Hostname: page-component-848d4c4894-xfwgj Total loading time: 0 Render date: 2024-06-22T11:08:21.390Z Has data issue: false hasContentIssue false

Dissolution of diopside and basaltic glass: the effect of carbonate coating

Published online by Cambridge University Press:  05 July 2018

G. Stockmann*
Affiliation:
Institute of Earth Sciences, University of Iceland, Askja, Sturlugata 7, 101 Reykjavik, Iceland
D. Wolff-Boenisch
Affiliation:
Institute of Earth Sciences, University of Iceland, Askja, Sturlugata 7, 101 Reykjavik, Iceland
S. R. Gíslason
Affiliation:
Institute of Earth Sciences, University of Iceland, Askja, Sturlugata 7, 101 Reykjavik, Iceland
E. H. Oelkers
Affiliation:
Géochimie et Biogéochimie Expérimentale, Université Paul Sabatier, CNRS-UMR 5563, 14 rue Edouard Belin, 31400 Toulouse, France
*
*E-mail: gjs3@hi.is

Abstract

Far-from-equilibrium dissolution experiments with diopside and basaltic glass in mixed-flow reactors at 70°C and pH 8.2 show that solute concentrations do not reach steady state over the experimental duration of 45—60 days. Chemical modelling indicates that during the dissolution experiments, solutions have become supersaturated with respect to carbonates in the case of diopside, and carbonates, clay minerals and zeolites in the case of the basaltic glass. Decreasing dissolution is therefore interpreted as a result of secondary surface precipitates blocking the reactive surface area. Calcite formation was supported in both experiments by a significant increase in Ca (and Sr) concentrations as pH was abruptly lowered from 8.2 to 7 because this change increased carbonate solubility and caused all potential carbonate precipitates to re-dissolve. The reduction in pH also led to an increase in Si concentration for diopside and a decrease in Si concentration for basaltic glass. This observation is in accordance with previous experiments on the pH-dependent dissolution rates of pyroxenes and basaltic glass.

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

Broecker, W.S. (2005) Global warming: Take action or wait. Jökull, 55, 1–16.Google Scholar
Cubillas, P., Köhler, P., Prieto, P. and Oelkers, E. H. (2005) Experimental determination of the dissolution rates of calcite, aragonite and bivalves. Chemical Geology, 216, 59–67.CrossRefGoogle Scholar
Gislason, S.R. and Oelkers, E.H. (2003) The mechanism, rates and consequences of basaltic glass dissolution: II. An experimental study of the dissolution rates of basaltic glass as a function of pH and temperature. Geochimica et Cosmochimica Ada, 67, 3817–3832.CrossRefGoogle Scholar
Gislason, S.R., Oelkers, E.H. and Snorrason, Á. (2006) The role of river suspended material in the global carbon cycle. Geology, 34, 49–52.CrossRefGoogle Scholar
Golubev, S.V., Pokrovsky, O.S. and Schott, J. (2005) Experimental determination of the effect of dissolved CO2 on the dissolution kinetics of Mg and Ca silicates at 25°C. Chemical Geology, 217, 227–238.CrossRefGoogle Scholar
Gysi, A.P. and Stefansson, A. (2008) Numerical simulation of CO2-water basalt interaction. Mineralogical Magazine, 72, 55–59.CrossRefGoogle Scholar
Knauss, K.G., Nguyen, S.N. and Weed, H.C. (1993) Diopside dissolution kinetics as a function of pH, CO2, temperature, and time. Geochimica et Cosmochimica Ada, 57, 285–294.CrossRefGoogle Scholar
Hodson, M. (2002) The influence of Fe-rich coatings on the dissolution of anorthite at pH 2.6. Geochimica et Cosmochimica Ada, 67, 3355–3363.Google Scholar
Hoffert, M.I., Caldeira, K., Benford, G., Criswell, D.R., Green, C, Herzog, H., Jain, A.K., Kheshgi, H.S., Lackner, K.S. Lewis, J.S., Lightfoot, H.D., Manheimer, W. Mankins, J.C., Mauel, M.E., Perkins, L.J., Schlesinger, M.E., Volk, T. and Wigley, T.M.L. (2002) Advanced technology paths to global climate stability. Energy for a greenhouse planet. Science, 298, 981–987.CrossRefGoogle ScholarPubMed
McGrail, B.P., Schaef, H.T., Ho, A.M., Chien, Y.-J., Dooley, J.J. and Davidson, C.L. (2006) Potential for carbon dioxide sequestration in flood basalts. Journal of Geophysical Research, 111, B12201, doi:10.1029/2005JB004169.CrossRefGoogle Scholar
Oelkers, E.H. and Gislason, S.R. (2001) The mechanism, rates and consequences of basaltic glass dissolution: I. An experimental study of the dissolution rates of basaltic glass as a function of aqueous Al, Si and oxalic acid concentration at 25°C and pH = 3 and 11. Geochimica et Cosmochimica Ada, 65, 3671–3681.Google Scholar
Oelkers, E.H. and Schott, J. (2001) An experimental study of enstatite dissolution rates as a function of pH, temperature, and aqueous Mg and Si concentration, and the mechanism of pyroxene/pyroxenoid dissolution. Geochimica et Cosmochimica Ada, 65, 1219–1231.Google Scholar
Parkhurst, D. L. and Appelo, C.A.J. (1999) User's guide to PHREEQC (Version 2) — A computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations. USGS, Water Resources Investigation Report, 99–4259.Google Scholar
Wolff-Boenisch, D., Gislason, S.R., Oelkers, E.H. and Putnis, C.V. (2004) The dissolution rates of natural glasses as a function of their composition at pH 4 and 10.6, and temperatures from 25 to 74°C. Geochimica et Cosmochimica Ada, 68, 4843–4858.CrossRefGoogle Scholar