Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-07T20:21:44.104Z Has data issue: false hasContentIssue false

Part II - Environmental Analysis

Published online by Cambridge University Press:  28 July 2022

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

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2022

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

Adgate, JL, Goldstein, BD, and McKenzie, LM. (2014). Potential public health hazards, exposures amd health effects from unconventional natural gas development. Environmental Science & Technology. 48: 83078320.Google Scholar
Ahmadi, M and John, K. (2015). Statistical evaluation of the impact of shale gas activities on ozone pollution in North Texas. Science of the Total Environment. 536: 457467. https://doi.org/http://dx.doi.org/10.1016/j.scitotenv.2015.06.114CrossRefGoogle ScholarPubMed
Allen, DT. (2016). Emissions from oil and gas operations in the United States and their air quality implications. Journal of the Air Waste Management Association. 66: 549575. https://doi.org/10.1080/10962247.2016.1171263Google Scholar
Allen, DT, Pacsi, AP, Sullivan, DW, Zavala-Araiza, D, Harrison, M, Keen, K, Fraser, MP, Daniel Hill, A, Sawyer, RF, and Seinfeld, JH. (2014a). Methane emissions from process equipment at natural gas production sites in the United States: Pneumatic controllers. Environmental Science & Technology. 49: 633640. https://doi.org/10.1021/es5040156CrossRefGoogle ScholarPubMed
Allen, DT, Sullivan, DW, Zavala-Araiza, D, Pacsi, AP, Harrison, M, Keen, K, Fraser, MP, Daniel Hill, A, Lamb, BK, Sawyer, RF, and Seinfeld, JH. (2014b). Methane emissions from process equipment at natural gas production sites in the United States: Liquid unloadings. Environmental Science & Technology. 49: 641648. https://doi.org/10.1021/es504016rCrossRefGoogle ScholarPubMed
Allen, DT, Torres, VM, Thomas, J, Sullivan, DW, Harrison, M, Hendler, A, Herndon, SC, Kolb, CE, Fraser, MP, Hill, AD, Lamb, BK, Miskimins, J, Sawyer, RF, and Seinfeld, JH. (2013). Measurements of methane emissions at natural gas production sites in the United States. Proceedings of the National Academy of Sciences. 110: 1776817773. https://doi.org/10.1073/pnas.1304880110CrossRefGoogle ScholarPubMed
Allshouse, WB, McKenzie, LM, Barton, K, Brindley, S, and Adgate, JL. (2019). Community noise and air pollution exposure during the development of a multi-well oil and gas pad. Environmental Science & Technology. 53: 71267135. https://doi.org/10.1021/acs.est.9b00052Google Scholar
Alvarez, RA, Zavala-Araiza, D, Lyon, DR, Allen, DT, Barkley, ZR, Brandt, AR, Davis, KJ, Herndon, SC, Jacob, DJ, Karion, A, Kort, EA, Lamb, BK, Lauvaux, T, Maasakkers, JD, Marchese, AJ, Omara, M, Pacala, SW, Peischl, J, Robinson, AL, Shepson, PB, Sweeney, C, Townsend-Small, A, Wofsy, SC, and Hamburg, SP. (2018). Assessment of methane emissions from the U.S. oil and gas supply chain. Science. 361: 186188. https://doi.org/10.1126/science.aar7204Google Scholar
Anenberg, SC, Henze, DK, Tinney, V, Kinney, PL, Raich, W, Fann, N, Malley, CS, Roman, H, Lamsal, L, Duncan, B, Martin, RV, Van Donkelaar, A, Brauer, M, Doherty, R, Jonson, JE, Davila, Y, Sudo, K, and Kuylenstierna, JCI. (2018). Estimates of the global burden of ambient PM 2.5, Ozone, and NO2 on asthma incidence and emergency room visits. Environmental Health Perspectives. 126. https://doi.org/10.1289/EHP3766Google Scholar
Baker, AK, Beyersdorf, AJ, Doezema, LA, Katzenstein, A, Meinardi, S, Simpson, IJ, Blake, DR, and Sherwood Rowland, F. (2008). Measurements of nonmethane hydrocarbons in 28 United States cities. Atmospheric Environment. 42: 170182. https://doi.org/10.1016/j.atmosenv.2007.09.007Google Scholar
Bean, JK, Bhandari, S, Bilotto, A, and Hildebrandt Ruiz, L. (2018). Formation of particulate matter from the oxidation of evaporated hydraulic fracturing wastewater. Environmental Science & Technology. 52: 49604968. https://doi.org/10.1021/acs.est.7b06009Google Scholar
Bishop, GA, Schuchmann, BG, and Stedman, DH. (2013). Heavy-duty truck emissions in the South Coast Air Basin of California. Environmental Science & Technology. 47: 95239529. https://doi.org/10.1021/es401487bGoogle Scholar
Bolden, AL, Kwiatkowski, CF, and Colborn, T. (2015). New look at BTEX: Are ambient levels a problem? Environmental Science & Technology. 49: 52615276. https://doi.org/10.1021/es505316fGoogle Scholar
Brandt, AR, Heath, GA, Kort, EA, O’Sullivan, F, Petron, G, Jordaan, SM, Tans, P, Wilcox, J, Gopstein, AM, Arent, D, Wofsy, S, Brown, NJ, Bradley, R, Stucky, GD, Eardley, D, and Harriss, RC. (2014). Methane leaks from North American natural gas systems. Science. 343: 733735. https://doi.org/10.1126/science.1247045Google Scholar
Brauer, M, Amann, M, Burnett, RT, Cohen, A, Dentener, F, Ezzati, M, Henderson, SB, Krzyzanowski, M, Martin, RV, Van Dingenen, R, van Donkelaar, A, and Thurston, GD. (2012). Exposure assessment for estimation of the global burden of disease attributable to outdoor air pollution. Environmental Science & Technology. 46: 652660. https://doi.org/10.1021/es2025752Google Scholar
Brauer, M, Freedman, G, Frostad, J, van Donkelaar, A, Martin, RV, Dentener, F, Dingenen, R. van, Estep, K, Amini, H, Apte, JS, Balakrishnan, K, Barregard, L, Broday, D, Feigin, V, Ghosh, S, Hopke, PK, Knibbs, LD, Kokubo, Y, Liu, Y, Ma, S, Morawska, L, Sangrador, JLT, Shaddick, G, Anderson, HR, Vos, T, Forouzanfar, MH, Burnett, RT, and Cohen, A. (2016). Ambient air pollution exposure estimation for the global burden of disease 2013. Environmental Science & Technology. 50: 7988. https://doi.org/10.1021/acs.est.5b03709Google Scholar
Bunch, AG, Perry, CS, Abraham, L, Wikoff, DS, Tachovsky, JA, Hixon, JG, Urban, JD, Harris, MA, and Haws, LC. (2014). Evaluation of impact of shale gas operations in the Barnett Shale region on volatile organic compounds in air and potential human health risks. Science Of the Total Environment. 468–469: 832842. https://doi.org/10.1016/J.SCITOTENV.2013.08.080Google Scholar
Burnett, RT, Pope III, CA, Ezzati, M, Olives, C, Lim, SS, Mehta, S, Shin, HH, Singh, G, Hubbell, B, Brauer, M, Anderson, HR, Smith, KR, Balmes, JR, Bruce, N, Kan, H, Laden, F, Pruss-Ustun, A, Turner, MC, Gapstur, SM, Diver, WR, and Cohen, A. (2014). An integrated risk function for estimating the global burden of disease attributable to ambient fine particulate matter exposure. Environmental Health Perspectives. 122: 397403. https://doi.org/http://dx.doi.org/10.1289/ehp.1307049Google Scholar
Carlton, AG, Little, E, Moeller, M, Odoyo, S, and Shepson, PB. (2014). The data gap: Can a lack of monitors obscure loss of Clean Air Act benefits in fracking areas? Environmental Science & Technology. 48: 893894. https://doi.org/dx.doi.org/10.1021/es405672tGoogle Scholar
Carter, WPL and Seinfeld, JH. (2012). Winter ozone formation and VOC incremental reactivites in the Upper Green River Basin of Wyoming. Atmospheric Environment. 50: 255266.CrossRefGoogle Scholar
Cheadle, LC, Oltmans, SJ, Petron, G, Schnell, RC, Mattson, EJ, Herndon, SC, Thompson, AM, Blake, DR, and McClure-Begley, A. (2017). Surface ozone in the Colorado northern Front Range and the influence of oil and gas development during FRAPPE/DISCOVER-AQ in summer 2014. Elementa: Science of the Anthropocene. 5: 61. https://doi.org/10.1525/elementa.254Google Scholar
Clark, NA, Demers, PA, Karr, CJ, Koehoorn, M, Lencar, C, Tamburic, L, Brauer, M. (2010). Effect of early life exposure to air pollution on development of childhood asthma. Environmental Health Perspectives. 118: 284290.CrossRefGoogle ScholarPubMed
Correia, AW, Pope III, CA, Dockery, DW, Wang, Y, Ezzati, M, and Dominici, F. (2013). Effect of air pollution control on life expectancy in the United States: An analysis of 545 U.S. counties for the period from 2000 to 2007. Epidemiology, 24: 2331. https://doi.org/10.1097/EDE.0b013e3182770237CrossRefGoogle ScholarPubMed
Dallmann, TR, DeMartini, SJ, Kirchstetter, TW, Herndon, SC, Onasch, TB, Wood, EC, and Harley, RA. (2012). On-road measurements of gas and particle phase pollutant emission factors for individual heavy-duty diesel trucks. Environmental Science & Technology. 46: 85118518. https://doi.org/10.1021/es301936cGoogle Scholar
Di, Q, Wang, Yan, Zanobetti, A, Wang, Yun, Koutrakis, P, Choirat, C, Dominici, F, and Schwartz, JD. (2017). Air pollution and mortality in the Medicare population. New England Journal of Medicine. 376: 25132522. https://doi.org/10.1056/NEJMoa1702747Google Scholar
Edie, R, Robertson, AM, Soltis, J, Field, RA, Snare, D, Burkhart, MD, and Murphy, SM. (2020). Off-site flux estimates of volatile organic compounds from oil and gas production facilities using fast-response instrumentation. Environmental Science & Technology. 54: 13851394. https://doi.org/10.1021/acs.est.9b05621Google Scholar
Edwards, PM, Young, CJ, Aikin, K, DeGouw, JA, Dube, WP, Geiger, F, Gilman, JB, Helmig, D, Holloway, JS, Kercher, J, Lerner, B, Martin, R, McLaren, R, Parrish, DD, Peischl, J, Roberts, JM, Ryerson, TB, Thornton, J, Warneke, C, Williams, EJ, and Brown, SS. (2013). Ozone photochemistry in an oil and natural gas extraction region during winter: Simulations of a snow-free season in the Uintah Basin, Utah. Atmospheric Chemistry and Physics. 13: 89558971. https://doi.org/10.5194/acp-13-8955-2013CrossRefGoogle Scholar
Edwards, PM, Brown, SS, Roberts, JM, Ahmadov, R, Banta, RM, de Gouw, JA, Dube, WP, Field, RA, Flynn, JH, Gilman, JB, Graus, M, Helmig, D, Koss, A, Langford, AO, Lefer, BL, Lerner, BM, Li, R, Li, S-M, McKeen, SA, Murphy, SM, Parrish, DD, Senff, CJ, Soltis, J, Stutz, J, Sweeney, C, Thompson, CR, Trainer, MK, Tsai, C, Veres, P, Washenfelder, RA, Warneke, C, Wild, RJ, Young, CJ, Yuan, B, and Zamora, R. (2014). High winter ozone production from carbonyl photolysis in an oil and gas basin. Nature. 514: 351354. https://doi.org/10.1038/nature13767Google Scholar
Fann, N, Baker, KR, Chan, EAW, Eyth, A, Macpherson, A, Miller, E, and Snyder, J. (2018). Assessing human health PM 2.5 and ozone impacts from U.S. oil and natural gas sector emissions in 2025. Environmental Science & Technology. 52: 80958103. https://doi.org/10.1021/acs.est.8b02050Google Scholar
Field, RA, Soltis, J, McCarthy, MC, Murphy, S, and Montague, DC. (2015). Influence of oil and gas field operations on spatial and temporal distributions of atmospheric non-methane hydrocarbons and their effect on ozone formation in winter. Atmospheric Chemistry and Physics. 15: 35273542. https://doi.org/10.5194/acp-15-3527-2015Google Scholar
Fine, PM, Sioutas, C, and Solomon, PA. (2008). Secondary particulate matter in the United States: Insights from the particulate matter supersites program and related studies secondary particulate matter in the United States: Insights from the particulate matter supersites program and related studies. Journal of the Air Waste Management Association. 58: 234253. https://doi.org/10.3155/1047-3289.58.2.234Google Scholar
Gehring, U, Gruzieva, O, Agius, RM, Beelen, R, Custovic, A, Cyrys, J, Eeftens, M, Flexeder, C, Fuertes, E, Heinrich, J, Hoffmann, B, de Jongste, JC, Kerkhof, M, Klumper, C, Korek, M, Molter, A, Schultz, ES, Simpson, A, Sugiri, D, Svartengren, M, von Berg, A, Wijga, AH, Pershagen, G, and Brunekreef, B. (2013). Air Pollution Exposure and Lung Function in Children: The ESCAPE Project. Environmental Health Perspectives. 121: 13571364. https://doi.org/10.1289/ehp.1306770Google Scholar
Gilman, JB, Lerner, BM, Kuster, WC, and de Gouw, J. (2013). Source signature of volatile organic compounds (VOCs) from oil and natural gas operations in northeastern Colorado. Environmental Science & Technology. 47: 12971305. https://doi.org/10.1021/es304119aCrossRefGoogle ScholarPubMed
Goetz, JD, Floerchinger, C, Fortner, EC, Wormhoudt, J, Massoli, P, Knighton, WB, Herndon, SC, Kolb, CE, Knipping, E, Shaw, SL, and DeCarlo, PF. (2015). Atmospheric Emission Characterization of Marcellus Shale Natural Gas Development Sites. Environmental Science & Technology. 49: 70127020. https://doi.org/10.1021/acs.est.5b00452Google Scholar
Grahame, TJ and Schlesinger, RB. (2010). Cardiovascular health and particulate vehicular emissions: A critical evaluation of the evidence. Air Quality and Atmospheric Health. 3: 327. https://doi.org/10.1007/s11869–009-0047-xGoogle Scholar
Halliday, HS, Thompson, AM, Wisthaler, A, Blake, DR, Hornbrook, RS, Mikoviny, T, Müller, M, Eichler, P, Apel, EC, and Hills, AJ. (2016). Atmospheric benzene observations from oil and gas production in the Denver-Julesburg Basin in July and August 2014. Journal of Geophysical Research. 121: 11,05511,074. https://doi.org/10.1002/2016JD025327Google Scholar
Hanninen, O, Knol, AB, Jantunen, M, Lim, T-A, Conrad, A, Rappolder, M, Carrer, P, Fanetti, A-C, Kim, R, Buekers, J, Torfs, R, Iavarone, I, Classen, T, Hornberg, C, Mekel, OC, and Group EbW. (2014). Environmental burden of disease in Europe: Assessing Nine Risk Factors in Six Countries. Environmental Health Perspectives. 122: 439446. https://doi.org/http://dx.doi.org/10.1289/ehp.1206154CrossRefGoogle ScholarPubMed
Hecobian, A, Clements, AL, Shonkwiler, KB, Zhou, Y, MacDonald, LP, Hilliard, N, Wells, BL, Bibeau, B, Ham, JM, Pierce, JR, and Collett, JL. (2019). Air toxics and other volatile organic compound emissions from unconventional oil and gas development. Environmental Science & Technology Letters. 6: 720726. https://doi.org/10.1021/acs.estlett.9b00591Google Scholar
Helmig, D, Thompson, CR, Evans, J, Boylan, P, Hueber, J, and Park, J-H. (2014). Highly elevated atmospheric levels of volatile organic compounds in the Uintah Basin, Utah. Environmental Science and Technology. 48: 47074715. https://doi.org/dx.doi.org/10.1021/es405046rCrossRefGoogle ScholarPubMed
Hildebrandt, L, Donahue, NM, and Panids, SN. (2009). High formation of secondary organic aerosol from the photo-oxidation of toluene. Atmospheric Chemistry and Physics. 9 29732986.CrossRefGoogle Scholar
Holder, C, Hader, J, Avanasi, R, Hong, T, Carr, E, Mendez, B, Wignall, J, Glen, G, Guelden, B, and Wei, Y. (2019). Evaluating potential human health risks from modeled inhalation exposures to volatile organic compounds emitted from oil and gas operations. Journal of the Air Waste Management Association. 69: 15031524. https://doi.org/10.1080/10962247.2019.1680459Google Scholar
Jerrett, M, Burnett, RT, Beckerman, BS, Turner, MC, Krewski, D, Thurston, G, Martin, RV, van Donkelaar, A, Hughes, E, Shi, Y, Gapstur, SM, Thun, MJ, and Pope III, CA. (2013). Spatial analysis of air pollution and mortality in California. American Journal of Respiratory Critical Care Medicine. 188: 593599. https://doi.org/10.1164/rccm.201303-0609OCCrossRefGoogle ScholarPubMed
Jimenez, JL, Canagaratna, MR, Donahue, NM, Prevot, ASH, Zhang, Q, Kroll, JH, DeCarlo, PF, Allan, JD, Coe, H, Ng, NL, Aiken, AC, Docherty, KS, Ulbrich, IM, Grieshop, AP, Robinson, AL, Duplissy, J, Smith, JD, Wilson, KR, Lanz, VA, Hueglin, C, Sun, YL, Tian, J, Laaksonen, A, Raatikainen, T, Rautiainen, J, Vaattovaara, P, Ehn, M, Kulmala, M, Tomlinson, JM, Collins, DR, Cubison, MJE, Dunlea, J, Huffman, JA, Onasch, TB, Alfarra, MR, Williams, PI, Bower, K, Kondo, Y, Schneider, J, Drewnick, F, Borrmann, S, Weimer, S, Demerjian, K, Salcedo, D, Cottrell, L, Griffin, R, Takami, A, Miyoshi, T, Hatakeyama, S, Shimono, A, Sun, JY, Zhang, YM, Dzepina, K, Kimmel, JR, Sueper, D, Jayne, JT, Herndon, SC, Trimborn, AM, Williams, LR, Wood, EC, Middlebrook, AM, Kolb, CE, Baltensperger, U, and Worsnop, DR. (2009). Evolution of organic aerosols in the atmosphere. Science. 80(326): 15251529. https://doi.org/10.1126/science.1180353Google Scholar
Johnson, D, Heltzel, R, and Oliver, D. (2019). Temporal variations in methane emissions from an unconventional well site. ACS Omega. 4: 37083715. https://doi.org/10.1021/acsomega.8b03246CrossRefGoogle ScholarPubMed
Kanakidou, M, Seinfeld, JH, Pandis, SN, Barnes, I, Dentener, FJ, Facchini, MC, Van Dingenen, R, Ervens, B, Nenes, A, Nielsen, CJ, Swietlicki, E, Putaud, JP, Balkanski, Y, Fuzzi, S, Horth, J, Moortgat, GK, Winterhalter, R, Myhre, CEL, Tsigaridis, K, Vignati, E, Stephanou, EG, and Wilson, J. (2005). Organic aerosol and global climate modelling: A review. Atmospheric Chemistry and Physics. 5: 10531123.Google Scholar
Koss, A, Yuan, B, Warneke, C, Gilman, JB, Lerner, BM, Veres, PR, Peischl, J, Eilerman, S, Wild, R, Brown, SS, Thompson, CR, Ryerson, T, Hanisco, T, Wolfe, GM, St Clair, JM, Thayer, M, Keutsch, FN, Murphy, S, and De Gouw, J. (2017). Observations of VOC emissions and photochemical products over US oil- and gas-producing regions using high-resolution H3O+ CIMS (PTR-ToF-MS). Atmospheric Measurement Techniques. 10: 29412968. https://doi.org/10.5194/amt-10-2941-2017CrossRefGoogle Scholar
Kroepsch, AC, Maniloff, PT, Adgate, JL, McKenzie, LM, and Dickinson, KL. (2019). Environmental Justice in Unconventional Oil and Natural Gas Drilling and Production: A Critical Review and Research Agenda. Environmental Science & Technology. 53: 66016615. https://doi.org/10.1021/acs.est.9b00209Google Scholar
Lefler, JS, Higbee, JD, Burnett, RT, Ezzati, M, Coleman, NC, Mann, DD, Marshall, JD, Bechle, M, Wang, Y, Robinson, AL, and Arden Pope, C. (2019). Air pollution and mortality in a large, representative U.S. cohort: Multiple-pollutant analyses, and spatial and temporal decompositions. Environmental Health. 18: 101. https://doi.org/10.1186/s12940–019-0544-9Google Scholar
Li, SM, Leithead, A, Moussa, SG, Liggio, J, Moran, MD, Wang, D, Hayden, K, Darlington, A, Gordon, M, Staebler, R, Makar, PA, Stroud, CA, McLaren, R, Liu, PSK, O’Brien, J, Mittermeier, RL, Zhang, J, Marson, G, Cober, SG, Wolde, M, and Wentzell, JJB. (2017). Differences between measured and reported volatile organic compound emissions from oil sands facilities in Alberta, Canada. Proceedings of the National Academy of Sciences U.S.A. 114: E3756E3765. https://doi.org/10.1073/pnas.1617862114Google Scholar
Lyon, DR, Alvarez, RA, Zavala-Araiza, D, Brandt, AR, Jackson, RB, and Hamburg, SP. (2016). Aerial surveys of elevated hydrocarbon emissions from oil and gas production sites. Environmental Science & Technology. 50: 48774886. https://doi.org/10.1021/acs.est.6b00705Google Scholar
Marchese, AJ, Vaughn, TL, Zimmerle, DJ, Martinez, DM, Williams, LL, Robinson, AL, Mitchell, AL, Subramanian, R, Tkacik, DS, Roscioli, JR, and Herndon, SC. (2015). Methane emissions from United States natural gas gathering and processing. Environmental Science & Technology. 49: 1071810727. https://doi.org/10.1021/acs.est.5b02275CrossRefGoogle ScholarPubMed
Marrero, JE, Townsend-Small, A, Lyon, DR, Tsai, TR, Meinardi, S, and Blake, DR. (2016). Estimating emissions of toxic hydrocarbons from natural gas production sites in the Barnett Shale region of Northern Texas. Environmental Science & Technology. 50: 1075610764. https://doi.org/10.1021/acs.est.6b02827Google Scholar
May, AA, Nguyen, NT, Presto, AA, Gordon, TD, Lipsky, EM, Karve, M, Gutierrez, A, Robertson, WH, Zhang, M, Brandow, C, Chang, O, Chen, S, Cicero-Fernandez, P, Dinkins, L, Fuentes, M, Huang, S-M, Ling, R, Long, J, Maddox, C, Massetti, J, McCauley, E, Miguel, A, Na, K, Ong, R, Pang, Y, Rieger, P, Sax, T, Truong, T, Vo, T, Chattopadhyay, S, Maldonado, H, Maricq, MM, and Robinson, AL. (2014). Gas- and particle-phase primary emissions from in-use, on-road gasoline and diesel vehicles. Atmospheric Environment. 88: 247260. https://doi.org/10.1016/j.atmosenv.2014.01.046Google Scholar
McDuffie, EE, Edwards, PM, Gilman, JB, Lerner, BM, Dubé, WP, Trainer, M, Wolfe, DE, Angevine, WM, deGouw, J, Williams, EJ, Tevlin, AG, Murphy, JG, Fischer, EV, McKeen, S, Ryerson, TB, Peischl, J, Holloway, JS, Aikin, K, Langford, AO, Senff, CJ, Alvarez, RJ, Hall, SR, Ullmann, K, Lantz, KO, and Brown, SS. (2016). Influence of oil and gas emissions on summertime ozone in the Colorado Northern Front Range. Journal of Geophysical Research: Atmospheres. 121: 87128729. https://doi.org/10.1002/2016JD025265CrossRefGoogle Scholar
McKenzie, LM, Blair, B, Hughes, J, Allshouse, WB, Blake, NJ, Helmig, D, Milmoe, P, Halliday, H, Blake, DR, and Adgate, JL. (2018). Ambient nonmethane hydrocarbon levels along Colorado’s Northern Front Range: Acute and chronic health risks. Environmental Science & Technology. 52: 45144525. https://doi.org/10.1021/acs.est.7b05983CrossRefGoogle ScholarPubMed
Mitchell, AL, Tkacik, DS, Roscioli, JR, Herndon, SC, Yacovitch, TI, Martinez, DM, Vaughn, TL, Williams, LL, Sullivan, MR, Floerchinger, C, Omara, M, Subramanian, R, Zimmerle, D, Marchese, AJ, and Robinson, AL. (2015). Measurements of methane emissions from natural gas gathering facilities and processing plants: Measurement results. Environmental Science & Technology. 49: 32193227. https://doi.org/10.1021/es5052809Google Scholar
Moore, CW, Zielinska, B, Petron, G, and Jackson, RB. (2014). Air impacts of increased natural gas acquisition, processing, and use: A critical review. Environmental Science & Technology. 48: 83498359. https://doi.org/dx.doi.org/10.1021/es4053472CrossRefGoogle ScholarPubMed
Ng, NL, Kroll, JH, Chan, AWH, Chhabra, PS, Flagan, RC, and Seinfeld, JH. (2007). Secondary organic aerosol formation from m-xylene, toluene, and benzene. Atmospheric Chemistry and Physics. 7: 39093922.Google Scholar
Oltmans, S, Schnell, R, Johnson, B, Pétron, G, Mefford, T, and Neely, R. (2014). Anatomy of wintertime ozone associated with oil and natural gas extraction activity in Wyoming and Utah. Elementa: Science of the Anthropecene. 2: 000024. https://doi.org/10.12952/journal.elementa.000024Google Scholar
Omara, M, Sullivan, MR, Li, X, Subramian, R, Robinson, AL, and Presto, AA. (2016). Methane emissions from conventional and unconventional natural gas production sites in the Marcellus Shale Basin. Environmental Science & Technology. 50: 20992107. https://doi.org/10.1021/acs.est.5b05503CrossRefGoogle ScholarPubMed
Omara, M, Zimmerman, N, Sullivan, MR, Li, X, Ellis, A, Cesa, R, Subramanian, R, Presto, AA, and Robinson, AL. (2018). Methane emissions from natural gas production sites in the United States: Data synthesis and national estimate. Environmental Science & Technology. 52: 1291512925. https://doi.org/10.1021/acs.est.8b03535Google Scholar
Pacsi, AP, Alhajeri, NS, Zavala-Araiza, D, Webster, MD, and Allen, DT. (2013). Regional air quality impacts of increased natural gas production and use in Texas. Environmental Science & Technology. 47. https://doi.org/10.1021/es3044714CrossRefGoogle ScholarPubMed
Pacsi, AP, Kimura, Y, McGaughey, G, McDonald-Buller, EC, and Allen, DT. (2015). Regional ozone impacts of increased natural gas use in the Texas power sector and development in the Eagle Ford shale. Environmental Science & Technology. 49: 39663973. https://doi.org/DOI: 10.1021/es5055012Google Scholar
Penard-Morand, C, Schillinger, C, Armengaud, A, Debotte, G, Chretien, E, and Annesi-Maesano, I. (2006). Assessment of schoolchildren’s exposure to traffic-related air pollution in the French Six Cities using a dispersion model. Atmospheric Environment. 40: 22742287.Google Scholar
Petron, G, Frost, G, Miller, BR, Hirsch, AI, Montzka, SA, Karion, A, Trainer, M, Sweeney, C, Andrews, AE, Miller, L, Kofler, J, Bar-lian, A, Dlugokencky, EJ, Patrick, L, Moore, CT Jr., Ryerson, TB, Siso, C, Kolodzey, W, Lang, PM, Conway, T, Novelli, P, Masarie, K, Hall, B, Guenther, D, Kitzis, D, Miller, J, Welsh, D, Wolfe, D, Neff, W, and Tans, P. (2012). Hydrocarbon emissions characterization in the Colorado Front Range: A pilot study. Journal of Geophysical Research. 117: D04304. https://doi.org/10.1029/2011JD016360Google Scholar
Pétron, G, Karion, A, Sweeney, C, Miller, BR, Montzka, SA, Frost, GJ, Trainer, M, Tans, P, Andrews, A, Kofler, J, Helmig, D, Guenther, D, Dlugokencky, E, Lang, P, Newberger, T, Wolter, S, Hall, B, Novelli, P, Brewer, A, Conley, S, Hardesty, M, Banta, R, White, A, Noone, D, Wolfe, D, and Schnell, R. (2014). A new look at methane and nonmethane hydrocarbon emissions from oil and natural gas operations in the Colorado Denver-Julesburg Basin. Journal of Geophysical Research. 119: 68366852. https://doi.org/10.1002/2013JD021272Google Scholar
Pope, CA, Coleman, N, Pond, ZA, and Burnett, RT. (2019). Fine particulate air pollution and human mortality: 25+ years of cohort studies. Environmental Research. 108924. https://doi.org/10.1016/j.envres.2019.108924Google Scholar
Pope, CA III, Burnett, RT, Thun, MJ, Calle, EE, Krewski, D, Ito, K, and Thurston, GD. (2002). Lung cancer, cardiopulmonary mortality, and long-term exposure to fine particulate air pollution. Journal of the American Medical Association. 287: 11321141.Google Scholar
Pope, CA III, Burnett, RT, Thurston, GD, Thun, MJ, Calle, EE, Krewski, D, and Godleski, JJ. (2004). Cardiovascular mortality and long-term exposure to particulate air pollution: Epidemiological evidence of general pathophysiological pathways of disease. Circulation. 109: 7177.Google Scholar
Pope, CA III and Dockery, DW. (2006). Health effects of fine particulate air pollution: lines that connect. Journal of the Air Waste Management Association. 56: 709742.Google Scholar
Pope, CA III, Ezzati, M, and Dockery, DW. (2009). Fine-particulate air pollution and life expectancy in the United States. New England Journal of Medicine. 360: 376386.Google Scholar
Raaschou-Nielsen, O, Andersen, ZJ, Beelen, R, Samoli, E, Stafoggia, M, Weinmayr, G, Hoffmann, B, Fischer, P, Nieuwenhuijsen, MJ, Brunekreef, B, Xun, WW, Katsouyanni, K, Dimakopoulou, K, Sommar, J, Forsberg, B, Modig, L, Oudin, A, Oftedal, B, Schwarze, PE, Nafstad, P, De Faire, U, Pedersen, NL, Östenson, C-G, Fratiglioni, L, Penell, J, Korek, M, Pershagen, G, Eriksen, KT, Sørensen, M, Tjønneland, A, Ellermann, T, Eeftens, M, Peeters, PH, Meliefste, K, Wang, M, Bueno-de-Mesquita, B, Key, TJ, de Hoogh, K, Concin, H, Nagel, G, Vilier, A, Grioni, S, Krogh, V, Tsai, M-Y, Ricceri, F, Sacerdote, C, Galassi, C, Migliore, E, Ranzi, A, Cesaroni, G, Badaloni, C, Forastiere, F, Tamayo, I, Amiano, P, Dorronsoro, M, Trichopoulou, A, Bamia, C, Vineis, P, and Hoek, G. (2013). Air pollution and lung cancer incidence in 17 European cohorts: Prospective analyses from the European Study of Cohorts for Air Pollution Effects (ESCAPE). Lancet Oncology. 14: 813822. https://doi.org/http://dx.doi.org/10.1016/S1470–2045(13)70279-1Google Scholar
Rappenglück, B, Ackermann, L, Alvarez, S, Golovko, J, Buhr, M, Field, RA, Soltis, J, Montague, DC, Hauze, B, Adamson, S, Risch, D, Wilkerson, G, Bush, D, Stoeckenius, T, and Keslar, C. (2014). Strong wintertime ozone events in the Upper Green River basin, Wyoming. Atmospheric Chemistry and Physics. 14: 49094934. https://doi.org/10.5194/acp-14-4909-2014Google Scholar
Roohani, YH, Roy, AA, Heo, J, Robinson, AL, and Adams, PJ. (2017). Impact of natural gas development in the Marcellus and Utica shales on regional ozone and fine particulate matter levels. Atmospheric Environment. 155: 1120. https://doi.org/10.1016/j.atmosenv.2017.01.001CrossRefGoogle Scholar
Roy, AA, Adams, PJ, and Robinson, AL. (2014). Air pollutant emissions from the development, production, and processing of Marcellus Shale natural gas. Journal Of the Air Waste Management Association. 64: 1937. https://doi.org/10.1080/10962247.2013.826151Google Scholar
Schildcrout, JS, Sheppard, L, Lumley, T, Slaughter, JC, Koenig, JQ, and Shapiro, GG. (2006). Ambient air pollution and asthma exacerbations in children: An eight-city analysis. American Journal of Epidemiology. 164: 505517.CrossRefGoogle ScholarPubMed
Schnell, RC, Oltmans, SJ, Neely, RR, Endres, MS, Molenar, JV, and White, AB. (2009). Rapid photochemical production of ozone at high concentrations in a rural site during winter. Nature Geoscience. 2: 120122. https://doi.org/10.1038/NGEO415Google Scholar
Seinfeld, JH and Pandis, SN. (2006). Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 2nd ed. John Wiley & Sons.Google Scholar
Shekarrizfard, M, Valois, M-F, Weichenthal, S, Goldberg, MS, Fallah-Shorshani, M, Cavellin, LD, Crouse, D, Parent, M-E, and Hatzopoulou, M. (2018). Investigating the effects of multiple exposure measures to traffic-related air pollution on the risk of breast and prostate cancer. Journal of Transport and Health. 11: 3446. https://doi.org/10.1016/J.JTH.2018.09.006Google Scholar
Su, JG, Jerrett, M, Beckerman, B, Wilhelm, M, Ghosh, JK, and Ritz, B. (2009). Predicting traffic-related air pollution in Los Angeles using a distance decay regression selection strategy. Environmental Research. 109: 657670. https://doi.org/10.1016/j.envres.2009.06.001Google Scholar
Subramanian, R, Williams, LL, Vaughn, TL, Zimmerle, D, Roscioli, JR, Herndon, SC, Yacovitch, TI, Floerchinger, C, Tkacik, DS, Mitchell, AL, Sullivan, MR, Dallmann, TR, and Robinson, AL. (2015). Methane emissions from natural gas compressor stations in the transmission and storage sector: Measurements and comparisons with the EPA greenhouse gas reporting program protocol. Environmental Science & Technology. 49: 32523261. https://doi.org/10.1021/es5060258Google Scholar
Swarthout, RF, Russo, RS, Zhou, Y, Miller, BM, Mitchell, B, Horsman, E, Lipsky, E, McCabe, DC, Baum, E, and Sive, BC. (2015). Impact of Marcellus Shale natural gas development in Southwest Pennsylvania on volatile organic compound emissions and regional air quality. Environmental Science & Technology. 49: 31753184. https://doi.org/10.1021/es504315fGoogle Scholar
Tan, Y, Dallmann, TR, Robinson, AL, and Presto, AA. (2016). Application of plume analysis to build land use regression models from mobile sampling to improve model transferability. Atmospheric Environment. 134: 5160. https://doi.org/10.1016/j.atmosenv.2016.03.032Google Scholar
Tan, Y, Lipsky, EM, Saleh, R, Robinson, AL, and Presto, AA. (2014). Characterizing the spatial variation of air pollutants and the contributions of high emitting vehicles in Pittsburgh, PA. Environmental Science & Technology. 48: 1418614194. https://doi.org/10.1021/es5034074CrossRefGoogle ScholarPubMed
Vinciguerra, T, Yao, S, Dadzie, J, Chittams, A, Deskins, T, Ehrman, S, and Dickerson, RR. (2015). Regional air quality impacts of hydraulic fracturing and shale natural gas activity: Evidence from ambient VOC observations. Atmospheric Environment. 110: 144150. https://doi.org/http://dx.doi.org/10.1016/j.atmosenv.2015.03.056CrossRefGoogle Scholar
Warneke, C, Geiger, F, Edwards, PM, Dube, W, Petron, G, Kofler, J, Zahn, A, Brown, SS, Graus, M, Gilman, J, Lerner, B, Peischl, J, Ryerson, TB, de Gouw, JA, and Roberts, JM. (2014). Volatile organic compound emissions from the oil and natural gas industry in the Uinta Basin, Utah: Oil and gas well pad emissions compared to ambient air composition. Atmospheric Chemistry and Physics. 14: 1097710988. https://doi.org/10.5194/acp-14-10977-2014Google Scholar
Zavala-Araiza, D, Lyon, D, Alvarez, RA, Palacios, V, Harriss, R, Lan, X, Talbot, R, and Hamburg, SP. (2015). Toward a functional definition of methane super-emitters: Application to natural gas production sites. Environmental Science & Technology. 49: 81678174. https://doi.org/10.1021/acs.est.5b00133CrossRefGoogle Scholar
Zimmerle, DJ, Williams, LL, Vaughn, TL, Quinn, C, Subramanian, R, Duggan, GP, Willson, B, Opsomer, JD, Marchese, AJ, Martinez, DM, and Robinson, AL. (2015). Methane emissions from the natural gas transmission and storage system in the United States. Environmental Science & Technology. 49: 93749383. https://doi.org/10.1021/acs.est.5b01669Google Scholar

References

Alvarez, RA, Pacala, SW, Winebrake, JJ, Chameides, WL, and Hamburg, SP. (2012). Greater focus needed on methane leakage from natural gas infrastructure. Proceedings of the National Academy of Sciences. 109: 64356440, doi:10.1073/pnas.1202407109Google Scholar
Alvarez, RA, Zavalao-Araiza, D, Lyon, DR, Allen, DT, Barkley, ZR, Brandt, AR, Davis, KJ, Herndon, SC, Jacob, DJ, Karion, A, Korts, EA, Lamb, BK, Lauvaux, T, Maasakkers, JD, Marchese, AJ, Omara, M, Pacala, JW, Peischl, J, Robinson, AJ, Shepson, PB, Sweeney, C, Townsend-Small, A, Wofsy, SC, and Hamburg, SP. (2018). Assessment of methane emissions from the U.S. oil and gas supply chain. Science. 361: 186188, doi:10.1126/science.aar7204Google ScholarPubMed
Begon, M, Howarth, RW, and Townsend, C. (2014). Essentials of Ecology, 4th Edition. Wiley. ISBN-13: 978-0470909133Google Scholar
Caulton, DR, Shepson, PD, Santoro, RL, Sparks, JP, Howarth, RW, Ingraffea, A, Camaliza, MO, Sweeney, C, Karion, A, Davis, KJ, Stirm, BH, Montzka, SA, and Miller, B. (2014). Toward a better understanding and quantification of methane emissions from shale gas development. Proceedings of the National Academy of Sciences. 111: 62376242, doi:10.1073/pnas.1316546111Google Scholar
Collins, WJ, Webber, CP, Cox, PM, Huntingford, C, Lowe, J, Sitch, S, Chadburn, SE, Comyn-Platt, E, Harper, AB, Hayman, G, and Powell, T. (2018). Increased importance of methane reduction for a 1.5 degree target. Environmental Research Letters. 13: 054003, doi:10.1088/1748-9326/aab89cGoogle Scholar
Fesenfeld, LP, Schmidt, TS, and Schrode, A. (2018). Climate policy for short- and long-lived pollutants. Nature Climate Change. 8: 933936, doi:10.1038/s41558-018-0328-1Google Scholar
Hausfather, Z. (2018). Analysis: How much ‘carbon budget’ is left to limit global warming to 1.5C? Carbon Brief www.carbonbrief.org/analysis-how-much-carbon-budget-is-left-to-limit-global-warming-to-1-5cGoogle Scholar
Hayhoe, K, Kheshgi, HS, Jain, AK, Wuebbles, DJ. (2002). Substitution of natural gas for coal: Climatic effects of utility sector emissions. Climatic Change. 54: 107139Google Scholar
Hmiel, B, Petrenko, VV, Dyonisius, MN et al. (2020). Preindustrial 14CH4 indicates greater anthropogenic fossil CH4 emissions. Nature. 578: 409412, doi:10.1038/s41586-020-1991-8Google Scholar
Hong, B and Howarth, RW. (2016). Greenhouse gas emissions from domestic hot water: heat pumps compared to most commonly used systems. Energy Science & Engineering. 4: 123133, doi:10.1002/ese3.112Google Scholar
Howarth, RW. (2014). A bridge to nowhere: Methane emissions and the greenhouse gas footprint of natural gas. Energy Science & Engineering. 2: 4760, doi:10.1002/ese3.35CrossRefGoogle Scholar
Howarth, RW. (2019). Ideas and perspectives: Is shale gas a major driver of recent increase in global atmospheric methane? Biogeosciences. 16: 30333046, doi:10.5194/bg-16-3033-2019Google Scholar
Howarth, RW. (2019a). Interactive comment on “Is shale gas a major driver of recent increase in global atmospheric methane” by Robert W. Howarth et al. Biogeosciences Discussion, doi.org/10.5194/bg-2019-131-AC3CrossRefGoogle Scholar
Howarth, RW. (2020). Methane emissions from fossil fuels: Exploring recent changes in greenhouse-gas reporting requirements for the State of New York. Journal of Integrative Environmental Sciences, doi.org/10.1080/1943815X.2020.1789666Google Scholar
Howarth, RW, Santoro, R, and Ingraffea, A. (2011). Methane and the greenhouse gas footprint of natural gas from shale formations. Climatic Change Letters. 106: 679690, doi:10.1007/s10584-011-0061-5Google Scholar
Howarth, RW, Santoro, R, and Ingraffea, A. (2012). Venting and leakage of methane from shale gas development: Reply to Cathles et al. Climatic Change. 113: 537549, doi:10.1007/s10584-012-0401-0CrossRefGoogle Scholar
Ingraffea, AR, Wawrzynek, PA, Santoro, R, Wells, M. (2020). Reported methane emissions from active oil and gas wells in Pennsylvania, 2014–2018. Environmental Science & Technology. 54: 57835789, doi:10.1021/acs.est.0c00863Google Scholar
Intergovernmental Panel on Climate Change (IPCC). (2013). Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change,. www.ipcc.ch/report/ar5/wg1/Google Scholar
Intergovernmental Panel on Climate Change (IPCC). (2018). Global warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty. Cambridge, United Kingdom and New York, NY, USA, Cambridge University Press.Google Scholar
Karion, A, Sweeney, C, Pétron, G, Frost, G, Hardesty, RM, Kofler, J, Miller, BR, Newberger, T, Wolter, S, Banta, R, and Brewer, A. (2013). Methane emissions estimate from airborne measurements over a western United States natural gas field. Geophysical Research Letters. 40: 43934397, doi:10.1002/grl.50811, 2013.Google Scholar
Karion, A, Sweeney, C, Kort, EA, Shepson, PB, Brewer, A, Cambaliza, M, et al. (2015). Aircraft-based estimate of total methane emissions from the Barnett Shale region. Environmental Science & Technology. 49: 81248131, doi:10.1021/acs.est.5b00217Google Scholar
Kessel, JM and Tabuchi, H. (2019). It’s a vast, invisible climate menace: We made it visible. New York Times, December 12, 2019, www.nytimes.com/interactive/2019/12/12/climate/texas-methane-super-emitters.htmlGoogle Scholar
Kirschke, S, Bousquest, P, Ciais, P, Saunois, M, Canadell, J, Dlugokencky, EJ, Beramaschi, P, Beergmann, D, Blake, D, et al. (2013). Three decades of global methane sources and sinks. Nature Geosciences. 6: 813823, doi:10.1038/ngeo1955Google Scholar
Lamb, BK, Cambaliza, M, Davis, K, Edburg, S, Ferrara, T, Floerchinger, C, Heimburger, A, Herndon, S, Lauvaux, T, Lavoie, T, Lyon, D, Miles, N, Prasad, K, Richardson, S, Roscioli, J, Salmon, O, Shepson, P, Stirm, B, and Whetstone, J. (2016). Direct and indirect measurements and modeling of methane emissions in Indianapolis, Indiana. Environmental Science & Technology. 50: 89108917, doi:10.1021/acs.est.6b01198Google Scholar
Lassey, KR, Etheridge, DM, Lowe, DC, Smith, AM, and Ferretti, DF. (2007). Centennial evolution of the atmospheric methane budget: What do the carbon isotopes tell us? Atmospheric Chemistry and Physics. 7: 21192139, doi:10.5194/acp-7-2119-2007Google Scholar
Lavelle, M. (2019). Trump EPA tries again to roll back methane rules for oil and gas industry. Inside Climate News, August 30, 2019. https://insideclimatenews.org/news/29082019/methane-regulation-oil-gas-storage-pipelines-epa-rollback-trump-wheelerGoogle Scholar
McKain, K, Down, A, Raciti, S, Budney, J, Hutyra, LR, Floerchinger, C, Herndon, SC, Nehrkorn, T, Zahniser, M, Jackson, R, Phillips, N, and Wofsy, S. (2015). Methane emissions from natural gas infrastructure and use in the urban region of Boston, Massachusetts. Proceedings of the National Academy of Sciences. 112: 19411946, doi:10.1073/pnas.1416261112Google Scholar
Milkov, AV, Schwietzke, S, Allen, G, Sherwood, OA, and Etiope, G. (2020). Using global isotopic data to constrain the role of shale gas production in recent increases in atmospheric methane. Scientific Reports. 10: 4199.Google Scholar
Miller, SM, Wofsy, SC, Michalak, AM, Kort, EA, Andrews, AE, Biraud, SC, Dlugokencky, EJ, Janusz Eluszkiewicz, J, Fischer, ML, Janssens-Maenhout, G, Miller, BR, Miller, JB, Montzka, SA, Nehrkorn, T, and Sweeney, C. (2013). Anthropogenic emissions of methane in the United States. Proceedings of the National Academy of Sciences. 110: 2001820022, doi:10.1073/pnas.1314392110Google Scholar
Ocko, IB, Hamburg, SP, Jacob, DJ, Keith, DW, Keohane, NO et al. (2017). Unmask temporal trade-offs in climate policy debates. Science. 356: 492493, doi:10.1126/science.aaj2350Google Scholar
Pandey, S, Gautam, R, Houweling, S, van der Gon, HD, Sadavarte, P, Borsdorff, T, Hasekamp, O, Landgraf, J, Tol, P, van Kempen, T, Hoogeveen, R, van Hees, R, Hamburg, SP, Maasakkers, JD, and Ilse, Aben. (2019). Satellite observations reveal extreme methane leakage from a natural gas well blowout. Proceedings of the National Academy of Sciences. 116: 2637626381. doi.org/10.1073/pnas.1908712116Google Scholar
Peischl, J, Ryerson, T, Brioude, J, Aikin, K, Andrews, A, Atlas, E et al. (2013). Quantifying sources of methane using light alkanes in the Los Angeles basin. California. Journal of Geophysical Research: Atmospheres. 118: 49744990, doi:10.1002/jgrd.50413Google Scholar
Peischl, J, Ryerson, T, Aikin, K, de Gouw, J, Gilman, J, Holloway, J et al. (2015). Quantifying atmospheric methane emissions from the Haynesville, Fayetteville, and northeastern Marcellus Shale gas production regions. Journal of Geophysical Research: Atmospheres. 120: 21192139, doi:10.1002/2014JD022697Google Scholar
Peischl, J, Karion, A, Sweeney, C, Kort, E, Smith, M, Brandt, A et al. (2016). Quantifying atmospheric methane emissions from oil and natural gas production in the Bakken Shale region of North Dakota. Journal of Geophysical Research: Atmospheres. 121: 61016111, doi:10.1002/2015JD024631Google Scholar
Peischl, J, Eilerman, S, Neuman, J, Aikin, K, de Gouw, J, Gilman, J et al. (2018). Quantifying methane and ethane emissions to the atmosphere from central and western U.S. oil and natural gas production regions. Journal of Geophysical Research: Atmospheres. 123: 77257740, doi:10.1029/2018JD028622Google Scholar
Petrenko, V, Smith, A, Schaefer, H et al. (2017). Minimal geological methane emissions during the Younger Dryas–Preboreal abrupt warming event. Nature. 548: 443446, doi:10.1038/nature23316Google Scholar
Plant, G, Kort, EA, Floerchinger, C, Gvakharia, A, Vimont, I, and Sweeney, C. (2019). Large fugitive methane emissions from urban centers along the US east coast. Geophysical Research Letters. 46: 85008507, doi:10.1029/2019GL082635Google Scholar
Ren, X, Hall, D, Vinciguerra, T, Benish, S, Stratton, P, Ahn, D et al. (2019). Methane emissions from the Marcellus Shale in Southwestern Pennsylvania and northern West Virginia based on airborne measurements. Journal of Geophysical Research: Atmospheres. 124: 18621878, doi:10.1029/2018JD029690Google Scholar
Schaefer, H, Mikaloff-Fletcher, S, Veid, C. Lassey, K, Brailsford, G, Bromley, T, Dlubokenck, E, Michel, S, Miller, J, Levin, I, Lowe, D, Martin, R, Vaugn, B, and White, J. (2016). A 21st century shift from fossil-fuel to biogenic methane emissions indicated by 13CH4. Science. 352: 8084, doi:10.1126/science.aad2705.Google Scholar
Schwietzke, S, Sherwood, O, Bruhwiler, L, Miller, J, Etiiope, G, Dlugokencky, E, Michel, S, Arling, V, Vaughn, B, White, J, and Tans, P. (2016). Upward revision of global fossil fuel methane emissions based on isotope database. Nature. 538: 8891, doi:10.1038/nature19797CrossRefGoogle ScholarPubMed
Schneising, O, Burrows, JP, Dickerson, RR, Buchwitz, M, Reuter, M, and Bovensmann, H. (2014). Remote sensing of fugitive emissions from oil and gas production in North American tight geological formations. Earth’s Future. 2: 548558, doi:10.1002/2014EF000265Google Scholar
Schneising, O, Buchwitz, M, Reuter, M, Vanselow, S, Bovensmann, H, and Burrows, JP. (2020). Remote sensing of methane leakage from natural gas and petroleum systems revisited. Atmospheric Chemistry and Physics. 20: 91699183.Google Scholar
Shindell, D, Kuylenstierna, JC, Vignati, E, van Dingenen, R, Amann, M, Klimont, Z, Anenberg, SC, Muller, N, Janssens-Maenhout, G, Raes, R, Schwartz, J, Falvegi, G, Pozzoli, L, Kupiainent, K, Höglund-Isaksson, L, Emberson, L, Streets, D, Ramanathan, V, Kicks, K, Oanh, NT, Milly, G, Williams, M, Demkine, V, and Fowler, D. (2012). Simultaneously mitigating near-term climate change and improving human health and food security. Science. 335: 183189, doi:10.1126/science.1210026Google Scholar
Townsend-Small, A. (2022). Isotopes as tracers of atmospheric and groundwater methane sources. In Stolz, JF, Griffin, WM, and Bain, DJ (eds.) Environmental Impacts from the Development of Unconventional Oil and Gas Reserves. Cambridge University Press.Google Scholar
Turner, AJ, Jacob, DJ, Benmergui, J, Wofsy, SC, Maasakker, JD, Butz, A, Haekamp, O, and Biraud, SC. (2016). A large increase in US methane emissions over the past decade inferred from satellite data and surface observations, Geophysical Research Letters. 43: 22182224, doi:10.1002/2016GL067987Google Scholar
US Department of Agriculture (USDA). (2020). Cattle Inventory. National Agricultural Statistics Service, US Department of Agriculture. www.nass.usda.gov/Surveys/Guide_to_NASS_Surveys/Cattle_Inventory/ downloaded August 28, 2020.Google Scholar
US Energy Information Administration (US EIA). (2016). Shale gas production drives world natural gas production growth. Energy Information Administration, U.S. Department of Energy. www.eia.gov/todayinenergy/detail.php?id=27512, downloaded September 13, 2018.Google Scholar
US EIA. (2020a). How much shale gas is produced in the United States? Energy Information Agency, U.S. Department of Energy. www.eia.gov/tools/faqs/faq.php?id=907&t=8, downloaded November 13, 2020.Google Scholar
US EIA. (2020b). Natural Gas: Dry Shale Gas Production Estimates by Play. Energy Information Agency, U.S. Department of Energy. www.eia.gov/naturalgas/data.php, downloaded September 9, 2020.Google Scholar
US EPA. (2011). Regulatory Impact Analysis: Proposed New Source Performance Standards and Amendments to the National Emissions Standards for Hazardous Air Pollutants for the Oil and Gas Industry. July 2011. U.S. Environmental Protection Agency, Office of Air and Radiation.Google Scholar
US EPA. (2016). Oil and Natural Gas Sector: Emission Standards for New, Reconstructed, and Modified Sources. U.S. Environmental Protection Agency, final rule, 40 CFR Part 60, EPA–HQ–OAR–2010–0505; FRL–9944–75– OAR, RIN 2060–AS30. Federal Register 81 (#107): 35824-35942. www.govinfo.gov/content/pkg/FR-2016-06-03/pdf/2016-11971.pdfGoogle Scholar
Vaughn, TL, Bella, CS, Picering, CK, Schwietzke, S, Heath, GA, Pétron, G, Zimmerle, DJ, Schnell, RC, and Nummedal, D. (2018). Temporal variability largely explains top-down/bottom-up difference in methane emission estimates from a natural gas production region. Proceedings of the National Academy of Sciences. 115: 1171211717, doi:10.1073/pnas.1805687115Google Scholar
Worden, J, Bloom, A, Pandey, S, Jiang, Z, Worden, H, Walter, T, Houweling, S, and Röckmann, T. (2017). Reduced biomass burning emissions reconcile conflicting estimates of the post-2006 atmospheric methane budget. Nature Communications. 8: 2227, doi:10.1038/s41467-017-02246-0, 2017.Google Scholar
Wunch, D, Toon, G, Hedelius, J, Vizenor, N, Roehl, C, Saad, K, Blavier, J, Blake, D, and Wennberg, P. (2016). Quantifying the loss of processed natural gas within California’s South Coast Air Basin using long-term measurements of the ethane and methane. Atmospheric Chemistry and Physics. 16: 1409114105, doi:10.5194/acp-16-14091-2016Google Scholar
Zhang, Y, Gautam, R, Pandey, S, Omara, M, Maasakkers, J, Sadavarte, P, Lyon, D et al. (2020). Quantifying methane emissions from the largest oil-producing basin in the United States from space. Science Advances. 6(17): eaaz5120 doi: 10.1126/sciadv.aaz5120Google Scholar

References

Allison, E and Mandler, B. (2018). Petroleum and the Environment. American Geosciences Institute. https://www.americangeosciences.org/critical-issues/petroleum-environmentGoogle Scholar
Ates, N, Yetis, U, and Kitis, M. (2007). Effects of bromide ion and natural organic matter fractions on the formation and speciation of chlorination by-products. Journal of Environmental Engineering. 133: 947954.Google Scholar
AWWA. (2013). Water and hydraulic fracturing. Available at www.spe.org/jpt/print/archives/2010/12/10Hydraulic.pdf. [Accessed February 24, 2020].Google Scholar
AWWA Research Foundation. (1996). Internal Corrosion of Water Distribution Systems. Second Ed., Denver.Google Scholar
Bair, ES and Digel, RK. (1990). Subsurface transport of inorganic and organic solutes from experimental road spreading of oil-field brine. Ground Water Monitoring & Remediation. 10: 94105. Available at: http://doi.wiley.com/10.1111/j.1745-6592.1990.tb00008.x.CrossRefGoogle Scholar
Bidwell, JR, Farris, JL, and Cherry, DS. (1995). Comparative response of the zebra mussel, Dreissena polymorpha, and the Asian clam, Corbicula fluminea, to DGH/QUAT, a nonoxidizing molluscicide. Aquatic Toxicology. 33: 183200.Google Scholar
Chang, EE, Lin, YP, and Chiang, PC. (2001). Effects of bromide on the formation of THMs and HAAs. Chemosphere. 43: 10291034.Google Scholar
Clark, CE and Veil, JA (2009). Produced Water Volumes and Management Practices in the United States. U.S. Department of Energy.Google Scholar
Clark, JE, Bonura, DK, and Vorhees, RF. (2006). An overview of injection well history in the United States of America. In Tsang, C-F and Apps, JA (eds.) Underground Injection: Science and Technology. Elsevier, pp. 312.Google Scholar
Cowman, GA and Singer, PC. (1996). Effect of bromide ion on haloacetic acid speciation resulting from chlorination and chloramination of aquatic humic substances. Environmental Science & Technology. 30: 1624.Google Scholar
Eaton, LJ, Hoyle, J, and King, A. (1999). Effect of deicing salt on lowbush blueberry flowering and yield. Canadian Journal of Plant Science. 79: 125–128.CrossRefGoogle Scholar
Eckstein, Y. (2011). Is use of oil-field brine as a dust-abating agent really benign? Tracing the source and flowpath of contamination by oil brine in a shallow phreatic aquifer. Environmental Earth Science. 63: 201214.Google Scholar
Ellsworth, WL (2013) Injection-induced earthquakes. Science. 341(6142). Available at https://doi.org/10.1126/science.1225942 [Accessed March 12, 2020].Google Scholar
Elshahed, MS, Najar, FZ, Roe, BA, Oren, A, Dewers, TA, and Krumholz, LR. (2004). Survey of archaeal diversity reveals an abundance of halophilic Archaea in a low-salt, sulfide- and sulfur-rich spring. Applied Environmental Microbiology. 70: 22302239.Google Scholar
Ferrar, KJ, Michanowicz, DR, Christen, CL, Mulcahy, N, Malone, SL, and Sharma, RK. (2013). Assessment of effluent contaminants from three facilities discharging Marcellus shale wastewater to surface waters in Pennsylvania. Environmental Science & Technology. 47(7): 3472–3481.Google Scholar
Gallegos, TJ, Varela, BA, Haines, SS, and Engle, MA. (2015). Hydraulic fracturing water use variability in the United States and potential environmental implications. Water Resources Research. 51: 58395845.Google Scholar
Greenstone, M. (2018). Fracking Has Its Costs And Benefits: The Trick Is Balancing Them. Forbes. Available at www.forbes.com/sites/ucenergy/2018/02/20/fracking-has-its-costs-and-benefits-the-trick-is-balancing-them/#1075fe8e19b4 [Accessed February 24, 2020].Google Scholar
Griswold, E. (2011). The Fracturing of Pennsylvania. New York Times. Available at www.nytimes.com/2011/11/20/magazine/fracking-amwell-township.html [Accessed February 24, 2020].Google Scholar
Groffman, PM, Gold, AJ, and Howard, G. (1995). Hydrologic tracer effects on soil microbial activities. Soil Science Society of America Journal. 59: 478481.Google Scholar
Hart, BT, Bailey, P, Edwards, R, Hortle, K, James, K, McMahon, A, Meredith, C, and Swadling, K. (1991). A review of the salt sensitivity of the Australian freshwater biota. Hydrobiologia. 210: 105144.Google Scholar
Hayes, T. (2009). Sampling and Analysis of Water Streams Associated with the Development of Marcellus Shale gas. Prepared for Marcellus Shale Coalition. Available at https://edx.netl.doe.gov/dataset/sampling-and-analysis-of-water-streams-associated-with-the-development-of-marcellus-shale-gas/resource/4a092e1c-f824-4ecf-8562-0556cd52e353/download/MSCommission-Report.pdf.Google Scholar
Hellergrossman, L, Manka, J, Limonirelis, B, and Rebhun, M. (1993). Formation and distribution of haloacetic acids, THM, and TOX in chlorination of bromide-rich lake water. Water Research. 27: 13231331.Google Scholar
Hong, PKA and Macauley, Y-Y. (1998). Corrosion and leaching of copper tubing exposed to chlorinated drinking water. Water, Air, & Soil Pollution. 108: 457471.Google 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. Seismology Research Letters. 83: 250260.Google Scholar
Institute of Medicine. (2014). Health Impact Assessment of Shale Gas Extraction. National Academies Press.Google Scholar
Kappel, WM, Williams, JH, and Szabo, Z. (2013). Water Resources and Shale Gas/Oil Production in the Appalachian Basin-Critical Issues and Evolving Developments. U.S. Geological Survey Open-File Report 2013-1137. 12p. Available at: https://pubs.usgs.gov/of/2013/1137/pdf/ofr2013-1137.pdfGoogle Scholar
Kaushal, S, Groffman, PM, Likens, GE, Belt, KT, Stack, WP, Kelly, VR, Band, LE, and Fisher, GT. (2005). Increased salinization of fresh water in the northeastern United States. Proceedings of the National Acadamy of Sciences. 102: 1351713520.Google Scholar
Kondash, AJ, Albright, E, and Vengosh, A. (2017). Quantity of flowback and produced waters from unconventional oil and gas exploration. Science of the Total Environment. 574: 314321.Google Scholar
Kondash, A. J. Lauer, NE, and Vengosh, A. (2018). The intensification of the water footprint of hydraulic fracturing. Science Advances. 4(8).Google Scholar
Koplos, J. Tuccillo, ME, and Ranalli, B. (2014). Hydraulic fracturing overview: How, where, and its role in oil and gas. Journal of the American Water Works Association. 106: 3846.Google Scholar
Kuwayama, Y, Olmstead, S, and Krupnick, A. (2015). Water quality and quantity impacts of hydraulic fracturing. Current Sustainable/Renewable Energy Reports. 2: 1724.CrossRefGoogle Scholar
Liang, L and Singer, PC. (2003). Factors influencing the formation and relative distribution of haloacetic acids and trihalomethanes in drinking water. Environmental Science & Technology. 37: 29202928.Google Scholar
Lin, Z, Lin, T, Lim, SH, Hove, MH, and Schuh, WM. (2018). Impacts of Bakken shale oil development on regional water uses and supply. Journal of the American Water Resources Association. 54: 225239.Google Scholar
Lutz, BD, Lewis, AN, and Doyle, MW. (2013). Generation, transport, and disposal of wastewater associated with Marcellus Shale gas development. Water Resources Research. 49: 647656.Google Scholar
Maloney, KO, Young, JA, Faulkner, SP, Hailegiorgis, A, Slonecker, ET, and Milheim, LE. (2018). A detailed risk assessment of shale gas development on headwater streams in the Pennsylvania portion of the Upper Susquehanna River Basin, U.S.A. Science of the Total Environment. 610611: 154166.Google Scholar
Mitchell, AL, Small, M, and Casman, EA. (2013). Surface water withdrawals for Marcellus shale gas development: Performance of alternative regulatory approaches in the upper Ohio river basin. Environmental Science & Technology. 47: 1266912678.Google Scholar
Murray, KE. (2013). State-scale perspective on water use and production associated with oil and gas operations, Oklahoma, U.S. Environmental Science & Technology. 47: 49184925.Google Scholar
Muylwyk, Q, Sandvig, A, and Snoeyink, V. (2014). Developing corrosion control for drinking water systems. Opflow. 40: 2427.Google Scholar
Nagpal, NK, Levy, SA, MacDonald, DD, and Ministry of Environment Canada B. C. (2003). Water Quality: Ambient water quality guidelines for choride – overview report. Available at www.env.gov.bc.ca/wat/wq/BCguidelines/chloride/chloride.html.Google Scholar
National Academies of Sciences. (2017). Flowback and Produced Waters: Opportunities and Challenges for Innovation: Proceedings of a Workshop. National Academies Press.Google Scholar
National Geographic. (2013). How hydraulic fracturing works. National Geographic Magazine Available at www.nationalgeographic.org/media/how-hydraulic-fracturing-works/ [Accessed February 24, 2020].Google Scholar
National Research Council. (2014). Risks and Risk Governance in Shale Gas Development. National Academies Press.Google Scholar
NGWA. (2013). Water wells in proximity to natural gas or oil development. Available at www.ntllabs.com [Accessed February 24, 2020].Google Scholar
Nicot, JP and Scanlon, BR. (2012). Water use for shale-gas production in Texas, U.S. Environmental Science & Technology. 46: 35803586.Google Scholar
Nielsen, DL, Brock, MA, Rees, GN, and Baldwin, DS. (2003). Effects of increasing salinity on freshwater ecosystems in Australia. Australian Journal of Botany. 51: 655665. Available at www.publish.csiro.au/paper/BT02115.htm.Google Scholar
North Dakota State Water Commission. (2019). North Dakota Fracking & Water Facts. Available at www.swc.nd.gov/pdfs/fracking_water_use.pdf [Accessed February 24, 2020].Google Scholar
Nunez, C. (2015). Fracking, quakes, and drinking water: Your questions answered. National Geographic Magazine Available at www.nationalgeographic.com/news/energy/2015/07/150723-fracking-questions-answered/#close [Accessed February 24, 2020].Google Scholar
Nunez, C. (2013). How has fracking changed our future? National Geographic Magazine Available at: www.nationalgeographic.com/environment/energy/great-energy-challenge/big-energy-question/how-has-fracking-changed-our-future/#close [Accessed February 24, 2020].Google Scholar
NYS DEC. (1999). An Investigation of Naturally Occurring Radioactive Materials (NORM) in Oil and Gas Wells in New York State. Available at www.dec.ny.gov/docs/materials_minerals_pdf/normrpt.pdf.Google Scholar
PA DEP. (2018a). 2018 Oil and Gas Annual Report. Available at www.depgis.state.pa.us/2018OilGasAnnualReport/index.html [Accessed January 2, 2020].Google Scholar
PA DEP. (2018b). Oil and gas reports. Available at www.dep.pa.gov/DataandTools/Reports/Oil and Gas Reports/Pages/default.aspx [Accessed January 2, 2020].Google Scholar
PA DEP. (2011). Proposed Rulemaking [25 PA. CODE CH. 95] Wastewater treatment requirements [39 Pa.B. 6467]. Available at www.pabulletin.com/secure/data/vol39/39-45/2065.html.Google Scholar
Richardson, SD, Plewa, MJ, Wagner, ED, Schoeny, R, and DeMarini, DM. (2007). Occurrence, genotoxicity, and carcinogenicity of regulated and emerging disinfection by-products in drinking water: A review and roadmap for research. Mutation Research. 636: 178242.Google Scholar
Rosa, L, Rulli, MC, Davis, KF, and D’Odorico, P. (2018). The water-energy nexus of hydraulic fracturing: A global hydrologic analysis for shale oil and gas extraction. Earth’s Future. 6: 745756.Google Scholar
Rosenblum, J, Nelson, AW, Ruyle, B, Schultz, MK, Ryan, JN, and Linden, KG. (2017). Temporal characterization of flowback and produced water quality from a hydraulically fractured oil and gas well. Science of the Total Environment. 596597: 369377.Google Scholar
Ross, N and Luu, P. (2012). Hydraulic fracturing and water resources: Separating the frack from the fiction. Available at www.pacinst.orgphone:510.251.1600Facsimile:510.251.2203 [Accessed February 24, 2020].Google Scholar
Scanlon, BR, Reedy, RC, and Nicot, JP. (2014). Comparison of water use for hydraulic fracturing for unconventional oil and gas versus conventional oil. Environmental Science & Technology. 48: 1238612393.Google Scholar
Soeder, DJ and Kappel, WM. (2009). Water resources and natural gas production from the Marcellus Shale. Available at http://geology.com/articles/marcellus-shale.shtml [Accessed February 18, 2020].Google Scholar
States, S, Cyprych, G, Stoner, M, Wydra, F, Kuchta, J, Monnell, J, and Casson, L. (2013). Brominated THMs in drinking water: A possible link to Marcellus Shale and other wastewaters. Journal of the American Water Works Association. 105: E432E448. Available at www.awwa.org/publications/journal-awwa/abstract/articleid/38156568.aspx [Accessed June 8, 2017].Google Scholar
Stone, J. (2017). Fracking is dangerous to your health: Here’s why. Forbes. Available at www.forbes.com/sites/judystone/2017/02/23/fracking-is-dangerous-to-your-health-heres-why/#45a60fd75945 [Accessed February 24, 2020].Google Scholar
Tasker, TL, Burgos, WD, Piotrowski, P, Castillo-Meza, L, Blewett, TA, Ganow, KB, Stallworth, A, Delompré, PLM, Goss, GG, Fowler, LB, Vanden Heuvel, JP, Dorman, F, and Warner, NR. (2018). Environmental and human health impacts of spreading oil and gas wastewater on roads. Environmental Science & Technology. 52: 70817091.Google Scholar
US EPA. (1988). Ambient Aquatic Life Water Quality Criteria for Chloride. Washington, DCGoogle Scholar
US EPA. (2015). Assessment of the Potential Impacts of Hydraulic Fracturing for Oil and Gas on Drinking Water Resources. Washington, DC.Google Scholar
US EPA. (2001). Class I underground injection control program: Study of the risks associated with class I underground injection wells. Available at www.epa.gov/safewater [Accessed January 2, 2020].Google Scholar
US EPA Science Advisory Board. (2016). SAB review of the EPA’s draft assessment of the potential impacts of hydraulic fracturing for oil and gas on drinking water resources.Google Scholar
VanBriesen, JM and Hammer, R. (2012). In Fracking’s Wake: New Rules Are Needed to Protect our Health and Environment from Contaminated Wastewater. NRDC.Google Scholar
VanBriesen, JM, Wilson, JM, and Wang, Y. (2014). Management of produced water in Pennsylvania: 2010–2012. In Proceedings of the 2014 Shale Energy Engineering Conference. Pittsburgh, pp. 17.Google Scholar
Veil, J. (2015). U.S. Produced Water Volumes and Management Practices in 2012. Available at www.gwpc.org/sites/default/files/Produced Water Report 2014-GWPC_0.pdf [Accessed February 24, 2020].Google Scholar
Veil, J. (2020). U.S. Produced Water Volumes and Management Practices in 2017. Available at www.veilenvironmental.com/publications/pw/pw_report_2017_final.pdf [Accessed February 24, 2020]Google Scholar
Walsh, FR and Zoback, MD. (2015). Oklahoma’s recent earthquakes and saltwater disposal. Science Advances. 1: e1500195.Google Scholar
Wang, Y, Small, MJ, and VanBriesen, JM. (2017). Assessing the risk associated with increasing bromide in drinking water sources in the Monongahela River. Pennsylvania Journal of Environmental Engineering. 143: 04016089.Google Scholar
Warner, NR, Christie, CA, Jackson, RB, and Vengosh, A. (2013). Impacts of shale gas wastewater disposal on water quality in Western Pennsylvania. Environmental Science & Technology. 47: 1184911857.CrossRefGoogle ScholarPubMed
Warwick, NWM and Bailey, PCE. (1997). The effect of increasing salinity on the growth and ion content of three non-halophytic wetland macrophytes. Aquatic Botany. 58: 7388.Google Scholar
Wilson, JM and VanBriesen, JM. (2012). Oil and gas produced water management and surface drinking water sources in Pennsylvania. Environmental Practice 14: 288300.Google Scholar
Wilson, JM and VanBriesen, JM. (2013). Source water changes and energy extraction activities in the Monongahela River, 2009–2012. Environmental Science & Technology. 47: 1257512582.Google Scholar
Young, WF, Horth, H, Crane, R, Ogden, T, and Arnott, M. (1996). Taste and odour threshold concentrations of potential potable water contaminants. Water Research. 30: 331340.Google Scholar

References

AER (Alberta Energy Regulator). (2019). Subsurface Order No. 6. www.aer.ca/documents/orders/subsurface-orders/SO6.pdfGoogle Scholar
Aki, K. (1965). Maximum likelihood estimate of b in the formula log N = abM and its confidence limits. Bulletin of the Earthquake Research Institute. 43(1): 237239.Google Scholar
Anderson, EM. (1951). The Dynamics of Faulting and Dyke Formation with Applications to Britain. Hafner Pub. Co.Google Scholar
Atkinson, GM, Eaton, DW, Ghofrani, H, Walker, D, Cheadle, B, Schultz, R, Shcherbakov, R, Tiampo, K, Gu, J, Harrington, RM, and Liu, Y. (2016). Hydraulic fracturing and seismicity in the western Canada sedimentary basin. Seismological Research Letters. 87(3): 631647.Google Scholar
Atkinson, GM, Eaton, DW, Igonin, N. (2020). Developments in understanding seismicity triggered by hydraulic fracturing. Nature Reviews Earth & Environment. 1(1): 264277.Google Scholar
Babaie Mahani, AB, Schultz, R, Kao, H, Walker, D, Johnson, J, and Salas, C. (2017). Fluid injection and seismic activity in the northern Montney play, British Columbia, Canada, with special reference to the 17 August 2015 Mw 4.6 induced earthquake. Bulletin of the Seismological Society of America. 107(2): 542552.Google Scholar
Babaie Mahani, A, Kao, H, Atkinson, GM, Assatourians, K, Addo, K, and Liu, Y. (2019). Ground‐motion characteristics of the 30 November 2018 injection‐induced earthquake sequence in Northeast British Columbia, Canada.Seismological Research Letters. 90(4): 14571467.Google Scholar
Babaie Mahani, A, Esfahani, F, Kao, H, Gaucher, M, Hayes, M, Visser, R, and Venables, S. (2020). A systematic study of earthquake source mechanism and regional stress field in the Southern Montney Unconventional Play of northeast British Columbia, Canada. Seismological Research Letters. 91(1): 195206.Google Scholar
Bao, X, and Eaton, DW. (2016). Fault activation by hydraulic fracturing in western Canada. Science. 354(6318): 14061409.Google Scholar
Barbour, AJ, Norbeck, JH, and Rubinstein, JL. (2017). The effects of varying injection rates in Osage County, Oklahoma, on the 2016 Mw 5.8 Pawnee earthquake. Seismological Research Letters. 88(4): 10401053.Google Scholar
Barclay, JE, Krause, FF, Campbell, RI, and Utting, J. (1990). Dynamic casting and growth faulting: Dawson Creek graben complex, Carboniferous-Permian Peace River embayment, western Canada. Bulletin of Canadian Petroleum Geology. 38(1): 115145.Google Scholar
BC OGC (British Columbia Oil and Gas Commission). (2014a). Montney Formation Play Atlas NEBC. www.bcogc.ca/node/8131/downloadGoogle Scholar
BC OGC (British Columbia Oil and Gas Commission). (2014b). Investigation of observed seismicity in the Montney trend. www.bcogc.ca/sites/default/files/documentation/technical-reports/investigation-observed-seismicity-montney-trend.pdfGoogle Scholar
BC OGC (British Columbia Oil and Gas Commission). (2017). Application guideline for deep well disposal of produced water. www.bcogc.ca/application-guideline-deep-well-disposal-produced-water-non-hazardous-waste.Google Scholar
Bhattacharya, P and Viesca, RC. (2019). Fluid-induced aseismic fault slip outpaces pore-fluid migration. Science. 364(6439): 464468.Google Scholar
Biot, MA. (1941). General theory of three‐dimensional consolidation. Journal of Applied Physics. 12(2): 155164.Google Scholar
Bommer, JJ, Oates, S, Cepeda, JM, Lindholm, C, Bird, J, Torres, R, Marroquín, G, and Rivas, J. (2006). Control of hazard due to seismicity induced by a hot fractured rock geothermal project. Engineering Geology. 83(4): 287306.Google Scholar
Cardott, BJ. (2012). Thermal maturity of Woodford Shale gas and oil plays, Oklahoma, USA. International Journal of Coal Geology. 103(1): 109119.Google Scholar
CCA (Council of Canadian Academies). (2014). Environmental Impacts of Shale Gas Extraction in Canada. Council of Canadian Academies.Google Scholar
Clarke, H, Eisner, L, Styles, P, and Turner, P. (2014). Felt seismicity associated with shale gas hydraulic fracturing: The first documented example in Europe. Geophysical Research Letters. 41(23): 83088314.Google Scholar
Clarke, H, Verdon, JP, Kettlety, T, Baird, AF, and Kendall, JM. (2019). Real‐time imaging, forecasting, and management of human‐induced seismicity at Preston New Road, Lancashire, England. Seismological Research Letters. 90(5): 19021915.Google Scholar
Darold, A, Holland, AA, Chen, C, and Youngblood, A. (2014). Preliminary Analysis of Seismicity near Eagleton 1–29, Carter County, July 2014. Oklahoma Geological Survey Open File Report, OF2–2014.Google Scholar
Deng, K, Liu, Y, and Harrington, RM. (2016). Poroelastic stress triggering of the December 2013 Crooked Lake, Alberta, induced seismicity sequence. Geophysical Research Letters. 43(16): 84828491.Google Scholar
Dieterich, J. (1994). A constitutive law for rate of earthquake production and its application to earthquake clustering. Journal of Geophysical Research: Solid Earth. 99(B2): 26012618.Google Scholar
Ducros, M, Sassi, W, Vially, R, Euzen, T, and Crombez, V. (2017). 2-D basin modeling of the Western Canada Sedimentary Basin across the Montney-Doig system: Implications for hydrocarbon migration pathways and unconventional resources potential. In AbuAli, MA Moretti, I, and Bolås, HMN (eds.) Memoir 114: Petroleum Systems Analysis: Case Studies, Tulsa, OK. American Association of Petroleum Geologists.Google Scholar
Dziewonski, AM, Chou, TA, and Woodhouse, JH. (1981). Determination of earthquake source parameters from waveform data for studies of global and regional seismicity. Journal of Geophysical Research: Solid Earth. 86(B4): 28252852.Google Scholar
Eaton, DW. (2018). Passive Seismic Monitoring of Induced Seismicity: Fundamental Principles and Application to Energy Technologies. Cambridge University Press.Google Scholar
Eaton, DW and Igonin, N. (2018). What controls the maximum magnitude of injection-induced earthquakes? The Leading Edge. 37(2): 135140.Google Scholar
Eaton, DW anf Maghsoudi, S. (2015). 2b… or not 2b? Interpreting magnitude distributions from microseismic catalogs. First Break. 33(10): 7986.CrossRefGoogle Scholar
Eaton, DW and Schultz, R. (2018). Increased likelihood of induced seismicity in highly overpressured shale formations. Geophysical Journal International. 214(1): 751757.Google Scholar
Eaton, DW, Igonin, N, Poulin, A, Weir, R, Zhang, H, Pellegrino, S, and Rodriguez, G. (2018). Induced seismicity characterization during hydraulic‐fracture monitoring with a shallow‐wellbore geophone array and broadband sensors. Seismological Research Letters. 89(5): 16411651.Google Scholar
Eaton, DW, Milkereit, B, Ross, GM, Kanasewich, ER, Geis, W, Edwards, DJ, Kelsch, L, and Varsek, J. (1995). Lithoprobe basin-scale seismic profiling in central Alberta: Influence of basement on the sedimentary cover. Bulletin of Canadian Petroleum Geology. 43(1): 6577.Google Scholar
Edwards, DE, Barclay, J, Gibson, D, Kvill, G, and Halton, E. (1994). Triassic strata of the Western Canada Sedimentary Basin. In Mossop, GD and Shetsen, I (eds.) Geological Atlas of the Western Canada Sedimentary Basin. Calgary, AB: Canadian Society of Petroleum Geologists and Alberta Research Council.Google Scholar
EIA (Energy Information Administration). (2015). World Shale Resource Assessments, www.eia.gov/analysis/studies/worldshalegas.Google Scholar
Ekström, G, Nettles, M, and Dziewoński, AM. (2012). The global CMT project 2004–2010: Centroid-moment tensors for 13,017 earthquakes. Physics of the Earth and Planetary Interiors. 200(1): 19.Google Scholar
Ellsworth, WL. (2013). Injection-induced earthquakes. Science: 341(6142), 1225942, https://doi.org/10.1126/science.1225942.Google Scholar
Ellsworth, WL, Llenos, AL, McGarr, AF, Michael, AJ, Rubinstein, JL, Mueller, CS, Petersen, MD, and Calais, E. (2015). Increasing seismicity in the US midcontinent: Implications for earthquake hazard. The Leading Edge. 34(6): 618626.Google Scholar
Eyre, TS, Eaton, DW, Zecevic, M, D’Amico, D, and Kolos, D. (2019a). Microseismicity reveals fault activation before MW 4.1 hydraulic-fracturing induced earthquake. Geophysical Journal International. 218(1): 534546.Google Scholar
Eyre, TS, Eaton, DW, Garagash, DI, Zecevic, M, Venieri, M, Weir, R, and Lawton, DC. (2019b). The role of aseismic slip in hydraulic fracturing–induced seismicity. Science Advances. 5(8): eaav7172. https://doi.org/10.1126/sciadv.aav7172Google Scholar
Foulger, GR, Wilson, M, Gluyas, J, Julian, BR, and Davies, R. (2018). Global review of human-induced earthquakes. Earth-Science Reviews. 178(1), 438514.Google Scholar
Friberg, PA, Besana‐Ostman, GM, and Dricker, I. (2014). Characterization of an earthquake sequence triggered by hydraulic fracturing in Harrison County, Ohio. Seismological Research Letters. 85(6), 12951307.Google Scholar
Frohlich, C, DeShon, H, Stump, B, Hayward, C, Hornbach, M, and Walter, JI. (2016). A historical review of induced earthquakes in Texas. Seismological Research Letters. 87(4): 10221038.Google Scholar
Galis, M, Ampuero, JP, Mai, PM, and Cappa, F. (2017). Induced seismicity provides insight into why earthquake ruptures stop. Science Advances. 3(12): eaap7528, https://doi.org/10.1126/sciadv.aap7528.Google Scholar
Gan, W and Frohlich, C. (2013). Gas injection may have triggered earthquakes in the Cogdell oil field, Texas. Proceedings of the National Academy of Sciences. 110(47): 1878618791.Google Scholar
Ghofrani, H and Atkinson, GM. (2020). Activation rate of seismicity for hydraulic fracture wells in the Western Canada Sedimentary Basin. Bulletin of the Seismological Society of America, in press.Google Scholar
Goebel, THW, Weingarten, M, Chen, X, Haffener, J, and Brodsky, EE. (2017). The 2016 Mw5. 1 Fairview, Oklahoma earthquakes: Evidence for long-range poroelastic triggering at > 40 km from fluid disposal wells. Earth and Planetary Science Letters. 472(1): 5061.Google Scholar
Gómez-Alba, S, Vargas, C, and Zang, A. (2020). Evidencing the relationship between injected volume of water and maximum expected magnitude during the Puerto Gaitán (Colombia) earthquake sequence from 2013 to 2015. Geophysical Journal International. 220(1): 335344.Google Scholar
Guglielmi, Y, Cappa, F, Avouac, JP, Henry, P, and Elsworth, D. (2015). Seismicity triggered by fluid injection–induced aseismic slip. Science. 348(6240): 12241226.Google Scholar
Gutenberg, B. and Richter, CF. (1944). Frequency of earthquakes in California. Bulletin of the Seismological Society of America. 34(4): 185188.Google Scholar
Hanks, TC and Kanamori, H. (1979). A moment magnitude scale. Journal of Geophysical Research: Solid Earth. 84(B5): 23482350.Google Scholar
Häring, MO, Schanz, U, Ladner, F, and Dyer, BC. (2008). Characterisation of the Basel 1 enhanced geothermal system. Geothermics. 37(5): 469495.Google Scholar
Healy, JH, Rubey, WW, Griggs, DT, and Raleigh, CB. (1968). The Denver earth-quakes. Science. 161(3848): 13011310.Google Scholar
Heidbach, O, Rajabi, M, Reiter, K, and Ziegler, M. (2016): World Stress Map Database Release 2016. GFZ Data Services, https://doi.org/10.5880/WSM.2016.001.Google Scholar
Heidbach, O, Rajabi, M, Cui, X, Fuchs, K, Müller, B, Reinecker, J, Reiter, K, Tingay, M, Wenzel, F, Xie, F, and Ziegler, MO. (2018). The World Stress Map database release 2016: Crustal stress pattern across scales. Tectonophysics. 744(1): 484-498.Google Scholar
Hennings, PH, Lund Snee, JE, Osmond, JL, DeShon, HR, Dommisse, R, Horne, E, Lemons, C, and Zoback, MD (2019). Injection‐induced seismicity and fault‐slip potential in the Fort Worth Basin, Texas. Bulletin of the Seismological Society of America. 109(5): 16151634.Google Scholar
Herrmann, RB.(2020). North America Moment Tensors. www.eas.slu.edu/eqc/eqc_mt/MECH.NA/.Google Scholar
Herrmann, RB, Park, SK, and Wang, CY. (1981). The Denver earthquakes of 1967–1968. Bulletin of the Seismological Society of America. 71(3), 731745.Google Scholar
Herrmann, RB, Benz, H, and Ammon, CJ. (2011). Monitoring the earthquake source process in North America. Bulletin of the Seismological Society of America. 101(6): 26092625.Google Scholar
Holland, A. (2013). Earthquakes triggered by hydraulic fracturing in south-central Oklahoma. Bulletin of the Seismological Society of America. 103(3): 17841792.Google Scholar
Holmgren, JM, Atkinson, GM, and Ghofrani, H. (2019). Stress drops and directivity of induced earthquakes in the Western Canada Sedimentary Basin. Bulletin of the Seismological Society of America. 109(5): 16351652.Google Scholar
Hosseini, BK and Eaton, DW. (2018). Fluid flow and thermal modeling for tracking induced seismicity near the Graham disposal well, British Columbia, Canada. SEG Technical Program Expanded Abstracts 2018 (pp. 4987–4991). Society of Exploration Geophysicists, https://doi.org/10.1190/segam2018–2996360.1.Google Scholar
Igonin, N, Zecevic, M, and Eaton, DW. (2018). Bilinear magnitude‐frequency distributions and characteristic earthquakes during hydraulic fracturing. Geophysical Research Letters. 45(23): 1286612874.Google Scholar
Ishimoto, M and Iida, K. (1939). Observations of earthquakes registered with the micro-seismograph constructed recently. Bulletin of the Earthquake Research Institute. 17(1): 443478.Google Scholar
Jia, SQ, Wong, RCK, Eaton, DW, and Eyre, TS. (2018). Investigating fracture growth and source mechanisms in shale using acoustic emission technique. In 52nd US Rock Mechanics/Geomechanics Symposium. American Rock Mechanics Association, ARMA-2018-136.Google Scholar
Kanamori, H and Anderson, DL. (1975). Theoretical basis of some empirical relations in seismology. Bulletin of the Seismological Society of America. 65(5): 10731095.Google Scholar
Kang, JQ, Zhu, JB, and Zhao, J. (2019). A review of mechanisms of induced earthquakes: from a view of rock mechanics. Geomechanics and Geophysics for Geo-Energy and Geo-Resources. 5(2): 171196.Google Scholar
Kao, H, Visser, R, Smith, B, and Venables, S. (2018). Performance assessment of the induced seismicity traffic light protocol for northeastern British Columbia and western Alberta. The Leading Edge. 37(2): 117126.Google Scholar
Keranen, KM and Weingarten, M. (2018). Induced seismicity. Annual Review of Earth and Planetary Sciences. 46(1): 149174.Google Scholar
Keranen, KM, Savage, HM, Abers, GA, and Cochran, ES. (2013). Potentially induced earthquakes in Oklahoma, USA: Links between wastewater injection and the 2011 Mw 5.7 earthquake sequence. Geology. 41(6): 699702.Google Scholar
Keranen, KM, Weingarten, M, Abers, GA, Bekins, BA, and Ge, S. (2014). Sharp increase in central Oklahoma seismicity since 2008 induced by massive wastewater injection. Science. 345, 448451.Google Scholar
Kettlety, T, Verdon, JP, Werner, MJ, and Kendall, JM. (2020). Stress transfer from opening hydraulic fractures controls the distribution of induced seismicity. Journal of Geophysical Research: Solid Earth. 125: e2019JB018794. https://doi.org/10.1029/2019JB018794Google Scholar
Kettlety, T, Verdon, JP, Werner, MJ, Kendall, JM, and Budge, J. (2019). Investigating the role of elastostatic stress transfer during hydraulic fracturing-induced fault activation. Geophysical Journal International. 217(2): 12001216.Google Scholar
Kim, YS, Peacock, DC, and Sanderson, DJ. (2004). Fault damage zones. Journal of Structural Geology. 26(3): 503517.Google Scholar
Knopoff, L. (2000). The magnitude distribution of declustered earthquakes in Southern California. Proceedings of the National Academy of Sciences. 97(22): 1188011884.Google Scholar
Kohli, AH and Zoback, MD. (2013). Frictional properties of shale reservoir rocks. Journal of Geophysical Research: Solid Earth. 118(9): 51095125.Google Scholar
Kozłowska, M, Brudzinski, MR, Friberg, P, Skoumal, RJ, Baxter, ND, and Currie, BS. (2018). Maturity of nearby faults influences seismic hazard from hydraulic fracturing. Proceedings of the National Academy of Sciences. 115(8): E1720E1729.Google Scholar
Kwiatek, G, Saarno, T, Ader, T, Bluemle, F, Bohnhoff, M, Chendorain, M, Dresen, G, Heikkinen, P, Kukkonen, I, Leary, P, and Leonhardt, M. (2019). Controlling fluid-induced seismicity during a 6.1-km-deep geothermal stimulation in Finland. Science Advances. 5(5): eaav7224, https://doi.org/10.1126/sciadv.aav7224.Google Scholar
Langenbruch, C and Zoback, MD. (2016). How will induced seismicity in Oklahoma respond to decreased saltwater injection rates? Science Advances. 2(11): e1601542. https://doi.org/10.1126/sciadv.1601542.Google Scholar
Lavrov, A. (2016). Dynamics of stresses and fractures in reservoir and cap rock under production and injection. Energy Procedia. 86(1): 381390.Google Scholar
Lei, X, Wang, Z, and Su, J. (2019). The December 2018 ML 5.7 and January 2019 ML 5.3 Earthquakes in South Sichuan Basin Induced by Shale Gas Hydraulic Fracturing. Seismological Research Letters. 90(3): 10991110.Google Scholar
Lengliné, O, Lamourette, L, Vivin, L, Cuenot, N, and Schmittbuhl, J. (2014). Fluid‐induced earthquakes with variable stress drop. Journal of Geophysical Research: Solid Earth. 119(12): 89008913.Google Scholar
Ludwig, J, Nafe, J, and Drake, C. (1971). Seismic refraction. In Maxwell, AE (ed.) The Sea, Vol. 4 . Wiley, pp. 5384.Google Scholar
MacKay, MK, Eaton, DW, Pedersen, PK, and Clarkson, CR. (2018). Integration of outcrop, subsurface, and microseismic interpretation for rock-mass characterization: An example from the Duvernay Formation, Western Canada. Interpretation 6(4): T919T936.Google Scholar
Mahani, AB, Schultz, R, Kao, H, Walker, D, Johnson, J, and Salas, C. (2017). Fluid injection and seismic activity in the northern Montney play, British Columbia, Canada, with special reference to the 17 August 2015 MW 4.6 induced earthquake. Bulletin of the Seismological Society of America. 107(2): 542552.Google Scholar
Marone, C. (1998). Laboratory-derived friction laws and their application to seismic faulting. Annual Review of Earth and Planetary Sciences. 26(1): 643696.Google Scholar
Marsh, S and Holland, A. (2016). Comprehensive Fault Database and Interpretive Fault Map of Oklahoma. Oklahoma Geological Survey Open‐File Rep. OF22016.Google Scholar
McGarr, A. (2014). Maximum magnitude earthquakes induced by fluid injection. Journal of Geophysical Research: Solid Earth. 119(2): 10081019.Google Scholar
McGarr, A and Barbour, AJ. (2017). Wastewater disposal and the earthquake sequences during 2016 near Fairview, Pawnee, and Cushing, Oklahoma. Geophysical Research Letters. 44(18): 93309336.Google Scholar
McNamara, DE, Hayes, GP, Benz, HM, Williams, RA, McMahon, ND, Aster, RC, Holland, A, Sickbert, T, Herrmann, R, Briggs, R, and Smoczyk, G. (2015). Reactivated faulting near Cushing, Oklahoma: Increased potential for a triggered earthquake in an area of United States strategic infrastructure. Geophysical Research Letters. 42(20): 83288332.Google Scholar
NRC (National Research Council). (2013). Induced Seismicity Potential in Energy Technologies. National Academies Press.Google Scholar
Ohnaka, M, Akatsu, M, Mochizuki, H, Odedra, A, Tagashira, F, and Yamamoto, Y. (1997). A constitutive law for the shear failure of rock under lithospheric conditions. Tectonophysics. 277(1–3): 127.Google Scholar
Okada, Y. (1992). Internal deformation due to shear and tensile faults in a half-space. Bulletin of the Seismological Society of America. 82(2): 10181040.Google Scholar
OCC (Oklahoma Corporation Commission). (2018). Moving forward: New protocol to further address seismicity in state’s largest oil and gas play, www.occeweb.com/og/02-27-18PROTOCOL.pdf.Google Scholar
Oklahoma Produced Water Working Group. 2017. Oklahoma Water for 2060: Produced Water Reuse and Recycling. www.owrb.ok.gov/2060/pwwg.php.Google Scholar
Parry, RH. (2004). Mohr Circles, Stress Paths and Geotechnics, 2nd Edition, CRC Press.Google Scholar
Pawley, S, Schultz, R, Playter, T, Corlett, H, Shipman, T, Lyster, S, and Hauck, T. (2018). The geological susceptibility of induced earthquakes in the Duvernay play. Geophysical Research Letters. 45(4): 17861793.Google Scholar
Peterie, SL, Miller, RD, Intfen, JW, and Gonzales, JB. (2018). Earthquakes in Kansas induced by extremely far‐field pressure diffusion. Geophysical Research Letters. 45(3): 13951401.Google Scholar
Ponomarev, AV, Zavyalov, AD, Smirnov, VB, and Lockner, DA. (1997). Physical modeling of the formation and evolution of seismically active fault zones. Tectonophysics. 277(1–3). 5781.Google Scholar
Poulin, A, Weir, R, Eaton, D, Igonin, N, Chen, Y, Lines, L, and Lawton, D. (2019). Focal-time analysis: A new method for stratigraphic depth control of microseismicity and induced seismic events. Geophysics. 84(6): KS173-KS182.Google Scholar
Raleigh, CB, Healy, JH, and Bredehoeft, JD. (1976). An experiment in earthquake control at Rangely, Colorado. Science. 191(4233): 12301237.Google Scholar
Richter, CF. (1935). An instrumental earthquake magnitude scale. Bulletin of the Seismological Society of America: 25(1): 132.Google Scholar
Roche, V and Van der Baan, M. (2015). The role of lithological layering and pore pressure on fluid‐induced microseismicity. Journal of Geophysical Research: Solid Earth. 120(2): 923943.Google Scholar
Rokosh, CD, Lyster, S, Anderson, SDA, Beaton, AP, Berhane, H, Brazzoni, T, Chen, D, Cheng, Y, Mack, T, Pana, C, and Pawlowicz, JG. (2012). Summary of Alberta’s Shale- and Siltstone-Hosted Hydrocarbon Resource Potential. Energy Resources Conservation Board, ERCB/AGS Open File Report, 2012-06.Google Scholar
Rubinstein, JL and Mahani, AB. (2015). Myths and facts on wastewater injection, hydraulic fracturing, enhanced oil recovery, and induced seismicity. Seismological Research Letters. 86(4): 10601067.Google Scholar
Schoenball, M and Ellsworth, WL. (2017). Waveform‐relocated earthquake catalog for Oklahoma and southern Kansas illuminates the regional fault network. Seismological Research Letters. 88(5): 12521258.Google Scholar
Schoenball, M, Walsh, FR, Weingarten, M, and Ellsworth, WL. (2018). How faults wake up: The Guthrie-Langston, Oklahoma earthquakes. The Leading Edge. 37(2): 100106.Google Scholar
Scholz, CH. (1998). Earthquakes and friction laws. Nature. 391(6662): 3742.Google Scholar
Schultz, R. and Pawley, S. (2019). Induced Earthquakes Geological Susceptibility Model for the Duvernay Formation, Central Alberta: Version 2. AER/AGS Open File Report 2019-02.Google Scholar
Schultz, R. and Wang, R. (2020). Newly emerging cases of hydraulic fracturing induced seismicity in the Duvernay East Shale Basin. Tectonophysics. 228393, https://doi.org/10.1016/j.tecto.2020.228393.Google Scholar
Schultz, R, Stern, V, Novakovic, M, Atkinson, G, and Gu, YJ. (2015). Hydraulic fracturing and the Crooked Lake Sequences: Insights gleaned from regional seismic networks. Geophysical Research Letters. 42(8): 27502758.Google Scholar
Schultz, R, Atkinson, G, Eaton, DW, Gu, YJ, and Kao, H. (2018). Hydraulic fracturing volume is associated with induced earthquake productivity in the Duvernay play. Science. 359(6373): 304308.Google Scholar
Schultz, R, Wang, R, Gu, YJ, Haug, K, and Atkinson, G. (2017). A seismological overview of the induced earthquakes in the Duvernay play near Fox Creek, Alberta. Journal of Geophysical Research: Solid Earth. 122(1): 492505.Google Scholar
Schultz, R, Skoumal, RJ, Brudzinski, MR, Eaton, DW, Baptie, B, and Ellsworth, WL. (2020). Hydraulic Fracturing Induced Seismicity. Reviews of Geophysics, submitted.Google Scholar
Schultz, R, Corlett, H, Haug, K, Kocon, K, MacCormack, K, Stern, V, and Shipman, T. (2016). Linking fossil reefs with earthquakes: Geologic insight to where induced seismicity occurs in Alberta. Geophysical Research Letters. 43(6), 25342542.Google Scholar
Segall, P and Lu, S. (2015). Injection‐induced seismicity: Poroelastic and earthquake nucleation effects. Journal of Geophysical Research: Solid Earth. 120(7): 50825103.Google Scholar
Shah, AK and Keller, GR. (2017). Geologic influence on induced seismicity: Constraints from potential field data in Oklahoma. Geophysical Research Letters. 44(1): 152161.Google Scholar
Shapiro, SA and Dinske, C. (2009). Fluid‐induced seismicity: Pressure diffusion and hydraulic fracturing. Geophysical Prospecting. 57(2): 301310.Google Scholar
Shapiro, SA, Huenges, E, and Borm, G. (1997). Estimating the crust permeability from fluid-injection-induced seismic emission at the KTB site. Geophysical Journal International. 131(2): F15F18.Google Scholar
Shapiro, SA, Dinske, C, Langenbruch, C, and Wenzel, F. (2010). Seismogenic index and magnitude probability of earthquakes induced during reservoir fluid stimulations. The Leading Edge. 29(3): 304309.Google Scholar
Shapiro, SA, Krüger, OS, Dinske, C, and Langenbruch, C. (2011). Magnitudes of induced earthquakes and geometric scales of fluid-stimulated rock volumes. Geophysics. 76(6): WC55WC63.Google Scholar
Shapiro, SA, Patzig, R, Rothert, E, and Rindschwentner, J. (2003). Triggering of seismicity by pore-pressure perturbations: Permeability-related signatures of the phenomenon. Pure and Applied Geophysics. 160(5–6): 10511066.Google Scholar
Shapiro, SA, Rothert, E, Rath, V, and Rindschwentner, J. (2002). Characterization of fluid transport properties of reservoirs using induced microseismicity. Geophysics. 67(1): 212220.Google Scholar
Shearer, PM. (2019). Introduction to Seismology, 2nd Edition. Cambridge University Press.Google Scholar
Shemeta, JE, Brooks, CE, and Lord, CC. (2019). Well Stimulation Seismicity in Oklahoma: Cataloging Earthquakes Related to Hydraulic Fracturing. Unconventional Resources Technology Conference (URTEC), https://doi.org/10.15530/AP-URTEC-2019-198283.Google Scholar
Shipman, T, MacDonald, R, and Byrnes, T. (2018). Experiences and learnings from induced seismicity regulation in Alberta. Interpretation. 6(2): SE15SE21.Google Scholar
Skoumal, RJ, Brudzinski, MR, and Currie, BS. (2015). Earthquakes induced by hydraulic fracturing in Poland Township, Ohio. Bulletin of the Seismological Society of America. 105(1): 189197.Google Scholar
Skoumal, RJ, Ries, R, Brudzinski, MR, Barbour, AJ, and Currie, BS. (2018). Earthquakes induced by hydraulic fracturing are pervasive in Oklahoma. Journal of Geophysical Research: Solid Earth. 123(12): 1091810935.Google Scholar
Stoakes, FA. (1980). Nature and control of shale basin fill and its effect on reef growth and termination: upper Devonian Duvernay and Ireton formations of Alberta, Canada, Bulletin of Canadian Petroleum Geology. 28(3): 345410.Google Scholar
Switzer, SB, Holland, WG, Christie, DS, Graf, GC, Hedinger, AS, McAuley, RJ, Wierzbicki, RA, Packard, JJ, Mossop, GD, and Shetsen, I. (1994). Devonian Woodbend-Winterburn strata of the Western Canada Sedimentary Basin. In Mossop, GD and Shetsen, I (eds.) Geological Atlas of the Western Canada Sedimentary Basin. Canadian Society of Petroleum Geologists and Alberta Research Council, pp. 165202.Google Scholar
Tenthorey, E and Cox, SF. (2006). Cohesive strengthening of fault zones during the interseismic period: An experimental study. Journal of Geophysical Research: Solid Earth. 111: B09202, https://doi:10.1029/2005JB004122.Google Scholar
Townend, J and Zoback, MD. (2000). How faulting keeps the crust strong. Geology. 28(5): 399402.Google Scholar
Trutnevyte, E and Wiemer, S. (2017). Tailor-made risk governance for induced seismicity of geothermal energy projects: An application to Switzerland. Geothermics. 65: 295312.Google Scholar
Van der Baan, M and Calixto, F. (2017). Human-induced seismicity and large-scale hydrocarbon production in the USA and Canada. Geochemistry, Geophysics, Geosystems. 18(7): 24672485.Google Scholar
Van der Elst, NJ, Page, MT, Weiser, DA, Goebel, TH, and Hosseini, SM. (2016). Induced earthquake magnitudes are as large as (statistically) expected. Journal of Geophysical Research: Solid Earth. 121(6): 45754590.Google Scholar
Vermilye, JM and Scholz, CH. (1998). The process zone: A microstructural view of fault growth. Journal of Geophysical Research: Solid Earth. 103(B6): 1222312237.Google Scholar
Walsh, FR and Zoback, MD. (2015). Oklahoma’s recent earthquakes and saltwater disposal. Science Advances. 1(5): e1500195. https://doi.org/10.1126/sciadv.1500195Google Scholar
Walsh, FRI, Zoback, MD, Pais, D, Weingartern, M, and Tyrell, T. (2017). FSP 1.0: A Program for Probabilistic Estimation of Fault Slip Potential Resulting from Fluid Injection, scits.stanford.edu/software.Google Scholar
Wang, R, Gu, YJ, Schultz, R, and Chen, Y. (2018). Faults and non‐double‐couple components for induced earthquakes. Geophysical Research Letters. 45(17): 89668975.Google Scholar
Wang, R, Gu, YJ, Schultz, R, Kim, A, and Atkinson, G. (2016). Source analysis of a potential hydraulic‐fracturing‐induced earthquake near Fox Creek, Alberta. Geophysical Research Letters. 43(2): 564573.Google Scholar
Wang, R, Gu, YJ, Schultz, R, Zhang, M, and Kim, A. (2017). Source characteristics and geological implications of the January 2016 induced earthquake swarm near Crooked Lake, Alberta. Geophysical Journal International. 210(2): 979988.Google Scholar
Warpinski, NR, Mayerhofer, M, Agarwal, K, and Du, J. (2013). Hydraulic-fracture geomechanics and microseismic-source mechanisms. SPE Journal. 18(04): 766780.Google Scholar
Weingarten, M, Ge, S, Godt, JW, Bekins, BA, and Rubinstein, JL. (2015). High-rate injection is associated with the increase in US mid-continent seismicity. Science. 348(6241): 13361340.Google Scholar
Weir, RM, Eaton, DW, Lines, LR, Lawton, DC, and Ekpo, E. (2018). Inversion and interpretation of seismic-derived rock properties in the Duvernay play. Interpretation. 6(2): SE1SE14.Google Scholar
Wessel, P, Luis, JF, Uieda, L, Scharroo, R, Wobbe, F, Smith, WHF, and Tian, D. (2019). The Generic Mapping Tools version 6. Geochemistry, Geophysics, Geosystems. 20(1): 55565564.Google Scholar
Wilson, MP, Foulger, GR, Gluyas, JG, Davies, RJ, and Julian, BR. (2017). HiQuake: The human‐induced earthquake database. Seismological Research Letters. 88(6): 15601565.Google Scholar
Yao, Y. (2012). Linear elastic and cohesive fracture analysis to model hydraulic fracture in brittle and ductile rocks. Rock Mechanics and Rock Engineering. 45(3): 375387.Google Scholar
Yeck, WL, Weingarten, M, Benz, HM, McNamara, DE, Bergman, EA, Herrmann, RB, Rubinstein, JL, and Earle, PS. (2016). Far‐field pressurization likely caused one of the largest injection induced earthquakes by reactivating a large pre-existing basement fault structure. Geophysical Research Letters. 43(19): 10198.Google Scholar
Yu, H, Harrington, RM, Liu, Y, and Wang, B. (2019). Induced Seismicity Driven by Fluid Diffusion Revealed by a Near‐Field Hydraulic Stimulation Monitoring Array in the Montney Basin, British Columbia. Journal of Geophysical Research: Solid Earth. 124(5): 46944709.Google Scholar
Zhang, J. and Van der Baan, M. (2019). Depth-dependent fault slip potential. In SEG Technical Program Expanded Abstracts 2019 (pp. 30163020), Society of Exploration Geophysicists, https://doi.org/10.1190/segam2019–3214231.1.Google Scholar
Zhang, H, Eaton, DW, Rodriguez, G, and Jia, SQ. (2019). Source‐mechanism analysis and stress inversion for hydraulic‐fracturing‐induced event sequences near Fox Creek, Alberta. Bulletin of the Seismological Society of America. 109(2): 636651.Google Scholar
Zoback, MD. (2010). Reservoir Geomechanics, 1st edition, Cambridge University Press.Google Scholar
Zoback, MD and Harjes, HP. (1997). Injection‐induced earthquakes and crustal stress at 9 km depth at the KTB deep drilling site, Germany. Journal of Geophysical Research: Solid Earth. 102(B8): 1847718491.Google Scholar

References

Ajemigbitse, MA, Cannon, FS, Klima, MS, Furness, JC, Wunz, C, and Warner, NR. (2019). Raw material recovery from hydraulic fracturing residual solid waste with implications for sustainability and radioactive waste disposal. Environmental Science Process. Impacts. 21: 308323. https://doi.org/10.1039/c8em00248gGoogle Scholar
Akob, DM, Mumford, AC, Orem, W, Engle, MA, Klinges, JG, Kent, DB, and Cozzarelli, IM. (2016). Wastewater Disposal from Unconventional Oil and Gas Development Degrades Stream Quality at a West Virginia Injection Facility. Environmental Science & Technology. 50: 55175525. https://doi.org/10.1021/acs.est.6b00428Google Scholar
Alhajji, E, Al-Masri, MS, Khalily, H, Naoum, BE, Khalil, HS, and Nashawati, A. (2016). A Study on Sorption of 226Ra on Different Clay Matrices. Bulletin of Environmental Contaminants and Toxicology. 97: 255260. https://doi.org/10.1007/s00128–016-1852-1Google Scholar
Alnuaim, S.. (2018). What Does Sustainability Mean for Oil and Gas? Soc. Pet. Eng.Google Scholar
Back, W, Hanshaw, BB. (1970). Comparison of chemical hydrogeology of the carbonate peninsulas of Florida and Yucatan. Journal of Hydrology. 10: 330368. https://doi.org/10.1016/0022-1694(70)90222-2Google Scholar
Barry, B and Klima, MS. (2013). Characterization of Marcellus Shale natural gas well drill cuttings. Jouranl of Unconventional Oil and Gas Resources. 1–2: 917. https://doi.org/10.1016/J.JUOGR.2013.05.003Google Scholar
Bateman, H. (1910). The solution of a system of differential equations occurring in the theory of radio-active transformations. Cambridge Philosophical Society. 15: 423427.Google Scholar
Blewett, TA, Delompré, PLM, Glover, CN, and Goss, GG. (2018). Physical immobility as a sensitive indicator of hydraulic fracturing fluid toxicity towards Daphnia magna. Science of the Total Environment. 635: 639643. https://doi.org/10.1016/j.scitotenv.2018.04.165Google Scholar
Blewett, TA, Weinrauch, AM, Delompré, PLM, and Goss, GG. (2017). The effect of hydraulic flowback and produced water on gill morphology, oxidative stress and antioxidant response in rainbow trout (Oncorhynchus mykiss). Scientific Reports. 7: 46582. https://doi.org/10.1038/srep46582Google Scholar
Bloch, S and Key, RM. (1953). Modes of Formation of Anomalously High Radioactivity in Oil-Field Brines.Google Scholar
Blondes, MS, Gans, KD, Engle, MA, Kharaka, YK, Reidy, ME, Saraswathula, V, Thordsen, JJ, Rowan, EL, and Morrissey, EA. (2018). U.S. Geological Survey National Produced Waters Geochemical Database (ver. 2.3, January 2018) [WWW Document]. U.S. Geol. Surv. data release. https://doi.org/doi.org/10.5066/F7J964W8Google Scholar
Bolivar, JP, García-León, M, and García-Tenorio, R. (1997). On self-attenuation corrections in gamma-ray spectrometry. Applied Radiation and Isotopes. 48: 11251126. https://doi.org/10.1016/S0969–8043(97)00034-1Google Scholar
Brandt, F, Curti, E, Klinkenberg, M, Rozov, K, Bosbach, D. (2015). Replacement of barite by a (Ba,Ra)SO4 solid solution at close-to-equilibrium conditions: A combined experimental and theoretical study. Geochimica et Cosmochimica Acta. 155: 115. https://doi.org/10.1016/j.gca.2015.01.016Google Scholar
Burgos, WD, Castillo-Meza, L, Tasker, TL, Geeza, TJ, Drohan, PJ, Liu, X, Landis, JD, Blotevogel, J, McLaughlin, M, Borch, T, and Warner, NR. (2017). Watershed-scale impacts from surface water disposal of oil and gas wastewater in Western Pennsylvania. Environmental Science & Technology. 51: 88518860. https://doi.org/10.1021/acs.est.7b01696Google Scholar
Burkhardt, A, Gawde, A, Cantrell, CL, Baxter, HL, Joyce, BL, Stewart, CN, and Zheljazkov, VD. (2015). Effects of produced water on soil characteristics, plant biomass, and secondary metabolites. Journal of Environmental Quality. 44: 19381947. https://doi.org/10.2134/jeq2015.06.0299Google Scholar
Carvalho, F, Chambers, D, Fesenko, S, Moore, WS, Porcelli, D, Vandenhoven, H, and Yankovich, T. (2014). Environmental Pathways and Corresponding Models: The Environmental Behaviour of Radium, Revised Edition. Vienna.Google Scholar
Cavazza, M. (2016). Reducing freshwater consumption in the marcellus shale play by recycling flowback with acid mine drainage, in: Proceedings – SPE Annual Technical Conference and Exhibition. https://doi.org/10.2118/184499-stuGoogle Scholar
Clark, C and Veil, J. (2015). U.S. Produced water volumes and management practices. Groundwater Protection Council. 119.Google Scholar
Coleman, JL, Milici, RC, Cook, TA, Charpentier, RR, Kirschbaum, M, Klett, TR, Pollastro, RM, and Schenk, CJ. (2011). Assessment of Undiscovered Oil and Gas Resources of the Devonian Marcellus Shale of the Appalachian Basin Province, 2011.Google Scholar
Considine, T, D. Robert Watson, P, E. Seth Blumsack, P.. (2010). The economic impacts of the Pennsylvania Marcellus Shale natural gas play: An update.Google Scholar
Coonrod, CL, Yin, YB, Hanna, T, Atkinson, A, Alvarez, PJJ, Tekavec, TN, Reynolds, MA, and Wong, MS. (2020). Fit-for-purpose treatment goals for produced waters in shale oil and gas fields. Water Research. 173: 115467. https://doi.org/10.1016/j.watres.2020.115467Google Scholar
Cozzarelli, IM, Skalak, KJ, Kent, DB, Engle, MA, Benthem, A, Mumford, AC, Haase, K, Farag, A, Harper, D, Nagel, SC, Iwanowicz, LR, Orem, WH, Akob, DM, Jaeschke, JB, Galloway, J, Kohler, M, Stoliker, DL, and Jolly, GD. (2016). Environmental signatures and effects of an oil and gas wastewater spill in the Williston Basin, North Dakota. Science of the Total Environment. 579: 17811793. https://doi.org/10.1016/j.scitotenv.2016.11.157Google Scholar
Cravotta, CA. (2008a). Dissolved metals and associated constituents in abandoned coal-mine discharges, Pennsylvania, USA. Part 1: Constituent quantities and correlations. Applied Geochemistry. 23: 166202. https://doi.org/10.1016/j.apgeochem.2007.10.011Google Scholar
Cravotta, CA. (2008b). Dissolved metals and associated constituents in abandoned coal-mine discharges, Pennsylvania, USA. Part 2: Geochemical controls on constituent concentrations. Applied Geochemistry. 23: 203226. https://doi.org/10.1016/j.apgeochem.2007.10.003Google Scholar
Curti, E. (1999). Coprecipitation of radionuclides with calcite: Estimation of partition coefficients based on a review of laboratory investigations and geochemical data. Applied Geochemistry. 14: 433445. https://doi.org/10.1016/S0883–2927(98)00065-1Google Scholar
Curtright, AE and Giglio, K. (2012). Coal Mine Drainage for Marcellus Shale Natural Gas Extraction, Proceedings and Recommendations from a Roundtable on Feasibility and Challenges.Google Scholar
Dolan, FC, Cath, TY, and Hogue, TS. (2018). Assessing the feasibility of using produced water for irrigation in Colorado. Science of the Total Environment. 640–641: 619628. https://doi.org/10.1016/j.scitotenv.2018.05.200Google Scholar
Dorich, A. (2017). Eureka Resources recovers vital resources from flowback and production wastewater. [WWW Document]. Energy Min. Int. URL www.emi-magazine.com/sections/profiles/1460-eureka-resourcesGoogle Scholar
Dove, PM and Platt, FM. (1996). Compatible real-time rates of mineral dissolution by Atomic Force Microscopy (AFM). Chemical Geology. 127: 331338. https://doi.org/10.1016/0009-2541(95)00127-1Google Scholar
Dresel, P and Rose, A. (2010). Chemistry and origin of oil and gas well brines in western Pennsylvania. Pennsylvania Geological Survey. 48. https://doi.org/Open-File%20Report%20OFOG%201001.0Google Scholar
Dunne, EJ. (2017). Flowback and Produced Waters. National Academies Press. https://doi.org/10.17226/24620Google Scholar
Echchelh, A, Hess, T, Sakrabani, R, de Paz, JM, and Visconti, F. (2019). Assessing the environmental sustainability of irrigation with oil and gas produced water in drylands. Agricultural Water Management. 223. https://doi.org/10.1016/j.agwat.2019.105694Google Scholar
Eitrheim, ES, May, D, Forbes, TZ, and Nelson, AW. (2016). Disequilibrium of Naturally Occurring Radioactive Materials (NORM) in drill cuttings from a horizontal drilling operation. Environmental Science & Technology Letters. 3: 425429. https://doi.org/10.1021/acs.estlett.6b00439Google Scholar
Ellsworth, WL. (2013). Injection-Induced Earthquakes. Science. 341: 1225942. https://doi.org/10.1126/SCIENCE.1225942Google Scholar
Fisher, RS (1998) Geologic and geochemical controls on naturally occurring radioactive materials (NORM) in produced water from oil, gas, and geothermal operations. Environmental Geosciences. 5: 139150.Google Scholar
Fleischer, RL and Raabe, OG. (1978). Recoiling alpha-emitting nuclei. Mechanisms for uranium-series disequilibrium. Geochimica et Cosmochimica Acta. 42: 973978. https://doi.org/10.1016/0016-7037(78)90286-7Google Scholar
Flexer, V, Fernando Baspineiro, C, and Galli, CI. (2018). Lithium recovery from brines: A vital raw material for green energies with a potential environmental impact in its mining and processing. Science of the Total Environment. 639: 11881204. https://doi.org/10.1016/j.scitotenv.2018.05.223Google Scholar
Geeza, TJ, Gillikin, DP, McDevitt, B, Van Sice, K, and Warner, NR. (2018). Accumulation of Marcellus Formation oil and gas wastewater metals in freshwater mussel shells. Environmental Science & Technology. 52: 1088310892. https://doi.org/10.1021/acs.est.8b02727Google Scholar
Gilmore, GR. (2008). Practical Gamma-Ray Spectrometry. John Wiley & Sons, Ltd. https://doi.org/10.1002/9780470861981Google Scholar
Glynn, P. (2000). Solid-solution solubilities and thermodynamics: Sulfates, Carbonates and Halides: Reviews in Mineralogy and Geochemistry. 40: 481511. https://doi.org/10.2138/rmg.2000.40.10Google Scholar
Goodman, C. (2017). Beneficial use of produced water for roadspreading: perspectives for Colorado policymakers. Denver. https://doi.org/10.1017/CBO9781107415324.004Google Scholar
Grandia, F, Merino, J, andBruno, J. (2008). Assessment of the radium-barium co-precipitation and its potentialinfluence on the solubility of Ra in the near-field TR-08-07, 52.Google Scholar
Gregory, KB, Vidic, RD, and Dzombak, DA. (2011). Water management challenges associated with the production of shale gas by hydraulic fracturing. Elements. 7: 181186. https://doi.org/10.2113/gselements.7.3.181Google Scholar
Growitz, BDJ, Reed, LA, and Beard, MM. (1985). Reconnaissance of Mine Drainage in the Coal Fields of Eastern Pennsylvania. Harrisburg. USGS 83-4274 https://doi.org/10.3133/wri834274Google Scholar
Guerra, K, Dahm, K, and Dundorf, S. (2011). Oil and gas produced water management and beneficial use in the Western United States. Denver. https://doi.org/www.usbr.gov/pmts/water/publications/reports.htmlGoogle Scholar
Hanshaw, BB and Back, W. (1979). Major geochemical processes in the evolution of carbonate-aquifer systems. Developments in Water Science. 12: 287312. https://doi.org/10.1016/S0167–5648(09)70022-XGoogle Scholar
Harkness, JS, Dwyer, GS, Warner, NR, Parker, KM, Mitch, WA, and Vengosh, A. (2015). Iodide, bromide, and ammonium in hydraulic fracturing and oil and gas wastewaters: Environmental implications. Environmental Science & Technology. 49: 19551963. https://doi.org/10.1021/es504654nGoogle Scholar
He, C, Zhang, T, and Vidic, RD. (2013). Use of Abandoned Mine Drainage for the Development of Unconventional Gas Resources. Disruptive Science and Technology. 1: 169176. https://doi.org/10.1089/dst.2013.0014Google Scholar
He, C, Zhang, T, and Vidic, RD. (2016). Co-treatment of abandoned mine drainage and Marcellus Shale flowback water for use in hydraulic fracturing. Water Research. 104: 425431. https://doi.org/10.1016/j.watres.2016.08.030Google Scholar
He, C, Li, M, Liu, W, Barbot, E, and Vidic, RD. (2014). Kinetics and equilibrium of barium and strontium sulfate formation in Marcellus Shale flowback water. Journal of Environmental Engineering. 140: B4014001–19. https://doi.org/10.1061/(ASCE)EE.1943-7870.0000807Google Scholar
Hubbell, JH. (1982). Photon mass attenuation and energy-absorption coefficients. The International Journal of Applied Radiation and Isotopes. 33: 12691290. https://doi.org/10.1016/0020-708X(82)90248-4Google Scholar
International Atomic Energy Agency. (2014). The Environmental Behaviour of Radium, Revised Edition. Vienna.Google Scholar
Jang, Y and Chung, E. (2018). Adsorption of lithium from shale gas produced water using titanium based adsorbent. Industrial & Engineering Chemistry Research. 57: 83818387. https://doi.org/10.1021/acs.iecr.8b00805Google Scholar
Jodłowski, P, Macuda, J, Nowak, J, and Nguyen, Dinh C. (2017) Radioactivity in wastes generated from shale gas exploration and production: North-Eastern Poland. Journal of Environmental Radioactivity. 175–176: 3438. doi: 10.1016/j.jenvrad.2017.04.006.Google Scholar
Jones, MJ, Butchins, LJ, Charnock, JM, Pattrick, RAD, Small, JS, Vaughan, DJ, Wincott, PL, and Livens, FR. (2011). Reactions of radium and barium with the surfaces of carbonate minerals. Applied Geochemistry. 26: 12311238. https://doi.org/10.1016/j.apgeochem.2011.04.012Google Scholar
Kassotis, CD, Iwanowicz, LR, Akob, DM, Cozzarelli, IM, Mumford, AC, Orem, WH, and Nagel, SC. (2016). Endocrine disrupting activities of surface water associated with a West Virginia oil and gas industry wastewater disposal site. Science of the Total Environment. 557–558: 901910. https://doi.org/10.1016/j.scitotenv.2016.03.113Google Scholar
Kerr, RA. (2010). Energy: Natural gas from shale bursts onto the scene. Science. 328: 16241626. https://doi.org/10.1126/science.328.5986.1624Google Scholar
Kirby, HW and Salutsky, ML. (1964). The Radiochemistry of Radium. NAS-NRC Nucl. Sci. Ser.Google Scholar
Kondash, AJ, Albright, E, and Vengosh, A. (2017). Quantity of flowback and produced waters from unconventional oil and gas exploration. Science of the Total Environment. 574: 314321. https://doi.org/10.1016/j.scitotenv.2016.09.069Google Scholar
Kondash, AJ, Lauer, NE, and Vengosh, A. (2018). The intensification of the water footprint of hydraulic fracturing. Science Advances. 4. https://doi.org/10.1126/sciadv.aar5982Google Scholar
Kondash, AJ, Warner, NR, Lahav, O, and Vengosh, A. (2014a). Radium and barium removal through blending hydraulic fracturing fluids with acid mine drainage. Environmental Science & Technology. 48: 13341342. https://doi.org/10.1021/es403852hGoogle Scholar
Kondash, AJ, Warner, NR, Lahav, O, and Vengosh, A. (2014b). Radium and barium removal through blending hydraulic fracturing fluids with acid mine drainage. Environmental Science & Technology. 48: 13341342. https://doi.org/10.1021/es403852hGoogle Scholar
Kondash, AJ, Redmon, J, Lambertini, L, Feinstein, L, Weinthal, E, Cabrales, ., and Vengosh, A. (2020). The impact of using low-saline oilfield produced water for irrigation on water and soil quality in California, Science of The Total Environment, Volume 733, 139392.Google Scholar
Kraemer, TF and Reid, DF. (1984). The occurrence and behavior of radium in saline formation water of the U.S. Gulf Coast region. Chemical Geology. 46: 153174. https://doi.org/10.1016/0009-2541(84)90186-4Google Scholar
Krishnaswami, S, Graustein, WC, Turekian, KK, and Dowd, JF. (1982). Radium, thorium and radioactive lead isotopes in groundwaters: Application to the in situ determination of adsorption‐desorption rate constants and retardation factors. Water Resources Research. 18: 16631675. https://doi.org/10.1029/WR018i006p01663Google Scholar
Landis, JD, Sharma, M, Renock, D, and Niu, D. (2018). Rapid desorption of radium isotopes from black shale during hydraulic fracturing. 1. Source phases that control the release of Ra from Marcellus Shale, Chemical Geology, Volume 496, 2018, Pages 1-13.Google Scholar
Langmuir, D and Melchior, D. (1985). The geochemistry of Ca, Sr, Ba and Ra sulfates in some deep brines from the Palo Duro Basin, Texas. Geochimica et Cosmochimica Acta. 49: 24232432. https://doi.org/10.1016/0016-7037(85)90242-XGoogle Scholar
Lauer, NE, Harkness, JS, and Vengosh, A. (2016). Brine Spills Associated with Unconventional Oil Development in North Dakota. Environ. Sci. Technol. acs.est.5b06349. https://doi.org/10.1021/acs.est.5b06349Google Scholar
Lauer, NE, Warner, NR, and Vengosh, A. (2018). Sources of radium accumulation in stream sediments near disposal sites in Pennsylvania: implications for disposal of conventional oil and gas wastewater. Environmental Science & Technology. 52: 955962. https://doi.org/10.1021/acs.est.7b04952Google Scholar
Lowson, RT. (1985). The thermochemistry of radium. Thermochimica Acta. 91: 185212. https://doi.org/10.1016/0040-6031(85)85214-XGoogle Scholar
Lutz, BD, Lewis, AN, and Doyle, MW (2013). Generation, transport, and disposal of wastewater associated with Marcellus Shale gas development. Water Resources Research. 49: 647656. https://doi.org/10.1002/wrcr.20096Google Scholar
Lyons, W, Plisga, G, and Lorenz, M. (2015). Standard Handbook of Petroleum and Natural Gas Engineering, 3rd ed. Gulf Professional Publishing.Google Scholar
Lyons, WC, Pilsga, GJ, and Lorenz, MD. (2016). Standard Handbook of petroleum and Natural Gas Engineering. Gulf Professional Publishing.Google Scholar
Maloney, JA. (1937a). Radium-Nature’s Oddest Child Pt. 4. 157, 212215. https://doi.org/10.2307/26070913Google Scholar
Maloney, JA. (1937b). Radium-Nature’s Oddest Child. 157: 1820. https://doi.org/10.2307/24997429Google Scholar
Maloney, KO and Yoxtheimer, DA. (2012). Production and disposal of waste materials from gas and oil extraction from the Marcellus Shale play in Pennsylvania. Environmental Practice. 14: 278287. https://doi.org/10.1017/S146604661200035XGoogle Scholar
McDevitt, B, Cavazza, M, Beam, R, Cavazza, ED, Burgos, W, Li, L, and Warner, NR. (2020a). Maximum removal efficiency of barium, strontium, radium, and sulfate with optimum AMD-Marcellus flowback mixing ratios for beneficial use in the Northern Appalachian Basin. Environmental Science & Technology. 54: 48294839. https://doi.org/10.1021/acs.est.9b07072Google Scholar
McDevitt, B, McLaughlin, MC, Vinson, DS, Geeza, TJ, Blotevogel, J, Borch, T, and Warner, NR. (2020b). Isotopic and element ratios fingerprint salinization impact from beneficial use of oil and gas produced water in the Western U.S. Science of the Total Environment. 716. https://doi.org/10.1016/j.scitotenv.2020.137006Google Scholar
McDevitt, B, Mclaughlin, M, Cravotta, CA, Ajemigbitse, MA, Sice, KJ, Van Blotevogel, J, Borch, T, and Warner, NR. (2018). Emerging investigator series: radium accumulation in carbonate river sediments at oil and gas produced water discharges: implications for beneficial use as disposal management. Environmental Science: Processes & Impacts. 21: 324338. https://doi.org/10.1039/c8em00336jGoogle Scholar
McDevitt, B., McLaughlin, M. C., Blotevogel, J., Borch, T., & Warner, N. R. (2021). Oil & gas produced water retention ponds as potential passive treatment for radium removal and beneficial reuse. Environmental Science: Processes & Impacts. 23(3), 501518.Google Scholar
McLaughlin, M, Borch, T, McDevitt, B, Warner, NR, and Blotevogel, J. (2020). Water quality assessment downstream of oil and gas produced water discharges intended for beneficial reuse in arid regions. Science of the Total Environment. 713: 136607.Google Scholar
McLaughlin, MC, Blotevogel, J, Watson, RA, Schell, B, Blewett, TA, Folkerts, EJ, Goss, GG, Truong, L, Tanguay, RL, Lucas, J, and Borch, T. (2020). Mutagenicity assessment downstream of oil and gas produced water discharges intended for agricultural beneficial reuse. Science of the Total Environment. 715: 136944. https://doi.org/10.1016/j.scitotenv.2020.136944Google Scholar
MGX Minerals Inc. (2019). MGX Minerals and Eureka Resources Announce Joint Venture to Recover Lithium from Produced Water in Eastern United States [WWW Document]. PW Newswire.Google Scholar
Montgomery, CT and Smith, MB. (2010). Hydraulic fracturing: History of an enduring technology. Journal of Petroleum Technology. 62: 2640. https://doi.org/10.2118/1210-0026-JPTGoogle Scholar
Nathwani, JS and Phillips, CR. (1979). Adsorption of 226Ra by soils in the presence of Ca2+ ions. Specific adsorption (II). Chemosphere. 8: 293299. https://doi.org/10.1016/0045-6535(79)90112-7Google Scholar
National Research Council. (1999). Health Effects of Exposure to Radon. National Academies Press. https://doi.org/10.17226/5499Google Scholar
Nelson, AW May, D, Knight, AW, Eitrheim, ES, Mehrhoff, M, Shannon, R, Litman, R, Schultz, MK. (2014). Matrix Complications in the Determination of Radium Levels in Hydraulic Fracturing Flowback Water from Marcellus Shale. Environmental Science & Technology Letters. 1: 204208. https://doi.org/10.1021/ez5000379Google Scholar
Nelson, AW, Eitrheim, ES, Knight, AW, May, D, Mehrhoff, MA, Shannon, R, Litman, R, Burnett, WC, Forbes, TZ, and Schultz, MK. (2015). Understanding the radioactive ingrowth and decay of naturally occurring radioactive materials in the environment: An analysis of produced fluids from the marcellus shale. Environ. Health Perspect. https://doi.org/10.1289/ehp.1408855Google Scholar
Ouyang, B, Akob, DM, Dunlap, D, and Renock, D. (2017). Microbially mediated barite dissolution in anoxic brines. Applied Geochemistry. 76: 5159. https://doi.org/10.1016/j.apgeochem.2016.11.008Google Scholar
Paranthaman, MP, Li, L, Luo, J, Hoke, T, Ucar, H, Moyer, BA, and Harrison, S. (2017). Recovery of lithium from geothermal brine with lithium–aluminum layered double hydroxide chloride sorbents. Environmental Science & Technology. 51: 1348113486. https://doi.org/10.1021/acs.est.7b03464Google Scholar
Parker, KM, Zeng, T, Harkness, J, Vengosh, A, and Mitch, WA (2014). Enhanced formation of disinfection byproducts in shale gas wastewater-impacted drinking water supplies. Environmental Science & Technology. 48, 1116111169. https://doi.org/10.1021/es5028184Google Scholar
Paukert Vankeuren, AN, Hakala, JA, Jarvis, K, and Moore, JE. (2017). Mineral reactions in shale gas reservoirs: Barite scale formation from reusing produced water as hydraulic fracturing fluid. Environmental Science & Technology. 51: 93919402. https://doi.org/10.1021/acs.est.7b01979Google Scholar
Pedersen, KS, Christensen, PL, and Shaikh, JA. (2015). Phase Behavior of Petroleum Reservoir Fluids. CRC Press.Google Scholar
PA DEP. (2016). 2016 Oil and Gas Annual Report. https://gis.dep.pa.gov/oilgasannualreport/index.htmlGoogle Scholar
Pennsylvania Department of Environmental Protection. (2016). Technologically Enhanced Naturally Occurring Radioactive Materials (TENORM) Study Report. www.depgreenport.state.pa.us/elibrary/getdocument?docid=5815&docname=01%20pennsylvania%20department%20of%20environmental%20protection%20tenorm%20study%20report%20rev%201.pdfGoogle Scholar
Phillips, EJP, Landa, ER, Kraemer, T, and Zielinski, R. (2001). Sulfate-reducing bacteria release barium and radium from naturally occurring radioactive material in oil-field barite. Geomicrobiology Journal. 18: 167182. https://doi.org/10.1080/01490450120549Google Scholar
Pichtel, J. (2016). Oil and gas production wastewater: Soil contamination and pollution prevention. Applied Environmental Soil Science. 2016: 124. https://doi.org/10.1155/2016/2707989Google Scholar
Ramirez, PJ. (2002). Oil Field Produced Water Discharges into Wetlands in Wyoming. US Fish Wildl. Serv. Report.Google Scholar
Renock, D, Landis, JD, and Sharma, M. (2016). Reductive weathering of black shale and release of barium during hydraulic fracturing. Applied Geochemistry. 65: 7386. https://doi.org/10.1016/j.apgeochem.2015.11.001Google Scholar
Röntgen, WC. (1970). On a new kind of rays. By W.C. Rontgen. Translated by Arthur Stanton from the Sitzungsberichte der Würzburger Physic-medic. Gesellschaft, 1895. Nature, January 23, 1896. Radiography. 36: 185188.Google Scholar
Rosenblum, J, Nelson, AW, Ruyle, B, Schultz, MK, Ryan, JN, Linden, KG. (2017). Temporal characterization of flowback and produced water quality from a hydraulically fractured oil and gas well. Science of the Total Environment. 596–597: 369377. https://doi.org/10.1016/j.scitotenv.2017.03.294Google Scholar
Rowan, EL, Engle, Ma, Kirby, CS, Kraemer, TF. (2011a). Radium Content of Oil- and Gas-Field Produced Waters in the Northern Appalachian Basin (USA): Summary and Discussion of Data. USGS Sci. Investig. Rep. 38 pp.Google Scholar
Rowan, EL, Engle, Ma, Kirby, CS, Kraemer, TF. (2011b). Radium Content of Oil- and Gas-Field Produced Waters in the Northern Appalachian Basin (USA): Summary and Discussion of Data. USGS Sci. Investig. Rep. 38 pp.Google Scholar
Rowan, EL, Engle, MA, Kraemer, TF, Schroeder, KT, Hammack, RW, and Doughten, MW. (2015). Geochemical and isotopic evolution of water produced from Middle Devonian Marcellus shale gas wells, Appalachian basin, Pennsylvania: American Association of Petroleum Geological Bulletin. 99: 181206. https://doi.org/10.1306/07071413146Google Scholar
Rutherford, E. (1899). VIII. Uranium radiation and the electrical conduction produced by it. London, Edinburgh, Dublin Philosophical Magazine and Journal of Science. 47: 109163. https://doi.org/10.1080/14786449908621245Google Scholar
Sajih, M, Bryan, ND, Livens, FR, Vaughan, DJ, Descostes, M, Phrommavanh, V, Nos, J, and Morris, K. (2014). Adsorption of radium and barium on goethite and ferrihydrite: A kinetic and surface complexation modelling study. Geochimica et Cosmochimica Acta. 146: 150163. https://doi.org/10.1016/j.gca.2014.10.008Google Scholar
Scanlon, BR, Reedy, RC, and Nicot, JP. (2014). Will water scarcity in semiarid regions limit hydraulic fracturing of shale plays? Environmental Research Letters. 9. https://doi.org/10.1088/1748-9326/9/12/124011Google Scholar
Schaller, J, Headley, T, Prigent, S, and Breuer, R. (2014). Potential mining of lithium, beryllium and strontium from oilfield wastewater after enrichment in constructed wetlands and ponds. Science of the Total Environment. 493: 910913. https://doi.org/10.1016/j.scitotenv.2014.06.097Google Scholar
Sedlacko, EM, Jahn, CE, Heuberger, AL, Sindt, NM, Miller, HM, Borch, T, Blaine, AC, Cath, TY, and Higgins, CP. (2019). Potential for beneficial reuse of oil‐and‐gas‐derived produced water in agriculture: Physiological and morphological responses in spring wheat ( Triticum aestivum). Environmental Toxicology and Chemistry. 4449. https://doi.org/10.1002/etc.4449Google Scholar
Sekhran, N, Sheahan, B, and Sullivan, B. (2017). Mapping the oil and gas industry to the SDGs: An Atlas. New York.Google Scholar
Shao, H, Kulik, DA, Berner, U, Kosakowski, G, and Kolditz, O. (2009). Modeling the competition between solid solution formation and cation exchange on the retardation of aqueous radium in an idealized bentonite column. Geochemical Journal. 43: e37-e42. https://doi.org/10.2343/geochemj.1.0069Google Scholar
Soddy, F. (1913). Intra-atomic Charge. Nature. 92: 399400.Google Scholar
Stallworth, A.M., Chase, E.H., Burgos, W.D., Warner, N.R. (2019). Laboratory Method to Assess Efficacy of Dust Supressants for Dirt and Gravel Roads, in: Transportation Research Record. Transportation Research Board.Google Scholar
Stewart, BW, Chapman, EC, Capo, RC, Johnson, JD, Graney, JR, Kirby, CS, and Schroeder, KT. (2015). Origin of brines, salts and carbonate from shales of the Marcellus Formation: Evidence from geochemical and Sr isotope study of sequentially extracted fluids. Applied Geochemistry. 60: 7888. https://doi.org/10.1016/j.apgeochem.2015.01.004Google Scholar
Swain, B. (2017). Recovery and recycling of lithium: A review. Separation and Purification Technology. 172: 388403. https://doi.org/10.1016/j.seppur.2016.08.031Google Scholar
Tasker, TL, Burgos, WD, Ajemigbitse, MA, Lauer, NE, Gusa, AV, Kuatbek, M, May, D, Landis, JD, Alessi, DS, Johnsen, AM, Kaste, JM, Headrick, KL, Wilke, FDH, McNeal, M, Engle, M, Jubb, AM, Vidic, RD, Vengosh, A, and Warner, NR. (2019). Accuracy of methods for reporting inorganic element concentrations and radioactivity in oil and gas wastewaters from the Appalachian Basin, U.S. based on an inter-laboratory comparison. Environmental Science Processes & Impacts. 21: 224241. https://doi.org/10.1039/C8EM00359AGoogle Scholar
Tasker, TL, Burgos, WD, Piotrowski, P, Castillo-Meza, L, Blewett, TA, Ganow, KB, Stallworth, A, Delompré, PLM, Goss, GG, Fowler, LB, Vanden Heuvel, JP, Dorman, F, and Warner, NR. (2018). Environmental and human health impacts of spreading oil and gas wastewater on roads. Environmental Science & Technology. 52: 70817091. https://doi.org/10.1021/acs.est.8b00716Google Scholar
Tesoriero, AJ and Pankow, JF. (1996). Solid solution partitioning of Sr2+, Ba2+, and Cd2+to calcite. Geochimica et Cosmochimica Acta. 60: 10531063. https://doi.org/10.1016/0016-7037(95)00449-1Google Scholar
US Energy Information Administration (US EIA). (2015). Top 100 U.S. Oil and Gas Fields. www.eia.gov/naturalgas/crudeoilreserves/top100/pdf/top100.pdfGoogle Scholar
US EIA. (2018). Annual energy outlook. www.eia.gov/outlooks/aeo/section_issues.phpGoogle Scholar
US EIA. (2019). Appalachia Region Drilling Productivity Report. www.eia.gov/petroleum/drilling/pdf/appalachia.pdfGoogle Scholar
US Environmental Protection Agency (US EPA). (1976). National Interim Primary Drinking Water Regulations. Washington, DC.Google Scholar
US EPA. (2005). A System’s Guide to the Management of Radioactive Residuals from Drinking Water. Epa 816-R-05-004.Google Scholar
US EPA. (2018a). Detailed Study of the Centralized Waste Treatment Point Source Category for Facilities Managing Oil and Gas Extraction Wastes. Washington, DC.Google Scholar
US EPA. (2018b). Study of Oil and Gas Extraction Wastewater Management. www.epa.gov/eg/study-oil-and-gas-extraction-wastewater-managementGoogle Scholar
US EPA (2019). Study of Oil and Gas Extraction Wastewater Management Under the Clean Water Act EPA-821-R19–001. Washington, DC.Google Scholar
United Nations. (2011). Sources and Effects of Ionizing Radiation United Nations Scientific Committee on the Effects of Atomic Radiation. New York.Google Scholar
US Nuclear Regulatory Commission. (2017). Appendix B to Part 20 Annual Limits on Intake (ALIs) and Derived Air Concentrations (DACs) of Radionuclides for Occupational Exposure; Effluent Concentrations; Concentrations for Release to Sewerage.Google Scholar
Van Sice, K, Cravotta, CA, McDevitt, B, Tasker, TL, Landis, JD, Puhr, J, and Warner, NR. (2018). Radium attenuation and mobilization in stream sediments following oil and gas wastewater disposal in western Pennsylvania. Applied Geochemistry. 98: 393403. https://doi.org/10.1016/j.apgeochem.2018.10.011Google Scholar
Veil, J. (2015). U.S. Produced Water Volumes and Management Practices in 2012.Google Scholar
Vengosh, A, Jackson, RB, Warner, N, Darrah, TH, and Kondash, A. (2014a). A critical review of the risks to water resources from unconventional shale gas development and hydraulic fracturing in the United States. Environmental Science & Technology. 48: 83348348. https://doi.org/10.1021/es405118yGoogle Scholar
Vengosh, A, Jackson, RB, Warner, N, Darrah, TH, and Kondash, A. (2014b). A Critical Review of the Risks to Water Resources from Unconventional Shale Gas Development and Hydraulic Fracturing in the United States. Environmental Science & Technology. 48: 83348348. https://doi.org/10.1021/es405118yGoogle Scholar
Vinograd, VL, Brandt, F, Rozov, K, Klinkenberg, M, Refson, K, Winkler, B, and Bosbach, D. (2013). Solid-aqueous equilibrium in the BaSO4-RaSO4-H2O system: First-principles calculations and a thermodynamic assessment. Geochimica et Cosmochimica Acta. 122: 398417. https://doi.org/10.1016/j.gca.2013.08.028Google Scholar
Vinograd, VL, Kulik, DA, Brandt, F, Klinkenberg, M, Weber, J, Winkler, B, and Bosbach, D. (2018a). Thermodynamics of the solid solution – Aqueous solution system (Ba,Sr,Ra)SO4+ H2O: I. The effect of strontium content on radium uptake by barite. Applied Geochemistry. 89: 5974. https://doi.org/10.1016/j.apgeochem.2017.11.009Google Scholar
Vinograd, VL, Kulik, DA, Brandt, F, Klinkenberg, M, Weber, J, Winkler, B, and Bosbach, D. (2018b). Thermodynamics of the solid solution – Aqueous solution system (Ba,Sr,Ra)SO4+ H2O: II. Radium retention in barite-type minerals at elevated temperatures. Applied Geochemistry. 93: 190208. https://doi.org/10.1016/j.apgeochem.2017.10.019Google Scholar
Vinson, DS, Lundy, JR, Dwyer, GS, and Vengosh, A. (2012). Implications of carbonate-like geochemical signatures in a sandstone aquifer: Radium and strontium isotopes in the Cambrian Jordan aquifer (Minnesota, USA). Chemical Geology. 334: 280294. https://doi.org/10.1016/j.chemgeo.2012.10.030Google Scholar
Wang, Y, Tavakkoli, S, Khanna, V, Vidic, RD, and Gilbertson, LM. (2018). Life cycle impact and benefit trade-offs of a produced water and abandoned mine drainage cotreatment process. Environmental Science & Technology. 52: 1399514005. https://doi.org/10.1021/acs.est.8b03773Google Scholar
Warner, NR, Christie, CA, Jackson, RB, and Vengosh, A. (2013a). Impacts of shale gas wastewater disposal on water quality in Western Pennsylvania. Environmental Science & Technology. 47: 1184911857. https://doi.org/10.1021/es402165bGoogle Scholar
Warner, NR, Christie, CA, Jackson, RB, and Vengosh, A. (2013b). Impacts of shale gas wastewater disposal on water quality in Western Pennsylvania. Environmental Science & Technology. 47: 1184911857. https://doi.org/10.1021/es402165bGoogle Scholar
Webster, IT, Hancock, GJ, and Murray, AS. (1995). Modelling the effect of salinity on radium desorption from sediments. Geochimica et Cosmochimica Acta. 59: 24692476. https://doi.org/10.1016/0016-7037(95)00141-7Google Scholar
Wiegand, JW and Sebastian, F. (2002). Origin of radium in high-mineralised waters. Technol. Enhanc. Nat. Radiat. (TENR II) - Proc. an Int. Symp. Held Rio Janeiro; IAEA - TECDOC - 1271 107–111.Google Scholar
Wilson, JM and Vanbriesen, JM. (2012). Oil and gas produced water management and surface Pennsylvania. Environmental Practice. 14: 288301.Google Scholar
Zhang, T, Gregory, K, Hammack, RW, and Vidic, RD. (2014) Co-precipitation of radium with barium and strontium sulfate and its impact on the fate of radium during treatment of produced water from unconventional gas extraction. Environmental Science & Technology. 48: 45964603. https://doi.org/10.1021/es405168bGoogle Scholar
Zhang, T, Bain, D, Hammack, R, and Vidic, RD. (2015a). Analysis of Radium-226 in High Salinity Wastewater from Unconventional Gas Extraction by Inductively Coupled Plasma-Mass Spectrometry. Environmental Science & Technology. 49: 29692976. https://doi.org/10.1021/es504656qGoogle Scholar
Zhang, T, Hammack, RW, and Vidic, RD. (2015b). Fate of radium in Marcellus Shale flowback water impoundments and assessment of associated health risks. Environmental Science & Technology. 49: 93479354. https://doi.org/10.1021/acs.est.5b01393Google Scholar

References

Andersson, PS, Wasserburg, GJ, and Ingri, J. (1992). The sources and transport of Sr and Nd isotopes in the Baltic Sea. Earth and Planetary Science Letters. 113: 459472.Google Scholar
Bagheri, R, Nadri, A, Raeisi, E, Eggenkamp, HGM, Kazemi, GA, and Montaseri, A. (2014). Hydrochemical and isotopic (δ18O, δ2H, 87Sr/86Sr, δ37Cl and δ81Br) evidence for the origin of saline formation water in a gas reservoir. Chemical Geology. 384: 6275.Google Scholar
Balashov, VN, Engelder, T, Gu, X, Fantle, MS, and Brantley, SL. (2015). A model describing flowback chemistry changes with time after Marcellus Shale hydraulic fracturing. American Association of Petroleum Geologists Bulletin. 99: 143154.Google Scholar
Banner, JL. (2004). Radiogenic isotopes: Systematics and applications to earth surface processes and chemical stratigraphy. Earth Science Reviews. 65: 141194.Google Scholar
Barbot, E, Vidic, NS, Gregory, KB, and Vidic, RD. (2013). Spatial and temporal correlation of water quality parameters of produced waters from Devonian-age shale following hydraulic fracturing. Environmental Science and Technology. 47: 25622569.Google Scholar
Bloomberg.com. (2019). MGX Minerals and Eureka Resources Announce Operation of First Commercial Rapid Petrolithium Recovery System in Pennsylvania. www.bloomberg.com/press-releases/2019-10-24/mgx-minerals-and-eureka-resources-announce-operation-of-first-commercial-rapid-petrolithium-recovery-system-in-pennsyl.Google Scholar
Böttcher, ME, Geprägs, P, Neubert, N, von Allmen, K, Pretet, C, Samankassou, E, and Nägler, TF. (2012). Barium isotope fractionation during experimental formation of the double carbonate BaMn[CO3]2 at ambient temperature. Isotopes in Environmental and Health Studies. 48: 457463.Google Scholar
Bottomley, DJ, Chan, LH, Katz, A, Starinsky, A, and Clark, ID. (2003). Lithium isotope geochemistry and origin of Canadian Shield brines. Ground Water. 41, 847856.Google Scholar
Brandt, JE, Lauer, NE, Vengosh, A, Bernhardt, ES, and Di Giulio, RT. (2018). Strontium isotope ratios in fish otoliths as biogenic tracers of coal combustion residual inputs to freshwater ecosystems. Environmental Science & Technology Letters. 5: 718723.Google Scholar
Brantley, SL, Yoxtheimer, D, Arjmand, S., Grieve, P., Vidic, R., Pollack, J., Llewellyn, GT, Abad, J, and Simon, S. (2014). Water resource impacts during unconventional shale gas development: The Pennsylvania experience. International Journal of Coal Geology. 126: 140-156.Google Scholar
Brinck, EL and Frost, CD. (2007). Detecting infiltration and impacts of introduced water using strontium isotopes. Ground Water. 45: 554568.Google Scholar
Bullen, TD. (2014). Metal stable isotopes in weathering and hydrology. In Holland, HD and Turekian, KK (eds.) Treatise on Geochemistry 7. Elsevier, 329359.Google Scholar
Bullen, TD and Eisenhauer, A. (2009). Metal stable isotopes in low-temperature systems: A primer. Elements. 5: 349352.Google Scholar
Bullen, T and Chadwick, O. (2016). Ca, Sr and Ba stable isotopes reveal the fate of soil nutrients along a tropical climosequence in Hawaii. Chemical Geology. 422: 2545.Google Scholar
Burgos, WD, Castillo-Meza, L, Tasker, TL, Geeza, TJ, Drohan, PJ, Liu, X, Landis, JD, Blotevogel, J, McLaughlin, M, Borch, T, and Warner, NR. (2017). Watershed-scale impacts from surface water disposal of oil and gas wastewater in western Pennsylvania. Environmental Science and Technology. 51: 88518860.Google Scholar
Campbell, CE, Pearson, BN, and Frost, CD. (2008). Strontium isotopes as indicators of aquifer communication in an area of coal bed natural gas production, Powder River Basin, Wyoming and Montana. Rocky Mountain Geology. 43: 149175.Google Scholar
Capo, RC, Stewart, BW, and Chadwick, OA. (1998). Strontium isotopes as tracers of ecosystem processes: Theory and methods. Geoderma. 82: 197225.Google Scholar
Capo, RC, Stewart, BW, Rowan, EL, Kolesar Kohl, CA, Wall, AJ, Chapman, EC, Hammack, RW, and Schroeder, KT. (2014). The strontium isotopic evolution of Marcellus Formation produced waters, southwestern Pennsylvania. International Journal of Coal Geology. 126: 5763.Google Scholar
Chan, L-H, Starinsky, A, and Katz, A. (2002). The behavior of lithium and its isotopes in oilfield brines: Evidence from the Heletz–Kokhav field, Israel. Geochimica et Cosmochimica Acta. 66: 615623.Google Scholar
Chan, LH, Edmond, JM, Thompson, G, and Gillis, K. (1992). Lithium isotopic composition of submarine basalts: implications for the lithium cycle in the oceans. Earth and Planetary Science Letters. 108: 151160.Google Scholar
Chan, L-H, Gieskes, JM, You, C-F, and Edmond, JM. (1994). Lithium isotope geochemistry of sediments and hydrothermal fluids of the Guaymas Basin, Gulf of California. Geochimica et Cosmochimica Acta. 58: 44434454.Google Scholar
Chapman, EC, Capo, RC, Stewart, BW, Kirby, CS, Hammack, RW, Schroeder, KT, and Edenborn, H.M. (2012). Geochemical and strontium isotope characterization of produced waters from Marcellus Shale natural gas extraction. Environmental Science and Technology. 46: 35453553.Google Scholar
Chaudhuri, S and Clauer, N. (1993). Stontium isotopic compositions and potassium and rubidium contents of formation waters in sedimentary basins: Clues to the origin of the solutes. Geochimica et Cosmochimica Acta. 57: 429437.Google Scholar
Choi, H-B, Ryu, J-S, Shin, W-J, Vigier, N. (2019). The impact of anthropogenic inputs on lithium content in river and tap water. Nature Communications. 105371. https://doi.org/10.1038/s41467-019-13376-y.Google Scholar
Collins, AG. (1978). Geochemistry of anomalous lithium in oil-field brines. Oklahoma Geological Survey Circular. 79: 9598.Google Scholar
DePaolo, DJ and Wasserburg, GJ. (1976). Nd isotopic variations and petrogenetic models. Geophysical Research Letters. 3: 249252.Google Scholar
DOE-NETL. (2010). Carbon dioxide enhance oil recovery: Untapped domestic energy supply and long term carbon storage solution. Department of Energy – National Energy Technology Laboratory, www.netl.doe.gov/file%20library/research/oil-gas/small_CO2_EOR_Primer.pdf.Google Scholar
Eiler, JM, Bergquist, BA, Bourg, IC, Cartigny, P, Farquhar, J, Gagnon, A, Guo, W, Halevy, I, and Hofmann, A et al. (2014). Frontiers of stable isotope geoscience. Chemical Geology. 372: 119143.Google Scholar
Engle, MA and Rowan, EL. (2013). Interpretation of Na–Cl–Br systematics in sedimentary basin brines: Comparison of concentration, element ratio, and isometric log-ratio approaches. Mathematical Geosciences. 45: 87101.Google Scholar
Engle, MA and Rowan, EL. (2014). Geochemical evolution of produced waters from hydraulic fracturing of the Marcellus Shale, northern Appalachian Basin: A multivariate compositional data analysis approach. International Journal of Coal Geology. 126: 4556.Google Scholar
Eureka Resources. (2019). MGX Minerals and Eureka Resources announce joint venture to recover lithium from produced water in eastern United States. www.eureka-resources.com/blog1.Google Scholar
Faure, G and Mensing, TM. (2005). Isotopes: Principles and Applications. John Wiley & Sons.Google Scholar
Ferrar, KJ, Michanowicz, DR, Christen, CL, Mulcahy, N, Malone, SL, and Sharma, RK. (2013). Assessment of effluent contaminants from three facilities discharging Marcellus shale wastewater to surface waters in Pennsylvania. Environmental Science and Technology. 47: 34723481.Google Scholar
Frost, CD, Pearson, BN, Ogle, KM, Heffern, EL, and Lyman, RM. (2002). Sr isotope tracing of aquifer interactions in an area of accelerating coal-bed methane production, Powder River Basin, Wyoming. Geology. 30: 923926.Google Scholar
Geeza, TJ, Gillikin, DP, McDevitt, B, Van Sice, K, and Warner, NR. (2018). Accumulation of Marcellus Formation oil and gas wastewater metals in freshwater mussel shells. Environmental Science & Technology. 52: 1088310892.Google Scholar
Haluszczak, LO, Rose, AW, and Kump, LR. (2013). Geochemical evaluation of flowback brine from Marcellus gas wells in Pennsylvania, USA. Applied Geochemistry. 28: 5561.Google Scholar
Hammack, R, Zorn, E, Harbert, W, Capo, R, Sharma, S, and Siriwardane, H. (2013). An evaluation of zonal isolation after hydraulic fracturing; Results from horizontal Marcellus Shale gas wells at NETL’s Greene County Test Site in southwestern Pennsylvania. Society of Petroleum Engineers Conference Paper DOI: 10.2118/165720-MS, SPE-165720-MS.Google Scholar
Hayes, T. (2009). Sampling and Analysis of Water Streams Associated with the Development of Marcellus Shale Gas. Report by the Gas Technology Institute, Des Plaines, IL. Marcellus Shale Coalition.Google Scholar
Hemsing, F, Hsieh, Y-T, Bridgestock, L, Spooner, PT, Robinson, LF, Frank, N, and Henderson, GM. (2018). Barium isotopes in cold-water corals. Earth and Planetary Science Letters. 491: 183192.Google Scholar
Hsieh, Y-T and Henderson, GM. (2017). Barium stable isotopes in the global ocean: Tracer of Ba inputs and utilization. Earth and Planetary Science Letters. 473: 269278.Google Scholar
Huang, T, Pang, Z, Li, Z, Li, Y, and Hao, Y. (2020). A framework to determine sensitive inorganic monitoring indicators for tracing groundwater contamination by produced formation water from shale gas development in the Fuling Gasfield, SW China. Journal of Hydrology. 581: 124403.Google Scholar
Huh, Y, Chan, L-H, Zhang, L, and Edmond, JM. (1998). Lithium and its isotopes in major world rivers: Implications for weathering and the oceanic budget. Geochimica et Cosmochimica Acta. 62: 20392051.Google Scholar
Johnson, CM, Beard, BL, and Albarède, F. (2004). Geochemistry of non-traditional stable isotopes. In Rosso, JJ (ed.) Reviews in Mineralogy & Geochemistry. Mineralogical Society of America, p. 454.Google Scholar
Johnson, JD, Graney, JR, Capo, RC, and Stewart, BW. (2015). Identification and quantification of regional brine and road salt sources in watersheds along the New York / Pennsylvania border, USA. Applied Geochemistry. 60: 3750.Google Scholar
Kolesar Kohl, CA, Capo, RC, Stewart, BW, Wall, AJ, Schroeder, KT, Hammack, RW, and Guthrie, GD. (2014). Strontium isotopes test long-term zonal isolation of injected and Marcellus Formation water after hydraulic fracturing. Environmental Science and Technology. 48: 98679873.Google Scholar
Lauer, NE, Harkness, JS, and Vengosh, A. (2016). Brine Spills Associated with Unconventional Oil Development in North Dakota. Environmental Science & Technology. 50: 53895397.Google Scholar
Lauer, NE, Warner, NR, and Vengosh, A. (2018). Sources of radium accumulation in stream sediments near disposal sites in Pennsylvania: Implications for disposal of conventional oil and gas wastewater. Environmental Science and Technology 52, 955-962.Google Scholar
Macpherson, GL, Capo, RC, Stewart, BW, Phan, TT, Schroeder, KT, and Hammack, RW. (2014). Temperature-dependent Li isotope ratios in Appalachian Plateau and Gulf Coast Sedimentary Basin saline water. Geofluids. 14: 419429.Google Scholar
Mavromatis, V, van Zuilen, K, Blanchard, M, van Zuilen, M, Dietzel, M, and Schott, J. (2020). Experimental and theoretical modelling of kinetic and equilibrium Ba isotope fractionation during calcite and aragonite precipitation. Geochimica et Cosmochimica Acta. 269: 566580.Google Scholar
Mavromatis, V, van Zuilen, K, Purgstaller, B, Baldermann, A, Nägler, TF, and Dietzel, M. (2016). Barium isotope fractionation during witherite (BaCO3) dissolution, precipitation and at equilibrium. Geochimica et Cosmochimica Acta. 190: 7284.Google Scholar
McIntosh, JC, Hendry, MJ, Ballentine, C, Haszeldine, RS, Mayer, B, Etiope, G, Elsner, M, Darrah, TH, and Prinzhofer, A et al. (2019). A critical review of state-of-the-art and emerging approaches to identify fracking-derived gases and associated contaminants in aquifers. Environmental Science and Technology. 53: 10631077.Google Scholar
Millot, R, Guerrot, C, Innocent, C, Négrel, P, and Sanjuan, B. (2011). Chemical, multi-isotopic (Li–B–Sr–U–H–O) and thermal characterization of Triassic formation waters from the Paris Basin. Chemical Geology. 283: 226241.Google Scholar
Naftz, DL, Peterman, ZE, and Spangler, LE. (1997). Using δ87Sr values to identify sources of salinity to a freshwater aquifer, Greater Aneth Oil Field, Utah, USA. Chemical Geology. 141: 195209.Google Scholar
Neymark, LA, Premo, WR, and Emsbo, P. (2018). Combined radiogenic (87Sr/86Sr, 234U/238U) and stable (δ88Sr) isotope systematics as tracers of anthropogenic groundwater contamination within the Williston Basin, USA. Applied Geochemistry. 96: 1123.Google Scholar
Ni, Y, Zou, C, Cui, H, Li, J, Lauer, NE, Harkness, JS, Kondash, AJ, Coyte, RM, and Dwyer, GS, et al. (2018). Origin of flowback and produced waters from Sichuan Basin, China. Environmental Science & Technology. 52: 1451914527.Google Scholar
Ono, S. (2017). Photochemistry of sulfur dioxide and the origin of mass-independent isotope fractionation in Earth’s atmosphere. Annual Review of Earth and Planetary Sciences. 45: 301329.Google Scholar
Osborn, SG, McIntosh, JC, Hanor, JS, and Biddulph, D. (2012). Iodine-129, 87Sr/86Sr, and trace elemental geochemistry of northern Appalachian Basin brines: Evidence for basinal-scale fluid migration and clay mineral diagenesis. American Journal of Science. 312: 263287.Google Scholar
Ouyang, B, Akob, DM, Dunlap, D, and Renock, D. (2017). Microbially mediated barite dissolution in anoxic brines. Applied Geochemistry. 76: 5159.Google Scholar
Pfister, S, Capo, RC, Stewart, BW, Macpherson, GL, Phan, TT, Gardiner, JB, Diehl, JR, Lopano, C, and Hakala, JA. (2017). Geochemical and lithium isotope tracking of dissolved solid sources in Permian Basin carbonate reservoir and overlying aquifer waters at an enhanced oil recovery site, northwest Texas, USA. Applied Geochemistry. 87: 122135.Google Scholar
Phan, TT, Hakala, JA, and Sharma, S. (2020). Application of isotopic and geochemical signals in unconventional oil and gas reservoir produced waters toward characterizing in situ geochemical fluid-shale reactions. Science of the Total Environment. 714: 136867.Google Scholar
Phan, TT, Capo, RC, Stewart, BW, Macpherson, GL, Rowan, EL, and Hammack, RW. (2016). Factors controlling Li concentration and isotopic composition in formation waters and host rocks of Marcellus Shale, Appalachian Basin. Chemical Geology. 420: 162179.Google Scholar
Phan, TT, Capo, RC, Stewart, BW, Graney, JR, Johnson, JD, Sharma, S, and Toro, J. (2015). Trace metal distribution and mobility in drill cuttings and produced waters from Marcellus shale gas extraction: uranium, arsenic, barium. Applied Geochemistry. 60: 89103.Google Scholar
Pollastro, RM, Roberts, LNR, and Cook, TA. (2010). Geologic assessment of technically recoverable oil in the Devonian and Mississippian Bakken Formation. Assessment of Undiscovered Oil and Gas Resources of the Williston Basin Province of North Dakota, Montana, and South Dakota, 2010. U.S. Geological Survey Digital Data Series, pp. 134.Google Scholar
Porcelli, D and Baskaran, M. (2012). An overview of isotope geochemistry in environmental studies. In Baskaran, M (ed.) Handbook of Environmental Isotope Geochemistry 1. Springer, pp. 1132.Google Scholar
Preston, TM, Thamke, J., Smith, BD, and Peterman, ZE. (2014). Chapter B: Brine contamination of Prairie Pothole environments at three study sites in the Williston Basin, United States. In Gleason, RA and Tangen, BA (eds.) Brine Contamination to Aquatic Resources from Oil and Gas Development in the Williston Basin, United States. U.S. Geological Survey Scientific Investigations Report 2014-5017, pp. 2162.Google Scholar
Pretet, C, van Zuilen, K, Nägler, TF, Reynaud, S, Böttcher, ME, and Samankassou, E. (2016). Constraints on barium isotope fractionation during aragonite precipitation by corals. The Depositional Record. 1: 118129.Google Scholar
Qi, HP, Coplen, TB, Wang, QZ, and Wang, YH. (1997). Unnatural isotopic composition of lithium reagents. Analytical Chemistry. 69: 40764078.Google Scholar
Romanak, KD, Smyth, RC, Yang, C, Hovorka, SD, Rearick, M, and Lu, J. (2012). Sensitivity of groundwater systems to CO2: Application of a site-specific analysis of carbonate monitoring parameters at the SACROC CO2-enhanced oil field. International Journal of Greenhouse Gas Control. 6: 142152.Google Scholar
Rotenberg, E, Davis, DW, Amelin, Y, Ghosh, S, and Bergquist, BA. (2012). Determination of the decay-constant of 87Rb by laboratory accumulation of 87Sr. Geochimica et Cosmochimica Acta. 85: 4157.Google Scholar
Rowan, EL, Engle, MA, Kirby, CS, and Kraemer, TF. (2011). Radium content of oil- and gas-field produced waters in the northern Appalachian Basin (USA): Summary and discussion of data. United States Geological Survey Scientific Investigations Report. 2011–5135: 131.Google Scholar
Rowan, EL, Engle, MA, Kraemer, TF, Schroeder, KT, Hammack, RW, and Doughten, MW. (2015). Geochemical and isotopic evolution of water produced from Middle Devonian Marcellus Shale gas wells, Appalachian Basin, Pennsylvania. American Association of Petroleum Geologists Bulletin. 99: 181206.Google Scholar
Schauble, EA. (2007). Role of nuclear volume in driving equilibrium stable isotope fractionation of mercury, thallium, and other very heavy elements. Geochimica et Cosmochimica Acta. 71: 21702189.Google Scholar
Shrestha, N, Chilkoor, G, Wilder, J, Gadhamshetty, V, and Stone, JJ. (2017). Potential water resource impacts of hydraulic fracturing from unconventional oil production in the Bakken shale. Water Research. 108: 124.Google Scholar
Skalak, KJ, Engle, MA, Rowan, EL, Jolly, GD, Conko, KM, Benthem, AJ, and Kraemer, TF. (2014). Surface disposal of produced waters in western and southwestern Pennsylvania: Potential for accumulation of alkali-earth elements in sediments. International Journal of Coal Geology. 126: 162170.Google Scholar
Soeder, DJ, Sharma, S, Pekney, N, Hopkinson, L, Dilmore, R, Kutchko, B, Stewart, B, Carter, C, Hakala, A, and Capo, R. (2014). An approach for assessing engineering risk from shale gas wells in the United States. International Journal of Coal Geology. 126: 419.Google Scholar
Stewart, BW, Capo, RC, and Chadwick, OA. (1998). Quantitative strontium isotope models for weathering, pedogenesis and biogeochemical cycling. Geoderma. 82: 173195.Google Scholar
Stewart, BW, Chapman, EC, Capo, RC, Johnson, JD, Graney, JR, Kirby, CS, and Schroeder, KT. (2015). Origin of brines, salts and carbonate from shales of the Marcellus Formation: Evidence from geochemical and Sr isotope study of sequentially extracted fluids. Applied Geochemistry. 60: 7888.Google Scholar
Tasker, TL, Warner, NR, and Burgos, WD. (2020). Geochemical and isotope analysis of produced water from the Utica/Point Pleasant Shale, Appalachian Basin. Environmental Science: Processes & Impacts. 22: 12241232.Google Scholar
Tasker, TL, Burgos, WD, Piotrowski, P, Castillo-Meza, L, Blewett, TA, Ganow, KB, Stallworth, A, Delompré, PLM, and Goss, GG, et al. (2018). Environmental and human health impacts of spreading oil and gas wastewater on roads. Environmental Science & Technology. 52: 70817091.Google Scholar
Tieman, ZG, Stewart, BW, Capo, RC, Phan, TT, Lopano, CL, and Hakala, JA. (2020). Barium isotopes track the source of dissolved solids in produced water from the unconventional Marcellus Shale gas play. Environmental Science and Technology. 54: 42754285.Google Scholar
Tisherman, R and Bain, DJ. (2019). Alkali earth ratios differentiate conventional and unconventional hydrocarbon brine contamination. Science of the Total Environment. 695: 133944.Google Scholar
Tomascak, PB. (2004). Developments in the understanding and application of lithium isotopes in the Earth and planetary sciences. In Johnson, CM, Beard, BL, and Albarède, F (eds.) Geochemistry of Non-traditional Stable Isotopes, Reviews in Mineralogy & Geochemistry 55. Mineralogical Society of America and Geochemical Society, 153195.Google Scholar
van Zuilen, K, Müller, T, Nägler, TF, Dietzel, M, and Küsters, T. (2016). Experimental determination of barium isotope fractionation during diffusion and adsorption processes at low temperatures. Geochimica et Cosmochimica Acta. 186: 226241.Google Scholar
Veil, JA. (2010). Final Report: Water Management Technologies Used by Marcellus Shale Gas Producers. U. S. Department of Energy-NETL-Argonne National Lab. www.evs.anl.gov/pub/dsp_detail.cfm?PubID=2537.Google Scholar
Vengosh, A, Jackson, RB, Warner, NR, Darrah, TH, and Kondash, A. (2014). A critical review of the risks to water resources from unconventional shale gas development and hydraulic fracturing in the United States. Environmental Science and Technology. 48: 83348348.Google Scholar
Vigier, N, Decarreau, A, Millot, R, Carignan, J, Petit, S, and France-Lanord, C. (2008). Quantifying Li isotope fractionation during smectite formation and implications for the Li cycle. Geochimica et Cosmochimica Acta. 72: 780792.Google Scholar
von Allmen, K, Böttcher, ME, Samankassou, E, and Nägler, TF. (2010). Barium isotope fractionation in the global barium cycle: First evidence from barium minerals and precipitation experiments. Chemical Geology. 277: 7077.Google Scholar
Warner, NR, Christie, CA, Jackson, RB, and Vengosh, A. (2013). Impacts of shale gas wastewater disposal on water quality in western Pennsylvania. Environmental Science and Technology. 47: 1184911857.Google Scholar
Warner, NR, Darrah, TH, Jackson, RB, Millot, R, Kloppmann, W, and Vengosh, A. (2014). New tracers identify hydraulic fracturing fluids and accidental releases from oil and gas operations. Environmental Science and Technology. 48: 1255212560.Google Scholar
Warner, NR, Jackson, RB, Darrah, TH, Osborn, SG, Down, A, Zhao, K, White, A, and Vengosh, A. (2012). Geochemical evidence for possible natural migration of Marcellus Formation brine to shallow aquifers in Pennsylvania. Proceedings of the National Academy of Sciences. 109: 1196111966.Google Scholar
Weiss, DJ, Rehkämper, M, Schoenberg, R, McLaughlin, M, Kirby, J, Campbell, PGC, Arnold, T, Chapman, J, Peel, K, and Gioia, S. (2008)Application of nontraditional stable-isotope systems to the study of sources and fate of metals in the environment. Environmental Science and Technology. 42: 655664.Google Scholar
Wiederhold, JG. (2015). Metal Stable Isotope Signatures as Tracers in Environmental Geochemistry. Environmental Science & Technology. 49: 26062624.Google Scholar
Williams, LB, Elliott, WC, and Hervig, RL. (2015). Tracing hydrocarbons in gas shale using lithium and boron isotopes: Denver Basin USA, Wattenberg Gas Field. Chemical Geology. 417: 404413.Google Scholar
Zagorski, WA, Wrightstone, GR, and Bowman, DC. (2012). The Appalachian Basin Marcellus gas play: Its history of development, geologic controls on production, and future potential as a world-class reservoir. In Breyer, JA (ed.) Shale Reservoirs: Giant Resources for the 21st Century. AAPG Memoir, pp. 172200.Google Scholar
Zhang, L, Chan, L-H, and Gieskes, JM. (1998). Lithium isotope geochemistry of pore waters from Ocean Drilling Program Sites 918 and 919, Irminger Basin. Geochimica et Cosmochimica Acta. 62: 24372450.Google Scholar
Zheng, Z, Zhang, H, Chen, Z, Li, X, Zhu, P, and Cui, X. (2017). Hydrogeochemical and isotopic indicators of hydraulic fracturing flowback fluids in shallow groundwater and stream water, derived from Dameigou shale gas extraction in the northern Qaidam Basin. Environmental Science & Technology. 51: 58895898.Google Scholar

References

Alvarez, RA, Zavala-Araiza, D, Lyon, DR, Allen, DT, Barkley, ZR, Brandt, AR, Davis, KJ, Herndon, SC, Jacob, D., Karion, A, Kort, EA, Lamb, BK, Lauvaux, T, Maasakkers, JD, Marchese, AJ, Omara, M, Pacala, SW, Peischl, J, Robinson, AL, Shepson, PB, Sweeney, C, Townsend-Smal, A, Wofsy, SC, and Hamburg, SP. (2018). Assessment of methane emissions from the U.S. oil and gas supply chain. Science. 7204. https://doi.org/10.1126/science.aar7204Google Scholar
Aydin, M, Verhulst, KR, Saltzman, ES, Battle, MO, Montzka, SA, Blake, DR, Tang, Q, and Prather, MJ. (2011). Recent decreases in fossil-fuel emissions of ethane and methane derived from firn air. Nature. 476: 198201. https://doi.org/10.1038/nature10352Google Scholar
Baldassare, F and Chapman, E. (2018). Chapter 4 - The application of isotope geochemistry in stray gas investigations: Case studies. In Stout, SA and Wang, Z (eds.) Oil Spill Environmental Forensics Case Studies. Butterworth-Heinemann, pp. 6786. https://doi.org/10.1016/B978–0-12-804434-6.00004-5Google Scholar
Barth-Naftilan, E, Sohng, J, and Saiers, JE. (2018). Methane in groundwater before, during, and after hydraulic fracturing of the Marcellus Shale. PNAS. 115: 69706975. https://doi.org/10.1073/pnas.1720898115Google Scholar
Botner, EC, Townsend-Small, A, Nash, DB, Xu, X, Schimmelmann, A, and Miller, JH. (2018). Monitoring concentration and isotopic composition of methane in groundwater in the Utica Shale hydraulic fracturing region of Ohio. Environmental Monitoring and Assessment. 190: 322. https://doi.org/10.1007/s10661–018-6696-1Google Scholar
Clayton, JL. (1998). Geochemistry of coalbed gas – A review. International Journal of Coal Geology. 35: 159173. https://doi.org/10.1016/S0166–5162(97)00017-7Google Scholar
Coleman, DD, Liu, C-L, Hackley, KC, and Pelphrey, SR. (1995). Isotopic Identification of Landfill Methane. Environmental Geosciences. 2: 95103.Google Scholar
Golding, SD, Boreham, CJ, and Esterle, JS. (2013). Stable isotope geochemistry of coal bed and shale gas and related production waters: A review. International Journal of Coal Geology. 120: 2440. https://doi.org/10.1016/j.coal.2013.09.001Google Scholar
Hammond, PA. (2016). The relationship between methane migration and shale-gas well operations near Dimock, Pennsylvania, USA. Hydrogeology Journal. 24: 503519. https://doi.org/10.1007/s10040–015-1332-4Google Scholar
Howarth, RW. (2019). Ideas and perspectives: is shale gas a major driver of recent increase in global atmospheric methane? Biogeosciences. 16: 30333046. https://doi.org/10.5194/bg-16-3033-2019Google Scholar
Howarth, RW. (2021). Methane and climate change. In Stolz, JF, Griffin, WM, and Bain, DJ (eds.) Environmental Impacts from the Development of Unconventional Oil and Gas Reserves. Cambridge University Press.Google Scholar
Jackson, RB, Vengosh, A, Darrah, TH, Warner, NR, Down, A, Poreda, RJ, Osborn, SG, Zhao, K, and Karr, JD. (2013). Increased stray gas abundance in a subset of drinking water wells near Marcellus shale gas extraction. PNAS. 110: 1125011255. https://doi.org/10.1073/pnas.1221635110Google Scholar
Kai, FM, Tyler, SC, Randerson, JT, and Blake, DR. (2011). Reduced methane growth rate explained by decreased Northern Hemisphere microbial sources. Nature. 476: 194197. https://doi.org/10.1038/nature10259Google Scholar
Keeling, CD. (1958). The concentration and isotopic abundances of atmospheric carbon dioxide in rural areas. Geochimica et Cosmochimica Acta. 13: 322334. https://doi.org/10.1016/0016-7037(58)90033-4Google Scholar
Keeling, CD. (1961) The concentration and isotopic abundances of carbon dioxide in rural and marine air. Geochimica et Cosmochimica Acta. 24: 277298. https://doi.org/10.1016/0016-7037(61)90023-0Google Scholar
Kendall, C and Doctor, DH. (2003). 5.11 - Stable Isotope Applications in Hydrologic Studies. In Holland, HD and Turekian, KK (eds.), Treatise on Geochemistry. Pergamon, pp. 319364. https://doi.org/10.1016/B0–08-043751-6/05081-7Google Scholar
Lamb, BK, Edburg, SL, Ferrara, TW, Howard, T, Harrison, MR, Kolb, CE, Townsend-Small, A, Dyck, W, Possolo, A, and Whetstone, JR. (2015). Direct measurements show decreasing methane emissions from natural gas local distribution systems in the United States. Environmental Science & Technology. 49: 51615169. https://doi.org/10.1021/es505116pGoogle Scholar
Lassey, KR, Etheridge, DM, Lowe, DC, Smith, AM, and Ferretti, DF. (2007a). Centennial evolution of the atmospheric methane budget: What do the carbon isotopes tell us? Atmospheric Chemistry and Physics. 7: 21192139. https://doi.org/10.5194/acp-7-2119-2007Google Scholar
Lassey, KR, Lowe, DC, and Smith, AM. (2007b). The atmospheric cycling of radiomethane and the “fossil fraction” of the methane source. Atmospheric Chemistry and Physics. 7: 21412149. https://doi.org/10.5194/acp-7-2141-2007Google Scholar
Mace, EK, Aalseth, CE, Day, AR, Hoppe, EW, Keillor, ME, Moran, JJ, Panisko, ME, Seifert, A, Tatishvili, G, and Williams, RM. (2016). First results of a simultaneous measurement of tritium and 14C in an ultra-low-background proportional counter for environmental sources of methane. Journal of Environmental Radioactivity. 155–156: 122129. https://doi.org/10.1016/j.jenvrad.2016.02.001Google Scholar
Martini, AM, Walter, LM, Budai, JM, Ku, TCW, Kaiser, CJ, and Schoell, M. (1998). Genetic and temporal relations between formation waters and biogenic methane: Upper Devonian Antrim Shale, Michigan Basin, USA. Geochimica et Cosmochimica Acta. 62: 16991720. https://doi.org/10.1016/S0016–7037(98)00090-8Google Scholar
McIntosh, JC, Hendry, MJ, Ballentine, C, Haszeldine, RS, Mayer, B, Etiope, G, Elsner, M, Darrah, TH, Prinzhofer, A, Osborn, S, Stalker, L, Kuloyo, O, Lu, Z-T, Martini, A, and Lollar, BS. (2019). A critical review of state-of-the-art and emerging approaches to identify fracking-derived gases and associated contaminants in aquifers. Environmental Science & Technology. 53: 10631077. https://doi.org/10.1021/acs.est.8b05807Google Scholar
Milkov, AV, Schwietzke, S, Allen, G, Sherwood, OA, and Etiope, G. (2020). Using global isotopic data to constrain the role of shale gas production in recent increases in atmospheric methane. Scientific Reports. 10: 17. https://doi.org/10.1038/s41598–020-61035-wGoogle Scholar
Nisbet, EG et al. (2016). Rising atmospheric methane: 2007–2014 growth and isotopic shift. Global Biogeochemical Cycles. 30: 13561370. https://doi.org/10.1002/2016GB005406Google Scholar
Nisbet, et al. (2019). Very strong atmospheric methane growth in the 4 Years 2014–2017: Implications for the Paris Agreement. Global Biogeochemical Cycles. 33: 318342. https://doi.org/10.1029/2018GB006009Google Scholar
NOAA Global Monitoring Division. (2020). NOAA ESRL Global Monitoring Division - FTP Navigator [WWW Document]. URL www.esrl.noaa.gov/gmd/dv/data/index.php?parameter_name=C13%252FC12%2Bin%2BMethane (accessed March 1, 2020).Google Scholar
Osborn, SG and McIntosh, JC. (2010). Chemical and isotopic tracers of the contribution of microbial gas in Devonian organic-rich shales and reservoir sandstones, northern Appalachian Basin. Applied Geochemistry. 25: 456471. https://doi.org/10.1016/j.apgeochem.2010.01.001Google Scholar
Osborn, SG, Vengosh, A, Warner, NR, and Jackson, RB. (2011). Methane contamination of drinking water accompanying gas-well drilling and hydraulic fracturing. PNAS. 108: 81728176. https://doi.org/10.1073/pnas.1100682108Google Scholar
Pataki, DE, Ehleringer, JR, Flanagan, LB, Yakir, D, Bowling, DR, Still, CJ, Buchmann, N, Kaplan, JO, and Berry, JA. (2003). The application and interpretation of Keeling plots in terrestrial carbon cycle research. Global Biogeochemical Cycles. 17. https://doi.org/10.1029/2001GB001850Google Scholar
Peischl, J. et al. (2013). Quantifying sources of methane using light alkanes in the Los Angeles basin, California. Journal of Geophysical Research: Atmospheres. 118: 49744990. https://doi.org/10.1002/jgrd.50413Google Scholar
Pétron, G. et al. (2014) A new look at methane and nonmethane hydrocarbon emissions from oil and natural gas operations in the Colorado Denver-Julesburg Basin. Journal of Geophysical Research: Atmospheres. 119: 68366852. https://doi.org/10.1002/2013JD021272Google Scholar
Reeburgh, WS. (2007). Global methane biogeochemistry. In Holland, HD and Turekian, KK (eds.) Treatise on Geochemistry. Pergamon, pp. 132. https://doi.org/10.1016/B0–08-043751-6/04036-6Google Scholar
Rice, AL, Butenhoff, CL, Teama, DG, Röger, FH, Khalil, MAK, and Rasmussen, RA. (2016). Atmospheric methane isotopic record favors fossil sources flat in 1980s and 1990s with recent increase. PNAS. 113: 1079110796. https://doi.org/10.1073/pnas.1522923113Google Scholar
Schlegel, ME, McIntosh, JC, Bates, BL, Kirk, MF, and Martini, AM. (2011). Comparison of fluid geochemistry and microbiology of multiple organic-rich reservoirs in the Illinois Basin, USA: Evidence for controls on methanogenesis and microbial transport. Geochimica et Cosmochimica Acta. 75: 19031919. https://doi.org/10.1016/j.gca.2011.01.016Google Scholar
Schwietzke, S, Sherwood, OA, Bruhwiler, LMP, Miller, JB, Etiope, G, Dlugokencky, EJ, Michel, SE, Arling, VA, Vaughn, BH, White, JWC, and Tans, PP. (2016). Upward revision of global fossil fuel methane emissions based on isotope database. Nature. 538: 8891. https://doi.org/10.1038/nature19797Google Scholar
Scott, AR, Kaiser, WR, and Ayers, WB. (1994). Thermogenic and secondary biogenic gases, San Juan Basin, Colorado and New Mexico: Implications for coalbed gas producibility. AAPG Bulletin. 78: 11861209. https://doi.org/10.1306/A25FEAA9–171B-11D7–8645000102C1865DGoogle Scholar
Sherwood, OA, Schwietzke, S, Arling, VA, and Etiope, G. (2017). Global inventory of gas geochemistry data from fossil fuel, microbial and burning sources, version 2017. Earth System Science Data. 9: 639656. https://doi.org/10.5194/essd-9-639-2017Google Scholar
Siegel, DI, Azzolina, NA, Smith, BJ, Perry, AE, and Bothun, RL. (2015). Methane Concentrations in Water Wells Unrelated to Proximity to Existing Oil and Gas Wells in Northeastern Pennsylvania. Environmental Science & Technology. 49: 41064112. https://doi.org/10.1021/es505775cGoogle Scholar
Smith, JW and Pallasser, RJ. (1996). Microbial Origin of Australian Coalbed Methane. AAPG Bulletin. 80: 891897. https://doi.org/10.1306/64ED88FE-1724-11D7–8645000102C1865DGoogle Scholar
Thomas, MA. (2018). Chemical and isotopic characteristics of methane in groundwater of Ohio, 2016, U.S. Geological Survey Scientific Investigations Report 2018–5097.Google Scholar
Townsend-Small, A, Tyler, SC, Pataki, DE, Xu, X, and Christensen, LE. (2012). Isotopic measurements of atmospheric methane in Los Angeles, California, USA: Influence of “fugitive” fossil fuel emissions. Journal of Geophysical Research: Atmospheres. 117. https://doi.org/10.1029/2011JD016826Google Scholar
Townsend-Small, A, Botner, EC, Jimenez, KL, Schroeder, JR, Blake, NJ, Meinardi, S, Blake, DR, Sive, BC, Bon, D, Crawford, JH, Pfister, G, and Flocke, FM. (2016). Using stable isotopes of hydrogen to quantify biogenic and thermogenic atmospheric methane sources: A case study from the Colorado Front Range: Hydrogen Isotopes in the Front Range. Geophysical Research Letters. 43: 11,462-11,471. https://doi.org/10.1002/2016GL071438Google Scholar
United States Department of the Interior. (2001). Technical Measures for the Investigation and Mitigation of Fugitive Methane Hazards in Areas of Coal Mining. Office of Surface Mining Reclamation and Enforcement.Google Scholar
US EPA National Center for Environmental Assessment, I.O. (2016). Hydraulic Fracturing for Oil and Gas: Impacts from the Hydraulic Fracturing Water Cycle on Drinking Water Resources in the United States (Final Report) [WWW Document]. URL https://cfpub.epa.gov/ncea/hfstudy/recordisplay.cfm?deid=332990 (accessed 5.7.20).Google Scholar
Vengosh, A, Jackson, RB, Warner, N, Darrah, TH, and Kondash, A. (2014). A critical review of the risks to water resources from unconventional shale gas development and hydraulic fracturing in the United States. Environmental Science & Technology. 48: 83348348. https://doi.org/10.1021/es405118yGoogle Scholar
Vidic, RD, Brantley, SL, Vandenbossche, JM, Yoxtheimer, D, and Abad, JD. (2013). Impact of shale gas development on regional water quality. Science. 340. https://doi.org/10.1126/science.1235009Google Scholar
Vigneron, A, Bishop, A, Alsop, EB, Hull, K, Rhodes, I, Hendricks, R, Head, IM, and Tsesmetzis, N. (2017). Microbial and Isotopic Evidence for Methane Cycling in Hydrocarbon-Containing Groundwater from the Pennsylvania Region. Frontiers in Microbiology. 8. https://doi.org/10.3389/fmicb.2017.00593Google Scholar
Vinson, DS, Blair, NE, Martini, AM, Larter, S, Orem, WH, and McIntosh, JC. (2017). Microbial methane from in situ biodegradation of coal and shale: A review and reevaluation of hydrogen and carbon isotope signatures. Chemical Geology. 453: 128145. https://doi.org/10.1016/j.chemgeo.2017.01.027Google Scholar
Wennberg, PO, Mui, W, Wunch, D, Kort, EA, Blake, DR, Atlas, EL, Santoni, GW, Wofsy, SC, Diskin, GS, Jeong, S, and Fischer, ML. (2012). On the Sources of Methane to the Los Angeles Atmosphere. Environmental Science & Technology. 46: 92829289. https://doi.org/10.1021/es301138yGoogle Scholar
Whiticar, MJ. (1999). Carbon and hydrogen isotope systematics of bacterial formation and oxidation of methane. Chemical Geology. 161: 291314. https://doi.org/10.1016/S0009–2541(99)00092-3Google Scholar
Whiticar, M and Schaefer, H. (2007). Constraining past global tropospheric methane budgets with carbon and hydrogen isotope ratios in ice. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 365: 17931828. https://doi.org/10.1098/rsta.2007.2048Google Scholar
Worden, JR, Bloom, AA, Pandey, S., Jiang, Z., Worden, H.M., Walker, T.W., Houweling, S, and Röckmann, T. (2017). Reduced biomass burning emissions reconcile conflicting estimates of the post-2006 atmospheric methane budget. Nature Communication. 8: 2227.Google Scholar

References

Akob, DM, Cozzarelli, IM, Dunlap, DS, Rowan, EL, and Lorah, MM. (2015). Organic and inorganic composition and microbiology of produced waters from Pennsylvania shale gas wells. Applied Geochemistry. 60: 116125.Google Scholar
Booker, AE, Borton, MA, Daly, RA et al. (2017). Sulfide generation by dominant Halanaerobium microorganisms in hydraulically fractured shales. mSphere. 2(4): e00257–17.Google Scholar
Booker, AE, Hoyt, DW, Meulia, T et al. (2019). Deep-subsurface pressure stimulates metabolic plasticity in shale-colonizing Halanaerobium spp. Applied and Environmental Microbiology. 85(12): e00018–19.Google Scholar
Borton, MA, Daly, RA, O’Banion, B et al. (2018a). Comparative genomics and physiology of the genus Methanohalophilus, a prevalent methanogen in hydraulically fractured. Environmental Microbiology. 20(12): 45964611.Google Scholar
Borton, MA, Hoyt, DW, Roux, S et al. (2018b). Coupled laboratory and field investigations resolve microbial interactions that underpin persistence in hydraulically fractured shales. Proceedings of the National Academy of Sciences. 115(28): E6585E6594.Google Scholar
Bowman, JP, McCammon, SA, Nichols, DS et al. (1997). Shewanella gelidimarina sp. nov. and Shewanella frigidimarina sp. nov., novel Antarctic species with the ability to produce eicosapentaenoic acid (20:5ω3) and grow anaerobically by dissimilatory Fe(III) reduction. International Journal of Systematic Bacteriology. 47(4): 10401047.Google Scholar
Campa, MF, Wolfe, AK, Techtmann, SM, Harik, A, and Hazen, TC. (2019). Unconventional oil and gas energy systems: an unidentified hotspot of antimicrobial resistance? Frontiers in Microbiology. 10: 2392.Google Scholar
Cliffe, L, Nixon, SL, Daly, RA et al. (2020). Identification of persistent sulfidogenic bacteria in shale gas produced waters. Frontiers in Microbiology. 11: 286.Google Scholar
Cluff, MA, Hartsock, A, MacRae, JD, Carter, K, and Mouser, PJ. (2014). Temporal changes in microbial ecology and geochemistry in produced water from hydraulically fractured Marcellus shale gas wells. Environmental Science & Technology. 48(11): 65086517.Google Scholar
Colwell, FS, Onstott, TC, Delwiche, ME et al. (1997). Microorganisms from deep, high temperature sandstones: constraints on microbial colonization. FEMS Microbiology Reviews. 20: 425435.Google Scholar
Craciun, S and Balskus, EP. (2012). Microbial conversion of choline to trimethylamine requires a glycyl radical enzyme. Proceedings of the National Academy of Sciences. 109(52): 2130721312.Google Scholar
Daly, RA, Borton, MA, Wilkins, MJ et al. (2016). Microbial metabolism in a 2.5-km-deep ecosystem created by hydraulic fracturing in shales. Nature Microbiology. 1(10): 16146.Google Scholar
Daly, RA, Roux, S, Borton, MA et al. (2019). Viruses control dominant bacteria colonizing the terrestrial deep biosphere after hydraulic fracturing. Nature Microbiology. 4: 352361.Google Scholar
Davis, JP, Struchtemeyer, CG, and Elshahed, MS. (2012). Bacterial communities associated with production facilities of two newly drilled thermogenic natural gas wells in the Barnett Shale (Texas, USA). Microbial Ecology. 64(4): 942954.Google Scholar
Elsner, M and Hoelzer, K. (2016). Quantitative survey and structural classification of hydraulic fracturing chemicals reported in unconventional gas production. Environmental Science & Technology. 50: 32903314.Google Scholar
Evans, MV, Getzinger, G, Luek, JL et al. (2019). In situ transformation of ethoxylate and glycol surfactants buu shale-colonizing microorganisms during hydraulic fracturing. ISME Journal. 13: 26902700.Google Scholar
Evans, MV, Panescu, J, Hanson, AJ et al. (2018). Members of Marinobacter and Arcobacter influence system biogeochemistry during early production of hydraulically fractured natural gas wells in the Appalachian basin. Frontiers in Microbiology. 9: 2646.Google Scholar
Fichter, J, Wunch, K, Moore, R et al. (2012). How hot is too hot for bacteria? A technical study assessing bacterial establishment in downhole drilling, fracturing and stimulation operations. In CORROSION 2012, March 11–15, Salt Lake City, Utah. NACE International.Google Scholar
Fredrickson, JK, McKinley, JP, Bjornstad, BN et al. (1997). Pore-size constraints on the activity and survival of subsurface bacteria in a late Cretaceous shale-sandstone sequence, northwestern New Mexico. Geomicrobiology Journal. 14: 182202.Google Scholar
Huang, R. (2008). Shale-Derived Dissolved Organic Matter as a Substrate for Subsurface Methanogenic Communities in the Antrim Shale Michigan Basin, USA. Masters thesis. Department of Geosciences, University of Massachusetts Amherst.Google Scholar
Johnson, K, French, K, Fichter, JK, and Oden, R. (2008). Use of microbiocides in Barnett Shale gas well fracturing fluids to control bacteria related problems. In CORROSION 2008, March 16–20, New Orleans, Louisiana. NACE International.Google Scholar
Kahrilas, GA, Blotevogul, J, Stewart, PS, and Borch, T. (2015) Biocides in hydraulic fracturing fluids: A critical review of their usage, mobility, degradation and toxicity. Environmental Science & Technology. 49(1): 1632.Google Scholar
Kashefi, K and Lovley, DR. (2003) Extending the upper temperature limit for life. Science. 301(5635): 934.Google Scholar
Kekacs, D, Drollette, BD, Brooker, M, Plata, DL, and Mouser, PJ. (2015). Aerobic biodegradation of organic compounds in hydraulic fracturing fluids. Biodegradation. 26(4): 271287.Google Scholar
Kirk, MF, Martini, AM, Breecker, SO et al. (2012). Impact of commercial natural gas production on geochemistry and microbiology in a shale-gas reservoir. Chemical Geology. 332–333: 1525.Google Scholar
Krumholz, LR, McKinley, JP, Ulrich, GA, and Suflita, JM. (1997). Confined subsurface microbial communities in cretaceous rock. Nature. 386: 6466.Google Scholar
Liang, R, Davidova, IA, Marks, CR et al. (2016). Metabolic capability of a predominant Halanaerobium sp. in hydraulically fractured gas wells and its implication in pipeline corrosion. Frontiers in Microbiology. 7: 988.Google Scholar
Lipus, D, Ross, D, Bibby, K, and Gulliver, D. (2017a). Draft genome sequence of Pseudomonas sp. BDAL1 reconstructed from a Bakken shale hydraulic fracturing-produced water storage tank metagenome. Genome Announcements. 5: e00033–17.Google Scholar
Lipus, D, Roy, D, Khan, E et al. (2018). Microbial communities in Bakken region produced water. FEMS Microbiology Letters. 365(12): fny107.Google Scholar
Lipus, D, Vikram, A, Ross, D et al. (2017b). Predominance and metabolic potential of Halanaerobium spp. in produced water in hydraulically fractured Marcellus shale wells. Applied Environmental Microbiology. 83(8): e02659–16.Google Scholar
Martini, AM, Walter, LM, Budai, JM et al. (1998). Genetic and temporal relations between formation waters and biogenic methane: Upper Devonian Antrim Shale, Michigan Basin, USA. Geochimica et Cosmochimica Acta. 62(10): 16991720.Google Scholar
Mouser, PJ, Borton, M, Darrah, TH, Hartsock, A, and Wrighton, KC. (2016). Hydraulic fracturing offers a view of microbial life in the deep terrestrial subsurface. FEMS Microbiology Ecology. 92(11).Google Scholar
Murali Mohan, A, Hartsock, A, Bibby, KJ et al. (2013a). Microbial community changes in hydraulic fracturing fluids and produced water from shale gas extraction. Environmental Science & Technology. 47(22): 1314113150.Google Scholar
Murali Mohan, A, Hartsock, A, Hammack, RW, Vidic, RD, and Gregory, KB. (2013b). Microbial communities in flowback water impoundments from hydraulic fracturing for recovery of shale gas. FEMS Microbiology Ecology. 86(3): 567580.Google Scholar
Nixon, SL, Daly, RA, Borton, MA et al. (2019). Genome-resolved metagenomics extends the environmental distribution of the Verrucomicrobia phylum to the deep terrestrial subsurface. mSphere. 4: e00613–19.Google Scholar
Nixon, SL, Walker, L, Streets, MDT et al. (2017). Guar gum stimulates biogenic sulfide production at elevated pressures: Implications for shale gas extraction. Frontiers in Microbiology. 8: 679.Google Scholar
Onstott, TC, Phelps, TJ, Colwell, FS et al. (1998). Observations pertaining to the origin and ecology of microorganisms recovered from the deep subsurface of Taylorsville Basin, Virginia. Geomicrobiology Journal. 15: 353385.Google Scholar
Santillan, EFU, Choi, W, Bennett, PC, and Leyris, JD. (2015). The effects of biocide use on the microbiology and geochemistry of produced water in the Eagle Ford formation, Texas, U.S.A. Journal of Petroleum Science and Engineering. 135: 19.Google Scholar
Schlegel, ME, McIntosh, JC, Bates, BL, Kirk, MF, and Martini, AM. (2011). Comparison of fluid geochemistry and microbiology of multiple organic-rich reservoirs in the Illinois Basin, USA: Evidence for controls on methanogenesis and microbial transport. Geochimica et Cosmochimica Acta. 75: 19031919.Google Scholar
Strong, LC, Gould, T, Kasinkas, L et al. (2013). Biodegradation in waters from hydraulic fracturing: Chemistry, microbiology, and engineering. Journal of Environmental Engineering. 140(5): B4013001.Google Scholar
Struchtemeyer, CG. (2018). Microbiology of oil- and natural gas-producing shale formations: An overview. In Steffan, R (ed) Consequences of Microbial Interactions with Hydrocarbons, Oils, and Lipids: Biodegradation and Bioremediation. Springer Nature Switzerland AG, pp. 215232.Google Scholar
Struchtemeyer, CG and Elshahed, MS. (2012). Bacterial communities associated with hydraulic fracturing fluids in thermogenic natural gas wells in North Central Texas, USA. FEMS Microbiol Ecology. 81(1): 1325.Google Scholar
Struchtemeyer, CG, Davis, JP, and Elshahed, MS. (2011). Influence of drilling mud formation process on the bacterial communities in thermogenic natural wells of the Barnett Shale. Applied and Environmental Microbiology. 77(14): 47444753.Google Scholar
Struchtemeyer, CG, Morrison, MD, and Elshahed, MS. (2012). A critical assessment of the efficacy of biocides used during the hydraulic fracturing process in shale natural gas wells. International Biodeterioration & Biodegradation. 71: 1521.Google Scholar
Struchtemeyer, CG, Youssef, NH, and Elshahed, MS. (2017). Protocols for investigating the microbiology of drilling fluids, hydraulic fracturing fluids, and formations in unconventional natural gas reservoirs. In McGenity, TJ, Timmis, KN, Fernandez, BN (eds.) Hydrocarbon and Lipid Microbiology Protocols. Springer-Verlag, pp. 125.Google Scholar
Tucker, YT, Kotcon, J, and Mroz, T. (2015). Methanogenic archaea in Marcellus shale: A possible mechanism for enhanced gas recovery in unconventional shale resources. Environmental Science & Technology. 49(11): 70487055.Google Scholar
Vandenbroucke, M and Largeau, C. (2007). Kerogen origin, evolution and structure. Organic Geochemistry. 38: 719833.Google Scholar
Venkateswaran, K, Moser, DP, Dollhopf, ME et al. (1999). Polyphasic taxonomy of the genus Shewanella and description of Shewanella oneidensis sp. nov. International Journal of Systematic Bacteriology. 49: 705724.Google Scholar
Vikram, A, Lipus, D, and Bibby, K. (2014). Produced water exposure alters bacterial response to biocides. Environmental Science & Technology. 48(21): 1300113009.Google Scholar
Vikram, A, Lipus, D, and Bibby, K. (2016). Metatranscriptome analysis of active microbial communities in produced water samples from the Marcellus Shale. Microbial Ecology. 72: 571.Google Scholar
Waldron, PJ, Petsch, ST, Martini, AM, and Nüsslein, K. (2007). Salinity constrains on subsurface archaeal diversity and methanogenesis in sedimentary rock rich in organic matter. Applied and Environmental Microbiology. 73: 41714179.Google Scholar
Wang, H, Lu, L, Chen, X, Bian, Y, and Ren, ZJ. (2019). Geochemical and microbial characterizations of flowback and produced water in three shale oil and gas plays in the central and western United States. Water Res. 164: 114942.Google Scholar
Wuchter, C, Banning, E, Mincer, T, Drenzek, NJ, and Coolen, MJ. (2013). Microbial diversity and methanogenic activity of Antrim Shale formation waters from recently fractured wells. Frontiers in Microbiology. 4: 367.Google Scholar
Zhang, Y, Yu, Z, Zhang, H, and Thompson, IP. (2017). Microbial distribution and variation in produced water from separators to storage tanks of shale gas wells in Sichuan Basin, China. Environmental Science: Water Research & Technology. 3(2): 340351.Google Scholar
Zhong, C, Li, J, Flynn, SL et al. (2019). Temporal changes in microbial community composition and geochemistry in flowback and produced water from the Duvernay formation. ACS Earth Space Chem. 3: 10471057.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.

  • Environmental Analysis
  • Edited by John Stolz, Duquesne University, Pittsburgh, Daniel Bain, University of Pittsburgh, Michael Griffin, Carnegie Mellon University, Pennsylvania
  • Book: Environmental Impacts from the Development of Unconventional Oil and Gas Reserves
  • Online publication: 28 July 2022
  • Chapter DOI: https://doi.org/10.1017/9781108774178.007
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.

  • Environmental Analysis
  • Edited by John Stolz, Duquesne University, Pittsburgh, Daniel Bain, University of Pittsburgh, Michael Griffin, Carnegie Mellon University, Pennsylvania
  • Book: Environmental Impacts from the Development of Unconventional Oil and Gas Reserves
  • Online publication: 28 July 2022
  • Chapter DOI: https://doi.org/10.1017/9781108774178.007
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.

  • Environmental Analysis
  • Edited by John Stolz, Duquesne University, Pittsburgh, Daniel Bain, University of Pittsburgh, Michael Griffin, Carnegie Mellon University, Pennsylvania
  • Book: Environmental Impacts from the Development of Unconventional Oil and Gas Reserves
  • Online publication: 28 July 2022
  • Chapter DOI: https://doi.org/10.1017/9781108774178.007
Available formats
×