Hostname: page-component-7bb8b95d7b-pwrkn Total loading time: 0 Render date: 2024-09-23T17:24:00.819Z Has data issue: false hasContentIssue false

Np-Incorporation Into K-boltwoodite

Published online by Cambridge University Press:  01 February 2011

Lindsay C. Shuller
Affiliation:
Materials Science and Engineering, University of Michigan
Rodney C. Ewing
Affiliation:
Materials Science and Engineering, University of Michigan Geological Sciences, University of Michigan
Udo Becker
Affiliation:
Geological Sciences, University of Michigan
Get access

Abstract

Np-237 (τ1/2 = 2.1 million years) is a potentially important contributor to the total dose for a geologic repository under oxidizing conditions. Further, the Np5+-complexes are mobile aqueous species. Several processes may limit the transport of Np, as well as other actinides: i) the precipitation of Np-solids, ii) the incorporation of Np into secondary uranium phases, and iii) the sorption and reduction of Np-complexes on Fe-oxide surfaces. This study utilizes quantum-mechanical calculations to determine the most energetically favorable Np5+-incorporation mechanisms into uranyl phases, where Np5+-substitution for U6+ requires a charge-balancing mechanism, such as the addition of H+ into the structure. Experimental results suggest that uranyl structures with charged interlayer cations have a greater affinity for Np5+ than uranyl structures without interlayer cations. Therefore, the uranyl silicate phase boltwoodite (KUO2(SiO3OH)(H2O)1.5) is selected for this computational investigation. The charge-balancing mechanisms considered to occur with substitution include: i) addition of H+, ii) substitution of Ca2+ for K+, and iii) substitution of P5+ for Si4+. While the incorporation energy results (1-3 eV)are higher than energies expected based on current experimental studies, solid-solution calculations are used to estimate the limit of Np incorporation for the P5+ substitution mechanism (10 ppm at ̃100°C). The electronic structure of the boltwoodite structure provides insight into the electron density that may be involved in the incorporation of Np into the structure.

Type
Research Article
Copyright
Copyright © Materials Research Society 2008

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

1 Hedin, A. (1997) Spent Nuclear Fuel – How dangerous is it? SKB technical Report 97-13, Swedish Nuclear Fuel and Waste Management Co.: 60.Google Scholar
2 Finch, R.J. and Ewing, R.C. (1992) The corrosion of uraninite under oxidizing conditions. Journal of Nuclear Materials, 190: 133156.Google Scholar
3 Bruno, J. and Ewing, R.C. (2006) Spent nuclear fuel. Elements, 2: 343349.Google Scholar
4 Burns, P.C., Ewing, R.C., and Miller, M.L. (1997) Incorporation mechanisms of actinide elements into the structures of U6+ phases formed during the oxidation of spent nuclear fuel. Journal of Nuclear Materials, 245: 19.Google Scholar
5 Buck, E.C., Finch, R.J., Finn, P.A., and Bates, J.K. (1998) Retention of neptunium in uranyl alteration phases formed during spent fuel corrosion. Materials Research Society Symposium Proceedings, 506: 8794.Google Scholar
6 Burns, P.C., Deely, K.M., and Skanthakumar, S. (2004) Neptunium incorporation into uranyl compounds that form as alteration products of spent nuclear fuel: Implications for geologic repository performance. Radiochimica Acta, 92: 151159.Google Scholar
7 Klingensmith, A.L., Deely, K.M., Kinman, W.S., Kelly, V., and Burns, P.C. (2007) Neptunium incorporation into sodium-substituted schoepite. American Mineralogist, 92: 662669.Google Scholar
8 Douglas, M., Clark, S.B., Friese, J.I., Arey, B.W., Buck, E.C., and Hanson, B.D. (2005) Neptunium(V) partitioning to uranium(VI) oxide and peroxide solids. Environmental Science and Technology, 39: 41174124.Google Scholar
9 Burns, P.C., Miller, M.L., and Ewing, R.C. (1996) U6+minerals and inorganic phases: A comparison and hierarchy of crystal structures. Canadian Mineralogist, 34: 845880.Google Scholar
10 Burns, P.C. (2005) U6+ minerals and inorganic compounds: insights into an expanded structural hierarchy of crystal structures. Canadian Mineralogist, 43: 18391894.Google Scholar
11 Payne, M.C., Teter, M.P., Allan, D.C., Arias, T.A., and Joannopoulos, J.D. (1992) Iterative minimization techniques for abinitio total-energy calculations - molecular-dynamics and conjugate gradients. Reviews of Modern Physics, 64(4): 10451097.Google Scholar
12 Perdew, J.P., Burke, K., and Ernzerhof, M. (1996) Generalized gradient approximation made simple. Physical Review Letters, 77(18): 38653868.Google Scholar
13 Robie, R.A., and Hemingway, B.S. (1995) Thermodynamic properties of minerals and related substances at 298.15 K and 1 bar (105 pascals) pressure and at higher temperatures. U.S. Geological Survey Bulletin, 2131.Google Scholar
14 Forbes, T.Z., and Burns, P.C. (2006) Ba(NpO2)(PO4)(H2O), its relationship to the uranophane group, and implications for Np incorporation in uranyl minerals. American Mineralogist, 91(7): 10891093.Google Scholar