Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-18T22:54:37.323Z Has data issue: false hasContentIssue false

9 - Naturally Occurring Radioactive Material (NORM)

from Part II - Environmental Analysis

Published online by Cambridge University Press:  28 July 2022

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

Summary

The recognition of naturally occurring radioactive material (NORM) associated with oil and gas fields started nearly at the same time as the discovery of radioactivity itself. Radium in produced water is typically the source of the majority of the NORM. Four processes within oil and gas formations, solubility, alpha recoil, cation exchange, and coprecipitation lead to high radium activity in pore fluids. These processes occur regardless of the type of reservoir (conventional high permeability oil and gas reservoirs or unconventional low permeability organic-rich shale source rocks). Following well stimulation via hydraulic fracturing fluids and some solids that return to the surface contain elevated radium. The data on radium from oil and gas wells across the USA is severely lacking relative to the volumes of produced water, especially considering that large volumes are beneficially used or disposed of to surface waters. Novel treatment and accurate measurements of radium are necessary prior to beneficially reuse or dispose produced water in order to protect human and environmental health. More measurements of radium in produced water should be obtained and made publicly available.

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

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/c8em00248gCrossRefGoogle ScholarPubMed
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.6b00428CrossRefGoogle 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/srep46582CrossRefGoogle ScholarPubMed
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)00034-1CrossRefGoogle 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.2018.05.200CrossRefGoogle ScholarPubMed
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.2018.05.223CrossRefGoogle 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)90186-4CrossRefGoogle 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

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×