Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-19T02:36:37.823Z Has data issue: false hasContentIssue false

3 - Earthquake-Induced Landforms in the Context of Ice-Sheet Loading and Unloading

from 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

In this chapter we present examples of earthquake-induced geomorphology in Northern Europe ranging from the readily visible surface expression to more subtle and complex landforms.

Stress changes in the subsurface created by loading and unloading of the ice sheets can result in reactivation of deep-seated faults. Glacially induced faulting can happen during the glaciation in a proglacial or subglacial setting, in a distal setting away from the ice margin or in a postglacial setting after the ice sheet has melted away. Thus, the timing and the location of the tectonic event is important for the resulting landform creation or landform change. Identification of earthquake-induced landforms can be used in interpretations of palaeoseismic events, for location of previously unrecognized fault zones and in evaluations of the likelihood of future seismic events. Interpretations of earthquake-induced landforms in and around former glaciated areas can therefore add important information to interpretations of both the Quaternary geology and the deep structural framework.

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

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.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.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.Google 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.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
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.2005.05.032Google 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.Google 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.CrossRefGoogle 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.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. (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.xGoogle 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.Google 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.Google 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.Google 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.Google 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.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
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.Google 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.Google 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.Google 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.Google 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.Google 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.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
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.Google 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.Google 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.CrossRefGoogle 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
×