Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-20T05:23:17.037Z Has data issue: false hasContentIssue false

8 - Seismicity Induced by the Development of Unconventional Oil and Gas Resources

from Part II - Environmental Analysis

Published online by Cambridge University Press:  28 July 2022

John Stolz
Affiliation:
Duquesne University, Pittsburgh
Daniel Bain
Affiliation:
University of Pittsburgh
Michael Griffin
Affiliation:
Carnegie Mellon University, Pennsylvania
Get access

Summary

Resource development in unconventional oil and gas plays is sometimes accompanied by unintended earthquakes, known as induced seismicity. To date, the largest such induced events have been the September 2016 5.8 MW Pawnee earthquake in Oklahoma, and the December 2018 5.2 MW earthquake in the Sichuan Basin. These earthquakes were triggered by different industrial processes, namely saltwater disposal (Pawnee) and hydraulic fracturing (Sichuan Basin). Current models indicate that such induced earthquakes occur by activation of a pre-existing fault system due to some combination of increased pore pressure, a change in fault-loading conditions arising from poroelastic effects, or precursory slow fault slip. This chapter provides a tutorial and review of basic underlying principles of induced seismicity and an overview of regulatory measures, along with several current research themes including tools for screening risk and forecasting maximum magnitude. These concepts are illustrated by case studies from the USA and western Canada.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2022

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

AER (Alberta Energy Regulator). (2019). Subsurface Order No. 6. www.aer.ca/documents/orders/subsurface-orders/SO6.pdfGoogle Scholar
Aki, K. (1965). Maximum likelihood estimate of b in the formula log N = abM and its confidence limits. Bulletin of the Earthquake Research Institute. 43(1): 237239.Google Scholar
Anderson, EM. (1951). The Dynamics of Faulting and Dyke Formation with Applications to Britain. Hafner Pub. Co.Google Scholar
Atkinson, GM, Eaton, DW, Ghofrani, H, Walker, D, Cheadle, B, Schultz, R, Shcherbakov, R, Tiampo, K, Gu, J, Harrington, RM, and Liu, Y. (2016). Hydraulic fracturing and seismicity in the western Canada sedimentary basin. Seismological Research Letters. 87(3): 631647.CrossRefGoogle Scholar
Atkinson, GM, Eaton, DW, Igonin, N. (2020). Developments in understanding seismicity triggered by hydraulic fracturing. Nature Reviews Earth & Environment. 1(1): 264277.Google Scholar
Babaie Mahani, AB, Schultz, R, Kao, H, Walker, D, Johnson, J, and Salas, C. (2017). Fluid injection and seismic activity in the northern Montney play, British Columbia, Canada, with special reference to the 17 August 2015 Mw 4.6 induced earthquake. Bulletin of the Seismological Society of America. 107(2): 542552.Google Scholar
Babaie Mahani, A, Kao, H, Atkinson, GM, Assatourians, K, Addo, K, and Liu, Y. (2019). Ground‐motion characteristics of the 30 November 2018 injection‐induced earthquake sequence in Northeast British Columbia, Canada.Seismological Research Letters. 90(4): 14571467.Google Scholar
Babaie Mahani, A, Esfahani, F, Kao, H, Gaucher, M, Hayes, M, Visser, R, and Venables, S. (2020). A systematic study of earthquake source mechanism and regional stress field in the Southern Montney Unconventional Play of northeast British Columbia, Canada. Seismological Research Letters. 91(1): 195206.Google Scholar
Bao, X, and Eaton, DW. (2016). Fault activation by hydraulic fracturing in western Canada. Science. 354(6318): 14061409.Google Scholar
Barbour, AJ, Norbeck, JH, and Rubinstein, JL. (2017). The effects of varying injection rates in Osage County, Oklahoma, on the 2016 Mw 5.8 Pawnee earthquake. Seismological Research Letters. 88(4): 10401053.Google Scholar
Barclay, JE, Krause, FF, Campbell, RI, and Utting, J. (1990). Dynamic casting and growth faulting: Dawson Creek graben complex, Carboniferous-Permian Peace River embayment, western Canada. Bulletin of Canadian Petroleum Geology. 38(1): 115145.Google Scholar
BC OGC (British Columbia Oil and Gas Commission). (2014a). Montney Formation Play Atlas NEBC. www.bcogc.ca/node/8131/downloadGoogle Scholar
BC OGC (British Columbia Oil and Gas Commission). (2014b). Investigation of observed seismicity in the Montney trend. www.bcogc.ca/sites/default/files/documentation/technical-reports/investigation-observed-seismicity-montney-trend.pdfGoogle Scholar
BC OGC (British Columbia Oil and Gas Commission). (2017). Application guideline for deep well disposal of produced water. www.bcogc.ca/application-guideline-deep-well-disposal-produced-water-non-hazardous-waste.Google Scholar
Bhattacharya, P and Viesca, RC. (2019). Fluid-induced aseismic fault slip outpaces pore-fluid migration. Science. 364(6439): 464468.Google Scholar
Biot, MA. (1941). General theory of three‐dimensional consolidation. Journal of Applied Physics. 12(2): 155164.Google Scholar
Bommer, JJ, Oates, S, Cepeda, JM, Lindholm, C, Bird, J, Torres, R, Marroquín, G, and Rivas, J. (2006). Control of hazard due to seismicity induced by a hot fractured rock geothermal project. Engineering Geology. 83(4): 287306.Google Scholar
Cardott, BJ. (2012). Thermal maturity of Woodford Shale gas and oil plays, Oklahoma, USA. International Journal of Coal Geology. 103(1): 109119.Google Scholar
CCA (Council of Canadian Academies). (2014). Environmental Impacts of Shale Gas Extraction in Canada. Council of Canadian Academies.Google Scholar
Clarke, H, Eisner, L, Styles, P, and Turner, P. (2014). Felt seismicity associated with shale gas hydraulic fracturing: The first documented example in Europe. Geophysical Research Letters. 41(23): 83088314.Google Scholar
Clarke, H, Verdon, JP, Kettlety, T, Baird, AF, and Kendall, JM. (2019). Real‐time imaging, forecasting, and management of human‐induced seismicity at Preston New Road, Lancashire, England. Seismological Research Letters. 90(5): 19021915.Google Scholar
Darold, A, Holland, AA, Chen, C, and Youngblood, A. (2014). Preliminary Analysis of Seismicity near Eagleton 1–29, Carter County, July 2014. Oklahoma Geological Survey Open File Report, OF2–2014.Google Scholar
Deng, K, Liu, Y, and Harrington, RM. (2016). Poroelastic stress triggering of the December 2013 Crooked Lake, Alberta, induced seismicity sequence. Geophysical Research Letters. 43(16): 84828491.Google Scholar
Dieterich, J. (1994). A constitutive law for rate of earthquake production and its application to earthquake clustering. Journal of Geophysical Research: Solid Earth. 99(B2): 26012618.CrossRefGoogle Scholar
Ducros, M, Sassi, W, Vially, R, Euzen, T, and Crombez, V. (2017). 2-D basin modeling of the Western Canada Sedimentary Basin across the Montney-Doig system: Implications for hydrocarbon migration pathways and unconventional resources potential. In AbuAli, MA Moretti, I, and Bolås, HMN (eds.) Memoir 114: Petroleum Systems Analysis: Case Studies, Tulsa, OK. American Association of Petroleum Geologists.Google Scholar
Dziewonski, AM, Chou, TA, and Woodhouse, JH. (1981). Determination of earthquake source parameters from waveform data for studies of global and regional seismicity. Journal of Geophysical Research: Solid Earth. 86(B4): 28252852.Google Scholar
Eaton, DW. (2018). Passive Seismic Monitoring of Induced Seismicity: Fundamental Principles and Application to Energy Technologies. Cambridge University Press.Google Scholar
Eaton, DW and Igonin, N. (2018). What controls the maximum magnitude of injection-induced earthquakes? The Leading Edge. 37(2): 135140.Google Scholar
Eaton, DW anf Maghsoudi, S. (2015). 2b… or not 2b? Interpreting magnitude distributions from microseismic catalogs. First Break. 33(10): 7986.Google Scholar
Eaton, DW and Schultz, R. (2018). Increased likelihood of induced seismicity in highly overpressured shale formations. Geophysical Journal International. 214(1): 751757.Google Scholar
Eaton, DW, Igonin, N, Poulin, A, Weir, R, Zhang, H, Pellegrino, S, and Rodriguez, G. (2018). Induced seismicity characterization during hydraulic‐fracture monitoring with a shallow‐wellbore geophone array and broadband sensors. Seismological Research Letters. 89(5): 16411651.Google Scholar
Eaton, DW, Milkereit, B, Ross, GM, Kanasewich, ER, Geis, W, Edwards, DJ, Kelsch, L, and Varsek, J. (1995). Lithoprobe basin-scale seismic profiling in central Alberta: Influence of basement on the sedimentary cover. Bulletin of Canadian Petroleum Geology. 43(1): 6577.Google Scholar
Edwards, DE, Barclay, J, Gibson, D, Kvill, G, and Halton, E. (1994). Triassic strata of the Western Canada Sedimentary Basin. In Mossop, GD and Shetsen, I (eds.) Geological Atlas of the Western Canada Sedimentary Basin. Calgary, AB: Canadian Society of Petroleum Geologists and Alberta Research Council.Google Scholar
EIA (Energy Information Administration). (2015). World Shale Resource Assessments, www.eia.gov/analysis/studies/worldshalegas.Google Scholar
Ekström, G, Nettles, M, and Dziewoński, AM. (2012). The global CMT project 2004–2010: Centroid-moment tensors for 13,017 earthquakes. Physics of the Earth and Planetary Interiors. 200(1): 19.Google Scholar
Ellsworth, WL. (2013). Injection-induced earthquakes. Science: 341(6142), 1225942, https://doi.org/10.1126/science.1225942.Google Scholar
Ellsworth, WL, Llenos, AL, McGarr, AF, Michael, AJ, Rubinstein, JL, Mueller, CS, Petersen, MD, and Calais, E. (2015). Increasing seismicity in the US midcontinent: Implications for earthquake hazard. The Leading Edge. 34(6): 618626.Google Scholar
Eyre, TS, Eaton, DW, Zecevic, M, D’Amico, D, and Kolos, D. (2019a). Microseismicity reveals fault activation before MW 4.1 hydraulic-fracturing induced earthquake. Geophysical Journal International. 218(1): 534546.Google Scholar
Eyre, TS, Eaton, DW, Garagash, DI, Zecevic, M, Venieri, M, Weir, R, and Lawton, DC. (2019b). The role of aseismic slip in hydraulic fracturing–induced seismicity. Science Advances. 5(8): eaav7172. https://doi.org/10.1126/sciadv.aav7172CrossRefGoogle ScholarPubMed
Foulger, GR, Wilson, M, Gluyas, J, Julian, BR, and Davies, R. (2018). Global review of human-induced earthquakes. Earth-Science Reviews. 178(1), 438514.Google Scholar
Friberg, PA, Besana‐Ostman, GM, and Dricker, I. (2014). Characterization of an earthquake sequence triggered by hydraulic fracturing in Harrison County, Ohio. Seismological Research Letters. 85(6), 12951307.CrossRefGoogle Scholar
Frohlich, C, DeShon, H, Stump, B, Hayward, C, Hornbach, M, and Walter, JI. (2016). A historical review of induced earthquakes in Texas. Seismological Research Letters. 87(4): 10221038.Google Scholar
Galis, M, Ampuero, JP, Mai, PM, and Cappa, F. (2017). Induced seismicity provides insight into why earthquake ruptures stop. Science Advances. 3(12): eaap7528, https://doi.org/10.1126/sciadv.aap7528.Google Scholar
Gan, W and Frohlich, C. (2013). Gas injection may have triggered earthquakes in the Cogdell oil field, Texas. Proceedings of the National Academy of Sciences. 110(47): 1878618791.Google Scholar
Ghofrani, H and Atkinson, GM. (2020). Activation rate of seismicity for hydraulic fracture wells in the Western Canada Sedimentary Basin. Bulletin of the Seismological Society of America, in press.Google Scholar
Goebel, THW, Weingarten, M, Chen, X, Haffener, J, and Brodsky, EE. (2017). The 2016 Mw5. 1 Fairview, Oklahoma earthquakes: Evidence for long-range poroelastic triggering at > 40 km from fluid disposal wells. Earth and Planetary Science Letters. 472(1): 5061.Google Scholar
Gómez-Alba, S, Vargas, C, and Zang, A. (2020). Evidencing the relationship between injected volume of water and maximum expected magnitude during the Puerto Gaitán (Colombia) earthquake sequence from 2013 to 2015. Geophysical Journal International. 220(1): 335344.CrossRefGoogle Scholar
Guglielmi, Y, Cappa, F, Avouac, JP, Henry, P, and Elsworth, D. (2015). Seismicity triggered by fluid injection–induced aseismic slip. Science. 348(6240): 12241226.Google Scholar
Gutenberg, B. and Richter, CF. (1944). Frequency of earthquakes in California. Bulletin of the Seismological Society of America. 34(4): 185188.Google Scholar
Hanks, TC and Kanamori, H. (1979). A moment magnitude scale. Journal of Geophysical Research: Solid Earth. 84(B5): 23482350.Google Scholar
Häring, MO, Schanz, U, Ladner, F, and Dyer, BC. (2008). Characterisation of the Basel 1 enhanced geothermal system. Geothermics. 37(5): 469495.Google Scholar
Healy, JH, Rubey, WW, Griggs, DT, and Raleigh, CB. (1968). The Denver earth-quakes. Science. 161(3848): 13011310.Google Scholar
Heidbach, O, Rajabi, M, Reiter, K, and Ziegler, M. (2016): World Stress Map Database Release 2016. GFZ Data Services, https://doi.org/10.5880/WSM.2016.001.Google Scholar
Heidbach, O, Rajabi, M, Cui, X, Fuchs, K, Müller, B, Reinecker, J, Reiter, K, Tingay, M, Wenzel, F, Xie, F, and Ziegler, MO. (2018). The World Stress Map database release 2016: Crustal stress pattern across scales. Tectonophysics. 744(1): 484-498.Google Scholar
Hennings, PH, Lund Snee, JE, Osmond, JL, DeShon, HR, Dommisse, R, Horne, E, Lemons, C, and Zoback, MD (2019). Injection‐induced seismicity and fault‐slip potential in the Fort Worth Basin, Texas. Bulletin of the Seismological Society of America. 109(5): 16151634.Google Scholar
Herrmann, RB.(2020). North America Moment Tensors. www.eas.slu.edu/eqc/eqc_mt/MECH.NA/.Google Scholar
Herrmann, RB, Park, SK, and Wang, CY. (1981). The Denver earthquakes of 1967–1968. Bulletin of the Seismological Society of America. 71(3), 731745.Google Scholar
Herrmann, RB, Benz, H, and Ammon, CJ. (2011). Monitoring the earthquake source process in North America. Bulletin of the Seismological Society of America. 101(6): 26092625.Google Scholar
Holland, A. (2013). Earthquakes triggered by hydraulic fracturing in south-central Oklahoma. Bulletin of the Seismological Society of America. 103(3): 17841792.Google Scholar
Holmgren, JM, Atkinson, GM, and Ghofrani, H. (2019). Stress drops and directivity of induced earthquakes in the Western Canada Sedimentary Basin. Bulletin of the Seismological Society of America. 109(5): 16351652.Google Scholar
Hosseini, BK and Eaton, DW. (2018). Fluid flow and thermal modeling for tracking induced seismicity near the Graham disposal well, British Columbia, Canada. SEG Technical Program Expanded Abstracts 2018 (pp. 4987–4991). Society of Exploration Geophysicists, https://doi.org/10.1190/segam2018–2996360.1.Google Scholar
Igonin, N, Zecevic, M, and Eaton, DW. (2018). Bilinear magnitude‐frequency distributions and characteristic earthquakes during hydraulic fracturing. Geophysical Research Letters. 45(23): 1286612874.Google Scholar
Ishimoto, M and Iida, K. (1939). Observations of earthquakes registered with the micro-seismograph constructed recently. Bulletin of the Earthquake Research Institute. 17(1): 443478.Google Scholar
Jia, SQ, Wong, RCK, Eaton, DW, and Eyre, TS. (2018). Investigating fracture growth and source mechanisms in shale using acoustic emission technique. In 52nd US Rock Mechanics/Geomechanics Symposium. American Rock Mechanics Association, ARMA-2018-136.Google Scholar
Kanamori, H and Anderson, DL. (1975). Theoretical basis of some empirical relations in seismology. Bulletin of the Seismological Society of America. 65(5): 10731095.Google Scholar
Kang, JQ, Zhu, JB, and Zhao, J. (2019). A review of mechanisms of induced earthquakes: from a view of rock mechanics. Geomechanics and Geophysics for Geo-Energy and Geo-Resources. 5(2): 171196.Google Scholar
Kao, H, Visser, R, Smith, B, and Venables, S. (2018). Performance assessment of the induced seismicity traffic light protocol for northeastern British Columbia and western Alberta. The Leading Edge. 37(2): 117126.Google Scholar
Keranen, KM and Weingarten, M. (2018). Induced seismicity. Annual Review of Earth and Planetary Sciences. 46(1): 149174.Google Scholar
Keranen, KM, Savage, HM, Abers, GA, and Cochran, ES. (2013). Potentially induced earthquakes in Oklahoma, USA: Links between wastewater injection and the 2011 Mw 5.7 earthquake sequence. Geology. 41(6): 699702.Google Scholar
Keranen, KM, Weingarten, M, Abers, GA, Bekins, BA, and Ge, S. (2014). Sharp increase in central Oklahoma seismicity since 2008 induced by massive wastewater injection. Science. 345, 448451.Google Scholar
Kettlety, T, Verdon, JP, Werner, MJ, and Kendall, JM. (2020). Stress transfer from opening hydraulic fractures controls the distribution of induced seismicity. Journal of Geophysical Research: Solid Earth. 125: e2019JB018794. https://doi.org/10.1029/2019JB018794Google Scholar
Kettlety, T, Verdon, JP, Werner, MJ, Kendall, JM, and Budge, J. (2019). Investigating the role of elastostatic stress transfer during hydraulic fracturing-induced fault activation. Geophysical Journal International. 217(2): 12001216.Google Scholar
Kim, YS, Peacock, DC, and Sanderson, DJ. (2004). Fault damage zones. Journal of Structural Geology. 26(3): 503517.CrossRefGoogle Scholar
Knopoff, L. (2000). The magnitude distribution of declustered earthquakes in Southern California. Proceedings of the National Academy of Sciences. 97(22): 1188011884.Google Scholar
Kohli, AH and Zoback, MD. (2013). Frictional properties of shale reservoir rocks. Journal of Geophysical Research: Solid Earth. 118(9): 51095125.Google Scholar
Kozłowska, M, Brudzinski, MR, Friberg, P, Skoumal, RJ, Baxter, ND, and Currie, BS. (2018). Maturity of nearby faults influences seismic hazard from hydraulic fracturing. Proceedings of the National Academy of Sciences. 115(8): E1720E1729.Google Scholar
Kwiatek, G, Saarno, T, Ader, T, Bluemle, F, Bohnhoff, M, Chendorain, M, Dresen, G, Heikkinen, P, Kukkonen, I, Leary, P, and Leonhardt, M. (2019). Controlling fluid-induced seismicity during a 6.1-km-deep geothermal stimulation in Finland. Science Advances. 5(5): eaav7224, https://doi.org/10.1126/sciadv.aav7224.Google Scholar
Langenbruch, C and Zoback, MD. (2016). How will induced seismicity in Oklahoma respond to decreased saltwater injection rates? Science Advances. 2(11): e1601542. https://doi.org/10.1126/sciadv.1601542.Google Scholar
Lavrov, A. (2016). Dynamics of stresses and fractures in reservoir and cap rock under production and injection. Energy Procedia. 86(1): 381390.Google Scholar
Lei, X, Wang, Z, and Su, J. (2019). The December 2018 ML 5.7 and January 2019 ML 5.3 Earthquakes in South Sichuan Basin Induced by Shale Gas Hydraulic Fracturing. Seismological Research Letters. 90(3): 10991110.Google Scholar
Lengliné, O, Lamourette, L, Vivin, L, Cuenot, N, and Schmittbuhl, J. (2014). Fluid‐induced earthquakes with variable stress drop. Journal of Geophysical Research: Solid Earth. 119(12): 89008913.Google Scholar
Ludwig, J, Nafe, J, and Drake, C. (1971). Seismic refraction. In Maxwell, AE (ed.) The Sea, Vol. 4 . Wiley, pp. 5384.Google Scholar
MacKay, MK, Eaton, DW, Pedersen, PK, and Clarkson, CR. (2018). Integration of outcrop, subsurface, and microseismic interpretation for rock-mass characterization: An example from the Duvernay Formation, Western Canada. Interpretation 6(4): T919T936.Google Scholar
Mahani, AB, Schultz, R, Kao, H, Walker, D, Johnson, J, and Salas, C. (2017). Fluid injection and seismic activity in the northern Montney play, British Columbia, Canada, with special reference to the 17 August 2015 MW 4.6 induced earthquake. Bulletin of the Seismological Society of America. 107(2): 542552.Google Scholar
Marone, C. (1998). Laboratory-derived friction laws and their application to seismic faulting. Annual Review of Earth and Planetary Sciences. 26(1): 643696.Google Scholar
Marsh, S and Holland, A. (2016). Comprehensive Fault Database and Interpretive Fault Map of Oklahoma. Oklahoma Geological Survey Open‐File Rep. OF22016.Google Scholar
McGarr, A. (2014). Maximum magnitude earthquakes induced by fluid injection. Journal of Geophysical Research: Solid Earth. 119(2): 10081019.Google Scholar
McGarr, A and Barbour, AJ. (2017). Wastewater disposal and the earthquake sequences during 2016 near Fairview, Pawnee, and Cushing, Oklahoma. Geophysical Research Letters. 44(18): 93309336.Google Scholar
McNamara, DE, Hayes, GP, Benz, HM, Williams, RA, McMahon, ND, Aster, RC, Holland, A, Sickbert, T, Herrmann, R, Briggs, R, and Smoczyk, G. (2015). Reactivated faulting near Cushing, Oklahoma: Increased potential for a triggered earthquake in an area of United States strategic infrastructure. Geophysical Research Letters. 42(20): 83288332.Google Scholar
NRC (National Research Council). (2013). Induced Seismicity Potential in Energy Technologies. National Academies Press.Google Scholar
Ohnaka, M, Akatsu, M, Mochizuki, H, Odedra, A, Tagashira, F, and Yamamoto, Y. (1997). A constitutive law for the shear failure of rock under lithospheric conditions. Tectonophysics. 277(1–3): 127.Google Scholar
Okada, Y. (1992). Internal deformation due to shear and tensile faults in a half-space. Bulletin of the Seismological Society of America. 82(2): 10181040.Google Scholar
OCC (Oklahoma Corporation Commission). (2018). Moving forward: New protocol to further address seismicity in state’s largest oil and gas play, www.occeweb.com/og/02-27-18PROTOCOL.pdf.Google Scholar
Oklahoma Produced Water Working Group. 2017. Oklahoma Water for 2060: Produced Water Reuse and Recycling. www.owrb.ok.gov/2060/pwwg.php.Google Scholar
Parry, RH. (2004). Mohr Circles, Stress Paths and Geotechnics, 2nd Edition, CRC Press.Google Scholar
Pawley, S, Schultz, R, Playter, T, Corlett, H, Shipman, T, Lyster, S, and Hauck, T. (2018). The geological susceptibility of induced earthquakes in the Duvernay play. Geophysical Research Letters. 45(4): 17861793.Google Scholar
Peterie, SL, Miller, RD, Intfen, JW, and Gonzales, JB. (2018). Earthquakes in Kansas induced by extremely far‐field pressure diffusion. Geophysical Research Letters. 45(3): 13951401.Google Scholar
Ponomarev, AV, Zavyalov, AD, Smirnov, VB, and Lockner, DA. (1997). Physical modeling of the formation and evolution of seismically active fault zones. Tectonophysics. 277(1–3). 5781.Google Scholar
Poulin, A, Weir, R, Eaton, D, Igonin, N, Chen, Y, Lines, L, and Lawton, D. (2019). Focal-time analysis: A new method for stratigraphic depth control of microseismicity and induced seismic events. Geophysics. 84(6): KS173-KS182.Google Scholar
Raleigh, CB, Healy, JH, and Bredehoeft, JD. (1976). An experiment in earthquake control at Rangely, Colorado. Science. 191(4233): 12301237.Google Scholar
Richter, CF. (1935). An instrumental earthquake magnitude scale. Bulletin of the Seismological Society of America: 25(1): 132.CrossRefGoogle Scholar
Roche, V and Van der Baan, M. (2015). The role of lithological layering and pore pressure on fluid‐induced microseismicity. Journal of Geophysical Research: Solid Earth. 120(2): 923943.Google Scholar
Rokosh, CD, Lyster, S, Anderson, SDA, Beaton, AP, Berhane, H, Brazzoni, T, Chen, D, Cheng, Y, Mack, T, Pana, C, and Pawlowicz, JG. (2012). Summary of Alberta’s Shale- and Siltstone-Hosted Hydrocarbon Resource Potential. Energy Resources Conservation Board, ERCB/AGS Open File Report, 2012-06.Google Scholar
Rubinstein, JL and Mahani, AB. (2015). Myths and facts on wastewater injection, hydraulic fracturing, enhanced oil recovery, and induced seismicity. Seismological Research Letters. 86(4): 10601067.Google Scholar
Schoenball, M and Ellsworth, WL. (2017). Waveform‐relocated earthquake catalog for Oklahoma and southern Kansas illuminates the regional fault network. Seismological Research Letters. 88(5): 12521258.Google Scholar
Schoenball, M, Walsh, FR, Weingarten, M, and Ellsworth, WL. (2018). How faults wake up: The Guthrie-Langston, Oklahoma earthquakes. The Leading Edge. 37(2): 100106.Google Scholar
Scholz, CH. (1998). Earthquakes and friction laws. Nature. 391(6662): 3742.Google Scholar
Schultz, R. and Pawley, S. (2019). Induced Earthquakes Geological Susceptibility Model for the Duvernay Formation, Central Alberta: Version 2. AER/AGS Open File Report 2019-02.Google Scholar
Schultz, R. and Wang, R. (2020). Newly emerging cases of hydraulic fracturing induced seismicity in the Duvernay East Shale Basin. Tectonophysics. 228393, https://doi.org/10.1016/j.tecto.2020.228393.Google Scholar
Schultz, R, Stern, V, Novakovic, M, Atkinson, G, and Gu, YJ. (2015). Hydraulic fracturing and the Crooked Lake Sequences: Insights gleaned from regional seismic networks. Geophysical Research Letters. 42(8): 27502758.Google Scholar
Schultz, R, Atkinson, G, Eaton, DW, Gu, YJ, and Kao, H. (2018). Hydraulic fracturing volume is associated with induced earthquake productivity in the Duvernay play. Science. 359(6373): 304308.Google Scholar
Schultz, R, Wang, R, Gu, YJ, Haug, K, and Atkinson, G. (2017). A seismological overview of the induced earthquakes in the Duvernay play near Fox Creek, Alberta. Journal of Geophysical Research: Solid Earth. 122(1): 492505.Google Scholar
Schultz, R, Skoumal, RJ, Brudzinski, MR, Eaton, DW, Baptie, B, and Ellsworth, WL. (2020). Hydraulic Fracturing Induced Seismicity. Reviews of Geophysics, submitted.Google Scholar
Schultz, R, Corlett, H, Haug, K, Kocon, K, MacCormack, K, Stern, V, and Shipman, T. (2016). Linking fossil reefs with earthquakes: Geologic insight to where induced seismicity occurs in Alberta. Geophysical Research Letters. 43(6), 25342542.Google Scholar
Segall, P and Lu, S. (2015). Injection‐induced seismicity: Poroelastic and earthquake nucleation effects. Journal of Geophysical Research: Solid Earth. 120(7): 50825103.Google Scholar
Shah, AK and Keller, GR. (2017). Geologic influence on induced seismicity: Constraints from potential field data in Oklahoma. Geophysical Research Letters. 44(1): 152161.Google Scholar
Shapiro, SA and Dinske, C. (2009). Fluid‐induced seismicity: Pressure diffusion and hydraulic fracturing. Geophysical Prospecting. 57(2): 301310.Google Scholar
Shapiro, SA, Huenges, E, and Borm, G. (1997). Estimating the crust permeability from fluid-injection-induced seismic emission at the KTB site. Geophysical Journal International. 131(2): F15F18.CrossRefGoogle Scholar
Shapiro, SA, Dinske, C, Langenbruch, C, and Wenzel, F. (2010). Seismogenic index and magnitude probability of earthquakes induced during reservoir fluid stimulations. The Leading Edge. 29(3): 304309.Google Scholar
Shapiro, SA, Krüger, OS, Dinske, C, and Langenbruch, C. (2011). Magnitudes of induced earthquakes and geometric scales of fluid-stimulated rock volumes. Geophysics. 76(6): WC55WC63.Google Scholar
Shapiro, SA, Patzig, R, Rothert, E, and Rindschwentner, J. (2003). Triggering of seismicity by pore-pressure perturbations: Permeability-related signatures of the phenomenon. Pure and Applied Geophysics. 160(5–6): 10511066.Google Scholar
Shapiro, SA, Rothert, E, Rath, V, and Rindschwentner, J. (2002). Characterization of fluid transport properties of reservoirs using induced microseismicity. Geophysics. 67(1): 212220.Google Scholar
Shearer, PM. (2019). Introduction to Seismology, 2nd Edition. Cambridge University Press.Google Scholar
Shemeta, JE, Brooks, CE, and Lord, CC. (2019). Well Stimulation Seismicity in Oklahoma: Cataloging Earthquakes Related to Hydraulic Fracturing. Unconventional Resources Technology Conference (URTEC), https://doi.org/10.15530/AP-URTEC-2019-198283.Google Scholar
Shipman, T, MacDonald, R, and Byrnes, T. (2018). Experiences and learnings from induced seismicity regulation in Alberta. Interpretation. 6(2): SE15SE21.Google Scholar
Skoumal, RJ, Brudzinski, MR, and Currie, BS. (2015). Earthquakes induced by hydraulic fracturing in Poland Township, Ohio. Bulletin of the Seismological Society of America. 105(1): 189197.Google Scholar
Skoumal, RJ, Ries, R, Brudzinski, MR, Barbour, AJ, and Currie, BS. (2018). Earthquakes induced by hydraulic fracturing are pervasive in Oklahoma. Journal of Geophysical Research: Solid Earth. 123(12): 1091810935.Google Scholar
Stoakes, FA. (1980). Nature and control of shale basin fill and its effect on reef growth and termination: upper Devonian Duvernay and Ireton formations of Alberta, Canada, Bulletin of Canadian Petroleum Geology. 28(3): 345410.Google Scholar
Switzer, SB, Holland, WG, Christie, DS, Graf, GC, Hedinger, AS, McAuley, RJ, Wierzbicki, RA, Packard, JJ, Mossop, GD, and Shetsen, I. (1994). Devonian Woodbend-Winterburn strata of the Western Canada Sedimentary Basin. In Mossop, GD and Shetsen, I (eds.) Geological Atlas of the Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council, pp. 165202.Google Scholar
Tenthorey, E and Cox, SF. (2006). Cohesive strengthening of fault zones during the interseismic period: An experimental study. Journal of Geophysical Research: Solid Earth. 111: B09202, https://doi:10.1029/2005JB004122.Google Scholar
Townend, J and Zoback, MD. (2000). How faulting keeps the crust strong. Geology. 28(5): 399402.Google Scholar
Trutnevyte, E and Wiemer, S. (2017). Tailor-made risk governance for induced seismicity of geothermal energy projects: An application to Switzerland. Geothermics. 65: 295312.Google Scholar
Van der Baan, M and Calixto, F. (2017). Human-induced seismicity and large-scale hydrocarbon production in the USA and Canada. Geochemistry, Geophysics, Geosystems. 18(7): 24672485.Google Scholar
Van der Elst, NJ, Page, MT, Weiser, DA, Goebel, TH, and Hosseini, SM. (2016). Induced earthquake magnitudes are as large as (statistically) expected. Journal of Geophysical Research: Solid Earth. 121(6): 45754590.Google Scholar
Vermilye, JM and Scholz, CH. (1998). The process zone: A microstructural view of fault growth. Journal of Geophysical Research: Solid Earth. 103(B6): 1222312237.Google Scholar
Walsh, FR and Zoback, MD. (2015). Oklahoma’s recent earthquakes and saltwater disposal. Science Advances. 1(5): e1500195. https://doi.org/10.1126/sciadv.1500195Google Scholar
Walsh, FRI, Zoback, MD, Pais, D, Weingartern, M, and Tyrell, T. (2017). FSP 1.0: A Program for Probabilistic Estimation of Fault Slip Potential Resulting from Fluid Injection, scits.stanford.edu/software.Google Scholar
Wang, R, Gu, YJ, Schultz, R, and Chen, Y. (2018). Faults and non‐double‐couple components for induced earthquakes. Geophysical Research Letters. 45(17): 89668975.Google Scholar
Wang, R, Gu, YJ, Schultz, R, Kim, A, and Atkinson, G. (2016). Source analysis of a potential hydraulic‐fracturing‐induced earthquake near Fox Creek, Alberta. Geophysical Research Letters. 43(2): 564573.Google Scholar
Wang, R, Gu, YJ, Schultz, R, Zhang, M, and Kim, A. (2017). Source characteristics and geological implications of the January 2016 induced earthquake swarm near Crooked Lake, Alberta. Geophysical Journal International. 210(2): 979988.CrossRefGoogle Scholar
Warpinski, NR, Mayerhofer, M, Agarwal, K, and Du, J. (2013). Hydraulic-fracture geomechanics and microseismic-source mechanisms. SPE Journal. 18(04): 766780.Google Scholar
Weingarten, M, Ge, S, Godt, JW, Bekins, BA, and Rubinstein, JL. (2015). High-rate injection is associated with the increase in US mid-continent seismicity. Science. 348(6241): 13361340.Google Scholar
Weir, RM, Eaton, DW, Lines, LR, Lawton, DC, and Ekpo, E. (2018). Inversion and interpretation of seismic-derived rock properties in the Duvernay play. Interpretation. 6(2): SE1SE14.Google Scholar
Wessel, P, Luis, JF, Uieda, L, Scharroo, R, Wobbe, F, Smith, WHF, and Tian, D. (2019). The Generic Mapping Tools version 6. Geochemistry, Geophysics, Geosystems. 20(1): 55565564.Google Scholar
Wilson, MP, Foulger, GR, Gluyas, JG, Davies, RJ, and Julian, BR. (2017). HiQuake: The human‐induced earthquake database. Seismological Research Letters. 88(6): 15601565.Google Scholar
Yao, Y. (2012). Linear elastic and cohesive fracture analysis to model hydraulic fracture in brittle and ductile rocks. Rock Mechanics and Rock Engineering. 45(3): 375387.Google Scholar
Yeck, WL, Weingarten, M, Benz, HM, McNamara, DE, Bergman, EA, Herrmann, RB, Rubinstein, JL, and Earle, PS. (2016). Far‐field pressurization likely caused one of the largest injection induced earthquakes by reactivating a large pre-existing basement fault structure. Geophysical Research Letters. 43(19): 10198.Google Scholar
Yu, H, Harrington, RM, Liu, Y, and Wang, B. (2019). Induced Seismicity Driven by Fluid Diffusion Revealed by a Near‐Field Hydraulic Stimulation Monitoring Array in the Montney Basin, British Columbia. Journal of Geophysical Research: Solid Earth. 124(5): 46944709.Google Scholar
Zhang, J. and Van der Baan, M. (2019). Depth-dependent fault slip potential. In SEG Technical Program Expanded Abstracts 2019 (pp. 30163020), Society of Exploration Geophysicists, https://doi.org/10.1190/segam2019–3214231.1.Google Scholar
Zhang, H, Eaton, DW, Rodriguez, G, and Jia, SQ. (2019). Source‐mechanism analysis and stress inversion for hydraulic‐fracturing‐induced event sequences near Fox Creek, Alberta. Bulletin of the Seismological Society of America. 109(2): 636651.Google Scholar
Zoback, MD. (2010). Reservoir Geomechanics, 1st edition, Cambridge University Press.Google Scholar
Zoback, MD and Harjes, HP. (1997). Injection‐induced earthquakes and crustal stress at 9 km depth at the KTB deep drilling site, Germany. Journal of Geophysical Research: Solid Earth. 102(B8): 1847718491.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
×