Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-qs9v7 Total loading time: 0 Render date: 2024-07-10T20:07:43.716Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  09 December 2021

Craig M. Bethke
Affiliation:
University of Illinois, Urbana-Champaign
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

Aagaard, P. and Helgeson, H. C., 1982, Thermodynamic and kinetic constraints on reaction rates among minerals and aqueous solutions, I. Theoretical considerations. American Journal of Science 282, 237– 285.Google Scholar
Adamson, A. W. and Gast, A. P., 1997, Physical Chemistry of Surfaces, 6th ed. Wiley, New York.Google Scholar
Alemi, M. H., Goldhamer, D. A. and Nelson, D. R., 1991, Modeling selenium transport in steady-state, unsaturated soil columns. Journal of Environmental Quality 20, 8995.Google Scholar
Allison, J. D., Brown, D. S. and Novo-Gradac, K. J., 1991, MINTEQA2/ PRODEFA2, a geochemical assessment model for environmental systems, version 3.0 user’s manual. US Environmental Protection Agency Report EPA/600/3–91/021.Google Scholar
Alvarez, P. J. J., Anid, P. J. and Vogel, T. M., 1991, Kinetics of aerobic biodegradation of benzene and toluene in sandy aquifer material. Biodegradation 2, 4351.CrossRefGoogle ScholarPubMed
Anderson, G. M., 2017, Thermodynamics of Natural Systems, 3rd ed. Cambridge University Press.CrossRefGoogle Scholar
Anderson, G. M. and Crerar, D. A., 1993, Thermodynamics in Geochemistry, The Equilibrium Model. Oxford University Press.Google Scholar
Anderson, G. M. and Garven, G., 1987, Sulfate-sulfide-carbonate associations in Mississippi Valley-type leadzinc deposits. Economic Geology 82, 482488.Google Scholar
Aplin, A. C. and Warren, E. A., 1994, Oxygen isotopic indications of the mechanisms of silica transport and quartz cementation in deeply buried sandstones. Geology 22, 847850.Google Scholar
Appelo, C. A. J., Parkhurst, D. L. and Post, V. E. A., 2014, Equations for calculating hydrogeochemical reactions of minerals and gases such as CO2 at high pressures and temperatures. Geochimica et Cosmochimica Acta 125, 4967.Google Scholar
Appelo, C. A. J. and Postma, D., 1999, A consistent model for surface complexation on birnessite (ı-MnO2) and its application to a column experiment. Geochimica et Cosmochimica Acta 63, 30393048.Google Scholar
Appelo, C. A. J. and Postma, D., 2005, Geochemistry, Groundwater, and Pollution, 2nd ed. Balkema, Rotterdam.Google Scholar
Arnórsson, S., Gunnlaugsson, E. and Svavarsson, H., 1983, The chemistry of geothermal waters in Iceland, II. Mineral equilibria and independent variables controlling water compositions. Geochimica et Cosmochimica Acta 47, 547566.Google Scholar
Arnórsson, S., Sigurdsson, S. and Svavarsson, H., 1982, The chemistry of geothermal waters in Iceland. I. Calculation of aqueous speciation from 0° to 370°C. Geochimica et Cosmochimica Acta 46, 15131532.Google Scholar
Atkinson, K., Han, W. and Steward, D., 2009, Numerical Solution of Ordinary Differential Equations. Wiley, New York.Google Scholar
Baccar, M. B. and Fritz, B., 1993, Geochemical modelling of sandstone diagenesis and its consequences on the evolution of porosity. Applied Geochemistry 8, 285295.Google Scholar
Baes, C. F., Jr. and Mesmer, R. E., 1976, The Hydrolysis of Cations. Wiley, New York.Google Scholar
Bahr, J. M. and Rubin, J., 1987, Direct comparison of kinetic and local equilibrium formulations for solute transport affected by surface reactions. Water Resources Research 23, 438452.CrossRefGoogle Scholar
Ball, , J. W, Jenne, E. A. and Nordstrom, D. K., 1979, WATEQ2—a computerized chemical model for trace and major element speciation and mineral equilibria of natural waters. In Jenne, E. A. (ed.), Chemical Modeling in Aqueous Systems, American Chemical Society, Washington DC, pp. 815835.Google Scholar
Ball, J. W. and Nordstrom, D. K., 1991, User’s manual for WATEQ4F, with revised thermodynamic data base and test cases for calculating speciation of major, trace, and redox elements in natural waters. US Geological Survey Open File Report 91–183.Google Scholar
Banfield, J. F. and Nealson, K. H. (eds.), 1997, Geomicrobiology: Interactions Between Microbes and Minerals. Reviews in Mineralogy 35, Mineralogical Society of America, Washington DC.Google Scholar
Barton, P. B., Jr., Bethke, P. M. and Toulmin, P., 3rd, 1963, Equilibrium in ore deposits. Mineralogical Society of America Special Paper 1, 171185.Google Scholar
Bear, J., 1972, Dynamics of Fluids in Porous Media. Elsevier, Amsterdam.Google Scholar
Bear, J., 1979, Hydraulics of Groundwater. McGraw-Hill, New York.Google Scholar
Bekins, B. A., Godsy, E. M. and Warren, E., 1999, Distribution of microbial physiologic types in an aquifer contaminated by crude oil. Microbial Ecology 37, 263275.Google Scholar
Bekins, B. A., Warren, E. and Godsy, E. M., 1998, A comparison of zero-order, first-order, and Monod bio-transformation models. Ground Water 36, 261268.Google Scholar
Belitz, K. and Bredehoeft, J. D., 1988, Hydrodynamics of Denver basin, explanation of subnormal fluid pressures. American Association of Petroleum Geologists Bulletin 72, 13341359.Google Scholar
Benjamin, M. A., 2002, Modeling the mass-action expression for bidentate adsorption. Environmental Science and Technology 36, 307313.Google Scholar
Berger, T., Mathurin, F., Gustafsson, J. P., Peltola, P. and Åström, M. E., 2015, The impact of fluoride on Al abundance and speciation in boreal streams. Chemical Geology 409, 118124.Google Scholar
Berner, R. A., 1980, Early Diagenesis, A Theoretical Approach. Princeton University Press, Princeton, New Jersey.Google Scholar
Bethke, C. M., 1992, The question of uniqueness in geochemical modeling. Geochimica et Cosmochimica Acta 56, 43154320.Google Scholar
Bethke, C. M., 1997, Modelling transport in reacting geochemical systems. Comptes Rendus de l’Académie des Sciences 324, 513528.Google Scholar
Bethke, C. M. and Brady, P. V., 2000, How the Kd approach undermines groundwater cleanup. Ground Water 38, 435443.Google Scholar
Bethke, C. M., Ding, D., Jin, Q. and Sanford, R. A., 2008, Origin of microbiological zoning in groundwater flows. Geology 36, 739742.Google Scholar
Bethke, C. M., Farrell, B. and Sharifi, M., 2021, The Geochemist’s Workbench® Release 16 (five volumes). Aqueous Solutions LLC, Champaign, Illinois.Google Scholar
Bethke, C. M., Harrison, W. J., Upson, C. and Altaner, S. P., 1988, Supercomputer analysis of sedimentary basins. Science 239, 261267.Google Scholar
Bethke, C. M., Lee, M.-K. and Wendlandt, R. F., 1992, Mass transport and chemical reaction in sedimentary basins, natural and artificial diagenesis. In Quintard, M. and Todorovic, M. S. (eds.), Heat and Mass Transfer in Porous Media. Elsevier, Amsterdam, pp. 421434.Google Scholar
Bethke, C. M. and Marshak, S., 1990, Brine migrations across North America—the plate tectonics of groundwater. Annual Review Earth and Planetary Sciences 18, 287315.Google Scholar
Bethke, C. M., Sanford, R. A., Kirk, M. F., Jin, Q. and Flynn, T. M., 2011, The thermodynamic ladder in geomicrobiology. American Journal of Science 311, 183210.Google Scholar
Bibby, R., 1981, Mass transport of solutes in dual-porosity media. Water Resources Research 17, 10751081.Google Scholar
Bird, R. B., Stewart, W. E. and Lightfoot, E. N., 2006, Transport Phenomena, revised 2nd ed. Wiley, New York.Google Scholar
Bischoff, J. L., Fitzpatrick, J. A. and Rosenbauer, R. J., 1993, The solubility and stabilization of ikaite (CaCO3 · 6H2O) from 0 °C to 25 °C, environmental and paleoclimatic implications for thinalite tufa. Journal of Geology 101, 2133.Google Scholar
Bischoff, J. L., Herbst, D. B. and Rosenbauer, R. J., 1991, Gaylussite formation at Mono Lake, California. Geochimica et Cosmochimica Acta 55, 17431747.CrossRefGoogle Scholar
Bjorlykke, K. and Egeberg, P. K., 1993, Quartz cementation in sedimentary basins. American Association of Petroleum Geologists Bulletin 77, 15381548.Google Scholar
Blesa, M. A., Figliolia, N. M., Maroto, A. J. G. and Regazzoni, A. E., 1984, The influence of temperature on the interface magnetite–aqueous electrolyte solution. Journal of Colloid and Interface Science 101, 410418.Google Scholar
Block, J. and Waters, O. B., Jr., 1968, The CaSO4–Na2SO4–NaCl–H2O system at 25 °C to 100 °C. Journal of Chemical and Engineering Data 13, 336344.Google Scholar
Blowes, D. W., Ptacek, C. J., Jambor, J. L. and Weisener, C. G., 2005, The geochemistry of acid mine drainage. In Lollar, B. S. (ed.), Environmental Geochemistry, Elsevier, Amsterdam, pp. 149204.Google Scholar
Blum, A. E. and Stillings, L. L., 1995, Feldspar dissolution kinetics. Reviews in Mineralogy 31, 291351.Google Scholar
Blum, J. S., Bindi, A. B., Buzelli, J., Stolz, J. F. and Oremland, R. S., 1998, Bacillus arsenicoselenatis, sp. nov., and Bacillus selenitireducens, sp. nov.: two haloalkaliphiles from Mono Lake, California that respire oxyanions. Archives Microbiology 171, 1930.Google Scholar
Bogard, M. J. and del Giorgio, P. A., 2016, The role of metabolism in modulating CO2 fluxes in boreal lakes. Global Biogeochemical Cycles 30, 15091525.Google Scholar
Bolt, G. H. and Van Riemsdijk, W. H., 1982, Ion adsorption on inorganic variable charge constituents. In G. H. Bolt (ed.), Soil Chemistry, Physico-Chemical Models, B., 2nd ed. Elsevier, Amsterdam, pp. 459504.Google Scholar
Boudart, M, 1976, Consistency between kinetics and thermodynamics. Journal of Physical Chemistry 80, 28692870.Google Scholar
Boulding, J. R., 1990, Assessing the geochemical fate of deep-well-injected hazardous waste. US Environmental Protection Agency Report EPA/625/6–89/025a.Google Scholar
Bourcier, W. L., 1985, Improvements to the solid solution modeling capabilities of the EQ3/6 geochemical code. Lawrence Livermore National Laboratory Report UCID-20587.Google Scholar
Bowers, T. S. and Taylor, H. P., Jr., 1985, An integrated chemical and stable isotope model of the origin of midocean ridge hot spring systems. Journal of Geophysical Research 90, 1258312606.Google Scholar
Bowers, T. S., Von Damm, K. L. and Edmond, J. H., 1985, Chemical evolution of mid-ocean ridge hot springs. Geochimica et Cosmochimica Acta 49, 22392252.Google Scholar
Boynton, F. P., 1960, Chemical equilibrium in multicomponent polyphase systems. Journal of Chemical Physics 32, 18801881.Google Scholar
Brady, J. B., 1975, Chemical components and diffusion. American Journal of Science 275, 10731088.Google Scholar
Brady, P. V. and Bethke, C. M., 2000, Beyond the Kd approach. Ground Water 38, 321322.Google Scholar
Brady, P. V., Brady, M. V. and Borns, D. J., 1998, Natural Attenuation, CERCLA, RBCA’s, and the Future of Environmental Remediation. CRC Press, Boca Raton, Florida.Google Scholar
Brady, P. V. and Walther, J. V., 1989, Controls on silicate dissolution rates in neutral and basic pH solutions at 25 °C. Geochimica et Cosmochimica Acta 53, 28232830.Google Scholar
Brantley, S. L., 1992, Kinetics of dissolution and precipitation—experimental and field results. In Kharaka, Y. K. and Maest, A. S. (eds.), Water–Rock Interaction. Balkema, Rotterdam, pp. 36.Google Scholar
Brantley, S. L., Crerar, D. A., Møller, N. E. and Weare, J. H., 1984, Geochemistry of a modern marine evaporite, Bocana de Virrilá, Peru. Journal of Sedimentary Petrology 54, 447462.Google Scholar
Brantley, S. L., Kubicki, J. D. and White, A. F. (eds.), 2008, Kinetics of Water-Rock Interaction. Springer-Verlag, New York.Google Scholar
Brendler, V., Vahle, A., Arnold, T., Bernhard, G. and Fanghänel, T., 2003, RES3T-Rossendorf expert system for surface and sorption thermodynamics. Journal of Contaminant Hydrology 61, 281291.Google Scholar
Brewer, P. G., 1975, Minor elements in sea water. In Riley, J. P. and Skirrow, G. (eds.), Chemical Oceanography. Academic Press, New York.Google Scholar
Brezonik, P. L., 1994, Chemical Kinetics and Process Dynamics in Aquatic Systems. Lewis Publishers, Boca Raton, Florida.Google Scholar
Brinkley, S. R., Jr., 1947, Calculation of the equilibrium composition of systems of many components. Journal of Chemical Physics 15, 107110.Google Scholar
Brinkley, S. R., Jr., 1960, Discussion of “A brief survey of past and curent methods of solution for equilibrium composition” by H. E. Brandmaier and J. J. Harnett. In Bahn, G. S. and Zukoski, E. E. (eds.), Kinetics, Equilibria and Performance of High Temperature Systems. Butterworths, Washington DC, p. 73.Google Scholar
Brower, R. D., Visocky, A. P., Krapac, I. G. et al., 1989, Evaluation of Underground Injection of Industrial Waste in Illinois. Illinois Scientific Surveys Joint Report 2.Google Scholar
Brown, T. H. and Skinner, B. J., 1974, Theoretical prediction of equilibrium phase assemblages in multicomponent systems. American Journal of Science 274, 961986.Google Scholar
Bunge, A. L. and Radke, C. J., 1982, Migration of alkaline pulses in reservoir sands. Society of Petroleum Engineers Journal 22, 9981012.Google Scholar
Butler, G. P., 1969, Modern evaporite deposition and geochemistry of coexisting brines, the sabkha, Trucial Coast, Arabian Gulf. Journal of Sedimentary Petrology 39, 7089.Google Scholar
Carnahan, B., Luther, H. A. and Wilkes, J. O., 1969, Applied Numerical Methods. Wiley, New York.Google Scholar
Carpenter, A. B., 1980, The chemistry of dolomite formation I: the stability of dolomite. Society of Economic Paleontologists and Mineralogists Special Publication 28, 111121.Google Scholar
Carroll, S. A. and Walther, J. V., 1990, Kaolinite dissolution at 25 °C, 60 °C, and 80 °C. American Journal of Science 290, 797810.Google Scholar
Casey, W. H. and Rustad, J. R., 2007, Reaction dynamics, molecular clusters, and aqueous geochemistry. Annual Review of Earth and Planetary Science 35, 2146.Google Scholar
Casey, W. H. and Swaddle, T. W., 2003, Why small? The use of small inorganic clusters to understand mineral surface and dissolution reactions in geochemistry. Reviews of Geophysics 41, 4-1–4-20.Google Scholar
Cederberg, G. A., Street, R. L. and Leckie, J. O., 1985, A groundwater mass transport and equilibrium chemistry model for multicomponent systems. Water Resources Research 21, 10951104.Google Scholar
Chapelle, F. H., 2001, Ground-Water Microbiology and Geochemistry, 2nd ed. Wiley, New York.Google Scholar
Chapelle, F. H. and Lovley, D. R., 1992, Competitive exclusion of sulfate reduction by Fe(III)-reducing bacteria: a mechanism for producing discrete zones of high-iron ground water. Ground Water 30, 2936.CrossRefGoogle Scholar
Chapman, B. M., Jones, D. R. and Jung, R. F., 1983, Processes controlling metal ion attenuation in acid mine drainage streams. Geochimica et Cosmochimica Acta 47, 19571973.Google Scholar
Cheng, H. P. and Yeh, G. T., 1998, Development of a three-dimensional model of subsurface flow, heat transfer, and reactive chemical transport: 3DHYDROGEOCHEM. Journal of Contaminant Hydrology 34, 4783.Google Scholar
Chorover, J. and Brusseau, M. L., 2008, Kinetics of sorption–desorption. In Brantley, S. L., Kubicky, J. D. and White, A. F. (eds.), Kinetics of Water-Rock Interaction. Springer, New York, pp. 109149.Google Scholar
Cicconi, M. R., Moretti, R. and Neuville, D. R., 2020, Earth’s electrodes. Elements 16, 157160.Google Scholar
Clayton, J. L. and Swetland, P. J., 1980, Petroleum generation and migration in Denver basin. American Association of Petroleum Geologists Bulletin 64, 16131633.Google Scholar
Cole, D. R. and Ohmoto, H., 1986, Kinetics of isotopic exchange at elevated temperatures and pressures. Reviews in Mineralogy 16, 4190.Google Scholar
Coudrain-Ribstein, A. and Jamet, P., 1989, Choix des composantes et spéciation d’une solution. Comptes Rendus de l’Académie des Sciences 309-II, 239–244.Google Scholar
Cox, B. G., 1994, Modern Liquid Phase Kinetics. Oxford University Press.Google Scholar
Crank, J., 1975, The Mathematics of Diffusion, 2nd ed. Oxford University Press.Google Scholar
Crerar, D. A., 1975, A method for computing multicomponent chemical equilibria based on equilibrium constants. Geochimica et Cosmochimica Acta 39, 13751384.Google Scholar
Criaud, A., Fouillac, C. and Marty, B., 1989, Low enthalpy geothermal fluids from the Paris basin. 2. Oxidation-reduction state and consequences for the prediction of corrosion and sulfide scaling. Geothermics 18, 711727.Google Scholar
Davies, C. W., 1962, Ion Association. Butterworths, Washington DC.Google Scholar
Davis, A., Olsen, R. L. and Walker, D. R., 1991, Distribution of metals between water and entrained sediment in streams impacted by acid mine drainage, Clear Creek, Colorado, U.S.A. Applied Geochemistry 6, 333348.Google Scholar
Davis, J. A., James, R. O. and Leckie, J. O., 1978, Surface ionization and complexation at the oxide/water interface. 1. Computation of electrical double layer properties in simple electrolytes. Journal of Colloid and Interface Science 63, 480499.Google Scholar
Davis, J. A. and Kent, D. B., 1990, Surface complexation modeling in aqueous geochemistry. Reviews in Mineralogy 23, 177260.Google Scholar
Davis, J. A. and Leckie, J. O., 1980, Surface ionization and complexation at the oxide/water interface. 3. Adsorption of anions. Journal of Colloid and Interface Science 74, 3243.Google Scholar
Degens, E. T. and Ross, D. A. (eds.), 1969, Hot Brines and Recent Heavy Metal Deposits in the Red Sea. Springer-Verlag, New York.CrossRefGoogle Scholar
Delany, J. M. and Lundeen, S. R., 1989, The LLNL thermochemical database. Lawrence Livermore National Laboratory Report UCRL-21658.Google Scholar
Delany, J. M. and Wolery, T. J., 1984, Fixed-fugacity option for the EQ6 geochemical reaction path code. Lawrence Livermore National Laboratory Report UCRL-53598.Google Scholar
Deloule, E., 1982, The genesis of fluorspar hydrothermal deposits at Montroc and Le Burc, The Tarn, as deduced from fluid inclusion analysis. Economic Geology 77, 18671874.Google Scholar
Denbigh, K., 1981, The Principles of Chemical Equilibrium, 4th ed. Cambridge University Press.Google Scholar
Domenico, P. A. and Schwartz, F. W., 1998, Physical and Chemical Hydrogeology, 2nd ed. Wiley, New York.Google Scholar
Dongarra, J. J., Moler, C. B., Bunch, J. R. and Stewart, G. W., 1979, Linpack Users’ Guide. Society for Industrial and Applied Mathematics, Philadelphia.CrossRefGoogle Scholar
Dove, P. M. and Crerar, D. A., 1990, Kinetics of quartz dissolution in electrolyte solutions using a hydrothermal mixed flow reactor. Geochimica et Cosmochimica Acta 54, 955969.Google Scholar
Drever, J. I., 1988, The Geochemistry of Natural Waters, 2nd ed. Prentice-Hall, Englewood Cliffs, New Jersey.Google Scholar
Dria, M. A., Schedchter, R. S. and Lake, L. W., 1988, An analysis of reservoir chemical treatments. SPE Production Engineering 3, 5262.Google Scholar
Driscoll, C. T. Jr., Baker, J. P., Bisogni, J. J., Jr. and Schofieldt, C. L., 1980, Effect of aluminium speciation on fish in dilute acidified waters. Nature 284, 161164.CrossRefGoogle Scholar
Druhan, J. and Tournassat, C. (eds.), 2019, Reactive Transport in Natural and Engineered Systems. Reviews in Mineralogy and Geochemistry 85, Mineralogical Society of America, Washington DC.Google Scholar
Drummond, S. E. and Ohmoto, H., 1985, Chemical evolution and mineral deposition in boiling hydrothermal systems. Economic Geology 80, 126147.Google Scholar
Dzombak, D. A. and Morel, F. M. M., 1987, Adsorption of inorganic pollutants in aquatic systems. Journal of Hydraulic Engineering 113, 430475.Google Scholar
Dzombak, D. A. and Morel, F. M. M., 1990, Surface Complexation Modeling. Wiley, New York.Google Scholar
Esters, L., Landwehr, S., Sutherland, G. et al., 2017, Parameterizing air-sea gas transfer velocity with dissipation. Journal of Geophysical Research: Oceans 122, 30413056.Google Scholar
Eugster, H. P., Harvie, C. E. and Weare, J. H., 1980, Mineral equilibria in the six-component seawater system, Na–K–Mg–Ca–SO4–Cl–H2O, at 25 °C. Geochimica et Cosmochimica Acta 44, 13351347.Google Scholar
Eugster, H. P. and Jones, B. F., 1979, Behavior of major solutes during closed-basin brine evolution. American Journal of Science 279, 609631.Google Scholar
Faure, G., 1986, Principles of Isotope Geology, 2nd ed. Wiley, New York.Google Scholar
Felmy, A. R. and Weare, J. H., 1986, The prediction of borate mineral equilibria in natural waters: application to Searles Lake, California. Geochimica et Cosmochimica Acta 50, 27712783.Google Scholar
Ficklin, W. H., Plumlee, G. S., Smith, K. S. and McHugh, J. B., 1992, Geochemical classification of mine drainages and natural drainages in mineralized areas. In Kharaka, Y. K. and Maest, A. S. (eds.), Water– Rock Interaction. Balkema, Rotterdam, pp. 381384.Google Scholar
Fouke, B. W., Bonheyo, G. T., Sanzenbacher, E. and Frias-Lopez, J., 2003, Partitioning of bacterial communities between travertine depositional facies at Mammoth Hot Springs, Yellowstone National Park, USA. Canadian Journal Earth Sciences 40, 15311548.Google Scholar
Fournier, R. O., 1977, Chemical geothermometers and mixing models for geothermal systems. Geothermics 5, 4150.Google Scholar
Fournier, , R. O and Potter, R. W. II, 1979, Magnesium correction to the Na-K-Ca chemical geothermometer. Geochimica et Cosmochimica Acta 43, 15431550.Google Scholar
Fournier, R. O. and Rowe, J. J., 1966, Estimation of underground temperatures from the silica content of water from hot springs and wet-steam wells. American Journal of Science 264, 685697.Google Scholar
Fournier, R. O. and Truesdell, A. H., 1973, An empirical Na-K-Ca geothermometer for natural waters. Geochimica et Cosmochimica Acta 37, 12551275.Google Scholar
Freeze, R. A. and Cherry, J. A., 1979, Groundwater. Prentice Hall, Englewood Cliffs, New Jersey.Google Scholar
Ganguly, J., 2020, Thermodynamics in Earth and Planetary Sciences, 2nd ed., Cambridge University Press.Google Scholar
Garrels, R. M. and Mackenzie, F. T., 1967, Origin of the chemical compositions of some springs and lakes. Equilibrium Concepts in Natural Waters, Advances in Chemistry Series 67, American Chemical Society, Washington DC, pp. 222–242.Google Scholar
Garrels, R. M. and Thompson, M. E., 1962, A chemical model for sea water at 25 °C and one atmosphere total pressure. American Journal of Science 260, 5766.Google Scholar
Garven, G. and Freeze, R. A., 1984, Theoretical analysis of the role of groundwater flow in the genesis of stratabound ore deposits: 2, Quantitative results. American Journal of Science 284, 11251174.Google Scholar
Gelhar, L. W., 1986, Stochastic subsurface hydrology from theory to applications. Water Resources Research 22, 135-S–145-S.Google Scholar
Gerke, H. H. and van Genuchten, M. T., 1993, A dual-porosity model for simulating the preferential movement of water and solutes in structured porous media. Water Resources Research 29, 305319.Google Scholar
Giffaut, E., Grivé, M., Blanc, Ph. et al., 2014, Andra thermodynamic database for performance assessment: ThermoChimie. Applied Geochemistry 49, 225236.Google Scholar
Giggenbach, W. F., 1988, Geothermal solute equilibria, derivation of Na-K-Mg-Ca geoindicators. Geochimica et Cosmochimica Acta 52, 27492765.Google Scholar
Glynn, P. D., Reardon, E. J., Plummer, L. N. and Busenberg, E., 1990, Reaction paths and equilibrium end-points in solid-solution aqueous-solution systems. Geochimica et Cosmochimica Acta 54, 267282.Google Scholar
Greenberg, J. P. and Møller, N., 1989, The prediction of mineral solubilities in natural waters, a chemical equilibrium model for the Na–K–Ca–Cl–SO4–H2O system to high concentration from 0 to 250 °C. Geochimica et Cosmochimica Acta 53, 25032518.Google Scholar
Greenwood, H. J., 1975, Thermodynamically valid projections of extensive phase relationships. American Mineralogist 60, 18.Google Scholar
Grenthe, I., Plyasunov, A. V. and Spahiu, K., 1997, Estimations of medium effects on thermodynamic data. In Grenthe, I. and Puigdomènech, I. (eds.), Modelling in Aquatic Chemistry, NEA OECD Publications, Paris, 325426.Google Scholar
Guggenheim, E. A., 1967, Thermodynamics, an Advanced Treatment for Chemists and Physicists, 5th ed. North-Holland, Amsterdam.Google Scholar
Gunnarsson, M., Abbas, Z., Ahlberg, E., Gobom, S. and Nordholm, S., 2002, Corrected Debye–Hückel analysis of surface complexation, II. A theory of surface charging. Journal of Colloid and Interface Science 249, 5261.Google Scholar
Gupta, S. S. and Bhattacharyya, K. G., 2011, Kinetics of adsorption of metal ions on inorganic materials: a review. Advances in Colloid and Interface Science 162, 3958.Google Scholar
Haas, J. L., Jr. and Fisher, J. R., 1976, Simultaneous evaluation and correlation of thermodynamic data. American Journal of Science 276, 525545.Google Scholar
Hardie, L. A., 1987, Dolomitization, a critical view of some current views. Journal of Sedimentary Petrology 57, 166183.Google Scholar
Hardie, L. A., 1991, On the significance of evaporites. Annual Review Earth and Planetary Sciences 19, 131168.Google Scholar
Hardie, L. A. and Eugster, H. P., 1970, The evolution of closed-basin brines. Mineralogical Society of America Special Paper 3, 273290.Google Scholar
Harrison, W. J., 1990, Modeling fluid/rock interactions in sedimentary basins. In Cross, T. A. (ed.), Quantitative Dynamic Stratigraphy. Prentice Hall, Englewood Cliffs, New Jersey, pp. 195231.Google Scholar
Harvie, C. E., Greenberg, J. P. and Weare, J. H., 1987, A chemical equilibrium algorithm for highly non-ideal multiphase systems: free energy minimization. Geochimica et Cosmochimica Acta 51, 10451057.Google Scholar
Harvie, C. E., Møller, N. and Weare, J. H., 1984, The prediction of mineral solubilities in natural waters: the Na–K–Mg–Ca–H–Cl–SO4–OH–HCO3-CO3–CO2–H2O system to high ionic strengths at 25 °C. Geochimica et Cosmochimica Acta 48, 723751.Google Scholar
Harvie, C. E. and Weare, J. H., 1980, The prediction of mineral solubilities in natural waters: the Na–K–Mg– Ca–Cl–SO4–H2O system from zero to high concentration at 25 °C. Geochimica et Cosmochimica Acta 44, 981997.Google Scholar
Harvie, C. E., Weare, J. H., Hardie, L. A. and Eugster, H. P., 1980, Evaporation of seawater: calculated mineral sequences. Science 208, 498500.Google Scholar
Hay, R. L., 1963, Stratigraphy and zeolitic diagenesis of the John Day formation of Oregon. University of California Publications in Geological Sciences, Berkeley, California, pp. 199261.Google Scholar
Hay, R. L., 1966, Zeolites and zeolitic reactions in sedimentary rocks. Geological Society of America Special Paper 85.Google Scholar
Hayes, J. B., 1979, Sandstone diagenesis—the hole truth. Society of Economic Paleontologists and Mineralogists Special Publication 26, 127139.Google Scholar
Hayes, K. F. and Leckie, J. O., 1987, Mechanism of lead ion adsorption at the goethite–water interface. ACS Symposium Series 323, 114141.Google Scholar
Hayes, K. F., Papelis, C. and Leckie, J. O., 1988, Modeling ionic strength effects on anion adsorption at hydrous oxide/solution interfaces. Journal of Colloid and Interface Science 125, 717726.Google Scholar
Hayes, K. F., Redden, G., Wendell, E. and Leckie, J. O., 1991, Surface complexation models: an evaluation of model parameter estimation using FITEQL and oxide mineral titration data. Journal of Colloid and Interface Science 142, 448469.Google Scholar
He, S., Oddo, J. E. and Tomson, M. B., 1994, The inhibition of gypsum and barite nucleation in NaCl brines at temperatures from 25 to 90 °C. Applied Geochemistry 9, 561567.Google Scholar
Healy, R. W., Haile, S. S., Parkhurst, D. L. and Charlton, S. R., 2018, VS2DRTI simulating heat and reactive solute transport in variably saturated porous media. Groundwater 56, 810815.Google Scholar
Helgeson, H. C., 1968, Evaluation of irreversible reactions in geochemical processes involving minerals and aqueous solutions, I. Thermodynamic relations. Geochimica et Cosmochimica Acta 32, 853877.Google Scholar
Helgeson, H. C., 1969, Thermodynamics of hydrothermal systems at elevated temperatures and pressures. American Journal of Science 267, 729804.Google Scholar
Helgeson, H. C., 1970, A chemical and thermodynamic model of ore deposition in hydrothermal systems. Mineralogical Society of America Special Paper 3, 155186.Google Scholar
Helgeson, H. C., Brown, T. H., Nigrini, A. and Jones, T. A., 1970, Calculation of mass transfer in geochemical processes involving aqueous solutions. Geochimica et Cosmochimica Acta 34, 569592.Google Scholar
Helgeson, H. C., Delany, J. M., Nesbitt, H. W. and Bird, D. K., 1978, Summary and critique of the thermodynamic properties of rock-forming minerals. American Journal of Science 278-A, 1229.Google Scholar
Helgeson, H. C., Garrels, R. M. and Mackenzie, F. T., 1969, Evaluation of irreversible reactions in geochemical processes involving minerals and aqueous solutions, II. Applications. Geochimica et Cosmochimica Acta 33, 455481.Google Scholar
Helgeson, H. C. and Kirkham, D. H., 1974, Theoretical prediction of the thermodynamic behavior of aqueous electrolytes at high pressures and temperatures, II. Debye–Hückel parameters for activity coefficients and relative partial molal properties. American Journal of Science 274, 11991261.Google Scholar
Helgeson, H. C., Kirkham, D. H. and Flowers, G. C., 1981, Theoretical prediction of the thermodynamic behavior of aqueous electrolytes at high temperatures and pressures, IV. Calculation of activity coefficients, osmotic coefficients, and apparent molal and standard and relative partial molal properties to 600 °C and 5 kB. American Journal of Science 281, 12491516.Google Scholar
Hem, J. D., 1985, Study and interpretation of the chemical characteristics of natural water. US Geological Survey Water-Supply Paper 2254.Google Scholar
Hemley, J. J., Cygan, G. L. and d’Angelo, W. M., 1986, Effect of pressure on ore mineral solubilities under hydrothermal conditions. Geology 14, 377379.Google Scholar
Henley, R. W., 1984, Chemical structure of geothermal systems. Reviews in Economic Geology 1, 928.Google Scholar
Hiemstra, T. and Van Riemsdijk, W. H., 1996, A surface structural approach to ion adsorption: the charge distribution (CD) model. Journal of Colloid and Interface Science 179, 488508.Google Scholar
Hill, C. G., Jr., 1977, An Introduction to Chemical Engineering Kinetics and Reactor Design. Wiley, New York.Google Scholar
Hoffmann, R., 1991, Hot brines in the Red Sea. American Scientist 79, 298299.Google Scholar
Holland, H. D., 1978, The Chemistry of the Atmosphere and Oceans. Wiley, New York.Google Scholar
Hostettler, J. D., 1984, Electrode electrons, aqueous electrons, and redox potentials in natural waters. American Journal of Science 284, 734759.Google Scholar
Hubbert, M. K., 1940, The theory of ground-water motion. Journal of Geology 48, 785944.Google Scholar
Huber, J. A., Butterfield, D. A. and Baross, J. A., 2003, Bacterial diversity in a subseafloor habitat following a deep-sea volcanic eruption. FEMS Microbiology Ecology 43, 393409.Google Scholar
Hubert, J. F., 1960, Petrology of the Fountain and Lyons formations, Front Range, Colorado. Colorado School of Mines Quarterly 55.Google Scholar
Hunt, J. M., 1990, Generation and migration of petroleum from abnormally pressured fluid compartments. American Association of Petroleum Geologists Bulletin 74, 112.Google Scholar
Huser, B. A., Wuhrmann, K. and Zehnder, A. J. B., 1982, Methanothrix soehngenii gen. nov. sp. nov., a new acetotrophic non-hydrogen-oxidizing methane bacterium. Archives of Microbiology 132, 19.Google Scholar
Hutcheon, I., 1984, A review of artificial diagenesis during thermally enhanced recovery. In MacDonald, D. A. and Surdam, R. C. (eds.), Clastic Diagenesis. American Association of Petroleum Geologists, Tulsa, Oklahoma, pp. 413429.Google Scholar
HZDR, 2021, RES3T-Rossendorf Expert System for Surface and Sorption Thermodynamics. Helmholtz-Zentrum Dresden-Rossendorf, www.hzdr.de/res3t.Google Scholar
Ingvorsen, K., Zehnder, A. J. B. and Jørgensen, B. B., 1984, Kinetics of sulfate and acetate uptake by Desulfobacter postgatei. Applied and Environmental Microbiology 47, 403408.Google Scholar
Publishers, Interscience, 1954, The Collected Papers of P. J. W. Debye. Interscience Publishers, Inc., New York.Google Scholar
ISO, 1985, International Standard ISO–7888: Water quality–Determination of electrical conductivity. International Organization for Standardization, Geneva, Switzerland.Google Scholar
Janecky, D. R. and Seyfried, W. E., Jr., 1984, Formation of massive sulfide deposits on oceanic ridge crests: incremental reaction models for mixing between hydrothermal solutions and seawater. Geochimica et Cosmochimica Acta 48, 27232738.Google Scholar
Janecky, D. R. and Shanks, W. C., III, 1988, Computational modeling of chemical and sulfur isotopic reaction processes in seafloor hydrothermal systems, chimneys, massive sulfides, and subjacent alteration zones. Canadian Mineralogist 26, 805825.Google Scholar
Jankowski, J. and Jacobson, G., 1989, Hydrochemical evolution of regional groundwaters to playa brines in central Australia. Journal of Hydrology 108, 123173.Google Scholar
Jarraya, F. and El Mansar, M., 1987, Modelisation Simplifié du Gisement de Saumare á Sebkhat el Melah á Zarzis. Projet de fin d’etudes, Ecole Nationale d’Ingénieurs de Tunis, Tunis, Tunisia.Google Scholar
Javandel, I., Doughty, C. and Tsang, C. F., 1984, Groundwater Transport: Handbook of Mathematical Models. American Geophysical Union, Washington DC.Google Scholar
Jenne, E. A. (ed.), 1998, Adsorption of Metals by Geomedia. Academic Press, New York.Google Scholar
Jennings, H. Y., Jr., Johnson, C. E., Jr. and McAuliffe, C. D., 1974, A caustic waterflooding process for heavy oils. Journal of Petroleum Technology 26, 13441352.Google Scholar
Jin, Q., 2007, Control of hydrogen partial pressures on the rates of syntrophic microbial metabolisms: a kinetic model for butyrate fermentation. Geobiology 5, 3548.Google Scholar
Jin, Q. and Bethke, C. M., 2002, Kinetics of electron transfer through the respiratory chain. Biophysical Journal 83, 17971808.Google Scholar
Jin, Q. and Bethke, C. M., 2003, A new rate law describing microbial respiration. Applied and Environmental Microbiology 69, 23402348.Google Scholar
Jin, Q. and Bethke, C. M., 2005, Predicting the rate of microbial respiration in geochemical environments. Geochimica et Cosmochimica Acta 69, 11331143.Google Scholar
Jin, Q. and Bethke, C. M., 2007, The thermodynamics and kinetics of microbial metabolism. American Journal of Science 307, 643677.Google Scholar
Jin, Q. and Bethke, C. M., 2009, Cellular energy conservation and the rate of microbial sulfate reduction. Geology 37, 10271030.Google Scholar
Johnson, C. A., 1986, The regulation of trace element concentrations in river and estuarine waters with acid mine drainage, the adsorption of Cu and Zn on amorphous Fe oxyhydroxides. Geochimica et Cosmochimica Acta 50, 24332438.Google Scholar
Johnson, J. W., Oelkers, E. H. and Helgeson, H. C., 1991, SUPCRT92: a software package for calculating the standard molal thermodynamic properties of minerals, gases, aqueous species, and reactions from 1 to 5000 bars and 0° to 1000 °C. Earth Sciences Department, Lawrence Livermore Laboratory.Google Scholar
Jurkevitch, E., 2007, Predatory behaviors in bacteria—diversity and transitions. Microbe 2, 6773.Google Scholar
Karamalidis, A. K. and Dzombak, D. A., 2010, Surface Complexation Modeling: Gibbsite. Wiley, New York.Google Scholar
Karpov, I. K. and Kaz’min, L. A., 1972, Calculation of geochemical equilibria in heterogeneous multicomponent systems. Geochemistry International 9, 252262.Google Scholar
Karpov, I. K., Kaz’min, L. A. and Kashik, S. A., 1973, Optimal programming for computer calculation of irreversible evolution in geochemical systems. Geochemistry International 10, 464470.Google Scholar
Kashefi, K. and Lovley, D. R., 2003, Extending the upper temperature limit for life. Science 301, 934.Google Scholar
Kastner, M., 1984, Control of dolomite formation. Nature 311, 410411.Google Scholar
Keenan, J. H., Keyes, F. G., Hill, P. G. and Moore, J. G., 1969, Steam Tables, Thermodynamic Properties of Water Including Vapor, Liquid, and Solid Phases. Wiley, New York.Google Scholar
Kelley, D. S., Baross, J. A. and Delaney, J. R., 2002, Volcanoes, fluids, and life at mid-ocean ridge spreading centers. Annual Review of Earth and Planetary Sciences 30, 385491.Google Scholar
Kennedy, V. C., Zellweger, G. W. and Jones, B. F., 1974, Filter pore-size effects on the analysis of Al, Fe, Mn, and Ti in water. Water Resources Research 10, 785790.Google Scholar
Kharaka, Y. K. and Barnes, I., 1973, SOLMNEQ: solution-mineral equilibrium computations. US Geological Survey Computer Contributions Report PB-215-899.Google Scholar
Kharaka, Y. K., Gunter, W. D., Aggarwal, P. K., Perkins, E. H. and DeBraal, J. D., 1988, SOLMINEQ.88, a computer program for geochemical modeling of water–rock interactions. US Geological Survey Water Resources Investigation Report 88–4227.Google Scholar
Kim, C., Zhou, Q., Deng, B., Thornton, E. C. and Xu, H., 2001, Chromium(VI) reduction by hydrogen sulfide in aqueous media: stoichiometry and kinetics. Environmental Science and Technology 35, 22192225.Google Scholar
Kim, S.-B., Hwang, I., Kim, D.-J., Lee, S. and Jury, W. A., 2003, Effect of sorption on benzene biodegradation in sandy soil. Environmental Toxicology and Chemistry 22, 23062311.Google Scholar
King, T. V. V. (ed.), 1995, Environmental considerations of active and abandoned mine lands: lessons from Summitville, Colorado. US Geological Survey Bulletin 2220.Google Scholar
Kirk, M. F., Holm, T. R., Park, J. et al., 2004, Bacterial sulfate reduction limits natural arsenic contamination of groundwater. Geology 32, 953956.Google Scholar
Knapp, R. B., 1989, Spatial and temporal scales of local equilibrium in dynamic fluid-rock systems. Geochimica et Cosmochimica Acta 53, 19551964.Google Scholar
Knauss, K. G. and Wolery, T. J., 1986, Dependence of albite dissolution kinetics on pH and time at 25 °C and 70 °C. Geochimica et Cosmochimica Acta 50, 24812497.Google Scholar
Knauss, K. G. and Wolery, T. J., 1988, The dissolution kinetics of quartz as a function of pH and time at 70 °C. Geochimica et Cosmochimica Acta 52, 4353.Google Scholar
Konhauser, K., 2007, Introduction to Geomicrobiology. Blackwell, Malden, Massachusetts.Google Scholar
Kulik, D. A., Wagner, T., Dmytrieva, S. V. et al., 2013, GEM-SELEKTOR geochemical modeling package: numerical kernel GEMS3K for coupled simulation codes. Computational Geosciences 17, 124.Google Scholar
Lafon, G. M., Otten, G. A. and Bishop, A. M., 1992, Experimental determination of the calcite-dolomite equilibrium below 200 °C; revised stabilities for dolomite and magnesite support near-equilibrium dolomitization models. Geological Society of America Abstracts with Programs 24, A210-A211.Google Scholar
Land, L. S. and Macpherson, G. L., 1992, Geothermometry from brine analyses: lessons from the Gulf Coast, U.S.A. Applied Geochemistry 7, 333340.Google Scholar
Lasaga, A. C., 1981a, Rate laws of chemical reactions. In Lasaga, A. C. and Kirkpatrick, R. J. (eds.), Kinetics of Geochemical Processes. Mineralogical Society of America, Washington DC, pp. 168.Google Scholar
Lasaga, A. C., 1981b, Transition state theory. In Lasaga, A. C. and Kirkpatrick, R. J. (eds.), Kinetics of Geochemical Processes. Mineralogical Society of America, Washington DC, pp. 135169.Google Scholar
Lasaga, A. C., 1984, Chemical kinetics of water–rock interactions. Journal of Geophysical Research 89, 40094025.Google Scholar
Lasaga, A. C., 1998, Kinetic Theory in the Earth Sciences. Princeton University Press, Princeton, New Jersey.Google Scholar
Lasaga, A. C. and Rye, D. M., 1993, Fluid flow and chemical reaction kinetics in metamorphic systems. American Journal of Science 293, 361404.Google Scholar
Lasaga, A. C., Soler, J. M., Ganor, J., Burch, T. E. and Nagy, K. L., 1994, Chemical weathering rate laws and global geochemical cycles. Geochimica et Cosmochimica Acta 58, 23612386.Google Scholar
Leach, D. L., Plumlee, G. S., Hofstra, A. H. et al., 1991, Origin of late dolomite cement by CO2-saturated deep basin brines: evidence from the Ozark region, central United States. Geology 19, 348351.Google Scholar
Leamnson, R. N., Thomas, J., Jr. and Ehrlinger, H. P., III, 1969, A study of the surface areas of particulate microcrystalline silica and silica sand. Illinois State Geological Survey Circular 444.Google Scholar
Lee, M.-K. and Bethke, C. M., 1994, Groundwater flow, late cementation, and petroleum accumulation in the Permian Lyons sandstone, Denver basin. American Association of Petroleum Geologists Bulletin 78, 217237.Google Scholar
Lee, M.-K. and Bethke, C. M., 1996, A model of isotope fractionation in reacting geochemical systems. American Journal of Science 296, 965988.Google Scholar
Levandowski, D. W., Kaley, M. E., Silverman, S. R. and Smalley, R. G., 1973, Cementation in Lyons sandstone and its role in oil accumulation, Denver basin, Colorado. American Association of Petroleum Geologists Bulletin 57, 22172244.Google Scholar
Levenspeil, O., 1972, Chemical Reaction Engineering, 2nd ed. Wiley, New York.Google Scholar
Lichtner, P. C., 1985, Continuum model for simultaneous chemical reactions and mass transport in hydrothermal systems. Geochimica et Cosmochimica Acta 49, 779800.Google Scholar
Lichtner, P. C., 1988, The quasi-stationary state approximation to coupled mass transport and fluid–rock interaction in a porous medium. Geochimica et Cosmochimica Acta 52, 143165.Google Scholar
Lichtner, P. C., 1996, Continuum formulation of multicomponent–multiphase reactive transport. Reviews in Mineralogy 34, 181.Google Scholar
Lichtner, P. C., Oelkers, E. H. and Helgeson, H. C., 1986, Interdiffusion with multiple precipitation/dissolution reactions: transient model and the steady-state limit. Geochimica et Cosmochimica Acta 50, 19511966.Google Scholar
Lichtner, P. C., Steefel, C. I. and Oelkers, E. H. (eds.), 1996, Reactive Transport in Porous Media. Reviews in Mineralogy 34, Mineralogical Society of America, Washington DC.Google Scholar
Lico, M. S., Kharaka, Y. K., Carothers, W. W. and Wright, V. A., 1982, Methods for collection and analysis of geopressured geothermal and oil field waters. US Geological Survey Water Supply Paper 2194.Google Scholar
Liger, E., Charlet, L. and Van Cappellen, P., 1999, Surface catalysis of uranium(VI) reduction by iron(II). Geochimica et Cosmochimica Acta 63, 29392955.Google Scholar
Limousin, G., Gaudet, J.-P., Charlet, L., Szenknect, S., Barthès, V. and Krimissa, M., 2007, Sorption isotherms: a review on physical bases, modeling and measurement. Applied Geochemistry 22, 249275.Google Scholar
Lind, C. J. and Hem, J. D., 1993, Manganese minerals and associated fine particulates in the streambed of Pinal Creek, Arizona, U.S.A.: a mining-related acid drainage problem. Applied Geochemistry 8, 6780.Google Scholar
Lindberg, R. D. and Runnells, D. D., 1984, Groundwater redox reactions: an analysis of equilibrium state applied to Eh measurements and geochemical modeling. Science 225, 925927.Google Scholar
Liss, P. S., 1983, Gas transfer: experiments and geochemical implications. In P. S. Liss and W. G. N. Slinn (eds.), Air-Sea Exchange of Gases and Particles. NATO Science Series C-108, Springer Netherlands, Heidelberg, pp. 241–298.Google Scholar
Liu, C. W. and Narasimhan, T. N., 1989a, Redox-controlled multiple-species reactive chemical transport, 1. Model development. Water Resources Research 25, 869882.Google Scholar
Liu, C. W. and Narasimhan, T. N., 1989b, Redox-controlled multiple-species reactive chemical transport, 2. Verification and application. Water Resources Research 25, 883910.Google Scholar
Lützenkirchen, J. (ed.), 2006, Surface Complexation Modelling, Academic Press, San Diego, California.Google Scholar
Madigan, M. and Martinko, J., 2017, Brock Biology of Microorganisms, 15th ed. Pearson, New York.Google Scholar
Maher, K. and Mayer, K. U. (eds.), 2019, Reactive transport modeling. Elements 15, Mineralogical Society of America, Washington DC.Google Scholar
Malmberg, C. G. and Maryott, A. A., 1956, Dielectric constant of water from 0° to 100°C. Journal of Research of the National Bureau of Standards 56, 18.Google Scholar
Malmstrom, M. E., Destouni, G. and Martinet, P., 2004, Modeling expected solute concentration in randomly heterogeneous flow systems with multicomponent reactions. Environmental Science and Technology 38, 26732679.Google Scholar
March, R., Doster, F. and Geiger, S., 2018, Assessment of CO2 storage potential in naturally fractured reservoirs with dual-porosity models. Water Resources Research 54, 16501668.Google Scholar
Marshall, W. L. and Slusher, R., 1966, Thermodynamics of calcium sulfate dihydrate in aqueous sodium chloride solutions, 0–110 °. Journal of Physical Chemistry 70, 40154027.Google Scholar
Martin, C. A., 1965, Denver basin. American Association of Petroleum Geologists Bulletin 49, 19081925.Google Scholar
Mathur, S. S. and Dzombak, D. A., 2006, Surface complexation modeling: goethite. In Lützenkirchen, J. (ed.), Surface Complexation Modelling, Academic Press, San Diego, California, pp. 443468.Google Scholar
Mattes, B. W. and Mountjoy, E. W., 1980, Burial dolomitization of the Upper Devonian Miette Buildup, Jasper National Park, Alberta. In Zenger, D. H., Dunham, J. B. and Effington, R. L. (eds.), Concepts and Models of Dolomitization. SEPM Special Publication 28, 259297.Google Scholar
May, H., 1992, The hydrolysis of aluminum, conflicting models and the interpretation of aluminum geochemistry. In Kharaka, Y. K. and Maest, A. S. (eds.), Water–Rock Interaction. Balkema, Rotterdam, pp. 1321.Google Scholar
Mayer, K. U., Frind, E. O. and Blowes, D. W., 2002, Multicomponent reactive transport modeling in variably saturated porous media using a generalized formulation for kinetically controlled reactions. Water Resources Research 38, 13-1–13-21.Google Scholar
Mayo, A. L., Nielsen, P. J., Loucks, M. and Brimhall, W. H., 1992, The use of solute and isotopic chemistry to identify flow patterns and factors which limit acid mine drainage in the Wasatch Range, Utah. Ground Water 30, 243249.Google Scholar
McCleskey, R. B., Nordstrom, D. K., Ryan, J. N. and Ball, J. W., 2012, A new method of calculating electrical conductivity with applications to natural waters. Geochimica et Cosmochimica Acta 77, 369382.Google Scholar
McCollom, T. M. and Shock, E. L., 1997, Geochemical constraints on chemolithoautotrophic metabolism by microorganisms in seafloor hydrothermal systems. Geochimica et Cosmochimica Acta 61, 43754391.Google Scholar
McConaghy, J. A., Chase, G. H., Boettcher, A. J. and Major, T. J., 1964, Hydrogeologic data of the Denver basin, Colorado. Colorado Groundwater Basic Data Report 15.Google Scholar
McCoy, A. W., III, 1953, Tectonic history of Denver basin. American Association of Petroleum Geologists Bulletin 37, 18731893.Google Scholar
McDuff, R. E. and Morel, F. M. M., 1980, The geochemical control of seawater (Sillen revisited). Environmental Science and Technology 14, 11821186.Google Scholar
Meeussen, J. C. L., 2003, ORCHESTRA: an object-oriented framework for implementing chemical equilibrium models. Environmental Science and Technology 37, 11751182.Google Scholar
Mehnert, E., Gendron, C. R. and Brower, R. D., 1990, Investigation of the hydraulic effects of deep-well injection of industrial wastes. Illinois State Geological Survey Environmental Geology 135, 100 p.Google Scholar
Merino, E., Nahon, D. and Wang, Y., 1993, Kinetics and mass transfer of pseudomorphic replacement, application to replacement of parent minerals and kaolinite by Al, Fe, and Mn oxides during weathering. American Journal of Science 293, 135155.Google Scholar
Meyers, W. J. and Lohmann, K. C., 1985, Isotope geochemistry of regional extensive calcite cement zones and marine components in Mississippian limestones, New Mexico. In Schneidermann, N. and Harris, P. M. (eds.), Carbonate Cements. SEPM Special Publication 36, 223239.Google Scholar
Michard, G., Fouillac, C., Grimaud, D. and Denis, J., 1981, Une méthode globale d’estimation des températures des réservoirs alimentant les sources thermales, exemple du Massif Centrale Français. Geochimica et Cosmochimica Acta 45, 11991207.Google Scholar
Michard, G. and Roekens, E., 1983, Modelling of the chemical composition of alkaline hot waters. Geothermics 12, 161169.Google Scholar
Miron, G. D., Kulik, D. A., Dmytrieva, S. V. and Wagner, T., 2015, GEMSFITS: code package for optimization of geochemical model parameters and inverse modeling. Applied Geochemistry 55, 2845.Google Scholar
Møller, N., 1988, The prediction of mineral solubilities in natural waters: a chemical equilibrium model for the Na–Ca–Cl–SO4–H2O system, to high temperature and concentration. Geochimica et Cosmochimica Acta 52, 821837.Google Scholar
Morel, F. M. M., 1983, Principles of Aquatic Chemistry. Wiley, New York.Google Scholar
Morel, F. and Morgan, J., 1972, A numerical method for computing equilibria in aqueous chemical systems. Environmental Science and Technology 6, 5867.Google Scholar
Morgan, J. J., 1967, Chemical equilibria and kinetic properties of manganese in natural waters. In Faust, S. D. and Hunter, J. V. (eds.), Principles and Applications of Water Chemistry. Wiley, New York.Google Scholar
Morse, J. W. and Casey, W. H., 1988, Ostwald processes and mineral paragenesis in sediments. American Journal of Science 288, 537560.Google Scholar
Moses, C. O., Nordstrom, D. K., Herman, J. S. and Mills, A. L., 1987, Aqueous pyrite oxidation by dissolved oxygen and by ferric iron. Geochimica et Cosmochimica Acta 51, 15611571.Google Scholar
Mottl, M. J. and McConachy, T. F., 1990, Chemical processes in buoyant hydrothermal plumes on the East Pacific Rise near 21 ° N. Geochimica et Cosmochimica Acta 54, 19111927.Google Scholar
Müller, B., 2015, CHEMEQL V3.2, A program to calculate chemical speciation equilibria, titrations, dissolution, precipitation, adsorption, kinetics, pX-pY diagrams, solubility diagrams. Limnological Research Center EAWAG, Kastanienbaum, Switzerland.Google Scholar
Nagy, K. L., 1995, Dissolution and precipitation kinetics of sheet silicates. Reviews in Mineralogy 31, 173– 233.Google Scholar
Nagy, K. L., Blum, A. E. and Lasaga, A. C., 1991, Dissolution and precipitation kinetics of kaolinite at 80 °C and pH 3: the dependence on solution saturation state. American Journal of Science 291, 649686.Google Scholar
Nagy, K. L. and Lasaga, A. C., 1992, Dissolution and precipitation kinetics of gibbsite at 80 °C and pH 3, the dependence on solution saturation state. Geochimica et Cosmochimica Acta 56, 30933111.Google Scholar
Neuman, S. P., 1990, Universal scaling of hydraulic conductivities and dispersivities in geologic media. Water Resources Research 26, 17491758Google Scholar
Newman, J. and Thomas-Alyea, K. E., 2004, Electrochemical Systems, 3rd ed. Wiley, Hoboken, New Jersey.Google Scholar
Nordeng, S. H. and Sibley, D. F., 1994, Dolomite stoichiometry and Ostwald’s step rule. Geochimica et Cosmochimica Acta 58, 191196.Google Scholar
Nordstrom, D. K., 1982, Aqueous pyrite oxidation and the consequent formation of secondary iron minerals. In Acid Sulfate Weathering. Soil Science Society of America Special Publication 10, 3756.Google Scholar
Nordstrom, D. K., Jenne, E. A. and Ball, J. W., 1979, Redox equilibria of iron in acid mine waters. In Jenne, E. A. (ed.), Chemical Modeling in Aqueous Systems, American Chemical Society, Washington DC, pp. 5179.Google Scholar
Nordstrom, D. K., McNutt, R. H., Puigdoménech, I., Smellie, J. A. T. and Wolf, M., 1992, Ground water chemistry and geochemical modeling of water–rock interactions at the Osamu Utsumi mine and the Morro do Ferro analogue study sites, Poços de Caldas, Minas Gerais, Brazil. Journal of Geochemical Exploration 45, 249287.Google Scholar
Nordstrom, D. K. and Munoz, J. L., 1994, Geochemical Thermodynamics, 2nd ed. Blackwell, Boston.Google Scholar
O’Connell, J., and Haile, J., 2005, Thermodynamics: Fundamentals for Applications. Cambridge University Press.Google Scholar
Okereke, A. and Stevens, S. E., Jr., 1991, Kinetics of iron oxidation by Thiobacillus ferrooxidans. Applied and Environmental Microbiology 57, 10521056.Google Scholar
O’Neil, J. R., 1987, Preservation of H, C, and O isotopic ratios in the low temperature environment. In Kyser, T. K. (ed.), Stable Isotope Geochemistry of Low Temperature Processes. Mineralogical Society of Canada Short Course 13, 85128.Google Scholar
Oreskes, N., Shrader-Frechette, K. and Belitz, K., 1994, Verification, validation, and confirmation of numerical models in the Earth sciences. Science 263, 641646.Google Scholar
Ortoleva, P. J., Merino, E., Moore, C. and Chadam, J., 1987, Geochemical self-organization, I: Reaction-transport feedbacks and modeling approach. American Journal of Science 287, 9791007.Google Scholar
Pace, M. L. and Prairie, Y. T., 2005, Respiration in lakes. In P. A. del Giorgio and J. le B. Williams (eds.), Respiration in Aquatic Ecosystems, Oxford University Press, pp. 103–121.Google Scholar
Pačes, T., 1975, A systematic deviation from Na–K–Ca geothermometer below 75 °C and above 10 4 atm PCO2 . Geochimica et Cosmochimica Acta 39, 541544.Google Scholar
Pačes, T., 1983, Rate constants of dissolution derived from the measurements of mass balance in hydrological catchments. Geochimica et Cosmochimica Acta 47, 18551863.Google Scholar
Pallud, C. and Van Cappellen, P., 2006, Kinetics of microbial sulfate reduction in estuarine sediments. Geochimica et Cosmochimica Acta 70, 11481162.Google Scholar
Panikov, N. S., 1995, Microbial Growth Kinetics. Chapman and Hall, London.Google Scholar
Park, J., Sanford, R. A. and Bethke, C. M., 2006, Geochemical and microbiological zonation of the Middendorf aquifer, South Carolina. Chemical Geology 230, 88104.Google Scholar
Park, J., Sanford, R. A. and Bethke, C. M., 2009, Microbial activity and chemical weathering in the Middendorf Aquifer, South Carolina. Chemical Geology 258, 232241.Google Scholar
Parkhurst, D. L., 1995, User’s guide to PHREEQC, a computer model for speciation, reaction-path, advective-transport and inverse geochemical calculations. US Geological Survey Water-Resources Investigations Report 95–4227.Google Scholar
Parkhurst, D. L. and Apello, C. A. J., 2013, Description of input and examples for PHREEQC version 3–A computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations. US Geological Surve Techniques and Methods 6–A43.Google Scholar
Parkhurst, D. L., Kipp, K. L. and Charlton, S. R., 2010, PHAST version 2. A program for simulating groundwater flow, solute transport, and multicomponent geochemical reactions. US Geological Survey Techniques and Methods 6–A35.Google Scholar
Parkhurst, D. L., Thorstenson, D. C. and Plummer, L. N., 1980, PHREEQE—a computer program for geochemical calculations. US Geological Survey Water-Resources Investigations Report 80–96.Google Scholar
Pawlowicz, R., 2008, Calculating the conductivity of natural waters. Limnology and Oceanography: Methods 6, 489501.Google Scholar
Peaceman, D. W., 1977, Fundamentals of Numerical Reservoir Simulation. Elsevier, Amsterdam.Google Scholar
Peng, D.-Y. and Robinson, D. B., 1976, A new two-constant equation of state. Industrial and Engineering Chemistry Fundamentals 15, 5964.Google Scholar
Perkins, E. H. and Brown, T. H., 1982, Program PATH, calculation of isothermal and isobaric mass transfer. University of British Columbia, unpublished manuscript.Google Scholar
Perthuisot, J. P., 1980, Sebkha el Melah near Zarzis: a recent paralic salt basin (Tunisia). In G. Busson (ed.), Evaporite Deposits, Illustration and Interpretation of Some Environmental Sequences, Editions Technip, Paris, pp. 11–17, 9295.Google Scholar
Phillips, O. M., 1991, Flow and Reaction in Permeable Rocks. Cambridge University Press.Google Scholar
Phillips, S. L., Igbene, A., Fair, J. A. and Ozbek, H., 1981, A technical databook for geothermal energy utilization. Lawrence Berkeley Laboratory Report LBL-12810.Google Scholar
Pitzer, K. S., 1975, Thermodynamics of electrolytes, V. Effects of higher order electrostatic terms. Journal of Solution Chemistry 4, 249265.Google Scholar
Pitzer, K. S., 1979, Theory: ion interaction approach. In Pytkowitz, R. M. (ed.), Activity Coefficients in Electrolyte Solutions, vol. 1. CRC Press, Boca Raton, Florida, pp. 157208.Google Scholar
Pitzer, K. S., 1987, A thermodynamic model for aqueous solutions of liquid-like density. Reviews in Mineralogy 17, 97142.Google Scholar
Pitzer, K. S., 1991, Ion interaction approach: theory and data correlation. In Pitzer, K. S. (ed.), Activity Coefficients in Electrolyte Solutions, 2nd ed., CRC Press, Boca Raton, Florida, pp. 75154.Google Scholar
Pitzer, K. S. and Brewer, L., 1961, Revised edition of Thermodynamics, by G. N. Lewis and M. Randall, 2nd ed., McGraw-Hill, New York.Google Scholar
Plankey, B. J., Patterson, H. H. and Cronan, C. S., 1986, Kinetics of aluminum fluoride complexation in acidic waters. Environmental Science and Technology 20, 160165.Google Scholar
Plumlee, G., 1994a, USGS assesses the impact of Summitville. USGS Office of Mineral Resources Newsletter 5 (2), 12.Google Scholar
Plumlee, G., 1994b, Environmental geology models of mineral deposits. Society of Economic Geologists Newsletter 16, 56.Google Scholar
Plumlee, G. S., Goldhaber, M. B. and Rowan, E. L., 1995, The potential role of magmatic gases in the genesis of Illinois-Kentucky fluorspar deposits, implications from chemical reaction path modeling. Economic Geology 90, 9991011.Google Scholar
Plumlee, G. S., Smith, K. S., Ficklin, W. H. and Briggs, P. H., 1992, Geological and geochemical controls on the composition of mine drainages and natural drainages in mineralized areas. In Kharaka, Y. K. and Maest, A. S. (eds.), Water–Rock Interaction. Balkema, Rotterdam, pp. 419422.Google Scholar
Plummer, L. N., 1992, Geochemical modeling of water–rock interaction: past, present, future. In Kharaka, Y. K. and Maest, A. S. (eds.), Water–Rock Interaction. Balkema, Rotterdam, pp. 2333.Google Scholar
Plummer, L. N., Parkhurst, D. L., Fleming, G. W. and Dunkle, S. A., 1988, PHRQPITZ, a computer program incorporating Pitzer’s equations for calculation of geochemical reactions in brines. US Geological Survey Water-Resources Investigations Report 88–4153.Google Scholar
Plyasunov, A. B. and Popova, E. S., 2013, Temperature dependence of the parameter of the SIT model for activity coefficients of 1:1 electrolytes. Journal of Solution Chemistry 42, 13201335.Google Scholar
Poling, B. E., Prausnitz, J. M. and O’Connell, J. P., 2001, The Properties of Gases and Liquids, 5th ed. McGraw-Hill, New York.Google Scholar
Prigogine, I. and Defay, R., 1954, Chemical Thermodynamics, D. H. Everett (trans.). Longmans, London.Google Scholar
Puigdomènech, I., Colàs, E., Grivé, M., Campos, I. and García, D., 2014, A tool to draw chemical equilibrium diagrams using SIT: applications to geochemical systems and radionuclide solubility. In Duro, L., Giménez, J., Casas, I. and de Pablo, J. (eds.), Scientific Basis for Nuclear Waste Management XXXVII, Materials Research Society, Warrendale, Pennsylvania, 111116.Google Scholar
Rabus, R., T. A. Hansen and F. Widdel, 2006, Dissimilatory sulfate- and sulfur-reducing prokaryotes. In The Prokaryotes: An Evolving Electronic Resource for the Microbiological Community, http://www.springerlink.com/content/n1084686101028pj/fulltext.pdf, Springer, New York.Google Scholar
Reardon, E. J., 1981, Kd’s—can they be used to describe reversible ion sorption reactions in contaminant migration? Ground Water 19, 279286.Google Scholar
Reed, M. and Spycher, N., 1984, Calculation of pH and mineral equilibria in hydrothermal waters with application to geothermometry and studies of boiling and dilution. Geochimica et Cosmochimica Acta 48, 14791492.Google Scholar
Reed, M. H., 1977, Calculations of hydrothermal metasomatism and ore deposition in submarine volcanic rocks with special reference to the West Shasta district, California. Ph.D. dissertation, University of California, Berkeley.Google Scholar
Reed, M. H., 1982, Calculation of multicomponent chemical equilibria and reaction processes in systems involving minerals, gases and an aqueous phase. Geochimica et Cosmochimica Acta 46, 513528.Google Scholar
Reed, M. H., Spycher, N. F. and Palandri, J., 2016, Users guide for CHIM-XPT: a program for computing reaction processes in aqueous-mineral-gas systems and MINTAB guide. Version 2.50. pages.uoregon.edu/palandri/data/chim-xpt%20guide%20V.2.50.pdf.Google Scholar
Rice, E. W., Baird, R. B. and Eaton, A. D. (eds.), 2017, Standard Methods for the Examination of Water and Wastewater, 23rd ed. American Public Health Association, Washington DC.Google Scholar
Richtmyer, R. D., 1957, Difference Methods for Initial-Value Problems. Wiley-Interscience, New York.Google Scholar
Rimstidt, J. D., 2014, Geochemical Rate Models, Cambridge University Press.Google Scholar
Rimstidt, J. D. and Barnes, H. L., 1980, The kinetics of silica-water reactions. Geochimica et Cosmochimica Acta 44, 16831700.Google Scholar
Robie, R. A., Hemingway, B. S. and Fisher, J. R., 1979, Thermodynamic properties of minerals and related substances at 298.15 K and 1 bar (105 pascals) pressure and at higher temperatures. US Geological Survey Bulletin 1452 (corrected edition).Google Scholar
Robinson, R. A. and Stokes, R. H., 1968, Electrolyte Solutions. Butterworths, London.Google Scholar
Roden, E. E. and Wetzel, R. G., 2003, Competition between Fe(III)-reducing and methanogenic bacteria for acetate in iron-rich freshwater sediments. Microbial Ecology 45, 252258.Google Scholar
Rosing, M. T., 1993, The buffering capacity of open heterogeneous systems. Geochimica et Cosmochimica Acta 57, 22232226.Google Scholar
Rowan, E., 1991, Un modéle géochimique, thérmique-hydrogéologique et tectonique pour la genése des gisements filoniens de fluorite de l’Albigeois, sud-ouest du Massif Central, France. Mémoire de DEA, Université Pierre et Marie Curie (Paris IV), Paris.Google Scholar
Rowland, D., Königsberger, E., Hefner, G. and P. M. May, 2015, Aqueous electrolyte solution modelling: some limitations of the Pitzer equations. Applied Geochemistry 55, 170183.Google Scholar
Rubin, J., 1983, Transport of reacting solutes in porous media, relationship between mathematical nature of problem formulation and chemical nature of reactions. Water Resources Research 19, 12311252.Google Scholar
Runnells, D. D., 1969, Diagenesis, chemical sediments, and the mixing of natural waters. Journal of Sedimentary Petrology 39, 11881201.Google Scholar
Runnells, D. D. and Lindberg, R. D., 1990, Selenium in aqueous solutions: the impossibility of obtaining a meaningful Eh using a platinum electrode, with implications for modeling of natural waters. Geology 18, 212215.Google Scholar
Sahai, N. and Sverjensky, D. A., 1997, Evaluation of internally consistent parameters for the triple-layer model by systematic analysis of oxide surface titration data. Geochimica et Cosmochimica Acta 61, 28012826.Google Scholar
Santore, R. C., Ryan, A. C., Kroglund, F. et al., 2018, Development and application of a biotic ligand model for predicting the chronic toxicity of dissolved and precipitated aluminum to aquatic organisms. Environmental Toxicology and Chemistry 37, 70–79.Google Scholar
Schecher, W. D. and McAvoy, D. C., 1994, MINEQL+, A Chemical Equilibrium Program for Personal Computers, User’s Manual, version 3.0. Environmental Research Software, Inc., Hallowell, Maine.Google Scholar
Schrenk, M. O., Kelley, D. S., Delaney, J. R. and Baross, J. A., 2003, Incidence and diversity of microorganisms within the walls of an active deep-sea sulfide chimney. Applied and Environmental Microbiology 69, 35803592.Google Scholar
Shaff, J. E., Schultz, B. A., Craft, E. J., Clark, R. T. and Kochian, L. V., 2010, GEOCHEM-EZ: a chemical speciation program with greater power and flexibility. Plant and Soil 330, 207214.Google Scholar
Shanks, W. C., III, and Bischoff, J. L., 1977, Ore transport and deposition in the Red Sea geothermal system, a geochemical model. Geochimica et Cosmochimica Acta 41, 15071519.Google Scholar
Sherlock, E. J., Lawrence, R. W. and Poulin, R., 1995, On the neutralization of acid rock drainage by carbonate and silicate minerals. Environmental Geology 25, 4354.Google Scholar
Shock, E. L., 1988, Organic acid metastability in sedimentary basins. Geology 16, 886890 (correction, 17, 572–573).Google Scholar
Sibley, D. F. and Blatt, H., 1976, Intergranular pressure solution and cementation of the Tuscarora orthoquartzite. Journal of Sedimentary Petrology 46, 881896.Google Scholar
Sillén, L. G., 1967, The ocean as a chemical system. Science 156, 11891197.Google Scholar
Simoes, M. C., Hughes, K. J., Ingham, D. B., Ma, L. and Pourkashanian, M., 2017, Temperature dependence of the parameters in the Pitzer equations. Journal of Chemical and Engineering Data 62, 20002013.Google Scholar
Sin, I., Lagneau, V. and Corvisier, J., 2017, Integrating a compressible multicomponent two-phase flow into an existing reactive transport simulator. Advances in Water Resources 100, 6277.Google Scholar
Skirrow, G., 1965, The dissolved gases—carbon dioxide. In Riley, J. P. and Skirrow, G. (eds.), Chemical Oceanography, Academic Press, London, pp. 227322.Google Scholar
Smith, G. D., 1986, Numerical Solution of Partial Differential Equations: Finite Difference Methods, 3rd ed. Oxford University Press.Google Scholar
Smith, K. S., Ficklin, W. H., Plumlee, G. S. and Meier, A. L., 1992, Metal and arsenic partitioning between water and suspended sediment at mine-drainage sites in diverse geologic settings. In Kharaka, Y. K. and Maest, A. S. (eds.), Water–Rock Interaction. Balkema, Rotterdam, pp. 443447.Google Scholar
Smith, W. R. and Missen, R. W., 1982, Chemical Reaction Equilibrium Analysis: Theory and Algorithms. Wiley, New York.Google Scholar
Snoeyink, V. L. and Jenkins, D., 1980, Water Chemistry. Wiley, New York.Google Scholar
Sorbie, K. S., Yuan, M. and Jordan, M. M., 1994, Application of a scale inhibitor squeeze model to improve field squeeze treatment design (SPE paper 28885). European Petroleum Conference Proceedings Volume (vol. 2 of 2), Society of Petroleum Engineers, Richardson, Texas, pp. 179–191.Google Scholar
Søreide, I. and Whitson, C. H., 1992, Peng–Robinson predictions for hydrocarbons, CO2,N2, and H2S with pure water and NaCl brine. Fluid Phase Equilibria 77, 217240.Google Scholar
Sørensen, S. P. L., 1909, Enzymstudier, II., Om Maalingen og Betydningen af Brintionkoncentration ved enzymatiske Processer. Meddelelser fra Carlsberg Laboratoriet 8, 1153.Google Scholar
Sposito, G., 2016, The Chemistry of Soils, 3rd ed. Oxford University Press.Google Scholar
Spycher, N. F. and Reed, M. H., 1988, Fugacity coefficients of H2,CO2,CH4,H2O and of H2O–CO2–CH4 mixtures, a virial equation treatment for moderate pressures and temperatures applicable to calculations of hydrothermal boiling. Geochimica et Cosmochimica Acta 52, 739749.Google Scholar
Steefel, C. I., Appelo, C. A. J., Arora, B. et al., 2015, Reactive transport codes for subsurface environmental simulation. Computational Geosciences 19, 445478.Google Scholar
Steefel, C. I., DePaolo, D. J. and Lichtner, P. C., 2005, Reactive transport modeling: an essential tool and a new research approach for the Earth sciences. Earth and Planetary Science Letters 240, 539558.Google Scholar
Steefel, C. I. and Lasaga, A. C., 1992, Putting transport into water–rock interaction models. Geology 20, 680– 684.Google Scholar
Steefel, C. I. and Lasaga, A. C., 1994, A coupled model for transport of multiple chemical species and kinetic precipitation/dissolution reactions with application to reactive flow in single phase hydrothermal systems. American Journal of Science 294, 529592.Google Scholar
Steefel, C. I. and MacQuarrie, K. T. B., 1996, Approaches to modeling of reactive transport in porous media. Reviews in Mineralogy 34, 85129.Google Scholar
Steefel, C. I. and Van Cappellen, P., 1990, A new kinetic approach to modeling water–rock interaction, the role of nucleation, precursors, and Ostwald ripening. Geochimica et Cosmochimica Acta 54, 26572677.Google Scholar
Steefel, C. I. and Yabusaki, S. B., 1996, OS3D/GIMRT, Software for multicomponent-multidimensional reactive transport: user’s manual and programmer’s guide. Report PNL-11166, Pacific Northwest National Laboratory, Richland, Washington.Google Scholar
Stumm, W., 1992, Chemistry of the Solid-Water Interface. Wiley, New York.Google Scholar
Stumm, W. and Morgan, J. J., 1996, Aquatic Chemistry, Chemical Equilibria and Rates in Natural Waters, 3rd ed. Wiley, New York.Google Scholar
Stumm, W. and Wollast, R., 1990, Coordination chemistry of weathering, kinetics of the surface-controlled dissolution of oxide minerals. Reviews of Geophysics 28, 5369.Google Scholar
Sudicky, E. A. and Frind, E. O., 1982, Contaminant transport in fractured porous media: analytical solutions for a system of parallel fractures. Water Resources Research 18, 16341642.Google Scholar
Sung, W. and Morgan, J. J., 1981, Oxidative removal of Mn(II) from solution catalysed by the ɣ-FeOOH (lepidocrocite) surface. Geochimica et Cosmochimica Acta 45, 23772383.Google Scholar
Surdam, R. C. and Boles, J. R., 1979, Diagenesis of volcanic sandstones. Society of Economic Paleontologists and Mineralogists Special Publication 26, 227242.Google Scholar
Sverjensky, D. A., 1984, Oil field brines as ore-forming solutions. Economic Geology 79, 2337.Google Scholar
Sverjensky, D. A., 1987, The role of migrating oil field brines in the formation of sediment-hosted Cu-rich deposits. Economic Geology 82, 11301141.Google Scholar
Sverjensky, D. A., 1993, Physical surface-complexation models for sorption at the mineral–water interface. Nature 364, 776780.Google Scholar
Sverjensky, D. A., 2003, Standard states for the activities of mineral surface sites and species. Geochimica et Cosmochimica Acta 67, 1728.Google Scholar
Sverjensky, D. A., 2006, Prediction of the speciation of alkaline earths adsorbed on mineral surfaces in salt solutions. Geochimica et Cosmochimica Acta 70, 24272453.Google Scholar
Sydansk, R. D., 1982, Elevated-temperature caustic/sandstone interaction, implications for improving oil recovery. Society of Petroleum Engineers Journal 22, 453462.Google Scholar
Tan, K. L. and Hameed, B. H., 2017, Insight into the adsorption kinetics models for the removal of contaminants from aqueous solutions. Journal of the Taiwan Institute of Chemical Engineers 74, 2548.Google Scholar
Tang, D. H., Sudicky, E. A. and Frind, E. O., 1981, Contaminant transport in fractured porous media: analytical solution for a single fracture. Water Resources Research 17, 555564.Google Scholar
Taylor, B. E., Wheeler, M. C. and Nordstrom, D. K., 1984, Isotope composition of sulphate in acid mine drainage as measure of bacterial oxidation. Nature 308, 538541.Google Scholar
Thauer, R. K., Jungerman, K. and Decker, K., 1977, Energy conservation in chemotrophic anaerobic bacteria. Bacteriological Reviews 41, 100180.Google Scholar
Thompson, J. B., Jr., 1959, Local equilibrium in metasomatic processes. In Abelson, P. H. (ed.), Researches in Geochemistry, Wiley, New York, pp. 427457.Google Scholar
Thompson, J. B., Jr., 1970, Geochemical reaction and open systems. Geochimica et Cosmochimica Acta 34, 529551.Google Scholar
Thompson, J. B., Jr., 1982, Composition space: an algebraic and geometric approach. Reviews in Mineralogy 10, 131.Google Scholar
Thompson, M. E., 1992, The history of the development of the chemical model for seawater. Geochimica et Cosmochimica Acta 56, 29852987.Google Scholar
Thorstenson, D. C., 1984, The concept of electron activity and its relation to redox potentials in aqueous geochemical systems. US Geological Survey Open File Report 84–072, 45 p.Google Scholar
Thorstenson, D. C., Fisher, D. W. and Croft, M. G., 1979, The geochemistry of the Fox Hills-Basal Hell Creek aquifer in southwestern North Dakota and northwestern South Dakota. Water Resources Research 15, 14791498.Google Scholar
Todd, A. C. and Yuan, M., 1990, Barium and strontium sulphate solid-solution formation in relation to North Sea scaling problems. SPE Production Engineering 5, 279285.Google Scholar
Todd, A. C. and Yuan, M., 1992, Barium and strontium sulphate solid-solution scale formation at elevated temperatures. SPE Production Engineering 7, 8592.Google Scholar
Truesdell, A. H. and Jones, B. F., 1974, WATEQ, a computer program for calculating chemical equilibria of natural waters. US Geological Survey Journal of Research 2, 233248.Google Scholar
Tsonopoulos, C., 1974, An empirical correlation of second virial coefficients. AIChE Journal 20, 263272.Google Scholar
Tsonopoulos, C. and Heidman, J. L., 1990, From the virial to the cubic equation of state. Fluid Phase Equilibria 57, 261276.Google Scholar
Turner, D. R. and Sassman, S. A., 1996, Approaches to sorption modeling for high-level waste performance assessment. Journal of Contaminant Hydrology 21, 311332.Google Scholar
Valocchi, A. J., 1985, Validity of the local equilibrium assumption for modeling sorbing solute transport through homogeneous soils. Water Resources Research 21, 808820.Google Scholar
Valocchi, , A. J, Street, R. L. and Roberts, P. V., 1981, Transport of ion-exchanging solutes in groundwater, chromatographic theory and field simulation. Water Resources Research 17, 15171527.Google Scholar
van der Lee, J., De Windt, L., Lagneau, V. and Goblet, P., 2003, Module oriented modeling of reactive transport with HYTEC. Computers and Geosciences 29, 265275.Google Scholar
Van Dover, C. L., 2000, The Ecology of Deep-Sea Hydrothermal Vents. Princeton University Press, Princeton, New Jersey.Google Scholar
Van Zeggeren, F. and Storey, S. H., 1970, The Computation of Chemical Equilibria. Cambridge University Press.Google Scholar
Verweij, W., 2017, Manual for CHEAQS NEXT, a program for calculating CHemical Equilibria in AQuatic Systems. http://www.cheaqs.eu/manual.pdf.Google Scholar
Verweij, W. and Simonin, J.-P., 2020, Implementing the mean spherical approximation model in the speciation code CHEAQS NEXT at high salt concentrations. Journal of Solution Chemistry 49, 13191327.Google Scholar
Viani, B. E. and Bruton, C. J., 1992, Modeling ion exchange in clinoptilolite using the EQ3/6 geochemical modeling code. In Y. K Kharaka and A. Maest, S (eds.), Water–Rock Interaction. Balkema, Rotterdam, pp. 7377.Google Scholar
Von Damm, K. L., Edmond, J. M., Grant, B. and Measures, C. I., 1985, Chemistry of submarine hydrothermal solutions at 21 °N, East Pacific Rise. Geochimica et Cosmochimica Acta 49, 21972220.Google Scholar
Wagner, T., Kulik, D. A., Hingerl, F. F. and Dmytrieva, S. V., 2012, GEM-SELEKTOR geochemical modeling package: TSOLMOD library and data interface for multicomponent phase models. Canadian Mineralogist 50, 11731195.Google Scholar
Wang, Z. and Giammar, D. E., 2013, Mass action expressions for bidentate adsorption in surface complexation modeling: theory and practice. Environmental Science and Technology 47, 39823996.Google Scholar
Warga, J., 1963, A convergent procedure for solving the thermo-chemical equilibrium problem. Journal Society Industrial and Applied Mathematicians 11, 594606.Google Scholar
Wat, R. M. S., Sorbie, K. S., Todd, A. C., Chen, P. and Jiang, P., 1992, Kinetics of BaSO4 crystal growth and effect in formation damage (SPE paper 23814). Proceedings of the Society of Petroleum Engineers International Symposium on Formation Damage Control, Lafayette, Louisiana, February 26–27, 1992, pp. 429437.Google Scholar
Weare, J. H., 1987, Models of mineral solubility in concentrated brines with application to field observations. Reviews in Mineralogy 17, 143176.Google Scholar
Webb, J. R., Santos, I. R., Maher, D. T. and Finlay, K., 2019, The importance of aquatic carbon fluxes in net ecosystem carbon budgets: a catchment-scale review. Ecosystems 22, 508527.Google Scholar
Wedepohl, K. H., Correns, C. W., Shaw, D. M., Turekian, K. K. and Zemann, J. (eds.), 1978, Handbook of Geochemistry, volumes II/1 and II/2, Springer-Verlag, Berlin.Google Scholar
Wehrli, B. and Stumm, W., 1988, Oxygenation of vanadyl(IV). Effect of coordinated surface hydroxyl groups and OH. Langmuir 4, 753758.Google Scholar
Wehrli, B. and Stumm, W., 1989, Vanadyl in natural waters: adsorption and hydrolysis promote oxygenation. Geochimica et Cosmochimica Acta 53, 6977.Google Scholar
Weng, L., Van Riemsdijk, W. H. and Hiemstra, T., 2008, Cu2C and Ca2C adsorption to goethite in the presence of fulvic acids. Geochimica et Cosmochimica Acta 72, 58575870.Google Scholar
Westall, J., 1980, Chemical equilibrium including adsorption on charged surfaces. In Kavanaugh, M. C. and Leckie, J. O. (eds.), Advances in Chemistry Series 189, American Chemical Society, Washington DC, pp. 3344.Google Scholar
Westall, J., 2002, Geochemical equilibrium and the interpretation of Eh. In Wilkin, R. T., R. D. Ludwig and R. G. Ford (eds.), Workshop on Monitoring Oxidation-Reduction Processes for Ground-water Restoration, Workshop Summary, Report EPA/600/R-02/002, US Environmental Protection Agency, Ada, Oklahoma, 21–23.Google Scholar
Westall, J. C. and Hohl, H., 1980, A comparison of electrostatic models for the oxide/solution interface. Advances in Colloid Interface Science 12, 265294.Google Scholar
Westall, J. C., Zachary, J. L. and Morel, F. F. M., 1976, MINEQL, a computer program for the calculation of chemical equilibrium composition of aqueous systems. Technical Note 18, R. M. Parsons Laboratory, Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts.Google Scholar
White, A. F., 1995, Chemical weathering rates of silicate minerals in soils. Reviews in Mineralogy 31, 407– 461.Google Scholar
White, D., 2007, The Physiology and Biochemistry of Prokaryotes, 3rd ed. Oxford University Press.Google Scholar
White, D. E., 1970, Geochemistry applied to the discovery, evaluation, and exploitation of geothermal energy resources. Geothermics 2, 5880.Google Scholar
White, W. B., 1967, Numerical determination of chemical equilibrium and the partitioning of free energy. Journal of Chemical Physics 46, 41714175.Google Scholar
White, W. B., Johnson, S. M. and Dantzig, G. B., 1958, Chemical equilibrium in complex mixtures. Journal of Chemical Physics 28, 751755.Google Scholar
Williamson, M. A. and Rimstidt, J. D., 1994, The kinetics and electrochemical rate-determining step of aqueous pyrite oxidation. Geochimica et Cosmochimica Acta 58, 54435454.Google Scholar
Wolery, T. J., 1978, Some chemical aspects of hydrothermal processes at mid-ocean ridges, a theoretical study, I., Basalt-sea water reaction and chemical cycling between the oceanic crust and the oceans, II. Calculation of chemical equilibrium between aqueous solutions and minerals. Ph.D. dissertation, Northwestern University, Evanston, Illinois.Google Scholar
Wolery, T. J., 1979, Calculation of chemical equilibrium between aqueous solution and minerals: the EQ3/6 software package. Lawrence Livermore National Laboratory Report UCRL-52658.Google Scholar
Wolery, T. J., 1983, EQ3NR, a computer program for geochemical aqueous speciation-solubility calculations: user’s guide and documentation. Lawrence Livermore National Laboratory Report UCRL-53414.Google Scholar
Wolery, T. J., 1992a, EQ3/EQ6, a software package for geochemical modeling of aqueous systems, package overview and installation guide (version 7.0). Lawrence Livermore National Laboratory Report UCRL-MA-110662 (1).Google Scholar
Wolery, T. J., 1992b, EQ3NR, a computer program for geochemical aqueous speciation-solubility calculations: theoretical manual, user’s guide, and related documentation (version 7.0). Lawrence Livermore National Laboratory Report UCRL-MA-110662 (3).Google Scholar
Wolery, T. J., 1995, Letter report: progress in developing EQ3/6 for modeling boiling processes. Lawrence Liverment National Laboratory Report UCRL-ID-130356.Google Scholar
Wolery, T. J., 2003, EQ3/6 version 8.0 software users manual. Document 10813-UM-8.0-00, U.S. Department of Energy, Office of Civilian Radioactive Waste Management, Las Vegas, Nevada.Google Scholar
Wolery, T. J. and Walters, L. J., Jr., 1975, Calculation of equilibrium distributions of chemical species in aqueous solutions by means of monotone sequences. Mathematical Geology 7, 99115.Google Scholar
Xu, T., Sonnenthal, E., Spycher, N. and Liange, L., 2014, TOUGHREACT V3.0 reference manual: a parallel simulation program for non-isothermal multiphase geochemical reactive transport. Report LBNL-DRAFT, Lawrence Berkeley National Laboratory, Berkeley, California.Google Scholar
Yabusaki, S. B., Steefel, C. I. and Wood, B. D., 1998, Multidimensional, multicomponent subsurface reactive transport in non-uniform velocity fields: code verification using an advective reactive streamtube approach. Journal of Contaminant Hydrology 30, 299331.Google Scholar
Yang, S. T. and Okos, M. R., 1987, Kinetic study and mathematical modeling of methanogenesis of acetate using pure cultures of methanogens. Biotechnology and Bioengineering 30, 661667.Google Scholar
Yeh, G. T. and Tripathi, V. S., 1989, A critical evaluation of recent developments in hydrogeochemical transport models of reactive multi-chemical components. Water Resources Research 25, 93108.Google Scholar
Yoneda, H., 1958, Stability of cobalt (III) and chromium (III) ammine complexes in a strongly alkaline solution. Bulletin of the Chemical Society of Japan 31, 209213.Google Scholar
Yuan, D. and Todd, A. C., 1991, Prediction of sulphate scaling tendency in oilfield operations. SPE Petroleum Engineering 6, 6372.Google Scholar
Yuan, M., Todd, A. C. and Sorbie, K. S., 1994, Sulphate scale precipitation arising from seawater injection, a prediction study. Marine and Petroleum Geology 11, 2430.Google Scholar
Zabala (de), E. F., Vislocky, J. M., Rubin, E. and Radke, C. J., 1982, A chemical theory for linear alkaline flooding. Society of Petroleum Engineers Journal 22, 245258.Google Scholar
Zehnder, A. J. B., Huser, B. A., Brock, T. D. and Wuhrmann, K., 1980, Characterization of an acetate decarboxylating, non-hydrogen-utilizing methane bacterium. Archives of Microbiology 124, 111.Google Scholar
Zeleznik, F. J. and Gordon, S., 1960, An analytical investigation of three general methods of calculating chemical equilibrium compositions. NASA Technical Note D-473, Washington DC.Google Scholar
Zeleznik, F. J. and Gordon, S., 1968, Calculation of complex chemical equilibria. Industrial and Engineering Chemistry 60, 2757.Google Scholar
Zhang, W. and Bouwer, E. J., 1997, Biodegradation of benzene, toluene and naphthalene in soil-water slurry microcosms. Biodegradation 8, 167175.Google Scholar
Zhang, Y., 2008, Geochemical Kinetics. Princeton University Press, Princeton, New Jersey.Google Scholar
Zhao, H., Liu, H. and Qu, J., 2009, Effect of pH on the aluminum salts hydrolysis during coagulation process: formation and decomposition of polymeric aluminum species. Journal of Colloid and Interface Science 330, 105112.Google Scholar
Zheng, C. and Bennett, G. D., 2002, Applied Contaminant Transport Modeling, 2nd ed. Wiley, New York.Google Scholar
Zhu, C. and Anderson, G., 2002, Environmental Applications of Geochemical Modeling. Cambridge University Press.Google Scholar
Zierenberg, R. A., Adams, M. W. W. and Arp, A. J., 2000, Life in extreme environments: hydrothermal vents. Proceedings National Academy of Sciences 97, 12 961–12 962.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.

  • References
  • Craig M. Bethke, University of Illinois, Urbana-Champaign
  • Book: Geochemical and Biogeochemical Reaction Modeling
  • Online publication: 09 December 2021
  • Chapter DOI: https://doi.org/10.1017/9781108807005.049
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.

  • References
  • Craig M. Bethke, University of Illinois, Urbana-Champaign
  • Book: Geochemical and Biogeochemical Reaction Modeling
  • Online publication: 09 December 2021
  • Chapter DOI: https://doi.org/10.1017/9781108807005.049
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.

  • References
  • Craig M. Bethke, University of Illinois, Urbana-Champaign
  • Book: Geochemical and Biogeochemical Reaction Modeling
  • Online publication: 09 December 2021
  • Chapter DOI: https://doi.org/10.1017/9781108807005.049
Available formats
×