Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-7drxs Total loading time: 0 Render date: 2024-07-20T07:30:09.175Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  14 June 2019

Mark D. Zoback
Affiliation:
Stanford University, California
Arjun H. Kohli
Affiliation:
Stanford University, California
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
Unconventional Reservoir Geomechanics
Shale Gas, Tight Oil, and Induced Seismicity
, pp. 442 - 478
Publisher: Cambridge University Press
Print publication year: 2019

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

Abercrombie, R. E. (1995). Earthquake source scaling relationships from -1 to 5 ML using seismograms recorded at 2.5-km depth. Journal of Geophysical Research, 100(B12), 2401524036.Google Scholar
Adachi, J., Siebrits, E., Peirce, A., & Desroches, J. (2007). Computer simulation of hydraulic fractures. International Journal of Rock Mechanics and Mining Sciences, 44(5), 739757. https://doi.org/10.1016/j.ijrmms.2006.11.006Google Scholar
Agarwal, K., Mayerhofer, M. J., & Warpinski, N. R. (2012). Impact of Geomechanics on Microseismicity. SPE/EAGE European Unconventional Resources Conference and Exhibition, (March), 2022. https://doi.org/10.2118/152835-MSGoogle Scholar
Aguilera, R. (2016). Shale gas reservoirs: Theoretical, practical and research issues. Petroleum Research, 1(1), 1026. https://doi.org/10.1016/S2096-2495(17)30027–3Google Scholar
Ahmadov, R. (2011). Microtextural, elastic and transport properties of source rocks. Stanford University.Google Scholar
Ahmed, U. & Meehan, D. N. (eds.). (2016). Unconventional Oil and Gas Resources: Exploitation and Development. Boca Raton, FL: CRC Press.Google Scholar
Ajani, A. A. & Kelkar, M. G. (2012). Interference Study in Shale Plays. SPE Hydraulic Fracturing Technology Conference, (February), 68. https://doi.org/10.2118/151045-MSGoogle Scholar
Ajisafe, F., Solovyeva, I., Morales, A., Ejofodomi, E., & Marongiu-Porcu, M. (2017). Impact of Well Spacing and Interference on Production Performance in Unconventional Reservoirs, Permian Basin. Unconventional Resources Technology Conference, 2426. https://doi.org/10.15530/urtec-2017–2690466CrossRefGoogle Scholar
Aki, K., Fehler, M., & Das, S. (1977). Source mechanism of volcanic tremor: fluid-driven crack models and their application to the 1963 kilauea eruption. Journal of Volcanology and Geothermal Research, 2(3), 259287. https://doi.org/10.1016/0377–0273(77)90003–8Google Scholar
Aki, K. & Richards, P. G. (2002). Quantitative Seismology (2nd edn.). Sausalito, CA: University Science Books.Google Scholar
Akono, A. T. & Kabir, P. (2016). Microscopic fracture characterization of gas shale via scratch testing. Mechanics Research Communications, 78, 8692. https://doi.org/10.1016/j.mechrescom.2015.12.003Google Scholar
Al Alalli, A. (2018). Multiscale investigation of porosity and permeability relationship of shales with application towards hydraulic fracture fluid chemistry on shale matrix permeability. Stanford University.Google Scholar
Alalli, A. & Zoback, M. D. (2018). Microseismic evidence for horizontal hydraulic fractures in the Marcellus Shale, southeastern West Virginia. Leading Edge, 37(5). https://doi.org/10.1190/tle37050356.1Google Scholar
Aljamaan, H. (2015). Multi-Component Physical Sorption Investigation of Gas Shales at the Core Level. SPE Annual Technical Conference and Exhibition. https://doi.org/10.2118/178736-STUGoogle Scholar
Aljamaan, H., Ross, C. M., & Kovscek, A. R. (2017). Multiscale Imaging of Gas Adsorption in Shales. In SPE Canada Unconventional Resources Conference. https://doi.org/SPE-185054-MSGoogle Scholar
Allmann, B. P. & Shearer, P. M. (2009). Global variations of stress drop for moderate to large earthquakes. Journal of Geophysical Research: Solid Earth, 114(1), 122. https://doi.org/10.1029/2008JB005821Google Scholar
Alnoaimi, K. R., Duchateau, C., & Kovscek, A. R. (2015). Characterization and measurement of multiscale gas transport in shale core samples. SPE Journal, 116. https://doi.org/10.15530/urtec-2014–1920820Google Scholar
Alt, R. C. & Zoback, M. D. (2017). In situ stress and active faulting in Oklahoma. Bulletin of the Seismological Society of America, 107(1), 113. https://doi.org/10.1785/0120160156CrossRefGoogle Scholar
Alvarez, R. A., Alvarez, R. A., Zavala-Araiza, D., Lyon, D. R., Allen, D. T., Barkley, Z. R., … Pacala, S. W. (2018). Assessment of methane emissions from the U.S. oil and gas supply chain, Science, 361(6398), 186188. https://doi.org/10.1126/science.aar7204Google ScholarPubMed
Alzate, J. H. (2012). Integration of surface seismic, micro- seismic, and production logs for shale gas characterization: Methodology and field application. The University of Oklahoma.Google Scholar
Amer, A., Primio, R., & Ondrak, R. (2013). The impact of heat flow variations in shale gas evaluation: A Haynesville Shale case study. Society of Petroleum Engineers. https://doi.org/10.2118/164347-MSCrossRefGoogle Scholar
An, C., Guo, X., & Killough, J. (2018). Impacts of Kerogen and Clay on Stress-Dependent Permeability Measurements of Shale Reservoirs. SPE/AAPG/SEG Unconventional Resources Technology Conference. https://doi.org/10.15530/urtec-2018–2902756Google Scholar
Anderson, E. M. (1951). The Dynamics of Faulting and Dyke Formation with Applications to Britain. Edinburgh: Oliver and Boyd.Google Scholar
Angelier, J. (1990). Inversion of field data in fault tectonics to obtain the regional stress – III. A new rapid direct inversion method by analytical means. Geophysical Journal International, 103, 363376.CrossRefGoogle Scholar
Anovitz, L. M. & Cole, D. R. (2015). Characterization and Analysis of Porosity and Pore Structures. Reviews in Mineralogy and Geochemistry, 80, 61164.Google Scholar
Anovitz, L. M., Cole, D. R., Sheets, J. M., Swift, A., Elston, H. W., Welch, S., … Wasbrough, M. J. (2015). Effects of maturation on multiscale (nanometer to millimeter) porosity in the Eagle Ford Shale. Interpretation, 3(3), SU59SU70. https://doi.org/10.1190/INT-2014–0280.1Google Scholar
Aplin, A. C. & Macquaker, J. H. S. (2011). Mudstone diversity: Origin and implications for source, seal, and reservoir properties in petroleum systems. AAPG Bulletin, 95(12), 20312059. https://doi.org/10.1306/03281110162CrossRefGoogle Scholar
Araujo, O., Lopez-Bonetti, E., Garza, D., & Salinas, G. (2014). Successful Extended Injection Test for Obtaining Reservoir Data in a Gas-Oil Shale Formation in Mexico, SPE 169345-MS, Society of Petroleum Engineers. doi:10.2118/169345-MSCrossRefGoogle Scholar
Arch, J. & Maltman, A. (1990). Anisotropic permeability and tormosity in deformed wet sediments structure in accrefionary tectonic features fluids flow from accrefionary within D6collement Project (DSDP)/ ODP drill holes which have penetrated major so the examination of microscopic s. Journal of Geophysical Research, 95(89), 90359045. https://doi.org/10.1029/JB095iB06p09035Google Scholar
Arnold, R. & Townend, J. (2007). A Bayesian approach to estimating tectonic stress from seismological data. Geophysical Journal International, 170(3), 13361356. https://doi.org/10.1111/j.1365-246X.2007.03485.xGoogle Scholar
Armitage, P. J., Faulkner, D. R., Worden, R. H., Aplin, A. C., Butcher, A. R., & Iliffe, J. (2011). Experimental measurement of, and controls on, permeability and permeability anisotropy of caprocks from the CO2storage project at the Krechba Field, Algeria. Journal of Geophysical Research: Solid Earth, 116(12). https://doi.org/10.1029/2011JB008385CrossRefGoogle Scholar
Ashby, M. F. & Sammis, C. G. (1990). The damage mechanics of brittle solids in compression. Pure and Applied Geophysics, 133(3), 489521. https://doi.org/10.1007/BF00878002Google Scholar
Atkinson, G. M., Eaton, D. W., Ghofrani, H., Walker, D., Cheadle, B., Schultz, R., … Kao, H. (2016). Hydraulic fracturing and seismicity in the Western Canada sedimentary basin. Seismological Research Letters, 87(3), 631647. https://doi.org/10.1785/0220150263Google Scholar
Aybar, U., Yu, W., Eshkalak, M. O., Sepehrnoori, K., & Patzek, T. (2015). Evaluation of production losses from unconventional shale reservoirs. Journal of Natural Gas Science and Engineering, 23, 509516. https://doi.org/10.1016/j.jngse.2015.02.030Google Scholar
Backeberg, N. R., Iacoviello, F., Rittner, M., Mitchell, T. M., Jones, A. P., Day, R., … Striolo, A. (2017). Quantifying the anisotropy and tortuosity of permeable pathways in clay-rich mudstones using models based on X-ray tomography. Scientific Reports, 7(1), 112. https://doi.org/10.1038/s41598-017–14810-1Google Scholar
Baihly, J. D., Altman, R. M., Malpani, R., & Luo, F. (2010). Shale Gas Production Decline Trend Comparison over Time and Basins. SPE Annual Technical Conference and Exhibition. https://doi.org/10.2118/135555-MSCrossRefGoogle Scholar
Bakulin, A., Grechka, V., & Tsvankin, I. (2000). Estimation of fracture parameters from reflection seismic data—Part I: HTI model due to a single fracture set. Geophysics, 65(6), 1788. https://doi.org/10.1190/1.1444863CrossRefGoogle Scholar
Bandyopadhyay, K. (2009). Seismic anisotropy: Geological causes and its implications to reservoir geophysics. Stanford University.Google Scholar
Bao, X. & Eaton, D. W. (2016). Fault activation by hydraulic fracturing in western Canada. Science, 354(6318), 14061409. https://doi.org/10.1126/science.aag2583Google Scholar
Barbour, A. J., Norbeck, J. H., & Rubinstein, J. L. (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. https://doi.org/10.1785/0220170003Google Scholar
Bardainne, T. (2011). Semi-Automatic Migration Location of Microseismic Events: A Good Trade-off between Efficiency and Reliability. In Third Passive Seismic Workshop – Actively Passive! (p. PSP09).Google Scholar
Barkley, Z. R., Lauvaux, T., Davis, K. J., Deng, A., Cao, Y., Sweeney, C., … Maasakkers, J. D. (2017). Quantifying methane emissions from natural gas production in northeastern Pennsylvania. Atmospheric Chemistry and Physics Discussions, 153. https://doi.org/10.5194/acp-2017–200Google Scholar
Baron, L. I., Longuntsov, B. M., & Pozin, E. Z. (1962). Determination of Properties of Rocks. Moscow: Gosgortekhizdat.Google Scholar
Barree, R. D., Barree, V. L., & Craig, D. P. (2007). Holistic Fracture Diagnostics. SPE Rocky Oil & Gas Technology Symposium. https://doi.org/10.2118/107877-PAGoogle Scholar
Barton, C. A., Zoback, M. D., & Moos, D. (1995). Fluid flow along potentially active faults in crystalline rock. Geology, 23, 683686.Google Scholar
Barton, C., Moos, D., & Tezuka, K. (2009). Geomechanical wellbore imaging: Implications for reservoir fracture permeability. AAPG Bulletin, 93(11), 15511569. https://doi.org/10.1306/06180909030Google Scholar
Belyadi, H., Yuyi, J., Ahmad, M., Wyatt, J., & Energy, C. (2016). Deep Dry Utica Well Spacing Analysis with Case Study Dry Utica Rock Properties / Geological Overview. SPE Eastern Regional Meeting, (September), 1315.Google Scholar
Bertier, P., Schweinar, K., Stanjek, H., Ghanizadeh, A., Clarkson, C. R., Busch, A., … Pipich, V. (2016). On the use and abuse of N2 physisorption for the characterization of the pore structure of shales. In The Clay Minerals Society Workshop Lectures Series (Vol. 21, pp. 151161). https://doi.org/10.1346/CMS-WLS-21.12Google Scholar
Bhandari, A. R., Flemings, P. B., Polito, P. J., Cronin, M. B., & Bryant, S. L. (2015). Anisotropy and stress dependence of permeability in the Barnett Shale. Transport in Porous Media, 108(2), 393411. https://doi.org/10.1007/s11242-015–0482-0Google Scholar
Biot, M. A. (1941). General theory of three-dimensional consolidation. Journal of Applied Physics, 12(2), 155164.Google Scholar
Bishop, A. (1967). Progressive Failure with Special Reference to the Mechanism Causing It. In Proc. Geotech. Conf. (pp. 142150). Oslo.Google Scholar
Bjørlykke, K. (1998). Clay mineral diagenesis in sedimentary basins – a key to the prediction of rock properties. Examples from the North Sea Basin. Clay Minerals, 33(1), 1534. https://doi.org/10.1180/000985598545390Google Scholar
Blackwell, D. D., Richards, M. C., Frone, Z. S., Batir, J. F., Williams, M. A., Ruzo, A. A., & Dingwall, R. K. (2011). SMU Geothermal Laboratory Heat Flow Map of the Conterminous United States, 2011.Google Scholar
Boness, N. L. & Zoback, M. D. (2006). A multiscale study of the mechanisms controlling shear velocity anisotropy in the San Andreas Fault Observatory at Depth. Geophysics, 71(5). https://doi.org/10.1190/1.2231107Google Scholar
Bonnelye, A., Schubnel, A., David, C., Henry, P., Guglielmi, Y., Gout, C., … Dick, P. (2016a). Elastic wave velocity evolution of shales deformed under uppermost-crustal conditions. Journal of Geophysical Research: Solid Earth, 122(1), 145. https://doi.org/10.1002/2016JB013540Google Scholar
Bonnelye, A., Schubnel, A., David, C., Henry, P., Guglielmi, Y., Gout, C., … Dick, P. (2016b). Strength anisotropy of shales deformed under uppermost crustal conditions. Journal of Geophysical Research: Solid Earth, 122(1), 110129. https://doi.org/10.1002/2016JB013040Google Scholar
Bott, M. H. P. (1959). The mechanics of oblique-slip faulting. Geological Magazine, 96, 109117.Google Scholar
Bowker, K. (2007). Barnett Shale gas production, Fort Worth Basin: Issues and discussion, AAPG Bulletin, 91(4), 523533.CrossRefGoogle Scholar
Brace, W. F., Walsh, J. B., & Frangos, W. T. (1968). Permeability of granite under high pressure. Journal of Geophysical Research, 73(6), 22252236. https://doi.org/10.1029/JB073i006p02225Google Scholar
Brandt, A. R., Heath, G. A., & Cooley, D. (2016). Methane leaks from natural gas systems follow extreme distributions. Environmental Science and Technology, 50(22), 1251212520. https://doi.org/10.1021/acs.est.6b04303CrossRefGoogle ScholarPubMed
Brannon, H. D. & Bell, C. E. (2011). Eliminating Slickwater Compromises for Improved Shale Stimulation. In SPE Annual Technical Conference and Exhibition (Vol. 6, pp. 4578–4588). Retrieved from www.scopus.com/inward/record.url?eid=2-s2.0–84856719865&partnerID=40&md5=6fb0c8cbf9a6b1b0869baf1782970a12CrossRefGoogle Scholar
Bratovich, M. W. & Walles, F. (2016). Formation evaluation and reservoir characterization of source rock reservoirs. In Ahmed, U. & Meehan, D. N. (eds.), Unconventional Oil and Gas Resources: Exploitation and Development. Boca Raton, FL: CRC Press.Google Scholar
Britt, L. K., Rock, B., Smith, M. B., & Klein, H. H. (2016). Production Benefits from Complexity – Effects of Rock Fabric, Managed Drawdown, and Propped Fracture Conductivity.Google Scholar
Burton, W. A. (2016). Multistage completion systems for unconventionals. In Ahmed, U. & Meehan, D. N. (eds.), Unconventional Oil and Gas Resources: Exploitation and Development. Boca Raton, FL: CRC Press.Google Scholar
Busch, A., Schweinar, K., Kampman, N., Coorn, A., Pipich, V., Feoktystov, A., … Bertier, P. (2017). Determining the porosity of mudrocks using methodological pluralism. Geological Society, London, Special Publications, 454(1), 1538. https://doi.org/10.1144/SP454.1Google Scholar
Byerlee, J. (1978). Friction of rocks. Pure and Applied Geophysics, 116(4–5), 615626. https://doi.org/10.1007/BF00876528Google Scholar
Caffagni, E., Eaton, D., van der Baan, M., & Jones, J. P. (2015). Regional seismicity: A potential pitfall for identification of long-period long-duration events. Geophysics, 80(1), A1A5. https://doi.org/10.1190/geo2014-0382.1Google Scholar
Caine, J. S., Evans, J. P., & Forster, C. B. (1996). Fault zone architecture and permeability structure, Geology, 24(11), 10251028.Google Scholar
Candela, T., Renard, F., Klinger, Y., Mair, K., Schmittbuhl, J., & Brodsky, E. E. (2012). Roughness of fault surfaces over nine decades of length scales. Journal of Geophysical Research: Solid Earth, 117(8), 130. https://doi.org/10.1029/2011JB009041Google Scholar
Cander, H. (2012). Sweet spots in shale gas and liquid plays: prediction of fluid composition and reservoir pressures. Search and Discovery, 40936, 29pp.Google Scholar
Cappa, F. & Rutqvist, J. (2011). Modeling of coupled deformation and permeability evolution during fault reactivation induced by deep underground injection of CO2. International Journal of Greenhouse Gas Control, 5(2), 336346. https://doi.org/10.1016/j.ijggc.2010.08.005Google Scholar
Carey, J. W. (2013). Geochemistry of wellbore integrity in CO2 sequestration: Portland cement-steel-brine-CO2 interactions. Reviews in Mineralogy and Geochemistry, 77(1), 505539. https://doi.org/10.2138/rmg.2013.77.15Google Scholar
Chae, K. S. & Lee, J. W. (2015). Risk Analysis and Simulation for Geologic Storage of CO2. In Proceedings of the World Congress on Advances in Civil, Environmental, and Materials Research, Incheon, Korea, 25–29 August 2015.Google Scholar
Chalmers, G. R., Bustin, R. M., & Power, I. M. (2012a). Characterization of gas shale pore systems by porosimetry, pycnometry, surface area, and field emission scanning electron microscopy/transmission electron microscopy image analyses: Examples from the Barnett, Woodford, Haynesville, Marcellus, and Doig uni. AAPG Bulletin, 96(6), 10991119. https://doi.org/10.1306/10171111052Google Scholar
Chalmers, G. R. L., Ross, D. J. K., & Bustin, R. M. (2012b). Geological controls on matrix permeability of Devonian Gas Shales in the Horn River and Liard basins, northeastern British Columbia, Canada. International Journal of Coal Geology, 103, 120131https://doi.org/10.1016/j.coal.2012.05.006Google Scholar
Chambers, K., Kendall, J. M., Brandsberg-Dahl, S., & Rueda, J. (2010). Testing the ability of surface arrays to monitor microseismic activity. Geophysical Prospecting, 58(5), 821830. https://doi.org/10.1111/j.1365–2478.2010.00893.xCrossRefGoogle Scholar
Chan, A. & Zoback, M. D. (2002). Deformation Analysis in Reservoir Space (DARS): A Simple Formalism for Prediction of Reservoir Deformation with Depletion – SPE 78174. In SPE/ISRM Rock Mechanics Conference. Irving, TX: Society of Petroleum Engineers.Google Scholar
Chandler, M. R., Meredith, P. G., Brantut, N., & Crawford, B. R. (2016). Fracture toughness anisotropy in shale. Journal of Geophysical Research: Solid Earth, 121(3), 17061729. https://doi.org/10.1002/2015JB012756Google Scholar
Chang, C., Zoback, M. D., & Khaksar, A. (2006). Empirical relations between rock strength and physical properties in sedimentary rocks. Journal of Petroleum Science and Engineering, 51(3–4), 223237. https://doi.org/10.1016/j.petrol.2006.01.003Google Scholar
Chang, C. & Zoback, M. D. (2009). Viscous creep in room-dried unconsolidated Gulf of Mexico shale (I): Experimental results. Journal of Petroleum Science and Engineering, 69(34). https://doi.org/10.1016/j.petrol.2009.08.018Google Scholar
Chang, K. W. & Segall, P. (2016). Injection-induced seismicity on basement faults including poroelastic stressing. Journal of Geophysical Research: Solid Earth, 121(4), 27082726. https://doi.org/10.1002/2015JB012561Google Scholar
Charlez, P. & Baylocq, P. (2015). The Shale Oil and Gas Debate. Paris, France: Technip.Google Scholar
Chen, Z., Zhou, L., Walsh, R., & Zoback, M. (2018). Case study Fault slip and Casing Deformation Induced by Hydraulic Fracturing in Sichuan Basin – URTeC-2882313 20180501. In URTeC 2882313.Google Scholar
Cheng, Y. (2012). Impact of Water Dynamics in Fractures on the Performance of Hydraulically Fractured Wells in Gas Shale Reservoirs. SPE International Symposium and Exhibition on Formation Damage Control, (May 2011), 10–12. https://doi.org/10.2118/127863-MSGoogle Scholar
Chester, F. M., Evans, J. P., & Biegel, R. L. (1993). Internal structure and weakening mechanisms of the San Andreas fault. Journal of Geophysical Research, 98, 771786.Google Scholar
Chester, F. M. & Logan, J. M. (1986). Implications for mechanical properties of brittle faults from observations of the Punchbowl fault zone, California. Pure and Applied Geophysics, 124, 79106.Google Scholar
Chiaramonte, L., Zoback, M. D., Friedmann, J., & Stamp, V. (2008). Seal integrity and feasibility of CO2 sequestration in the Teapot Dome EOR pilot: Geomechanical site characterization. Environmental Geology, 54(8). https://doi.org/10.1007/s00254-007–0948-7Google Scholar
Cho, Y., Ozkan, E., & Apaydin, O. G. (2013). Pressure-dependent natural-fracture permeability in shale and its effect on shale-gas well production. SPE Reservoir Evaluation & Engineering, 16(02), 216228. https://doi.org/10.2118/159801-PACrossRefGoogle Scholar
Chong, K. K., Grieser, W. V., Passman, A., Tamayo, C. H., Modeland, N., & Burke, B. (2010). A Completions Roadmap to Shale-Play Development: A Review of Successful Approaches toward Shale-Play Stimulation in the Last Two Decades. International Oil and Gas Conference and Exhibition in China, 1–27. https://doi.org/10.2118/130369-MSGoogle Scholar
Ciezobka, J., Courtier, J., & Wicker, J. (2018). Results, 1–9. https://doi.org/10.15530/urtec-2018–2902355Google Scholar
Cipolla, C. & Wallace, J. (2014). Stimulated Reservoir Volume: A Misapplied Concept? SPE Hydraulic Fracturing Technology Conference, (February), 4–6. https://doi.org/10.2118/168596-MSGoogle Scholar
Cipolla, C., Motiee, M., Kechemir, A., & Corporation, H. (2018a). Integrating Microseismic, Geomechanics, Hydraulic Fracture Modeling, and Reservoir Simulation to Characterize Parent Well Depletion and Infill Well Performance in the Bakken, 1–12. https://doi.org/10.15530/urtec-2018–2899721Google Scholar
Cipolla, C., Gilbert, C., Sharma, A., & Lebas, J. (2018b). Case History of Completion Optimization in the Utica. SPE Hydraulic Fracturing Technology Conference and Exhibition, (January), 2325.Google Scholar
Cipolla, C. L., Warpinski, N. R., Mayerhofer, M. J., Lolon, E. P., & Vincent, M. C. (2008). The Relationship between Fracture Complexity, Reservoir Properties, and Fracture Treatment Design. In 2008 SPE Annual Technical Conference and Exhibition, Denver, CO. 11–24 September 2008 (Vol. 1). Society of Petroleum Engineers. Retrieved from http://www.scopus.com/inward/record.url?eid=2-s2.0–79952850661&partnerID=40&md5=0139cfc6cf0b922340aca989490b6fe8Google Scholar
Clarke, H., Eisner, L., Styles, P., & Turner, P. (2014). Felt seismicity associated with shale gas hydraulic fracturing: The first documented example in Europe. Geophysical Research Letters, 41(23), 83088314. https://doi.org/10.1002/2014GL062047Google Scholar
Clarkson, C. R., Solano, N., Bustin, R. M., Bustin, A. M. M., Chalmers, G. R. L., He, L., … Blach, T. P. (2013). Pore structure characterization of North American shale gas reservoirs using USANS/SANS, gas adsorption, and mercury intrusion. Fuel, 103, 606616. https://doi.org/10.1016/j.fuel.2012.06.119Google Scholar
Clerc, F., Harrington, R. M., Liu, Y., & Gu, Y. J. (2016). Stress drop estimates and hypocenter relocations of induced seismicity near Crooked Lake, Alberta. Geophysical Research Letters, 43(13), 69426951. https://doi.org/10.1002/2016GL069800Google Scholar
Coates, D. F. & Parsons, R. C. (1966). Experimental criteria for classification of rock substances. International Journal of Rock Mechanics and Mining Sciences, 3(3), 181189. https://doi.org/10.1016/0148–9062(66)90022–2Google Scholar
Cocco, M., Tinti, E., & Cirella, A. (2016). On the scale dependence of earthquake stress drop. Journal of Seismology. https://doi.org/10.1007/s10950-016–9594-4Google Scholar
Council, N. R. (2013). Induced Seismicity Potential in Energy Technologies. Washington D.C.: National Academies Press.Google Scholar
Craig, D. P., Barree, R. D., Warpinski, N. R., & Blasingame, T. A. (2017). Fracture Closure Stress: Reexamining Field and Laboratory Experiments of Fracture Closure Using Modern Interpretation Methodologies, SPE187038-MS. In SPE Annual Technical Conference and Exhibition, 9–11 October, San Antonio, Texas, USA. Society of Petroleum Engineers. https://doi.org/10.2118/187038-MSGoogle Scholar
Crain, E. R. (2010). Crain’s Data Acquisition.Google Scholar
Cramer, D. D. & Nguyen, D. H. (2013).Diagnostic Fracture Injection Testing Tactics in Unconventional Reservoirs, SPE 163863. Society of Petroleum Engineers. doi:10.2118/163863-MSGoogle Scholar
Crandall, D., Moore, J., Gill, M., & Stadelman, M. (2017). CT scanning and flow measurements of shale fractures after multiple shearing events. International Journal of Rock Mechanics and Mining Sciences, 100 (November 2016), 177187. https://doi.org/10.1016/j.ijrmms.2017.10.016Google Scholar
Crawford, B. R., Faulkner, D. R., & Rutter, E. H. (2008). Strength, porosity, and permeability development during hydrostatic and shear loading of synthetic quartz-clay fault gouge. Journal of Geophysical Research: Solid Earth, 113(3), 114. https://doi.org/10.1029/2006JB004634Google Scholar
Crone, A. J. & Luza, K. V. (1990). Style and timing of Holocene surface faulting on the Meers fault, southwestern Oklahoma. Geological Society of America Bulletin, 102, 117.2.3.CO;2>CrossRefGoogle Scholar
Cui, X., Bustin, A. M. M., & Bustin, R. M. (2009). Measurements of gas permeability and diffusivity of tight reservoir rocks: Different approaches and their applications. Geofluids, 9(3), 208223. https://doi.org/10.1111/j.1468–8123.2009.00244.xGoogle Scholar
Curtis, C. D. (1980). Diagenetic alteration in black shales. Journal of the Geological Society, 137(2), 189194. https://doi.org/10.1144/gsjgs.137.2.0189Google Scholar
Curtis, J. B. (2002). Fractured shale-gas systems. AAPG Bulletin, 86(11), 19211938. https://doi.org/10.1306/61EEDDBE-173E-11D7-8645000102C1865DGoogle Scholar
Curtis, M. E., Ambrose, R. J., Sondergeld, C. H., & Rai, C. S. (2011). Investigation of the Relationship between Organic Porosity and Thermal Maturity in the Marcellus Shale. SPE Conference, SPE 114370. https://doi.org/10.2118/144370-msGoogle Scholar
Curtis, M. E., Cardott, B. J., Sondergeld, C. H., & Rai, C. S. (2012a). Development of organic porosity in the Woodford Shale with increasing thermal maturity. International Journal of Coal Geology, 103, 2631. https://doi.org/10.1016/j.coal.2012.08.004Google Scholar
Curtis, M. E., Sondergeld, C. H., Ambrose, R. J., & Rai, C. S. (2012b). Microstructural investigation of gas shales in two and three dimensions using nanometer-scale resolution imaging. AAPG Bulletin, 96(4), 665677. https://doi.org/10.1306/08151110188Google Scholar
Daigle, H. & Dugan, B. (2011). Permeability anisotropy and fabric development: A mechanistic explanation. Water Resources Research, 47(12), 111. https://doi.org/10.1029/2011WR011110Google Scholar
Daniels, J., Waters, G., Le Calvez, J., Bentley, D., & Lassek, J. (2007). Contacting More of the Barnett Shale Through an Integration of Real-Time Microseismic Monitoring, Petrophysics, and Hydraulic Fracture Design. Proceedings of SPE Annual Technical Conference and Exhibition. https://doi.org/10.2118/110562-MSGoogle Scholar
Darold, A. P. & Holland, A. A. (2015). Preliminary Oklahoma Optimal Fault Orientations. Oklahoma Open File Report oF 4–2015.Google Scholar
Das, I. & Zoback, M. D. (2013a). Long-period, long-duration seismic events during hydraulic stimulation of shale and tight-gas reservoirs – Part 1: Waveform characteristics. Geophysics, 78(6). https://doi.org/10.1190/GEO2013-0164.1Google Scholar
Das, I. & Zoback, M. D. (2013b). Long-period long-duration seismic events during hydraulic stimulation of shale and tight-gas reservoirs – Part 2: Location and mechanisms. Geophysics, 78(6). https://doi.org/10.1190/GEO2013-0165.1Google Scholar
Day, S., Sakurovs, R., & Weir, S. (2008). Supercritical gas sorption on moist coals. International Journal of Coal Geology, 74(3–4), 203214. https://doi.org/10.1016/j.coal.2008.01.003CrossRefGoogle Scholar
Detournay, E. (2004). Propagation regimes of fluid-driven fractures in impermeable rocks. International Journal of Geomechanics, 4(1), 3545. https://doi.org/10.1061/(ASCE)1532–3641(2004)4:1(35)Google Scholar
Detournay, E. (2016). Mechanics of hydraulic fractures. Annual Review of Fluid Mechanics, 48(1), 311339. https://doi.org/10.1146/annurev-fluid-010814–014736Google Scholar
Deutch, J., Holditch, S., Krupp, F., McGinty, K., Tierney, S., Yergin, D., & Zoback, M. D. (2011). Shale Gas Production Subcommittee 90-Day Report.Google Scholar
Dieterich, J. H. (1978). Time-dependent friction and the mechanics of stick-slip. In Rock Friction and Earthquake Prediction (pp. 790–806). Basel: Birkhäuser Basel. https://doi.org/10.1007/978–3-0348–7182-2_15Google Scholar
Dieterich, J. H. (1979). Modeling of rock friction 1. Experimental results and constitutive equations. Journal of Geophysical Research, 84, 21612168.Google Scholar
Dieterich, J. H. & Kilgore, B. D. (1994). Direct observation of frictional contacts: New insights for state-dependent properties. Pure and Applied Geophysics, 143(1–3), 283302. https://doi.org/10.1007/BF00874332Google Scholar
Dohmen, T., Zhang, J., Barker, L., & Blangy, J. P. (2017). Microseismic magnitudes and b-values for delineating hydraulic fracturing and depletion. SPE Journal, 22(5), 111. https://doi.org/10.2118/186096-PAGoogle Scholar
Dow, W. G. (1977). Kerogen studies and geological interpretations. Journal of Geochemical Exploration, 7(C), 7999. https://doi.org/10.1016/0375-6742(77)90078-4Google Scholar
Dunham, E. M., Belanger, D., Cong, L., & Kozdon, J. E. (2011). Earthquake ruptures with strongly rate-weakening friction and off-fault plasticity, part 2: Nonplanar faults. Bulletin of the Seismological Society of America, 101(5), 23082322. https://doi.org/10.1785/0120100076Google Scholar
Eaton, D. W. (2018). Passive Seismic Monitoring of Induced Seismicity. Cambridge University Press.Google Scholar
Eaton, D. W. & Forouhideh, F. (2011). Solid angles and the impact of receiver-array geometry on microseismic moment-tensor inversion. Geophysics, 76(6), WC77WC85. https://doi.org/10.1190/geo2011-0077.1CrossRefGoogle Scholar
Eaton, D. W. & Schultz, R. (2018). Increased likelihood of induced seismicity in highly overpressured shale formations, Geophysical Journal International, 214, 751757.Google Scholar
Eaton, D. W., Igonin, N., Poulin, A., Weir, R., Zhang, H., Pellegrino, S., & Rodriquez, G. (2018). Tony Creek Dual Microseismic Experiment (ToC2ME). In GeoConvention 2018. Calgary, Canada.Google Scholar
Economides, M. J. & Nolte, K. (2000). Reservoir Stimulation. Wiley.Google Scholar
Edwards, R. W. J. & Celia, M. A. (2018). Shale gas well, hydraulic fracturing, and formation data to support modeling of gas and water flow in shale formations. Water Resources Research, 111. https://doi.org/10.1002/2017WR022130Google Scholar
EIA (2013). Technically Recoverable Shale Oil and Shale Gas Resources: An Assessment of 137 Shale Formations in 41 Countries Outside the Unites States (June).Google Scholar
Eisner, L., Fischer, T., & Rutledge, J. T. (2009). Determination of S-wave slowness from a linear array of borehole receivers. Geophysical Journal International, 176(1), 3139.Google Scholar
Eisner, L., Gei, D., Hallo, M., Opršal, I., & Ali, M. Y. (2013). The peak frequency of direct waves for microseismic events. Geophysics, 78(6), A45A49. https://doi.org/10.1190/geo2013-0197.1Google Scholar
Eisner, L., Hulsey, B. J., Duncan, P. M., Jurick, D., Werner, H., & Keller, W. (2010). Comparison of surface and borehole locations of induced seismicity. Geophysical Prospecting, 58(5), 809820. https://doi.org/10.1111/j.1365–2478.2010.00867.xGoogle Scholar
Ellsworth, W. L. (2013). Injection-induced earthquakes. Science, 341 (1225942 12 July 2013). https://doi.org/10.1126/science.1225942Google Scholar
Engelder, T. & Fischer, M. P. (1994). Influence of poroelastic behavior on the magnitude of minimum horizontal stress, Sh, in overpressured parts of sedimentary basins. Geology, 22(10), 949952. https://doi.org/10.1130/0091–7613(1994)022<0949:IOPBOT>2.3.COGoogle Scholar
Engelder, T., Lash, G. G., & Uzcátegui, R. S. (2009). Joint sets that enhance production from Middle and Upper Devonian gas shales of the Appalachian Basin. AAPG Bulletin, 93(7), 857889. https://doi.org/10.1306/03230908032Google Scholar
Engle, M. A., Reyes, F. R., Varonka, M. S., Orem, W. H., Ma, L., Ianno, A. J., … Carroll, K. C. (2016). Geochemistry of formation waters from the Wolfcamp and “Cline” shales: Insights into brine origin, reservoir connectivity, and fluid flow in the Permian Basin, USA. Chemical Geology, 425, 7692. https://doi.org/10.1016/j.chemgeo.2016.01.025Google Scholar
English, J. M., English, K. L., Corcoran, D. V. & Toussaint, F. (2016) Exhumation charge: The last gasp of a petroleum source rock and implications for unconventional shah resources. AAPG Bulletin, 100 (1), pp. 116. doi: 10.1306/07271514224.Google Scholar
Environmental Protection Agency (2015). Assessment of the Potential Impacts of Hydraulic Fracturing for Oil and Gas on Drinking Water Resources: Executive Summary, (June), 1–3. Retrieved from http://www2.epa.gov/sites/production/files/2015–07/documents/hf_es_erd_jun2015.pdfGoogle Scholar
EPA (2014). Minimizing and Managing Potential Impacts of Injection-Induced Seismicity from Class II Disposal Wells: Practical Approaches.Google Scholar
EPA (2016). Hydraulic Fracturing for Oil and Gas: Impacts from the Hydraulic Fracturing Water Cycle on Drinking Water Resources in the United States. Retrieved from www.epa.gov/hfstudyGoogle Scholar
Esaki, T., Du, S., Mitani, Y., Ikusada, K., & Jing, L. (1999). Development of a shear-flow test apparatus and determination of coupled properties for a single rock joint. International Journal of Rock Mechanics and Mining Sciences, 36(5), 641650. https://doi.org/10.1016/S0148-9062(99)00044–3Google Scholar
Eyre, T. S. & van der Baan, M. (2017). The reliability of microseismic moment tensor solutions: Surface versus borehole monitoring. Geophysics, 82(6), 146. https://doi.org/10.1190/geo2017-0056.1Google Scholar
Fan, L., Harris, B., & Jamaluddin, A. (2005). Understanding gas-condensate reservoirs. Oilfield Review, (Winter 2005/2006), 14–27. https://doi.org/http://dx.doi.org/10.4043/25710-MSGoogle Scholar
Fan, Z., Eichhubl, P., & Gale, J. F. W. (2016). Geomechanical analysis of fluid injection and seismic fault slip for the Mw4.8 Timpson, Texas, earthquake sequence. Journal of Geophysical Research: Solid Earth, 121, 27982812. https://doi.org/10.1002/2016JB012821.ReceivedGoogle Scholar
Fang, Y., Elsworth, D., Wang, C., Ishibashi, T., & Fitts, J. P. (2017). Frictional stability-permeability relationships for fractures in shales. Journal of Geophysical Research: Solid Earth, 122(3), 17601776. https://doi.org/10.1002/2016JB013435CrossRefGoogle Scholar
Farghal, N. (2018). Fault and fracture identification and characterization in 3D seismic data from unconventional reservoirs, PhD Thesis, Stanford University.Google Scholar
Farghal, N. S. & Zoback, M. D. (2015). Identification of slowly slipping faults in the Barnett Shale utilizing ant tracking; Identification of slowly slipping faults in the Barnett Shale utilizing ant tracking. In SEG Technical Program Expanded Abstracts (Vol. 34), 49194923. https://doi.org/10.1190/segam2015-5811224.1Google Scholar
Faulkner, D. R. & Rutter, E. H. (1998). The gas permeability of clay-bearing fault gouge at 20 C. Geological Society, London, Special Publications, 147(1), 147156. https://doi.org/10.1144/GSL.SP.1998.147.01.10CrossRefGoogle Scholar
Faulkner, D. R., Lewis, A. C., & Rutter, E. H. (2003). On the internal structure and mechanics of large strike-slip fault zones: Field observations of the Carboneras fault in southeastern Spain. Tectonophysics, 367(3–4), 235251. https://doi.org/10.1016/S0040-1951(03)00134–3Google Scholar
Faulkner, D. R., Mitchell, T. M., Healy, D., & Heap, M. J. (2006). Slip on “weak” faults by the rotation of regional stress in the fracture damage zone. Nature, 444(7121), 922925. https://doi.org/10.1038/nature05353Google Scholar
Ferguson, W., Richards, G., Bere, A., Mutlu, U., & Paw, F. (2018). Modelling Near-Wellbore Hydraulic Fracture Branching, Complexity and Tortuosity: A Case Study Based on a Fully Coupled Geomechanical Modelling Approach. SPE Hydraulic Fracturing Technology Conference and Exhibition. https://doi.org/10.2118/189890-MSCrossRefGoogle Scholar
Fisher, M. K. & Warpinski, N. R. (2012). Hydraulic-Fracture-Height Growth: Real Data. SPE Production & Operations. https://doi.org/10.2118/145949-PAGoogle Scholar
Fisher, M. K., Heinze, J. R., Harris, C. D., Davidson, B. M., Wright, C. A., & Dunn, K. P. (2004). Optimizing Horizontal Completion Techniques in the Barnett Shale Using Microseismic Fracture Mapping. SPE Annual Technical Conference and Exhibition. https://doi.org/10.2118/90051-MSGoogle Scholar
Fisher, M. K., Wright, C. A., Davidson, B. M., Goodwin, A. K., Fielder, E. O., Buckler, W. S., & Steinsberger, N. P. (2002). Integrating Fracture Mapping Technologies to Optimize Stimulations in the Barnett Shale. In SPE Annual Technical Conference and Exhibition. Soc. Petr. Engr. https://doi.org/10.2118/77441-MSCrossRefGoogle Scholar
Fogden, A., Olson, T., Cheng, Q., Middleton, J., Kingston, A., Turner, M., … Armstrong, R. (2015). Dynamic Micro-CT Imaging of Diffusion in Unconventionals. Proceedings of the Unconventional Resources Technology Conference, 1–16. https://doi.org/10.15530/urtec-2015–2154822Google Scholar
Forand, D., Heesakkers, V., & Schwartz, K. (2017). Constraints on Natural Fracture and In-Situ Stress Trends of Unconventional Reservoirs in the Permian Basin, USA, 24–26. https://doi.org/10.15530/-urtec-2017–2669208Google Scholar
Foulger, G. R., Wilson, M. P., Gluyas, J. G., Julian, B. R., & Davies, R. J. (2018). Global review of human-induced earthquakes. Earth-Science Reviews, 178 (January 2017), 438514. https://doi.org/10.1016/j.earscirev.2017.07.008Google Scholar
Friberg, P. A., Besana-Ostman, G. M., & Dricker, I. (2014). Characterization of an earthquake sequence triggered by hydraulic fracturing in Harrison County, Ohio. Seismological Research Letters, 85(6), 12951307. https://doi.org/10.1785/0220140127Google Scholar
Friedrich, M. & Milliken, M. (2013). Determining the Contributing Reservoir Volume from Hydraulically Fractured Horizontal Wells in the Wolfcamp Formation in the Midland Basin. Unconventional Resources Technology Conference, Denver, Colorado, 12–14 August 2013, 14611468. https://doi.org/10.1190/urtec2013-149Google Scholar
Frohlich, C., Hayward, C., Stump, B., & Potter, E. (2011). The Dallas-Fort Worth earthquake sequence: October 2008 through May 2009. Bulletin of the Seismological Society of America, 101(1), 327340. https://doi.org/10.1785/0120100131Google Scholar
Frohlich, C., Ellsworth, W. L., Brown, W. A., Brunt, M., Luetgert, J., Macdonald, T., & Walter, S. (2014). The 17 May 2012M4.8 earthquake near Timpson, East Texas: An event possibly triggered by fluid injection. Journal of Geophysical Research, (August 1982), 581593. https://doi.org/10.1002/2013JB010755.ReceivedCrossRefGoogle Scholar
Gaarenstroom, L., Tromp, R. A. J., Jong, M. C. de, & Brandenburg, A. M. (1993). Overpressures in the Central North Sea: Implications for trap integrity and drilling safety. In Parker, J. R. (ed.), Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference (pp. 13051313). London.Google Scholar
Gale, J. F. W., Elliott, S. J., & Laubach, S. E. (2018). Hydraulic Fractures in Core From Stimulated Reservoirs: Core Fracture Description of HFTS Slant Core, Midland Basin, West Texas, (1993). https://doi.org/10.15530/urtec-2018–2902624Google Scholar
Gale, J. F. W., Laubach, S. E., Olson, J. E., Eichhubl, P., & Fall, A. (2014). Natural fractures in shale: A review and new observations. AAPG Bulletin, 98(11), 21652216.Google Scholar
Gale, J. F. W., Reed, R. M., & Holder, J. (2007). Natural fractures in the Barnett Shale and their importance for hydraulic fracture treatments. AAPG Bulletin, 91(4), 603622. https://doi.org/10.1306/11010606061CrossRefGoogle Scholar
Gasparik, M., Ghanizadeh, A., Bertier, P., Gensterblum, Y., Bouw, S., & Krooss, B. M. (2012). High-pressure methane sorption isotherms of black shales from the Netherlands. Energy and Fuels, 26(8), 49955004. https://doi.org/10.1021/ef300405gGoogle Scholar
Gasparik, M., Bertier, P., Gensterblum, Y., Ghanizadeh, A., Krooss, B. M., & Littke, R. (2014). Geological controls on the methane storage capacity in organic-rich shales. International Journal of Coal Geology, 123, 3451. https://doi.org/10.1016/j.coal.2013.06.010Google Scholar
Gdanski, R. D., Fulton, D. D., & Shen, C. (2009). Fracture-face-skin evolution during cleanup. SPE Production & Operations, 24(1), 2234. https://doi.org/10.2118/101083-paGoogle Scholar
Geng, Z., Bonnelye, A., Chen, M., Jin, Y., Dick, P., David, C., … Schubnel, A. (2017). Elastic anisotropy reversal during brittle creep in shale. Geophysical Research Letters, 44(21),10, 887–10, 895. https://doi.org/10.1002/2017GL074555Google Scholar
Gensterblum, Y., Merkel, A., Busch, A., & Krooss, B. M. (2013). High-pressure CH4 and CO2 sorption isotherms as a function of coal maturity and the influence of moisture. International Journal of Coal Geology, 118, 4557. https://doi.org/10.1016/j.coal.2013.07.024Google Scholar
Gensterblum, Y., Ghanizadeh, A., Cuss, R. J., Amann-Hildenbrand, A., Krooss, B. M., Clarkson, C. R., … Zoback, M. D. (2015). Gas transport and storage capacity in shale gas reservoirs – A review. Part A: Transport processes. Journal of Unconventional Oil and Gas Resources, 12, 87122. https://doi.org/10.1016/j.juogr.2015.08.001Google Scholar
Gephart, J. W. & Forsyth, D. W. (1984). An improved method for determining the regional stress tensor using earthquake focal mechanism data: application to the San Fernando earthquake sequence. Journal of Geophysical Research, 89, 93059320.Google Scholar
Ghanbarian, B. & Javadpour, F. (2017). Upscaling pore pressure-dependent gas permeability in shales. Journal of Geophysical Research: Solid Earth, 122(4), 25412552. https://doi.org/10.1002/2016JB013846Google Scholar
Ghanizadeh, A., Gasparik, M., Amann-Hildenbrand, A., Gensterblum, Y., & Krooss, B. M. (2013). Lithological controls on matrix permeability of organic-rich shales: An experimental study. Energy Procedia, 40, 127136. https://doi.org/10.1016/j.egypro.2013.08.016Google Scholar
Ghanizadeh, A., Amann-Hildenbrand, A., Gasparik, M., Gensterblum, Y., Krooss, B. M., & Littke, R. (2014b). Experimental study of fluid transport processes in the matrix system of the European organic-rich shales: II. Posidonia Shale (Lower Toarcian, northern Germany). International Journal of Coal Geology, 123, 2033. https://doi.org/10.1016/j.coal.2013.06.009Google Scholar
Gherabati, S. A., Hammes, U., Male, F., Browning, J., Smye, K., Ikonnikova, S. A., & McDaid, G. (2016). Assessment of Hydrocarbon-in-Place and Recovery Factors in the Eagle Ford Shale Play, URTeC: 2460252.Google Scholar
Godec, M., Koperna, G., Petrusak, R., & Oudinot, A. (2013). Potential for enhanced gas recovery and CO2 storage in the Marcellus Shale in the Eastern United States. International Journal of Coal Geology, 118, 95104. https://doi.org/10.1016/j.coal.2013.05.007Google Scholar
Goertz-Allmann, B. P., Goertz, A., & Wiemer, S. (2011). Stress drop variations of induced earthquakes at the Basel geothermal site. Geophysical Research Letters, 38(9), 15. https://doi.org/10.1029/2011GL047498Google Scholar
Goodway, B., Chen, T., & Downton, J. (1997). Improved AVO Fluid Detection and Lithology Discrimination Using Lame Petrophysical Parameters from P and S Inversions. SEG Annual Meeting. https://doi.org/10.1190/1.1885795Google Scholar
Goodway, B., Varsek, J., & Abaco, C. (2006). Practical applications of P-wave AVO for unconventional gas Resource Plays – Part 1: Seismic petrophysics. CSEG Recorder, 31 Special (March), 1–17.Google Scholar
Goodway, B., Perez, M., Varsek, J., & Abaco, C. (2010). Seismic petrophysics and itotropic-anisotropic AVO methods for unconventional gas exploration. The Leading Edge, December, 15001508. https://doi.org/10.1190/1.3525367Google Scholar
Gourjon, E. & Bertoncello, A. (2018). Impact of near Well-Bore Geology on Hydraulic Fractures Geometry and Well Productivity: A Statistical Look Back at the Utica Play. In SPE Hydraulic Fracturing Technology Conference & Exhibition.Google Scholar
Graham, S. A. & Williams, L. A. (1985). Tectonic, depositional, and diagenetic history of Monterey Formation (Miocene), Centra San Joaquin Basin, California. American Association of Petroleum Geologists Bulletin, 69(3), 385411. https://doi.org/10.1306/AD4624F7-16F7-11D7-8645000102C1865DGoogle Scholar
Grechka, V. & Heigl, W. M. (2017). Microseismic Monitoring. SEG.Google Scholar
Guo, B. (n.d.). Article in press.Google Scholar
Guo, Z., Chapman, M., & Li, X. (2012). Exploring the effect of fractures and microstructure on brittleness index in the Barnett Shale. In SEG Technical Program Expanded Abstracts (Vol. 2, pp. 15). https://doi.org/10.1190/segam2012-0771.1Google Scholar
Gupta, J. K., Zielonka, M. G., Albert, R. A., El-Rabaa, A. W., Burnham, H. A, & Choi, N. H. (2012). SPE 152224 Integrated methodology for optimizing development of unconventional gas resources. Stress: The International Journal on the Biology of Stress. https://doi.org/10.2118/152224-MSGoogle Scholar
Gutierrez, M. & Nyga, R. (2000). Stress-dependent permeability of a de-mineralised fracture in shale, Marine and Petroleum Geology, 17, 895907.Google Scholar
GWPC & IOGCC (2017). Potential Injection-Induced Seismicity Associated with Oil & Gas Development.Google Scholar
Haege, M., Maxwell, S., Sonneland, L., Norton, M., & Resources, P. E. (2012). Integration of Passive Seismic and 3D Reflection Seismic in an Unconventional Shale Gas Play: Relationship between Rock Fabric and Seismic Moment of Microseismic Events. 2012 SEG Annual Meeting, 15.Google Scholar
Hagin, P. N. & Zoback, M. D. (2004a). Viscous deformation of unconsolidated reservoir sands – Part 1: Time-dependent deformation, frequency dispersion, and attenuation. Geophysics, 69(3). https://doi.org/10.1190/1.1759459Google Scholar
Hagin, P. N. & Zoback, M. D. (2004b). Viscous deformation of unconsolidated reservoir sands – Part 2: Linear viscoelastic models. Geophysics, 69(3). https://doi.org/10.1190/1.1759460Google Scholar
Haimson, B. & Fairhurst, C. (1967). Initiation and extension of hydraulic fractures in rocks. Society of Petroleum Engineers Journal, Sept.: 310–318.Google Scholar
Hajiabdolmajid, V. & Kaiser, P. (2003). Brittleness of rock and stability assessment in hard rock tunneling. Tunnelling and Underground Space Technology, 18(1), 3548. https://doi.org/10.1016/S0886-7798(02)00100–1Google Scholar
Hakso, A. & Zoback, M. (2017). Utilizing multiplets as an independent assessment of relative microseismic location uncertainty. The Leading Edge, 36(10), 829836. https://doi.org/10.1190/tle36100829.1Google Scholar
Hakso, A. & Zoback, M. D. (2019). The relation between stimulated shear fractures and production in the Barnett Shale: Implications for unconventional oil and gas reservoirs. Geophysics, 84 (6), B461-B469. doi 10.1190/GEO2018-0545.1.Google Scholar
Hallo, M., Oprsal, I., Eisner, L., & Ali, M. Y. (2014). Prediction of magnitude of the largest potentially induced seismic event. Journal of Seismology, 18(3), 421431. https://doi.org/10.1007/s10950-014–9417-4Google Scholar
Harbaugh, A. W., Banta, E. R., Hill, M. C., & McDonald, M. G. (2010). Modflow-2000, the U.S. Geological Survey Modular Groundwater Model – User Guide to Modularization Concepts and the Groundwater Flow Process.Google Scholar
Healy, J. H., Rubey, W. W., Griggs, D. T., & Raleigh, C. B. (1968). The Denver earthquakes. Science, 161, 13011310.Google Scholar
Heller, R. & Zoback, M. (2014). Adsorption of methane and carbon dioxide on gas shale and pure mineral samples. Journal of Unconventional Oil and Gas Resources, 8 (C). https://doi.org/10.1016/j.juogr.2014.06.001Google Scholar
Heller, R., Vermylen, J. P., & Zoback, M. D. (2014). Experimental investigation of matrix permeability of gas shales. AAPG Bulletin, 98(5), 975995. https://doi.org/10.1306/09231313023Google Scholar
Hennings, P., Allwardt, P., Paul, P., Zahm, C., Reid, R., Alley, H., … Hough, E. (2012). Relationship between fractures,faultzones, stress,andreservoir productivity in the Suban gas field, Sumatra, Indonesia. AAPG Bulletin, 96(4), 753772. https://doi.org/10.1306/08161109084Google Scholar
Hennings, P., Lund Snee, J.-E., Osmond, J. L., Dommisse, R., DeShon, H. R., & Zoback, M. D. (2019). Slip potential of faults in the Fort Worth Basin of North-Central Texas, USA. Science Advances, in review.Google Scholar
Hill, R. (1963). Elastic properties of reinforced solids: Some theoretical principles. Journal of the Mechanics and Physics of Solids, 11(5), 357372. https://doi.org/10.1016/0022–5096(63)90036-XGoogle Scholar
Ho, N. C., Peacor, D. R., & Van Der Pluijm, B. A. (1999). Preferred orientation of phyllosilicates in Gulf Coast mudstones and relation to the smectite-illite transition. Clays and Clay Minerals, 47(4), 495504. https://doi.org/10.1346/CCMN.1999.0470412Google Scholar
Holland, A. A. (2013). Earthquakes triggered by hydraulic fracturing in south-central Oklahoma. Bulletin of the Seismological Society of America, 103(3), 17841792. https://doi.org/10.1785/0120120109Google Scholar
Hornbach, M. J., Deshon, H. R., Ellsworth, W. L., Stump, B. W., Hayward, C., Frohlich, C., … Luetgert, J. H. (2015). Causal factors for seismicity near Azle, Texas. Nature Communications, 6. https://doi.org/10.1038/ncomms7728Google Scholar
Hornbach, M. J., Jones, M., Scales, M., DeShon, H. R., Magnani, M. B., Frohlich, C., … Layton, M. (2016). Ellenburger wastewater injection and seismicity in North Texas. Physics of the Earth and Planetary Interiors, 261, 5468. https://doi.org/10.1016/j.pepi.2016.06.012Google Scholar
Horton, S. (2012). Disposal of hydrofracking waste fluid by injection into subsurface aquifers triggers earthquake swarm in Central Arkansas with potential for damaging earthquake. Seismological Research Letters, 83(2), 250260. https://doi.org/10.1785/gssrl.83.2.250Google Scholar
Hu, H., Li, A., & Zavala-Torres, R. (2017). Long-period long-duration seismic events during hydraulic fracturing: Implications for tensile fracture development. Geophysical Research Letters, 44(10), 48144819. https://doi.org/10.1002/2017GL073582Google Scholar
Huang, J., Morris, J. P., Fu, P., Settgast, R. R., Sherman, C. S., & Ryerson, F. J. (2018). Hydraulic Fracture Height Growth Under the Combined Influence of Stress Barriers and Natural Fractures. SPE Hydraulic Fracturing Technology Conference and Exhibition. https://doi.org/10.2118/189861-MSGoogle Scholar
Huang, Y., Beroza, G. C., & Ellsworth, W. L. (2016). Stress drop estimates of potentially induced earthquakes in the Guy-Greenbrier sequence. Journal of Geophysical Research: Solid Earth, 121(9), 65976607. https://doi.org/10.1002/2016JB013067Google Scholar
Hubbert, M. D. & Rubey, W. W. (1959). Role of fluid pressure in mechanics of overthrust faulting. Geological Society of America Bulletin, 70, 115205.Google Scholar
Hubbert, M. K. & Willis, D. G. (1957). Mechanics of hydraulic fracturing. Petr. Trans. AIME, 210, 153163.Google Scholar
Hucka, V. & Das, B. (1974). Brittleness determination of rocks by different methods. International Journal of Rock Mechanics and Mining Sciences, 11(10), 389392. https://doi.org/10.1016/0148–9062(74)91109–7Google Scholar
Hull, R. A., Leonard, P. A., & Maxwell, S. C. (2017a). Geomechanical Investigation of Microseismic Mechanisms Associated With Slip on Bed Parallel Fractures, 2426. https://doi.org/10.15530/urtec-2017–26888667Google Scholar
Hull, R. A., Meek, R., Bello, H., Resources, P. N., & Miller, D. (2017b). Case History of DAS Fiber-Based Microseismic and Strain Data, Monitoring Horizontal Hydraulic Stimulations Using Various Tools to Highlight Physical Deformation Processes (Part A).Google Scholar
Hurd, O. (2012). Geomechanical analysis of intraplate earthquakes and earthquakes induced during stimulation of low permeability gas reservoirs. Stanford University.Google Scholar
Hurd, O. & Zoback, M. D. (2012a). Intraplate earthquakes, regional stress and fault mechanics in the central and eastern U.S. and Southeastern Canada. Tectonophysics, 581. https://doi.org/10.1016/j.tecto.2012.04.002Google Scholar
Hurd, O. & Zoback, M. D. (2012b). Stimulated Shale Volume Characterization: Multiwell Case Study from the Horn River Shale: I. Geomechanics and Microseismicity. SPE Annual Technical Conference and Exhibition, C. https://doi.org/10.2118/159536-MSGoogle Scholar
Imanishi, K. & Ellsworth, W. L. (2006). Source scaling relationships of microearthquakes at Parkfield, CA, determined using the SAFOD pilot hole seismic array. In Earthquakes: Radiated Energy and the Physics of Faulting (pp. 81–90). American Geophysical Union. https://doi.org/10.1029/170GM10Google Scholar
Inamdar, A. A., Ogundare, T. M., Malpani, R., Atwood, W. K., Brook, K., Erwemi, A. M., & Purcell, D. (2010). Evaluation of Stimulation Techniques Using Microseismic Mapping in the Eagle Ford Shale. Tight Gas Completions Conference. https://doi.org/10.2118/136873-MSGoogle Scholar
Ingram, G. M. & Urai, J. L. (1999). Top-seal leakage through faults and fractures: the role of mudrock properties. Geological Society, London, Special Publications, 158(1), 125135. https://doi.org/10.1144/GSL.SP.1999.158.01.10Google Scholar
Ishibashi, T., Watanabe, N., Asanuma, H., & Tsuchiya, N. (2016). Linking microearthquakes to fracture permeability evolution. Crustal Permeability, 4964. https://doi.org/10.1002/9781119166573.ch7Google Scholar
Al Ismail, M. I. (2016). Influence of nanopores on the transport of gas and gas-condensate in unconventional resources. Stanford University.Google Scholar
Al Ismail, M. I. & Zoback, M. D. (2016). Effects of rock mineralogy and pore structure on extremely low stress-dependent matrix permeability of unconventional shale gas and shale oil samples. Royal Society Philosophical Transactions A. https://doi.org/10.1098/rsta.2015.0428Google Scholar
Al Ismail, M. I., Hol, S., Reece, J. S., & Zoback, M. D. (2014). The Effect of CO2 Adsorption on Permeability Anisotropy in the Eagle Ford Shale. Unconventional Resources Technology Conference, (1921520).Google Scholar
Ito, H. & Zoback, M. D. (2000). Fracture permeability and in situ stress to 7 km depth in the KTB scientific drillhole. Geophysical Research Letters, 27, 10451048.Google Scholar
Jackson, R. B., Vengosh, A., Carey, J. W., Darrah, T. H., O’Sullivan, F., & Pétron, G. (2014). The environmental costs and benefits of fracking. Ssrn, (August), 136. https://doi.org/10.1146/annurev-environ-031113–144051Google Scholar
Jaeger, J. C. & Cook, N. G. W. (1971). Fundamentals of Rock Mechanics (3rd edn.). New York: Chapman and Hall.Google Scholar
Jarvie, D. M., Hill, R. J., Ruble, T. E., & Pollastro, R. M. (2007). Unconventional shale-gas systems: The Mississippian Barnett Shale of north-central Texas as one model for thermogenic shale-gas assessment. AAPG Bulletin, 91(4), 475499. https://doi.org/10.1306/12190606068Google Scholar
Jin, G. & Roy, B. (2017). Hydraulic-fracture geometry characterization using low-frequency DAS signal. The Leading Edge, 36(12), 975980. https://doi.org/10.1190/tle36120975.1Google Scholar
Jin, L. & Zoback, M. D. (2017). Fully coupled nonlinear fluid flow and poroelasticity in arbitrarily fractured porous media: A hybrid‐dimensional computational model. Journal of Geophysical Research, 122, 33. https://doi.org/10.1002/2017JB014892Google Scholar
Jin, L. & Zoback, M. D. (2018)  Fully dynamic spontaneous rupture due to quasi-static pore pressure and poroelastic effects: An implicit nonlinear computational model of fluid-induced seismic events. Journal of Geophysical Research: Solid Earth, 123. https://doi.org/10.1029/2018JB015669Google Scholar
Jin, L. & Zoback, M. (2019). Depletion-induced poroelastic stress changes in naturally fractured unconventional reservoirs and implications for hydraulic fracture propagation, SPE 196215–MS, SPE Ann. Tech. Conf and Exhib., Calgary, Alberta, CA, 30 Sep–2 Oct 2019.Google Scholar
Jin, X., Shah, S. N., Roegiers, J.-C., & Zhang, B. (2014a). Fracability evaluation in shale reservoirs – An integrated petrophysics and geomechanics approach. SPE Journal, (October 2015). https://doi.org/10.2118/168589-MSGoogle Scholar
Jin, X., Shah, S. N., Truax, J. A., & Roegiers, J.-C. (2014b). A Practical Petrophysical Approach for Brittleness Prediction from Porosity and Sonic Logging in Shale Reservoirs. SPE Annual Technical Conference and Exhibition, 18. https://doi.org/10.2118/170972-MSGoogle Scholar
Johnston, J. E. & Christensen, N. I. (1995). Seismic anisotropy of shales. Journal of Geophysical Research, 100(B4), 59916003.Google Scholar
Johri, M., Dunham, E. M., Zoback, M. D., & Fang, Z. (2014a). Predicting fault damage zones by modeling dynamic rupture propagation and comparison with field observations. Journal of Geophysical Research: Solid Earth, 119(2). https://doi.org/10.1002/2013JB010335Google Scholar
Johri, M., Zoback, M. D., & Hennings, P. (2014b). A scaling law to characterize fault-damage zones at reservoir depths. AAPG Bulletin, 98(10), 20572079. https://doi.org/10.1306/05061413173Google Scholar
Jung, H., Sharma, Mukul M., Cramer, D., Oakes, S., & McClure, M. (2016). Re-examining interpretations of non-ideal behavior during diagnostic fracture injection tests, Journal of Petroleum Science and Engineering, 145, 114136, http://dx.doi.org/10.1016/j.petrol.2016.03.016Google Scholar
Jweda, J., Michael, E., Jokanola, O., Hofer, R., & Parisi, V. (2017). Optimizing Field Development Strategy Using Time-Lapse Geochemistry and Production Allocation in Eagle Ford – URTeC: 2671245 (pp. 24–26). Paper was prepared for presentation at the Unconventional Resources Technology Conference held in Austin, Texas, USA, 24–26 July 2017. https://doi.org/10.15530/urtec-2017–2671245Google Scholar
Kahn, D., Roberts, J., & Rich, J. (2017). Integrating Microseismic and Geomechanics to Interpret Hydraulic Fracture Growth. Proceedings of the 5th Unconventional Resources Technology Conference, (2016). https://doi.org/10.15530/urtec-2017–2697445Google Scholar
Kale, S., Rai, C., & Sondergeld, C. (2010). Rock Typing in Gas Shales. In SPE Annual Technical Conference and Exhibition. https://doi.org/10.2118/134539-MSGoogle Scholar
Kaluder, Z., Nikolaev, M., Davidenko, I., Leskin, F., Martynov, M., Shishmanidi, I., Platunov, A., Chong, K. K., Astafyev, V., Shnitiko, A., & Fedorenko, E. (2014) First High-Rate Hybrid Fracture in Em-Yoga Field, West Siberia, Russia. OTC 24712-MS, Offshore Technology Conference-Asia, http://www.onepetro.org/doi/10.4043/24712-MSGoogle Scholar
Kang, S. M., Fathi, E., Ambrose, R. J., Akkutlu, I. Y., & Sigal, R. F. (2011). Carbon dioxide storage capacity of organic-rich shales. SPE Journal, 16(04), 842855. https://doi.org/10.2118/134583-PAGoogle Scholar
Kanitpanyacharoen, W., Kets, F. B., Wenk, H. R., & Wirth, R. (2012). Mineral preferred orientation and microstructure in the Posidonia Shale in relation to different degrees of thermal maturity. Clays and Clay Minerals, 60(3), 315329. https://doi.org/10.1346/CCMN.2012.0600308Google Scholar
Kanitpanyacharoen, W., Wenk, H.-R., Kets, F., Lehr, C., & Wirth, R. (2011). Texture and anisotropy analysis of Qusaiba shales. Geophysical Prospecting, 59(3), 536556. https://doi.org/10.1111/j.1365–2478.2010.00942.xGoogle Scholar
Kassis, S. & Sondergeld, C. (2010). Fracture permeability of gas shale: Effects of roughness, fracture offset, proppant, and effective stress. Society of Petroleum Engineers Journal, (1), 117. https://doi.org/10.2118/131376-MSGoogle Scholar
Katz, D. (ed.). (1959). Handbook of Natural Gas Engineering. McGraw-Hill.Google Scholar
Keller, L. M. & Holzer, L. (2018). Image-based upscaling of permeability in Opalinus clay. Journal of Geophysical Research: Solid Earth, 123(1), 285295. https://doi.org/10.1002/2017JB014717Google Scholar
Kelly, S., El-Sobky, H., Torres-Verdín, C., & Balhoff, M. T. (2016). Assessing the utility of FIB-SEM images for shale digital rock physics. Advances in Water Resources, 95, 302316. https://doi.org/10.1016/j.advwatres.2015.06.010Google Scholar
Kennedy, R. L., Knecht, W. N., & Georgi, D. T. (2012). Comparisons and Contrasts of Shale Gas and Tight Gas Developments, North American Experience and Trends. SPE Saudi Arabia Section Technical Symposium and Exhibition, (August 2005). https://doi.org/10.2118/160855-MSGoogle Scholar
Kennedy, R., Luo, L., & Kusskra, V. (2016). The unconventional basins and plays – North America, the rest of the world and emerging basins. In Ahmed, U. & Meehan, D. N. (eds.), Unconventional Oil and Gas Resources: Exploitation and Development. Boca Raton, FL: CRC Press.Google Scholar
Keranen, K. M., Savage, H. M., Abers, G. A., & Cochran, E. S. (2013). Potentially induced earthquakes in Oklahoma, USA: Links between wastewater injection and the 2011 Mw5.7 earthquake sequence. Geology, 41(6), 699702. https://doi.org/10.1130/G34045.1Google Scholar
Kim, W. Y. (2013). Induced seismicity associated with fluid injection into a deep well in Youngstown, Ohio. Journal of Geophysical Research: Solid Earth, 118(7), 35063518. https://doi.org/10.1002/jgrb.50247Google Scholar
King, G. C. P., Stein, R. S., & Lin, J. (1994). Static stress changes and the triggering of earthquakes. Bulletin Of the Seismological Society of America, 84, 935953.Google Scholar
King, G. E. (2012). Hydraulic Fracturing 101: What Every Representative, Environmentalist, Regulator, Reporter, Investor, University Researcher, Neighbor and Engineer Should Know About Estimating Frac Risk and Improving Frac Performance in Unconventional Gas and Oil Wells. SPE Hydraulic Fracturing Technology Conference, 180. https://doi.org/10.2118/152596-MSGoogle Scholar
King, G. E. (2014). Improving recovery factors in liquids-rich resource plays requires new approaches. Editors Choice Magazine. Retrieved from http://www.aogr.com/magazine/editors-choice/improving-recovery-factors-in-liquids-rich-resource-plays-requires-new-apprGoogle Scholar
King, G. E. & King, D. E. (2013). Environmental risk arising from well-construction failure–Differences between barrier and well failure, and estimates of failure frequency across common well types, locations, and well age. SPE Production & Operations, 28(04), 323344. https://doi.org/10.2118/166142-PAGoogle Scholar
King, G. E., Rainbolt, M. F., Swanson, C., & Corporation, A. (2017). SPE-187192-MS Frac Hit Induced Production Losses: Evaluating Root Causes, Damage Location, Possible Prevention Methods and Success of Remedial Treatments.Google Scholar
King, H. E., Eberle, A. P. R., Walters, C. C., Kliewer, C. E., Ertas, D., & Huynh, C. (2015). Pore architecture and connectivity in gas shale. Energy and Fuels, 29(3), 13751390. https://doi.org/10.1021/ef502402eGoogle Scholar
Klaver, J., Desbois, G., Urai, J. L., & Littke, R. (2012). BIB-SEM study of the pore space morphology in early mature Posidonia Shale from the Hils area, Germany. International Journal of Coal Geology, 103, 1225. https://doi.org/10.1016/j.coal.2012.06.012Google Scholar
Klaver, J., Desbois, G., Littke, R., & Urai, J. L. (2015). BIB-SEM characterization of pore space morphology and distribution in postmature to overmature samples from the Haynesville and Bossier Shales. Marine and Petroleum Geology, 59, 451466. https://doi.org/10.1016/j.marpetgeo.2014.09.020Google Scholar
Klinkenberg, L. J. (1941). The permeability of porous media to liquids and gases. API Drilling and Production Practice, 200–2013. https://doi.org/10.5510/OGP20120200114Google Scholar
Kohli, A. H. & Zoback, M. D. (2013). Frictional properties of shale reservoir rocks. Journal of Geophysical Research: Solid Earth, 118(9), 51095125. https://doi.org/10.1002/jgrb.50346Google Scholar
Kohli, A. H. & Zoback, M. D. (2019). Frictional properties of shale reservoir rocks II: Calcareous shales and thermal controls on frictional stability. In Preparation.Google Scholar
Kramer, S. L. (1996). Geotechnical Earthquake Engineering. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Kranz, R. L., Saltzman, J. S., & Blacic, J. D. (1990). Hydraulic diffusivity measurements on laboratory rock samples using an oscillating pore pressure method. International Journal of Rock Mechanics and Mining Sciences, 27(5), 345352. https://doi.org/10.1016/0148–9062(90)92709-NGoogle Scholar
Kronenberg, A. K., Kirby, S., & Pinkston, J. (1990). Basal slip and mechanical anisotropy of biotite. Journal of Geophysical Research, 95, 1925719278.Google Scholar
Krupnick, A., Gordon, H., & Olmstead, S. (2013). What the Experts Say about the Environmental Risks of Shale Gas Development. Resources for the Future, Pathways to Dialogue (Vol. 2). Retrieved from http://scholar.google.com/scholar?hl=en&q=What+the+Experts+Say+about+the+Environmental+Risks+of+Shale+Gas+Development&btnG=≈sdt=1,5≈sdtp=#0Google Scholar
Kuang, W., Zoback, M., & Zhang, J. (2017). Estimating geomechanical parameters from microseismic plane focal mechanisms recorded during multistage hydraulic fracturing. Geophysics, 82(1), KS1KS11. https://doi.org/10.1190/geo2015-0691.1Google Scholar
Kubo, T. & Katayama, I. (2015). Effect of temperature on the frictional behavior of smectite and illite. Journal of Mineralogical and Petrological Sciences, 110(6), 293299. https://doi.org/10.2465/jmps.150421Google Scholar
Kuila, U., McCarty, D. K., Derkowski, A., Fischer, T. B., Topór, T., & Prasad, M. (2014). Nano-scale texture and porosity of organic matter and clay minerals in organic-rich mudrocks. Fuel, 135, 359373. https://doi.org/10.1016/j.fuel.2014.06.036Google Scholar
Kumar, A. & Hammack, R. (2016). Long period, long duration (LPLD) seismicity observed during hydraulic fracturing of the Marcellus Shale in Greene County, Pennsylvania, (Figure 1), 2684–2688. https://doi.org/10.1190/segam2016-13876878.1Google Scholar
Kumar, A., Chao, K., Hammack, R., & Harbert, W. (2018). Surface Seismic Monitoring of Hydraulic Fracturing Test Site (HFTS) in the Midland Basin, Texas National Energy Technology Laboratory, Department of Energy, Pittsburgh, PA, 1–11. https://doi.org/10.15530/urtec-2018–2902789Google Scholar
Kumar, A., Zorn, E., Hammack, R., & Harbert, W. (2017). Long-period, long-duration seismicity observed during hydraulic fracturing of the Marcellus Shale in Greene County, Pennsylvania, (July), 580–587. https://doi.org/10.1190/tle36070580.1Google Scholar
Kwon, O., Kronenberg, A. K., Gangi, A. F., Johnson, B., & Herbert, B. E. (2004). Permeability of illite-bearing shale: 1. Anisotropy and effects of clay content and loading. Journal of Geophysical Research, 109(B10), 119. https://doi.org/10.1029/2004JB003052Google Scholar
Kwon, O., Kronenberg, A. K., Gangi, A., & Johnson, B. (2001). Permeability of Wilcox shale and its effective pressure law. Journal of Geophysical Research, 106, 1933919353.Google Scholar
LaFollette, R. F., Holcomb, W. D., & Aragon, J. (2012). Practical Data Mining: Analysis of Barnett Shale Production Results with Emphasis on Well Completion and Fracture Stimulation: SPE 152531. Society of Petroleum Engineers. https://doi.org/10.2118/152531-MSGoogle Scholar
Lakes, R. S. (1999). Viscoelastic Solids. CRC Mechanical Engineering Series.Google Scholar
Lalehrokh, F. & Bouma, J. (2014). Well Spacing Optimization in Eagle Ford. SPE/CSUR Unconventional Resources Conference. https://doi.org/10.2118/171640-MSGoogle Scholar
Lan, Q., Xu, M., Binazadeh, M., Dehghanpour, H., & Wood, J. M. (2015). A comparative investigation of shale wettability: The significance of pore connectivity. Journal of Natural Gas Science and Engineering, 27, 11741188. https://doi.org/10.1016/j.jngse.2015.09.064Google Scholar
Langenbruch, C. & Shapiro, S. A. (2010). Decay rate of fluid-induced seismicity after termination of reservoir stimulations. Geophysics, 75(6).Google Scholar
Langenbruch, C. & Zoback, M. D. (2016). How will induced seismicity in Oklahoma respond to decreased saltwater injection rates? Science Advances, 2(11), 19. https://doi.org/10.1126/sciadv.1601542Google Scholar
Langenbruch, C. & Zoback, M. D. (2017). Response to comment on “How will induced seismicity in Oklahoma respond to decreased saltwater injection rates?“ Science Advances, 3(8). https://doi.org/10.1126/sciadv.aao2277Google Scholar
Langenbruch, C., Dinske, C., & Shapiro, S. A. (2011). Inter event times of fluid induced earthquakes suggest their Poisson nature. Geophysical Research Letters, 38(21), 16. https://doi.org/10.1029/2011GL049474Google Scholar
Langenbruch, C., Weingarten, M., & Zoback, M. D. (2018). Physics-based forecasting 1 of man-made earthquake hazards in Oklahoma and Kansas. Nature Communications, 9, 3946. DOI: 10.1038/s41467-018-06167-49Google Scholar
Langmuir, I. (1916). The constitution and fundamental properties of solids and liquids. Part I. Solids. Journal of the American Chemical Society, 38(11), 22212295. https://doi.org/10.1021/ja02268a002Google Scholar
Laplace, P. S., Bowditch, N., & Bowditch, N. I. (1829). Mécanique céleste. Meccanica.Google Scholar
Laubach, S. E., Olson, J. E., & Gale, J. F. W. (2004). Are open fractures necessarily aligned with maximum horizontal stress? Earth and Planetary Science Letters, 222(1), 191195.Google Scholar
Lecampion, B., Bunger, A., & Zhang, X. (2017). Numerical methods for hydraulic fracture propagation: A review of recent trends. Journal of Natural Gas Science and Engineering. https://doi.org/10.1016/j.jngse.2017.10.012Google Scholar
Lee, D., Shrivastava, K., & Sharma, M. M. (2017). Effect of Fluid Rheology on Proppant Transport in Hydraulic Fractures in Soft Sands. American Rock Mechanics Association.Google Scholar
Lee, S., Fischer, T. B., Stokes, M. R., Klingler, R. J., Ilavsky, J., McCarty, D. K., … Winans, R. E. (2014). Dehydration effect on the pore size, porosity, and fractal parameters of shale rocks: Ultrasmall-angle X-ray scattering study. Energy and Fuels, 28(11), 67726779. https://doi.org/10.1021/ef501427dGoogle Scholar
Letham, E. A. & Bustin, R. M. (2016). Klinkenberg gas slippage measurements as a means for shale pore structure characterization. Geofluids, 16(2), 264278. https://doi.org/10.1111/gfl.12147Google Scholar
Leu, L., Georgiadis, A., Blunt, M. J., Busch, A., Bertier, P., Schweinar, K., … Ott, H. (2016). Multiscale description of shale pore systems by scanning SAXS and WAXS microscopy. Energy and Fuels, 30(12), 1028210297. https://doi.org/10.1021/acs.energyfuels.6b02256Google Scholar
Li, J., Zhang, H., Sadi Kuleli, H., & Nafi Toksoz, M. (2011). Focal mechanism determination using high-frequency waveform matching and its application to small magnitude induced earthquakes. Geophysical Journal International, 184(3), 12611274. https://doi.org/10.1111/j.1365-246X.2010.04903.xGoogle Scholar
Li, N., Lolon, E., Mayerhofer, M., Cordts, Y., White, R., & Childers, A. (2017). Optimizing Well Spacing and Well Performance in the Piceance Basin Niobrara Formation. SPE Hydraulic Fracturing Technology Conference and Exhibition. https://doi.org/10.2118/184848-MSGoogle Scholar
Li, S., Yuan, Y., Sun, W., Sun, D., & Jin, Z. (2016). Formation and destruction mechanism as well as major controlling factors of the Silurian shale gas overpressure in the Sichuan Basin, China, Journal of Natural Gas Geoscience, 1(4), 287294. doi: 10.1016/j.jnggs.2016.09.002.Google Scholar
Liang, B., Du, M., Goloway, C., Hammond, R., Yanez, P. P., & Richey, M. (2017). Subsurface Well Spacing Optimization in the Permian Basin. Proceedings of the 5th Unconventional Resources Technology Conference. https://doi.org/10.15530/urtec-2017–2671346Google Scholar
Lindsay, G. J., White, D. J., Miller, G. A., Baihly, J. D., & Sinosic, B. (2016). Understanding the Applicability and Economic Viability of Refracturing Horizontal Wells in Unconventional Plays. SPE Hydraulic Fracturing Technology Conference. https://doi.org/10.2118/179113-MSGoogle Scholar
Linker, M. F. & Dieterich, J. H. (1992). Effects of variable normal stress on rock friction: Observations and constitutive equations. Journal of Geophysical Research, 97(92), 49234940.Google Scholar
Lisjak, A., Grasselli, G., & Vietor, T. (2014). Continuum-discontinuum analysis of failure mechanisms around unsupported circular excavations in anisotropic clay shales. International Journal of Rock Mechanics and Mining Sciences, 65, 96115. https://doi.org/10.1016/j.ijrmms.2013.10.006Google Scholar
Liu, F., Ellett, K., Xiao, Y., & Rupp, J. A. (2013). Assessing the feasibility of CO2 storage in the New Albany Shale (Devonian-Mississippian) with potential enhanced gas recovery using reservoir simulation. International Journal of Greenhouse Gas Control, 17, 111126. https://doi.org/10.1016/j.ijggc.2013.04.018Google Scholar
Liu, S. & Valko, P. P. (2015). An Improved Equilibrium-Height Model for Predicting Hydraulic Fracture Height Migration in Multi-Layered Formations. SPE Hydraulic Fracturing Technology Conference, (1), 116. https://doi.org/10.2118/173335-MSGoogle Scholar
Lockner, D. A., Byerlee, J. D., Kuksenko, V., Ponomarev, A., & Sidorin, A. (1992). Observations of quasistatic fault growth from acoustic emissions. In Fault Mechanics and Transport Properties of Rocks (pp. 129). Academic Press.Google Scholar
Lockner, D. A., Tanaka, H., Ito, H., Ikeda, R., Omura, K., & Naka, H. (2009). Geometry of the Nojima fault at Nojima-Hirabayashi, Japan – I. A simple damage structure inferred from borehole core permeability. Pure and Applied Geophysics, 166(10–11), 16491667. https://doi.org/10.1007/s00024-009–0515-0Google Scholar
Loucks, R. G. & Reed, R. M. (2014). Scanning-electron-microscope petrographic evidence for distinguishing organic-matter pores associated with depositional organic matter versus migrated organic matter in mudrocks. GCAGS Journal, 3, 5160.Google Scholar
Loucks, R. G. & Ruppel, S. C. (2007). Mississippian Barnett Shale: Lithofacies and depositional setting of a deep-water shale-gas succession in the Fort Worth Basin, Texas. AAPG Bulletin, 91(4), 579601. https://doi.org/10.1306/11020606059Google Scholar
Loucks, R. G., Reed, R. M., Ruppel, S. C., & Hammes, U. (2012). Spectrum of pore types and networks in mudrocks and a descriptive classification for matrix-related mudrock pores. AAPG Bulletin, 96(6), 10711098. https://doi.org/10.1306/08171111061Google Scholar
Loucks, R. G., Reed, R. M., Ruppel, S. C., & Jarvie, D. M. (2009). Morphology, genesis, and distribution of nanometer-scale pores in siliceous mudstones of the Mississippian Barnett Shale. Journal of Sedimentary Research, 79(12), 848861. https://doi.org/10.2110/jsr.2009.092Google Scholar
Lund, B. & Townend, J. (2007). Calculating horizontal stress orientations with full or partial knowledge of the tectonic stress tensor. Geophysical Journal International, 170(3), 13281335. https://doi.org/10.1111/j.1365-246X.2007.03468.xGoogle Scholar
Lund Snee, J. E. & Zoback, M. D. (2016). State of stress in Texas: Implications for induced seismicity. Geophysical Research Letters, 43(19),10, 208–10,214. https://doi.org/10.1002/2016GL070974Google Scholar
Lund Snee, J.-E. & Zoback, M. D. (2018a). State of stress in the Central and Eastern U.S. Submitted to Geology.Google Scholar
Lund Snee, J.-E. & Zoback, M. D. (2018b). State of stress in the Permian Basin, Texas and New Mexico: Implications for induced seismicity. The Leading Edge, February. Retrieved from https://doi.org/10.1190/tle37020127.1.Google Scholar
Luo, M., Baker, M. R., & Lemone, D. V. (1994). Distribution and generation of the overpressure system, eastern Delaware Basin, western Texas and southern New Mexico. American Association of Petroleum Geologists Bulletin, 78(9), 13861405. https://doi.org/10.1306/A25FECB1-171B-11D7-8645000102C1865DGoogle Scholar
Lupini, J. F., Skinner, A. E., & Vaughan, P. R. (1981). The drained residual strength of cohesive soils. Géotechnique, 31(2), 181213. https://doi.org/10.1680/geot.1981.31.2.181Google Scholar
Lyman, S. N., Watkins, C., Jones, C. P., Mansfield, M. L., McKinley, M., Kenney, D., & Evans, J. (2017). Hydrocarbon and carbon dioxide fluxes from natural gas well pad soils and surrounding soils in Eastern Utah. Environmental Science and Technology, 51(20), 1162511633. https://doi.org/10.1021/acs.est.7b03408Google Scholar
Lynn, H. B. (2004). The winds of change: Anisotropic rocks – Their preferred direction of fluid flow and their associated seismic signatures – Part 2. The Leading Edge, 23(12), 12581268. https://doi.org/10.1190/leedff.23.1258_1Google Scholar
Lyu, Q., Long, X., Ranjith, P. G., Tan, J., & Kang, Y. (2018). Experimental investigation on the mechanical behaviours of a low-clay shale under water-based fluids. Engineering Geology, 233 (July 2017), 124138. https://doi.org/10.1016/j.enggeo.2017.12.002Google Scholar
Lyu, Q., Ranjith, P. G., Long, X., Kang, Y., & Huang, M. (2015). A review of shale swelling by water adsorption. Journal of Natural Gas Science and Engineering, 27, 14211431. https://doi.org/10.1016/j.jngse.2015.10.004Google Scholar
Ma, L., Fauchille, A.-L., Dowey, P. J., Figueroa Pilz, F., Courtois, L., Taylor, K. G., & Lee, P. D. (2017). Correlative multi-scale imaging of shales: a review and future perspectives. Geological Society, London, Special Publications, 454. https://doi.org/10.1144/SP454.11Google Scholar
Ma, L., Slater, T., Dowey, P. J., Yue, S., Rutter, E. H., Taylor, K. G., & Lee, P. D. (2018). Hierarchical integration of porosity in shales. Scientific Reports, 8(1), 114. https://doi.org/10.1038/s41598-018–30153-xGoogle Scholar
Ma, L., Taylor, K. G., Lee, P. D., Dobson, K. J., Dowey, P. J., & Courtois, L. (2016). Novel 3D centimetre-to nano-scale quantification of an organic-rich mudstone: The Carboniferous Bowland Shale, Northern England. Marine and Petroleum Geology, 72, 193205. https://doi.org/10.1016/j.marpetgeo.2016.02.008Google Scholar
Ma, X. & Zoback, M. D. (2017a). Laboratory experiments simulating poroelastic stress changes associated with depletion and injection in low-porosity sedimentary rocks. Journal of Geophysical Research: Solid Earth, 122(4), 126. https://doi.org/10.1002/2016JB013668Google Scholar
Ma, X. & Zoback, M. D. (2017b). Lithology-controlled stress variations and pad-scale faults: A case study of hydraulic fracturing in the Woodford Shale, Oklahoma. Geophysics, 82(6). https://doi.org/10.1190/GEO2017-0044.1Google Scholar
Ma, X. & Zoback, M. D. (2019). Predicting lithology-controlled stress variations in the Woodford Shale from geophysical well log data, in review, SPE Journal.Google Scholar
Mack, M. & Warpinski, N. R. (1989). Mechanics of hydraulic fracturing. In Economides, M. J. & Nolte, K. (eds.), Reservoir Stimulation (3rd edn., p. 807). Wiley.Google Scholar
Marone, C. & Cox, S. J. D. (1994). Scaling of rock friction constitutive parameters: The effects of surface roughness and cumulative offset on friction of gabbro. Pure and Applied Geophysics, 143(1–3), 359385. https://doi.org/10.1007/BF00874335Google Scholar
Marone, C. & Kilgore, B. (1993). Scaling of the critical slip distance for seismic faulting with shear strain in fault zones. Nature, 362(6421), 618621. https://doi.org/10.1038/362618a0Google Scholar
Marone, C., Raleigh, C. B., & Scholz, C. H. (1990). Frictional behavior and constitutive modeling of simulated fault gouge. Journal of Geophysical Research, 95, 70077025.Google Scholar
Marongiu Porcu, M., Lee, D., Shan, D., & Morales, A. (2015). Advanced Modeling of Interwell Fracturing Interference: An Eagle Ford Shale Oil Study. SPE Hydraulic Fracturing Technology Conference. https://doi.org/10.2018/174902-MSGoogle Scholar
Martin, T., Kotov, S., Nelson, S. G., & Hughes, B. (2016). Stimulation of unconventional reservoirs. In Ahmed, U. & Meehan, D. N. (eds.), Unconventional Oil and Gas Resources: Exploitation and Development. Boca Raton, FL: CRC Press.Google Scholar
Martínez-Garzón, P., Ben-Zion, Y., Abolfathian, N., Kwiatek, G., & Bohnhoff, M. (2016). A refined methodology for stress inversions of earthquake focal mechanisms. Journal of Geophysical Research: Solid Earth, 121(12), 86668687. https://doi.org/10.1002/2016JB013493Google Scholar
Maury, J., Cornet, F. H., & Dorbath, L. (2013). A review of methods for determining stress fields from earthquakes focal mechanisms; Application to the Sierentz 1980 seismic crisis (Upper Rhine Graben). Bulletin de La Societe Geologique de France, 184(4–5), 319334. https://doi.org/10.2113/gssgfbull.184.4–5.319Google Scholar
Mavko, G., Mukerji, T., & Dvorkin, J. (2009). The Rock Physics Handbook, Second Edition: Tools for Seismic Analysis of Porous Media. Cambridge University Press.Google Scholar
Maxwell, S. C. (2009). Assessing the Impact of Microseismic Location Uncertainties On Interpreted Fracture Geometries. SPE Annual Technical Conference and Exhibition, 13. https://doi.org/10.2118/125121-MSGoogle Scholar
Maxwell, S. C. (2012). Statistical Evaluation for Comparative Microseismic Interpretation. 74th EAGE Conference & Exhibition Incorporating SPE EUROPEC 2012, (June 2012), D023.Google Scholar
Maxwell, S. C. (2014). Microseismic Imaging of Hydraulic Fracturing: Improved Engineering of Unconventional Shale Reservoirs. Society of Exploration Geophysicists. https://doi.org/10.1190/1.9781560803164Google Scholar
Maxwell, S. C. & Cipolla, C. (2011). SPE 146932 What Does Microseismicity Tell Us About Hydraulic Fracturing ? (November).Google Scholar
Maxwell, S. C., Raymer, D., Williams, M., & Primiero, P. (2012). Tracking microseismic signals fro the reservoir to the surface. The Leading Edge, (November).Google Scholar
Maxwell, S. C., Shemeta, J., Campbell, E., & Quirk, D. (2008). Microseismic Deformation Rate Monitoring. In EAGE Passive Seismic Workshop – Exploration and Monitoring Applications (p. SPE 116596). https://doi.org/10.2118/116596-MSGoogle Scholar
Maxwell, S. C., Urbancic, T. I., Steinsberger, N., & Zinno, R. (2002). Microseismic Imaging of Hydraulic Fracture Complexity in the Barnett shale, Paper 77440. In Society Petroleum Engineering Annual Technical Conference. San Antonio, TX: Society of Petroleum Engineers.Google Scholar
Mayerhofer, M. J., Lolon, E., Warpinski, N. R., Cipolla, C. L., Walser, D. W., & Rightmire, C. M. (2010). What is stimulated reservoir volume? SPE Production & Operations, 25(01), 8998. https://doi.org/10.2118/119890-PAGoogle Scholar
McClure, M. W. (2017). The Spurious deflection of log-log superposition-time derivative plots of diagnostic fracture-injection tests. SPE Reservoir Evaluation & Engineering, 20(04) (November 2016), 12. https://doi.org/10.2118/186098-PAGoogle Scholar
McClure, M. W. & Horne, R. N. (2011). Investigation of injection-induced seismicity using a coupled fluid flow and rate/state friction model. Geophysics, 76(6), WC181WC198. https://doi.org/10.1190/geo2011-0064.1Google Scholar
McClure, M. W. & Kang, C. A. (2017). A Three-Dimensional Reservoir, Wellbore, and Hydraulic Fracturing Simulator that is Compositional and Thermal, Tracks Proppant and Water Solute Transport, Includes Non-Darcy and Non-Newtonian Flow, and Handles Fracture Closure SPE-182593. In SPE Reservoir Simulation Conference in Montgomery, TX 20–22 February 2017. Society of Petroleum Engineering. https://doi.org/10.2118/182593-MSGoogle Scholar
McClure, M. W., Babazadeh, M., Shiozawa, S., & Huang, J. (2016). Fully coupled hydromechanical simulation of hydraulic fracturing in 3D discrete-fracture networks. SPE Journal, 21(04), 13021320. https://doi.org/10.2118/173354-PAGoogle Scholar
Mckernan, R., Mecklenburgh, J., Rutter, E., & Taylor, K. (2017). Microstructural controls on the pressure-dependent permeability of Whitby Mudstone. Geological Society of London Special Publication, 454. https://doi.org/10.1144/SP454.15Google Scholar
McNamara, D. E., Benz, H. M., Herrmann, R. B., Bergman, E. A., Earle, P., Holland, A., … Gassner, A. (2015). Earthquake hypocenters and focal mechanisms in central Oklahoma reveal a complex system of reactivated subsurface strike-slip faulting. Geophysical Research Letters, 42(8), 27422749. https://doi.org/10.1002/2014GL062730Google Scholar
McNutt, S. R. (1992). Volcanic tremor. In Encyclopedia of Earth System Science (pp. 417425). Academic Press Inc.Google Scholar
McPhee, C., Reed, J., & Zubizarreta, I. (2015). Best practice in coring and core analysis. Developments in Petroleum Science, 64, 115. https://doi.org/10.1016/B978-0–444-63533–4.00001–9Google Scholar
Meléndez-Martínez, J. & Schmitt, D. R. (2016). A comparative study of the anisotropic dynamic and static elastic moduli of unconventional reservoir shales: Implication for geomechanical investigations. Geophysics, 81(3), D245D261. https://doi.org/10.1190/geo2015-0427.1Google Scholar
Michael, A. J. (1984). Determination of stress from slip data: Faults and folds. Journal of Geophysical Research, 89(B13), 1151711526.Google Scholar
Middleton, R. S., Carey, J. W., Currier, R. P., Hyman, J. D., Kang, Q., Karra, S., … Viswanathan, H. S. (2015). Shale gas and non-aqueous fracturing fluids: Opportunities and challenges for supercritical CO2. Applied Energy, 147, 500509. https://doi.org/10.1016/j.apenergy.2015.03.023Google Scholar
Miller, B., Paneitz, J., Mullen, M., Meijs, R., Tunstall, K., & Garcia, M. (2008). The Successful Application of a Compartmental Completion Technique Used to Isolate Multiple Hydraulic-Fracture Treatments in Horizontal Bakken Shale Wells in North Dakota. In SPE Annual Technical Conf. and Exhibition.Google Scholar
Miller, D. E., Daley, T. M., White, D., Freifeld, B. M., Robertson, M., Cocker, J., & Craven, M. (2012). Simultaneous Acquisition of Distributed Acoustic Sensing VSP with Multi-mode and Single-mode Fiber-optic Cables and 3C-Geophones at the Aquistore CO2 Storage Site, CSEG Recorder, 2833.Google Scholar
Miller, G., Lindsay, G., Baihly, J., & Xu, T. (2016). Parent well refracturing: Economic safety nets in an uneconomic market. CC, (May), 56. https://doi.org/10.2118/180200-MSGoogle Scholar
Milliken, K. L. & Day-Stirrat, R. J. (2013). Cementation in mudrocks: Brief review with examples from cratonic basin mudrocks. AAPG Memoir, 103, 133150. https://doi.org/10.1306/13401729H5252Google Scholar
Milliken, K. L., Rudnicki, M., Awwiller, D. N., & Zhang, T. (2013). Organic matter-hosted pore system, Marcellus Formation (Devonian), Pennsylvania. AAPG Bulletin, 97(2), 177200. https://doi.org/10.1306/072312l2O48Google Scholar
Mitchell, J. K. & Soga, K. (2005). Fundamentals of Soil Behavior. New York: John Wiley & Sons.Google Scholar
Mitchell, T. M. & Faulkner, D. R. (2009). The nature and origin of off-fault damage surrounding strike-slip fault zones with a wide range of displacements: A field study from the Atacama fault system, northern Chile. Journal of Structural Geology, 31(8), 802816. https://doi.org/10.1016/j.jsg.2009.05.002Google Scholar
Mokhtari, M., Alqahtani, A. A., Tutuncu, A. N., & Yin, X. (2013). Stress-Dependent Permeability Anisotropy and Wettability of Shale Resources. In Unconventional Resources Technology Conference (pp. 27132728).Google Scholar
Montgomery, S. L., Jarvie, D. M., Bowker, K. A., & Pollastro, R. M. (2005). Mississippian Barnett Shale, Fort Worth basin, north-central Texas: Gas-shale play with multi-trillion cubic foot potential. AAPG Bulletin, 89(2), 155175. https://doi.org/10.1306/09170404042Google Scholar
Moore, D. E. & Lockner, D. A. (2004). Crystallographic controls on the frictional behavior of dry and water-saturated sheet structure minerals. Journal of Geophysical Research, 109(B3), 116. https://doi.org/10.1029/2003JB002582Google Scholar
Moos, D. & Zoback, M. D. (1990). Utilization of observations of well bore failure to constrain the orientation and magnitude of crustal stresses: Application to continental deep sea drilling project and ocean drilling program boreholes. Journal of Geophysical Research, 95, 93059325.Google Scholar
Moos, D. & Zoback, M. D. (1993). State of stress in the Long Valley caldera, California. Geology, 21, 837. https://doi.org/10.1130/0091–7613(1993)021<0837:SOSITL>2.3.CO;2Google Scholar
Moos, D., Vassilellis, G., Cade, R., Franquet, J., Lacazette, A., Bourtembourg, E., & Daniel, G. (2011). Predicting Shale Reservoir Response to Stimulation in the Upper Devonian of West Virginia. SPE Annual Technical Conference and Exhibition, 16.Google Scholar
Morales, A., Zhang, K., Gekhar, K., Marongiu Porcu, M., Lee, D., Shan, D., … Acock, A. (2015). Advanced Modeling of Interwell Fracturing Interference: An Eagle Ford Shale Oil Study – Refracturing. SPE Hydraulic Fracturing Technology Conference. https://doi.org/10.2116/179177-MSGoogle Scholar
Morales, A., Zhang, K., Gakhar, K., Porcu, M., Lee, D. Shan, D. Malpani, R. Pope, T, Sobernheim, D, Acock, S. (2016). Advanced Modelling of Interwell Fracturing Interference: An Eagle Ford Shale Oil Study – Refracturing, SPE 179177-MS, SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 9–11 February 2016.Google Scholar
Morrow, C. A., Bo-Chong, Z., & Byerlee, J. D. (1986). Effective pressure law for permeability of Westerly granite under cycling loading. Journal of Geophysical Research, 91(6), 38703876. https://doi.org/10.1029/JB091iB03p03870Google Scholar
National Academies Press (2017). Flowback and Produced Waters. National Academies Press. https://doi.org/10.17226/24620Google Scholar
National Research Council (2012). Induced Seismicity Potential in Energy Technologies. National Academies Press.Google Scholar
Nelson, P. H. (2003). A review of the multiwell experiment, Williams Fork and Iles Formations, Garfield County, Colorado. In U.S. Geological Survey Digital Data Series DDS–69–B (Chapter 15). U.S. Geological Survey.Google Scholar
Nelson, P. H. (2009). Pore-throat sizes in sandstones, tight sandstones, and shales. AAPG Bulletin, 93(3), 329340. https://doi.org/10.1306/10240808059Google Scholar
Nelson, P. H., Gianoutsos, N. J., & Drake, R. M. (2015). Underpressure in mesozoic and paleozoic rock units in the midcontinent of the United States. AAPG Bulletin, 99(10), 18611892. https://doi.org/10.1306/04171514169Google Scholar
Nicholson, C. & Wesson, R. L. (1992). Triggered earthquakes and deep well activities. Pure & Applied Geophysics, 139(3), 561578.Google Scholar
Nicot, J. P., Scanlon, B. R., Reedy, R. C., & Costley, R. A. (2014). Source and fate of hydraulic fracturing water in the Barnett Shale: A historical perspective. Environmental Science and Technology, 24642471. https://doi.org/dx.doi.org/10.1021/es404050rGoogle Scholar
Nolte, K. (1979). Determination of Fracture Parameters From Fracturing Pressure Decline, SPE-8341-MS. SPE Annual Technical Conference and Exhibition, Las Vegas, Nevada, USA, 23–26 September. https://doi.org/10.2118/8341-MS.Google Scholar
Nolte, K., Maniere, J. L., & Owens, K. A. (1979). After-Closure Analysis of Fracture Calibration Tests. Paper SPE 38676, Society of Petroleum Engineers.Google Scholar
Nur, A. & Byerlee, J. D. (1971). An exact effective sress law for elastic deformation of rock with fluids. Journal of Geophysical Research, 64146419.Google Scholar
Nuttall, B. C., Drahovzal, J. A., Eble, C., & Bustin, R. M. (2005). CO2 Sequestration in Gas Shales of Kentucky (Vol. 19).Google Scholar
Nygaard, K. J., Cardenas, J., Krishna, P. P., Ellison, T. K., & Templeton-Barrett, E. L. (2013). Technical Considerations Associated with Risk Management of Potential Induced Seismicity in Injection Operations. In 5to. Congreso de Producción y Desarrollo de Reservas. Retrieved from https://pangea.stanford.edu/researchgroups/scits/sites/default/files/Argentina_Congress_May2013_TechConRiskManIndSeismicity_Final.pdfGoogle Scholar
Odling, N. E., Gillespie, P., Bourgine, B., Castaing, C., Chiles, J. P., Christensen, N. P., … Watterson, J. (1999). Variations in fracture system geometry and their implications for fluid flow in fractures hydrocarbon reservoirs. Petroleum Geoscience, 5(4), 373384. https://doi.org/10.1144/petgeo.5.4.373Google Scholar
Ogwari, P. O., DeShon, H. R., & Hornbach, M. J. (2018). The Dallas-Fort Worth Airport earthquake sequence: Seismicity beyond injection period. Journal of Geophysical Research: Solid Earth, 123(1), 553563. https://doi.org/10.1002/2017JB015003Google Scholar
Ojha, S. P., Misra, S., Tinni, A., Sondergeld, C., & Rai, C. (2017). Pore connectivity and pore size distribution estimates for Wolfcamp and Eagle Ford shale samples from oil, gas and condensate windows using adsorption-desorption measurements. Journal of Petroleum Science and Engineering, 158, 454468. https://doi.org/10.1016/j.petrol.2017.08.070Google Scholar
Parshall, J. (2018). “Enormous” merge play resource rivals major world gas fields, largest discoveries. Journal of Petroleum Technology, 18 March.Google Scholar
Passey, Q. R., Bohacs, K. M., Esch, W. L., Klimentidis, R., & Sinha, S. (2010). From Oil-Prone Source Rock to Gas-Producing Shale Reservoir – Geologic and Petrophysical Characterization of Unconventional Shale-Gas Reservoirs. CPS/SPE International Oil & Gas Conference and Exhibition in China, SPE145849, 17071735. https://doi.org/10.2118/131350-MSGoogle Scholar
Patchen, D. G. & Carter, K. M., eds. (2015). Utica Shale Appalachian Basin Exploration Consortium Final Report. West Virginia University.Google Scholar
Patel, H., Johanning, J., & Fry, M. (2013). Borehole Microseismic, Completion and Production Data Analysis to Determine Future Wellbore Placement, Spacing and Vertical Connectivity, Eagle Ford Shale, South Texas. In Unconventional Resources Technology Conference, Denver, Colorado, 12–14 August 2013 (pp. 225236). https://doi.org/10.1190/urtec2013-026Google Scholar
Paterson, M. S. & Wong, T. (2005). Experimental Rock Deformation – The Brittle Field. Berlin: Springer.Google Scholar
Pathi, V. S. M. (2008). Factors Affecting the Permeability of Gas Shales. University of British Columbia.Google Scholar
Patzek, T., Male, F., & Marder, M. (2013). Gas production in the Barnett Shale obeys a simple scaling theory. Proceedings of the National Academy of Sciences, 110(49), 1973119736. https://doi.org/10.1073/pnas.1313380110Google Scholar
Patzek, T., Male, F., & Marder, M. (2014). A simple model of gas production from hydrofractured horizontal wells in shales. AAPG Bulletin, 98(12), 25072529. https://doi.org/10.1306/03241412125Google Scholar
Paul, P., Zoback, M., & Hennings, P. (2009). Fluid Flow in a Fractured Reservoir Using a Geomechanically Constrained Fault-Zone-Damage Model for Reservoir Simulation. SPE Reservoir Evaluation and Engineering, August (4).Google Scholar
Peltonen, C., Marcussen, Ø., Bjørlykke, K., & Jahren, J. (2009). Clay mineral diagenesis and quartz cementation in mudstones: The effects of smectite to illite reaction on rock properties. Marine and Petroleum Geology, 26(6), 887898. https://doi.org/10.1016/J.MARPETGEO.2008.01.021Google Scholar
Peng, S. & Xiao, X. (2017). Investigation of multiphase fluid imbibition in shale through synchrotron-based dynamic micro-CT imaging. Journal of Geophysical Research: Solid Earth, 122(6), 44754491. https://doi.org/10.1002/2017JB014253Google Scholar
Peng, S., Yang, J., Xiao, X., Loucks, B., Ruppel, S. C., & Zhang, T. (2015). An integrated method for upscaling pore-network characterization and permeability estimation: Example from the Mississippian Barnett Shale. Transport in Porous Media, 109(2), 359376. https://doi.org/10.1007/s11242-015–0523-8Google Scholar
Perez Altamar, R. & Marfurt, K. (2014). Mineralogy-based brittleness prediction from surface seismic data: Application to the Barnett Shale. Interpretation, 2(4), T255T271. https://doi.org/10.1190/INT-2013–0161.1Google Scholar
Perez Altamar, R. & Marfurt, K. J. (2015). Identification of brittle/ductile areas in unconventional reservoirs using seismic and microseismic data: Application to the Barnett Shale. Interpretation, 3(4), T233T243. https://doi.org/10.1190/INT-2013–0021.1Google Scholar
Peters, K. E., Walters, C. C., & Moldowan, J. M. (2005). Origin and preservation of organic matter. The Biomarker Guide, 1, 317.Google Scholar
Pettegrew, J. & Qiu, J. (2016). Understanding Wolfcamp Well Performance – A Workflow to Describe the Relationship Between Well Spacing and EUR, 13. https://doi.org/10.15530-urtec-2016–2464916Google Scholar
Poirier, J.-P. (1985). Creep of Crystals:High-Temperature Deformation Processes in Metals, Ceramics and Minerals. Cambridge University Press.Google Scholar
Pollastro, R. M. (2007). Total petroleum system assessment of undiscovered resources in the giant Barnett Shale continuous (unconventional) gas accumulation, Fort Worth Basin, Texas. AAPG Bulletin, 91(4), 551578. https://doi.org/10.1306/06200606007Google Scholar
Pollastro, R. M., Jarvie, D. M., Hill, R. J., & Adams, C. W. (2007). Geologic framework of the Mississippian Barnett Shale, Barnett-Paleozoic total petroleum system, Bend arch-Fort Worth Basin, Texas. AAPG Bulletin, 91(4), 405436. https://doi.org/10.1306/10300606008Google Scholar
Rahm, B. G. & Riha, S. J. (2012). Toward strategic management of shale gas development: Regional, collective impacts on water resources. Environmental Science and Policy, 17, 1223. https://doi.org/10.1016/j.envsci.2011.12.004Google Scholar
Rainbolt, M. F., Corporation, A., & Esco, J. (2018). SPE-189853-MS Paper Title: Frac Hit Induced Production Losses: Evaluating Root Causes, Damage Location, Possible Prevention Methods and Success of Remediation Treatments, Part II Case Study V, Wolfcamp, “ PARENT “ – ” CHILD ” Relationship, 187192 (partI).Google Scholar
Raleigh, C. B., Healy, J. H., & Bredehoeft, J. D. (1976). An experiment in earthquake control at Rangely, Colorado. Science, 191, 12301237.Google Scholar
Ramanathan, V., Boskovic, D., Zhmodik, A., Li, Q., Ansarizadeh, M., Michi, O. P., & Garcia, G. (2015). A Simulation Approach to Modelling and Understanding Fracture Geometry with Respect to Well Spacing in Multi Well Pads in the Duvernay – A Case.Google Scholar
Randen, T., Pedersen, S., & Sønneland, L. (2001). Automatic extraction of fault surfaces from three dimensional seismic data. In SEG Technical Program Expanded Abstracts (pp. 551554). https://doi.org/10.1190/1.1816675Google Scholar
Randolph, P. L., Soeder, D. J., & Chowdiah, P. (1984). Porosity and Permeability of Tight Sands. SPE Unconventional Gas Recovery Symposium. https://doi.org/10.2118/12836-MSGoogle Scholar
Rassouli, F. (2018). Laboratory study on the effects of carbonates and clay content on viscoplastic deformation of shales at reservoir conditions. Stanford University.Google Scholar
Rassouli, F. S. & Zoback, M. D. (2018). Comparison of short-term and long-term creep experiments in shales and carbonates from unconventional gas reservoirs. Rock Mechanics and Rock Engineering. https://doi.org/10.1007/s00603-018–1444-yGoogle Scholar
Rassouli, F. S., Ross, C. M., & Zoback, M. D. (2016). Shale Rock Characterization Using Multi-Scale Imaging. American Rock Mechanics Association.Google Scholar
Raterman, K. T., Farrell, H. E., Mora, O. S., Janssen, A. L., Gomez, G. A., Busetti, S., … Warren, M. (2017). Sampling a Stimulated Rock Volume: An Eagle Ford Example (pp. 2426). https://doi.org/10.15530/urtec-20172670034Google Scholar
Revil, A., Grauls, D., & Brevart, O. (2002). Mechanical compaction of sand/clay mixtures. Journal of Geophysical Research, 107(B11), 2293. https://doi.org/10.1029/2001JB000318Google Scholar
Reynolds, A. C., Bonnie, R. J. M., Kelly, S., Krumm, R., & Group, P. O. (2018). Quantifying Nanoporosity: Insights Revealed by Parallel and Multiscale Analyses, (July), 112. https://doi.org/10.15530/urtec-2018–2898355Google Scholar
Rickman, R., Mullen, M. J., Petre, J. E., Grieser, W. V., & Kundert, D. (2008). A Practical Use of Shale Petrophysics for Stimulation Design Optimization: All Shale Plays Are Not Clones of the Barnett Shale. SPE Annual Technical Conference and Exhibition, (Wang), 111. https://doi.org/10.2118/115258-MSGoogle Scholar
Rittenhouse, S., Currie, J., & Blumstein, R. (2016). Using Mud Weights, DST, and DFIT Data to Generate a Regional Pore Pressure Model for the Delaware Basin, New Mexico and Texas, 1243–1252.Google Scholar
Rong, G., Yang, J., Cheng, L., & Zhou, C. (2016). Laboratory investigation of nonlinear flow characteristics in rough fractures during shear process. Journal of Hydrology, 541, 13851394. https://doi.org/10.1016/j.jhydrol.2016.08.043Google Scholar
Ross, D. J. K. & Bustin, M. (2009). The importance of shale composition and pore structure upon gas storage potential of shale gas reservoirs. Marine and Petroleum Geology, 26(6), 916927. https://doi.org/10.1016/j.marpetgeo.2008.06.004Google Scholar
Roy, B., Hart, B., Mironova, A., Zhou, C., & Zimmer, U. (2014). Integrated characterization of hydraulic fracture treatments in the Barnett Shale: The Stocker geophysical experiment. Interpretation, 2(2), T111T127. https://doi.org/10.1190/INT-2013–0071.1Google Scholar
Rubinstein, J. L. & Mahani, A. B. (2015). Myths and facts on wastewater injection, hydraulic fracturing, enhanced oil recovery, and induced seismicity. Seismological Research Letters, 86(4), 10601067. https://doi.org/10.1785/0220150067Google Scholar
Rucker, W. K., Oil, E., States, U., Bobich, J., Oil, E., States, U., … States, U. (2016). Low Cost Field Application of Pressure Transient Communication for Rapid Determination of the Upper Limit of Horizontal Well Spacing, 2466–2474.Google Scholar
Rüger, A. & Gray, D. (2014). Analysis of anisotropic fractured reservoirs. Encyclopedia of Exploration Geophysics, N1N14.Google Scholar
Ruina, A. (1983). Slip instability and state variable friction laws. Journal of Geophysical Research. https://doi.org/10.1029/JB088iB12p10359Google Scholar
Rutledge, J. & Phillips, W. S. (2003). Hydraulic stimulation of natural fractures as revealed by induced microearthquakes, Carthage Cotton Valley gas field, east Texas. Geophysics, 68(2), 441452. https://doi.org/10.1190/1.1567214Google Scholar
Rutledge, J., Downie, R., Maxwell, S., & Drew, J. (2013). Geomechanics of Hydraulic Fracturing Inferred from Composite Radiation Patterns of Microseismicity. Proceedings of SPE Annual Technical Conference and Exhibition, SPE 166370. https://doi.org/10.2118/166370-MSGoogle Scholar
Rutledge, J., Phillips, W. S., & Mayerhofer, M. J. (2004). Faulting induced by forced fluid injection and fluid flow forced by faulting: An interpretation of hydraulic-fracture microseismicity, Carthage Cotton Valley gas field, Texas. Bulletin of the Seismological Society of America, 94(5), 18171830. https://doi.org/10.1785/012003257Google Scholar
Rutledge, J., Weng, X., Chapman, C., Yu, X., & Leaney, S. (2016). Bedding-Plane Slip as a Microseismic Source During Hydraulic Fracturing. SEG International Exposition and 86th Annual Meeting, 25552559. https://doi.org/10.1190/segam2016-13966680.1Google Scholar
Rutledge, J., Yu, X., Leaney, S., Bennett, L., & Maxwell, S. (2014). Microseismic shearing generated by fringe cracks and bedding-plane slip. SEG Technical Program Expanded Abstracts 2014, 22672272. https://doi.org/10.1190/segam2014-0896.1Google Scholar
Rutter, E. H. (1974). The influence of temperature, strain rate and interstitial water in the experimental deformation of calcite rocks. Tectonophysics, 22(3–4), 311334. https://doi.org/10.1016/0040–1951(74)90089–4Google Scholar
Rutter, E. H. & Mecklenburgh, J. (2018). Influence of normal and shear stress on the hydraulic transmissivity of thin cracks in a tight quartz sandstone, a granite, and a shale. Journal of Geophysical Research: Solid Earth, 123(2), 12621285. https://doi.org/10.1002/2017JB014858Google Scholar
Rutter, E. H., Hackston, A. J., Yeatman, E., Brodie, K. H., Mecklenburgh, J., & May, S. E. (2013). Reduction of friction on geological faults by weak-phase smearing, Journal of Structural Geology, 51, 5260.Google Scholar
Sahai, V., Jackson, G., & Rai, R. (2012). SPE 1557 Optimal Well W Spaccing Configurations for Unconventional Gas Reservoirs. Americas Unconventional Resources Conference, June 5–7, 2013, (2010 0).Google Scholar
Saldungaray, P. & Palisch, T. (2012). Hydraulic fracture optimization in unconventional reservoirs. World Oil, (Spec. Suppl.), 713.Google Scholar
Sandrea, R. & Sandrea, I. (2014). New well-productivity data provide US shale potential insights. Oil & Gas, 6677.Google Scholar
Sardinha, C., Petr, C., Lehmann, J., Pyecroft, J., Merkle, S., & Energy, N. (2014). Determining Interwell Connectivity and Reservoir Complexity Through Frac Introduction to the Horn River Shales, (Table 1), 115.Google Scholar
Saunois, M., Jackson, R. B., Bousquet, P., Poulter, B., & Candell, J. G. (2016). The growing role of methane in anthropogenic climate change. Environmental Research Letters, 11. https://doi.org/doi:10.1088/1748-9326/11/12/120207Google Scholar
Savage, H. M. & Brodsky, E. E. (2011). Collateral damage: Evolution with displacement of fracture distribution and secondary fault strands in fault damage zones. Journal of Geophysical Research: Solid Earth, 116(3). https://doi.org/10.1029/2010JB007665Google Scholar
Sayers, C. M. (1994). The elastic anisotrophy of shales. Journal of Geophysical Research, 99(B1), 767. https://doi.org/10.1029/93JB02579Google Scholar
Scales, M. M., DeShon, H. R., Magnani, M. B., Walter, J. I., Quinones, L., Pratt, T. L., & Hornbach, M. J. (2017). A decade of induced slip on the causative fault of the 2015 Mw4.0 Venus earthquake, Northeast Johnson County, Texas. Journal of Geophysical Research: Solid Earth, 122(10), 78797894. https://doi.org/10.1002/2017JB014460Google Scholar
Scanlon, B. R., Reedy, R. C., & Nicot, J. P. (2014). Comparison of water use for hydraulic fracturing for unconventional oil and gas versus conventional oil. Environmental Science & Technology, 48(20), 12386–93. https://doi.org/10.1021/es502506vGoogle Scholar
Scanlon, B. R., Reedy, R. C., Male, F., & Walsh, M. (2017). Water issues related to transitioning from conventional to unconventional oil production in the Permian Basin. Environmental Science & Technology, acs.est.7b02185. https://doi.org/10.1021/acs.est.7b02185Google Scholar
Schieber, J. (2010). Common Themes in the Formation and Preservation of Intrinsic Porosity in Shales and Mudstones – Illustrated with Examples Across the Phanerozoic. SPE Unconventional Gas Conference. https://doi.org/10.2118/132370-MSGoogle Scholar
Schoenball, M. & Ellsworth, W. L. (2017). Waveform‐relocated earthquake catalog for Oklahoma and Southern Kansas illuminates the regional fault network. Seismological Research Letters, 88(5), 12521258. https://doi.org/10.1785/0220170083Google Scholar
Schoenball, M., Walsh, F. R., Weingarten, M., & Ellsworth, W. L. (2018). How faults wake up: The Guthrie-Langston, Oklahoma earthquakes. The Leading Edge, 37(2), 100106. https://doi.org/10.1190/tle37020100.1Google Scholar
Scuderi, M. M. & Collettini, C. (2016). The role of fluid pressure in induced vs. triggered seismicity: insights from rock deformation experiments on carbonates. Scientific Reports, 6(24852). https://doi.org/10.1038/srep24852Google Scholar
Scuderi, M. M., Collettini, C., & Marone, C. (2017a). Fluid-injection and the mechanics of frictional stability of shale-bearing faults. In EGU General Assembly Conference Abstracts.Google Scholar
Scuderi, M. M., Collettini, C., & Marone, C. (2017b). Frictional stability and earthquake triggering during fluid pressure stimulation of an experimental fault. Earth and Planetary Science Letters, 477, 8496. https://doi.org/10.1016/j.epsl.2017.08.009Google Scholar
Segall, P. (1995). A note on induced stress changes in hydrocarbon and geothermal reservoirs.Google Scholar
Segall, P., Grasso, J. R., & Mossop, A. (1994). Poroelastic stressing and induced seismicity near the Lacq Gas Field, Southwestern France. Journal of Geophysical Research, 99(15), 423.Google Scholar
Segall, P., Rubin, A. M., Bradley, A. M., & Rice, J. R. (2010). Dilatant strengthening as a mechanism for slow slip events. Journal of Geophysical Research: Solid Earth, 115(12). https://doi.org/10.1029/2010JB007449Google Scholar
Sen, V., Min, K. S., Ji, L., & Sullivan, R. (2018). Completions and Well Spacing Optimization by Dynamic SRV Modeling for Multi-Stage Hydraulic Fracturing SPE 191571. In SPE ATCE.Google Scholar
Shaffer, D. L., Chavez, L. H., Ben-Sasson, M., Castrillon, S., Yip, N., & Elimelech, M. (2013). Critical review desalination and reuse of high-salinity shale gas produced water: Drivers, technologies, and future directions. Environmental Science & Technology, 47, 95699583. https://doi.org/dx.doi.org/10.1021/es401966eGoogle Scholar
Shaffner, J., Cheng, A., Simms, S., Keyser, E., & Yu, M. (2011). The Advantage of Incorporating Microseismic Data into Fracture Models. Canadian Unconventional Resources Conference, 112.Google Scholar
Shanley, K. W., Cluff, R. M., & Robinson, J. W. (2004). Factors controlling prolific gas production from low permeability sandstone reservoirs implications for resource assessment prospect development and risk analysis. AAPG Bulletin, 88(8), 10831121.Google Scholar
Shapiro, S. A. (2015). Fluid-Induced Seismicity. Cambridge University Press.Google Scholar
Shapiro, S. A., Dinske, C., Langenbruch, C., & Wenzel, F. (2010). Seismogenic index and magnitude probability of earthquakes induced during reservoir fluid stimulations. The Leading Edge, March, 304309.Google Scholar
Shemeta, J. & Anderson, P. (2010). It’s a matter of size: Magnitude and moment estimates for microseismic data. The Leading Edge, (March).Google Scholar
Shen, Y., Ge, H., Meng, M., Jiang, Z., & Yang, X. (2017). Effect of water imbibition on shale permeability and its influence on gas production. Energy and Fuels, 31(5), 49734980. https://doi.org/10.1021/acs.energyfuels.7b00338Google Scholar
Shrivastava, K., Hwang, J., & Sharma, M. M. (2018). Formation of Complex Fracture Networks in the Wolfcamp Shale: Calibrating Model Predictions with Core Measurements from the Hydraulic Fracturing Test Site SPE-191630. In SPE Annual Technical Conference and Exhibition.Google Scholar
Siddiqui, M. A. Q., Ali, S., Fei, H., & Roshan, H. (2018). Current understanding of shale wettability: A review on contact angle measurements. Earth-Science Reviews, 181 (October2017), 111. https://doi.org/10.1016/j.earscirev.2018.04.002Google Scholar
Siddiqui, S. & Kumar, A. (2016). Well Interference Effects for Multiwell Configurations in Unconventional Reservoirs. SPE.Google Scholar
Sigal, R. F. (2013). Mercury capillary pressure measurements on Barnett Core. SPE Reservoir Evaluation & Engineering, 16(04), 432442. https://doi.org/10.2118/167607-PAGoogle Scholar
Sileny, J., Hill, D. P., Eisner, L., & Cornet, F. H. (2009). Non-double-couple mechanisms of microearthquakes induced by hydraulic fracturing. Journal of Geophysical Research: Solid Earth, 114(8), 115. https://doi.org/10.1029/2008JB005987Google Scholar
Simpson, R. W. (1997). Quantifying Anderson’s fault types, Journal of Geophysical Research, 102, 909919.Google Scholar
Singh, H. (2016). A critical review of water uptake by shales. Journal of Natural Gas Science and Engineering, 34, 751766. https://doi.org/10.1016/j.jngse.2016.07.003Google Scholar
Sinha, S., Braun, E. M., Passey, Q. R., Leonardi, S. A., Wood, A. C., Zirkle, T., … Kudva, R. A. (2012). Advances in Measurement Standards and Flow Properties Measurements for Tight Rocks such as Shales. In SPE/EAGE European Unconventional Resources Conference and Exhibition. Society of Petroleum Engineers. https://doi.org/10.2118/152257-MSGoogle Scholar
Skempton, A. W. (1960). Effective Stress in Soils, Concrete, and Rock, Pore Pressure and Suction in Soils. London: Butterworths.Google Scholar
Skoumal, R. J., Brudzinski, M. R., & Currie, B. S. (2015). Earthquakes induced by hydraulic fracturing in Poland township, Ohio. Bulletin of the Seismological Society of America, 105(1), 189197. https://doi.org/10.1785/0120140168Google Scholar
Skoumal, R. J., Brudzinski, M. R., & Currie, B. S. (2018a). Proximity of Precambrian basement affects the likelihood of induced seismicity in the Appalachian, Illinois, and Williston Basins, central and eastern United States. Geosphere, 14(3), 13651379. https://doi.org/10.1130/GES01542.1Google Scholar
Skoumal, R. J., Brudzinski, M. R., Barbour, A., Ries, R., & Currie, B. S. (2018b). Earthquakes Induced by Hydraulic Fracturing Are Pervasive in Oklahoma. In 2018 Banff International Induced Seismicity Workshop, 24–27 October 2018.Google Scholar
Slatt, R. M. & Abousleiman, Y. (2011). Merging sequence stratigraphy and geomechanics for unconventional gas shales. The Leading Edge, 30(3), 274. https://doi.org/10.1190/1.3567258Google Scholar
Smith, M. B. & Montgomery, C. (2015). Hydraulic Fracturing: Emerging Trends and Technologies in Petroleum Engineering. Boca Raton, FL: CRC Press.Google Scholar
Sondergeld, C. H., Ambrose, R. J., Rai, C. S., & Moncrieff, J. (2010a). Micro-Structural Studies of Gas Shales. SPE Annual Technical Conference and Exhibition, SPE-131771. https://doi.org/10.2118/131771-MSGoogle Scholar
Sondergeld, C., Newsham, K., Comisky, J., Rice, M., & Rai, C. (2010b). Petrophysical Considerations in Evaluating and Producing Shale Gas Resources. SPE Unconventional Gas Conference, 134. https://doi.org/10.2118/131768-MSGoogle Scholar
Sone, H. (2012). Mechanical properties of shale gas reservoir rocks and its relation to the in-situ stress variation observed in shale gas reservoirs. PhD Thesis, Stanford University. https://doi.org/10.1017/CBO9781107415324.004Google Scholar
Sone, H. & Zoback, M. D. (2013a). Mechanical properties of shale-gas reservoir rocks – Part 1: Static and dynamic elastic properties and anisotropy. Geophysics, 78(5), D381D392. https://doi.org/10.1190/geo2013-0050.1Google Scholar
Sone, H. & Zoback, M. D. (2013b). Mechanical properties of shale-gas reservoir rocks – Part 2: Ductile creep, brittle strength, and their relation to the elastic modulus. Geophysics, 78(5), D393D402. https://doi.org/10.1190/geo2013-0051.1Google Scholar
Sone, H. & Zoback, M. D. (2014a). Time-dependent deformation of shale gas reservoir rocks and its long-term effect on the in situ state of stress. International Journal of Rock Mechanics and Mining Sciences, 69, 120132. https://doi.org/10.1016/j.ijrmms.2014.04.002Google Scholar
Sone, H. & Zoback, M. D. (2014b). Viscous relaxation model for predicting least principal stress magnitudes in sedimentary rocks. Journal of Petroleum Science and Engineering, 124, 416431. https://doi.org/10.1016/j.petrol.2014.09.022Google Scholar
Song, F. & Toksöz, M. N. (2011). Full-waveform based complete moment tensor inversion and source parameter estimation from downhole microseismic data for hydrofracture monitoring. Geophysics, 76(6), WC103WC116. https://doi.org/10.1190/geo2011-0027.1Google Scholar
Stanek, F. & Eisner, L. (2013). New model explaining inverted source mechanisms of microseismic events induced by hydraulic fracturing, 22012205. https://doi.org/10.1190/segam2013-0554.1Google Scholar
Stanek, F. & Eisner, L. (2017). Seismicity induced by Hydraulic Fracturing in Shales: A Bedding Plane Slip Model. Journal of Geophysical Research, 122. https://doi.org/10.1002/2017JB014213Google Scholar
Staněk, F., Eisner, L., & Vesnaver, A. (2017). Theoretical assessment of the full-moment-tensor resolvability for receiver arrays used in microseismic monitoring. Acta Geodynamica et Geomaterialia, 14(2), 235240. https://doi.org/10.13168/AGG.2017.0006Google Scholar
Stein, S. & Wysession, M. (2003). An Introduction to Seismology, Earthquakes and Earth Structure. Malden, MA: Blackwell Publishing.Google Scholar
Stephenson, B., Galan, E., Williams, W., Macdonald, J., Azad, A., Carduner, R., & Canada, S. (2018). SPE-189863-MS Geometry and Failure Mechanisms from Microseismic in the Duvernay Shale to Explain Changes in Well Performance with Drilling Azimuth, 1–20.Google Scholar
Stover, C. W. & Coffman, J. L. (1989). Seismicity of the United States 1568–1989 (Revised) U.S.G.S. Professional Paper 1527 (Vol. 1989).Google Scholar
Suarez-Rivera, R. & Fjær, E. (2013). Evaluating the poroelastic effect on anisotropic, organic-rich, mudstone systems. In Rock Mechanics and Rock Engineering (Vol. 46, pp. 569580). https://doi.org/10.1007/s00603-013–0374-yGoogle Scholar
Suarez-Rivera, R., Burghardt, J., Edelman, E., Stanchits, S., & Surdi, A. (2013). Geomechanics Considerations for Hydraulic Fracture Productivity. 47th US Rock Mechanics/Geomechanics Symposium. Retrieved from https://www.onepetro.org/conference-paper/ARMA-2013–666Google Scholar
Suter, M. (1991). State of stress and active deformation in Mexico and western Central America. In Neotectonics of North America (Vol. 1, pp. 401421). Geological Society of America Decade Map.Google Scholar
Takahashi, M., Mizoguchi, K., Kitamura, K., & Masuda, K. (2007). Effects of clay content on the frictional strength and fluid transport property of faults, Journal of Geophysical Research, 112 (March). https://doi.org/10.1029/2006JB004678Google Scholar
Tarasov, B. & Potvin, Y. (2013). Universal criteria for rock brittleness estimation under triaxial compression. International Journal of Rock Mechanics and Mining Sciences, 59, 5769. https://doi.org/10.1016/j.ijrmms.2012.12.011Google Scholar
Tembe, S., Lockner, D. A., & Wong, T.-F. (2010). Effect of clay content and mineralogy on frictional sliding behavior of simulated gouges: Binary and ternary mixtures of quartz, illite, and montmorillonite. Journal of Geophysical Research, 115(B3), 122. https://doi.org/10.1029/2009JB006383Google Scholar
Terzaghi, K. V. (1923). Die Berechnung der Durchassigkeitsziffer des Tones aus dem Verlauf der hydrodynamischen Spannungserscheinungen. Sitzungsber. Akad. Wiss. Math. Naturwiss, 132(105).Google Scholar
Teufel, L. W. & Logan, J. M. (1978). Effect of displacement rate on the real area of contact and temperatures generated during frictional sliding of Tennessee sandstone. Pure and Applied Geophysics, 116(4–5), 840865. https://doi.org/10.1007/BF00876541Google Scholar
The Academy of Medicine, Engineering and Science of Texas (2017). Environmental and Community Impacts of Shale Development in Texas. The Academy of Medicine, Engineering and Science of Texas. https://doi.org/10.25238/TAMESTstf.6.2017Google Scholar
Thomsen, L. (1986). Weak elastic anisotropy. Geophysics, 51, 19541966.Google Scholar
Tien, C. (1994). Adsorption Calculations and Modeling. Butterworth-Heinemann.Google Scholar
Tinni, A., Fathi, E., Agarwal, R., Sondergeld, C., Akkutlu, Y., & Rai, C. (2012). Shale permeability measurements on plugs and crushed samples. In SPE Canada Unconventional Resources Conference (pp. 1–14). https://doi.org/10.2118/162235-MSGoogle Scholar
Tinni, A., Sondergeld, C., & Rai, C. (2017). Pore Connectivity Between Different Wettability Systems in Organic-Rich Shales. SPE Reservoir Evaluation & Engineering, 20(04), 10201027. https://doi.org/10.2118/185948-PAGoogle Scholar
Todd, T. & Simmons, G. (1972). Effect of pore pressure on the velocity of compressional waves in low-porosity rocks. Journal of Geophysical Research, 77(20), 37313743. https://doi.org/10.1029/JB077i020p03731Google Scholar
Torsch, W. C. (2012). Thermal and pore pressure history of the Haynesville Shale in north Louisiana: a numerical study of hydrocarbon generation, overpressure, and natural hydraulic fractures. Master’s Thesis, Louisiana State University.Google Scholar
Townend, J. & Zoback, M. D. (2000). How faulting keeps the crust strong. Geology, 28(5), 399. https://doi.org/10.1130/0091–7613(2000)28<399:HFKTCS>2.0.CO;2Google Scholar
Ugueto, G., Huckabee, P., Reynolds, A., Somanchi, K., Wojtaszek, M., Nasse, D., … Ellis, D. (2018). SPE-189842-MS Hydraulic Fracture Placement Assessment in a Fiber Optic Compatible Coiled Tubing Activated Cemented Single Point Entry System.Google Scholar
Ulmer-Scholle, D. S., Scholle, P. A., Schieber, J., & Raine, R. J. (2015). A Color Guide to the Petrography of Sandstones, Siltstones, Shales and Associated Rocks. American Association of Petroleum Geologists. https://doi.org/10.1306/M1091304Google Scholar
US Energy Information Agency (2013). Annual Energy Outlook 2013. https://doi.org/DOE/EIA-0383(2015)Google Scholar
US Energy Information Agency (2017). Annual Energy Outlook 2017. Retrieved from http://www.eia.gov/outlooks/aeo/pdf/0383(2017).pdfGoogle Scholar
Vanorio, T., Prasad, M., & Nur, A. (2003). Elastic properties of dry clay mineral aggregates, suspensions and sandstones. Geophysical Journal International, 155(1), 319326. https://doi.org/10.1046/j.1365-246X.2003.02046.xGoogle Scholar
Vanorio, T., Scotellaro, C., & Mavko, G. (2008). The effect of chemical and physical processes on the acoustic properties of carbonate rocks. The Leading Edge, 27(8), 10401048. https://doi.org/10.1190/1.2967558Google Scholar
Vavryčuk, V. (2007). On the retrieval of moment tensors from borehole data. Geophysical Prospecting, 55(3), 381391. https://doi.org/10.1111/j.1365–2478.2007.00624.xGoogle Scholar
Vavryčuk, V. (2014). Iterative joint inversion for stress and fault orientations from focal mechanisms. Geophysical Journal International, 199(1), 6977. https://doi.org/10.1093/gji/ggu224Google Scholar
Vavrycuk, V., Bohnhoff, M., Jechumtálová, Z., Kolár, P., & Šileny, J. (2008). Non-double-couple mechanisms of microearthquakes induced during the 2000 injection experiment at the KTB site, Germany: A result of tensile faulting or anisotropy of a rock? Tectonophysics, 456, 7493. https://doi.org/10.1016/j.tecto.2007.08.019Google Scholar
Veatch, R. W. J. (1983). Overview of current hydraulic fracturing design and treatment technology – Part1. Journal of Petroleum Technology, 35(4), 677687. https://doi.org/10.2118/10039-PAGoogle Scholar
Vega, B., Dutta, A., & Kovscek, A. R. (2014). CT imaging of low-permeability, dual-porosity systems using high X-ray contrast gas. Transport in Porous Media, 101(1), 8197. https://doi.org/10.1007/s11242-013–0232-0Google Scholar
Vega, B., Ross, C. M., & Kovscek, A. R. (2015). Imaging-based characterization of calcite-filled fractures and porosity in shales. SPE Journal. https://doi.org/10.2118/2014–1922521-paGoogle Scholar
Verberne, B. A., Spiers, C. J., Niemeijer, A. R., De Bresser, J. H. P., De Winter, D. A. M., & Plümper, O. (2013). Frictional properties and microstructure of calcite-rich fault gouges sheared at sub-seismic sliding velocities. Pure and Applied Geophysics, 171(10), 26172640. https://doi.org/10.1007/s00024-013–0760-0Google Scholar
Vermylen, J. P. (2011). Geomechanical Studies of the Barnett Shale, Texas, USA. Stanford University.Google Scholar
Vermylen, J. P. & Zoback, M. D. (2011). Hydraulic Fracturing, Microseismic magnitudes, and Stress Evolution in the Barnett Shale, Texas, USA. In Society of Petroleum Engineers – SPE Hydraulic Fracturing Technology Conference 2011.Google Scholar
Vernik, L. (1994). Hydrocarbon-generation-induced microcracking of source rocks. Geophysics, 59(4), 555. https://doi.org/10.1190/1.1443616Google Scholar
Vernik, L. & Nur, A. (1992). Ultrasonic velocity and anisotropy of hydrocarbon source rocks. Geophysics, 57(5), 727735. https://doi.org/10.1190/1.1443286Google Scholar
Vernik, L. & Liu, X. (1997). Velocity anisotropy in shales: A petrophysical study. Geophysics, 62(2), 521. https://doi.org/10.1190/1.1444162Google Scholar
Waldhauser, F. & Ellsworth, W. L. (2000). A double-difference earthquake location algorithm: Method and application to the Northern Hayward Fault, California. Bulletin of the Seismological Society of America, 90(6), 13531368. https://doi.org/10.1785/0120000006Google Scholar
Wallace, R. E. (1951). Geometry of shearing stress and relation to faulting. Journal of Geology, 59(2), 118130. https://doi.org/10.1086/625831Google Scholar
Walls, J. & Nur, A. (1979). Pore Pressure and Confining Pressure Dependence of Permeability in Sandstone. In 7th Formation Evaluation Symposium. Calgary, Canada: Canadian Well Logging Society.Google Scholar
Walsh, F. R. (2017). Seismotectonics of the central United States and probabilistic assessment of injection induced earthquakes. Stanford University.Google Scholar
Walsh, F. R. & Zoback, M. D. (2015). Oklahoma’s recent earthquakes and saltwater disposal. Science Advances, 1(5). https://doi.org/10.1126/sciadv.1500195Google Scholar
Walsh, F. R. & Zoback, M. D. (2016). Probabilistic assessment of potential fault slip related to injectioninduced earthquakes: Application to north-central Oklahoma, USA. Geology, 44(12), 991994. https://doi.org/10.1130/G38275.1Google Scholar
Walsh, F. R., Zoback, M. D., Pais, D., Weingarten, M., & Tyrrell, T. (2017). FSP1.0: A program for probabilistic estimation of fault slip potential resulting from fluid injection.Google Scholar
Walsh, J. B. (1981). Effect of pore pressure and confining pressure on fracture permeability. International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts, 18, 429435.Google Scholar
Walters, R. J., Zoback, M. D., Baker, J. W., & Beroza, G. C. (2015). Characterizing and responding to seismic risk associated with earthquakes potentially triggered by fluid disposal and hydraulic fracturing. Seismological Research Letters, 86(4). https://doi.org/10.1785/0220150048Google Scholar
Walton, I. & McLennan, J. (2013). The role of natural fractures in shale gas production. In Effective and Sustainable Hydraulic Fracturing. https://doi.org/10.5772/56404Google Scholar
Wang, F. P. & Gale, J. F. W. (2009). Screening criteria for shale-gas systems. Gulf Coast Association of Geological Societies Transactions, 59, 779793. Retrieved from http://archives.datapages.com/data/gcags_pdf/2009/WangGale.pdfGoogle Scholar
Wang, F. P. & Reed, R. M. (2009). Pore Networks and Fluid Flow in Gas Shales, SPE 124253. In Proceedings of the SPE Annual Technical Conference and Exhibition. New Orleans, LA: Society of Petroleum Engineers.Google Scholar
Wang, H. & Sharma, M. M. (2018). Estimating un-propped fracture conductivity and fracture compliance from diagnostic fracture injection tests. SPE Journal, 23(05). Retrieved from http://arxiv.org/abs/1802.05112Google Scholar
Wang, Z. (2002a). Seismic anisotropy in sedimentary rocks, part 1: A single-plug laboratory method. Geophysics, 67(5), 1415. https://doi.org/10.1190/1.1512787Google Scholar
Wang, Z. (2002b). Seismic anisotropy in sedimentary rocks, part 2: A single-plug laboratory method. Geophysics, 67(5), 1415. https://doi.org/10.1190/1.1512787Google Scholar
Warpinski, N. R. & Branagan, P. T. (1989). Altered-stress fracturing. Journal of Petroleum Technology, SPE-17533-PA, 41(09), 990997. https://doi.org/10.2118/17533-PAGoogle Scholar
Warpinski, N. R. & Teufel, L. W. (1986). A Viscoelastic Constitutive Model for Determining In-Situ Stress Magnitudes from Anelastic Strain Recovery of Core. In SPE 61st ATCE New Orleans.Google Scholar
Warpinski, N. R., Mayerhofer, M. J., & Agarwal, K. (2013). Hydraulic fracture geomechanics and microseismic source mechanisms. SPE Journal, 18(4), 766780. https://doi.org/10.2118/158935-PAGoogle Scholar
Warpinski, N. R., Moschovidis, Z. A., Parker, C. D., & Abou-Sayed, I. S. (1994). Discussion of comparison study of hydraulic fracturing models – test case: GRI staged field experiment no. 3. SPE Production and Facilities (Society of Petroleum Engineers), 9 (SPE28158), 716.Google Scholar
Wenk, H. R., Voltolini, M., Mazurek, M., Van Loon, L. R., & Vinsot, A. (2008). Preferred orientations and anisotropy in shales: Callovo-oxfordian shale (France) and Opalinus clay (Switzerland). Clays and Clay Minerals, 56(3), 285306. https://doi.org/10.1346/CCMN.2008.0560301Google Scholar
Williams, L. B. & Hervig, R. L. (2005). Lithium and boron isotopes in illite-smectite: The importance of crystal size. Geochimica et Cosmochimica Acta, 69(24), 57055716. https://doi.org/10.1016/j.gca.2005.08.005Google Scholar
Wu, C.-H. & Sharma, M. M. (2016). Effect of Perforation Geometry and Orientation on Proppant Placement in Perforation Clusters in a Horizontal Well. Hydraulic Fracturing Technology Conference, 179117–MS. https://doi.org/10.2118/179117-MSGoogle Scholar
Wu, K. & Olson, J. (2013). Investigation of Critical In Situ and Injection Factors in Multi-Frac Treatments: Guidelines for Controlling Fracture Complexity. Proceedings of 2013 SPE Hydraulic Fracturing Technology Conference, (2000). https://doi.org/10.2118/163821-MSGoogle Scholar
Wu, W., Reece, J. S., Gensterblum, Y., & Zoback, M. D. (2017). Permeability evolution of slowly slipping faults in shale reservoirs. Geophysical Research Letters, 18. https://doi.org/10.1002/2017GL075506Google Scholar
Xu, S., Rassouli, F. S., & Zoback, M. D. (2017). Utilizing a Viscoplastic Stress Relaxation Model to Study Vertical Hydraulic Fracture Propagation in Permian Basin. Unconventional Resources Technology Conference. https://doi.org/10.15530/urtec-2017–2669793Google Scholar
Yang, F., Ning, Z., Zhang, R., Zhao, H., & Krooss, B. M. (2015). Investigations on the methane sorption capacity of marine shales from Sichuan Basin, China. International Journal of Coal Geology, 146, 104117. https://doi.org/10.1016/j.coal.2015.05.009Google Scholar
Yang, F., Xie, C., Ning, Z., & Krooss, B. M. (2016). High-pressure methane sorption on dry and moisture-equilibrated shales. Energy and Fuels, 31(1), 482492. https://doi.org/10.1021/acs.energyfuels.6b02999Google Scholar
Yang, Y. & Aplin, A. C. (2007). Permeability and petrophysical properties of 30 natural mudstones. Journal of Geophysical Research: Solid Earth, 112(3). https://doi.org/10.1029/2005JB004243Google Scholar
Yang, Y. & Mavko, G. (2018). Mathematical modeling of microcrack growth in source rock during kerogen thermal maturation. AAPG Bulletin. https://doi.org/10.1306/05111817062Google Scholar
Yang, Y. & Zoback, M. D. (2014). The role of preexisting fractures and faults during multistage hydraulic fracturing in the Bakken Formation. Interpretation, 2(3). https://doi.org/10.1190/INT-2013–0158.1Google Scholar
Yang, Y. & Zoback, M. D. (2016). Viscoplastic Deformation of the Bakken and Adjacent Formations and its Relation to Hydraulic Fracture Growth. Rock Mechanics and Rock Engineering, 49(2), 689698. https://doi.org/10.1007/s00603-015-0866-zGoogle Scholar
Yang, Y., Sone, H., Hows, A., & Zoback, M. D. (2013). Comparison of Brittleness Indices in Organic-Rich Shale Formations. In 47th US Rock Mechanics / Geomechanics Symposium (Vol. 13, pp. 13981404). Retrieved from http://www.scopus.com/inward/record.url?eid=2-s2.0–84892858564&partnerID=40&md5=f455e5d5e054a4c298d30166fdc0d351Google Scholar
Ye, Z., Janis, M., Ghassemi, A., & Riley, S. (2017). Laboratory Investigation of Fluid Flow and Permeability Evolution through Shale Fractures, 2426. https://doi.org/10.15530/urtec-2017–2674846Google Scholar
Yoon, C. E., Huang, Y., Ellsworth, W. L., & Beroza, G. C. (2017). Seismicity during the initial stages of the Guy-Greenbrier, Arkansas, earthquake sequence. Journal of Geophysical Research: Solid Earth, 122(11), 92539274. https://doi.org/10.1002/2017JB014946Google Scholar
Yu, W. & Sepehrnoori, K. (2014). Optimization of Well Spacing for Bakken Tight Oil Reservoirs. Proceedings of the 2nd Unconventional Resources Technology Conference, (2013), 1981–1996. https://doi.org/10.15530/urtec-2014–1922108Google Scholar
Zecevic, M., Daniel, G., & Jurick, D. (2016). On the nature of long-period long-duration seismic events detected during hydraulic fracturing. Geophysics, 81(3), KS113KS121. https://doi.org/10.1190/geo2015-0524.1Google Scholar
Zhang, D. & Yang, T. (2015). Environmental impacts of hydraulic fracturing in shale gas development in the United States. Petroleum Exploration and Development, 42(6), 876883. https://doi.org/10.1016/S1876-3804(15)30085–9Google Scholar
Zhang, D., Ranjith, P. G., & Perera, M. S. A. (2016). The brittleness indices used in rock mechanics and their application in shale hydraulic fracturing: A review. Journal of Petroleum Science and Engineering, 143(February), 158170. https://doi.org/10.1016/j.petrol.2016.02.011Google Scholar
Zhang, J., Kamenov, A., Zhu, D., & Hill, A. D. (2014). Laboratory Measurement of Hydraulic Fracture Conductivities in the Barnett Shale. SPE Hydraulic Fracturing Technology Conference.Google Scholar
Zhang, J., Ouyang, L., Zhu, D., & Hill, A. D. (2015). Experimental and numerical studies of reduced fracture conductivity due to proppant embedment in the shale reservoir. Journal of Petroleum Science and Engineering, 130, 3745. https://doi.org/10.1016/j.petrol.2015.04.004Google Scholar
Zhang, X. & Sanderson, D. J. (1995). Anisotropic features of geometry and permeability in fractured rock masses. Engineering Geology, 40(1–2), 6575. https://doi.org/10.1016/0013–7952(95)00040–2Google Scholar
Zhang, Y., Mostaghimi, P., Fogdon, A., Arena, A., Middleton, J., & Armstrong, R. T. (2017). Determination of Local Diffusion Coefficients and Directional Anisotropy in Shale From Dynamic Micro-CT Imaging and Microscopy. In Proceedings of the 5th Unconventional Resources Technology Conference (pp. 113). https://doi.org/10.15530/urtec-2017–2695407Google Scholar
Zhang, Y., Person, M., Rupp, J., Ellett, K., Celia, M. A., Gable, C. W., … Elliot, T. (2013). Hydrogeologic controls on induced seismicity in crystalline basement rocks due to fluid injection into basal reservoirs. Groundwater, 51(4), 525538. https://doi.org/10.1111/gwat.12071Google Scholar
Zhang, Z. X. (2002). An empirical relation between mode I fracture toughness and the tensile strength of rock. International Journal of Rock Mechanics and Mining Sciences, 39(3), 401406. https://doi.org/10.1016/S1365-1609(02)00032–1Google Scholar
Zhu, H. & Tomson, R. (2013). Exploring Water Treatment, Reuse and Alternative Sources in Shale Production. Retrieved from http://shaleplay.loewy.com/2013/11/exploring-water-treatment-reuse-and-alternative-sources-in-shale-production/Google Scholar
Ziarani, A. S. & Aguilera, R. (2012). Knudsen’s permeability correction for tight porous media. Transport in Porous Media, 91(1), 239260. https://doi.org/10.1007/s11242-011–9842-6Google Scholar
Zoback, M. D. (2007). Reservoir Geomechanics. Cambridge University. https://doi.org/10.1017/CBO9780511586477Google Scholar
Zoback, M. D. (2012). Managing the seismic risk posed by wastewater disposal. Earth, (April).Google Scholar
Zoback, M. D. & Arent, D. J. (2014). The opportunities and challenges of sustainable shale gas development. Elements, 10(4). https://doi.org/10.2113/gselements.10.4.251Google Scholar
Zoback, M. D. & Byerlee, J. D. (1975). The effect of cyclic differential stress on dilatancy in Westernly granite under uniaxial and triaxial conditions. Journal of Geophysical Research, 80, 15261530.Google Scholar
Zoback, M. D. & Harjes, H. P. (1997). Injection induced earthquakes and crustal stress at 9 km depth at the KTB deep drilling site, Germany. Journal of Geophysical Research, 102, 1847718491.Google Scholar
Zoback, M. D. & Lund Snee, J.-E. (2018). Predicted and observed shear on pre-existing faults during hydraulic fracture stimulation. In SEG Technical Program Expanded Abstracts. Society of Exploration Geophysicists.Google Scholar
Zoback, M. D. & Townend, J. (2001). Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere. Tectonophysics, 336, 1930.Google Scholar
Zoback, M. D. & Zoback, M. L. (1991). Tectonic stress field of North America and relative plate motions. In Slemmons, D. B., Engdahl, E. R., Zoback, M. D., & Blackwell, D. D. (eds.), Neotectonics of North America. Geol. Soc. Amer., Decade Map (Vol. 1). Boulder, Co.: Geological Society of America.Google Scholar
Zoback, M. D., Mastin, L., & Barton, C. (1987). In situ stress measurements in deep boreholes using hydraulic fracturing, wellbore breakouts and Stonely wave polarization. In Stefansson, O. (ed.), Rock Stress and Rock Stress Measurements, (pp. 289299). Stockholm, Sweden: Centrek Publ., Lulea.Google Scholar
Zoback, M. D., Kohli, A., Das, I., & McClure, M. (2012). The Importance of Slow Slip on Faults During Hydraulic Fracturing Stimulation of Shale Gas Reservoirs. In Society of Petroleum Engineers – SPE Americas Unconventional Resources Conference 2012.Google Scholar
Zoback, M. D., Townend, J., & Grollimund, B. (2002). Steady-state failure equilibrium and deformation of intraplate lithosphere. International Geology Review, 44(5). https://doi.org/10.2747/0020–6814.44.5.383Google Scholar
Zoback, M. L. (1992a). First and second order patterns of tectonic stress: The World Stress Map Project. Journal of Geophysical Research, 97, 11, 703711, 728.Google Scholar
Zoback, M. L. (1992b). Stress field constraints on intraplate seismicity in eastern North America. Journal of Geophysical Research, 97(B8), 1176111782.Google Scholar
Zoback, M. L. & Zoback, M. D. (1980). State of stress in the conterminous United States. Journal of Geophysical. Research, 85, 61136156.Google Scholar
Zoback, M. L. & Zoback, M. D. (1989). Tectonic stress field of the continental U.S. in geophysical framework of the continental United States. GSA Memoir, 172, 523539.Google Scholar
Zuo, J. Y., Guo, X., Liu, Y., Pan, S., Canas, J., & Mullins, O. C. (2018). Impact of capillary pressure and nanopore confinement on phase behaviors of shale gas and oil. Energy and Fuels, 32(4), 47054714. https://doi.org/10.1021/acs.energyfuels.7b03975.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
×