Association of dissolved aluminum with silica: Connecting molecular structure to surface reactivity using NMR

https://doi.org/10.1016/j.gca.2008.04.028Get rights and content

Abstract

We studied uptake mechanisms for dissolved Al on amorphous silica by combining bulk-solution chemistry experiments with solid-state Nuclear Magnetic Resonance techniques (27Al magic-angle spinning (MAS) NMR, 27Al{1H} cross-polarization (CP) MAS NMR and 29Si{1H} CP-MAS NMR). We find that reaction of Al (1 mM) with amorphous silica consists of at least three reaction pathways; (1) adsorption of Al to surface silanol sites, (2) surface-enhanced precipitation of an aluminum hydroxide, and (3) bulk precipitation of an aluminosilicate phase. From the NMR speciation and water chemistry data, we calculate that 0.20 (±0.04) tetrahedral Al atoms nm−2 sorb to the silica surface. Once the surface has sorbed roughly half of the total dissolved Al (∼8% site coverage), aluminum hydroxides and aluminosilicates precipitate from solution. These precipitation reactions are dependent upon solution pH and total dissolved silica concentration. We find that the Si:Al stoichiometry of the aluminosilicate precipitate is roughly 1:1 and suggest a chemical formula of NaAlSiO4 in which Na+ acts as the charge compensating cation. For the adsorption of Al, we propose a surface-controlled reaction mechanism where Al sorbs as an inner-sphere coordination complex at the silica surface. Analogous to the hydrolysis of Al(OH2)63+, we suggest that rapid deprotonation by surface hydroxyls followed by dehydration of ligated waters results in four-coordinate (>SiOH)2Al(OH)2 sites at the surface of amorphous silica.

Introduction

Aluminum plays a key role in mineral reactivity in a wide range of environments that include the formation of hydrous aluminosilicates from natural waters, preservation of deep sea sediments, and precipitation of geothermal scales. Doucet et al. (2001) found that hydroxyaluminosilicate (HAS) phases form by competitive condensation of silica on a hydroxy-aluminum template and that these HAS phases have similar 27Al MAS NMR spectra to HAS found in soil horizons (Barron et al., 1982). The structural similarities between precipitated HAS and those found in soils suggest that the precipitation of HAS may regulate the bioavailability and thus toxicity of aqueous Al in natural waters (Doucet et al., 2001). As another example, Van Cappellen et al. (2002) examined the solubility of diatom cultures and siliceous sediments containing different amounts of Al and found that bulk solubilities typically decreased by an order of magnitude when the Al/Si increased by a factor of two. It was shown using X-ray Absorption Near Edge Spectroscopy (XANES) that Al in diatoms had been incorporated into 4-fold coordination (Gehlen et al., 2002). This suggests that incorporation of tetrahedral Al into the diatom framework during synthesis is most likely the reason that dissolution rates of biogenic opal in deep-sea sediments are reduced by over an order of magnitude relative to surface waters. Lastly, Carroll et al. (1998) measured silica precipitation rates for amorphous silica in complex geothermal brines at the Wairakei geothermal energy plant (New Zealand) and found that rates were much faster than in laboratory studies (pH 7–8, reaction time 48 h). They suggested that the accelerated silica precipitation rates found in the field were due to precipitation of Al-containing phases. Gallup (1997) reported an 27Al MAS NMR spectrum for an Al-rich silica scale from Awibengkok geothermal field which showed only tetrahedrally coordinated Al (27Al MAS NMR, [4]Al = 52.5 ppm; uncorrected for second-order quadrupolar shift), most likely due to aluminosilicate precipitation in which the Al had been supplied by the geothermal brine.

These examples illustrate the importance of understanding the connection between structure and reactivity. Due to the complex structural arrangement of natural systems, however, spectroscopic studies are often difficult to perform and interpret. Much of our understanding of metal cation sorption and precipitation comes from changes in bulk, aqueous measurements. While extremely useful, bulk-chemistry studies lack molecular-scale information which allows for the identification of reaction mechanisms that control solubility.

The motivation behind our study is to explain silicate surface reactivity in terms of molecular structure using NMR spectroscopy. We use NMR to identify the structural form of Al associated with amorphous silica over a pH range similar to that of natural waters. Because amorphous silicas are often used as a proxy for natural silicas, we anticipate the surface chemistry to be analogous to that of natural phases. The speciation information we obtain from NMR combined with the water chemistry data allow us to identify three reaction mechanisms that describe the solubility of Al and surface reactivity of amorphous silica.

Section snippets

Sample preparation

Amorphous silica was chosen because it serves as a high-surface area proxy for naturally occurring silica. Silica gel (Mallinckrodt silicar: 306 m2/g surface area by BET, 75–100 μm particle size) was washed three times ultrasonically with distilled water and then dried at 50 °C overnight. Batch-sorption experiments were performed by equilibrating the silica gel in a background electrolyte of 0.1 M NaCl at 25 °C for ∼22 h with constant stirring (solid/solution = 5 g/L). The pH was adjusted with 100 mM

Silica gel dissolution

The pH and time dependence of silica gel dissolution kinetics were first investigated in the absence of Al to understand how dissolved Al affects the structure and reactivity of amorphous silica. Fig. 1 shows the change in aqueous silica over time for gel suspended in a 0.1 M NaCl solution at pH 5.4. Initially, dissolution kinetics are fast but decrease at longer reaction times as silica levels approach constant concentration (>3 days). The data show that during t = 22–24 h, which is the reaction

Discussion

Reaction of dissolved Al with amorphous silica consists of at least three reactions. These reactions include sorption of [4]Al as an inner-sphere complex to silanol sites,2>SiOH+Al(OH2)63+(>SiOH)2[Al(OH)2]++2H++4H2Osurface-enhanced precipitation of an aluminum hydroxide,>SiOH+Al3++3H2O>SiOH[Al(OH)3]+3H+and bulk precipitation of an aluminosilicate phase.SiO2(aq)+Al3++Na++2H2ONaAlSiO4+4H+The pH-dependence of all three reactions is shown in Fig. 11 and mass balance calculations are shown in

Acknowledgments

We thank Kevin Knauss, Sarah Roberts, Bill Bourcier, and Carol Bruton for helpful discussions throughout the course of this study. We also thank three anonymous reviewers for their insightful comments. This work was funded by the Department of Energy, Office of Basic Energy Science and performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract W-7405-Eng-48 and Contract DE-AC52-07NA27344.

References (40)

  • Y. Kim et al.

    23Na and 133Cs NMR study of cation adsorption on mineral surfaces: local environments, dynamics, and effects of mixed cations

    Geochim. Cosmochim. Acta

    (1997)
  • W.H. Kuan et al.

    Modeling and electrokinetic evidences on the processes of the Al(III) sorption continuum in SiO2(s) suspension

    J. Colloid Interface Sci.

    (2004)
  • P.A. O’Day et al.

    X-ray absorption spectroscopy of cobalt(II) multinuclear surface complexes and surface precipitates on kaolinite

    J. Colloid Interface Sci.

    (1994)
  • A.M. Scheidegger et al.

    Spectroscopic evidence for the formation of mixed-cation hydroxide phases upon metal sorption on clays and Aluminum oxides

    J. Colloid Interface Sci.

    (1997)
  • L.L. Stillings et al.

    Feldspar dissolution at 25 °C and pH 3. Reaction stoichiometry and the effect of cations

    Geochim. Cosmochim. Acta

    (1995)
  • T. Yokoyama et al.

    The effect of aluminum on the biodeposition of silica in hot spring water: chemical state of aluminum in siliceous deposits collected along the hot spring water stream of Steep Cone hot spring in Yellowstone National Park, USA

    Chem. Geol.

    (2004)
  • T. Yokoyama et al.

    A study of the alumina-silica gel adsorbent for the removal of silicic acid from geothermal water: increase in adsorption capacity of the adsorbent due to formation of amorphous aluminosilicate by adsorption of silicic acid

    J. Colloid Interface Sci.

    (2002)
  • C.F. Baes et al.

    The Hydrolysis of Cations

    (1976)
  • P.F. Barron et al.

    Detection of imogolite in soils using solid state 29Si NMR

    Nature

    (1982)
  • C.M. Bethke

    The Geochemist’s Workbench

    (1994)
  • Cited by (39)

    • Formation and mitigation of mineral scaling in geothermal power plants

      2022, Water-Formed Deposits: Fundamentals and Mitigation Strategies
    View all citing articles on Scopus

    This document was prepared as an account of work sponsored by an agency of the United States government. Neither the United States government nor Lawrence Livermore National Security, LLC, nor any of their employees makes any warranty, expressed or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States government or Lawrence Livermore National Security, LLC. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States government or Lawrence Livermore National Security, LLC, and shall not be used for advertising or product endorsement purposes.

    View full text