Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-27T15:23:32.568Z Has data issue: false hasContentIssue false

Part I - Introduction

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

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
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

Adams, J. (1981). Postglacial Faulting: A Literature Survey of Occurrences in Eastern Canada and Comparable Glaciated Areas. Atomic Energy of Canada Limited Technical Report, TR-142.Google Scholar
Adams, J. (1996). Paleoseismology in Canada: a dozen years of progress. Journal of Geophysical Research, 101, 61936207, doi.org/10.1029/95JB01817.Google Scholar
Adams, J., Wetmiller, R. J., Hasegawa, H. S. and Drysdale, J. (1991). The first surface faulting from a historical intraplate earthquake in North America. Nature, 352, 617619, doi.org/10.1038/352617a0.Google Scholar
Adams, J., Percival, J. A., Wetmiller, R. J., Drysdale, J. and Robertson, P. B. (1992). Geological controls on the 1989 Ungava surface rupture: a preliminary interpretation. Papers of the Geological Survey of Canada, 92-1C, 147155.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
Bentley, M. J. and McCulloch, R. D. (2005). Impact of neotectonics on the record of glacier and sea level fluctuations, Strait of Magellan, southern Chile. Geografiska Annaler: Series A, Physical Geography, 87(2), 393402, doi.org/10.1111/j.0435-3676.2005.00265.x.Google Scholar
Brandes, C. and Tanner, D. (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., Polom, U. and Winsemann, J. (2011). Reactivation of basement faults: interplay of ice-sheet advance, glacial lake formation and sediment loading. Basin Research, 23, 5364, doi.org/10.1111/j.1365-2117.2010.00468.x.Google 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.Google Scholar
Brandes, C., Plenefisch, T., Tanner, D. T., Gestermann, N. and Steffen, H. (2019). Evaluation of deep crustal earthquakes in northern Germany – possible tectonic causes. Terra Nova, 31(2), 8393, doi.org/10.1111/ter.12372.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
Brøgger, W. C. (1884). Spaltenverwerfungen in der Gegend Langesund-Skien [Crevasse faults in the Langesund-Skien area]. Nyt Magazin for Naturvidenskaberne, 28, 253419.Google Scholar
Brooks, G. R. and Adams, J. (2020). A review of evidence of glacially-induced faulting and seismic shaking in southeastern Canada. Quaternary Science Reviews, 228, 106070, doi.org/10.1016/j.quascirev.2019.106070.CrossRefGoogle Scholar
Byrd, J. O. D., Smith, R. B. and Geissman, J. W. (1994). The Teton fault, Wyoming: topographic signature, neotectonics, and mechanisms of deformation. Journal of Geophysical Research, 99(B10), 2009520122,doi.org/10.1029/94JB00281.Google Scholar
Clark, P. U., Dyke, A. S., Shakun, J. D., et al. (2009). The last glacial maximum. Science, 325(5941), 710714, doi.org/10.1126/science.1172873.Google Scholar
Fenton, C. (1991). Neotectonics and Palaeoseismicity in North West Scotland. PhD thesis, University of Glasgow, Glasgow.Google Scholar
Fenton, C. (1994). Postglacial Faulting in Eastern Canada. Geological Survey of Canada Open File, 2774.Google Scholar
Fenton, C. (1999). Glacio-isostatic (postglacial) faulting: criteria for recognition. In Hanson, K. L., Kelson, K. I., Angell, M. A. and Lettis, W. R., eds., Identifying Faults and Determining Their Origins. U.S. Nuclear Regulatory Commission, pp. A-51A-99.Google Scholar
Firth, C. R. and Stewart, I. S. (2000). Postglacial tectonics of the Scottish glacio-isostatic uplift centre. Quaternary Science Reviews, 19, 14691493, doi.org/10.1016/S0277-3791(00)00074-3.CrossRefGoogle Scholar
Greene, D. C. (1996). Quaternary reactivation of the Lost Lakes fault, a brittle fault zone containing pseudotachylite in the Tuolumne intrusive suite, Sierra Nevada, California. Geological Society of America Cordilleran Section, Spring Meeting, Abstracts with Program, 28, p. 70.Google Scholar
Gregersen, S., Leth, J., Lind, G. and Lykke-Andersen, H. (1996). Earthquake activity and its relationship with geologically recent motion in Denmark. Tectonophysics, 254, 265273, doi.org/10.1016/0040-1951(95)00193-X.Google Scholar
Grollimund, B. and Zoback, M. D. (2001). Did deglaciation trigger intraplate seismicity in the New Madrid seismic zone? Geology, 29, 175178, doi.org/10.1130/0091-7613(2001)029%3C0175:DDTISI%3E2.0.CO;2.Google Scholar
Hetzel, R. and Hampel, A. (2005). Slip rate variations on normal faults during glacial-interglacial changes in surface loads. Nature, 435, 8184, doi.org/10.1038/nature03562.Google Scholar
Hinz, N. H., Carson, R. J., Gardner, T. W. and McKenna, K. (1997). Late Quaternary deglaciation, flooding, and tectonism (?), upper Clarks Fork Valley, Park County, Wyoming. Geological Society of America Annual Meeting, Abstracts with Program, 29, p. 15.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
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
Jäckli, H. C. A. (1965). Pleistocene glaciation of the Swiss Alps and signs of postglacial differential uplift. Geological Society of America, Special Paper, 84, 153157, doi.org/10.1130/SPE84-p153.Google Scholar
Jakobsson, M., Björck, S., O’Regan, M. et al. (2014). Major earthquake at the Pleistocene–Holocene transition in Lake Vättern, southern Sweden. Geology, 42, 379382, doi.org/10.1130/G35499.1.Google Scholar
Johnston, A. C. (1987). Suppression of earthquakes by large continental ice sheets. Nature, 330, 467469, doi.org/10.1038/330467a0.Google Scholar
Karpov, N. N. (1960). Traces of postglacial tectonic faults in the Khibiny Mountains. Moscow University Bulletin, 5(4), 61.Google Scholar
Knight, J. (1999). Geological evidence for neotectonic activity during deglaciation of the southern Sperrin Mountains, Northern Ireland. Journal of Quaternary Science, 14, 4557, doi.org/10.1002/(SICI)1099-1417(199902)14:1<45::AID-JQS389>3.0.CO;2-3.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, 16, 3036 (in Finnish).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.CrossRefGoogle Scholar
Lagerbäck, R. (1992). Dating of Late Quaternary faulting in northern Sweden. Journal of the Geological Society, 149, 285291, doi.org/10.1144/gsjgs.149.2.0285.CrossRefGoogle 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
Lough, A. C., Wiens, D. A. and Nyblade, A. (2018). Reactivation of ancient Antarctic rift zones by intraplate seismicity. Nature Geoscience, 11(7), 515519, doi.org/10.1038/s41561-018-0140-6.Google Scholar
Ludwig, A. O. (1954/1955). Eistektonik und echte Tektonik in Ost-Rügen (Jasmund) [Ice tectonics and real tectonics in East Rügen Island (Jasmund)]. Wissenschaftliche Zeitschrift der E.-M.-A.-Universität Greifswald, 4, 251288.Google Scholar
Lund, B. (2015). Palaeoseismology of glaciated terrain. In Beer, M. et al., eds., Encyclopedia of Earthquake Engineering. Springer-Verlag, Berlin/Heidelberg, 17651779.Google Scholar
Lund, B. and Näslund, J.-O. (2009). Glacial isostatic adjustment – implications for glacially induced faulting and nuclear waste repositories. In Connor, C. B., Chapman, N. A. and Connor, L. J., eds., Volcanic and Tectonic Hazard Assessment for Nuclear Facilities. Cambridge University Press, Cambridge, pp. 142155.Google Scholar
Lund, B., Roberts, R. and Smith, C. A. (2017). Review of paleo-, historical and current seismicity in Sweden and surrounding areas with implications for the seismic analysis underlying SKI report 92:3. Swedish Radiation Safety Authority Report, 2017:35.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
Mantovani, M. and Scherneck, H.-G. (2013). DInSAR investigation in the Pärvie end-glacial fault region, Lapland, Sweden. International Journal of Remote Sensing, 34(23), 84918502, doi.org/10.1080/01431161.2013.843871.CrossRefGoogle Scholar
Mather, W. W. (1843). Geology of New-york. Part 1, Comprising Geology of the First Geological District. Carroll & Cook, Albany, New York.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, 334338, doi.org/10.1080/11035897.2015.1036360.Google Scholar
Mohr, P. (1986). Possible Late Pleistocene faulting in Iar (west) Connacht, Ireland. Geological Magazine, 123, 545552, doi.org/10.1017/S0016756800035135.Google Scholar
Mörner, N.-A. (2005). An interpretation and catalogue of paleoseismicity in Sweden. Tectonophysics, 408, 265307, doi.org/10.1016/j.tecto.2005.05.039.Google Scholar
Muir Wood, R. (1993). A Review of the Seismotectonics of Sweden. SKB Technical Report TR-93-13, Swedish Nuclear Fuel and Waste Management Co., Stockholm.Google Scholar
Munier, R. and Fenton, C. (2004). Review of postglacial faulting. In Munier, R. and Hökmark, H., eds., Respect Distances. SKB Technical Report TR-04-17, Swedish Nuclear Fuel and Waste Management Co., Stockholm, pp. 157218.Google Scholar
Munier, R., Adams, J., Brandes, C. et al. (2020). International Database of Glacially-Induced Faults, PANGAEA, doi.org/10.1594/PANGAEA.922705.Google Scholar
Munthe, H. (1905). Om en sen- eller postglacial förkastning vid Allebergsände i Västergötland och om en postglacial rubbning i silurlagren SV om Visby [About a Lateglacial or postglacial fault at Allebergsände in Västergötland and about a postglacial disturbance in the Silurian deposits SW of Visby]. Geologiska Föreningens i Stockholm Förhandlingar, 27(6), 346.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
Olesen, O. (1988). The Stuoragurra fault, evidence of neotectonics in the Precambrian of Finnmark, northern Norway. Norsk Geologisk Tidskrift, 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., 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 Vol. 13, pp. 145174.Google Scholar
Oliver, J., Johnston, T. and Dorman, J. (1970). Postglacial faulting and seismicity in New York and Quebec. Canadian Journal of Earth Sciences, 7, 579590, doi.org/10.1139/e70-059.Google Scholar
Olsen, L., Olesen, O. and Høgaas, F. (2020). Dating of the Stuoragurra Fault at Finnmarksvidda, northern Norway. Abstracts and Proceedings of the Geological Society of Norway 1The 34th Nordic Geological Winter Meeting, Oslo, pp. 157–158.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
Peltier, W. R. and Andrews, J. T. (1976). Glacial‐isostatic adjustment – I. The forward problem. Geophysical Journal of the Royal Astronomical Society, 46, 605646, doi.org/10.1111/j.1365-246X.1976.tb01251.x.CrossRefGoogle Scholar
Reusch, H. (1888). Bømmeløen og Karmøen med omgivelser [Bømmeløen and Karmøen with Surroundings]. Norges geologiske undersøkelse, Norway.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
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.CrossRefGoogle 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
Steffen, R., Wu, P., Steffen, H. and Eaton, D. W. (2014). 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., Sleep, N. H., Geller, R. J., Wang, S.‐C. and Kroeger, G. C. (1979). Earthquakes along the passive margin of eastern Canada. Geophysical Research Letters, 6, 537540, doi.org/10.1029/GL006i007p00537.CrossRefGoogle Scholar
Stewart, I. S., Firth, C. R., Rust, D. J., Collins, P. E. F. and Firth, J. A. (2001). Postglacial fault movement and palaeoseismicity in western Scotland: a reappraisal of the Kinloch Hourn fault, Kintail. Journal of Seismology, 5 , 307328, doi.org/10.1023/A:1011467307511.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.CrossRefGoogle Scholar
Tanner, V. (1930). Om nivåförändringarna och grunddragen av den geografiska utvecklingen efter istiden i Ishavsfinland samt om homotaxin av Fennoskandias kvartära marina avlagringar. Studier över kvartärsystemet i Fennoskandias nordliga delar – IV. [On level changes and basic features of the geographical development after the ice age in the Polar Sea of Finland and on the homotaxis of Fennoscandia’s Quaternary marine deposits. Studies of the Quaternary System in Northern Fennoscandia – IV]. Bulletin de la Commission Géologique de Finlande, 88, 594 pp.Google Scholar
van Loon, A. J. 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., 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 Vliet-Lanoë, B., Bonnet, S., Hallegouët, B. and Laurent, M. (1997). Neotectonic and seismic activity in the Armorican and Cornubian Massifs: regional stress field with glacio-isostatic influence? Journal of Geodynamics, 24(1–4), 219239, doi.org/10.1016/S0264-3707(96)00035-X.Google Scholar
Wessel, P., Smith, W. H. F., Scharroo, R., Luis, J. F. and Wobbe, F. (2013). Generic Mapping Tools: improved version released. EOS Transactions American Geophysical Union, 94, 409410, doi.org/10.1002/2013EO450001.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.CrossRefGoogle 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
Wu, P. and Peltier, W. R. (1982). Viscous gravitational relaxation. Geophysical Journal of the Royal Astronomical Society, 70, 435486, doi.org/10.1111/j.1365-246X.1982.tb04976.x.Google Scholar

References

Anderson, E. M. (1951). The Dynamics of Faulting and Dyke Formation with Applications to Britain, 2nd ed., Oliver & Boyd, Edinburgh.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., 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. 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
Byerlee, J. D. (1978). Friction of rock. Pure and Applied Geophysics, 116, 615626, doi.org/10.1007/BF00876528.Google Scholar
Cathles, L. M. III (1975). The Viscosity of the Earth’s Mantle. Princeton University Press, Princeton.Google Scholar
Coulomb, C. A. (1776). Essai sur une application des règles des maximis et minimis à quelquels problemes de statique relatifs, à la architecture. Mémoires de mathématique et physique presenté à l’Acádemie des sciences par savantes étrangères, 7, 343382.Google Scholar
Di Toro, G., Han, R., Hirose, T. et al. (2011). Fault lubrication during earthquakes. Nature, 471, 494498, doi.org/10.1038/nature09838.Google Scholar
Gordon, R. G. (1998). The plate tectonic approximation: plate non-rigidity, diffuse plate boundaries, and global plate reconstructions. Annual Review of Earth and Planetary Sciences, 26, 615642, doi.org/10.1146/annurev.earth.26.1.615.Google Scholar
Handy, M. R. and Brun, J.-P. (2004). Seismicity, structure and strength of the continental lithosphere. Earth and Planetary Science Letters, 223, 427441, doi.org/10.1016/j.epsl.2004.04.021.Google Scholar
Harris, R. A. (1998). Introduction to special section: stress triggers, stress shadows, and implications for seismic hazard. Journal of Geophysical Research, 103(B10), 2434724358, doi.org/10.1029/98JB01576.Google Scholar
Heidbach, O., Rajabi, M., Cui, X. et al. (2018). The World Stress Map database release 2016: crustal stress pattern across scales. Tectonophysics, 744, 484498, doi.org/10.1016/j.tecto.2018.07.007.Google Scholar
Heyman, J. (1972). Coulomb’s Memoir on Statics. Cambridge University Press, London.Google Scholar
Hjartardóttir, Á. R., Einarsson, P. and Brandsdóttir, B. (2011). The Kerlingar fault, Northeast Iceland: a Holocene normal fault east of the divergent plate boundary. Jökull, 60, 103116.Google Scholar
Ishibe, T., Shimazaki, K., Tsuruoka, H., Yamanaka, Y. and Satake, K. (2011). Correlation between Coulomb stress changes imparted by large historical strike-slip earthquakes and current seismicity in Japan. Earth, Planets and Space, 63, 12, doi.org/10.5047/eps.2011.01.008.CrossRefGoogle Scholar
Jaeger, J. C., Cook, N. G. W. and Zimmerman, R. W. (2007). Fundamentals of Rock Mechanics. Blackwell Publishing, Malden, Massachusetts.Google Scholar
Johnson, R. B. and DeGraff, J. V. (1988). Principles of Engineering Geology. John Wiley & Sons, New York.Google Scholar
Johnston, A. C. (1987). Suppression of earthquakes by large continental ice sheets. Nature, 330, 467469, doi.org/10.1038/330467a0.Google Scholar
Kaufmann, G., Wu, P. and Ivins, E. R. (2005). Lateral viscosity variations beneath Antarctica and their implications on regional rebound motions and seismotectonics. Journal of Geodynamics, 39, 165181, doi.org/10.1016/j.jog.2004.08.009.Google Scholar
King, G. C. P., Stein, R. S. and Lin, J. (1994). Static stress changes and the triggering of earthquakes. Bulletin of the Seismological Society of America, 84(3), 935953.Google Scholar
Klemann, V. and Wolf, D. (1998). Modelling of stresses in the Fennoscandian lithosphere induced by Pleistocene glaciations. Tectonophysics, 294, 291303, doi.org/10.1016/S0040-1951(98)00107-3.Google Scholar
Lagerbäck, R. and Sundh, M. (2008). Early Holocene Faulting and Paleoseismicity in Northern Sweden. SGU Research Paper, C386, 80 pp.Google Scholar
Lund, B. (2005). Effects of Deglaciation on the Crustal Stress Field and Implications for Endglacial Faulting: A Parametric Study of Simple Earth and Ice Models. SKB Technical Report TR-05-04, Swedish Nuclear Fuel and Waste Management Co., Stockholm, 68 pp.Google Scholar
Lund, B., Schmidt, P. and Hieronymus, C. (2009). Stress Evolution and Fault Stability during the Weichselian Glacial Cycle. SKB Technical Report TR-09-15, Swedish Nuclear Fuel and Waste Management Co., Stockholm, 106 pp.Google Scholar
Minster, J. B. and Anderson, D. L. (1980). Dislocations and nonelastic processes in the mantle. Journal of Geophysical Research, 85(B11), 63476352, doi.org/10.1029/JB085iB11p06347.Google Scholar
Mohr, O. (1914). Abhandlungen aus dem Gebiete der Technische Mechanik. Treatise on Topics in Engineering Mechanics, 2nd ed., Ernst und Sohn, Berlin.Google Scholar
Nur, A. and Booker, J. R. (1972). Aftershocks caused by pore fluid flow? Science, 175, 885887, doi.org/10.1126/science.175.4024.885.Google Scholar
Peltier, W. R. (1974). The impulse response of a Maxwell Earth. Reviews of Geophysics and Space Physics, 12, 649669, doi.org/10.1029/RG012i004p00649.Google Scholar
Ranalli, G. (1995). Rheology of the Earth, 2nd ed., Chapman & Hall, London.Google Scholar
Sauber, J. M. and Ruppert, N. (2013). Rapid ice mass loss: does it have an influence on earthquake occurrence in Southeast Alaska? In Freymueller, J. T., Haeussler, P. J., Wesson, R. L. and Ekström, G., eds., Active Tectonics and Seismic Potential of Alaska. American Geophysical Union, Geophysical Monograph Series, Vol. 179, doi.org/10.1029/179GM21.Google Scholar
Sauber, J., Plafker, G., Molnia, B. F. and Bryant, M. A. (2000). Crustal deformation associated with glacial fluctuations in the eastern Chugach Mountains, Alaska. Journal of Geophysical Research, 105, 80558077, doi.org/10.1029/1999JB900433.Google Scholar
Scholz, C. H. (2019). The Mechanics of Earthquakes and Faulting, 3rd ed., Cambridge University Press, doi.org/10.1017/9781316681473.CrossRefGoogle Scholar
Sibson, R. H. (1985). A note on fault reactivation. Journal of Structural Geology, 7, 751754, doi.org/10.1016/0191-8141(85)90150-6.Google Scholar
Steffen, R., Wu, P., Steffen, H. and Eaton, D.W. (2014a). 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
Steffen, R., Wu, P., Steffen, H. and Eaton, D. W. (2014b). The effect of earth rheology and ice-sheet size on fault slip and magnitude of postglacial earthquakes. Earth and Planetary Science Letters, 388, 7180, doi.org/10.1016/j.epsl.2013.11.058.Google Scholar
Steffen, H., Steffen, R. and Tarasov, L. (2019). Modelling of glacially-induced stress changes in Latvia, Lithuania and the Kaliningrad District of Russia. Baltica, 32(1), 7890, doi.org/10.5200/baltica.2019.1.7.Google Scholar
Stein, R. S. (1999). The role of stress transfer in earthquake occurrence. Nature, 402, 605609, doi.org/10.1038/45144.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
Turcotte, D. L. and Schubert, G. (1982). Geodynamics – Applications of Continuum Physics to Geological Problems. Wiley, New York.Google Scholar
Twiss, R. J. and Moores, E. M. (2007). Structural Geology, 2nd ed., W. H. Freeman, New York.Google Scholar
Walcott, R. I. (1970). Isostatic response to loading of the crust in Canada. Canadian Journal of Earth Sciences, 7, 716727, doi.org/10.1139/e70-070.Google Scholar
Watts, A. B. and Burov, E. B. (2003). Lithospheric strength and its relationship to the elastic and seismogenic layer thickness. Earth and Planetary Science Letters, 213, 113131, doi.org/10.1016/S0012-821X(03)00289-9.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
Yin, Z.-M. and Ranalli, G. (1995). Estimation of the frictional strength of faults from inversion of fault-slip data: a new method. Journal of Structural Geology, 17, 13271335, doi.org/10.1016/0191-8141(95)00028-C.Google Scholar
Zoback, M. L. (1992). First and second order patterns of stress in the lithosphere: The World Stress Map Project. Journal of Geophysical Research, 97, 1170311728, doi.org/10.1029/92jb00132.Google Scholar
Zoback, M. D. and Townend, J. (2001). Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere. Tectonophysics, 336, 1930, doi.org/10.1016/S0040-1951(01)00091-9.CrossRefGoogle Scholar
Zoback, M. L. and Zoback, M. (2015). Lithosphere stress and deformation. Treatise on Geophysics, 6, 255271, doi.org/10.1016/B978-0-444-53802-4.00115-9.Google Scholar
Zoback, M. L., Zoback, M. D., Adams, J. et al. (1989). Global patterns of tectonic stress. Nature, 341, 291298, doi.org/10.1038/341291a0.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
×