Revealing Transition States during the Hydration of ... - ACS Publications

Jun 20, 2019 - important role in many environmental and industrial systems, including ..... tools, such as density functional theory or a reactive for...
0 downloads 0 Views 5MB Size
Letter pubs.acs.org/JPCL

Cite This: J. Phys. Chem. Lett. 2019, 10, 3704−3709

Revealing Transition States during the Hydration of Clay Minerals Tuan A. Ho,* Louise J. Criscenti, and Jeffery A. Greathouse Geochemistry Department, Sandia National Laboratories, Albuquerque, New Mexico 87185, United States

Downloaded via BUFFALO STATE on July 31, 2019 at 08:15:12 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: A molecular-scale understanding of the transition between hydration states in clay minerals remains a challenging problem because of the very fast stepwise swelling process observed from X-ray diffraction (XRD) experiments. XRD profile modeling assumes the coexistence of multiple hydration states in a clay sample to fit the experimental XRD pattern obtained under humid conditions. While XRD profile modeling provides a macroscopic understanding of the heterogeneous hydration structure of clay minerals, a microscopic model of the transition between hydration states is still missing. Here, for the first time, we use molecular dynamics simulation to investigate the transition states between a dry interlayer, one-layer hydrate, and two-layer hydrate. We find that the hydrogen bonds that form across the interlayer at the clay particle edge make an important contribution to the energy barrier to interlayer hydration, especially for initial hydration.

S

swelling of clay progresses in a stepwise fashion: the hydration state varies abruptly from dry to one water layer (0W−1W) and from one water layer to two water layers (1W−2W) as RH increases. These hydration states are easily identified by measuring basal spacing (d-spacing) using powder X-ray diffraction (XRD).46−52 However, the transition between hydration states is very difficult to characterize at the molecular level because of abrupt d-spacing changes over small RH intervals. XRD profile modeling has revealed the coexistence of multiple interlayer states of a clay sample at a single RH value.51−55 For example, to fit the XRD results for Nasaturated synthetic saponites obtained at 49% RH, Ferrage et al. estimated the coexistence of multiple hydration states (17.9% 2W, 79.6% 1W, and 2.5% 0W) in the sample.51 While XRD profile modeling provides a useful macroscopic understanding of the heterogeneous hydration structure of clay minerals, a molecular model of the transition between hydration states is still missing. Our goals in this work are to understand (i) how water molecules enter the interlayer of MMT, thereby initiating the expansion from the 0W to 1W state and (ii) the molecular structure of the 0W−1W and 1W− 2W transition states. In the structure of bulk MMT, layer hydroxyl (−OH) groups are associated with the Al-centered octahedral sheet and do not participate in forming hydrogen bonds (HBs) between TOT layers. However, −OH groups associated with edge Si atoms might initiate HBs between TOT layers. Many prior molecular simulation studies12−19 modeled the bulk

mectite clays including montmorillonite (MMT) are naturally occurring phyllosilicate minerals that play an important role in many environmental and industrial systems, including cement and concrete, health and beauty products, underground disposal of waste streams, and energy resources.1−7 In the MMT structure, two Si-centered tetrahedral sheets stack above and below an Al-centered octahedral sheet to form a TOT (tetrahedral−octahedral−tetrahedral) layer. Isomorphic metal substitutions result in negatively charged TOT layers that are balanced by interlayer cations.8 The interlayer cations readily attract water, enabling MMT swelling with increasing relative humidity (RH). In addition to RH, temperature, pressure, net layer charge, substitution location, and interlayer cation have been documented to affect the hydration process of MMT and other clay phases.9−21 The influence of these variables on clay mineral hydration has been the focus of numerous molecular simulation studies of bulk clay models, as detailed previously.22−25 The stepwise swelling process of clays due to interlayer hydration and expansion likely occurs as a series of rare events at time scales outside the scope of standard molecular dynamics (MD) simulation methods. Enhanced sampling methods26−28 have been used in MD simulations of other rare events, including ion adsorption−desorption,29−32 nucleation,33−35 bulk solution properties,36,37 and hydration of layered materials.38,39 However, to our knowledge such methods have not been used to determine the free-energy profile of clay swelling. The hydration of expandable clays strongly affects the retention and transport of water and ions.40−44 Additionally, interlayer expansion results in very complex chemo-mechanical properties in clay-containing subsurface systems.22,45 The © 2019 American Chemical Society

Received: May 31, 2019 Accepted: June 17, 2019 Published: June 20, 2019 3704

DOI: 10.1021/acs.jpclett.9b01565 J. Phys. Chem. Lett. 2019, 10, 3704−3709

Letter

The Journal of Physical Chemistry Letters

Figure 1. Potential of mean force (PMF) profile of the hydration process (top left). The simulation snapshots A, B, C, D, and E represent the hydration states for the corresponding points on the PMF profile. Simulation snapshot showing the Na-MMT particle (0W state) immersed in water (A). Simulation snapshots demonstrate possible structures for the 0W−1W transition state (B), 1W state (C), 1W−2W transition state (D), and 2W state (E). The d-spacing (i.e., the distance between centers of mass of two TOT layers) values for A, B, C, D and E are 9.85, 10.95, 12.55, 13.25, and 15.13 Å, respectively. Na+ ions are yellow spheres. Water molecules are shown as red sticks (external fluid) and red and white spheres (interlayer).

MMT structure (without edges) using the ClayFF56 force field. Recent ClayFF development allows us to simulate clay particles with protonated edges57,58 and to model the HBs between TOT layers in a MMT particle. Our simulations indicate that these HBs significantly affect the swelling behavior of MMT. In the top left panel of Figure 1 we report the potential of mean force (PMF) profile between two TOT layers of a NaMMT particle immersed in water (Figure 1A). The particle is finite in the x and z directions and infinite in the y direction. Umbrella sampling was used to constrain the interlayer distance (z direction), allowing water molecules and Na+ ions to exchange between the bulk solution and the interlayer through the particle edges in the x direction. The Na-MMT nanoparticle was simulated using the swell-ClayFF force field, a modified version of ClayFF56 with new interaction parameters between Na+ ions and bridging oxygen atoms of the MMT structure. For other simulation details and force field parameters, see the Supporting Information. The PMF profile in Figure 1 provides thermodynamic insight into the hydration process. When placed in a simulation box full of water (Figure 1A), the Na-MMT particle is expected to undergo swelling (i.e., water molecules intercalate into the interlayer) as evidenced by experiment.47 The minimum in the PMF profile at 9.85 Å (point A) corresponds to the d-spacing of dry Na-MMT, in agreement with XRD results (9.6−9.8 Å).59−63 The shoulder at 12.55 Å (point C) corresponds to the 1W hydration state. The global minimum

observed at 15.13 Å (point E) corresponds to the 2W state, which agrees with XRD results under aqueous conditions.47 The PMF profile also reveals the energy barrier for the 0W− 1W transition (point B, which has the highest energy on the PMF profile). Because each TOT layer possesses a negative charge, there will be a repulsive force between two TOT layers. Strong electrostatic attraction with interlayer Na+ ions keeps the TOT layers together. The repulsive short-range LennardJones (LJ) interaction between Na+ ions and atoms belonging to the TOT layer plays a crucial role in separating two TOT layers during the hydration from 0W to 1W state as we discuss in the Supporting Information. Briefly, we found that ClayFF56 overpredicts the interaction between Na+ ions and surface O atoms, making the Na-MMT dry state too stable in water. We modified the LJ interaction between Na+ ions and bridging O atoms of the MMT structure to obtain stronger repulsive forces at short distances, resulting in a lower energy barrier for interlayer hydration and swelling of the Na-MMT particle. Note that the new Na+−bridging O atoms LJ parameters are well validated for dry Na-MMT as described in the Supporting Information, but a full validation for the Na+ ion structure in the hydrated interlayers or Na+ ion structure near the MMT external surface is beyond the scope of this work. However, because interlayer cations are fully hydrated and thus farther removed from surface O atoms, we expect that these parameters will have a minor effect on the properties of hydrated Na-MMT simulated by ClayFF. Also contributing are HB interactions between −OH groups of adjacent TOT layers 3705

DOI: 10.1021/acs.jpclett.9b01565 J. Phys. Chem. Lett. 2019, 10, 3704−3709

Letter

The Journal of Physical Chemistry Letters

Figure 2. Density profiles of water oxygen atoms in the 1W (A), 1W−2W transition (B, C, and D), 2W (E), and expanded 17 Å (F) states. The fixed d-spacing values are shown in each panel.

Figure 3. Simulation snapshots show HBs at the particle edge when full HB interactions are included in the simulations (A) and the expanded pore opening when HBs between TOT layers are not included (B). Hydroxyl oxygen and hydrogen atoms are shown as silver and green spheres, respectively. The edge −OH groups are enlarged compared to the −OH groups in the octahedral sheet. Na+ ions are in yellow. These simulations were conducted for the dry state (i.e., no interlayer water). Water molecules in the external fluid were removed for clarity. These snapshots were rotated around the z axis to visualize more edge sites. Comparison of the PMF profiles with and without HBs between TOT layers (C).

In Figure 2, we present the density profiles of water oxygen atoms at different d-spacings to further visualize the 1W−2W transition. The results indicate that this transition is associated with a gradual change in the disposition and growth of water layers. For the 1W state (Figure 2A), water molecules form a single peak at the midplane. As the d-spacing increases, this peak splits, and shoulders form near the basal surfaces (Figure 2B). The intensities of the midplane peaks continue to decrease while those of the shoulders increase (Figure 2C). Eventually, the inner peaks and shoulders merge (Figure 2D) to form two pronounced peaks in the 2W state (Figure 2E). When the d-spacing is increased further to 17 Å (Figure 2F), no additional peaks are observed as the water density reaches a constant value at the midplane. For periodic MMT structures (i.e., without edges), there are no HBs between TOT layers. However, our MMT model contains edge−OH groups that can form HBs between TOT layers in the 0W state (Figure 3A). These HBs act like a gate, preventing the passage of water molecules and ions into the interlayer. When these HBs are prevented from forming (i.e., by turning off force field interactions between edge hydroxyl hydrogen atoms from adjacent TOT layers), the interlayer is much more accessible to external fluid (Figure 3B). We hypothesize that the swelling process begins by breaking those HBs so that water molecules can diffuse into the interlayer. To understand the energy barrier associated with HB breaking, in Figure 3C we compare PMF profiles for the hydration process

(discussed below). The 1W−2W transition (i.e., point D) occurs without a significant energy barrier. The snapshot in Figure 1B illustrates a possible 0W−1W transition state (i.e., corresponding to point B on the PMF profile) of the swelling process: water molecules move from the external fluid into the dry interlayer region to solvate the Na+ ions. The difference in interlayer spacing between the edge (partially hydrated) and center (dry) regions results in a water concentration gradient across the lateral dimension of the NaMMT particle and mechanical bending of the TOT layer. Water molecules continue to fill the interlayer region until all Na+ ions are partially hydrated in the 1W state (Figure 1C). Figure 1D reveals a possible 1W−2W transition. As we discuss in Figure 2, the 1W−2W transition is a complicated process with changes in the water structure as the d-spacing increases. At the stable 2W state (Figure 1E), two distinguishable water layers are observed in the interlayer. Note that interpretation of thermodynamic properties from a PMF calculation depends on the choice of the initial and final states and the associated reaction coordinate. While the selection of the 0W, 1W, and 2W states seems obvious in our PMF calculation, one can select a collective variable other than the d-spacing to study the free-energy path. For example, one can vary the number of interlayer water molecules or construct a two-dimensional free-energy map as a function of d-spacing and number of interlayer water molecules to provide a better picture of the hydration transition. 3706

DOI: 10.1021/acs.jpclett.9b01565 J. Phys. Chem. Lett. 2019, 10, 3704−3709

The Journal of Physical Chemistry Letters with and without the HB interactions between TOT layers. The results indicate that without these HBs, the 0W−1W energy barrier is significantly reduced. The HBs also strongly affect the 0W−1W free-energy difference, while the impact of the HBs on the swelling process beyond the 1W state is negligible. The presence of interlayer water and expanded dspacing (i.e., 1W state) prevents HB interactions between TOT layers. It is also possible that proton transfer could significantly affect the mechanism and energy differences between hydration states, but this can be investigated only using smaller and less realistic models with other simulation tools, such as density functional theory or a reactive force field. Note that the contribution of HB interactions on the hydration energy of clay platelets depends on the size and shape of the particle. In our model we consider only HBs on the edge of the MMT ribbon (i.e., the particle is infinite in the y direction). The number of HBs in a realistic hexagonal MMT particle increases significantly, compared to the MMT ribbon on the basis of the same basal surface area. For the realistic hexagonal MMT particle, the lateral dimension is much larger (up to several hundred nanometers 64 ) than can be simulated.65,66 Obviously, the number of HBs at edge increases with the increasing lateral dimension. It is reasonable to postulate that the contribution of HB interactions on the hydration energy depends on the ratio of the perimeter and surface area of the hexagon (i.e., ∼1/b, where b is the edge of the hexagon). Therefore, when increasing the edge size b by a factor of n, the contribution of the HB interaction on hydration energy decreases by a factor of n. Future MD studies using hexagonal particles can help verify this simple geometric analysis. Our MD simulations address a critical gap in understanding the hydration process of smectite clays. Importantly, our simulations include the transition from the dry (0W) to onelayer hydrate (1W) states, which is often omitted from MD simulation studies of clay swelling. Our slightly modified ClayFF force field correctly reproduces the expected freeenergy trend of interlayer hydration. The hydration process begins by breaking the HB network between TOT layers at the particle edges, allowing water molecules to diffuse into the dry interlayer. These layer−layer HBs significantly contribute to the energy barrier of the 0W−1W transition. Changes in interlayer water structure also characterize the 1W−2W transition.





ACKNOWLEDGMENTS



REFERENCES

Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC., a wholly owned subsidiary of Honeywell International, Inc., for the U.S. Department of Energy’s National Nuclear Security Administration under contract DE-NA0003525. The views expressed in this Letter do not necessarily represent the views of the U.S. Department of Energy or the United States Government. This research was funded by the Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and Biosciences Division.

(1) Amato, I. Concrete Solutions. Nature 2013, 494, 300−301. (2) Carretero, M. I. Clay Minerals and Their Beneficial Effects Upon Human Health. A Review. Appl. Clay Sci. 2002, 21, 155−163. (3) Carretero, M. I.; Pozo, M. Clay and Non-Clay Minerals in the Pharmaceutical and Cosmetic Industries Part II. Active Ingredients. Appl. Clay Sci. 2010, 47, 171−181. (4) Colón, C. F. J., et al. Evaluation of Used Fuel Disposition in ClayBearing Rock (FCRD-UFD-2014-000056); SAND2014-18303 R; Sandia National Laboratories: Albuquerque, NM, 2014; p 434. (5) Busch, A.; Alles, S.; Gensterblum, Y.; Prinz, D.; Dewhurst, D. N.; Raven, M. D.; Stanjek, H.; Krooss, B. M. Carbon Dioxide Storage Potential of Shales. Int. J. Greenhouse Gas Control 2008, 2, 297−308. (6) Arthur, M. A.; Cole, D. R. Unconventional Hydrocarbon Resources: Prospects and Problems. Elements 2014, 10, 257−264. (7) Wu, L. M.; Zhou, C. H.; Keeling, J.; Tong, D. S.; Yu, W. H. Towards an Understanding of the Role of Clay Minerals in Crude Oil Formation, Migration and Accumulation. Earth-Sci. Rev. 2012, 115, 373−386. (8) Odom, I. E. Smectite Clay-Minerals - Properties and Uses. Philos. Trans. R. Soc., A 1984, 311, 391−409. (9) Sun, L.; Hirvi, J. T.; Schatz, T.; Kasa, S.; Pakkanen, T. A. Estimation of Montmorillonite Swelling Pressure: A Molecular Dynamics Approach. J. Phys. Chem. C 2015, 119, 19863−19868. (10) Sun, L. L.; Ling, C. Y.; Lavikainen, L. P.; Hirvi, J. T.; Kasa, S.; Pakkanen, T. A. Influence of Layer Charge and Charge Location on the Swelling Pressure of Dioctahedral Smectites. Chem. Phys. 2016, 473, 40−45. (11) Hsiao, Y. W.; Hedstrom, M. Swelling Pressure in Systems with Na-Montmorillonite and Neutral Surfaces: A Molecular Dynamics Study. J. Phys. Chem. C 2017, 121, 26414−26423. (12) Dazas, B.; Lanson, B.; Delville, A.; Robert, J. L.; Komarneni, S.; Michot, L. J.; Ferrage, E. Influence of Tetrahedral Layer Charge on the Organization of Inter Layer Water and Ions in Synthetic NaSaturated Smectites. J. Phys. Chem. C 2015, 119, 4158−4172. (13) Teich-McGoldrick, S. L.; Greathouse, J. A.; Jove-Colon, C. F.; Cygan, R. T. Swelling Properties of Montmorillonite and Beidellite Clay Minerals from Molecular Simulation: Comparison of Temperature, Interlayer Cation, and Charge Location Effects. J. Phys. Chem. C 2015, 119, 20880−20891. (14) Smith, D. E.; Wang, Y.; Chaturvedi, A.; Whitley, H. D. Molecular Simulations of the Pressure, Temperature, and Chemical Potential Dependencies of Clay Swelling. J. Phys. Chem. B 2006, 110, 20046−20054. (15) Rotenberg, B.; Morel, J. P.; Marry, V.; Turq, P.; MorelDesrosiers, N. On the Driving Force of Cation Exchange in Clays: Insights from Combined Microcalorimetry Experiments and Molecular Simulation. Geochim. Cosmochim. Acta 2009, 73, 4034−4044. (16) Skipper, N. T.; Sposito, G.; Chang, F. R. C. Monte-Carlo Simulation of Interlayer Molecular-Structure in Swelling ClayMinerals. 2. Monolayer Hydrates. Clays Clay Miner. 1995, 43, 294− 303. (17) Morrow, C. P.; Yazaydin, A. O.; Krishnan, M.; Bowers, G. M.; Kalinichev, A. G.; Kirkpatrick, R. J. Structure, Energetics, and

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpclett.9b01565.



Letter

Simulation methods and force field modification (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Tuan A. Ho: 0000-0002-8129-1027 Louise J. Criscenti: 0000-0002-5212-7201 Jeffery A. Greathouse: 0000-0002-4247-3362 Notes

The authors declare no competing financial interest. 3707

DOI: 10.1021/acs.jpclett.9b01565 J. Phys. Chem. Lett. 2019, 10, 3704−3709

Letter

The Journal of Physical Chemistry Letters Dynamics of Smectite Clay Interlayer Hydration: Molecular Dynamics and Metadynamics Investigation of Na-Hectorite. J. Phys. Chem. C 2013, 117, 5172−5187. (18) Liu, X. D.; Lu, X. C.; Wang, R. C.; Zhou, H. Q. Effects of Layer-Charge Distribution on the Thermodynamic and Microscopic Properties of Cs-Smectite. Geochim. Cosmochim. Acta 2008, 72, 1837−1847. (19) Chavez-Paez, M.; dePablo, L.; dePablo, J. J. Monte Carlo Simulations of Ca-Montmorillonite Hydrates. J. Chem. Phys. 2001, 114, 10948−10953. (20) Honorio, T.; Brochard, L.; Vandamme, M. Hydration Phase Diagram of Clay Particles from Molecular Simulations. Langmuir 2017, 33, 12766−12776. (21) Ngouana, W. B. F.; Kalinichev, A. G. Structural Arrangements of Isomorphic Substitutions in Smectites: Molecular Simulation of the Swelling Properties, Inter Layer Structure, and Dynamics of Hydrated Cs-Montmorillonite Revisited with New Clay Models. J. Phys. Chem. C 2014, 118, 12758−12773. (22) Anderson, R. L.; Ratcliffe, I.; Greenwell, H. C.; Williams, P. A.; Cliffe, S.; Coveney, P. V. Clay Swelling - a Challenge in the Oilfield. Earth-Sci. Rev. 2010, 98, 201−216. (23) Greathouse, J. A.; Cygan, R. T. Molecular Simulation of Clay Minerals. In Handbook of Clay Science, 2nd ed.; Bergaya, F., Lagaly, G., Eds.; Elsevier: Amsterdam, 2013; Vol. B. (24) Bourg, I. C.; Ajo-Franklin, J. B. Clay, Water, and Salt: Controls on the Permeability of Fine-Grained Sedimentary Rocks. Acc. Chem. Res. 2017, 50, 2067−2074. (25) Ma, Z. Y.; Gamage, R. P.; Rathnaweera, T.; Kong, L. Review of Application of Molecular Dynamic Simulations in Geological HighLevel Radioactive Waste Disposal. Appl. Clay Sci. 2019, 168, 436− 449. (26) Abrams, C.; Bussi, G. Enhanced Sampling in Molecular Dynamics Using Metadynamics, Replica-Exchange, and TemperatureAcceleration. Entropy 2014, 16, 163−199. (27) Bernardi, R. C.; Melo, M. C. R.; Schulten, K. Enhanced Sampling Techniques in Molecular Dynamics Simulations of Biological Systems. Biochim. Biophys. Acta, Gen. Subj. 2015, 1850, 872−877. (28) Yang, L. J.; Liu, C. W.; Shao, Q.; Zhang, J.; Gao, Y. Q. From Thermodynamics to Kinetics: Enhanced Sampling of Rare Events. Acc. Chem. Res. 2015, 48, 947−955. (29) Leung, K.; Criscenti, L. J.; Knight, A. W.; Ilgen, A. G.; Ho, T. A.; Greathouse, J. A. Concerted Metal Cation Desorption and Proton Transfer on Deprotonated Silica Surfaces. J. Phys. Chem. Lett. 2018, 9, 5379−5385. (30) Sena, M. M.; Morrow, C. P.; Kirkpatrick, R. J.; Krishnan, M. Supercritical Carbon Dioxide at Smectite Mineral-Water Interfaces: Molecular Dynamics and Adaptive Biasing Force Investigation of CO2/H2O Mixtures Nanoconfined in Na-Montmorillonite. Chem. Mater. 2015, 27, 6946−6959. (31) Underwood, T.; Erastova, V.; Greenwell, H. C. Ion Adsorption at Clay-Mineral Surfaces: The Hofmeister Series for Hydrated Smectite Minerals. Clays Clay Miner. 2016, 64, 472−487. (32) Wright, L. B.; Walsh, T. R. Efficient Conformational Sampling of Peptides Adsorbed onto Inorganic Surfaces: Insights from a Quartz Binding Peptide. Phys. Chem. Chem. Phys. 2013, 15, 4715−4726. (33) Sosso, G. C.; Chen, J.; Cox, S. J.; Fitzner, M.; Pedevilla, P.; Zen, A.; Michaelides, A. Crystal Nucleation in Liquids: Open Questions and Future Challenges in Molecular Dynamics Simulations. Chem. Rev. 2016, 116, 7078−7116. (34) Malolepsza, E.; Kim, J.; Keyes, T. Entropic Description of Gas Hydrate Ice-Liquid Equilibrium Via Enhanced Sampling of Coexisting Phases. Phys. Rev. Lett. 2015, 114, 170601. (35) Salvalaglio, M.; Perego, C.; Giberti, F.; Mazzotti, M.; Parrinello, M. Molecular-Dynamics Simulations of Urea Nucleation from Aqueous Solution. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, E6−E14. (36) Brancato, G.; Barone, V. Free Energy Landscapes of Ion Coordination in Aqueous Solution. J. Phys. Chem. B 2011, 115, 12875−12878.

(37) Damjanovic, A.; Miller, B. T.; Okur, A.; Brooks, B. R. Reservoir pH Replica Exchange. J. Chem. Phys. 2018, 149, 072321. (38) Cao, W.; Wang, J.; Ma, M. Exfoliation of Two-Dimensional Materials: The Role of Entropy. J. Phys. Chem. Lett. 2019, 10, 981− 986. (39) Yoshida, H.; Kaiser, V.; Rotenberg, B.; Bocquet, L. Dripplons as Localized and Superfast Ripples of Water Confined between Graphene Sheets. Nat. Commun. 2018, 9, 1496. (40) Higgo, J. J. W. Clay as a Barrier to Radionuclide Migration. Prog. Nucl. Energy 1987, 19, 173−207. (41) Pusch, R. Use of Bentonite for Isolation of Radioactive-Waste Products. Clay Miner. 1992, 27, 353−361. (42) Yuan, G. D.; Theng, B. K. G.; Churchman, G. J.; Gates, W. P. Clays and Clay Minerals for Pullution Control, 2nd ed.; Elsever: Amsterdam, 2013; Vol. B. (43) Rotenberg, B.; Marry, V.; Vuilleumier, R.; Malikova, N.; Simon, C.; Turq, P. Water and Ions in Clays: Unraveling the Interlayer/ Micropore Exchange Using Molecular Dynamics. Geochim. Cosmochim. Acta 2007, 71, 5089−5101. (44) Tournassat, C.; Bourg, I. C.; Holmboe, M.; Sposito, G.; Steefel, C. I. Molecular Dynamics Simulations of Anion Exclusion in Clay Interlayer Nanopores. Clays Clay Miner. 2016, 64, 374−388. (45) Boek, E. S.; Coveney, P. V.; Skipper, N. T. Monte Carlo Molecular Modeling Studies of Hydrated Li-, Na-, and K-Smectites: Understanding the Role of Potassium as a Clay Swelling Inhibitor. J. Am. Chem. Soc. 1995, 117, 12608−12617. (46) Brindley, G. W.; Brown, G. Crystal Structures of Clay Minerals and Their X-Ray Indentification; 1980. (47) Morodome, S.; Kawamura, K. Swelling Behavior of Na- and CaMontmorillonite up to 150 Degrees C by in Situ X-Ray Diffraction Experiments. Clays Clay Miner. 2009, 57, 150−160. (48) Ferrage, E.; Sakharov, B. A.; Michot, L. J.; Delville, A.; Bauer, A.; Lanson, B.; Grangeon, S.; Frapper, G.; Jiménez-Ruiz, M.; Cuello, G. J. Hydration Properties and Interlayer Organization of Water and Ions in Synthetic Na-Smectite with Tetrahedral Layer Charge. Part 2. Toward a Precise Coupling between Molecular Simulations and Diffraction Data. J. Phys. Chem. C 2011, 115, 1867−1881. (49) Tamura, K.; Yamada, H.; Nakazawa, H. Stepwise Hydration of High-Quality Synthetic Smectite with Various Cations. Clays Clay Miner. 2000, 48, 400−404. (50) Yamada, H.; Nakazawa, H.; Hashizume, H.; Shimomura, S.; Watanabe, T. Hydration Behavior of Na-Smectite Crystals Synthesized at High-Pressure and High-Temperature. Clays Clay Miner. 1994, 42, 77−80. (51) Ferrage, E.; Lanson, B.; Michot, L. J.; Robert, J. L. Hydration Properties and Interlayer Organization of Water and Ions in Synthetic Na-Smectite with Tetrahedral Layer Charge. Part 1. Results from XRay Diffraction Profile Modeling. J. Phys. Chem. C 2010, 114, 4515− 4526. (52) Holmboe, M.; Wold, S.; Jonsson, M. Porosity Investigation of Compacted Bentonite Using XRD Profile Modeling. J. Contam. Hydrol. 2012, 128, 19−32. (53) Lanson, B.; Ferrage, E.; Hubert, F.; Pret, D.; Mareschal, L.; Turpault, M. P.; Ranger, J. Experimental Aluminization of Vermiculite Interlayers: An X-Ray Diffraction Perspective on Crystal Chemistry and Structural Mechanisms. Geoderma 2015, 249, 28−39. (54) Ferrage, E.; Vidal, O.; Mosser-Ruck, R.; Cathelineau, M.; Cuadros, J. A Reinvestigation of Smectite Illitization in Experimental Hydrothermal Conditions: Results from X-Ray Diffraction and Transmission Electron Microscopy. Am. Mineral. 2011, 96, 207−223. (55) Ferrage, E.; Tournassat, C.; Rinnert, E.; Charlet, L.; Lanson, B. Experimental Evidence for Ca-Chloride Ion Pairs in the Interlayer of Montmorillonite. An XRD Profile Modeling Approach. Clays Clay Miner. 2005, 53, 348−360. (56) Cygan, R. T.; Liang, J. J.; Kalinichev, A. G. Molecular Models of Hydroxide, Oxyhydroxide, and Clay Phases and the Development of a General Force Field. J. Phys. Chem. B 2004, 108, 1255−1266. (57) Pouvreau, M.; Greathouse, J. A.; Cygan, R. T.; Kalinichev, A. G. Structure of Hydrated Gibbsite and Brucite Edge Surfaces: DFT 3708

DOI: 10.1021/acs.jpclett.9b01565 J. Phys. Chem. Lett. 2019, 10, 3704−3709

Letter

The Journal of Physical Chemistry Letters Results and Further Development of the ClayFF Classical Force Field with Metal-O-H Angle Bending Terms. J. Phys. Chem. C 2017, 121, 14757−14771. (58) Pouvreau, M.; Greathouse, J. A.; Cygan, R. T.; Kalinichev, A. G. Structure of Hydrated Kaolinite Edge Surfaces: DFT Results and Further Development of the ClayFF Classical Force Field with Metal−O−H Angle Bending Terms. J. Phys. Chem. C 2019, 123, 11628−11638. (59) Mooney, R. W.; Keenan, A. G.; Wood, L. A. Adsorption of Water Vapor by Montmorillonite 0.2. Effect of Exchangeable Ions and Lattice Swelling as Measured by X-Ray Diffraction. J. Am. Chem. Soc. 1952, 74, 1371−1374. (60) Berend, I.; Cases, J. M.; Francois, M.; Uriot, J. P.; Michot, L.; Masion, A.; Thomas, F. Mechanism of Adsorption and Desorption of Water-Vapor by Homoionic Montmorillonites. II. The Li+, Na+, K+, Rb+ and Cs+-Exchanged Forms. Clays Clay Miner. 1995, 43, 324−336. (61) Chiou, C. T.; Rutherford, D. W. Effects of Exchanged Cation and Layer Charge on the Sorption of Water and Egme Vapors on Montmorillonite Clays. Clays Clay Miner. 1997, 45, 867−880. (62) Rutherford, D. W.; Chiou, C. T.; Eberl, D. D. Effects of Exchanged Cation on the Microporosity of Montmorillonite. Clays Clay Miner. 1997, 45, 534−543. (63) Calvet, R. Hydration of Montmorillonite and Diffusion of Exchangeable Cations. 1. Hydration of Montmorillonite Saturated by Monovalent Cations. Ann. Agron 1973, 24, 77−133. (64) Cadene, A.; Durand-Vidal, S.; Turq, P.; Brendle, J. Study of Individual Na-Montmorillonite Particles Size, Morphology, and Apparent Charge. J. Colloid Interface Sci. 2005, 285, 719−730. (65) Ho, T. A.; Greathouse, J. A.; Lee, A. S.; Criscenti, L. J. Enhanced Ion Adsorption on Mineral Nanoparticles. Langmuir 2018, 34, 5926−5934. (66) Ho, T. A.; Greathouse, J. A.; Wang, Y. F.; Criscenti, L. J. Atomistic Structure of Mineral Nano-Aggregates from Simulated Compaction and Dewatering. Sci. Rep. 2017, 7, 15286.

3709

DOI: 10.1021/acs.jpclett.9b01565 J. Phys. Chem. Lett. 2019, 10, 3704−3709