.Google Scholar
Fiske, I. & Chandler, R. B. 2011. unmarked: an R package for fitting hierarchical models of wildlife occurrence and abundance. Journal of Statistical Software 43, 123.Google Scholar
Flodén, T. 1980. Seismic stratigraphy and bedrock geology of the central Baltic. Stockholm Contributions in Geology 35, 1240.Google Scholar
Flodén, T., Bjerkéus, M., Tuuling, I. & Eriksson, M. 2001. A Silurian reefal succession in the Gotland area, Baltic Sea. GFF 123, 137–52.Google Scholar
Fortey, R. A. & Cocks, L. R. M. 2005. Late Ordovician global warming–The Boda event. Geology 33, 405–8.Google Scholar
Frisk, Å. M. & Harper, D. A. T. 2013. Late Ordovician brachiopod distribution and ecospace partitioning in the Tvären crater system, Sweden. Palaeogeography, Palaeoclimatology, Palaeoecology 369, 114–24.Google Scholar
Grahn, Y. & Nõlvak, J. 2007. Ordovician chitinozoa and biostratigraphy from Skåne and Bornholm, southernmost Scandinavia – an overview and update. Bulletin of Geosciences 82, 1126.CrossRefGoogle Scholar
Hammer, Ø. 2003. Biodiversity curves for the Ordovician of Baltoscandia. Lethaia 36, 305–14.Google Scholar
Hanken, N.-M. & Owen, A. W. 1982. The Upper Ordovician (Ashgill) of Ringerike. In Field Excursion Guide. IVth International Symposium on the Ordovician System (eds Bruton, D. L. & Williams, S. H.), pp. 122–31. Oslo: University of Oslo.Google Scholar
Hansen, T. 2011. Mid to Late Ordovician trilobite Palaeoecology in a mud dominated epicontinental sea, Southern Norway. Ordovician of the World: 11th International Symposium on the Ordovician System, Cuadernos del Museo Geominero 14, 157–65.Google Scholar
Hansen, J. & Harper, D. A. T. 2008. The late Sandbian – earliest Katian (Ordovician) brachiopod immigration and its influence on the brachiopod fauna in the Oslo Region, Norway. Lethaia 41, 2535.Google Scholar
Hansen, J., Kresten Nielsen, J. & Hanken, N.-M. 2009. The relationships between Late Ordovician sea-level changes and faunal turnover in western Baltica: geochemical evidence of oxic and dysoxic bottom-water conditions. Palaeogeography, Palaeoclimatology, Palaeoecology 271, 268–78.Google Scholar
Harland, T. L. 1981. Middle Ordovician reefs of Norway. Lethaia 14, 169–88.Google Scholar
Harris, M. T., Sheehan, P. M., Ainsaar, L., Hints, L., Männik, P., Nõlvak, J. & Rubel, J. 2004. Upper Ordovician sequences of western Estonia. Palaeogeography, Palaeoclimatology, Palaeoecology 210, 135–48.Google Scholar
Hints, L., Hints, O., Kaljo, D., Kiipli, T., Männik, P., Nõlvak, J. & Pärnaste, H. 2010. Hirnantian (latest Ordovician) bio-and chemostratigraphy of the Stirnas-18 core, western Latvia. Estonian Journal of Earth Sciences 59, 124.Google Scholar
Hints, L. & Meidla, T. 1997a. Keila Stage. In Geology and Mineral Resources of Estonia (eds Raukas, A. & Teedumäe, A.), pp. 234–41. Tallinn: Estonian Academy Publishers.Google Scholar
Hints, L. & Meidla, T. 1997b. Porkuni Stage. In Geology and Mineral Resources of Estonia (eds Raukas, A. & Teedumäe, A.), pp. 282303. Tallinn: Estonian Acaemy Publishers.Google Scholar
Hints, L., Meidla, T., Nõlvak, J. & Sarv, L. 1989. Some specific features of the Late Ordovician evolution in the Baltic basin. Proceedings of the Academy of Sciences of the Estonian SSR, Geology 38, 8792.Google Scholar
Hints, L., Oraspõld, A. & Kaljo, D. 2000. Stratotype of the Porkuni Stage with comments on the Röa Member, uppermost Ordovician, Estonia. Proceedings of the Estonian Academy of Sciences, Geology 49, 177–99.Google Scholar
Hints, L., Oraspõld, A. & Nõlvak, J. 2005. The Pirgu Regional Stage (Upper Ordovician) in the East Baltic: lithostratigraphy, biozonation, and correlation. Proceedings of the Estonian Academy of Sciences, Geology 54, 224–59.Google Scholar
Holland, C. H. & Patzkowsky, M. E. 1996. Sequence stratigraphy and long-term paleoceanographic change in the Middle and Upper Ordovician of the Eastern United States. In Paleozoic Sequence Stratigraphy: Views from the North American Craton (eds Witzke, B. J., Ludvigson, G. A. & Day, J.), pp. 117–30. Boulder, CO: Geological Society of America.Google Scholar
Hucke, K. & Voigt, E. 1967. Einführung in die Geschiebeforschung (Sedimentärgeschiebe). Nederlandse Geologische Vereniging, Oldenzaal, 132 pp.Google Scholar
Jaanusson, V. 1973. Aspects of carbonate sedimentation in the Ordovician of Baltoscandia. Lethaia 6, 1134.CrossRefGoogle Scholar
Jaanusson, V. 1976. Faunal dynamics in the Middle Ordovician (Viruan) of Baltoscandia. The Ordovician System: Proceedings of the Palaeontological Association, Birmingham, September 1974, 301–26.Google Scholar
Jaanusson, V. 1978. Ordovician Meiklejohn carbonate mound, Nevada. Geological Magazine 115, 467–8.Google Scholar
Jaanusson, V. 1979. Karbonatyne postrojki v ordovike shvetsii. Izvestyia Akademii Nauk Kazakhskoy SSR, Seriya Geologicheskaya 4, 92–9.Google Scholar
Jaanusson, V. 1982. The Siljan district. IV International Symposium on the Ordovician System, Field Excursion Guide, Palaeontological Contributions from the University of Oslo 279, 1542.Google Scholar
Jourdan, F., Reimold, W. U. & Deutsch, A. 2012. Dating terrestrial impact strctures. Elements 8, 4953.Google Scholar
Juhlin, C. & Pedersen, L. B. 1987. Reflection seismic investigations of the Siljan impact structure, Sweden. Journal of Geophysical Research 92, 14113–22.CrossRefGoogle Scholar
Jux, U. 1966. Palaeoporella im Boda-Kalk von Dalarne. Palaeontographica B 118, 153–65.Google Scholar
Kaljo, D. 1957. Codonophyllacea ordovika i llandoveri Pribaltiki. Loodusuurijate Seltsi Aastaraamat 50, 153–68.Google Scholar
Kaljo, D., Hints, L., Hints, O., Männik, P., Martma, T. O. N. & Nõlvak, J. 2011. Katian prelude to the Hirnantian (Late Ordovician) mass extinction: a Baltic perspective. Geological Journal 46, 464–77.CrossRefGoogle Scholar
Kaljo, D., Hints, L., Männik, P. & Nõlvak, J. 2008. The succession of Hirnantian events based on data from Baltica: brachiopods, chitinozoans, conodonts, and carbon isotopes. Estonian Journal of Earth Sciences 57, 197218.Google Scholar
Kaljo, D., Martma, T. & Saadre, T. 2007. Post-Hunnebergian Ordovician carbon isotope trend in Baltoscandia, its environmental implications and some similarities with that of Nevada. Palaeogeography, Palaeoclimatology, Palaeoecology 245, 138–55.Google Scholar
Kaljo, D., Nõlvak, J. & Uutela, A. 1996. More about Ordovician microfossil diversity patterns in the Rapla section, northern Estonia. Proceedings of the Estonian Academy of Sciences, Geology 45, 131–48.Google Scholar
Kaminskas, D., Michelevičius, D. & Blažauskas, N. 2015 New evidence of an early Pridoli barrier reef in the southern part of the Baltic Silurian basin based on three-dimensional seismic survey, Lithuania. Estonian Journal of Earth Sciences 64, 4755.Google Scholar
Kiessling, W. 2005. Long-term relationships between ecological stability and biodiversity in Phanerozoic reefs. Nature 433, 410–3.CrossRefGoogle ScholarPubMed
Kiipli, E. 1997. Geochemistry of Llandovery black shales in the Aizpute-41 core, West Latvia. Proceedings of the Estonian Academy of Sciences, Geology 46, 127–45.Google Scholar
Kiipli, E., Kiipli, T. & Kallaste, T. 2009. Reconstruction of currents in the Mid-Ordovician-Early Silurian central Baltic Basin using geochemical and mineralogical indicators. Geology 37, 271–4.Google Scholar
Kiær, J. 1897. Faunistische Uebersicht der Etage 5 des norwegischen Silursystems. Norsk Videnskabsselkabets Skrifter 1 Mathematisk-Naturvidenskab Klasse 1897, 176.Google Scholar
Klaamann, E. R. 1966. Inkommunikatnye tabulaty Estonii. Trudy Instituta Geologii AN ESSR 1966, 197.Google Scholar
Klimenko, S., Anischenko, L. & Antoshkina, A. 2011. The Timan–Pechora sedimentary basin: palaeozoic reef formations and petroleum systems. Research Reports. Göttinger Arbeiten zur Geologie und Paläontologie, Sonderband 3, University of Göttingen, 223–36.Google Scholar
Kozlowski, R. & Kazmierczak, J. 1968. On two Ordovician calcareous algae. Acta Palaeontogica Polonica 8, 325–46.Google Scholar
Kröger, B. & Ebbestad, J. O. R. 2014. Palaeoecology and palaeogeography of Late Ordovician (Katian–Hirnantian) cephalopods of the Boda Limestone, Siljan district, Sweden. Lethaia 47, 1530.CrossRefGoogle Scholar
Kröger, B., Ebbestad, J. O. R. & Lehnert, O. 2016. Accretionary mechanisms and temporal sequence of formation of the Boda Limestone Mud Mounds (Upper Ordovician), Siljan District, Sweden. Journal Sedimentary Research 86, 117.Google Scholar
Kröger, B., Ebbestad, J. O. R., Lehnert, O., Ullmann, C. V., Korte, C., Frei, R. & Rasmussen, C. M. Ø. 2015. Subaerial speleothems and deep karst in central Sweden linked to Hirnantian glaciations. Journal of the Geological Society 172, 349–56.Google Scholar
Kröger, B., Hints, L. & Lehnert, O. 2014. Age, facies, and geometry of the Sandbian/Katian (Upper Ordovician) pelmatozoan-bryozoan-receptaculitid reefs of the Vasalemma Formation, northern Estonia. Facies 60, 963–86.Google Scholar
Kröger, B., Hints, L., Lehnert, O., Männik, P. & Joachimski, M. 2014. The early Katian (Late Ordovician) reefs near Saku, northern Estonia and the age of the Saku Member, Vasalemma Formation. Estonian Journal of Earth Sciences 63, 271–6.Google Scholar
Laškovas, J. 2000. Sedimentation Environments of the Ordovician Basin in the Southwestern Margin of the East European Platform and Lithogenesis of Deposits. Vilnius: Institute of Geology, 314 pp.Google Scholar
Lavoie, D. 1995. A Late Ordovician high-energy temperate-water carbonate ramp, southern Quebec, Canada: implications for Late Ordovician oceanography. Sedimentology 42, 95116.Google Scholar
Lindström, M., Flodén, T., Grahn, Y. & Kathol, B. 1994. Post-impact deposits in Tvären, a marine Ordovician crater south of Stockholm, Sweden. Geological Magazine 131, 91103.Google Scholar
Liow, L. H. & Nichols, J. D. 2010. Estimating rates and probabilities of origination and extinction using taxonomic occurrence data: capture-mark-recapture (CMR) approaches. In The Paleontological Society Short Course, October 30th 2010 (eds Alroy, J. & Hunt, G.), pp. 8194. The Palaeontological Society.Google Scholar
Mackenzie, D. I., Bailey, L. L. & Nichols, J. D. 2004. Investigating species co-occurrence patterns when species are detected imperfectly. Journal of Animal Ecology 73, 546–55.Google Scholar
Mackenzie, D. I., Nichols, J. D., Hines, J. E., Knutson, M. G. & Franklin, B. 2003. Estimating site occupancy, colonization, and local extinction when a species is detected imperfectly. Ecology 84, 2200–7.Google Scholar
Männil, R. 1966. Istoria razvitiya Baltiyskogo basseina v ordovike. Tallinn: Valgus, 200 pp.Google Scholar
Manten, A. A. 1971. Silurian Reefs of Gotland. Amsterdam: Elsevier, 537 pp.Google Scholar
Meidla, L. 1996. Late Ordovician ostracods of Estonia. Fossilia Baltica 2, 1222.Google Scholar
Melchin, M. J., Mitchell, C. E., Holmden, C. & Štorch, P. 2013. Environmental changes in the Late Ordovician–early Silurian: Review and new insights from black shales and nitrogen isotopes. Geological Society of America Bulletin 125, 1635–70.Google Scholar
Modliński, Z. & Szymański, B. 1997. The Ordovician lithostratigraphy of the Peribaltic Depression (NE Poland). Geological Quarterly 41, 273–88.Google Scholar
Nakrem, H. A. & Rasmussen, J. A. 2013. Oslo Region, Norway. In Field Guide for the 3rd Annual Meeting of the IGCP Project 591, The Lower Palaeozoic of Southern Sweden and the Oslo Region, Norway (eds Calner, M., Ahlberg, P., Lehnert, O. & Erlström, M.), pp. 5886. Uppsala: Sveriges Geologiska Undersöking.Google Scholar
Nestor, H. 1964. Stromatoporoidei ordovika i llandoveri Estonii. Trudy Institut Geologii AN ESSR, 1964, 1111.Google Scholar
Nestor, H. 1999. Community structure and succession of Baltoscandian Early Palaeozoic stromatoporoids. Proceedings of the Estonian Academy of Sciences, Geology 48, 123–39.Google Scholar
Nestor, H. & Webby, B. D. 2013. Biogeography of the Ordovician and Silurian stromatoporoidea. In Early Paalaeozoic Biogeography and Palaeogeography (eds Harper, D.A.T. & Servais, T.), pp. 6779. Geological Society of London, Memoir no. 38.Google Scholar
Nielsen, A. T. 2004. Ordovician sea level changes: a baltoscandian perspective. In The Great Ordovician Biodiversification Event (eds Webby, B. D., Paris, F., Droser, M. & Percival, I.), pp. 8493. New York: Columbia University Press.Google Scholar
Nielsen, A. T. 2011. A re-calibrated revised sea-level curve for the Ordovician of Baltoscandia. Ordovician of the World: 11th International Symposium on the Ordovician System, Cuadernos del Museo Geominero 14, 399402.Google Scholar
Nitecki, M. H., Webby, B. D., Spjeldnaes, N. & Yong-Yi, Z. 2004. Receptaculitids and Algae. In The Great Ordovician Biodiversification Event (eds Webby, B. D., Paris, F., Droser, M. & Percival, I.). pp. 336347. New York: Columbia University Press.Google Scholar
Opalinski, P. R. & Harland, T. 1981. The Middle Ordovician of the Oslo Region, Norway, 29, Stratigraphy of the Mjøsa Limestone in the Toten and Nes-Hamar areas. Norsk Geologisk Tidsskrift 61, 5978.Google Scholar
Oraspõld, A. 1975. On the lithology of the Porkuni Stage in Estonia. Acta et Commentationes Universitatis Tartuensis (Töid Geoloogia Alalt, VII) 359, 3375.Google Scholar
Oraspõld, A. 1982. Lithological characterisation of the boundary beds of the Vormsi and Pirgu stages in Central Estonia. Acta et Commentationes Universitatis Tartuensis (Töid Geoloogia Alalt, VIII) 527, 7590.Google Scholar
Owen, A., Bruton, D. L., Bockelie, J. F. & Bockelie, T. G. 1990. The Ordovician successions of the Oslo Region, Norway. Norges Geologiske Undersøgelse 4, 354.Google Scholar
Pancost, R. D., Freeman, K. H., Herrmann, A. D., Patzkowsky, M. E., Ainsaar, L. & Martma, T. 2013. Reconstructing Late Ordovician carbon cycle variations. Geochimica et Cosmochimica Acta 105, 433–54.Google Scholar
Paškevičius, J. 1997. The Geology of the Baltic Republics. Vilnius: Geological Survey of Lithuania, 388 pp.Google Scholar
Patzkowsky, M. E., Slupik, L. M., Arthur, M. A., Pancost, R. D. & Freeman, K. H. 1997. Late Middle Ordovician environmental change and extinction: Harbinger of the Late Ordovician or continuation of Cambrian patterns? Geology 25, 911–4.Google Scholar
Paul, C. & Bockelie, J. 1983. Evolution and functional morphology of the cystoid Sphaeronites in Britain and Scandinavia. Palaeontology 26, 687734.Google Scholar
Perens, H. 1995. Upper Ordovician sequence on the Põltsamaa–Jõlgeva–Rusakvere Line (in Estonian with English summary). In Geology of Livonia (eds Meidla, T., Jõeleht, A. & Kirs, J.), pp. 4551. Tartu: University of Tartu.Google Scholar
Pope, M. C. & Steffen, J. B. 2003. Widespread, prolonged late Middle to Late Ordovician upwelling in North America: a proxy record of glaciation? Geology 31, 63–6.Google Scholar
Poprawa, P., Sliaupa, S., Stephenson, R. & Lazauskiene, J. 1999. Late Vendian–Early Palaeozoic tectonic evolution of the Baltic Basin: regional tectonic implications from subsidence analysis. Tectonophysics 314, 219–39.Google Scholar
Raukas, A. & Teedumäe, A. (eds) 1997. Geology and Mineral Resources of Estonia. Tallinn: Estonian Academy Publishers, 436 pp.Google Scholar
Riding, R. 2002. Structure and composition of organic reefs and carbonate mud mounds: concepts and categories. Earth-Science Reviews 58, 163231.Google Scholar
Roberts, D. 2003. The Scandinavian Caledonides: event chronology, palaeogeographic settings and likely modern analogues. Tectonophysics 365, 283–99.Google Scholar
Rosenau, N. A., Herrmann, A. D. & Leslie, S. A. 2012. Conodont apatite δ18O values from a platform margin setting, Oklahoma, USA: implications for initiation of Late Ordovician icehouse conditions. Palaeogeography, Palaeoclimatology, Palaeoecology 315–316, 172–80.Google Scholar
Ross, R. J. J., Jaanusson, V. & Friedman, I. 1975. Lithology and Origin of Middle Ordovician Calcareous Mudmound at Meiklejohn Peak, Southern Nevada. US Geological Survey, Washington, Professional Paper no. 871, 48 pp.Google Scholar
Saltzman, M. R. & Young, S. A. 2005. Long-lived glaciation in the Late Ordovician? Isotopic and sequence-stratigraphic evidence from western Laurentia. Geology 33, 109–12.Google Scholar
Sell, B., Ainsaar, L. & Leslie, S. 2013. Precise timing of the Late Ordovician (Sandbian) super-eruptions and associated environmental, biological, and climatological events. Journal of the Geological Society 170, 711–4.Google Scholar
Sell, B. K., Samson, S. D., Mitchell, C. E., Mclaughlin, P. I., Koenig, A. E. & Leslie, S. A. 2015. Stratigraphic correlations using trace elements in apatite from Late Ordovician (Sandbian-Katian) K-bentonites of eastern North America. Geological Society of America Bulletin, published online 3 April 2015, doi: 10.1130/B31194.1.Google Scholar
Servais, T. & Harper, D. A. T. (eds) 2013. Cambrian–Ordovician Cephalopod Palaeogeography and Diversity. London: The Geological Society of London.Google Scholar
Sivhed, U., Erlström, M., Bojesen-Koefoed, J. A. & Löfgren, A. 2004. Upper Ordovician carbonate mounds on Gotland, central Baltic Sea: distribution, composition and reservoir characteristics. Journal of Petroleum Geology 27, 115–40.Google Scholar
Sliaupa, S., Fokin, P., Lazauskiene, J. & Stephenson, R. A. 2006. The Vendian-Early Palaeozoic sedimentary basins of the East European Craton. In European Lithosphere Dynamics (eds Gee, D.G. & Stephenson, R.A.), pp. 449–62. Geological Society of London, Memoir no. 32.Google Scholar
Sliaupa, S. & Hoth, P. 2011. Geological evolution and resources of the Baltic Sea area from the Precambrian to the Quaternary. In The Baltic Sea Basin (eds Harff, J., Björck, S. & Hoth, P.), pp. 1351. Berlin, Heidelberg: Springer.Google Scholar
Söderberg, P. & Hagenfeldt, S. E. 1995. Upper Proterozoic and Ordovician submarine outliers in the archipelago northeast of Stockholm, Sweden. GFF 117, 153–61.Google Scholar
Spjeldnæs, N. & Nitecki, M. H. 1994. Baltic Ordovician lithographic limestones. Geobios 27, Supplement 1, 267–73.Google Scholar
Stolley, E. 1898. Die silurische algenfacies und ihre verbreitung im Skandinavisch-Baltischen Silurgebiet. Schriften des Naturwissenschaftlichen Vereins Schleswig-Holstein 11, 109–31.Google Scholar
Thorslund, P. 1936. Siljanområdets brännkalkstenar och kalkindustri. Sveriges Geologiska Undersökning, Afhandlingar och Uppsater C 398, 157.Google Scholar
Tobin, K. J., Bergström, S. M. & De La Garza, P. 2005. A mid-Caradocian (453 Ma) drawdown in atmospheric pCO2 without ice sheet development? Palaeogeography, Palaeoclimatology, Palaeoecology 226, 187204.Google Scholar
Tolmacheva, T., Fedorov, P. & Egerquist, E. 2003. Conodonts and brachiopods from the Volkhov Stage (Lower Ordovician) microbial mud mound at Putilovo Quarry, north-western Russia. Bulletin of the Geological Society of Denmark 50, 6374.Google Scholar
Toomey, D. F. & Nitecki, M. N. 1979. Organic buildups in the Lower Ordovician (Canadian) of Texas and Oklahoma. Fieldiana Geology, New Series 2, 1181.Google Scholar
Tuuling, I. & Flodén, T. 2000. Late Ordovician carbonate buildups and erosional features northeast of Gotland, northern Baltic Sea. GFF 122, 237–49.Google Scholar
Tuuling, I. & Flodén, T. 2010. The Ordovician–Silurian boundary beds between Saaremaaa and Gotland, Baltic Sea, based on high resolution seismic data. Geological Quarterly 51, 217–29.Google Scholar
Ulst, R. Z. 1972. A local subdivision of Upper Ordovician of the middle Baltic region. In Regionalnaja Geologija Pribaltiki i Belorussii (ed. Ulst, R. Ž.), pp. 719. Riga: Zinatne (in Russian with English summary).Google Scholar
Ulst, R. Z., Gailite, L. K. & Yakovleva, V. I. 1982. Ordovik Latvii. Riga: Zinatne, 294 pp.Google Scholar
Uutela, A. 1998. Extent of the northern Baltic Sea during the early Palaeozoic area – new evidence from Ostrobothnia, Western Finland. Bulletin of the Geological Society of Finland 70, 5168.Google Scholar
Verbruggen, H., Ashworth, M., Loduca, S. T., Vlaeminck, C., Cocquyt, E., Sauvage, T., Zechman, F. W., Littler, D. S., Littler, M. M., Leliaert, F. & De Clerck, O. 2009. A multi-locus time-calibrated phylogeny of the siphonous green algae. Molecular Phylogenetics and Evolution 50, 642–53.Google Scholar
Webby, B. D. 1979. Ordovician stromatoporoids from the Mjøsa district, Norway. Norsk Geologisk Tidsskrift 59, 199211.Google Scholar
Webby, B. D. 1984. Ordovician reefs and climate: a review. In Aspects of the Ordovician System (ed. Bruton, D. L.), pp. 89100. Palaeontological Contributions from the University of Oslo no. 295.Google Scholar
Webby, B. D. 2002. Patterns of Ordovician reef development. In Phanerozoic Reef Patterns (eds Kiessling, W., Flügel, E. K. & Golonka, J.), pp. 129–79. Tulsa: SEPM (Society for Sedimentary Geology).Google Scholar
Webby, B. D., Elias, R. J., Young, G. A., Neumann, B. E. E. & Kaljo, D. 2004. Corals. In The Great Ordovician Biodiversification Event (eds Webby, B. D., Paris, F., Droser, M. & Percival, I.), pp. 124–46. New York: Columbia University Press.Google Scholar
Wilmsen, M. 2012. Origin and significance of Late Cretaceous bioevents. Examples from the Cenomanian. Acta Palaeontologica Polonica 57, 759–71.Google Scholar
Wiman, C. 1908. Studien über das Nordbaltische Silurgebiet, II. Bulletin of the Geological Institute University Uppsala 8, 73166.Google Scholar
Winterhalter, B. 1972. On the geology of the Bothnian Sea, an epeiric sea that has undergone Pleistocene glaciation. Geological Survey of Finland Bulletin 258, 166.Google Scholar
Wright, D. F. & Stigall, A. L. 2013. Geologic drivers of Late Ordovician faunal change in Laurentia: investigating links between tectonics, speciation, and biotic invasions. PLoS ONE 8, e68353.Google Scholar
Young, S. A., Saltzman, M. R., Bergström, S. M., Leslie, S. A. & Xu, C. 2008. Paired δ13Ccarb and δ13Corg records of Upper Ordovician (Sandbian‚ Katian) carbonates in North America and China: implications for paleoceanographic change. Palaeogeography, Palaeoclimatology, Palaeoecology 270, 166–78.Google Scholar
Supplementary material: File

Kröger supplementary material

Kröger supplementary material

Download Kröger supplementary material(File)
File 127.5 KB