Abstract
Density functional theory calculations were performed on an extended silicate molecular cluster to test the hypothesis that inclusion of Cl− to charge-balance protonation of bridging O atoms in siloxane bridges, and H2O attack via an SN2 mechanism, could explain the observed activation energy of quartz dissolution under acidic conditions. The results are consistent with this hypothesis and help explain the observed solution pH and salt-dependence of quartz dissolution rates.
Subject terms: Environmental chemistry, Environmental impact, Theory and computation
Introduction
Dissolution of silica and silicates is a topic that relates to biogeochemical cycles1,2, soil productivity3,4, corrosion of man-made materials5, and geothermal energy extraction6–10. Consequently, myriad field, laboratory, and simulation studies have been published on the rates and mechanisms of silica and silicate dissolution. However, field-laboratory11–14, laboratory15–20, field-laboratory-simulation studies21–23, and simulation studies24–41 result in significant discrepancies in rates.
There is no consensus molecular mechanism for quartz dissolution. The current scientific hypothesis is that hydrolysis of Si-Obr-Si linkages (Obr for “bridging O atom” hereafter) is the rate-limiting step. No other reaction has been posited that can transform SiO4-tetrahedra with 4 Si- Obr-Si linkages in the solid into the aqueous species Si(OH)4. This study focuses on modeling the Si- Obr-Si hydrolysis reaction with the goal of determining a reaction pathway consistent with observed activation energies (ΔEa) of silica dissolution. With such a reaction pathway determined for realistic models of surface SiO4-tetrahedra, we posit that it will become possible to predict rates of hydrolysis and dissolution over a range of temperatures and solution compositions for use in weathering where rates are slow and subsurface reactions at higher T and ionic strength solutions.
Field and laboratory experiments related to silica weathering or dissolution are too numerous to cover so we reference a recent review that covers the literature on this topic23. We cite three key papers that provide reliable ΔEa values, however, as these observables are the main targets of our simulation efforts. Dove (1999) performed dissolution experiments on quartz at pH ≈5.7 over a temperature range of 175 to 295 °C in distilled water and with chloride salts at 0.04 to 0.15 M to derive ΔEa values of 71 to 75 kJ·mol− 1.14 Icenhower and Dove (2000) studied fused silica dissolution rates from 25 to 250 °C in distilled and 0.05 M NaCl solutions and determined ΔEa to be 74.5 ± 1.4 kJ·mol− 1.20 Crundwell (2017) analyzed quartz dissolution data and concluded the ΔEa = 70 kJ·mol− 1.22 Because simulations do not represent all potential steps in the dissolution process, we consider these values to be upper limits for the acceptable ΔEa values determined in our DFT calculations.
Quantum mechanical studies of Si-O-Si hydrolysis and silica dissolution began with Casey et al. (1990) who determined that H+-transfer to the Obr was not the rate-limiting step because modeled H-D kinetic isotopic effects were significantly different in experiment and calculation30. These types of studies continued for three decades, but the results were either based on models that did not represent silica surface chemistry (e.g., the molecule H3SiOSiH3) or resulted in activation energies significantly higher than observation (e.g., Criscenti et al., 2006, where ΔEa = 112 kJ·mol− 1).32 More recent studies have used machine-learning methods to derive ΔEavalues from experimental data21, and kinetic Monte Carlo techniques based on observed ΔEa values to predict behavior consistent (i.e., etch pit formation) with experimental dissolution rates40. Periodic DFT simulations have focused more on the quartz-water interface as a function of solution composition25,27,28,41, and these studies have indicated changes in the intrasurface H-bonding (i.e., H-bonding between SiOH and Obr sites) that could explain the observed salt-dependence of silica dissolution rates14.
Periodic DFT simulations have revealed that association of the Cl− with the quartz surface is probable and that this near approach to the surface can help lower the calculated energy of a H+ attached to an Obr. Previous work had not examined this potential ion-pairing of Cl− and H+ at the silica surface, but lower dielectric constants at mineral-water interfaces are likely to help stabilize such ion pairing42. This paper focuses on such ion-paired structures and the subsequent hydrolysis reaction pathway. In addition, simulations predicted that rather than direct attack of H2O from solution to form a 5-coordinate surface Si ([5]Si), H2O could diffuse just below the silica surface and form the [5]Si from an SN2 mechanism. The hypothesized reaction mechanism is shown in Fig. 1. These two variations from previous modeling studies result in a significant lowering of the calculated ΔEa which is more consistent with observed values.
Fig. 1.
Hypothesized reaction mechanism. From left to right, a lone pair of electrons from the bridging O bonds to an H+, and a lone pair of electrons from H2O bonds beneath a Q1 Si to form the middle structure that will lead to a transitions state. Finally, a lone pair of electrons from the bridging O−[5]Si bond in the middle structure moves to the bridging O to form the products in the right structure. The bonds on the left-most Si in each model show the generic bonds of a Si to the remainder of the silica model.
Results and discussion
Figure 2 shows the quartz cluster model used for this work. The model was chosen to represent a general silicate that exhibited a Q1 Si, where we hypothesize that dissolution can occur with hydration of that Si from below and breaking of the Si-Obr bond, while a Cl− stabilizes the H+ on Obr. After energy minimizing the model, the Si-O bond distances were 1.64 ± 0.02 Å and the Si-O-Si bond angles were 134.4 ± 7.3º, which agree with experimental data for those measurements of 1.61 to 1.62 Å for Si-O bond lengths in α-quartz43, and 144º for Si-O-Si angles in α-quartz44. Our Si-O-Si bond angle average is less than that for α-quartz; however, in lower-pressure crystalline silicate minerals and glasses, the Si-O-Si angle can range from 134 to 150º, which would include our calculated results45.
Fig. 2.

The labeled and enlarged spheres in this model show the areas of interest discussed in this work. The model above shows a pentavalent Q1 Si bonded to three OH−, one H2O, and a Obr between the Q1 Si of interest and a Q4 Si. A Cl− is adjacent to the bridging H that is bonded to the Obr. Atom colors are white for H, red for O, yellow for Si, and green for Cl.
Figure 3 shows the relative electronic energy (E) compared with the lowest energy model configuration (Fig. 3a), the change in the distance between Obr and Hbr (Fig. 3b), and the distance between the Hbr on the Obr and Cl− (Fig. 3c), versus the distance between Obr and the Q1 Si. The Si(OH)3(H2O) – Obr distance was held constant during the energy minimization calculations, except for the first (1.76 Å) and last (3.15 Å) calculations, which were minimized with no constraints.
Fig. 3.

Relative electronic energy (E) (a), Obr-Hbr bond length (b), and the Hbr-Cl− distance as the bond between O(H)br and Si(OH)3(OH2) of the model increases and breaks (c).
The plot in Fig. 3a shows the trend in the energy change relative to the final model as the Si(OH)3(H2O) – Obr bond distance increases, breaks, and dissociates. The initial model had an energy that was 14 kJ·mol− 1 higher than the lowest energy (final) model. The model with the Si(OH)3(H2O) – Obr distance of 2.21 Å had a relative energy of 56 kJ·mol− 1 higher than the energy of the final model (at 0 kJ·mol− 1); the higher energy model also had one imaginary (negative) vibrational frequency of −66 cm− 1 for the Si(OH)3(H2O) – Obr vibrational mode, making it the transition state for the reaction and provide the ΔEa. The ΔEaresult is consistent with the experimental results from the literature that we cited in the introduction14,20,22 and does not exceed those results dramatically like prior calculated results from the literature.
In addition, a recent paper fit data obtained from 285 quartz and silica dissolution experiments and found an ΔEa of 70 kJ·mol− 1,22 which agrees well with our result, suggesting that dissolution mechanism presented here could be prominent. We posit that the flexibility of the reaction is greater in model clusters compared to real surfaces where lattice constraints could increase the ΔEa.
Modeling periodic quartz surfaces using the proposed reaction mechanism would be a logical step to test this hypothesis; however, a prior study showed that the cluster structures such as those used here can give similar results to condensed phases used in periodic model calculations46. Using a small molecular model, (OH)3Si-O-Si(OH)3, Gibbs’ et al., (2006) calculated bond lengths, bond angles, bond critical points, local potentials, and kinetic energy densities that were similar to quartz46. (OH)3Si-O-Si(OH)3 and quartz also showed close similarity of electron density distribution between their model and quartz and showed that Si-O bond energies were similar between (OH)3Si-O-Si(OH)3 (462 kJ·mol− 1) and quartz (465 kJ·mol− 1)46. Therefore, we agree with those authors’ assertion that these similarities justify the use of small molecular models to act as proxies for quartz polymorphs.
We also ran single-point energy calculations using implicit solvation to test the effect on the calculated activation and reaction energies on the initial, transition state, and final structures. The results show a −5 kJ/mol lower ΔEa(51 vs. 56 kJ/mol) and the reaction energy difference is +5 (19 vs. 14 kJ/mol). We consider these differences to be insignificant as we do not claim accuracy of better than ± 10 kJ/mol in our calculations. Replicate DFT calculations would give 0 kJ/mol differences, but those results could differ from experiment by as much as 10 kJ/mol. Furthermore, the implicit solvation calculations were performed with a permittivity (ε) of 78.4 for bulk water, but the dielectric constant at the interface is likely to be significantly lower23, so the effect of long-range solvation on reaction energetics would be less.
Figure 3b shows that as the Si(OH)3(H2O) – Obr distance increases, the Obr-Hbr bond length decreases, making the latter bond stronger and more stable. The Obr-Hbr bond length decreased by nearly 0.2 Å due to Si(OH)3(H2O) dissolution.
As the Obr-Hbr bond length decreases, the distance between the Hbr and Cl− increases (Fig. 3c) by nearly 0.3 Å as the Si(OH)3(H2O) group dissociates. This result suggests that Cl− is shielding the Hbr from attack by H2O to form H3O+, as well as stabilizing and lowering the energy of the cluster. As the Si(OH)3(H2O) dissociates and the Obr-Hbr distance decreases, the Cl− moves away from the Hbr.
To evaluate this potential phenomenon, we also calculated the effect of Cl− proximity to the Hbr, using the two models shown in Fig. 4. We compared that model in Fig. 4a, which is the same model shown in Fig. 2, with the model shown in Fig. 4b, where we moved the Cl− away from the Hbr, so that no interaction could occur between the Hbr and Cl−. After energy minimizing the structure in Fig. 4b, the E for that model was 196 kJ·mol− 1 higher in free energy than the minimized model in Fig. 4a. The distance between the H+ on the Obr and Cl- in the model in Fig. 4a is 1.78 Å and 3.11 Å in Fig. 4b model. Therefore, Cl− interacting with the proton on the Obr is lowering the model energy by stabilizing the proton on the Obr. Coupled with the results showing that Cl− moves away from Hbr as the Si(OH)3(H2O) group dissociates, the results suggest that Cl- stabilizes the H+ on Obr and is necessary for our proposed mechanism.
Fig. 4.

(a) is the same the model in Fig. 1, whereas in (b), the Cl− was moved far from the protonated ObrHbr. The hydrating H2O molecules were also moved away from ObrHbr, because without the Cl− near the Hbr, H2O deprotonates ObrHbr to form H3O+.
The calculated rate constants, kr, shown in Table 1 were obtained using Eq. (1) and are compared with a select calculated result33. We calculated the rate constants at 298.15 K and 1.01 bar and at 468.15 K and 206 bar; the latter were to make the result pertinent to well UO_FU_5832_PT from U.S. Department of Energy (US DOE) Forge (Frontier Observatory for Research in Geothermal Energy) from a sampling depth of 7450 m. Our kr result at 298.15 K suggests that the reaction we hypothesize would proceed 1000 times slower than that using the calculation from the literature33.; that work used a model that was smaller and charged (Si2O8H9+), whereas our model was larger and neutral (Si8O31H31Cl). These differences, coupled with the lack of Cl− counter ion for charge balance and stability in the other work, could account for the differences in model results.
Table 1.
Rate constant, kr, results from this work and calculations performed in reference 44*. The Q results for the reactant model (Q) and transition state model (Q‡) are from the Gaussian 16 log file results at the temperatures shown. Reference 44 did not report Q results.
| Temperature (K) | kr (s− 1) | Q | Q‡ |
|---|---|---|---|
| 298.15 | 9.2 × 105 | 1.98 × 1043 | 1.86 × 1046 |
| 468.15 | 1.2 × 1010 | 1.30 × 1055 | 2.61 × 1058 |
| 298* | 8.6 × 108* |
There is a large discrepancy between our results and model predictions reported in the literature (Table 1)33; our proposed mechanism where H2O attacks a Q1 Si by SN2 attack coupled with Cl− stabilizing a proton on the adjacent Obr is different that prior mechanism, so we expect our results to differ from those of previous work that used a H6Si2O7 + H3O+ model. Furthermore, the previous work assumed the protonated Obr as the reactant when this state is significantly higher in energy than the non-protonated state. Good agreement between our ΔEa results and experimental data14,22 suggest that our results are more realistic.
Using our calculated rate constants shown in Table 1, a surface site density of 8 Obr per nm2, and Eq. 2, the rates were calculated at 298.15 and 498.15 K (Table 2). As discussed in the methods section, the % Obr as ObrH site density was based on XPS data47.
Table 2.
Rate of dissolution reaction as a function of H+ site density percentage (pH).
| pH | % Obr as ObrH | T 298 K | T 468 K |
|---|---|---|---|
| Rate (Obr sites·nm− 2·s− 1)) | Rate (Obr sites·nm− 2·s− 1)) | ||
| 1 | 7.29 | 5.35 × 105 | 6.75 × 109 |
| 2 | 5.14 | 3.78 × 105 | 4.77 × 109 |
| 3 | 3.93 | 2.88 × 105 | 3.64 × 109 |
| 4 | 3.21 | 2.36 × 105 | 2.98 × 109 |
| 5 | 2.86 | 2.10 × 105 | 2.65 × 109 |
| 6 | 2.14 | 1.57 × 105 | 1.99 × 109 |
The results in Table 2 show that as the % Obr as ObrH decreased (pH increase), the rate of the reaction decreased at both temperatures, as expected for a mechanism that would occur under acidic conditions. The rates are 468.15 K (i.e., geothermal conditions) are much greater than the rates at 298.15 K.
Conclusion
This work has shown that silicate dissolution under acidic conditions could occur at Q1 Si atoms by SN2 attack, if H2O attacks from beneath the Si atom to form a [5]Si. Simultaneously, the Obr adjacent to the Q1 Si would be protonated under acidic conditions, and the H+ on the Obr would be stabilized by an anion, such as Cl−. The calculated energy of activation of 56 kJ·moL− 1 for this work agrees more accurately with data from the literature than prior computational study results have. We reported rate constant and rate results at 298.15 and 498.15 K for the calculated Ea result, which could be used to guide experiments and to determine if our proposed mechanism is occurring. Our results show that as pH increases, the rate of dissolution by the proposed SN2 mechanism decreases at both temperatures, as would be expected for a H+-driven mechanism.
Methods
Models were built using Materials Studio 2016 (Biovia, Inc.; San Diego, CA - https://www.3ds.com/products/biovia/materials-studio) and energy minimized using Gaussian 16 (ES64L-G16RevB.01 http://www.gaussian.com48. The model stoichiometry was Si8O31 H31Cl. The initial structure for the cluster was extracted from an unpublished periodic DFT simulation. We added 5 H2O molecules to our model to hydrate and stabilize the Cl− that was added to stabilize a proton that was added to one of the bridging O atoms (Obr). Although the small silica model that we used might not be able to mimic the chemistry of quartz, earlier authors found that small models can capture quartz chemistry accurately46. Our goal for this work is to explore a reaction mechanism using a silica model where H2O attacks a Q1 Si from below and breaks away from a protonated Obr to form Si(OH)3(H2O). If activation energy results from this work match those from experiment, this method could then be applied to larger silica models and periodic models that better model quartz.
Energy minimization calculations were performed using density functional theory (DFT)49,50 with the M06-2X functional51. We chose the M06-2X functional due to its ability predict benchmark level (CCSD/6–31 + G(d)) geometries for the transition states structures of Diels Alder reactions52. The 6-311G(d, p) basis set was used for all calculations53. For the energy minimization calculations, a fine CPHF grid49, and ultrafine integration grid48 were used. The models were first geometry optimized then subjected to frequency calculation to confirm that each model had reached a potential energy minimum and that each model exhibited only real (> 0 cm− 1) vibrational frequencies.
Single-point energy calculations were run with the SMD implicit solvation model on the reactants55, transition state and product structures to evaluate the potential effect of long-range solvation on the energies. We note that a permittivity (ε) of 78.4 for bulk water was used although the value of ε is likely to be significantly lower at the quartz-water interface. Hence, these results will represent the maximum potential impact of long-range solvation on our results.
For the stepwise dissolution calculations, the minimized structure in Fig. 1 was modified by incrementally increasing the distance between the Q1 Si and the Obr in that model in ca. 0.1 Å steps until the Si(OH)3(OH2) group was dissociated from the cluster. For each of the stepwise calculations, the Si-Obr distance was held constant by freezing those atoms and the distance between them using the “−1” method in Gaussian 16; all other atoms were minimized without constraints. The end member models of the stepwise calculations were energy minimized without constraining the distance between the Obr and the Q1 Si. The transition state model underwent a frequency calculation without constraints. This model was used to find the energy of activation (Ea) of the dissolution and had one and only one imaginary vibrational frequency.
For the stepwise calculations, we report electronic energies (E) that were relative to the E of the model having the fully dissociated Si(OH)3(OH2); that is, the E of that model was set to 0 kJ·mol− 1 and E of the other stepwise dissociation models were calculated relative to that model.
Rate constants were calculated using Eq. 156:
![]() |
1 |
Here, kr is the rate constant, kB is the Boltzmann constant, h is the Planck constant, T is the temperature, and ΔEa is the change in energy between the minimized reactant (initial) structure and the minimized transition state structure. Q‡ and Q are the v = 0 total partition functions of the transition state and reactant structures, respectively. obtained from Gaussian 16.
Reaction rates were used by using the rate law in Eq. (2):
![]() |
2 |
Here, kr is the rate constant, Obr site density is estimated as 8 Obr sites per nm2from the quartz (101) model upon which the model in this work was based25,41, and % Obr as ObrH was estimated using a plot of SiOH2+ surface site density percentage versus pH based on X-ray photoelectron spectroscopy (XPS) data47. Previous work showed that the proton affinity of SiOH2+ is approximately equal to that of ObrH surface density27,56, so we halved the site density present on the plot from the literature (based on SiOH2+) to obtain percentage of H+ site density for Eq. 2.
The bond distance for Obr-Hbr and the distance between Hbr and Cl− were obtained from each of the stepwise, energy-minimized models using the bond length using Materials Studio 2016.
Acknowledgements
This work was funded by the U.S. Department of Energy, Office of Science, Energy Earthshots Initiatives under Award Number DE-SC0024703. Computational support provided by the Research Academic Data Center at the University of Texas at El Paso and by National Energy Research Scientific Computing (NERSC) U.S. Department of Energy.
Author contributions
JDK and HDW designed the study. HDW performed the calculations and wrote the first draft of the manuscript Methods and Results and Discussion sections. JDK wrote the first draft of the Introduction. JDK and HDW edited the manuscript.
Funding
This work was funded by the U.S. Department of Energy, Office of Science, Energy Earthshots Initiatives under Award Number DE-SC0024703.
Data availability
Data is provided within the manuscript.
Declarations
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- 1.Tréguer, P. J. et al. Reviews and syntheses: The biogeochemical cycle of silicon in the modern ocean. Biogeosciences vol. 18 1269–1289 Preprint at (2021). 10.5194/bg-18-1269-2021
- 2.Sutton, J. N. et al. A review of the stable isotope bio-geochemistry of the global silicon cycle and its associated trace elements. Frontiers in Earth Science vol. 5 Preprint at (2018). 10.3389/feart.2017.00112
- 3.Swoboda, P., Döring, T. F. & Hamer, M. Remineralizing soils? The agricultural usage of silicate rock powders: A review. Sci. Total Environ. Vol. 10.1016/j.scitotenv.2021.150976 (2022). 807 150976 Preprint at. [DOI] [PubMed] [Google Scholar]
- 4.Schaller, J., Puppe, D., Kaczorek, D., Ellerbrock, R. & Sommer, M. Silicon cycling in soils revisited. Plants vol. 10 1–36 Preprint at (2021). 10.3390/plants10020295 [DOI] [PMC free article] [PubMed]
- 5.Golovchak, R. et al. Remedial insight on ageing of glass through the study of ancient man-made artefacts. Archaeometry63, 312–326 (2021). [Google Scholar]
- 6.Spitzmüller, L. et al. Selective silica removal in geothermal fluids: implications for applications for geothermal power plant operation and mineral extraction. Geothermics95, 102141 (2021). [Google Scholar]
- 7.Kamaratou, M., Spinthaki, A., Disci, D. & Demadis, K. D. Ferric silicate precipitates relevant to geothermal systems: delineation of their complex formation. Geothermics123, 103118 (2024). [Google Scholar]
- 8.Spinthaki, A., Petratos, G., Matheis, J., Hater, W. & Demadis, K. D. The precipitation of magnesium silicate under geothermal stresses. Formation Charact. Geothermics. 74, 172–180 (2018). [Google Scholar]
- 9.Liu, S. & Ott, W. K. Sodium silicate applications in oil, gas & geothermal well operations. Journal of Petroleum Science and Engineering vol. 195 107693 Preprint at (2020). 10.1016/j.petrol.2020.107693
- 10.den Heuvel, D. B. et al. Understanding amorphous silica scaling under well-constrained conditions inside geothermal pipelines. Geothermics76, 231–241 (2018). [Google Scholar]
- 11.Dove, P. M. Kinetic and thermodynamic controls on silica in weathering environments. In Reviews in Mineralogy, A.F. White and S.L. Brantley (editors) vol. 31 235–290 (Mineralogical Society of America, 1995).
- 12.Rimstidt, J. D. & Barnes, H. L. The kinetics of silica-water reactions. Geochim. Cosmochim. Acta. 44, 1683–1699 (1980). [Google Scholar]
- 13.Zheng, X. Y., Beard, B. L. & Johnson, C. M. Constraining silicon isotope exchange kinetics and fractionation between aqueous and amorphous Si at room temperature. Geochim. Cosmochim. Acta. 253, 267–289 (2019). [Google Scholar]
- 14.Dove, P. M. The dissolution kinetics of quartz in aqueous mixed cation solutions. Geochim. Cosmochim. Acta. 63, 3715–3727 (1999). [Google Scholar]
- 15.Hamilton, J. P., Pantano, C. G. & Brantley, S. L. Dissolution of albite glass and crystal. Geochim. Cosmochim. Acta. 64, 2603–2615 (2000). [Google Scholar]
- 16.Dove, P. M., Han, N., Wallace, A. F. & De Yoreo, J. J. Kinetics of amorphous silica dissolution and the paradox of the silica polymorphs. Proc. Natl. Acad. Sci. U S A. 105, 9903–9908 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Dove, P. M., Han, N. & De Yoreo, J. J. Mechanisms of classical crystal growth theory explain quartz and silicate dissolution behavior. Proc. Natl. Acad. Sci. U S A. 102, 15357–15362 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Arvidson, R. S. & Luttge, A. Mineral dissolution kinetics as a function of distance from equilibrium - New experimental results. Chem. Geol.269, 79–88 (2010). [Google Scholar]
- 19.Davis, M. C., Wesolowski, D. J., Rosenqvist, J., Brantley, S. L. & Mueller, K. T. Solubility and near-equilibrium dissolution rates of quartz in dilute NaCl solutions at 398-473K under alkaline conditions. Geochim. Cosmochim. Acta. 75, 401–415 (2011). [Google Scholar]
- 20.Icenhower, J. P. & Dove, P. M. The dissolution kinetics of amorphous silica into sodium chloride solutions: effects of temperature and ionic strength. Geochim. Cosmochim. Acta. 64, 4193–4203 (2000). [Google Scholar]
- 21.Gong, K., Aytas, T., Zhang, S. Y. & Olivetti, E. A. Data-driven prediction of quartz dissolution rates at near-neutral and alkaline environments. Front. Mater.9, 924834 (2022). [Google Scholar]
- 22.Crundwell, F. K. On the mechanism of the dissolution of quartz and silica in aqueous solutions. ACS Omega. 2, 1116–1127 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Bañuelos, J. L. et al. Oxide- and silicate-water interfaces and their roles in technology and the environment. Chemical Reviews vol. 123 6413–6544 Preprint at (2023). 10.1021/acs.chemrev.2c00130 [DOI] [PubMed]
- 24.Adeagbo, W. A., Doltsinis, N. L., Klevakina, K. & Renner, J. Transport processes at α-Quartz–water interfaces: insights from first‐principles molecular dynamics simulations. ChemPhysChem9, 994–1002 (2008). [DOI] [PubMed] [Google Scholar]
- 25.Dellostritto, M. J., Kubicki, J. & Sofo, J. O. Density functional theory simulation of hydrogen-bonding structure and vibrational densities of States at the quartz (1 0 1)-water interface and its relation to dissolution as a function of solution pH and ionic strength. J. Phys. Condens. Matter. 26, 244101 (2014). [DOI] [PubMed] [Google Scholar]
- 26.Nangia, S. & Garrison, B. J. Role of intrasurface hydrogen bonding on silica dissolution. J. Phys. Chem. C. 114, 2267–2272 (2010). [Google Scholar]
- 27.Kubicki, J. D., Sofo, J. O., Skelton, A. A. & Bandura, A. V. A new hypothesis for the dissolution mechanism of silicates. J. Phys. Chem. C. 116, 17479–17491 (2012). [Google Scholar]
- 28.Pfeiffer-Laplaud, M. & Gaigeot, M. P. Adsorption of singly charged ions at the hydroxylated (0001) α-quartz/water interface. J. Phys. Chem. C. 120, 4866–4880 (2016). [Google Scholar]
- 29.Pelmenschikov, A., Leszczynski, J. & Pettersson, L. G. M. Mechanism of dissolution of neutral silica surfaces: including effect of self-healing. J. Phys. Chem. A. 105, 9528–9532 (2001). [Google Scholar]
- 30.Casey, W. H., Lasaga, A. C. & Gibbs, G. V. Mechanisms of silica dissolution as inferred from the kinetic isotope effect. Geochim. Cosmochim. Acta. 54, 3369–3378 (1990). [Google Scholar]
- 31.Xiao, Y. & Lasaga, A. C. Ab initio quantum mechanical studies of the kinetics and mechanisms of silicate dissolution: H+(H3O+) catalysis. Geochim. Cosmochim. Acta. 58, 5379–5400 (1994). [Google Scholar]
- 32.Criscenti, L. J., Kubicki, J. D. & Brantley, S. L. Silicate glass and mineral dissolution: calculated reaction paths and activation energies for hydrolysis of a Q3 Si by H3O+ using ab initio methods. J. Phys. Chem. A. 110, 198–206 (2006). [DOI] [PubMed] [Google Scholar]
- 33.Nangia, S. & Garrison, B. J. Reaction rates and dissolution mechanisms of quartz as a function of pH. J. Phys. Chem. A. 112, 2027–2033 (2008). [DOI] [PubMed] [Google Scholar]
- 34.Mahadevan, T. S. & Garofalini, S. H. Dissociative chemisorption of water onto silica surfaces and formation of hydronium ions. J. Phys. Chem. C. 112, 1507–1515 (2008). [Google Scholar]
- 35.Nangia, S. & Garrison, B. J. Ab initio study of dissolution and precipitation reactions from the edge, kink, and terrace sites of quartz as a function of pH. Mol. Phys.107, 831–843 (2009). [Google Scholar]
- 36.Lopes, P. E. M., Demchuk, E. & Mackerell, A. D. Reconstruction of the (011) surface on α-quartz: A Semiclassical Ab initio molecular dynamics study. Int. J. Quantum Chem.109, 50–64 (2009). [Google Scholar]
- 37.Lopes, P. E. M., Murashov, V., Tazi, M., Demchuk, E. & MacKerell, A. D. Development of an empirical force field for silica. Application to the quartz-water interface. J. Phys. Chem. B. 110, 2782–2792 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Leung, K., Nielsen, I. M. B. & Criscenti, L. J. Elucidating the bimodal acid-base behavior of the water-silica interface from first principles. J. Am. Chem. Soc.131, 18358–18365 (2009). [DOI] [PubMed] [Google Scholar]
- 39.Wallace, A. F., Gibbs, G. V. & Dove, P. M. Influence of ion-associated water on the hydrolysis of Si-O bonded interactions. J. Phys. Chem. A. 114, 2534–2542 (2010). [DOI] [PubMed] [Google Scholar]
- 40.Martin, P., Gaitero, J. J., Dolado, J. S. & Manzano, H. New kinetic Monte Carlo model to study the dissolution of quartz. ACS Earth Space Chem.5, 516–524 (2021). [Google Scholar]
- 41.DelloStritto, M. J., Kubicki, J. D. & Sofo, J. O. Effect of ions on h-bond structure and dynamics at the quartz(101)-water interface. Langmuir32, 11353–11365 (2016). [DOI] [PubMed] [Google Scholar]
- 42.Criscenti, L. J. & Sverjensky, D. A. The role of electrolyte anions (ClO₄–,NO₃-, and Cl-) in divalent metal (M2+) adsorption on oxide and hydroxide surfaces in salt solutions. Am. J. Sci.299, 828–899 (1999). [Google Scholar]
- 43.Goumans, T. P. M., Wander, A., Brown, W. A. & Catlow, C. R. A. Structure and stability of the (001) α-quartz surface. Phys. Chem. Chem. Phys.9, 2146–2152 (2007). [DOI] [PubMed] [Google Scholar]
- 44.Coombs, P. G. et al. The nature of the Si-O-Si bond angle distribution in vitreous silica. Philos. Mag. B. 51, L39–L42 (1985). [Google Scholar]
- 45.Yuan, X. & Cormack, A. N. Si–O–Si bond angle and torsion angle distribution in vitreous silica and sodium silicate glasses. J. Non Cryst. Solids. 319, 31–43 (2003). [Google Scholar]
- 46.Gibbs, G. V., Jayatilaka, D., Spackman, M. A., Cox, D. F. & Rosso, K. M. Si-O bonded interactions in silicate crystals and molecules: A comparison. J. Phys. Chem. A. 110, 12678–12683 (2006). [DOI] [PubMed] [Google Scholar]
- 47.Duval, Y., Mielczarski, J. A., Pokrovsky, O. S., Mielczarski, E. & Ehrhardt, J. J. Evidence of the existence of three types of species at the quartz-aqueous solution interface at pH 0–10: XPS surface group quantification and surface complexation modeling. J. Phys. Chem. B. 106, 2937–2945 (2002). [Google Scholar]
- 48.Frisch, M. J. et al. Gaussian 16. Preprint at (2016).
- 49.Kohn, W. & Sham, L. J. Self-Consistent equations including exchange and correlation effects. Phys. Rev.140, A1133–A1138 (1965). [Google Scholar]
- 50.Hohenberg, P. & Kohn, W. Inhomogeneous Electron. Gas Phys. Rev.136, B864–B871 (1964). [Google Scholar]
- 51.Zhao, Y. & Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other function. Theor. Chem. Acc.120, 215–241 (2008). [Google Scholar]
- 52.Linder, M. & Brinck, T. On the method-dependence of transition state asynchronicity in Diels–Alder reactions. Phys. Chem. Chem. Phys.15, 5108 (2013). [DOI] [PubMed] [Google Scholar]
- 53.Krishnan, R., Brinkley, J. S., Seeger, R. & Pople, J. A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys.72, 650–654 (1980). [Google Scholar]
- 54.Marenich, A. V., Cramer, C. J. & Truhlar, D. G. Unviersal solvation model based on solute electron density and a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B. 113, 6378–6396 (2009). [DOI] [PubMed] [Google Scholar]
- 55.Felipe, M., Xiao, Y. & Kubicki, J. D. Molecular orbital modeling and transition state theory In the geosciences. In molecular modeling theory: applications In the geosciences. Rev. Mineral. Geochem.42 vol, 42 485–531 (2001). [Google Scholar]
- 56.Kubicki, J. D., Blake, G. A. & Apitz, S. E. Ab initio calculations on aluminosilicate Q3 species: implications for atomic structures of mineral surfaces and dissolution mechanisms of feldspars. Am. Mineral.81, 789–799 (1996). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
Data is provided within the manuscript.



