Skip to main content Accessibility help
×
Hostname: page-component-7c8c6479df-27gpq Total loading time: 0 Render date: 2024-03-29T02:26:52.640Z Has data issue: false hasContentIssue false

11 - Postglacial Faulting in Norway

Large Magnitude Earthquakes of the Late Holocene Age

from Part III - Glacially Triggered Faulting in the Fennoscandian Shield

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 90-km long Stuoragurra Fault Complex, part of the approximately 4–5-km wide Precambrian Mierojávri–Sværholt Shear Zone, constitutes the Norwegian part of the larger Lapland province of postglacial faults. It consists of three separate fault systems being 6–12 km apart. The faults dip 30–75° to the SE and can be traced to about 500 m depth. Deep seismic profiling shows that the shear zone dips at an angle of about 43° to the southeast and can be traced to about 3 km depth. A total of approximately 80 earthquakes were registered here between 1991 and 2019. Most of them occurred to the southeast of the fault scarps. The maximum moment magnitude was 4.0. The formation of postglacial faults in northern Fennoscandia has previously been associated with the deglaciation of the last inland ice. Dating of fault reactivation reveals, however, a late Holocene age (between around 700 and 4000 a BP). The reverse displacement of around 9 m and fault system lengths of 14 and 21 km of the two southernmost fault systems indicate a moment magnitude of about 7. The results from this study indicate that the expected maximum magnitude of future earthquakes in Fennoscandia is about 7.

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

ABEM (2012). ABEM Terrameter LS. Instruction Manual. ABEM 20120109, based on release 1.10. ABEM, Sweden.Google Scholar
Åm, M. (1994). Mineralogisk og petrologisk karakterisering av vitrings/sleppemateriale fra Stuoragurraforkastningen Finnmark [Mineralogical and Petrological Characterization of Weathering/Drag-Along Material from the Stuoragurra Fault, Finnmark]. MSc thesis, Norwegian University of Science and Technology, 102 pp.Google Scholar
Bakken, A. J. H. (1983). Nordmannvikdalen kvartærgeologi og geomorfologi [Quaternary Geology and Geomorphology of Nordmanvikdalen]. MSc thesis, University of Oslo, 126 pp.Google Scholar
Berthelsen, A. and Marker, M. (1986). 1.9–1.8 Ga old strike-slip megashears in the Baltic Shield, and their plate tectonic implications. In Galson, D. A. and Mueller, S., eds., The European Geotraverse, Part 2. Tectonophysics, 128(3–4), pp. 163–181, doi.org/10.1016/0040-1951(86)90292-1.Google Scholar
Bingen, B., Solli, A., Viola, G. et al. (2015). Geochronology of the Palaeoproterozoic Kautokeino Greenstone Belt, Finnmark, Norway: tectonic implications in a Fennoscandia context. Norwegian Journal of Geology, 95, 365396, doi.org/10.17850/njg95-3-09.Google Scholar
Bungum, H. and Lindholm, C. (1997). Seismo- and neotectonics in Finnmark, Kola Peninsula and the southern Barents Sea. Part 2: seismological analysis and seismotectonics. Tectonophysics, 270, 1528, doi.org/10.1016/S0040-1951(96)00139-4.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, 10,62110,637, doi.org/10.1002/2016GL070815.CrossRefGoogle Scholar
Clark, D., McPherson, A. and Van Dissen, R. (2012). Long-term behaviour of Australian stable continental region (SCR) faults. Tectonophysics, 566, 130, doi.org/10.1016/j.tecto.2012.07.004.CrossRefGoogle Scholar
Craig, T. J., Calais, E., Fleitout, L., Bollinger, L. and Scotti, O. (2016). Evidence for the release of long-term tectonic strain stored in continental interiors through intraplate earthquakes, Geophysical Research Letters, 43, 68266836, doi.org/10.1002/2016GL069359.Google Scholar
Dahlin, T. (1993). On the Automation of 2D Resistivity Surveying for Engineering and Environmental Applications. PhD thesis, Department of Engineering Geology, Lund Institute of Technology, Lund University.Google Scholar
Dalsegg, E. and Olesen, O. (2014). Resistivitetsmålinger ved Masi, Fitnajohka and Riednajávre og implikasjoner for malmleting [Resistivity Measurements at Masi, Fitnajohka and Riednajávre and Implications for Ore Exploration]. NGU Report 2014.021, 28 pp.Google Scholar
Dehls, J., Olesen, O., Olsen, L. and Blikra, L. H. (2000). Neotectonic faulting in northern Norway; the Stuoragurra and Nordmannvikdalen postglacial faults. Quaternary Science Reviews, 19, 14451460, doi.org/10.1016/S0277-3791(00)00073-1.Google Scholar
Eiken, O., Degutsch, M., Riste, P. and Rød, K. (1989). Snowstreamer: an efficient tool in seismic acquisition. First Break, 7(9), 374378, doi.org/10.3997/1365-2397.1989021.Google Scholar
Fundal, E. (1967). En undersøkelse i det prekambriske Biggevarre område i Finnmark, Nord-Norge med særlig henblikk på de såkalde albitdiabasers geologi og petrografi [An Investigation in the Precambrian Biggevarre Area in Finnmark, Northern Norway with Special Attention to the Geology and Petrography of the So-Called Albitdiabas]. NGU Report 680, 81 pp.Google Scholar
Gibbons, S. J. and Kværna, T. (2017). Illuminating the seismicity pattern of the October 8, 2005, M = 7.6 Kashmir earthquake aftershocks. Physics of the Earth and Planetary Interiors, 270, 18, doi.org/10.1016/j.pepi.2017.06.008.Google Scholar
Henderson, I. H. C., Viola, G. and Nasuti, A. (2015). A new tectonic model for the Kautokeino Greenstone Belt, northern Norway, based on high resolution airborne magnetic data and field structural analysis and implications for mineral potential. Norwegian Journal of Geology, 95, 339363, doi.org/10.17850/njg95-3-05.Google Scholar
Henkel, H. (1991). Magnetic crustal structures in northern Fennoscandia. In Wasilewski, P. and Hood, P., eds., Magnetic Anomalies – Land and Sea. Tectonophysics, 192, pp. 5779, doi.org/10.1016/0040-1951(91)90246-O.Google Scholar
Henriksen, H. (1986). Bedrock map Iddjajavri 2034 II M 1:50 000, preliminary edition. Geological Survey of Norway, Trondheim.Google Scholar
Johansen, T. A., Ruud, B. O., Bakke, N. E. et al. (2011). Seismic profiling on Arctic glaciers. First Break, 29(2), 2935.Google Scholar
Lagerbäck, R. and Sundh, M. (2008). Early Holocene Faulting and Paleoseismicity in Northern Sweden. Geological Survey of Sweden Research Paper, Series C, Vol. 836, 80 pp.Google Scholar
Liu, M. and Stein, S. (2016). Mid-continental earthquakes: spatiotemporal occurrences, causes, and hazards. Earth-Science Reviews, 162, 364386, doi.org/10.1016/j.earscirev.2016.09.016.CrossRefGoogle Scholar
Loke, M. H. (2010). RES2INV ver. 3.59. Geoelectrical Imaging 2D & 3D. Instruction Manual. 151 pp. www.geoelectrical.com.Google Scholar
Lund, B., Schmidt, P. and Hieronymus, C. (2009). Stress Evolution and Fault Instability during the Weichselian Glacial Cycle. SKB Technical Report TR-09-15, Swedish Nuclear Fuel and Waste Management Co., Stockholm, Sweden, 106 pp.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.2019.05.022CrossRefGoogle 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.CrossRefGoogle 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. NGU Report, 2019.009, 64 pp.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.Google Scholar
Myers, S. C., Johannesson, G. and Hanley, W. (2007). A Bayesian hierarchical method for multiple-event seismic location. Geophysical Journal International, 171, 10491063, doi.org/10.1111/j.1365-246X.2007.03555.x.CrossRefGoogle Scholar
Nasuti, A., Roberts, D., Dumais, M.-A. et al. (2015). New high-resolution aeromagnetic and radiometric surveys in Finnmark and North Troms: linking anomaly patterns to bedrock geology and structure. Norwegian Journal of Geology, 95, 217243, doi.org/10.17850/njg95-3-10.Google Scholar
Ojala, A. E. K., Mattila, J., Ruskeeniemi, T. et al. (2017). Postglacial seismic activity along the Isovaarae–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 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., Hämäläinen, J. and Sutinen, R. (2019). 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
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. and Sandstad, J. (1993). Interpretation of the Proterozoic Kautokeino Greenstone Belt, Finnmark, Norway from combined geophysical and geological data. Geological Survey of Norway Bulletin, 425, 4364.Google Scholar
Olesen, O., Roberts, D., Henkel, H., Lile, O. B. and Torsvik, T. H. (1990). Aeromagnetic and gravimetric interpretation of regional structural features in the Caledonides of West Finnmark and Northern Troms, north Norway. Geological Survey of Norway Bulletin, 419, 124.Google Scholar
Olesen, O., Henkel, H., Lile, O.B., Mauring, E. and Rønning, J. S. (1992a). 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., Henkel, H., Lile, O.B. et al. (1992b). Neotectonics in the Precambrian of Finnmark, northern Norway. Norsk Geologisk Tidsskrift, 72, 301306.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., Brönner, M., Ebbing, J. et al. (2010). New aeromagnetic and gravity compilations from Norway and adjacent areas – methods and applications. In Vining, B. A. and Pickering, S. C., eds., Petroleum Geology: From Mature Basins to New Frontiers. Proceedings of the 7th Petroleum Geology Conference. Petroleum Geology Conference Series 7, Geological Society of London, pp. 559–586, doi.org/10.1144/0070559.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
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
Olsen, L., Olesen, O. and Høgaas, F. (2020). Dating of the Stuoragurra Fault at Finnmarksvidda, northern Norway. In Nakrem, H. A. and Husås, A. M., eds., 34th Nordic Geological Winter Meeting January 8th–10th 2020, Oslo, Norway. Abstracts and Proceedings of the Geological Society of Norway, 1, pp. 157158.Google Scholar
Olsen, L., Olesen, O., Høgaas, F. and Tassis, G. (2021). A part of the Stuoragurra postglacial fault complex, at Máze in N-Norway, is less than 600 yrs old. In Nakrem, H. A. and Husås, A. M. (eds.), Vinterkonferansen 2021, Digital, January 6–8, 2020. Abstracts and Proceedings of the Geological Society of Norway, 1, p. 55.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
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.CrossRefGoogle Scholar
Siedlecka, A. (1985). Geology of the Iešjávri-Skoganvarre area, northern Finnmarksvidda, North Norway. Geological Survey of Norway Bulletin, 403, 103112.Google Scholar
Siedlecka, A. (1987). Berggrunnskart Iešjávri; 1934 II, foreløpig utgave, M 1:50 000 [Iešjávri bedrock map; 1934 II, preliminary edition, M 1:50,000]. Norges geologiske undersøkelse, Trondheim.Google Scholar
Siedlecka, A., Davidsen, B., Rice, A. H. N. and Townsend, C. (2011). Berggrunnskart; Skoganvarri 2034 IV, M 1:50 000, revidert foreløpig utgave [Bedrock map; Skoganvarri 2034 IV, M 1:50,000, revised preliminary edition]. Norges geologiske undersøkelse, Trondheim.Google Scholar
Siedlecka, A. and Roberts, D. (1996). Finnmark Fylke. Berggrunnsgeologi Finnmark Fylke M 1:500 000 [Finnmark Fylke; bedrock map M 1:500 000]. Norges geologiske undersøkelse, Trondheim.Google Scholar
Skaar, J. A. Å. (2014). 3D Geophysical and Geological Modelling of the Karasjok Greenstone Belt. PhD thesis, Norwegian University of Science and Technology, 170 pp.Google Scholar
Sletten, K., Olsen, L. and Blikra, L.H. (2000). Slides in low-gradient areas of Finnmarksvidda. In Dehls, J. and Olesen, O., eds., Neotectonics in Norway, Annual Technical Report 1999. NGU Report 2000.001, 41–42.Google Scholar
Smith, C. A., Grigull, S. and Mikko, H. (2018). Geomorphic evidence for multiple surface ruptures of the Merasjärvi “postglacial fault,” northern Sweden. GFF, 140, doi.org/10.1080/11035897.2018.1492963Google Scholar
Smith, C. A., Sundh, M. and Mikko, H. (2014). Surficial geologic evidence for early Holocene faulting and seismicity. International Journal of Earth Sciences, 103, 17111724, doi.org/10.1007/s00531-014-1025-6.Google Scholar
Solli, A. (1983). Precambrian stratigraphy in the Masi area, Southwestern Finnmark, Norway. Geological Survey of Norway Bulletin, 380, 97105.Google Scholar
Solli, A. (1988). Masi, 1933 IV – berggrunnsgeologisk kart – M 1:50,000 [Masi, 1933 IV – Map of bedrock geology – M 1:50,000]. Norges geologiske undersøkelse, Trondheim.Google Scholar
Steffen, R., Steffen, H., Wu, P. and Eaton, D. W. (2014a). Stress and fault parameters affecting fault slip magnitude and activation time during a glacial cycle. Tectonics, 33, doi.org/10.1002/2013TC003450.Google Scholar
Steffen, R., Wu, P., Steffen, H. and Eaton, D. W. (2014b). On the implementation of faults in finite-element glacial isostatic adjustment models. Computers & Geosciences, 62, 150159, doi.org/10.1016/j.cageo.2013.06.012.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
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
Tolgensbakk, J. and Sollid, J. L. (1988). Kåfjord, kvartærgeologi og geomorfologi 1:50,000, 1634 II [Kåfjord, Quaternary geology and geomorphology, 1:50,000, 1634 II]. Geografisk Institutt, University of Oslo.Google Scholar
Townsend, C., Rice, A. H. N. and Mackay, A. (1989). The structure and stratigraphy of the southwestern portion of the Gaissa Thrust Belt and the adjacent Kalak Nappe Complex, N Norway. In Gayer, R. A., ed., The Caledonide Geology of Scandinavia. Graham & Trotman, London, pp. 111126.Google Scholar
Vestøl, O., Ågren, J., Steffen, H., Kierulf, H. and Tarasov, L. (2019). NKG2016LU: a new land uplift model for Fennoscandia and the Baltic Region. Journal of Geodesy, 93(9), 17591779, doi.org/10.1007/s00190-019-01280-8.Google Scholar
Wells, D. L. and Coppersmith, K. J. (1994). New empirical relationships among magnitude, rupture length, rupture area, and surface displacement. Bulletin of the Seismological Society of America, 84, 9741002.Google Scholar
Wesnousky, S. G. (2008). Displacement and geometrical characteristics of earthquake surface ruptures: issues and implications for seismic-hazard analysis and the process of earthquake rupture. Bulletin of the Seismological Society of America, 98, 16091632, doi.org/10.1785/0120070111.Google Scholar
Wu, P., Johnston, P. and Lambeck, K. (1999). Postglacial rebound and fault instability in Fennoscandia. Geophysical Journal International, 139(3), 657670, doi.org/10.1046/j.1365-246x.1999.00963.x.Google Scholar
Zwaan, K. B. (1985). Berggrunnskart Suolovuopmi 1934 III, M 1:50 000, foreløpig utgave [Map of bedrock geology Suolovuompi 1934 III, M 1:50,000, preliminary edition]. Norges geologiske undersøkelse, Trondheim.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
×