Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-20T00:04:42.596Z Has data issue: false hasContentIssue false

Ionicity of Clay–Cation Bonds in Relation to Dispersive Behavior of Mg and K Soil Clays as Influenced by pH

Published online by Cambridge University Press:  01 January 2024

Yingcan Zhu*
Affiliation:
Centre for Sustainable Agricultural Systems, University of Southern Queensland, Toowoomba, QLD 4350, Australia
Alla Marchuk
Affiliation:
Centre for Sustainable Agricultural Systems, University of Southern Queensland, Toowoomba, QLD 4350, Australia
John McLean Bennett
Affiliation:
Centre for Sustainable Agricultural Systems, University of Southern Queensland, Toowoomba, QLD 4350, Australia
*
*E-mail address of corresponding author: Yingcan.Zhu@usq.edu.au

Abstract

The dispersive behavior dynamics of clay determine soil characteristics such as permeability and aggregate stability, and, consequently, crop productivity. Soil dispersion is heavily influenced by the ionicity of clay–cation bonds and has been shown to be related to the net negative charge and pH of the system. Little work has been done, however, which considers these factors together, especially for K and Mg clays. The objective of the present study was to investigate the effect of changing pH on the dispersive behavior of Mg and K homoionic clays, in comparison to Ca and Na clays under equivalent pH conditions. The clay fractions used here were extracted from three soils and have distinctly different mineralogies. These clays were treated to become homoionic with regard to Na, K, Ca, and Mg. Excess salts were removed by dialysis and pH was adjusted to 3, 4, 5, 6, 7, 8, 9, 10, and 11 for all clays, except Mg (pH range 3–7). Clay dispersion-flocculation dynamics were investigated, and the net negative charge, pH, electrical conductivity (EC), and turbidity were measured. Mg has a similar but less flocculative effect than Ca, while K has a similar but less dispersive effect than Na, under similar pH conditions. The dispersive behavior of Na, K, Mg, and Ca homoionic clays was correlated well with the ionicity of clay–cation bonds at equivalent pH, with the degree of clay dispersion being explained by the pH, EC, ionicity, ζ-potential, and mean particle size of the clay–cation system. A predictive model for dispersion was developed with its applicability and limitations discussed.

Type
Article
Copyright
Copyright © Clay Minerals Society 2020

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

Agassi, M., Shainberg, I., & Morin, J. (1981). Effect of electrolyte concentration and soil sodicity on infiltration rate and crust formation. Soil Science Society of America Journal, 45, 848851.CrossRefGoogle Scholar
Amezketa, E. (1999). Soil aggregate stability: A review. Journal of Sustainable Agriculture, 14, 83151.CrossRefGoogle Scholar
Arienzo, M., Christen, E., Quayle, W., & Kumar, A. (2009). A review of the fate of potassium in the soil–plant system after land application of wastewaters. Journal of Hazardous Materials, 164, 415422.CrossRefGoogle ScholarPubMed
Bennett, J. M., Marchuk, A., & Marchuk, S. (2016). An alternative index to ESP for explanation of dispersion occurring in soils. Soil Research, 54, 949957.CrossRefGoogle Scholar
Bergaya, F., Lagaly, G., & Vayer, M. (2006) Cation and anion exchange. Pp. 9791001 in: Handbook of Clay Science (Bergaya, F., Theng, B.K.G., and Lagaly, G., editors). Developments in Clay Science, 1, Elsevier Ltd., Amsterdam.CrossRefGoogle Scholar
Boström, M., Williams, D. R. M., & Ninham, B. W. (2001). Specific ion effects: why DLVO theory fails for biology and colloid systems. Physical Review Letters, 87, 168103.CrossRefGoogle ScholarPubMed
Bradfield, R. (1936). The value and limitations of calcium in soil structure. Soil Science Society of America Journal, 2001(17), 3132.CrossRefGoogle Scholar
Brigatti, M. F., Galan, E., & Theng, B. K. G. (2006). Structures and mineralogy of clay minerals. Developments in Clay Science, 1, 1986.CrossRefGoogle Scholar
Brown, M. A., Goel, A., & Abbas, Z. (2016). Effect of electrolyte concentration on the stern layer thickness at a charged interface. Angewandte Chemie International Edition, 55, 37903794.CrossRefGoogle Scholar
Chorom, M., & Rengasamy, P. (1995). Dispersion and zeta potential of pure clays as related to net particle charge under varying pH, electrolyte concentration and cation type. European Journal of Soil Science, 46, 657665.CrossRefGoogle Scholar
Churchman, G. J., & Weissmann, D. A. (1995). Separation of submicron particles from soils and sediments without mechanical disturbance. Clays and Clay Minerals, 43, 8591.CrossRefGoogle Scholar
Curtin, D., Steppuhn, H., & Selles, F. (1994a). Clay dispersion in relation to sodicity, electrolyte concentration, and mechanical effects. Soil Science Society of America Journal, 58, 955962.CrossRefGoogle Scholar
Curtin, D., Steppuhn, H., & Selles, F. (1994b). Effects of magnesium on cation selectivity and structural stability of sodic soils. Soil Science Society of America Journal, 58, 730737.CrossRefGoogle Scholar
Dang, A., Bennett, J. M., Marchuk, A., Biggs, A., & Raine, S. (2018). Quantifying the aggregation-dispersion boundary condition in terms of saturated hydraulic conductivity reduction and the threshold electrolyte concentration. Agricultural Water Management, 203, 172178.CrossRefGoogle Scholar
Delgado, A., Gonzalez-Caballero, F., & Bruque, J. M. (1986). On the zeta potential and surface charge density of montmorillonite in aqueous electrolyte solutions. Journal of Colloid and Interface Science, 113, 203211.CrossRefGoogle Scholar
Derjaguin, B. V. (1941). Theory of the stability of strongly charged lyophobic sols and the adhesion of strongly charged particles in solutions of electrolytes. Acta Physicochimica USSR, 14, 633662.Google Scholar
Emerson, W. W., & Chi, C. L. (1977). Exchangeable calcium, magnesium and sodium and the dispersion of illites in water. II. Dispersion of illites in water. Soil Research, 15, 255262.CrossRefGoogle Scholar
Frenkel, H., Goertzen, J. O., & Rhoades, J. D. (1978). Effects of clay type and content, exchangeable sodium percentage, and electrolyte concentration on clay dispersion and soil hydraulic conductivity. Soil Science Society of America Journal, 42, 3239.CrossRefGoogle Scholar
Goldberg, S., & Glaubig, R. A. (1987). Effect of saturating cation, pH, and aluminum and iron oxide on the flocculation of kaolinite and montmorillonite. Clays and Clay Minerals, 35, 220227.CrossRefGoogle Scholar
Goldberg, S., Kapoor, B. S., & Rhoades, J. D. (1990). Effect of aluminum and iron oxides and organic matter on flocculation and dispersion of arid zone soils. Soil Science, 150, 588593.CrossRefGoogle Scholar
Goldberg, S., Forster, H. S., & Heick, E. L. (1991). Flocculation of illite/kaolinite and illite/montmorillonite mixtures as affected by sodium adsorption ratio and pH. Clays and Clay Minerals, 39, 375380.CrossRefGoogle Scholar
Greene, R., Posner, A., & Quirk, J. (1973). Factors affecting the formation of quasi-crystals of montmorillonite. Soil Science Society of America Journal, 37, 457460.CrossRefGoogle Scholar
Greene, R., Posner, A., & Quirk, J. (1978). A study of the coagulation of montmorillonite and illite suspensions by calcium chloride using the electron microscope. In Emerson, W. W., Bond, R. D., & Dexter, A. R. (Eds.), Modification of soil structure. New York: John Wiley & Sons.Google Scholar
Gupta, V., & Miller, J. D. (2010). Surface force measurements at the basal planes of ordered kaolinite particles. Journal of Colloid and Interface Science, 344, 362371.CrossRefGoogle ScholarPubMed
Hartley, P. G., Larson, I., & Scales, P. J. (1997). Electrokinetic and direct force measurements between silica and mica surfaces in dilute electrolyte solutions. Langmuir, 13, 22072214.CrossRefGoogle Scholar
Hu, F., Liu, J., Xu, C., Du, W., Yang, Z., Liu, X., Liu, G., & Zhao, S. (2018a). Soil internal forces contribute more than raindrop impact force to rainfall splash erosion. Geoderma, 330, 9198.CrossRefGoogle Scholar
Hu, F., Liu, J., Xu, C., Wang, Z., Liu, G., Li, H., & Zhao, S. (2018b). Soil internal forces initiate aggregate breakdown and splash erosion. Geoderma, 320, 4351.CrossRefGoogle Scholar
Huheey, J., Keiter, E., & Keiter, R. (1993). Inorganic Chemistry: Principles of Structure and Reactivity. New York: HarperCollins College Publishers.Google Scholar
Ibanez, M., Wijdeveld, A., & Chassagne, C. (2014). The role of monoand divalent ions in the stability of kaolinite suspensions and fine tailings. Clays and Clay Minerals, 62, 374385.CrossRefGoogle Scholar
Karraker, K., & Radke, C. (2002). Disjoining pressures, zeta potentials and surface tensions of aqueous non-ionic surfactant/electrolyte solutions: theory and comparison to experiment. Advances in Colloid and Interface Science, 96, 231264.CrossRefGoogle ScholarPubMed
Keren, R. (1991). Specific effect of magnesium on soil erosion and water infiltration. Soil Science Society of America Journal, 55, 783787.CrossRefGoogle Scholar
Lagaly, G. (1981). Characterization of clays by organic compounds. Clay Minerals, 16, 121.CrossRefGoogle Scholar
Levy, G. J., & Van Der Watt, H. V. H. (1990). Effect of exchangeable potassium on the hydraulic conductivity and infiltration rate of some South African soils. Soil Science, 149, 6977.CrossRefGoogle Scholar
Liu, X., Li, H., Li, R., Xie, D., Ni, J., & Wu, L. (2014). Strong non-classical induction forces in ion-surface interactions: General origin of Hofmeister effects. Scientific Reports, 4, 5047.CrossRefGoogle ScholarPubMed
Manciu, M., & Ruckenstein, E. (2003). Specific ion effects via ion hydration: I. Surface tension. Advances in Colloid and Interface Science, 105, 63101.CrossRefGoogle ScholarPubMed
Marchuk, A., & Rengasamy, P. (2011). Clay behaviour in suspension is related to the ionicity of clay–cation bonds. Applied Clay Science, 53, 754759.CrossRefGoogle Scholar
Marchuk, A., Rengasamy, P., & McNeill, A. (2013a). Influence of organic matter, clay mineralogy, and pH onthe effects of CROSS on soil structure is related to the zeta potential of the dispersed clay. Soil Research, 51, 3440.CrossRefGoogle Scholar
Marchuk, A., Rengasamy, P., McNeill, A., & Kumar, A. (2013b). Nature of the clay–cation bond affects soil structure as verified by X-ray computed tomography. Soil Research, 50, 638644.CrossRefGoogle Scholar
Misono, M., Ochiai, E. I., Saito, Y., & Yoneda, Y. (1967). A new dual parameter scale for the strength of Lewis acids and bases with the evaluation of their softness. Journal of Inorganic and Nuclear Chemistry, 29, 26852691.CrossRefGoogle Scholar
Mitchell, J. K., & Soga, K. (2005). Fundamentals of Soil Behavior. New York: John Wiley & Sons.Google Scholar
Molina, F. V. (2014). Soil Colloids: Properties and Ion Binding. Boca Raton, Florida, USA: CRC Press.Google Scholar
Oster, J. D., Shainberg, I., & Wood, J. D. (1980). Flocculation value and gel structure of sodium/calcium montmorillonite and illite suspensions. Soil Science Society of America Journal, 44, 955959.CrossRefGoogle Scholar
Parameswaran, T., & Sivapullaiah, P. (2017). Influence of sodium and lithium monovalent cations on dispersivity of clay soil. Journal of Materials in Civil Engineering, 29, 04017042.CrossRefGoogle Scholar
Pashley, R. M., & Quirk, J. P. (1984). The effect of cation valency on DLVO and hydration forces between macroscopic sheets of muscovite mica in relation to clay swelling. Colloids and Surfaces, 9, 117.CrossRefGoogle Scholar
Pauling, L. (1960). The Nature of the Chemical Bond and the Structure of Molecules and Crystals: An Introduction to Modern Structural Chemistry. Ithaca, New York: Cornell University Press.Google Scholar
Quirk, J. P. (2001). The significance of the threshold and turbidity concentrations in relation to sodicity and microstructure. Soil Research, 39, 11851217.CrossRefGoogle Scholar
Quirk, J. P., & Marcelja, S. (1997). Application of double-layer theories to the extensive crystalline swelling of Li-montmorillonite. Langmuir, 13, 62416248.CrossRefGoogle Scholar
Rayment, G. E., & Lyons, D. J. (2011). Soil Chemical Methods: Australasia. Collingwood, Australia: CSIRO publishing.Google Scholar
Rengasamy, P. (1983). Clay dispersion in relation to changes in the electrolyte composition of dialysed red-brown earths. Journal of Soil Science, 34, 723732.CrossRefGoogle Scholar
Rengasamy, P., & Marchuk, A. (2011). Cation ratio of soil structural stability (CROSS). Soil Research, 49, 280285.CrossRefGoogle Scholar
Rengasamy, P., & Olsson, K. A. (1991). Sodicity and soil structure. Soil Research, 29, 935952.CrossRefGoogle Scholar
Rengasamy, P., Greene, R. S. B., Ford, G. W., & Mehanni, A. H. (1984). Identification of dispersive behaviour and the management of red-brown earths. Soil Research, 22, 413431.CrossRefGoogle Scholar
Rengasamy, P., Tavakkoli, E., & McDonald, G. K. (2016). Exchangeable cations and clay dispersion: net dispersive charge, a new concept for dispersive soil. European Journal of Soil Science.CrossRefGoogle Scholar
Saka, E. E., & Güler, C. (2006). The effects of electrolyte concentration, ion species and pH on the zeta potential and electrokinetic charge density of montmorillonite. Clay Minerals, 41, 853861.CrossRefGoogle Scholar
Semerjian, L., & Ayoub, G. M. (2003). High-pH–magnesium coagulation–flocculation in wastewater treatment. Advances in Environmental Research, 7, 389403.CrossRefGoogle Scholar
Shainberg, I., & Levy, G. J. (2005). Flocculation and dispersion. Pp. 2734 in: Encyclopedia of Soils in the Environment. Oxford: Elsevier.CrossRefGoogle Scholar
Smiles, D. (2006). Sodium and potassium in soils of the Murray–Darling Basin: a note. Soil Research, 44, 727730.CrossRefGoogle Scholar
Smith, C. J., Oster, J. D., & Sposito, G. (2015). Potassium and magnesium in irrigation water quality assessment. Agricultural Water Management, 157, 5964.CrossRefGoogle Scholar
Sposito, G. (1994). Chemical Equilibria and Kinetics in Soils. New York: Oxford University Press.CrossRefGoogle Scholar
Sposito, G. (2008). The Chemistry of Soils. New York: Oxford University Press.Google Scholar
Steudel, A., & Emmerich, K. (2013). Strategies for the successful preparation of homoionic smectites. Applied Clay Science, 75, 1321.CrossRefGoogle Scholar
Tisdall, J. M., & Oades, J. M. (1982). Organic matter and water-stable aggregates in soils. Journal of Soil Science, 33, 141163.CrossRefGoogle Scholar
Tombácz, E., & Szekeres, M. (2004). Colloidal behavior of aqueous montmorillonite suspensions: the specific role of pH in the presence of indifferent electrolytes. Applied Clay Science, 27, 7594.CrossRefGoogle Scholar
US Salinity Laboratory Staff (1954). Diagnosis and improvement of saline and alkali soils. Pp. 154. United States Department of Agriculture, Agriculture Handbook No. 60, Washington DC.Google Scholar
Verwey, E.J.W. & Overbeek, J.T.G. (1948). Theory of the Stability of Lyophobic Colloids. Elsevier Publishing Company.Google Scholar
Walkley, A., & Black, I. A. (1934). An examination of the Degtjareff method for determining soil organic matter, and a proposed modification of the chromic acid titration method. Soil Science, 37, 2938.CrossRefGoogle Scholar
Weil, R. R., & Brady, N. C. (2016). The Nature and Properties of Soils (15th ed.). Harlow, United Kingdom: Pearson Education Limited.Google Scholar
Williams, D. J. A., & Williams, K. P. (1978). Electrophoresis and zeta potential of kaolinite. Journal of Colloid and Interface Science, 65, 7987.CrossRefGoogle Scholar
Zhang, X. C., & Norton, L. D. (2002). Effect of exchangeable Mg on saturated hydraulic conductivity, disaggregation and clay dispersion of disturbed soils. Journal of Hydrology, 260, 194205.CrossRefGoogle Scholar
Zhou, Z., & Gunter, W. D. (1992). The nature of the surface charge of kaolinite. Clays and Clay Minerals, 40, 365368.CrossRefGoogle Scholar
Zhu, Y., Marchuk, A., & Bennett, J. M. (2016). Rapid method for assessment of soil structural stability by turbidimeter. Soil Science Society of America Journal, 80, 16291637.CrossRefGoogle Scholar
Zhu, Y., Bennett, J. M., & Marchuk, A. (2019). Reduction of hydraulic conductivity and loss of organic carbon in non-dispersive soils of different clay mineralogy is related to magnesium induced disaggregation. Geoderma, 349, 110.CrossRefGoogle Scholar