Hostname: page-component-848d4c4894-xfwgj Total loading time: 0 Render date: 2024-07-06T22:34:48.085Z Has data issue: false hasContentIssue false

Atomic force microscopy study of montmorillonite dissolution under highly alkaline conditions

Published online by Cambridge University Press:  01 January 2024

Shingo Yokoyama
Affiliation:
Graduate School of Natural Science and Technology, Kanazawa University, Kakuma, Kanazawa, Ishikawa 920-1192, Japan
Masato Kuroda
Affiliation:
Graduate School of Natural Science and Technology, Kanazawa University, Kakuma, Kanazawa, Ishikawa 920-1192, Japan
Tsutomu Sato*
Affiliation:
Institute of Nature and Environmental Technology, Kanazawa University, Kakuma, Kanazawa, Ishikawa 920-1192, Japan
*
*E-mail address of corresponding author: tomsato@earth.s.kanazawa-u.ac.jp

Abstract

Montmorillonite dissolution under highly alkaline conditions (pH = 13.3; I = 0.3 M) was investigated by bulk dissolution methods and in situ atomic force microscopy (AFM). In bulk dissolution experiments, initial SiO2 concentrations were high, and a steady state was reached after 136 h. The dissolution rates derived from the edge surface area (ESA) at the steady-state condition at 30, 50 and 70°C were 3.39 x 10−12, 1.75 × 10−11 and 5.81 × 10−11 mol/m2 s, respectively. The AFM observations were conducted under three conditions: (Run I) short-term in situ batch dissolution at RT; (Run II) long-term in situ flow-through dissolution at RT; and (Run III) long-term batch dissolution at 50°C. The observed reductions in montmorillonite particle volume for Runs I and II were due primarily to edge-surface dissolution. The ESA-based dissolution rate for Run I (10−9 mol/m2 s) was three orders of magnitude faster than that for Run II (10−12 mol/m2 s). The rate obtained for Run II corresponded to the rate at the steady-state conditions in the bulk dissolution experiments. A small number of etch pits developed in Run III slightly increased the ESA of montmorillonite since most of the montmorillonite particles were separated into monolayers lacking three-dimensional periodicity. The ESA-based dissolution rate for Run III was 2.26 × 10−11 mol/m2 s. Dissolution rates based on long-term AFM observations could be directly compared with steady-state rates obtained from bulk dissolution experiments. The AFM observations indicated that dissolution occurred at edge surfaces; therefore, the ESA should be used to calculate the dissolution rate for montmorillonite under alkaline conditions. Dissolution rates of individual particles with different morphologies estimated by AFM were similar to rates estimated from bulk dissolution experiments.

Type
Research Article
Copyright
Copyright © Clay Minerals Society 2005

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

Atkinson, A., (1985) The time-dependence of pH within a repository for radioactive waste disposal London HMSO.Google Scholar
Bauer, A. and Berger, G., (1998) Kaolinite and smectite dissolution rate in high molar KOH solutions at 35° and 80°C Applied Geochemistry 13 905916 10.1016/S0883-2927(98)00018-3.Google Scholar
Bauer, A. and Velde, B., (1999) Smectite transformation in high molar KOH solutions Clay Minerals 34 259273 10.1180/000985599546226.Google Scholar
Bickmore, B.R. Hochella, M.F. Jr. Bosbach, D. and Charlet, L., (1999) Methods for performing atomic force microscopy imaging of clay minerals in aqueous solutions Clays and Clay Minerals 47 573581 10.1346/CCMN.1999.0470504.Google Scholar
Bickmore, B.R. Bosbach, D. Hochella, M.F. Jr. Charlet, L. and Rufe, E., (2001) In situ atomic force microscopy study of hectorite and nontronite dissolution: Implications for phyllosilicate edge surface structures and dissolution mechanisms American Mineralogist 86 411423 10.2138/am-2001-0404.Google Scholar
Bickmore, B.R. Rosso, K.M. Nagy, K.L. Cygan, R.T. and Tadanier, C.J., (2003) Ab initio determination of edge surface structure for dioctahedral 2:1 phyllosilicates: Implications for acid-base reactivity Clays and Clay Minerals 51 359371 10.1346/CCMN.2003.0510401.Google Scholar
Bosbach, D. Charlet, L. Bickmore, B. Hochella, M.F. Jr., (2000) The dissolution of hectorite: In-situ, real-time observations using atomic force microscopy American Mineralogist 85 12091216 10.2138/am-2000-8-914.Google Scholar
Brandt, F. Bosbach, D. Krawczyk-Bärsch, E. Arnold, T. and Bernhard, G., (2003) Chlorite dissolution in acid pH-range: A combined microscopic and macroscopic approach Geochimica et Cosmochimica Acta 67 14511461 10.1016/S0016-7037(02)01293-0.Google Scholar
Cama, J. Ganor, J. Ayora, C. and Lasaga, A., (2000) Smectite dissolution kinetics at 80°C and pH 8.8 Geochimica et Cosmochimica Acta 64 27012717 10.1016/S0016-7037(00)00378-1.Google Scholar
Chou, L. and Wollast, R., (1984) Study of the weathering of albite at room temperature and pressure with a fluidized bed reactor Geochimica et Cosmochimica Acta 48 22052217 10.1016/0016-7037(84)90217-5.Google Scholar
Chou, L. and Wollast, R., (1985) Steady-state kinetics and dissolution mechanism of albite American Journal of Science 285 963993 10.2475/ajs.285.10.963.Google Scholar
Claret, F. Bauer, A. Schäfer, T. Griffault, L. and Lanson, B., (2002) Experimental investigation of the interaction of clays with high-pH solutions: A case study from the Callovo-Oxfordian formation, Meuse-Haute Marne underground laboratory (France) Clays and Clay Minerals 50 633646 10.1346/000986002320679369.Google Scholar
Dove, P.M. Chermak, J.A., Nagy, K.L. and Blum, A.E., (1994) Mineral-water interactions: fluid cell applications of scanning force microscopy Scanning Probe Microscopy of Clay Minerals Boulder, Colorado The Clay Minerals Society 139169.Google Scholar
Dove, P.M. and Platt, F.M., (1996) Compatible real-time rates of mineral dissolution by atomic force microscopy (AFM) Chemical Geology 127 331338 10.1016/0009-2541(95)00127-1.Google Scholar
Ganor, J. Mogollón, J.L. and Lasaga, A.C., (1995) The effect of pH on kaolinite dissolution rates and on activation energy Geochimica et Cosmochimica Acta 59 10371052 10.1016/0016-7037(95)00021-Q.Google Scholar
Gautier, J.-M. Oelkers, E.H. and Schott, J., (2001) Are quartz dissolution rates proportional to B.E.T. surface area? Geochimica et Cosmochimica Acta 65 10591070 10.1016/S0016-7037(00)00570-6.Google Scholar
Hayashi, H. and Yamada, M., (1990) Kinetics of dissolution of noncrystalline oxides and crystalline clay minerals in basic tiron solution Clays and Clay Minerals 38 308314 10.1346/CCMN.1990.0380310.Google Scholar
Holdren, G.R. Jr. and Berner, R.A., (1979) Mechanism of feldspar weathering. I. Experimental studies Geochimica et Cosmochimica Acta 43 11611171 10.1016/0016-7037(79)90109-1.Google Scholar
Huertas, F.J. Chou, L. and Wollast, R., (1999) Mechanism of kaolinite dissolution at room temperature and pressure Part II: Kinetic study Geochimica et Cosmochimica Acta 63 32613275 10.1016/S0016-7037(99)00249-5.Google Scholar
Huertas, F.J. Caballero, E. de Jimenez Cisneros, C. Huertas, F. and Linares, J., (2001) Kinetics of montmorillonite dissolution in granitic solutions Applied Geochemistry 16 397407 10.1016/S0883-2927(00)00049-4.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 10.1016/0016-7037(88)90055-5.Google Scholar
Knauss, K.G. and Wolery, T.J., (1989) Muscovite dissolution kinetics as a function of pH and time at 70°C Geochimica et Cosmochimica Acta 53 14931501 10.1016/0016-7037(89)90232-9.Google Scholar
MacEwan, D.M.C. and Brown, G., (1961) Montmorillonite minerals The X-ray Identification and Crystal Structure of Clay Minerals London Mineralogical Society 143.Google Scholar
Nagy, K.L., White, A.F. and Brantley, S.L., (1995) Dissolution and precipitation kinetics of sheet silicates Chemical Weathering Rates of Silicate Minerals Washington, D.C Mineralogical Society of America 173233 10.1515/9781501509650-007.Google Scholar
Rassineux, F. Griffault, L. Meunier, A. Berger, G. Petit, S. Vieillard, P. Zellagui, R. and Munoz, M., (2001) Expandability-layer stacking relationship during experimental alteration of a Wyoming bentonite in pH 13.5 solutions at 35 and 60°C Clay Minerals 36 197210 10.1180/000985501750177933.Google Scholar
Rufe, E. Hochella, M.F. Jr., (1999) Quantitative assessment of reactive surface area of phlogopite during acid dissolution Science 285 874876 10.1126/science.285.5429.874.Google Scholar
Schott, J. Berner, R.A. and Sjöberg, E.L., (1981) Mechanism of pyroxene and amphibole weathering — I. Experimental studies of iron-free minerals Geochimica et Cosmochimica Acta 45 21232135 10.1016/0016-7037(81)90065-X.Google Scholar
Sposito, G., (1984) The Surface Chemistry of Soils Oxford, UK Oxford University Press.Google Scholar
Stillings, L.L. and Brantley, S.L., (1995) Feldspar dissolution at 25°C and pH 3: Reaction stoichiometry and pH effect of cations Geochimica et Cosmochimica Acta 59 14831496 10.1016/0016-7037(95)00057-7.Google Scholar
Taubald, T. Bauer, A. Schäfer, T. Geckeis, H. Satir, M. and Kim, J.I., (2000) Experimental investigation of the effect of high-pH solutions on the Opalinus Shale and the Hammerschmiede smectite Clay Minerals 35 515524 10.1180/000985500546981.Google Scholar
Tournassat, C. Neaman, A. Villieras, F. Bosbach, D. and Charlet, L., (2003) Nanomorphology of montmorillonite particles: estimation of the clay edge sorption site density by low-pressure gas adsorption and AFM observations American Mineralogist 88 19891995 10.2138/am-2003-11-1243.Google Scholar
Zysset, M. and Schindler, P.W., (1996) The proton promoted dissolution kinetics of K-montmorillonite Geochimica el Cosmochimica Acta 60 921931 10.1016/0016-7037(95)00451-3.Google Scholar