Hostname: page-component-848d4c4894-p2v8j Total loading time: 0.001 Render date: 2024-06-04T12:40:00.125Z Has data issue: false hasContentIssue false

The Influence of Uranyl Hydrolysis and Multiple Site-Binding Reactions on Adsorption of U(VI) to Montmorillonite

Published online by Cambridge University Press:  28 February 2024

James P. McKinley
Affiliation:
Pacific Northwest Laboratory, MSIN K3-61, P.O. Box 999, Richland, Washington 99352
John M. Zachara
Affiliation:
Pacific Northwest Laboratory, MSIN K3-61, P.O. Box 999, Richland, Washington 99352
Steven C. Smith
Affiliation:
Pacific Northwest Laboratory, MSIN K3-61, P.O. Box 999, Richland, Washington 99352
Gary D. Turner
Affiliation:
Pacific Northwest Laboratory, MSIN K3-61, P.O. Box 999, Richland, Washington 99352

Abstract

Adsorption of uranyl to SWy-1 montmorillonite was evaluated experimentally and results were modeled to identify likely surface complexation reactions responsible for removal of uranyl from solution. Uranyl was contacted with SWy-1 montmorillonite in a NaCIO4 electrolyte solution at three ionic strengths (I = 0.001, 0.01, 0.1), at pH 4 to 8.5, in a N2(g) atmosphere. At low ionic strength, adsorption decreased from 95% at pH 4 to 75% at pH 6.8. At higher ionic strength, adsorption increased with pH from initial values less than 75%; adsorption edges for all ionic strengths coalesced above a pH of 7. A site-binding model was applied that treated SWy-1 as an aggregate of fixed-charge sites and edge sites analogous to gibbsite and silica. The concentration of fixed-charge sites was estimated as the cation exchange capacity, and non-preference exchange was assumed in calculating the contribution of fixed-charge sites to total uranyl adsorption. The concentration of edge sites was estimated by image analysis of transmission electron photomicrographs. Adsorption constants for uranyl binding to gibbsite and silica were determined by fitting to experimental data, and these adsorption constants were then used to simulate SWy-1 adsorption results. The best simulations were obtained with an ionization model in which AlOH2+ was the dominant aluminol surface species throughout the experimental range in pH. The pH-dependent aqueous speciation of uranyl was an important factor determining the magnitude of uranyl adsorption. At low ionic strength and low pH, adsorption by fixed-charge sites was predominant. The decrease in adsorption with increasing pH was caused by the formation of monovalent aqueous uranyl species, which were weakly bound to fixed-charge sites. At higher ionic strengths, competition with Na+ decreased the adsorption of UO22+ to fixed-charge sites. At higher pH, the most significant adsorption reactions were the binding of UO22+ to AlOH and of (UO2)3(OH)5+ to SiOH edge sites. Near-saturation of AlOH sites by UO22+ allowed significant contributions of SiOH sites to uranyl adsorption.

Type
Research Article
Copyright
Copyright © 1995, The Clay Minerals Society

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

Aberg, M., 1978. The crystal structure of hexaaqua-tri-μ-tri-μ3-oxo-triuranyl(VI) nitrate tetrahydrate,[(OU2)3O(OH)3(H2O)6]-NO34H2O. Acta Chemica Scandinavica 32: 101107.Google Scholar
Aberg, M., Ferri, D., Glaser, J., and Grenthe, I. 1983. Structure of the hydrated dioxyuranium (VI) ion in aqueous solution. An X-ray diffraction and 1H NMR study. Inorg. Chem. 22: 39863989.Google Scholar
Allard, B., Olofsson, U., Torstenfelt, B., and Kipatsi, H. 1983. Sorption behavior of actinides in well-defined oxidation states. SK BF KBS Technical Report 1983–05–15. Chalmers University of Technology, Goteborg, Sweden.Google Scholar
Babcock, K. L., and Schultz, R. K. 1970. Isotopic and conventional determination of exchangeable sodium percentage of soil in relation to plant growth. Soil Sci. 109: 1922.CrossRefGoogle Scholar
Bonotto, D. M., 1989. The behavior of dissolved uranium in groundwaters of the Morro do Ferro thorium deposit, Brazil. J. Hydrology 107: 155168.Google Scholar
Borovec, Z., 1981. The adsorption of uranyl species by fine clay. Chem. Geo. 32: 4558.Google Scholar
Chisholm-Brause, C., Conradson, S. D., Buscher, C. T., Eller, P. G., and Morris, D. E. 1994. Speciation of sites on mont-morillonite. Geochim. et Cosmichim. Acta 58: 36253631.Google Scholar
Davis, J. A., and Leckie, J. O. 1978. Surface ionization and complexation at the oxide/water interface II. Surface properties of amorphous iron oxyhydroxide and adsorption of metal ions. J. Colloid Interface Sci. 67: 90107.Google Scholar
Dent, A. J., Ramsay, J. D. F., and Swanton, S. W. 1992. An EXAFS study of uranyl ion in solution and sorbed onto silica and montmorillonite clay colloids. J. Colloid Interface Sci. 150: 4560.Google Scholar
Fiala, V., 1988. The significance of the clay minerals in the genesis of hydrothermal uranium deposits. Tenth Conference on Clay Mineralogy and Petrology, Ostrava (1986), 255260.Google Scholar
Fletcher, P., and Sposito, G. 1989. The chemical modeling of clay/electrolyte interaction for montmorillonite. Clay Miner. 24: 375391.Google Scholar
Giblin, A. M., 1980. The role of clay adsorption in genesis of uranium ores. Uranium in the Pine Creek Geosyncline. International Atomic Energy Agency, Vienna, Austria, 521529.Google Scholar
Grenthe, I. Chairman. 1992. Chemical Thermodynamics of Uranium. New York: North-Holland, 715 pp.Google Scholar
Hayes, K. F., and Leckie, J. O. 1987. Modeling ionic strength effects on cation adsorption at hydrous oxide/solution interfaces. J. Colloid Interface Sci. 115: 564572.Google Scholar
Hiemstra, T., van Riemsdijk, W. H., and Bruggenwert, M. G. M. 1987. Proton adsorption mechanism at the gibbsite and aluminum oxide solid/solution interface. Netherlands J. Agricultural Sci. 35: 281293.Google Scholar
Ho, C. H., and Miller, N. H. 1985. Effect of humic acid on uranium uptake by hematite particles. J. Colloid Interface Sci. 106: 281288.CrossRefGoogle Scholar
Hsi, C.-K. D., and Langmuir, D. 1985. Adsorption of uranyl onto ferric oxyhydroxides: Application of the surface complexation site-binding model. Geochim. Cosmochim. Acta 49: 19311941.Google Scholar
Lieser, K. H., Quandt-Klenk, S., and Thybusch, B. 1992. Sorption of uranyl ions on hydrous silicon dioxide. Radiochim. Acta 57: 4550.Google Scholar
Maya, L., 1982. Sorbed uranium (VI) species on hydrous titania, zirconia, and silica gel. Radiochim. Acta 31: 147151.Google Scholar
Morin, K. A., Cherry, J. A., Lim, T. P., and Vivyurka, A. J. 1982. Contaminant migration in a sand aquifer near an inactive uranium tailings impoundment, Elliot Lake, Ontario. Canadian Geotechnical J. 19: 4962.Google Scholar
Morin, K. A., and Cherry, J. A. 1988. Migration of acidic groundwater seepage from uranium-tailings impoundments, 3. Simulation of the conceptual model with application to seepage area A. J. Contaminant Hydrology 2: 323342.Google Scholar
Morris, D. E., Chisholm-Brause, C. J., Barr, M. E., Conradson, S. D., and Eller, P. G. 1994. Spectroscopic evidence for discrete multiple sorption sites for UO22+ species on a reference smectite. Geochim. et Cosmochim. Acta 58: 36133623.Google Scholar
Newman, A. C. D., and Brown, G. 1987. The chemical constitution of clays. In Chemistry of Clays and Clay Minerals, Newman, A. C. D., ed. Mineralogical Society Monograph No. 6. New York: John Wiley & Sons, 1128.Google Scholar
Newman, A. C. D., 1987. The interaction of water with clay mineral surfaces. In Chemistry of Clays and Clay Minerals, Newman, A. C. D., ed. Mineralogical Society Monograph No. 6. New York: John Wiley & Sons, 237274.Google Scholar
Payne, T. E., and Waite, T. D. 1991. Surface complexation modeling of uranium sorption data obtained by isotope exchange techniques. Radiochim. Acta 52: 487493.Google Scholar
Riese, A. C., 1982. Adsorption of radium and thorium on quartz and kaolinite: A comparison of solution/surface equilibrium models. Ph.D. dissertation, Colorado School of Mines.Google Scholar
Riley, R. G., Zachara, J. M., and Wobber, F. J. 1992. Chemical Contaminants on DOE Lands and Selection of Contaminant Mixtures for Subsurface Science Research. Report DOE/ER-0547T. U.S. Department of Energy, Office of Energy Research, Washington, D.C.Google Scholar
Schindler, P. W., Liechti, P., and Westall, J. C. 1987. Adsorption of copper, cadmium, and lead from aqueous solution to the kaolinite/water interface. Netherlands J. Agricultural Sci. 35: 219230.Google Scholar
Sposito, G., 1984. The Surface Chemistry of Soils. New York: Oxford University Press, 234 pp.Google Scholar
Sposito, G., 1989. The Chemistry of Soils. New York: Oxford University Press, 277 pp.Google Scholar
Sposito, G., Holtzclaw, K. M., Johnston, C. T., and LeVesque-Madore, C. S. 1981. Thermodynamics of sodium-copper exchange on Wyoming bentonite at 298°K. Soil Sci. Soc. Am. J. 45: 10791084.Google Scholar
Stadler, M., and Schindler, P. W. 1993. Modeling of H+ and Cu2+ adsorption on calcium-montmorillonite. Clays & Clay Miner. 41: 288296.Google Scholar
Thomson, B. M., Longmire, P. A., and Brookins, D. G. 1986. Geochemical constraints on underground disposal of uranium mill tailings. Applied Geochem. 1: 335343.CrossRefGoogle Scholar
Tripathi, V. J., 1984. Uranium (VI) transport modeling: Geochemical data and submodels. Ph.D. dissertation, Stanford University.Google Scholar
Tsunashima, A., Brindley, G. W., and Bastovanov, M. 1981. Adsorption of uranium from solutions by montmorillonite; Compositions and properties of uranyl montmorillonites. Clays & Clay Miner. 29: 1016.Google Scholar
Westall, J., 1982a. FITEQL. A computer program for determination of equilibrium constants from experimental data. Version 1.2. Report 82-01, Department of Chemistry, Oregon State University, Corvallis, Oregon.Google Scholar
Westall, J., 1982b. FITEQL. A computer program for determination of equilibrium constants from experimental data. Version 2.0. Report 82-02, Department of Chemistry, Oregon State University, Corvallis, Oregon.Google Scholar
Westall, J.C., and Herbellien, A.L. (1993) FITEQL. A computer program for determination of equilibrium constants from experimental data. Version 3.1. Report 94-01, Department of Chemistry, Oregon State University, Corvallis, Oregon.Google Scholar
White, G. N., and Zelazny, L. W. 1988. Analysis and implications of the edge structure of dioctahedral phyllosilicates. Clays & Clay Miner. 36: 141146.Google Scholar
Wieland, E., Kohler, M., and Leckie, J. O. 1995. Adsorption of neptunium [Np(V)] on mineral surfaces: Modeling of surface complexation. Geochim. Cosmochim. Acta (in press.)Google Scholar
Zachara, J. M., and Smith, S. C. 1994. Edge site contributions to Cd sorption on specimen and soil-derived smectite in Na+ and Ca2+ electrolytes. Soil Sci. Soc. Am. J. (in press).Google Scholar
Zachara, J. M., Smith, S. C., McKinley, J. P., and Resch, C. T. 1993. Cadmium sorption on specimen and soil smectites in Na+, Ca2+, and Na+/Ca2+ electrolytes. Soil Sci. Soc. Am. J. 57: 14911501.Google Scholar