Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-28T17:10:27.858Z Has data issue: false hasContentIssue false

Part II - Methods and Techniques for Fault Identification and Dating

Published online by Cambridge University Press:  02 December 2021

Holger Steffen
Affiliation:
Lantmäteriet, Sweden
Odleiv Olesen
Affiliation:
Geological Survey of Norway
Raimo Sutinen
Affiliation:
Geological Survey of Finland
Get access

Summary

The following sections introduce geological, geodetic and geophysical methods and techniques that specifically help in the identification of glacially induced faults. In addition, a summary of methods for dating of fault (re-)activation is presented, and the forthcoming drilling project into the Pärvie fault is introduced.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2021

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

References

Al Hseinat, M., Hübscher, C., Lang, J., Lüdmann, T., Ott, I., and Polom, U. (2016). Triassic to recent tectonic evolution of a crestal collapse graben above a salt-cored anticline in the Glückstadt Graben/North German Basin. Tectonophysics, 680, 5066, doi.org/10.1016/j.tecto.2016.05.008.Google Scholar
Brandes, C. and Winsemann, J. (2013). Soft-sediment deformation structures in NW Germany caused by Late Pleistocene seismicity. International Journal of Earth Sciences, 102, 22552274, doi.org/10.1007/s00531-013-0914-4.CrossRefGoogle Scholar
Brandes, C., Winsemann, J., Roskosch, J. et al. (2012). Activity along the Osning Thrust in Central Europe during the Lateglacial: ice-sheet and lithosphere interactions. Quaternary Science Reviews, 38, 4962, doi.org/10.1016/j.quascirev.2012.01.021.CrossRefGoogle Scholar
Brandes, C., Steffen, H., Steffen, R. and Wu, P. (2015). Intraplate seismicity in northern Central Europe is induced by the last glaciation. Geology, 43, 611614, doi.org/10.1130/G36710.1.CrossRefGoogle Scholar
Brandes, C., Steffen, H., Sandersen, P. B. E., Wu, P. and Winsemann, J. (2018). Glacially induced faulting along the NW segment of the Sorgenfrei-Tornquist Zone, northern Denmark: implications for neotectonics and Lateglacial fault-bound basin formation. Quaternary Science Reviews, 189, 149168, doi.org/10.1016/j.quascirev.2018.03.036.Google Scholar
Britze, P. and Japsen, P. (1991). Geological map of Denmark 1:400 000: the Danish Basin: «Top Zechstein» and the Triassic (two-way travel time and depth, thickness and interval velocity). Geological Survey of Denmark, Map Series, 31, 14.Google Scholar
Buldovicz, S. N., Khilimonyuk, V. Z., Bychkov, A. Y. et al. (2018). Cryovolcanism on the Earth: origin of a spectacular crater in the Yamal Peninsula (Russia). Scientific Reports, 8, 13534, doi.org/10.1038/s41598-018-31858-9.Google Scholar
Burbank, D. W. and Anderson, R. S. (2012). Tectonic Geomorphology, 2nd ed., Wiley-Blackwell, Hoboken, New Jersey.Google Scholar
Dramis, F. and Blumetti, A. M. (2005). Some considerations concerning seismic geomorphology and paleoseismology. Tectonophysics, 408, 177191, doi.org/10.1016/j.tecto.2005.05.032CrossRefGoogle Scholar
Ekström, G., Nettles, M. and Tsai, V. C. (2006). Seasonality and increasing frequency of Greenland glacial earthquakes. Science, 311(5768), 17561758, doi.org/10.1126/science.1122112.Google Scholar
French, H. M. (2017). The Periglacial Environment, 4th ed., John Wiley & Sons Ltd., Hoboken, New Jersey.Google Scholar
Grim, S. and Sirocko, F. (2012). Natural depressions on modern topography in Schleswig-Holstein (Northern Germany) – indicators for recent crustal movements or “only” kettle holes? Zeitschrift der deutschen Gesellschaft für Geowissenschaften, 163(4), 469481, doi.org/10.1127/1860-1804/2012/0163-000.CrossRefGoogle Scholar
Grube, A. (2019). Palaeoseismic structures in Quaternary sediments of Hamburg (NW Germany), earthquake evidence during the younger Weichselian and Holocene. International Journal of Earth Sciences, 108(3), 845861, doi.org/10.1007/s00531-019-01681-2.Google Scholar
Hoppe, G. (1952). Hummocky moraine regions with special reference to the interior of Norrbotten. Geografiska Annaler, 34, 172.Google Scholar
Hungr, O., Leroueil, S. and Picarelli, L. (2014). The Varnes classification of landslide types, an update. Landslides, 11, 167194, doi.org/10.1007/s10346-013-0436-y.Google Scholar
Johnson, M. D., Fredin, O., Ojala, A. E. K. and Peterson, G. (2015). Unraveling Scandinavian geomorphology: the LiDAR revolution. GFF, 137, 245251, doi.org/11035897.1111410.Google Scholar
Knudsen, C. G., Larsen, E., Sejrup, H. P. and Stalsberg, K. (2006). Hummocky moraine landscape on Jæren, SW Norway – implications for glacier dynamics during the last glaciations. Geomorphology, 77, 153168, doi.org/10.1016/j.geomorph.2005.12.011.CrossRefGoogle Scholar
Kuivamäki, A., Vuorela, P. and Paananen, M. (1998). Indications of Postglacial and Recent Bedrock Movements in Finland and Russian Karelia. Geological Survey of Finland Nuclear Waste Disposal Research Report YST-99, Espoo, Finland, 92 pp.Google Scholar
Kujansuu, R. (1967). On the deglaciation of western Finnish Lapland. Geological Survey of Finland Bulletin, 232.Google Scholar
Kujansuu, R. (1972). On landslides in Finnish Lapland. Geological Survey of Finland Bulletin, 256.Google Scholar
Lagerbäck, R. (1988). The Veiki moraines in northern Sweden – widespread evidence of an Early Weichselian deglaciation. Boreas 17, 469486, doi.org/10.1111/j.1502-3885.1988.tb00562.xCrossRefGoogle Scholar
Lagerbäck, R. and Sundh, M. (2008). Early Holocene faulting and paleoseismicity in northern Sweden. Geological Survey of Sweden Research Paper C 836, 80 pp.Google Scholar
Lang, J., Hampel, A., Brandes, C. and Winsemann, J. (2014). Response of salt structures to ice-sheet loading: implications for ice-marginal and subglacial processes. Quaternary Science Reviews, 101, 217233, doi.org/10.1016/j.quascirev.2014.07.022.Google Scholar
Lykke-Andersen, H., Madirazza, I. and Sandersen, P. B. E. (1996). Tektonik og landskabsdannelse i Midtjylland [Tectonics and landscape formation in Mid-Jutland]. Geologisk Tidsskrift 1996, 3, 132.Google Scholar
Markovaara-Koivisto, M., Ojala, A. E. K., Mattila, J. et al. (2020). Geomorphological evidence of paleoseismicity: surficial and underground structures of Pasmajärvi postglacial fault. Earth Surface Processes and Landforms, 45(12), 30113024, doi.org/10.1002/esp.4948.Google Scholar
Mattila, J., Ojala, A. E. K., Ruskeeniemi, T. et al. (2019). Evidence of multiple slip events on postglacial faults in northern Fennoscandia. Quaternary Science Reviews, 215, 242252, doi.org/10.1016/j.quascirev.2019.05.022.Google Scholar
McCalpin, J. P. (2009). Paleoseismology. International Geophysics Series Vol. 95, 2nd ed., Elsevier, Amsterdam, doi.org/10.1016/S0074-6142(09)95001-X.Google Scholar
McCalpin, J. P. and Nelson, A. R. (2009). Introduction to paleoseismology. In McCalpin, J. P., ed., Paleoseismology. International Geophysics Series Vol. 95, 2nd ed., Elsevier, Amsterdam, pp. 127, doi.org/10.1016/S0074-6142(09)95001-X.Google Scholar
Menzies, J. and Shilts, W. W. (2002). Subglacial environments. In Menzies, J., ed., Modern & Past Glacial Environments. Butterworth-Heinemann, Oxford, pp. 183278.Google Scholar
Michetti, A. M., Esposito, E., Guerrieri, L. et al. (2007). Environmental seismic intensity scale – ESI 2007. Memorie descrittive della carta geologica d’Italia, 74.Google Scholar
Middleton, M., Heikkonen, J., Nevalainen, P., Hyvönen, E. and Sutinen, R. (2020a). Machine learning-based mapping of micro-topographic earthquake-induced paleo Pulju moraines and liquefaction spreads. Geomorphology, 358, 107099, doi.org/10.1016/j.geomorph.2020.107099.Google Scholar
Middleton, M., Nevalainen, P., Hyvönen, E., Heikkonen, J. and Sutinen, R. (2020b). Pattern recognition of LiDAR data and sediment anisotropy advocate polygenetic subglacial mass-flow origin of the Kemijärvi hummocky moraine field in northern Finland. Geomorphology, 362, 107212, doi.org/10.1016/j.geomorph.2020.107212.CrossRefGoogle Scholar
Mikko, H., Smith, C. A., Lund, B., Ask, M. V. S. and Munier, R. (2015). LiDAR-derived inventory of post-glacial fault scarps in Sweden. GFF, 137, 334338, doi.org/10.1080/11035897.2015.1036360.CrossRefGoogle Scholar
Muscheler, R., Kromer, B., Björk, S. et al. (2008). Tree rings and ice cores reveal 14C calibration uncertainties during the Younger Dryas. Nature Geoscience, 1(4), 263267, doi.org/10.1038/ngeo128.Google Scholar
Nordkalott Project (1986). Geological Map, Northern Fennoscandia, 1:1 million. Geological Surveys of Finland, Norway and Sweden.Google Scholar
Obermaier, S. F. (2009). Using liquefaction-induced and other soft-sediment features for paleoseismic analysis. International Geophysics, 95, 497564, doi.org/10.1016/S0074-6142(09)95007-0.Google Scholar
Öhrling, C., Peterson, G. and Mikko, H. (2018). Detailed Geomorphological Analysis of LiDAR Derived Elevation Data, Forsmark. Searching for Indicatives of Late- and Postglacial Seismic Activity. SKB Report R-18-10, Swedish Nuclear Fuel and Waste Management Co., Stockholm, 38 pp.Google Scholar
Ojala, A. E. K., Markovaara-Koivisto, M., Middleton, M. et al. (2018). Dating of paleolandslides in western Finnish Lapland. Earth Surface Processes and Landforms, 43, 24492462, doi.org/10.1002/esp.4408.CrossRefGoogle Scholar
Ojala, A. E. K., Mattila, J., Markovaara-Koivisto, M. et al. (2019a). Distribution and morphology of landslides in northern Finland: an analysis of postglacial seismic activity. Geomorphology, 326, 190201, doi.org/10.1016/j.geomorph.2017.08.045.CrossRefGoogle Scholar
Ojala, A. E. K., Peterson, G., Mäkinen, J. et al. (2019b). Ice-sheet scale distribution and morphometry of triangular shaped hummocks (murtoos): a subglacial landform produced during rapid retreat of the Scandinavian Ice Sheet. Annals of Glaciology, 60(80), 115126, doi.org/10.1017/aog.2019.34.CrossRefGoogle Scholar
Ojala, A. E. K., Mattila, J., Middleton, M. et al. (2020). Earthquake-induced deformation structures in glacial sediments – evidence on fault reactivation and instability at the Vaalajärvi fault in northern Fennoscandia. Journal of Seismology, 24(3), doi.org/10.1007/s10950-020-09915-6.CrossRefGoogle Scholar
Olesen, O. (1988). The Stuoragurra Fault, evidence of neotectonics in the Precambrian of Finnmark, northern Norway. Norsk Geologisk Tidsskrift, 68, 107118.Google Scholar
Olesen, O., Blikra, L. H., Braathen, A. et al. (2004). Neotectonic deformation in Norway and its implications: a review. Norwegian Journal of Geology, 84, 334.Google Scholar
Olesen, O., Bungum, H., Dehls, J. et al. (2013). Neotectonics, seismicity and contemporary stress field in Norway – mechanisms and implications. In Olsen, L., Fredin, O. and Olesen, O., eds., Quaternary Geology of Norway, Geological Survey of Norway Special Publication, 13, pp. 145174.Google Scholar
Palmu, J.-P., Ojala, A. E. K., Ruskeeniemi, T., Sutinen, R. and Mattila, J. (2015). LiDAR DEM detection and classification of postglacial faults and seismically-induced landforms in Finland: a paleoseismic database. GFF, 137, 344352, doi.org/10.1080/11035897.2015.1068370.CrossRefGoogle Scholar
Påsse, T. (1998). Lake-tilting, a method for estimation of glacio-isostatic uplift. Boreas, 27, 6980, doi.org/10.1111/j.1502-3885.1998.tb00868.x.CrossRefGoogle Scholar
Pisarska-Jamroży, M., Belzyt, S., Börner, A. et al. (2018). Evidence from seismites for glacio-isostatically induced crustal faulting in front of an advancing land-ice mass (Rügen Island, SW Baltic Sea). Tectonophysics, 745, 338348, doi.org/10.1016/j.tecto.2018.08.004.Google Scholar
Rasmussen, S. O., Bigler, M., Blockey, S. P. et al. (2014). A stratigraphic framework for abrupt climatic changes during the last Glacial period based on three synchronized Greenland ice-core records: refining and extending the INTIMATE event stratigraphy. Quaternary Science Reviews, 106, 1428, doi.org/10.1016/j.quascirev.2014.09.007.Google Scholar
Sandersen, P. B. E. and Jørgensen, F. (2015). Neotectonic deformation of a Late Weichselian outwash plain by deglaciation-induced fault reactivation of a deep-seated graben structure. Boreas, 44, 413431, doi.org/10.1111/bor.12103.CrossRefGoogle Scholar
Schoof, C. (2010). Ice-sheet acceleration driven by melt supply variability. Nature, 468(7325), 803806, doi.org/10.1038/nature09618.Google Scholar
Schumm, S. A., Dumont, J. F. and Holbrook, J. M. (2002). Active Tectonics and Alluvial Rivers. Cambridge University Press, Cambridge.Google Scholar
Seppä, H., Tikkanen, M. and Mäkiaho, J.-P. (2012). Tilting of Lake Pielinen, eastern Finland – an example of extreme transgressions and regressions caused by differential post-glacial isostatic uplift. Estonian Journal of Earth Sciences, 61(3), 149161, doi.org/10.3176/earth.2012.3.02.CrossRefGoogle Scholar
Sirocko, F., Reicherter, K., Lehne, R. W. et al. (2008). Glaciation, salt and the present landscape. In Littke, R. et al., eds., Dynamics of Complex Intracontinental Basins: The Central European Basin System. Springer Verlag, Heidelberg, pp. 234245.Google Scholar
Sirocko, F., Szeder, T., Seelos, C. et al. (2002). Young tectonic and halokinetic movements in the North-German-Basin: its effect on formation of modern rivers and surface morphology. Netherlands Journal of Geosciences/Geologie en Mijnbouw, 81(3-4), 431441, doi.org/10.1017/S0016774600022708.Google Scholar
Smith, C. A., Grigull, S. and Mikko, H. (2018). Geomorphic evidence of multiple surface ruptures of the Merasjärvi “postglacial fault”, northern Sweden. GFF, 140(4), 318322, doi.org/10.1080/11035897.2018.1492963.Google Scholar
Stewart, I. S., Sauber, J. and Rose, J. (2000). Glacio-seismotectonics: ice sheets, crustal deformation and seismicity. Quaternary Science Reviews, 19, 13671389, doi.org/10.1016/S0277-3791(00)00094-9.CrossRefGoogle Scholar
Sutinen, R. (1985). On the subglacial sedimentation of hummocky moraines and eskers in northern Finland. Striae, 22, 2125.Google Scholar
Sutinen, R., Middleton, M., Hänninen, P. et al. (2007). Dielectric constant time stability of glacial till at a clear-cut site. Geoderma, 141, 311319, doi.org/10.1016/j.geoderma.2007.06.016.CrossRefGoogle Scholar
Sutinen, R., Piekkari, M. and Middleton, M. (2009a). Glacial geomorphology in Utsjoki, Finnish Lapland proposes Younger Dryas fault-instability. Global and Planetary Change, 69, 1628, doi.org/10.1016/j.gloplacha.2009.07.002.CrossRefGoogle Scholar
Sutinen, R., Middleton, M., Liwata, M., Piekkari, M. and Hyvönen, E. (2009b). Sediment anisotropy coincides with moraine ridge trend in south-central Finnish Lapland. Boreas, 38, 638646, doi.org/10.1111/j.1502-3885.2009.00089.x.Google Scholar
Sutinen, R., Aro, I., Närhi, P., Piekkari, M. and Middleton, M. (2014a). Maskevarri Ráhhpát in Finnmark, northern Norway – is it an earthquake-induced landform complex? Solid Earth, 5, 683691, doi.org/10.5194/se-5-683-2014.Google Scholar
Sutinen, R., Hyvönen, E., Middleton, M. and Ruskeeniemi, T. (2014b). Airborne LiDAR detection of postglacial faults and Pulju moraine in Palojärvi, Finnish Lapland. Global and Planetary Change, 115, 2432, doi.org/10.1016/j.gloplacha.2014.01.007.Google Scholar
Sutinen, R., Hyvönen, E. and Kukkonen, I. (2014c). LiDAR detection of paleolandslides in the vicinity of the Suasselkä postglacial fault, Finnish Lapland. International Journal of Applied Earth Observation and Geoinformation, 27, 9199, doi.org/10.1016/j.jag.2013.05.004.Google Scholar
Sutinen, R., Hyvönen, E., Middleton, M. and Airo, M.-L. (2018). Earthquake-induced deformations on ice-stream landforms in Kuusamo, eastern Finnish Lapland. Global and Planetary Change, 160, 4660, doi.org/10.1016/j.gloplacha.2017.11.011.Google Scholar
Sutinen, R., Andreani, L. and Middleton, M. (2019a). Post-Younger Dryas fault instability and de-formations on ice lineations in Finnish Lapland. Geomorphology, 326, 202212, doi.org/10.1016/j.geomorph.2018.08.034.CrossRefGoogle Scholar
Sutinen, R., Hyvönen, E., Liwata-Kenttälä, P. et al. (2019b). Electrical-sedimentary anisotropy of landforms adjacent to postglacial faults in Lapland. Geomorphology, 326, 213224, doi.org/10.1016/j.geomorph.2018.01.008.Google Scholar
Sutinen, R., Sutinen, A. and Middleton, M. (2021). Subglacial squeeze-up moraines adjacent to the Vaalajärvi-Ristonmännikkö glacially-induced fault system, Finnish Lapland. Geomorphology, 384, 107716, doi.org/10.1016/j.geomorph.2021.107716.Google Scholar
Ter-Borch, N. (1991). Geological map of Denmark, 1:500,000. Structural map of the top chalk group. Geological Survey of Denmark Map Series 7, 4 pp. Copenhagen.Google Scholar
van Balen, R. T., Bakker, M. A. J., Kasse, C. et al. (2019). A Late Glacial surface rupturing earthquake at the Peel Boundary fault zone, Roer Valley Rift System, the Netherlands. Quaternary Science Reviews, 218, 254266, doi.org/10.1016/j.quascirev.2019.06.033.CrossRefGoogle Scholar
Van Vliet-Lanoë, B., Brulhet, J., Combes, P. et al. (2016). Quaternary thermokarst and thermal erosion features in northern France: origin and palaeoenvironments. Boreas, 46 (3), 442461, doi.org/10.1111/bor.12221.CrossRefGoogle Scholar
Van Vliet-Lanoë, B., Maygari, A. and Meilliez, F. (2004). Distinguishing between tectonic and periglacial deformations of quaternary continental deposits in Europe. Global and Planetary Change, 43, 103127, doi.org/10.1016/j.gloplacha.2004.03.003.Google Scholar
Wu, P., Johnston, P. and Lambeck, K. (1999). Postglacial rebound and fault instability in Fennoscandia. Geophysical Journal International, 139, 657670, doi.org/10.1046/j.1365-246x.1999.00963.x.Google Scholar

References

Aber, J. S. and Ber, A. (2007). Glaciotectonism. Developments in Quaternary Science, Vol. 6. Elsevier, Amsterdam.Google Scholar
Alfaro, P., Delgado, J., Estévez, A. et al. (2002). Liquefaction and fluidization structures in Messinian storm deposits (Bajo Segura Basin, Betic Cordillera, southern Spain). International Journal of Earth Sciences, 91, 505513, doi.org/10.1007/s00531-001-0241-z.Google Scholar
Åmark, M. (1986). Clastic dikes formed beneath an active glacier. Geologiska Foereningen i Stockholm Förhandlingar, 108, 1320, doi.org/10.1080/11035898609453740.Google Scholar
Atkinson, G. M., Finn, W. L. and Charlwood, R. G. (1984). Simple computation of liquefaction probability for seismic hazard applications. Earthquake Spectra, 1, 107123, doi.org/10.1193/1.1585259.Google Scholar
Ballas, G., Fossen, H. and Soliva, R. (2015). Factors controlling permeability of cataclastic deformation bands and faults in porous sandstone reservoirs. Journal of Structural Geology, 76, 121, doi.org/10.1016/j.jsg.2015.03.013.Google Scholar
Belzyt, S., Nartišs, M., Pisarska-Jamroży, M., Woronko, B. and Bitinas, A. (2018). Large-scale glaciotectonically-deformed Pleistocene sediments with deformed layers sandwiched between undeformed layers, Baltmuiža site, Western Latvia. In Pisarska-Jamroży, M. and Bitinas, A., eds., Soft-Sediment Deformation Structures and Palaeoseismic Phenomena in the South-Eastern Baltic Region. Excursion Guide of International Palaeoseismological Field Workshop, 17–21st September 2018, Vilnius, Lithuania. Lithuanian Geological Survey, Lithuanian Geological Society, pp. 38–42.Google Scholar
Belzyt, S., Pisarska-Jamroży, M., Bitinas, A. et al. (2021). Repetitive Late Pleistocene soft-sediment deformation by seismicity-induced liquefaction in north-western Lithuania. Sedimentology, doi.org/10.1111/sed.12883.Google Scholar
Benn, D. I. and Evans, D. J. A. (2013). Glaciers and Glaciation, 2nd ed. Routledge, New York.Google Scholar
Bennett, M. R., Huddart, D., Waller, R. I. et al. (2004). Styles of ice-marginal deformation at Hagafellsjökull–Eystri, Iceland during the 1998/99 winter-spring surge. Boreas, 33, 97107, doi.org/10.1111/j.1502-3885.2004.tb01132.x.Google Scholar
Bertran, P. (1993). Deformation‐induced microstructures in soils affected by mass movements. Earth Surface Processes and Landforms, 18, 645660, doi.org/10.1002/esp.3290180707.Google Scholar
Bertran, P., Font, M., Giret, A., Manchuel, K. and Sicilia, D. (2019a). Experimental soft-sediment deformation caused by fluidization and intrusive ice melt in sand. Sedimentology, 66(3), 11021117, doi.org/10.1111/sed.12537.CrossRefGoogle Scholar
Bertran, P., Manchuel, K. and Sicilia, D. (2019b). Discussion on ‘Palaeoseismic structures in Quaternary sediments, related to an assumed fault zone north of the Permian Peissen–Gnutz salt structure (NW Germany) – neotectonic activity and earthquakes from the Saalian to the Holocene’ (Grube, 2019). Geomorphology, 365, 106704, doi.org/10.1016/j.geomorph.2019.03.010.Google Scholar
Bockheim, J. G. and Tarnocai, C. (1998). Recognition of cryoturbation for classifying permafrost-affected soils. Geoderma, 81, 281293, doi.org/10.1016/S0016-7061(97)00115-8.Google Scholar
Boulton, G. S. and Caban, P. (1995). Groundwater flow beneath ice sheets: part II – its impact on glacier tectonic structures and moraine formation. Quaternary Science Reviews, 14, 563587, doi.org/10.1016/0277-3791(95)00058-W.Google Scholar
Brandes, C. and Le Heron, D. P. (2010). The glaciotectonic deformation of Quaternary sediments by fault-propagation folding. Proceedings of the Geologists’ Association, 121, 270280, doi.org/10.1016/j.pgeola.2010.03.001.CrossRefGoogle Scholar
Brandes, C. and Tanner, D. C. (2012). Three-dimensional geometry and fabric of shear deformation-bands in unconsolidated Pleistocene sediments. Tectonophysics, 518 –521, 8492, doi.org/10.1016/j.tecto.2011.11.012.Google Scholar
Brandes, C. and Winsemann, J. (2013). Soft-sediment deformation structures in NW Germany caused by Late Pleistocene seismicity. International Journal of Earth Sciences, 102, 22552274, doi.org/10.1007/s00531-013-0914-4.Google Scholar
Brandes, C., Winsemann, J., Roskosch, J. et al. (2012). Activity along the Osning Thrust in Central Europe during the Lateglacial: ice-sheet and lithosphere interactions. Quaternary Science Reviews, 38, 4962, doi.org/10.1016/j.quascirev.2012.01.021.Google Scholar
Brandes, C., Steffen, H., Sandersen, P. B. E., Wu, P. and Winsemann, J. (2018a). Glacially induced faulting along the NW segment of the Sorgenfrei-Tornquist Zone, northern Denmark: implications for neotectonics and Lateglacial fault-bound basin formation. Quaternary Science Reviews, 189, 149168, doi.org/10.1016/j.quascirev.2018.03.036.Google Scholar
Brandes, C., Igel, J., Loewer, M. et al. (2018b). Visualisation and analysis of shear-deformation bands in unconsolidated Pleistocene sand using ground-penetrating radar: implications for paleoseismological studies. Sedimentary Geology, 367, 135145, doi.org/10.1016/j.sedgeo.2018.02.005.Google Scholar
Brooks, G. R. (2018). Deglacial record of palaeoearthquakes interpreted from mass transport deposits at three lakes near Rouyn–Noranda, north-western Quebec, Canada. Sedimentology, 65, 24392467, doi.org/10.1111/sed.12473.Google Scholar
Buldovicz, S. N., Khilimonyuk, V. Z., Bychkov, A. Y. et al. (2018). Cryovolcanism on the Earth: origin of a spectacular crater in the Yamal Peninsula (Russia). Scientific Reports, 8, 13534, doi.org/10.1038/s41598-018-31858-9.CrossRefGoogle ScholarPubMed
Cashman, S. M., Baldwin, J. N., Cashman, K. V., Swanson, K. and Crawford, R. (2007). Microstructures developed by coseismic and aseismic faulting in near-surface sediments, San Andreas Fault, California. Geology, 35, 611614, doi.org/10.1130/G23545A.1.Google Scholar
Chamley, H. (1990). Sedimentology. Springer, Berlin/Heidelberg.Google Scholar
Chen, J. and Lee, H. S. (2013). Soft-sediment deformation structures in Cambrian siliciclastic and carbonate storm deposits (Shandong Province, China): differential liquefaction and fluidization triggered by storm-wave loading. Sedimentary Geology, 288, 8194, doi.org/10.1016/j.sedgeo.2013.02.001.Google Scholar
Collinson, J. D., Mountney, N. P. and Thompson, D. B. (2006). Sedimentary Structures, 3rd ed., Terra Publishing, England.Google Scholar
Davenport, C. A., Ringrose, P. S., Becker, A., Hancock, P. and Fenton, C. (1989). Geological investigations of late and post glacial earthquake activity in Scotland. In Gregersen, S. and Basham, P. W., eds., Earthquakes at North-Atlantic Passive Margins: Neotectonics and Postglacial Rebound. NATO ASI Series 266. Springer, Dordrecht, pp. 175194.Google Scholar
Deynoux, M., Proust, J. N., Durand, J. and Merino, E. (1990). Water-transfer cylindrical structures in the Late Proterozoic eolian sandstones in the Taoudeni Basin, West Africa. Sedimentary Geology, 66, 227242, doi.org/10.1016/0037-0738(90)90061-W.CrossRefGoogle Scholar
Dobiński, W. (2011). Permafrost. Earth-Science Reviews, 108, 158169, doi.org/10.1016/j.earscirev.2011.06.007.Google Scholar
Druzhinina, O., Bitinas, A., Molodkov, A. and Kolesnik, T. (2017). Palaeoseismic deformations in the Eastern Baltic region (Kaliningrad District of Russia). Estonian Journal of Earth Sciences, 66, 119129, doi.org/10.3176/earth.2017.09.Google Scholar
Eden, D. J. and Eyles, N. (2001). Description and numerical model of Pleistocene iceberg scours and ice-keel turbated facies at Toronto, Canada. Sedimentology, 48, 10791102, doi.org/10.1046/j.1365-3091.2001.00409.x.Google Scholar
Fossen, H. (2010). Deformation bands formed during soft-sediment deformation: observations from SE Utah. Marine and Petroleum Geology, 27, 215222, doi.org/10.1016/j.marpetgeo.2009.06.005.Google Scholar
French, H. M. (2017). The Periglacial Environment, 4th ed., John Wiley and Sons, Chichester.Google Scholar
Frey, S. E., Gingras, M. K. and Dashtgard, S. E. (2009). Experimental studies of gas-escape and water-escape structures: mechanisms and morphologies. Journal of Sedimentological Research, 79(11), 808816, doi.org/10.2110/jsr.2009.087.Google Scholar
Galli, P. (2000). New empirical relationships between magnitude and distance for liquefaction. Tectonophysics, 324, 169187, doi.org/10.1016/S0040–1951(00)00118-9.Google Scholar
Gehrmann, A. and Harding, C. (2018). Geomorphological mapping and spatial analyses of an Upper Weichselian glacitectonic complex based on LiDAR data, Jasmund Peninsula (NE Rügen), Germany. Geosciences, 8(6), 208, doi.org/10.3390/geosciences8060208.Google Scholar
Giona Bucci, M., Almond, P., Villamor, P. et al. (2017). When the earth blisters: exploring recurrent liquefaction features in the coastal system of Christchurch, New Zealand. Terra Nova, 29, 162172, doi.org/10.1111/ter.12259.Google Scholar
Giona Bucci, M., Smith, C. M., Almond, P. C., Villamor, P. and Tuttle, M. P. (2019). Micromorphological analysis of liquefaction features in alluvial and coastal environments of Christchurch, New Zealand. Sedimentology, 66, 963982, doi.org/10.1111/sed.12526.Google Scholar
Grube, A. (2019a). Palaeoseismic structures in Quaternary sediments of Hamburg (NW Germany), earthquakes evidence during the younger Weichselian and Holocene. International Journal of Earth Sciences, 108, 845861, doi.org/10.1007/s00531-019-01681-2.Google Scholar
Grube, A. (2019b). Palaeoseismic structures in Quaternary sediments, related to an assumed fault zone north of the Permian Peissen-Gnutz salt structure (NW Germany) – neotectonic activity and earthquakes from the Saalian to the Holocene. Geomorphology, 328, 1527, doi.org/10.1016/j.geomorph.2018.12.004.Google Scholar
Gruszka, B. and van Loon, A. J. (2011). Genesis of a giant gravity-induced depression (gravifossum) in the Enköping esker, S. Sweden. Sedimentary Geology, 235, 304313, doi.org/10.1016/j.sedgeo.2010.10.004.Google Scholar
Harry, D. G. (1988). Ground ice and permafrost. In M. J. Clark, ed., Advances in Periglacial Geomorphology, Wiley, Chichester, pp. 113149.Google Scholar
Hart, J. K. and Boulton, G. S. (1991). The interrelation of glaciotectonic and glacio-depositional processes within the glacial environment. Quaternary Science Reviews, 10, 335350, doi.org/10.1016/0277-3791(91)90035-S.Google Scholar
Hoffmann, G. and Reicherter, K. (2012). Soft-sediment deformation of Late Pleistocene sediments along the southwestern coast of the Baltic Sea (NE Germany). International Journal of Earth Sciences, 101, 351363, doi.org/10.1007/s00531-010-0633-z.Google Scholar
Houtgast, R. F., van Balen, R. T. and Kasse, C. (2005). Late Quaternary evolution of the Feldbiss Fault (Roer Valley Rift System, the Netherlands) based on trenching, and its potential relation to glacial unloading. Quaternary Science Reviews, 24, 489508, doi.org/10.1016/j.quascirev.2004.01.012.Google Scholar
Hungr, O., Leroueil, S. and Picarelli, L. (2014). The Varnes classification of landslide types, an update. Landslides, 11, 167194, doi.org/10.1007/s10346-013-0436-y.Google Scholar
Hurst, A. and Cartwright, J. (2007). Relevance of sand injectites to hydrocarbon exploration and production. In Hurst, A. and Cartwright, J., eds., Sand Injectites: Implications for Hydrocarbon Exploration and Production, AAPG Memoir 87, Tulsa, pp. 119, doi.org/10.1306/1209846M871546.Google Scholar
Hurst, A., Cartwright, J. and Duranti, D. (2003). Fluidization structures produced by upward injection of sand through a sealing lithology. In P. Van Rensbergen, R. R. Hillis, A. J. Maltman and C. K. Morley, eds., Subsurface Sediment Mobilization. Geological Society, London, Special Publication, Vol. 216, pp. 123138, doi.org/10.1144/GSL.SP.2003.216.01.09.Google Scholar
Hurst, A., Scott, A. and Vigorito, M. (2011). Physical characteristics of sand injectites. Earth-Science Reviews, 106, 215246, doi.org/10.1016/j.earscirev.2011.02.004.Google Scholar
Kenzler, M., Obst, K., Hüneke, H. and Schütze, K. (2010). Glazitektonische Deformation der kretazischen und pleistozänen Sedimente an der Steilküste von Jasmund nördlich des Königsstuhls (Rügen) [Glaciotectonic deformation of the Cretaceous and Pleistocene sediments on the steep coast of Jasmund north of the Königsstuhl (Rügen Island)]. Brandenburger Geowissenschaftliche Beiträge, 17, 107122.Google Scholar
Kowalski, A., Makoś, M. and Pitura, M. (2018). New insights into the glacial history of southwestern Poland based on large-scale glaciotectonic deformations – a case study from the Czaple II Gravel Pit (Western Sudetes). Annales Societatis Geologorum Poloniae, 88, 341359, doi.org/10.14241/asgp.2018.022.Google Scholar
Lee, J. R. and Phillips, E. R. (2008). Progressive soft sediment deformation within a subglacial shear zone – a hybrid mosaic-pervasive deformation model for Middle Pleistocene glaciotectonised sediments from eastern England. Quaternary Science Reviews, 27, 13501362, doi.org/10.1016/j.quascirev.2008.03.009.Google Scholar
Li, Y., Craven, J., Schweig, E. S. and Obermeier, S. F. (1996). Sand boils induced by the 1993 Mississippi River flood: could they one day be misinterpreted as earthquake-induced liquefaction? Geology, 24, 171174, doi.org/10.1130/0091-7613(1996)024<0171:SBIBTM>2.3.CO;2.Google Scholar
Longva, O. and Bakkejord, K. J. (1990). Iceberg deformation and erosion in soft sediments, southeast Norway. Marine Geology, 92, 87104, doi.org/10.1016/0025-3227(90)90028 I.Google Scholar
Lowe, D. R. (1975). Water escape structures in coarse-grained sediments. Sedimentology, 22, 157204, doi.org/10.1111/j.1365-3091.1975.tb00290.x.Google Scholar
Maizels, J. K. (1992). Boulder ring structures produced during jökulhaups flows: origin and hydraulic significance. Geografiska Annaler, 74A, 2133, doi.org/10.1080/04353676.1992.11880346.Google Scholar
Matsuoka, N. (2001). Solifluction rates, processes and landforms: a global review. Earth-Science Reviews, 55, 107134, doi.org/10.1016/S0012-8252(01)00057-5.Google Scholar
Meinsen, J., Winsemann, J., Roskosch, J. et al. (2014). Climate control on the evolution of Late Pleistocene alluvial‐fan and aeolian sand‐sheet systems in NW Germany. Boreas, 43, 4266, doi.org/10.1111/bor.12021.Google Scholar
Molina, J. M., Alfaro, P., Moretti, M. and Soria, J. M. (1998). Soft-sediment deformation structures induced by cyclic stress of storm waves in tempestites (Miocene, Guadalquivir Basin, Spain). Terra Nova, 10, 145150, doi.org/10.1046/j.1365-3121.1998.00183.x.Google Scholar
Monecke, K., Anselmetti, F. S., Becker, A. et al. (2006). Earthquake-induced deformation structures in lake deposits: a Late Pleistocene to Holocene paleoseismic record for Central Switzerland. Eclogae Geologicae Helvetiae, 99, 343362, doi.org/10.1007/s00015-006-1193-x.CrossRefGoogle Scholar
Montenat, C., Barrier, P., d’Estevou, P. O. and Hibsch, C. (2007). Seismites: an attempt at critical analysis and classification. Sedimentary Geology, 196, 530, doi.org/10.1016/j.sedgeo.2006.08.004.Google Scholar
Moretti, M. and Sabato, L. (2007). Recognition of trigger mechanisms for soft-sediment deformation in the Pleistocene lacustrine deposits of the SantʻArcangelo Basin (Southern Italy): seismic shock vs. overloading. Sedimentary Geology, 196, 3145, doi.org/10.1016/j.sedgeo.2006.05.012.Google Scholar
Moretti, M., Miguel, J., Alfaro, P. and Walsh, N. (2001). Asymmetrical soft-sediment deformation structures triggered by rapid sedimentation in turbiditic deposits (Late Miocene, Guadix Basin, Southern Spain). Facies, 44, 283294, doi.org/10.1007/BF02668179.Google Scholar
Moretti, M., Alfaro, P. and Owen, G. (2016). The environmental significance of soft-sediment deformation structures: key signatures for sedimentary and tectonic processes. Sedimentary Geology, 344, 14, doi.org/10.1016/j.sedgeo.2016.10.002.Google Scholar
Morsilli, M., Giona Bucci, M., Gliozzi, E., Lisco, S. and Moretti, M. (2020). Sedimentary features influencing the occurrence and spatial variability of seismites (late Messinian, Gargano Promontory, southern Italy). Sedimentary Geology, 401, 105628, doi.org/10.1016/j.sedgeo.2020.105628.Google Scholar
Mörz, T., Karlik, E. A., Kreiter, S. and Kopf, A. (2007). An experiment setup for fluid venting in unconsolidated sediments: new insights to fluid mechanics and structures. Sedimentary Geology, 196, 251267, doi.org/10.1016/j.sedgeo.2006.07.006.Google Scholar
Naik, S. P., Mohanty, A., Porfido, S. et al. (2020). Intensity estimation for the 2001 Bhuj earthquake, India on ESI-07 scale and comparison with historical 16th June 1819 Allah Bund earthquake: a test of ESI-07 application for intraplate earthquakes. Quaternary International, 536, 127143, doi.org/10.1016/j.quaint.2019.12.024.Google Scholar
Nichols, R. J., Sparks, R. S. J. and Wilson, C. J. N. (1994). Experimental studies of fluidization of layered sediments and the formation of fluid escape structures. Sedimentology, 41, 233253, doi.org/10.1111/j.1365-3091.1994.tb01403.x.Google Scholar
Obermeier, S. F. (2009) Using liquefaction-induced and other soft-sediment features for paleoseismic analysis. In McCalpin, J. P., ed., Paleoseismology, Vol. 95. International Geophysics Series, Elsevier, Amsterdam, pp. 497564, doi.org/10.1016/S0074-6142(09)95007-0.Google Scholar
Obermeier, S. F., Olson, S. M. and Green, R. A. (2005). Field occurrences of liquefaction-induced features: a primer for engineering geologic analysis of paleoseismic shaking. Engineering Geology, 76, 209234, doi.org/10.1016/j.enggeo.2004.07.009.Google Scholar
Ogino, Y. and Matsuoka, N. (2007). Involutions resulting from annual freeze – thaw cycles: a laboratory simulation based on observations in northeastern Japan. Permafrost and Periglacial Processes, 18, 323335, doi.org/10.1002/ppp.597.Google Scholar
Ojala, A. E. K., Mattila, J., Virtasalo, J., Kuva, J. and Luoto, T. P. (2018). Seismic deformation of varved sediments in southern Fennoscandia at 7400 cal BP. Tectonophysics, 744, 5871, doi.org/10.1016/j.tecto.2018.06.015.CrossRefGoogle Scholar
Oliveira, C. M. M, Hodgson, D. M. and Flint, S. S. (2009). Aseismic controls on in situ soft-sediment deformation processes and products in submarine slope deposits of the Karoo Basin, South Africa. Sedimentology, 56, 12011225, doi.org/10.1111/j.1365-3091.2008.01029.x.Google Scholar
Owen, L. A. (1991). Mass movement deposits in the Karakoram Mountains: their sedimentary characteristics, recognition and role in Karakoram landform evolution. Zeitschrift für Geomorphologie, 35, 401424.Google Scholar
Owen, G. (2003). Load structures: gravity-driven sediment mobilization in the shallow subsurface. In Van Rensbergen, F., Hillis, R. R., Maltman, A. J. and Morley, C. K., eds., Subsurface Sediment Mobilization. Geological Society, London, Special Publication, Vol. 216, pp. 2134, doi.org/10.1144/GSL.SP.2003.216.01.03.Google Scholar
Owen, G. and Moretti, M. (2008). Determining the origin of soft‐sediment deformation structures: a case study from Upper Carboniferous delta deposits in south‐west Wales, UK. Terra Nova, 20, 237245, doi.org/10.1111/j.1365-3121.2008.00807.x.Google Scholar
Owen, G. and Moretti, M. (2011). Identifying triggers for liquefaction-induced soft-sediment deformation in sands. Sedimentary Geology, 235(3–4), 141147, doi.org/10.1016/j.sedgeo.2010.10.003.Google Scholar
Pedersen, S. A. S. (2014). Architecture of glaciotectonic complexes. Geosciences, 4(4), 269296, doi.org/10.3390/geosciences4040269.Google Scholar
Perucca, L. P., Godoy, E. and Pantano, A. (2014). Late Pleistocene–Holocene earthquake-induced slumps and soft-sediment deformation structures in the Acequion River Valley, Central Precordillera, Argentina. Geologos, 20(2), 147156, doi.org/10.2478/logos-2014-0007.Google Scholar
Phillips, E., Lee, J. R. and Burke, H. (2008). Progressive proglacial to subglacial deformation and syntectonic sedimentation at the margins of the Mid-Pleistocene British Ice Sheet: evidence from north Norfolk, UK. Quaternary Science Reviews, 27, 18481871, doi.org/10.1016/j.quascirev.2008.06.011.Google Scholar
Piotrowski, J. A., Larsen, N. K. and Junge, F. W. (2004). Reflections on soft subglacial beds as a mosaic of deforming and stable spots. Quaternary Science Reviews, 23, 9931000, doi.org/10.1016/j.quascirev.2004.01.006.Google Scholar
Pisarska-Jamroży, M. (2013). Varves and megavarves in the Eberswalde Valley (NE Germany) – a key for the interpretation of glaciolimnic processes. Sedimentary Geology, 291, 8496, doi.org/10.1016/j.sedgeo.2013.03.018.Google Scholar
Pisarska-Jamroży, M. and Woźniak, P. P. (2019). Debris-flow and glacio isostatic-induced soft-sediment deformation structures in a Pleistocene glaciolacustrine fan: the southern Baltic Sea coast, Poland. Geomorphology, 326, 225238, doi.org/10.1016/j.geomorph.2018.01.015.Google Scholar
Pisarska-Jamroży, M. and Zieliński, T. (2012). Specific erosional and depositional processes in a Pleistocene subglacial tunnel in the Wielkopolska region, Poland. Geografiska Annaler, 94A, 429443, doi.org/10.1111/j.1468-0459.2012.00466.x.Google Scholar
Pisarska-Jamroży, M., Belzyt, S., Börner, A. et al. (2018). Evidence from seismites for glacio-isostatically induced crustal faulting in front of an advancing land-ice mass (Rügen Island, SW Baltic Sea). Tectonophysics, 745, 338348, doi.org/10.1016/j.tecto.2018.08.004.Google Scholar
Pisarska-Jamroży, M., Belzyt, S., Bitinas, A., Jusienė, A. and Woronko, B. (2019a). Seismic shocks, periglacial conditions and glaciotectonics as causes of the deformation of a Pleistocene meandering river succession in central Lithuania. Baltica, 32, 6377, doi.org/10.5200/baltica.2019.1.6.Google Scholar
Pisarska-Jamroży, M., Belzyt, S, Börner, A. et al. (2019b). The sea cliff at Dwasieden: soft-sediment deformation structures triggered by glacial isostatic adjustment in front of the advancing Scandinavian Ice Sheet. DEUQUA Special Publications, 2, 6167, doi.org/10.5194/deuquasp-2-61-2019.Google Scholar
Pisarska-Jamroży, M. and Weckwerth, P. (2013). Soft‐sediment deformation structures in a Pleistocene glaciolacustrine delta and their implications for the recognition of sub-environments in delta deposits. Sedimentology, 60, 637665, doi.org/10.1111/j.1365-3091.2012.01354.x.Google Scholar
Rijsdijk, K. F. (2001). Density-driven deformation structures in glacigenic consolidated diamicts: examples from Traeth Y Mwnt, Cardiganshire, Wales, UK. Journal of Sedimentary Research, 71, 122135, doi.org/10.1306/042900710122.Google Scholar
Ringrose, P. S. (1989). Palaeoseismic (?) liquefaction event in late Quaternary lake sediment at Glen Roy, Scotland. Terra Nova, 1, 5762, doi.org/10.1111/j.1365-3121.1989.tb00326.x.Google Scholar
Rodrígues, N., Cobbold, P. R. and Løseth, H. (2009). Physical modeling of sand injectites. Tectonophysics, 474, 610632, doi.org/10.1016/j.tecto.2009.04.032.Google Scholar
Rodríguez-Pascua, M. A., Calvo, J. P., De Vicente, G. and Gómez-Gras, D. (2000). Soft-sediment deformation structures interpreted as seismites in lacustrine sediments of the Prebetic Zone, SE Spain, and their potential use as indicators of earthquake magnitudes during the Late Miocene. Sedimentary Geology, 135, 117135, doi.org/10.1016/S0037-0738(00)00067-1.Google Scholar
Ross, J. A., Peakall, J. and Keevil, G. M. (2011). An integrated model of extrusive sand injectites in cohesionless sediments. Sedimentology, 58, 16931715, doi.org/10.1111/j.1365-3091.2011.01230.x.Google Scholar
Rydelek, P. A. and Tuttle, M. (2004). Explosive craters and soil liquefaction. Nature, 427, 115116, doi.org/10.1038/427115a.Google Scholar
Seilacher, A. (1969). Fault-graded beds interpreted as seismites. Sedimentology, 13, 155159, doi.org/10.1111/j.1365-3091.1969.tb01125.x.Google Scholar
Shanmugam, G. (2016a). The seismite problem. Journal of Palaeogeography, 5, 318362, doi.org/10.1016/j.jop.2016.06.002.Google Scholar
Shanmugam, G. (2016b). Submarine fans: A critical retrospective (1950–2015). Journal of Palaeogeography, 5, 110184, doi.org/10.1016/j.jop.2015.08.011.Google Scholar
Shipton, Z. K., Meghraoui, M. and Monro, L. (2017). Seismic slip on the west flank of the Upper Rhine Graben (France–Germany): evidence from tectonic morphology and cataclastic deformation bands. In Landgraf, A., Kuebler, S., Hintersberger, E. and Stein, S., eds., Seismicity, Fault Rupture and Earthquake Hazards in Slowly Deforming Regions. Geological Society, London, Special Publication, Vol. 432, pp. 147161, doi.org/10.1144/SP432.12.Google Scholar
Strasser, M., Anselmetti, F. S., Fäh, D., Giardini, D. and Schnellmann, M. (2006). Magnitudes and source areas of large prehistoric northern Alpine earthquakes revealed by slope failures in lakes. Geology, 34, 10051008, doi.org/10.1130/G22784A.1.Google Scholar
Suter, F., Martínez, J. I. and Vélez, M. I. (2011). Holocene soft-sediment deformation of the Santa Fe–Sopetrán Basin, northern Colombian Andes: evidence for pre-Hispanic seismic activity? Sedimentary Geology, 235, 188199, doi.org/10.1016/j.sedgeo.2010.09.018.Google Scholar
Sutinen, R., Andreani, L. and Middleton, M. (2019a). Post-Younger Dryas fault instability and deformations on ice lineations in Finnish Lapland. Geomorphology, 326, 202212 doi.org/10.1016/j.geomorph.2018.08.034.Google Scholar
Sutinen, R., Hyvönen, E., Liwata-Kenttälä, P. et al. (2019b). Electrical-sedimentary anisotropy of landforms adjacent to postglacial faults in Lapland. Geomorphology, 326, 213224, doi.org/10.1016/j.geomorph.2018.01.008.Google Scholar
Szuman, I., Ewertowski, M. and Kasprzak, L. (2013). Thermo-mechanical facies representative of fast and slow flowing ice sheets: the Weichselian ice sheet, a central west Poland case study. Proceedings of the Geologists’ Association, 124, 818833, doi.org/10.1016/j.pgeola.2012.09.003.Google Scholar
Tuttle, M. P., Hartleb, R., Wolf, L. and Mayne, P. W. (2019). Paleoliquefaction studies and the evaluation of seismic hazard. Geosciences, 9, 311, doi.org/10.3390/geosciences9070311.Google Scholar
van Balen, R. T., Bakker, M. A. J., Kasse, C., Wallinga, J. and Woolderink, H. A. G. (2019). A Late Glacial surface rupturing earthquake at the Peel Boundary fault zone, Roer Valley Rift System, the Netherlands. Quaternary Science Reviews, 218, 254266, doi.org/10.1016/j.quascirev.2019.06.033.Google Scholar
Van der Wateren, F. M., Kluiving, S. J. and Bartek, L. R. (2000). Kinematic indicators of subglacial shearing. In Maltman, A. J., Hubbard, B. and Hambrey, M. J., eds., Deformation of Glacial Materials. Geological Society, London, Special Publication, Vol. 176, pp. 259278, doi.org/10.1144/GSL.SP.2000.176.01.20.Google Scholar
van Loon, A. J. T. (2009). Soft-sediment deformation structures in siliciclastic sediments: an overview. Geologos, 15, 355.Google Scholar
van Loon, A. J. T. and Pisarska-Jamroży, M. (2014). Sedimentological evidence of Pleistocene earthquakes in NW Poland induced by glacio-isostatic rebound. Sedimentary Geology, 300, 110, doi.org/10.1016/j.sedgeo.2013.11.006.Google Scholar
van Loon, A. J. T., Pisarska-Jamroży, M., Nartišs, M., Krievāns, M. and Soms, J. (2016). Seismites resulting from high-frequency, high-magnitude earthquakes in Latvia caused by Late Glacial glacio-isostatic uplift. Journal of Palaeogeography, 5, 363380, doi.org/10.1016/j.jop.2016.05.002.Google Scholar
van Loon, A. J. T., Soms, J., Nartišs, M., Krievāns, M. and Pisarska-Jamroży, M. (2019). Sedimentological traces of ice-raft grounding in a Weichselian glacial lake near Dukuli (NE Latvia). Baltica, 32, 170181, doi.org/10.5200/baltica.2019.2.4.Google Scholar
van Loon, A. J. T., Pisarska-Jamroży, M. and Woronko, B. (2020). Sedimentological distinction in glacigenic sediments between load casts induced by periglacial processes from those induced by seismic shocks. Geological Quarterly, 64, 626640, doi.org/10.7306/gq.1546.Google Scholar
Van Vliet-Lanoë, B., Magyar, I. A. and Meilliez, F. (2004). Distinguishing between tectonic and periglacial deformations of quaternary continental deposits in Europe. Global and Planetary Change, 43, 103127, doi.org/10.1016/j.gloplacha.2004.03.003.Google Scholar
Vandenberghe, J. (2013). Cryoturbation structures. In Elias, S. A. and Mock, C. J., eds., The Encyclopedia of Quaternary Science, 2nd ed., Elsevier, Amsterdam, pp. 430435, doi.org/10.1016/B978-0-444-53643-3.00096-0.Google Scholar
Vandekerkhove, E., Van Daele, M., Praet, N. et al. (2020). Flood‐triggered versus earthquake‐triggered turbidites: a sedimentological study in clastic lake sediments (Eklutna Lake, Alaska). Sedimentology, 67(1), 364–389, doi.org/10.1111/sed.12646.Google Scholar
Vanneste, K., Meghraoui, M. and Camelbeeck, T. (1999). Late Quaternary earthquake-related soft-sediment deformation along the Belgian portion of the Feldbiss Fault, Lower Rhine Graben system. Tectonophysics, 309, 5779, doi.org/10.1016/S0040-1951(99)00132-8.Google Scholar
Vanneste, K., Camelbeeck, T., Verbeeck, K. and Demoulin, A. (2018). Morphotectonics and past large earthquakes in Eastern Belgium. In Demoulin, A., ed., Landscapes and Landforms of Belgium and Luxembourg, World Geomorphological Landscapes. Springer, Cham, pp. 215236, doi.org/10.1007/978-3-319-58239-9_13.Google Scholar
Wheeler, R. L. (2002). Distinguishing seismic from nonseismic soft-sediment structures: criteria from seismic-hazard analysis. In F. R. Ettensohn, N. Rast and C. E. Brett, eds., Ancient Seismites. Geological Society of America, Special Paper 359, pp. 1–11, doi.org/10.1130/0-8137-2359-0.1.Google Scholar
Weise, O. R. (1983). Das Periglazial. Geomorphologie und Klima in gletscherfreien, kalten Regionen [The Periglacial. Geomorphology and Climate in Glacier-Free, Cold Regions], Gebrüder Borntraeger, Stuttgart.Google Scholar
Winsemann, J., Asprion, U., Meyer, T., Schultz, H. and Victor, P. (2003). Evidence of iceberg ploughing in a subaqueous ice-contact fan, glacial Lake Rinteln, NW Germany. Boreas, 32, 386398, doi.org/10.1111/j.1502-3885.2003.tb01092.x.Google Scholar
Winsemann, J, Alho, P., Laamanen, L. et al. (2016). Flow dynamics, sedimentation and erosion of glacial lake outburst floods along the Middle Pleistocene Scandinavian ice sheet (northern Central Europe). Boreas, 45, 260283, doi.org/10.1111/bor.12146.Google Scholar
Winsemann, J., Koopmann, H., Tanner, D. et al. (2020). Seismic interpretation and structural restoration of the Heligoland glaciotectonic thrust-fault complex: implications for multiple deformation during (pre-)Elsterian to Warthian ice advances into the southern North Sea Basin. Quaternary Science Reviews, 227, 106068, doi.org/10.1016/j.quascirev.2019.106068.Google Scholar
Woronko, B., Belzyt, S., Bujak, Ł. and Pisarska-Jamroży, M. (2018). Glaciotectonically deformed glaciofluvial sediments with ruptured pebbles (the Koczery study site, E Poland). Bulletin of the Geological Society of Finland, 90, 145159.Google Scholar
Worsley, P. (2014). Ice-wedge growth and casting in a Late Pleistocene periglacial, fluvial succession at Baston, Lincolnshire. Mercian Geologist, 18, 159170.Google Scholar
Woźniak, P. P. and Pisarska-Jamroży, M. (2018). Debris flows with soft-sediment clasts in a Pleistocene glaciolacustrine fan (Gdańsk Bay, Poland). Catena, 165, 178191, doi.org/10.1016/j.catena.2018.01.022.Google Scholar

References

Biland, J. and Çöltekin, A. (2017). An empirical assessment of the impact of the light direction on the relief inversion effect in shaded relief maps: NNW is better than NW. Cartography and Geographic Information Science, 44(4), 358372, doi.org/10.1080/15230406.2016.1185647.Google Scholar
Jenness, J. (2013). DEM surface tools. Jenness Enterprises. www.jennessent.com/arcgis/surface_area.htm.Google Scholar
Keefer, D. K. (1984). Landslides caused by earthquakes. Geological Society of America Bulletin, 95(4), 406421.Google Scholar
Keiding, M., Olesen, O. and Dehls, J. (2018). Neotectonic map of Norway and adjacent areas. Geological Survey of Norway, doi.org/10.13140/RG.2.2.32996.48005.Google Scholar
Kuivamäki, A. (1986). Havaintoja Venejärven ja Ruostejärven postglasiaalisista siirroksista. [English abstract: Observations of the Venejärvi and Ruostejärvi Postglacial Faults]. Geological Survey of Finland Espoo, Report YST-52, Espoo, Finland, 20 pp.Google Scholar
Kuivamäki, A., Paananen, M. and Kukkonen, I. (1988). The Pasmajärvi postglacial Fault – a reactivated pre-existing fracture zone. Nordiske geologiske vintermøde: abstracts af foredrag anmeldt til det 18. Nordiske Geologiske Vintermøde 12–14 January 1988, 240, København.Google Scholar
Kuivamäki, A., Vuorela, P. and Paananen, M. (1998). Indications of Postglacial and Recent Bedrock Movements in Finland and Russian Karelia. Geological Survey of Finland Nuclear Waste Disposal Research Report, YST-99, Espoo, Finland, 92 pp.Google Scholar
Kujansuu, R. (1964). Nuorista siirroksista Lapissa. [English summary: Recent faults in Lapland]. Geologi, 6, 3036.Google Scholar
Kujansuu, R. (1972). On landslides in Finnish Lapland. Bulletin de la Commission Géologique de Finlande, 256, 22 pp.Google Scholar
Kukkonen, I. and Kuivamäki, A. (1985). Geologisia ja geofysikaalisia havaintoja Pasmajärven ja Suasselän postglasiaalisista siirroksista. [English abstract: Geological and Geophysical Observations of the Pasmajärvi and Suasselkä Postglacial Faults]. Geological Survey of Finland Report YST-46, Espoo, Finland, 14 pp.Google Scholar
Leonard, M. (2010). Earthquake fault scaling: self-consistent relating of rupture length, width, average displacement, and moment release. Bulletin of the Seismological Society of America, 100, 19711988, doi.org/10.1785/0120090189.Google Scholar
Maanmittauslaitos (2016). Kansallisen maastotietokannan laatumalli, korkeusmallit, 1.3, biturl.top/yUBjQv.Google Scholar
Malamud, B. D., Turcotte, D. L., Guzzetti, F. and Reichenbach, P. (2004). Landslide inventories and their statistical properties. Earth Surface Processes and Landforms, 29, 687711, doi.org/10.1002/esp.1064.Google Scholar
Mattila, J., Ojala, A., Sutinen, R., Palmu, J.-P. and Ruskeeniemi, T. (2016). Digging deeper with LiDAR: vertical slip profiles of post-glacial faults. In LITHOSPHERE 2016 Symposium, November 9–11, 2016, Espoo, Finland, pp. 87–90.Google Scholar
Mattila, J., Ojala, A. E. K., Ruskeeniemi, T. et al. (2019). Evidence of multiple slip events on postglacial faults in northern Fennoscandia. Quaternary Science Reviews, 215, 242252, doi.org/10.1016/j.quascirev.2019.05.022.Google Scholar
Mikko, H., Smith, C. A., Lund, B., Ask, M. V. S. and Munier, R. (2015). LiDAR-derived inventory of post-glacial fault scarps in Sweden. GFF, 137, 334338, doi.org/10.1080/11035897.2015.1036360.Google Scholar
Moss, E. S. and Ross, Z. E. (2011). Probabilistic fault displacement hazard analysis for reverse faults. Bulletin of the Seismological Society of America, 101, 15421553, doi.org/10.1785/0120100248.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2017). Postglacial seismic activity along the Isovaara-Riikonkumpu fault complex. Global and Planetary Change, 157, 5972, doi.org/10.1016/j.gloplacha.2017.08.015.Google Scholar
Ojala, A. E. K., Markovaara-Koivisto, M., Middleton, M. et al. (2018). Dating of seismically-induced paleolandslides in western Finnish Lapland. Earth Surface Processes and. Landforms, 43, 24492462, doi.org/10.1002/esp.4408.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2019a). Postglacial Faults in Finland – A Review of PGSdyn Project Results. Posiva Report 2019-01, 118 pp., Posiva Oy, Eurajoki.Google Scholar
Ojala, A. E. K., Mattila, J., Markovaara-Koivisto, M. et al. (2019b). Distribution and morphology of landslides in northern Finland: an analysis of postglacial seismic activity. Geomorphology, 326, 190201, doi.org/10.1016/j.geomorph.2017.08.045.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2019c). Postglacial reactivation of the Suasselkä PGF complex in SW Finnish Lapland. International Journal of Earth Sciences, 108, 10491065, doi.org/10.1007/s00531-019-01695-w.Google Scholar
Palmu, J.-P., Ojala, A. E. K., Ruskeeniemi, T., Sutinen, R. and Mattila, J. (2015). LiDAR DEM detection and classification of postglacial faults and seismically-induced landforms in Finland: a paleoseismic database. GFF, 137, 344352, doi.org/10.1080/11035897.2015.1068370.Google Scholar
Smith, M. J. and Clark, C. D. (2005). Methods for the visualization of digital elevation models for landform mapping. Earth Surface Processes and Landforms, 30, 885900, doi.org/10.1002/esp.1210.Google Scholar
Smith, C. A., Sundh, M. and Mikko, H. (2014). Surficial geology indicates early Holocene faulting and seismicity, central Sweden. International Journal of Earth Sciences, 103(6), 17111724, doi.org/10.1007/s00531-014-1025-6.Google Scholar
Sutinen, R. (2005). Timing of early Holocene landslides in Kittilä, Finnish Lapland. Geological Survey of Finland, Special Paper, 40, 5358.Google Scholar
Sutinen, R., Piekkari, M. and Liwata, P. (2007). Time-transgressive evolution of landslides possibly induced by seismotectonic events in Lapland. Applied Quaternary research in the central part of glaciated terrain. Geological Survey of Finland, Special Paper, 46, 121128.Google Scholar
Sutinen, R., Hyvönen, E. and Kukkonen, I. (2014a). LiDAR detection of paleolandslides in the vicinity of the Suasselkä postglacial fault, Finnish Lapland. International Journal of Applied Earth Observation and Geoinformation, 27, 9199, doi.org/10.1016/j.jag.2013.05.004.CrossRefGoogle Scholar
Sutinen, R., Hyvönen, E., Middleton, M. and Ruskeeniemi, T. (2014b). Airborne LiDAR detection of postglacial faults and Pulju moraine in Palojärvi, Finnish Lapland. Global and Planetary Change, 115, 2432, doi.org/10.1016/j.gloplacha.2014.01.007.Google Scholar
Verduzco, B., Fairhead, J. D., Green, C. M. and MacKenzie, C. (2004). New insights into magnetic derivatives for structural mapping. The Leading Edge, 23(2), 116119, doi.org/10.1190/1.1651454.Google Scholar
Wells, D. L. and Coppersmith, K. J. (1994). New empirical relationships among magnitude, rupture length, rupture width, rupture area, and surface displacement. Bulletin of the Seismological Society of America, 84, 9741002.Google Scholar

References

Aki, K. and Richards, P. G. (1980). Quantitative Seismology: Theory and Methods. W. H. Freeman, San Francisco.Google Scholar
Barth, L. and Plenefisch, T. (2020). Focal mechanisms for small to intermediate earthquakes in the northern part of the Alps and their seismotectonic interpretation. EGU, Vienna, doi.org/10.5194/egusphere-egu2020-12066.Google Scholar
Bormann, P. and Dewey, J. W. (2014). IS 3.3: The new IASPEI standards for determining magnitudes from digital data and their relation to classical magnitudes. In P. Bormann, ed., New Manual of Seismological Observatory Practice (NMSOP-2). Deutsches GeoForschungsZentrum GFZ, Potsdam, pp. 1–44, doi.org/10.2312/GFZ.NMSOP-2_IS_3.3.Google Scholar
Brandes, C., Plenefisch, T., Tanner, D. C., Gestermann, N. and Steffen, H. (2019). Evaluation of deep crustal earthquakes in northern Germany – Possible tectonic causes. Terra Nova, 31, 8393, doi.org/10.1111/ter.12372.Google Scholar
Dahm, T. and Krüger, F. (2014). IS 3.9: Moment tensor inversion and moment tensor interpretation. In P. Bormann, ed., New Manual of Seismological Observatory Practice 2 (NMSOP-2). Deutsches GeoForschungsZentrum GFZ, Potsdam, pp. 1–37, doi.org/10.2312/GFZ.NMSOP-2_IS_3.9.Google Scholar
Di Stefano, R., Aldersons, F., Kissling, E. et al. (2006). Automatic seismic phase picking and consistent observation error assessment: application to the Italian seismicity. Geophysical Journal International, 165(1), 121134, doi.org/10.1111/j.1365-246X.2005.02799.x.Google Scholar
Donner, S., Igel, H., Hadziioannou, C. and the Romy group (2018). Retrieval of the seismic moment tensor from joint measurements of translational and rotational ground motions: sparse networks and single stations. In D’Amico, S., ed., Moment Tensor Solutions. Springer, Cham, pp. 263280, doi.org/10.1007/978-3-319-77359-9_12.Google Scholar
Duncan, P. M. (2005). Is there a future for passive seismic? First Break, 23(6), 111115.Google Scholar
Efron, B. (1980). The Jackknife, the Bootstrap, and Other Resampling Plans. Stanford University, Department of Statistics, Technical Report, NSF 163, 135 pp.Google Scholar
Geiger, L. (1910). Herdbestimmung bei Erdbeben aus den Ankunftszeiten. Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse, June 1910, 331–349; trans. (1912) Probability method for the determination of earthquake epicentres from the arrival time only. Bulletin of St. Louis University, 8(1), 56–71, eudml.org/doc/58769.Google Scholar
Gutenberg, B. and Richter, C. F. (1956). Magnitude and energy of earthquakes. Annali di Geofisica, 9, 115.Google Scholar
Harjes, H.-P. (1990). Design and siting of a new regional array in Central Europe. Bulletin of the Seismological Society of America, 80(6), 18011817.Google Scholar
Havskov, J. and Ottemöller, L. (2010). Routine Data Processing in Earthquake Seismology. Springer, Dordrecht, doi.org/10.1007/978-90-481-8697-6.Google Scholar
Jost, M. and Herrmann, R. (1989). A student’s guide to and review of moment tensor. Seismological Research Letters, 60(2), 3757, doi.org/10.1785/gssrl.60.2.37.Google Scholar
Joswig, M. (1992). System architecture of seismic networks and its implications to network automatization. Cahiers Centre Européen de Géodynamique et de Séismologie, 5, 7584.Google Scholar
Keiding, M., Kreemer, C., Lindholm, C. et al. (2015). A comparison of strain rates and seismicity for Fennoscandia: depth dependency of deformation from glacial isostatic adjustment. Geophysical Journal International, 202, 10211028, doi.org/10.1093/gji/ggv207.Google Scholar
Kilb, D. and Rubin, A. M. (2002). Implications of diverse fault orientations imaged in relocated aftershocks of the Mount Lewis, ML 5.7, California, earthquake. Journal of Geophysical Research, 107(B11), 2294, doi.org/10.1029/2001JB000149.Google Scholar
Kraft, T., Mignan, A. and Giardini, D. (2013). Optimization of a large-scale microseismic monitoring network in northern Switzerland. Geophysical Journal International, 195(1), 474490, doi.org/10.1093/gji/ggt225.Google Scholar
Li, L., Tan, J., Schwarz, B. et al. (2020). Recent advances and challenges of waveform‐based seismic location methods at multiple scales. Reviews of Geophysics, 58, e2019RG000667, doi.org/10.1029/2019RG000667.Google Scholar
Lomax, A., Virieux, J., Volant, P. and Berge, C. (2000). Probabilistic earthquake location in 3D and layered models: introduction of a Metropolis–Gibbs method and comparison with linear locations. In Thurber, C. H. and Rabinowitz, N., eds., Advances in Seismic Event Location. Kluwer, Amsterdam, pp. 101134, doi.org/10.1007/978-94-015-9536-0_5.Google Scholar
Meier, M.-A., Ross, Z. E., Ramachandran, A. et al. (2019). Reliable real‐time seismic signal/noise discrimination with machine learning. Journal of Geophysical Research: Solid Earth, 124(1), 788800, doi.org/10.1029/2018JB016661.Google Scholar
Omori, F. (1894). On after-shocks of earthquakes. The Journal of the College of Science, Imperial University of Tokyo, Japan, 7, 111200.Google Scholar
Provost, F., Hibert, C. and Malet, J.-P. (2017). Automatic classification of endogenous landslide seismicity using the Random Forest supervised classifier. Geophysical Research Letters, 44, 113120, doi.org/10.1002/2016GL070709.Google Scholar
Riggelsen, C. and Ohrnberger, M. (2014). A machine learning approach for improving the detection capabilities at 3C seismic stations. Pure and Applied Geophysics, 171, 395411, doi.org/10.1007/s00024–012-0592-3.Google Scholar
Rost, S. and Thomas, C. (2002). Array seismology: methods and applications. Reviews of Geophysics, 40(3), 2-1–2-27, doi.org/10.1029/2000RG000100.Google Scholar
Schmelzbach, C., Donner, S., Igel, H. et al. (2018). Advances in 6C seismology: applications of combined translational and rotational motion measurements in global and exploration seismology. Geophysics, 83(3), doi.org/10.1190/geo2017-0492.1.Google Scholar
Schuster, G. T., Yu, J. and Sheng, J. (2004). Interferometric/daylight seismic imaging. Geophysical Journal International, 157(2), 838852, doi.org/10.1111/j.1365-246X.2004.02251.x.Google Scholar
Schweitzer, J. (2001). HYPOSAT – an enhanced routine to locate seismic events. Pure and Applied Geophysics, 158, 277289, doi.org/10.1007/978-3-0348-8250-7_17.Google Scholar
Schweitzer, J., Fyen, J., Mykkeltveit, S. et al. (2012 online): seismic arrays. In Bormann, P., ed., New Manual of Seismological Observatory Practice 2 (NMSOP-2). Deutsches GeoForschungsZentrum GFZ, Potsdam, pp. 180, doi.org/10.2312/GFZ.NMSOP-2_ch9.Google Scholar
Shearer, P. M. (2009). Introduction to Seismology. Cambridge University Press, Cambridge.Google Scholar
Sick, B., Walter, M. and Joswig, M. (2014). Visual event screening of continuous seismic data by supersonograms. Pure and Applied Geophysics, 171, 549559, doi.org/10.1007/s00024-012-0618-xGoogle Scholar
Snoke, J. A. (2003). FOCMEC: FOCal MEChanism determinations. International Handbook of Earthquake and Engineering Seismology, pp. 1629–1630.Google Scholar
Stein, S. and Wysession, M. (2003). An Introduction to Seismology, Earthquakes, and Earth Structure. Blackwell Publishing, Malden, Massachusetts.Google Scholar
Tanner, D. and Brandes, C. (2019). Understanding Faults: Detecting, Dating, and Modeling. Elsevier, Amsterdam, doi.org/10.1016/B978-0-12-815985-9.00001-1.Google Scholar
Udías, A. and Buforn, E. (2017). Principles of Seismology. Cambridge University Press, Cambridge.Google Scholar
Waldhauser, F. and Ellsworth, W. E. (2000). A double-difference earthquake location algorithm: method and application to the Northern Hayward Fault, California. Bulletin of the Seismological Society of America, 90(6), 13531368, doi.org/10.1785/0120000006.Google Scholar
Wells, D. L. and Coppersmith, K. J. (1994). New empirical relationships among magnitude, rupture length, rupture width, rupture area, and surface displacement. Bulletin of the Seismological Society of America, 84, 9741002.Google Scholar

References

Abdi, A., Heinonen, S., Juhlin, C. and Karinen, T. (2015). Constraints on the geometry of the Suasselkä post-glacial fault, northern Finland, based on reflection seismic imaging. Tectonophysics, 649, 130138, doi.org/10.1016/j.tecto.2015.03.004.Google Scholar
Ahmadi, O., Juhlin, C., Ask, M. V. S. and Lund, B. (2015). Revealing the deeper structure of the endglacial Pärvie fault system in northern Sweden by seismic reflection profiling. Solid Earth, 6, 621632 doi.org/10.5194/se-6-621-2015.Google Scholar
Beckel, R. A. and Juhlin, C. (2019). The cross-dip correction as a tool to improve imaging of crooked-line seismic data: a case study from the post-glacial Burträsk fault, Sweden. Solid Earth, 10, 581598, doi.org/10.5194/se-10-581-2019.Google Scholar
Dalsegg, E. and Olesen, O. (2014). Resistivitetsmålinger ved Masi, Fiednajohka og Riednajavre og implikasjoner for malmleting, Kautokeino kommune, Finnmark. Report 2014.021, Geological Survey of Norway, Trondheim, Norway.Google Scholar
Henkel, H. (1987). Tectonic Studies in the Lansjärv Region. SKB Technical Report TR 88-07, Swedish Nuclear Fuel and Waste Management Co., Stockholm, 80 pp.Google Scholar
Henkel, H. and Guzmán, M. (1977). Magnetic features of fracture zones. Geoexploration, 15(3), 173181.Google Scholar
Henkel, H., Hult, K., Eriksson, L. and Johansson, L. (1983). Neotectonics in Northern Sweden – Geophysical Investigations. SKB Technical Report TR 83-57, Swedish Nuclear Fuel and Waste Management Co., Stockholm, Sweden.Google Scholar
Juhlin, C. (1995). Imaging of fracture zones in the Finnsjön area, central Sweden, using the seismic reflection method. Geophysics, 60(1), 6675, doi.org/10.1190/1.1443764.Google Scholar
Juhlin, C. and Lund, B. (2011). Reflection seismic studies over the end-glacial Burträsk fault, Skellefteå, Sweden. Solid Earth, 2, 916, doi.org/10.5194/se-2-9-2011.Google Scholar
Juhlin, C., Dehghannejad, M., Lund, B., Malehmir, A. and Pratt, G. (2010). Reflection seismic imaging of the end-glacial Pärvie Fault system, northern Sweden. Journal of Applied Geophysics, 70, 307316, doi.org/10.1016/j.jappgeo.2009.06.004.Google Scholar
Kuivamäki, A., Vuorela, P. and Paananen, M. (1998). Indications of Postglacial and Recent Bedrock Movements in Finland and Russian Karelia. Geological Survey of Finland, Nuclear Waste Disposal Research Report YST-99, Espoo, Finland, 92 pp.Google Scholar
Kujansuu, R. (1964). Nuorista sirroksista Lapissa [English summary: Recent faults in Lapland]. Geologi, 16, 3036 (in Finnish).Google Scholar
Kukkonen, I., Lahti, I., Heikkinen, P. et al. (2009). HIRE Seismic Reflection Survey in the Suurikuusikko Gold Mining and Exploration Area, North Finland. Report Q 23/2009/28, Geological Survey of Finland.Google Scholar
Lagerbäck, R. (1978). Neotectonic structures in northern Sweden. Geologiska Föreningen i Stockholm Förhandlingar, 100(3), 263269, doi.org/10.1080/11035897809452533.Google Scholar
Lindblom, E., Lund, B., Tryggvason, A. et al. (2015). Microearthquakes illuminate the deep structure of the endglacial Pärvie fault, northern Sweden. Geophysical Journal International, 201, 17041716, doi.org/10.1093/gji/ggv112.Google Scholar
Mair, J. A. and Green, A. G. (1981). High-resolution seismic reflection profiles reveal fracture zones within a ‘homogeneous’ granite batholith. Nature, 294, 439442, doi.org/10.1038/294439a0.Google Scholar
Malehmir, A., Andersson, M., Mehta, S. et al. (2016). Post-glacial reactivation of the Bollnäs fault, central Sweden – a multidisciplinary geophysical investigation. Solid Earth, 7, 509527, doi.org/10.5194/se-7-509-2016.Google Scholar
Mauring, E., Olesen, O., Rønning, J. S. and Tønnesen, J. F. (1997). Ground-Penetrating Radar Profiles across Postglacial Fault at Kåfjord, Troms and Fidnajohka, Finnmark. Report 97.174, Geological Survey of Norway, Trondheim, Norway.Google Scholar
McDowell, P. W. (1979). Geophysical mapping of water-filled fracture zones in rocks. Bulletin of the International Association of Engineering Geology, 19(1), 258264, doi.org/10.1007/BF02600485.Google Scholar
Mikko, H., Smith, C. A., Lund, B., Ask, M. V. S. and Munier, R. (2015). LiDAR-derived inventory of post-glacial fault scarps in Sweden. GFF, 137(4), 334338, doi.org/10.1080/11035897.2015.1036360.Google Scholar
Moos, D. and Zoback, M. D. (1983). In situ studies of velocity in fractured crystalline rocks. Journal of Geophysical Research, 88(B3), 23452358, doi.org/10.1029/JB088iB03p02345.Google Scholar
Mrope, F. M., Becken, M., Ruud, B. O. et al. (2019). Magnetotelluric 2D Inversion and Joint Interpretation of MT, Seismic, Magnetic and Gravity Data from Masi, Kautokeino Municipality, Finnmark. Report 2019.009, Geological Survey of Norway, Trondheim, Norway.Google Scholar
Olesen, O., Henkel, H., Lile, O. B., Mauring, E. and Rønning, J. S. (1992). Geophysical investigations of the Stuoragurra postglacial fault, Finnmark, northern Norway. Journal of Applied Geophysics, 29, 95118, doi.org/10.1016/0926-9851(92)90001-2.Google Scholar
Olsen, L., Olesen, O., Dehls, J. and Tassis, G. (2018). Late-/postglacial age and tectonic origin of the Nordmannvikdalen Fault, northern Norway. Norwegian Journal of Geology, 98, 483500, doi.org/10.17850/njg98-3-09.Google Scholar
Roberts, D., Olesen, O. and Karpuz, M. R. (1997). Seismo- and neotectonics in Finnmark, Kola Peninsula and the southern Barents Sea. Part 1: geological and neotectonic framework. Tectonophysics, 270, 113, doi.org/10.1016/S0040-1951(96)00173-4.Google Scholar
Schön, J. H. (2011). Physical Properties of Rocks, Vol. VIII of Handbook of Petroleum Exploration and Production. Elsevier, Oxford.Google Scholar
Sheriff, R. E. and Geldart, L. P. (1995). Exploration Seismology, 2nd ed., Cambridge University Press, Cambridge, doi.org/10.1017/CBO9781139168359.Google Scholar
Smith, C. A., Sundh, M. and Mikko, H. (2014). Surficial geology indicates early Holocene faulting and seismicity, central Sweden. International Journal of Earth Sciences, 103, 17111724, doi.org/10.1007/s00531-014-1025-6.Google Scholar

References

Benavente, C., Zerathe, S., Audin, L. et al. (2017). Active transpressional tectonics in the Andean forearc of southern Peru quantified by 10Be surface exposure dating of an active fault scarp. Tectonics, 36(9), 16621678, doi.org/10.1002/2017TC004523.Google Scholar
Benedetti, L., Manighetti, I., Gaudemer, Y. et al. (2013). Earthquake synchrony and clustering on Fucino faults (Central Italy) as revealed from in situ 36Cl exposure dating. Journal of Geophysical Research: Solid Earth, 118(9), 49484974, doi.org/10.1002/jgrb.50299.Google Scholar
Berglund, M. (2012). Early Holocene in Gästrikland, east central Sweden: shore displacement and isostatic recovery. Boreas, 41(2), 263276, doi.org/10.1111/j.1502-3885.2011.00228.x.Google Scholar
Berthet, T., Ritz, J.-F., Ferry, M. et al. (2014). Active tectonics of the eastern Himalaya: new constraints from the first tectonic geomorphology study in southern Bhutan. Geology, 42(5), 427430, doi.org/10.1130/G35162.1.Google Scholar
Björck, S. (1995). A review of the history of the Baltic Sea, 13.0–8.0 ka BP. Quaternary International, 27(94), 1940, doi.org/10.1016/1040-6182(94)00057-C.Google Scholar
Crozier, M. J. (1992). Determination of paleoseismicity from landslides. In Bell, D.H., ed., Landslides (Glissements de terrain). Proceedings of the 6th International Symposium, Christchurch, New Zealand, A. A. Balkema, Rotterdam, pp. 11731180.Google Scholar
Cruden, D. M. and Varnes, D. J. (1996). Landslide types and processes. In Turner, A. K. and Schuster, R. L., eds., Landslides: Investigation and Mitigation. Transportation Research Board, US National Research Council Special Report 247, Washington, DC, pp. 3675.Google Scholar
Gosse, J. C. and Phillips, F. M. (2001). Terrestrial in situ cosmogenic nuclides: theory and application. Quaternary Science Reviews, 20, 14751560, doi.org/10.1016/S0277-3791(00)00171-2.Google Scholar
Hughes, A. L. C., Gyllencreutz, R., Lohne, Ø. S., Mangerud, J. and Svendsen, J. I. (2016). The last Eurasian ice sheets – a chronological database and time-slice reconstruction, DATED-1. Boreas, 45, 145, doi.org/10.1111/bor.12142.Google Scholar
Hungr, O., Leroueil, S. and Picarelli, L. (2014). The Varnes classification of landslide types, an update. Landslides, 11, 167194, doi.org/10.1007/s10346–013-0436-y.Google Scholar
Jibson, R. W. (1996). Use of landslides for paleoseismic analysis. Engineering Geology, 43, 291323, doi.org/10.1016/S0013–7952(96)00039-7.Google Scholar
Jibson, R. W. and Keefer, D. K. (1989). Statistical analysis of factors affecting landslide distribution in the New Madrid seismic zone, Tennessee and Kentucky. Engineering Geology, 27, 509542, doi.org/10.1016/0013-7952(89)90044-6.Google Scholar
Jibson, R. W. and Keefer, D. K. (1993). Analysis of the seismic orgin of landslides—examples from the New Madrid seismic zone. Geological Society of America Bulletin, 105, 421436.Google Scholar
Kujansuu, R. (1964). Nuorista siirroksista Lapissa [English summary: Recent faults in Lapland]. Geologi, 16, 3036 (in Finnish).Google Scholar
Lagerbäck, R. (1978). Neotectonic structures in northern Sweden. GFF, 100(3), 263269, doi.org/10.1080/11035897809452533.Google Scholar
Lagerbäck, R. (1988). The Veiki moraines in northern Sweden – widespread evidence of an Early Weichselian deglaciation. Boreas, 17, 469486, doi.org/10.1111/j.1502-3885.1988.tb00562.x.Google Scholar
Lagerbäck, R. (1990). Late Quaternary faulting and paleoseismicity in northern Fennoscandia with particular reference to the Lansjärv area, Northern Sweden. GFF, 112, 333354, doi.org/10.1080/11035899009452733.Google Scholar
Lagerbäck, R. (1992). Dating of Late Quaternary faulting in northern Sweden. Journal of the Geological Society, 149(2), 285291, doi.org/10.1144/gsjgs.149.2.0285.Google Scholar
Lagerbäck, R. and Robertsson, A.-M. (1988). Kettle holes – stratigraphical archives for Weichselian geology and palaeoenvironment in northernmost Sweden. Boreas, 17, 439468, doi.org/10.1111/j.1502-3885.1988.tb00561.x.Google Scholar
Lagerbäck, R. and Sundh, M. (2008). Early Holocene Faulting and Paleoseismicity in Northern Sweden: Research Paper C 836. Geological Survey of Sweden pp.Google Scholar
Lindén, M., Möller, P., Björck, S. and Sandgren, P. (2006). Holocene shore displacement and deglaciation chronology in Norrbotten, Sweden. Boreas, 35(1), 122, doi.org/10.1111/j.1502-3885.2006.tb01109.x.Google Scholar
Lundqvist, J. and Lagerbäck, R. (1976). The Pärve Fault: a late-glacial fault in the Precambrian of Swedish Lapland. Geologiska Föreningens i Stockholm Förhandlingar, 98, 4551, doi.org/10.1080/11035897609454337.Google Scholar
Mattila, J., Ojala, A. E. K., Ruskeeniemi, T. et al. (2019). Evidence of multiple slip events on postglacial faults in northern Fennoscandia. Quaternary Science Reviews, 215, 242252, doi.org/10.1016/j.quascirev.2019.05.022.Google Scholar
Mikko, H., Smith, C. A., Lund, B., Ask, M. V. S. and Munier, R. (2015). LiDAR-derived inventory of post-glacial fault scarps in Sweden. GFF, 137, 334338, doi.org/10.1080/11035897.2015.1036360.Google Scholar
Ojala, A. E. K. and Alenius, T. (2005). 10000 years of interannual sedimentation recorded in the Lake Nautajärvi (Finland) clastic–organic varves. Palaeogeography, Palaeoclimatology, Palaeoecology, 219(3), 285302, doi.org/10.1016/j.palaeo.2005.01.002.Google Scholar
Ojala, A. E. K., Palmu, J.-P., Åberg, A., Åberg, S. and Virkki, H. (2013). Development of an ancient shoreline database to reconstruct the Litorina Sea maximum extension and the highest shoreline of the Baltic Sea basin in Finland. Bulletin of the Geological Society of Finland, 85(PART 2), 127144, doi.org/10.17741/bgsf/85.2.002.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2017). Postglacial seismic activity along the Isovaara–Riikonkumpu fault complex. Global and Planetary Change, 157(January), 5972, doi.org/10.1016/j.gloplacha.2017.08.015.Google Scholar
Ojala, A. E. K., Markovaara-Koivisto, M., Middleton, M. et al. (2018a). Dating of paleolandslides in western Finnish Lapland. Earth Surface Processes and Landforms, 43, 24492462, doi.org/10.1002/esp.4408.Google Scholar
Ojala, A. E. K., Mattila, J., Virtasalo, J. Kuva, J. and Luoto, T.P. (2018b). Seismic deformation of varved sediments in southern Fennoscandia at 7400 cal BP. Tectonophysics, 744, 5871, doi.org/10.1016/j.tecto.2018.06.015.Google Scholar
Ojala, A. E. K., Mattila, J., Hämäläinen, J. and Sutinen, R. (2019a). Lake sediment evidence of paleoseismicity: timing and spatial occurrence of late- and postglacial earthquakes in Finland. Tectonophysics, 771, 228227, doi.org/10.1016/j.tecto.2019.228227.Google Scholar
Ojala, A. E. K., Mattila, J., Markovaara-Koivisto, M. et al. (2019b). Distribution and morphology of landslides in northern Finland: an analysis of postglacial seismic activity. Geomorphology, 326, 190201, doi.org/10.1016/j.geomorph.2017.08.045.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2019c). Postglacial Faults in Finland – A Review of PGSdyn – Project Results, Posiva Report 2019-1, 118 pp., Posiva Oy, Eurajoki.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2019d). Postglacial reactivation of the Suasselkä PGF complex in SW Finnish Lapland. International Journal of Earth Sciences, 108(3), 10491065, doi.org/10.1007/s00531-019-01695-w.Google Scholar
Owen, G., Moretti, M. and Alfaro, P. (2011). Recognising triggers for soft-sediment deformation: current understanding and future directions. Sedimentary Geology, 235, 133140, doi.org/10.1016/j.sedgeo.2010.12.010.Google Scholar
Palmu, J. P., Ojala, A. E. K., Ruskeeniemi, T., Sutinen, R. and Mattila, J. (2015). LiDAR DEM detection and classification of postglacial faults and seismically-induced landforms in Finland: a paleoseismic database. GFF, 137(4), 344352, doi.org/10.1080/11035897.2015.1068370.Google Scholar
Salomaa, R. (1982). Post-glacial shoreline displacement in the Lauhanvuori area, Western Finland. Annales Academiæ Scientiarum Fennicæ A III, 134, 8197.Google Scholar
Sims, J. D. (1973). Earthquake-Induced Structures in Sediments of Van Norman Lake, San Fernando, California. Science, 182, 161163, doi.org/10.1126/science.182.4108.161.Google Scholar
Smith, C. A., Sundh, M. and Mikko, H. (2014). Surficial geology indicates early Holocene faulting and seismicity, central Sweden. International Journal of Earth Sciences, 103(6), 17111724, doi.org/10.1007/s00531-014-1025-6.Google Scholar
Smith, C. A., Nyberg, J. and Bergman, B. (2018). Comparison between hydroacoustical and terrestrial evidence of glacially induced faulting, Lake Voxsjön, central Sweden. International Journal of Earth Sciences, 107, 169175, doi.org/10.1007/s00531-017-1479-4.Google Scholar
Stroeven, A. P., Hättestrand, C., Kleman, J., et al. (2016). Deglaciation of Fennoscandia. Quaternary Science Reviews, 147, 91121, doi.org/10.1016/j.quascirev.2015.09.016.Google Scholar
Strömberg, B. (1989) Late Weichselian Deglaciation and Clay Varve Chronology in East-Central Sweden, Geological Survey of Sweden (SGU) Series Ca. 73, 70 pp.Google Scholar
Sutinen, R., Hyvönen, E., Middleton, M. and Ruskeeniemi, T. (2014). Airborne LiDAR detection of postglacial faults and Pulju moraine in Palojärvi, Finnish Lapland. Global and Planetary Change, 115, 2432, doi.org/10.1016/j.gloplacha.2014.01.007.Google Scholar
Tesson, J. and Benedetti, L. (2019). Seismic history from in situ 36Cl cosmogenic nuclide data on limestone fault scarps using Bayesian reversible jump Markov chain Monte Carlo. Quaternary Geochronology, 52, 120, doi.org/10.1016/j.quageo.2019.02.004.Google Scholar
Tikhomirov, D., Amiri, N. M., Ivy-Ochs, S. et al. (2019). Fault Scarp Dating Tool – a MATLAB code for fault scarp dating using in-situ chlorine-36 supplemented with datasets of Yavansu and Kalafat faults. Data in Brief, 26, 104476, doi.org/10.1016/j.dib.2019.104476.Google Scholar
Tiljander, M., Saarnisto, M., Ojala, A. E. K. and Saarinen, T. (2003). A 3000-year palaeoenvironmental record from annually laminated sediment of Lake Korttajärvi, central Finland. Boreas, 32(4), 566577, doi.org/10.1111/j.1502-3885.2003.tb01236.x.Google Scholar
Varnes, D. J. (1978). Slope movement types and processes. In Schuster, R. L. and Krizek, R. J., eds., Landslides: Analysis and Control, Special Report 176, Transportation Research Board, National Academy of Sciences, Washington, DC., pp. 1133.Google Scholar
Yang, H., Yang, X., Huang, X. et al. (2018). New constraints on slip rates of the Fodongmiao–Hongyazi fault in the Northern Qilian Shan, NE Tibet, from the 10Be exposure dating of offset terraces. Journal of Asian Earth Sciences, 151, 131147, doi.org/10.1016/j.jseaes.2017.10.034.Google Scholar

References

Ahmadi, O., Juhlin, C., Ask, M. V. S. and Lund, B. (2015). Revealing the deeper structure of the end-glacial Pärvie fault system in northern Sweden by seismic reflection profiling. Solid Earth, 6, 621632, doi.org/10.5194/se-6-621-2015.Google Scholar
Andersson, J. B. H. (2019). Structural Evolution of Two Ore-Bearing Palaeoproterozoic Metasupracrustal Belts in the Kiruna Area, Northwestern Fennoscandian Shield. Licentiate thesis, Luleå University of Technology, Sweden, 91 pp.Google Scholar
Araki, E., Saffer, D. M., Kopf, A. et al. (2017). Recurring and triggered slow-slip events near the trench at the Nankai Trough subduction megathrust. Science, 356, 11571160, doi.org/10.1126/science.aan3120.Google Scholar
Arvidsson, R. (1996). Fennoscandian earthquakes: whole crustal rupturing related to postglacial rebound. Science, 274, 744746, doi.org/10.1126/science.274.5288.744.Google Scholar
Bäckström, A., Giulio, V., Rantakokko, N., Jonsson, E. and Ask, M. (2013). Preliminary Results from Fault-Slip Analysis of the Pärvie Neotectonic Postglacial Fault Zone, Northern Sweden. EGU General Assembly 2013, 7–12 April 2013 in Vienna, Austria, id. EGU2013–1751.Google Scholar
Bauer, T., Andersson, J., Sarlus, Z., Lund, C. and Kearney, T. (2018). Structural controls on the setting, shape and hydrothermal alteration of the Malmberget IOA deposit, northern Sweden. Economic Geology, 113(2), 377395, doi.org/10.5382/econgeo.2018.4554.Google Scholar
Bergman, S., Kübler, L. and Martinsson, O. (2001). Regional geological and geophysical maps of northern Norrbotten County: bedrock map (east of the Caledonian orogen). Sveriges Geologiska Undersökning, Ba 56.Google Scholar
Bohnhoff, M., Dresen, G., Ceken, U. et al. (2017). GONAF – the borehole Geophysical Observatory at the North Anatolian Fault in the eastern Sea of Marmara. Scientific Drilling, 22, 1928, doi.org/10.5194/sd-22-19-2017.Google Scholar
Calais, E., Camelbeeck, T., Stein, S., Liu, M. and Craig, T. J. (2016). A new paradigm for large earthquakes in stable continental plate interiors. Geophysical Research Letters, 43, 1062110637, doi.org/10.1002/2016GL070815.Google Scholar
Campell, D. L. (1978). Investigation of the stress-concentration mechanism for intraplate earthquakes. Geophysical Research Letters, 5, 477479.Google Scholar
Carlson, L. and Lundqvist, A. (1984). Laukujärvi kopparfyndighet. Prospekteringsförslag 1984 och geologisk sammanfattning [Laukujärvi copper deposit. Exploration proposal 1984 and geological summary]. Sveriges Geologiska AB, PRAP 84067, 57 pp. (in Swedish).Google Scholar
Claesson, L.-Å. and Nilsson, G. (2005). Forsmark Site investigation. Drilling of the Borehole KFM01B at Drilling Site DS1. SKB Report P-04-302, Swedish Nuclear Fuel and Waste Management Co., Stockholm, 32 pp.Google Scholar
Claesson Liljedahl, L., Kontula, A., Harper, J. et al. (2016). The Greenland Analogue Project: Final Report, SKB Technical Report TR-14-13, Swedish Nuclear Fuel and Waste Management Co., Stockholm, 142 pp.Google Scholar
Costain, J. K. (2017). Groundwater recharge as the trigger of naturally occurring intraplate earthquakes. In Landgraf, A., Kuebler, S., Hintersberger, E. and Stein, S., eds., Seismicity, Fault Rupture and Earthquake Hazards in Slowly Deforming Regions. Geological Society, London, Special Publication, Vol. 432, pp. 91–118, doi.org/10.1144/SP432.9.Google Scholar
Doughty, C., Tsang, C.-F., Rosberg, J.-E. et al. (2017). Flowing fluid electrical conductivity logging of a deep borehole during and following drilling: estimation of transmissivity, water salinity and hydraulic head of conductive zones. Hydrogeology Journal, 25(2): 501517, doi.org/10.1007/s10040–016-1497-5.Google Scholar
England, R. W. and Ebbing, J. (2012). Crustal structure of central Norway and Sweden from integrated modelling of teleseismic receiver functions and the gravity anomaly. Geophysical Journal International, 191(1), 111, doi.org/10.1111/j.1365-246X.2012.05607.x.Google Scholar
England, P. and Jackson, J. (2011). Uncharted seismic risk. Nature Geoscience, 4, 348349, doi.org/10.1038/ngeo1168.Google Scholar
Gaál, G. and Gorbatschev, R. (1987). Precambrian geology and of the Central Baltic Shield. Precambrian Research, 35, 382 pp.Google Scholar
Gerdin, P. (1979). Vieto resultat av utförda prospekteringsarbeten [Vieto Results of Exploration Work Performed]. Rapport för NSG. Berggrundsbyrån, Geological Survey of Sweden.Google Scholar
Grad, M., Tiira, T. and the ESC Working Group (2009). The Moho depth map of the European Plate. Geophysical Journal International, 176, 279292, doi.org/10.1111/j.1365-246X.2008.03919.x.Google Scholar
Gupta, H. K., Arora, K., Rao, N. P. et al. (2017). Investigations of continued reservoir triggered seismicity at Koyna, India. In Mukherjee, S., Misra, A. A., Calvès, G. and Nemčok, M., eds., Tectonics of the Deccan Large Igneous Province. Geological Society, London, Special Publication, Vol. 445, pp. 151188, doi.org/10.1144/SP445.11.Google Scholar
Haimson, B. C. and Cornet, F. H. (2003). ISRM SM for rock stress estimation – part 3: hydraulic fracturing (HF) and/or hydraulic testing of pre-existing fractures (HTPF). International Journal of Rock Mechanics and Mining Sciences, 40, 10111020, doi.org/10.1016/j.ijrmms.2003.08.002.Google Scholar
Harper, J., Hubbard, A., Ruskeeniemi, T. et al. (2016). The Greenland Analogue Project. SKB Report R-14-13, 387 pp.Google Scholar
Hirose, T., Kawagucci, S. and Suzuki, K. (2011). Mechanoradical H2 generation during simulated faulting: implications for an earthquake‐driven subsurface biosphere. Geophysical Research Letters, 38, L17303, doi.org/10.1029/2011GL048850.Google Scholar
Johnson, C. W., Fu, Y. and Bürgmann, R. (2017). Seasonal water storage, stress modulation, and California seismicity. Science, 356(6343), 11611164, doi.org/10.1126/science.aak9547.Google Scholar
Juhlin, C., Dehghannejad, M., Lund, B., Malehmir, A. and Pratt, G. (2010). Reflection seismic imaging of the end-glacial Pärvie Fault system, Sweden. Journal of Applied Geophysics, 70, 307316, doi.org/10.1016/j.jappgeo.2009.06.004.Google Scholar
Kallmeyer, J. (2017). Contamination control for scientific drilling operations. Advances in Applied Microbiology, 98, 6191, doi.org/10.1016/bs.aambs.2016.09.003.Google Scholar
Keiding, M., Kreemer, C., Lindholm, C. D. et al. (2015). A comparison of strain rates and seismicity for Fennoscandia: depth dependency of deformation from glacial isostatic adjustment. Geophysical Journal International, 202, 10211028, doi.org/10.1093/gji/ggv207.Google Scholar
Kierulf, H. P., Steffen, H., Simpson, M. J. R. et al. (2014). A GPS velocity field for Fennoscandia and a consistent comparison to glacial isostatic adjustment models. Journal of Geophysical Research, 119(8), 66136629, doi.org/10.1002/2013JB010889.Google Scholar
Korja, T. (2007). How is the European lithosphere imaged by magnetotellurics? Surveys in Geophysics, 28(2–3), 239272, doi.org/10.1007/s10712-007-9024-9.Google Scholar
Kreemer, C., Blewitt, G. and Klein, E. C. (2014). A geodetic plate motion and Global Strain Rate Model. Geochemistry, Geophysics, Geosystems, 15, 38493889, doi.org/10.1002/2014GC005407.Google Scholar
Kuivamäki, A., Vuorela, P. and Paananen, M. (1998). Indications of Postglacial and Recent Bedrock Movements in Finland and Russian Karelia. Geological Survey of Finland Nuclear Waste Disposal Research Report YST-99, Espoo, Finland, 92 pp.Google Scholar
Kujansuu, R. (1972). The deglaciation of Finnish Lapland. In L. K. Kauranne, ed., Glacial Stratigraphy, Engineering Geology and Earth Construction. Geological Survey of Finland Special Paper 15, pp. 2131.Google Scholar
Kukkonen, I. T., Olesen, O., Ask, M. V. S. and the PFDP Working Group (2010). Postglacial faults in Fennoscandia: targets for scientific drilling. GFF, 132(1), 7181, doi.org/10.1080/11035891003692934.Google Scholar
Lagerbäck, R. (1978). Neotectonic structures in northern Sweden. Geologiska Föreningens i Stockholm Förhandlingar, 100(3), 263269, doi.org/10.1080/11035897809452533.Google Scholar
Lagerbäck, R. (1992). Dating of Late Quaternary faulting in northern Sweden. Journal of the Geological Society, London, 149, 285291, doi.org/10.1144/gsjgs.149.2.0285.Google Scholar
Lagerbäck, R. and Sundh, M. (2008). Early Holocene Faulting and Paleoseismicity in Northern Sweden. Geological Survey of Sweden Research Paper, C836, 84 pp.Google Scholar
Lagerbäck, R. and Witschard, F. (1983). Neotectonics in Northern Sweden – Geological Investigations. SKBF/KBS Technical Report 83–58, Svensk Kärnbränslehantering AB, Stockholm, 58 pp.Google Scholar
Lahtinen, R., Korja, A. and Nironen, M. (2005). Palaeoproterozoic tectonic evolution. In Lehtinen, M., Nurmi, P. and Rämö, T., eds., Precambrian Geology of Finland – Key to the Evolution of the Fennoscandian Shield. Elsevier Science Publishers, Amsterdam, pp. 481–531, doi.org/10.1016/S0166-2635(05)809012-X.Google Scholar
Leonard, M. (2010). Earthquake fault scaling: self-consistent relating of rupture length, width, average displacement, and moment release. Bulletin of the Seismological Society of America, 100, 19711988, doi.org/10.1785/0120090189.Google Scholar
Li, Q., Liu, M. and Stein, S. (2009). Spatiotemporal complexity of continental intraplate seismicity: insights from geodynamic modeling and implications for seismic hazard estimation. Bulletin of the Seismological Society of America, 99, 5299, doi.org/10.1785/0120080005.Google Scholar
Lin, Y.-Y., Ma, K.-F. and Oye, V. (2012). Observation and scaling of microearthquakes from the Taiwan Chelungpu-fault borehole seismometers. Geophysical Journal International, 190, 665676, doi.org/10.1111/j.1365-246X.2012.05513.x.Google Scholar
Lindblom, E., Lund, B., Tryggvason, A. et al. (2015). Microearthquakes illuminate the deep structure of the endglacial Pärvie fault, northern Sweden, Geophysical Journal International, 201, 17041716, doi.org/10.1093/gji/ggv112.Google Scholar
Lund, B. (2015). Palaeoseismology of glaciated terrain. In Beer, M., Kougioumtzoglou, I. A., Patelli, E. and Au, I. K., eds., Encyclopedia of Earthquake Engineering. Springer, Berlin/Heidelberg, doi.org/10.1007/978-3-642-36197-5_25-1.Google Scholar
Lundqvist, J. and Lagerbäck, R. (1976). The Pärve Fault: a late-glacial fault in the Precambrian of Swedish Lapland. Geologiska Föreningens i Stockholm Förhandlingar, 98, 4551, doi.org/10.1080/11035897609454337.Google Scholar
Mather, W. W. (1843). Geology of New-york. Part I. Comprising the Geology of the First Geological District. Carroll & Cook, Albany.Google Scholar
Matthew, G. F. (1894). Movements of the Earth’s crust at St. John, N. B., in post-glacial times. Bulletin of the Natural History Society of New Brunswick, 12, 3442.Google Scholar
Mikko, H., Smith, C. A., Lund, B., Ask, M. V. S. and Munier, R. (2015). LiDAR-derived inventory of post-glacial fault scarps in Sweden. GFF, 137, 344352, doi.org/10.1080/11035897.2015.1036360.Google Scholar
Muir Wood, R. (1989). Extraordinary deglaciation reverse faulting in northern Fennoscandia. In Gregersen, S. and Basham, P. W., eds., Earthquakes at North-Atlantic Passive Margins: Neotectonics and Postglacial Rebound. Kluwer Academic Publishers, Dordrecht, pp. 141173, doi.org/10.1007/978-94-009-2311-9_10.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2017). Postglacial seismic activity along the Isovaara–Riikonkumpu fault complex. Global and Planetary Change, 157, 5972, doi.org/10.1016/j.gloplacha.2017.08.015.Google Scholar
Ojala, A. E. K., Markovaa-Koivisto, M., Middleton, M. et al. (2018). Dating of paleolandslides in western Finnish Lapland. Earth Surface Processes and Landforms, 43, 24492462, doi.org/10.1002/esp.4408.Google Scholar
Ojala, A. E. K., Mattila, J., Markovaara-Koivisto, M. et al. (2019). Distribution and morphology of landslides in northern Finland: an analysis of postglacial seismic activity. Geomorphology, 326, 190201, doi.org/10.1016/j.geomorph.2017.08.045.Google Scholar
Olesen, O., Henkel, H., Lile, O. B., Mauring, E. and Rønning, J. S. (1992). Geophysical investigations of the Stuoragurra postglacial fault, Finnmark, northern Norway. Journal of Applied Geophysics, 29, 95118, doi.org/10.1016/0926-9851(92)90001-2.Google Scholar
Olesen, O., Bungum, H., Lindholm, C. et al. (2013). Neotectonics, seismicity and contemporary stress field in Norway – mechanisms and implications. In Olsen, L., Fredin, O. and Olesen, O., eds., Quaternary Geology of Norway. Geological Survey of Norway Special Publication 13, pp. 145174.Google Scholar
Page, M. T. and Hough, S. E. (2014). The New Madrid seismic zone: not dead yet. Science, 343(6172), 762764, doi.org/10.1126/science.1248215.Google Scholar
Peng, Z. and Gomberg, J. (2010). An integrated perspective of the continuum between earthquakes and slow-slip phenomena. Nature Geoscience, 3(9), 599607, doi.org/10.1038/ngeo940.Google Scholar
Plomerova, J. and Babuska, V. (2010). Long memory of mantle lithosphere fabric – European LAB constrained from seismic anisotropy. Lithos, 120, 131143, doi.org/10.1016/j.lithos.2010.01.008.Google Scholar
Riad, L. (1990). The Pärvie Fault, Northern Sweden. Research Report 63, Minerology, Department of Mineralogy and Petrology, Uppsala University, 48 pp.Google Scholar
Saar, M. O. and Manga, M. (2003). Seismicity induced by seasonal groundwater recharge at Mt. Hood, Oregon. Earth and Planetary Science Letters, 214(3–4), 605618, doi.org/10.1016/S0012-821X(03)00418-7.Google Scholar
Sarlus, Z., Andersson, U. B., Bauer, T. E. et al. (2018). Timing of plutonism in the Gällivare area: implications for Proterozoic crustal development in the northern Norrbotten ore district, Sweden. Geological Magazine, 155(6), 13511376, doi.org/10.1017/S0016756817000280.Google Scholar
Smith, C., Sundh, M. and Mikko, H. (2014). Surficial geology indicates early Holocene faulting and seismicity, central Sweden. International Journal of Earth Sciences, 103, 17111724, doi.org/10.1007/s00531–014-1025-6.Google Scholar
Stein, S. and Liu, M. (2009). Long aftershock sequences within continents and implications for earthquake hazard assessment. Nature, 462, 8789, doi.org/10.1038/nature08502.Google Scholar
Stein, S., Liu, M., Calais, E. and Li, Q. (2009). Mid-continent earthquakes as a complex system. Seismological Research Letters, 80(4), 551553, doi.org/10.1785/gssrl.80.4.551.Google Scholar
Stephansson, O. (1983). Rock Stress Measurements by Sleeve Fracturing. Proceedings of the 5th Congress of the International Society for Rock Mechanics and Rock Engineering, 10–15 April 1983, Melbourne, Australia, F129–F137.Google Scholar
Stephansson, O. (1989). Stress measurements and modeling of crustal rock mechanics in Fennoscandia. In Gregersen, S. and Basham, P. W., eds., Earthquake at North-Atlantic Passive Margins: Neotectonics and Postglacial Rebound. NATO Advanced Studies Institute Series, Series C, Vol. 266, pp. 213229.Google Scholar
Stroeven, A. P., Hättestrand, C., Kleman, J. et al. (2016). Deglaciation of Fennoscandia. Quaternary Science Reviews, 147, 91121, doi.org/10.1016/j.quascirev.2015.09.016.Google Scholar
Sugihara, T., Kinoshita, M., Araki, E. et al. (2014). Re-evaluation of temperature at the updip limit of locked portion of Nankai megasplay inferred from IODP Site C0002 temperature observatory. Earth, Planets and Space, 66, 107, doi.org/10.1186/1880-5981-66-107.Google Scholar
Sutinen, R. (2005). Timing of early Holocene landslides in Kittilä, Finnish Lapland. In Ojala, A. E. K., ed., Quaternary Studies in the Northern and Arctic Regions of Finland, Proceedings of the workshop organized within the Finnish National Committee for Quaternary Research (INQUA), Kilpisjärvi Biological Station, Finland, January 13–14th 2005. Geological Survey of Finland Special Paper 40, Espoo, Finland, pp. 53–58.Google Scholar
Sutinen, R., Hyvönen, E., Middleton, M. and Ruskeeniemi, T. (2014). Airborne LiDAR detection of postglacial faults and Pulju moraine in Palojärvi, Finnish Lapland. Global and Planetary Change, 115, 2432, doi.org/10.1016/j.gloplacha.2014.01.007Google Scholar
Townend, J., Sutherland, R., Toy, V. G. et al. (2017). Petrophysical, geochemical, and hydrological evidence for extensive fracture-mediated fluid and heat transport in the Alpine Fault’s hanging-wall damage zone. Geochemistry, Geophysics, Geosystems, 18(12), 47094732, doi.org/10.1002/2017GC007202.Google Scholar
Ueda, T. and Kato, A. (2019). Seasonal variations in crustal seismicity in San‐in district, southwest Japan. Geophysical Research Letters, 46, 31723179, doi.org/10.1029/2018GL081789.Google Scholar
Vestøl, O. (2006). Determination of postglacial land uplift in Fennoscandia from leveling, tide-gauges and continuous GPS stations using least squares collocation. Journal of Geodesy, 80(5), 248258, doi.org/10.1007/s00190-006-0063-7.Google Scholar
Wanke, A. and Melezhik, V. (2005). Sedimentary and volcanic facies recording the Neoarchaean continent breakup and decline of the positive δ13 Ccarb excursion. Precambrian Research, 140(1–2), 135, doi.org/10.1016/j.precamres.2005.05.003.Google Scholar
Westaway, R. (2006). Investigation of coupling between surface processes and induced flow in the lower continental crust as a cause of intraplate seismicity. Earth Surface Processes and Landforms, 31, 14801509, doi.org/10.1002/esp.1366.Google Scholar
Witschard, F. (1984). The geological and tectonic evolution of the Precambrian of northern Sweden – a case for basement reactivation? Precambrian Research, 23(3–4), 273315, doi.org/10.1016/0301-9268(84)90047-0.Google Scholar
Wu, P. and Hasegawa, H. S. (1996a). Induced stresses and fault potential in eastern Canada due to a disc load: a preliminary analysis. Geophysical Journal International, 125, 415430, doi.org/10.1111/j.1365-246X.1996.tb00008.x.Google Scholar
Wu, P. and Hasegawa, H. S. (1996b). Induced stresses and fault potential in eastern Canada due to a realistic load: a preliminary analysis. Geophysical Journal International, 127, 215229, doi.org/10.1111/j.1365-246X.1996.tb01546.x.Google Scholar
Zoback, M., Hickman, S., Ellsworth, W. and the SAFOD Science Team (2011). Scientific drilling into the San Andreas Fault Zone – an overview of SAFOD’s first five years. Scientific Drilling, 11, 1428, doi.org/10.2204/iodp.sd.11.02.2011.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×