Hostname: page-component-7479d7b7d-jwnkl Total loading time: 0 Render date: 2024-07-16T00:26:50.508Z Has data issue: false hasContentIssue false

Methods for handling redox-sensitive smectite dispersions

Published online by Cambridge University Press:  27 February 2018

J. W. Stucki*
Affiliation:
Department of Natural Resources and Environmental Sciences, University of Illinois, Urbana, Illinois, USA
K. Su*
Affiliation:
Department of Natural Resources and Environmental Sciences, University of Illinois, Urbana, Illinois, USA
L. Pentráková
Affiliation:
Department of Natural Resources and Environmental Sciences, University of Illinois, Urbana, Illinois, USA
M. Pentrák
Affiliation:
Department of Natural Resources and Environmental Sciences, University of Illinois, Urbana, Illinois, USA
*
§Address for correspondence: W-321 Turner Hall, 1102 S. Goodwin Avenue, Urbana, IL, 61801, USA
Current address: Faculty of Geosciences and Environmental Engineering, Southwest Jiaotong University, Chengdu, 610031, China

Abstract

Redox activation (reduction of structural Fe) of smectites greatly alters their chemical reactivity and physical properties, which may be exploited for various environmental, agricultural or industrial purposes. Their re-oxidation during preparation, characterization, and use is, however, a significant risk to their utility. In this study, methods and apparatus were developed and described which mitigated reoxidation. Ferruginous smectite (sample SWa-1, Na saturated) was used as the model smectite. It was reduced with sodium dithionite in a citrate-bicarbonate buffer solution at 70°C for 4 h, which achieved a maximum Fe(II)/total Fe ratio of 0.9113 ± 0.0048. The first step in rendering reduced samples useful is to remove from them the reducing agents and other solutes present during reduction. This was accomplished in the present study by reducing the sample in an inert-atmosphere reaction tube (IRT) (a 50 mL centrifuge tube equipped with a removable septum cap), then removing solutes from the suspension by centrifuge washing. The washing steps were performed with the aid of a controlled-atmosphere liquid exchanger (CALE) which provided connections between the sample suspension and deoxygenated solutions. The reduced state was measured by 1, 10-phenanthroline or by Mössbauer spectroscopy at 77 K to give Fe(II)/total Fe ratios. Some samples were freeze dried after washing. Results revealed that if reduced smectites are washed without protection from atmospheric O2, the extent of reoxidation is on the order of 40 to 60%. If the sample is subsequently dried, reoxidation increases to more than 76%. If the sample is protected using the IRT and the CALE, however, reoxidation is decreased to less than 2%. Freeze drying in a glove box allowed reoxidaton to increase to slightly more than 10%. These results indicate that more reoxidation occurred during the drying stage than during the washing stage. These observations lead to the conclusions that (1) protection of reduced samples from atmospheric O2 is essential if extensive reoxidation is to be prevented, and (2) the methods and apparatus described herein are effective for accomplishing that purpose in abiotically reduced smectites. They may also be effective if applied to microbially reduced smectites.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 2014

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

Bishop, M.E., Dong, H., Kukkadapu, R.K., Liu, C. & Edelman, R.E. (2011) Bioreduction of Fe-bearing clay minerals and their reactivity toward pertechnetate (Tc-99) Geochimica et Cosmochimica Acta, 75, 52295246.Google Scholar
Cervini-Silva, J., Hernández-Pineda, J., Rivas-Valdés, M.T., Cornejo-Garrido, H., Guzmán, J., Fernández-Lomelín, P. & Del Razo, L.M. (2010) Arsenic(III) methylation in betaine-nontronite clay-water suspensions under environmental conditions. Journal of Hazardous Materials, 178, 450454.CrossRefGoogle ScholarPubMed
Dong, H.L., Kostka, J.E. & Kim, J. (2003a) Microscopic evidence for microbial dissolution of smectite. Clays and Clay Minerals, 51, 502512.Google Scholar
Dong, H.L., Kukkadapu, R.K., Fredrickson, J.K., Zachara, J.M., Kennedy, D.W. & Kostandarithes, H.M. (2003b) Microbial reduction of structural Fe(III) in illite and goethite. Environmental Science and Technology, 37, 12681276.CrossRefGoogle Scholar
Dong, H.L., Jaisi, D.P., Kim, J. & Zhang, G.X. (2009) Microbe-clay mineral interactions. American Mineralogist, 94, 15051519.CrossRefGoogle Scholar
Dyar, M.D., Schaefer, M.W., Sklute, E.C. & Bishop, J.L. (2008) Mössbauer spectroscopy of phyllosilicates: Effects of fitting models on recoil-free fractions and redox ratios. Clay Minerals, 43, 333.CrossRefGoogle Scholar
Hofstetter, T.B., Schwarzenbach, R.P. & Haderlein, S.B. (2003) Reactivity of Fe(II) species associated with clay minerals. Environmental Science and Technology, 37, 519528.CrossRefGoogle ScholarPubMed
Jaisi, D.P., Kukkadapu, R.K., Eberl, D.D. & Dong, H. (2005) Control of Fe(III) site occupancy on the rate and extent of microbial reduction of Fe(III) in nontronite. Geochimica et Cosmochimica Acta, 69, 54295440.CrossRefGoogle Scholar
Jaisi, D.P., Dong, H.L. & Liu, C.X. (2007) Kinetic analysis of microbial reduction of Fe(III) in nontronite. Environmental Science and Technology, 41, 24372444.Google Scholar
Jaisi, D.P., Dong, H.L., Plymale, A.E., Fredrickson, J.K., Zachara, J.M., Heald, S. & Liu, C.X. (2009) Reduction and long-term immobilization of technetium by Fe(II) associated with clay mineral nontronite. Chemical Geology, 264, 127138.Google Scholar
Komadel, P. & Stucki, J.W. (1988) Quantitative assay of minerals for Fe2+ and Fe3+ using 1,10-phenanthroline: III. A rapid photochemical method. Clays and Clay Minerals, 36, 379381.Google Scholar
Komadel, P., Lear, P.R. & Stucki, J.W. (1990) Reduction and reoxidation of iron in nontronites: Rate of reaction and extent of reduction: Clays and Clay Minerals, 37, 203208.Google Scholar
Komadel, P., Madejová, J. & Stucki, J.W. (1995) Reduction and reoxidation of nontronites: questions of reversibility. Clays and Clay Minerals, 43, 105110.Google Scholar
Komadel, P., Madejová, J. & Stucki, J.W. (2006) Strucural Fe(III) reduction in smectites. Applied Clay Science, 34, 8894.Google Scholar
Kostka, J.E., Wu, J., Nealson, K.H. & Stucki, J.W. (1999) The impact of structural Fe(III) reduction by bacteria on the surface chemistry of smectite clay minerals. Geochimica et Cosmochimica Acta, 63, 37053713.CrossRefGoogle Scholar
Kukkadapu, R.K., Zachara, J.M., Fredrickson, J.K., McKinley, J.P., Kennedy, D.W., Smith, S.C. & Dong, H.L. (2006) Reductive biotransformation of Fe in shale-limestone saprolite containing Fe(III) oxides and Fe(II)/Fe(III) phyllosilicates. Geochimica et Cosmochimica Acta, 70, 36623676.Google Scholar
Lee, K., Kostka, J.E. & Stucki, J.W. (2006) Comparisons of structural Fe reduction in smectites by bacteria and dithionite: An infrared spectroscopic study. Clays and Clay Minerals, 54, 195208.CrossRefGoogle Scholar
Li, Y.L., Vali, H., Sears, S.K., Yang, J., Deng, B.L. & Zhang, C.L. (2004) Iron reduction and alteration of nontronite NAu-2 by a sulfate reducing bacterium. Geochimica et Cosmochimica Acta, 68, 32513260.CrossRefGoogle Scholar
Manceau, A., Lanson, B., Drits, V.A., Chateigner, D., Gates, W.P., Wu, J., Huo, D. & Stucki, J.W. (2000a) Oxidation-reduction mechanism of iron in dioctahedral smectites: I. Crystal chemistry of oxidized reference nontronites. American Mineralogist, 85, 133152.Google Scholar
Manceau, A., Drits, V.A., Lanson, B., Chateigner, D., Wu, J., Huo, D., Gates, W.P. & Stucki, J.W. (2000b) Oxidation-reduction mechanism of iron in dioctahedral smectites: II. Crystal chemistry of reduced Garfield nontronite. American Mineralogist, 85, 153172.CrossRefGoogle Scholar
Merola, R.B., Fournier, E.D. & McGuire, M.M. (2007) Spectroscopic investigations of Fe2+ complexation on nontronite clay. Langmuir, 23, 12231226.Google Scholar
Neumann, A., Hofstetter, T.B., Skarpeli-Liati, M. & Schwarzenbach, R.P. (2009) Reduction of polychlorinated ethanes and carbon tetrachloride by structural Fe(II) in smectites. Environmental Science and Technology, 43, 40824089.CrossRefGoogle ScholarPubMed
O’Reilly, S.E., Watkins, J. & Furukawa, Y. (2005) Secondary mineral formation associated with respiration of nontronite, NAu-1 by iron reducing bacteria. Geochemical Transactions, 6, 6776.CrossRefGoogle ScholarPubMed
O’Reilly, S.E., Furukawa, Y. & Newell, S. (2006) Dissolution and microbial Fe(III) reduction of nontronite (NAu-1). Chemical Geology, 235, 111.Google Scholar
Pentráková, L., Su, K., Pentrák, M. & Stucki, J.W. (2013) A review of microbial redox interactions with structural Fe in clay minerals. Clay Minerals, 48, 543560.Google Scholar
Peretyazhko, T., Zachara, J.M., Heald, S.M., Jeon, B.H., Kukkadapu, R.K., Liu, C., Moore, D. & Resch, C.T. (2008) Heterogeneous reduction of Tc(VII) by Fe(II) at the solid-water interface. Geochimica et Cosmochimica Acta, 72, 15211539.CrossRefGoogle Scholar
Ribeiro, F.R., Fabris, J.D., Kostka, J.E., Komadel, P. & Stucki, J.W. (2009) Comparisons of structural iron reduction in smectites by bacteria and dithionite: II. A variable-temperature Mö ssbauer spectroscopic study of Garfield nontronite. Pure and Applied Chemistry, 81, 14991509.CrossRefGoogle Scholar
Roth, C.B. & Tullock, R.J. (1973) Deprotonation of nontronite resulting from chemical reduction of structural ferric iron. Proceedings of the Sixth International Clay Conference, Madrid, 1972, 107114.Google Scholar
Roth, C.B., Jackson, M.L. & Syers, J.K. (1969) Deferration effect on structural ferrous-ferric iron ratio and CEC of vermiculites and soils. Clays and Clay Minerals, 17, 253264.CrossRefGoogle Scholar
Rozenson, I. & Heller-Kallai, L. (1976) Reduction and oxidation of Fe3+ in dioctahedral smectite. I: Reduction with hydrazine and dithionite: Clays and Clay Minerals, 24, 271282.CrossRefGoogle Scholar
Russell, J.D., Goodman, B.A. & Fraser, A.R. (1979) Infrared and Mössbauer studies of reduced nontronites. Clays and Clay Minerals, 27, 6371.Google Scholar
Schaefer, M.V., Gorski, C.A. & Scherer, M.M. (2011) Spectroscopic evidence for interfacial Fe(II)-Fe(III) electron transfer in a clay mineral. Environmental Science and Technology, 45, 540545.Google Scholar
Schilt, A.A. (1969). Analytical applications of 1,10- phenanthroline and related compounds. International Series Monographs in Analytical Chemistry, 32. Pergamon Press, Oxford, UK.Google Scholar
Southam, G. (2012) Minerals as substrates for life: The prokaryotic view. Elements, 8, 101105.Google Scholar
Stucki, J.W. (1975) Chemical and spectroscopic analysis of oxidation-reduction mechanisms for structural iron in nontronite. Ph.D. Dissertation, Purdue University, West Lafayette, Indiana, USA.Google Scholar
Stucki, J.W. (1988) Structural iron in smectites. Pp 625–675 in: Iron in Soils and Clay Minerals (J.W. Stucki, B.A. Goodman & U. Schwertmann, editors). D.Reidel, Dordrecht, The Netherlands.Google Scholar
Stucki, J.W. (2006) Properties and behavior of iron in clay minerals. Pp. 429–482 in: Handbook of Clay Science (F. Bergaya, B.K.G. Theng & G. Lagaly, editors.). Elsevier, Amsterdam.Google Scholar
Stucki, J.W. (2013) Properties and behavior of iron in clay minerals. Pp. 559–611 in: Handbook of Clay Science, 5A (F. Bergaya, B.K.G. Theng & G. Lagaly, editors). Elsevier, Amsterdam.Google Scholar
Stucki, J.W. & Roth, C.B. (1977) Oxidation-reduction mechanism for structural iron in nontronite. Soil Science Society of America Journal, 41, 808814.Google Scholar
Stucki, J.W. & Tessier, D. (1991) Effects of iron oxidation state on the texture and structural order of Na-nontronite gels. Clays and Clay Minerals, 39, 137143.Google Scholar
Stucki, J.W., Golden, D.C. & Roth, C.B. (1984) Preparation and handling of dithionite-reduced smectite suspensions. Clays and Clay Minerals, 32, 191197.Google Scholar
Stucki, J.W., Komadel, P. & Wilkinson, H.T. (1987) Microbial reduction of structural iron(III) in smectites. Soil Science Society of America Journal, 51, 16631665.Google Scholar
Stucki, J.W., Lee, K., Zhang, L.Z. & Larson, R.A. (2002) Effects of iron oxidation state on the surface and structural properties of smectites. Pure and Applied Chemistry, 74, 21452158.Google Scholar
Su, K., Radian, A., Mishael, Y., Yang, L. & Stucki, J.W. (2012) Nitrate reduction by redox-activated, polydiallyldimethylammonium- exchanged ferruginous smectite. Clays and Clay Minerals, 59, 464472.Google Scholar
Violet, C.E. & Pipkorn, D.N. (1971) Mössbauer line positions and hyperfine interactions in alpha iron. Journal of Applied Physics, 42, 43394342.CrossRefGoogle Scholar
Wu, J., Low, P.F. & Roth, C.B. (1989) Effects of octahedral iron reduction and swelling pressure on interlayer distances in Na-nontronite. Clays and Clay Minerals, 37, 211218.Google Scholar
Zhang, G., Kim, J., Dong, H. & Sommer, A.J. (2007) Microbial effects in promoting the smectite to illite reaction: Role of organic matter intercalated in the interlayer. American Mineralogist, 92, 14011410.Google Scholar
Zhang, G., Senko, J.M., Kelly, S.D., Tan, H. Kemner, K.M. & Burgos, W.D. (2009) Microbial reduction of iron(III)-rich nontronite and uranium(VI). Geochimica et Cosmochimica Acta, 73, 35233538.Google Scholar
Zhuang, Y., Fialips, C.I., White, M.L. & Perez Ferrandez, D.M. (2012) New redox-active material for permeable water remediation systems. Applied Clay Science, 59–60, 2635.Google Scholar