Byrne 2017
Byrne 2017
∼50 Å containing 4127 SPC/Fw16 water molecules plus the solubility of the sulfate phases was as well reproduced as
ions. The free energy perturbation method17,18 was used to possible. All intramolecular parameters for the sulfate anion
calculate the ion solvation free energies, as implemented in were kept fixed at the values proposed by Allan et al.29
LAMMPS. The solute−solvent interactions were gradually 2.3. Ab Initio Molecular Dynamics. In order to validate
switched off in 40 stages of 5 ns each; 20 stages for the the classical simulations, ab initio molecular dynamics (AIMD)
electrostatic interactions followed by a further 20 stages for the has also been performed. Here we use Gaussian plane-wave
van der Waals interactions; finite size19 and standard state approach as embodied in the QuickStep module of the CP2K
corrections20 have been included, while the volume correction code.30 All atoms were represented using the pseudopotentials
required for the NPT runs turned out to be negligible for the of Goedecker−Teter−Hutter form31 (with Ca having 10
chosen system size. The ion pairing free energy was calculated valence electrons) and a triple-ζ double polarized Gaussian
from well-tempered21 and multiple walker22 metadynamics basis set (TZ2P), while the electron density was expanded in an
simulations using the PLUMED 2.2 plug-in.23 The collective auxiliary basis set with a plane-wave cutoff of 400 Ry.
variables used were the calcium−sulfur distance and the Calculations were performed using the BLYP-D3 functional32,33
calcium−water oxygen coordination number, which was (i.e., GGA with Grimme’s dispersion corrections). To partially
calculated using a switching function of the form correct for the systematic overstructuring of water, an elevated
n temperature of 330 K was employed with deuterium replacing
f (d ) =
1−( ) d − d0
r0
hydrogen. This combination of functional, temperature, and
atomic mass has been found to give a more accurate
m
1−( ) d − d0 representation of the experimental water radial density
r0 (1) distribution function at ambient conditions.34 Self-consistent
where d0 = 2.1 Å, r0 = 1.0 Å, n = 4, and m = 10. In order to field was achieved using the orbital transformation algorithm.35
reconstruct the free energy surface, Gaussians of widths 0.2 Å Molecular dynamics was performed in the NVT ensemble
and 0.1 along the distance and coordination number, using a stochastic thermostat and a time step of 0.5 fs. A cell
respectively, were deposited every 1 ps with an initial height dimension of 16 Å was chosen containing the CaSO4 ion pair
of 0.026 eV. A total of 25 walkers were used, resulting in an and 134 water molecules for evaluation of the free energies,
aggregate run time of 250 ns. A bias factor of 5 was used to while a smaller system with a cell of 14.611 Å with SO4 and 100
gradually reduce the Gaussian heights and ensure convergence waters was used for the hydration behavior of the sulfate ion
of the free energy. alone. Initial configurations were taken from classical
All the simulations with the AMOEBA force field24 have simulations and then re-equilibrated with the ab initio
been performed with GULP25 using an in-house interface to molecular dynamics.
routines from the Tinker program26 in order to provide the Task farming was used to run metadynamics for the ion pair
energy and forces for the AMOEBA model. In order to improve with 10 walkers, each of which was run for at least 110 ps,
the efficiency, the calculation of the AMOEBA energy and resulting in a total simulation time of 1.14 ns. Two collective
forces was parallelized using MPI. Parameters for the sulfate− variables were sampled, namely, the Ca−S distance and the
water system have been taken from the work of Lambrecht et Ca−O(water) coordination number using parameters of n = 16
al.27 where the results of the AMOEBA model were and m = 32 for the powers of distances and an r0 of 3 Å (note
benchmarked against those from both density functional and that compared to eq 1 there is no d0 in the CP2K
ab initio methods. For the Ca2+ ion, the parameters were taken implementation of the coordination number). It is important
from the AMOEBA09 library, though we note that other to note that the use of the water coordination number for Ca2+
parameter sets for Ca2+ in water are also available.28 Full self- is essential for determining the ion pairing free energy
consistency was used for the multipole moments at each step of landscape in AIMD because the rate of water exchange at
the trajectory. All simulations were initially equilibrated in the this cation is still slow compared with the time scale that is
NPT ensemble for an isotropic cubic cell of approximately 25 Å accessible to each walker, even if it is relatively fast on an
in size containing 520 water molecules plus the ions using a experimental time scale. For the unbiased dynamics of the
stochastic barostat and thermostat with a time constant of 0.1 sulfate anion in water, a simulation of 54 ps was performed.
ps. The PLUMED 2.2 plug-in23 was also used to calculate the 2.4. Ab Initio Solvation Free Energy. To provide a
ion pairing free energy as a function of the same two collective further benchmark for the solvation free energy of the sulfate
variables using task farming of multiple walkers within GULP. anion, this quantity has also been determined using molecular
2.2. Lattice Dynamics. Calculations of the bulk properties quantum mechanical methods. Here, calculations were
of the solid phases were performed using lattice dynamics performed using QChem 4.136 at the M06/6-31+G** level
within the GULP code.25 The free energies of the crystalline of theory within the SM8 continuum solvation model.37 The
phases were computed using the quasi-harmonic approximation solvation free energy taken is that for the geometry of the
at 298.15 K with phonons sampled using a shrinking factor of 6 sulfate anion after optimization in the presence of the solvation
to define the Monkhorst−Pack mesh. For consistency with the model.
target application in molecular dynamics simulations, the zero-
point energy contribution was excluded from the free energy. 3. FORCE FIELD VALIDATION
Single-point calculation of the free energies was used at the As a starting point for the development of a thermodynamically
fully optimized bulk structure with respect to the internal accurate force field for CaSO4 in water, we took our previous
energy. force field for Ca2+ with SPC/Fw water12,38 and the parameters
Relaxed fitting was used to determine the calcium−sulfate for sulfate ions developed by Allan et al.29 The calcium−water
interaction parameters using the anhydrite structures and bulk interaction parameters were fitted to reproduce the solvation
moduli, where available, as observables. The free energy was free energy of the ion and consisted of a simple Lennard-Jones
then iteratively refined as a further constraint to ensure that the potential:
B DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article
⎡⎛ σ ⎞12 ⎛ σ ⎞6 ⎤ ΔG = ΔH − T ΔS (6)
Uvdw = 4ε⎢⎜ ⎟ − ⎜ ⎟ ⎥ 41
⎣⎝ r ⎠ ⎝r⎠ ⎦ (2)
However, in a later publication Marcus suggested that the
enthalpic (−1035 kJ/mol) and entropic (−200 J/mol/K) terms
The sulfate intramolecular interactions are described by a are more reliable than the free energy itself, and the reported
Morse bond stretching term and a harmonic angle bending value, ΔG = −975 kJ/mol at standard conditions, was indeed
term; simply obtained using the definition of the Gibbs free energy
and not actually measured. Because of this ambiguity in the
UMorse bond = D[1 − e−α(rij − r0)]2 (3) literature, we decided to develop two sets of force field
parameters (hereafter labeled as FF1 and FF2) that target the
1
Uharmonic angle = kθ(θijk − θ0)2 two extreme sets of values of the hydration free energy
2 (4) (−1080/−1090 and −975 kJ/mol) and use both to study the
where D is the potential well depth, α is the stiffness parameter, Ca−SO4 ion pairing process. The comparison between the
rij is the distance between atoms i and j, r0 is the equilibrium classical force fields, the polarizable force field (AMOEBA) and
bond length, kθ is the angle-bending force constant, θijk is the the ab initio calculations reported below, together with
angle between the atoms at the pivot atom j, and θ0 is the experimental data, does however suggest that the more negative
equilibrium angle. solvation free energy is a more reasonable value based on the
The remaining interactions between sulfate−water and Ca− properties of the sulfate ion in water that result from this
sulfate were described using a Buckingham potential: choice.
The calculated solvation free energies for the sulfate dianion
C
Ubuck = A e−rij / ρ − are shown in Figure 1 as a function of temperature and
rij 6 (5)
The parameters for these interactions were fitted to reproduce
the experimental solvation free energy and the dissolution free
energy of calcium sulfate crystalline phases, respectively. The
atomic charges (in atomic units) were set to qCa = +2.0, qS =
+1.36, qO = −0.84 and qHw = +0.41, qOw = −0.82. For
completeness, in Table 1 we report all the parameters used for
the classical MD simulations carried out for this work including
those taken from previous work.
C DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article
Table 4. Comparison with Experiment of the Structural Properties of the Crystalline Calcium Sulfate Phases Calculated from
Lattice Dynamicsa
experiment FF1 FF2 AMOEBA
3
gypsum volume (Å ) 488.829 498.087 565.825 492.137
I12/C1 a (Å) 5.674 5.517 5.702 5.886
b (Å) 15.105 15.168 16.174 14.596
c (Å) 6.491 6.656 6.884 6.494
β (deg) 118.513 116.577 116.980 118.091
K (GPa) 43.5 42.3 30.2 53.9
ΔGdiss (kJ/mol) 24.91/27.4257 25.38 25.00
bassanite volume (Å3) 1056.074 1087.560 1182.490 1060.329
I121 a (Å) 12.032 12.169 12.600 11.964
b (Å) 6.927 6.955 7.217 6.800
c (Å) 12.671 12.854 13.008 13.034
β (deg) 90.270 91.435 91.206 90.269
K (GPa) 41.4 37.3 59.1
ΔGdiss (kJ/mol) 20.07/19.11 19.25 53.05
insoluble anhydrite volume (Å3) 305.481 306.810 327.522 257.603
Cmcm a (Å) 6.993 6.908 7.056 6.470
b (Å) 6.245b 6.367 6.460 6.115
c (Å) 6.995b 6.976 7.186 6.511
K (GPa) 54.9 63.8 57.8 162.4
ΔGdiss (kJ/mol) 23.68/26.157 39.82 89.09
soluble anhydrite 1 volume (Å3) 530.856 535.398 570.146 549.439
C222 a (Å) 12.078 11.992 12.331 12.125
b (Å) 6.972 6.924 7.119 7.023
c (Å) 6.304 6.448 6.495 6.453
K (GPa) 59.1 54.3 71.9
ΔGdiss (kJ/mol) 15.31 31.33 80.45
soluble anhydrite 2 volume (Å3) 265.149 267.699 285.073 225.506
P6222 a(Å) 6.969 6.924 7.119 6.470
c (Å) 6.300 6.448 6.495 6.222
K (GPa) 59.1 54.3 150.0
ΔGdiss (kJ/mol) 10.87 31.19 80.45
a
The experimental values are taken from Schofield et al.65 (gypsum), Bezou et al.66 (bassanite, C222), Hawthorne et al.67 (Cmcm), and Lager et al.68
(P6222). The experimental values for the bulk moduli (K) of gypsum and Cmcm are taken from Comodi et al.69 and Schwerdtner et al.,70
respectively. The dissolution free energies are from Wagman et al.,43 unless otherwise specified. bThe cell parameters b and c have been switched
relative to the original experimental data, and the space group changed from Amma to the alternative setting of Cmcm.
solution to a crystalline environment. This problem is field, FF1, overall, and slightly superior in terms of the cell
particularly acute for hydrate salts where the water model is volume and angles (where not right angles). However, for the
optimized primarily for the liquid phase and often fails to anhydrous phases of calcium sulfate, there are substantial errors
capture the different properties of water when embedded in an in the cell parameters, in some cases of up to 7.5%.
ionic solid. This also has an effect on the structural properties of Interestingly, the results for anhydrous calcium sulfate in the
the CaSO4 phases predicted by FF2 that show errors of up to soluble anhydrite 1 structure are quite accurate, while both of
7.5% in the lattice parameters and up to 15.75% in the the other polymorphs show large errors. This suggests that
predicted volume of the solid phases. In this regard, both the further refinement of the AMOEBA parameters would be
AMOEBA force field and the AIMD simulations should be necessary for use in the solid-state under dry conditions.
superior and better able to reproduce the solubilities of all the Specifically, it may be necessary to derive individual pairwise
structures simultaneously. Because of the much larger terms for the Lennard-Jones interactions rather than rely on
computational cost of using these methods, we are unable to combination rules. Beyond structure, the AMOEBA model also
obtain solvation free energies at present, thus preventing the seems to systematically overestimate the mechanical hardness
calculation of the solubilities of mineral phases. However, of phases based on the current parameters, as judged using the
recent work has demonstrated that the determination of free bulk moduli. At present, we cannot comment on the bulk
solubilities for AMOEBA for reasons discussed above.
energies of solvation for ions using AMOEBA49 and AIMD is
However, it should be noted that even first-principles methods
becoming possible.50
have been shown to struggle to yield accurate thermodynamics
When analyzing the performance of the AMOEBA model for
for hydrated mineral phases.51
the structure of calcium sulfate bulk phases, the performance
appears to be somewhat variable. Given that the parameters
were developed with a focus on species in water, it is perhaps 4. ION PAIRING RESULTS
unsurprising that the reproduction of the structure of the Having developed two new candidate classical force fields for
hydrate phases is best, being comparable to the rigid ion force CaSO4 and thoroughly benchmarked them against the
F DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article
Figure 4. Ca−SO4 ion pairing free energy calculated using FF1 (a), FF2 (b), AMOEBA (c), and BLYP-D3 (d) MD as a function of the Ca−S
distance and of the Ca by water coordination number. The free energy maps have been shifted along the y-axis to correct any artifacts in the
calculation of the Ca coordination number due to use of two slightly different switching functions, which can either partially include neighbors in the
second shell or consider only fractional amounts for the water in the first solvation shell.
AMOEBA force field and ab initio MD, where possible, we now AIMD, it is not possible to unambiguously determine the
turn our attention to the ion pairing free energy as a relative free energy of the Ca water coordination states for a
thermodynamic property that is not explicitly fitted during truly isolated Ca2+ ion.
the force field parametrization. Here the free energy landscape In order to better appreciate the differences between the
has been calculated as a function of the calcium−sulfur distance classical force field and the AIMD, it is instructive to look at the
and the calcium−water coordination number. As in our 1D pairing free energy obtained by integrating out the water
previous work on the ion pairing of alkaline earth metals coordination number (Figure 5). Because of the lower
with carbonate,12 the analysis of the 2D free energy maps computational cost of the classical and polarizable MD
(Figure 4) provides a clear picture of the ion pair formation. simulations, it is feasible to use large enough cells to converge
Although the finer details of the free energy profiles are the long-range tail of the free energy and align them onto the
different, all the methods used here clearly show that the Ca2+ analytic solution for the pairing free energy of two point
ion in solution has a preferred water coordination number of 7 charges interacting via a screened electrostatic potential
and that the formation of a contact ion pair (CIP) involves a (dashed line). However, for the AIMD simulations, we could
progressive desolvation of the cation to be surrounded by six only explore the contact ion pair (CIP) and solvent-shared ion
and five water molecules to form a monodentate and bidentate pair (SSHIP) states, and therefore, we aligned the free energy
ion pair, respectively. The results reported here are qualitatively based on the position of the SSHIP minimum. All methods
consistent with previous DFT calculations of the Ca−Cl ion used in this work clearly indicate that the SSHIP state has a
pair formation by Baer and Mundy, which also considered the lower free energy than the CIP and has similar barriers for the
hydration of the Ca2+ ion alone.52 In both the BLYP-D3 results formation of the CIP and for the dissociation of the complex.
of this work and the force fields of the present study, the However, there are significant differences between the
preferred Ca by water coordination number is also 7. However, predicted stability of the CIP, with FF1 and the AMOEBA
Baer and Mundy52 found a Ca2+ coordination number of 6 to calculation indicating that the CIP is only marginally higher in
be more stable than 8 as the secondary favored hydration state, energy than the SSHIP (∼2 kJ/mol), while BLYP-D3 shows a
while our force field calculations suggest the reverse (Figure larger difference (∼7 kJ/mol), and FF2 only has a shallow
SI4). Due to the limited range of Ca−S distance sampled in our minimum for the CIP, which is approximately 15 kJ/mol less
G DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article
values for the CaCO3 system (29.3 J/mol/K)58 and other formation of the contact ion pair might be a prerequisite to
experimental estimates of the water entropy change due its subsequent formation of any pre- or postnucleation species.
localization in crystalline hydrated phases (32/33.5 J/mol/ Future work is required to probe the further ion association
K).59,60 This phenomenon of entropy-driven association, which processes of calcium sulfate, and this is now possible based on
has been experimentally observed in a variety of systems61 and the force field (FF1) developed within this study.
■
extensively studied with computational methods,62,63 may have
important consequences for the thermodynamic stability and ASSOCIATED CONTENT
abundance of prenucleation clusters, beyond that expected * Supporting Information
S
from classical nucleation theory. Among the many computa-
The Supporting Information is available free of charge on the
tional works in the literature that deal with the formation of ion
ACS Publications website at DOI: 10.1021/acs.jpcc.7b09820.
pairs between ionic species, a few recent ones have started
looking at the contribution of water and entropy to the pairing Comparison between radial pair distribution functions
free energy.62,63 In particular, the works by Ballard and computed with the potential described in this work and
Dellago62 and by Shi and Beck,63 which both focused on alkali literature ones; 3D water density map computed using a
halide ion pairs, conclude that the ion association is driven by force field from the literature; self-diffusion coefficient as
an increase in entropy due to the release of water molecules a function of the box size; molecular arrangement in the
from the ions’ coordination shell into the solution. This picture various states of the ion pair formation; all the pairing
is perfectly consistent with the results presented here for CaSO4 free energy curves for the two force fields discussed in
and with our previous work on the formation of CaCO3 ion this work; table with the dissolution free energy and
pairs, where ion pair formation is driven by an increase in solubilities of the calcium sulfate phases (PDF)
■
entropy proportional to the number of water molecules
released into solution and opposed by a positive enthalpic AUTHOR INFORMATION
term. The unfavorable enthalpy change can be ascribed to the
loss of electrostatic interactions with the released water that is Corresponding Author
not compensated by the cation−anion electrostatic interaction. *E-mail: p.raiteri@curtin.edu.au. Phone: + 61 8 9266 2687.
ORCID
5. CONCLUSIONS Paolo Raiteri: 0000-0003-0692-0505
In conclusion, we have developed and tested two rigid ion force Author Contributions
fields by targeting the two available experimental estimates of All authors contributed to performing the calculations,
the hydration free energy of the sulfate anion and the solubility analyzing the data, and writing the manuscript. All authors
of the most common hydrated solid phases of CaSO4. Because have given approval to the final version of the manuscript.
the two available hydration free energies differ by about 100 kJ/
mol, the two force fields gave markedly different predictions for Notes
The authors declare no competing financial interest.
■
most of the structural and dynamic properties computed in this
work. To discriminate between these force field models, we
have additionally performed calculations using both a more ACKNOWLEDGMENTS
sophisticated, polarizable force field using the AMOEBA model E.H.B. would like to acknowledge the contribution of an
and ab initio molecular dynamics with a dispersion-corrected Australian Government Research Training Program Scholar-
GGA functional. Through comparison of the two rigid ion force ship in supporting this research, and P.R. and J.D.G. thank the
fields against the results of the other methods, as well as other Australian Research Council for financial support through the
literature data, it is possible to identify FF1, which gives the Discovery Programme (FT130100463, DP160100677). This
more exothermic hydration free energy for sulfate, as being the work was also supported by resources provided by the Pawsey
superior representation of aqueous calcium sulfate systems. In Supercomputing Centre and National Computational Infra-
particular, the distribution of water around the sulfate ion is structure with funding from the Australian Government and the
markedly different between FF2 and all of the other results. Government of Western Australia.
■
Having identified the best rigid ion force field model for the
system, it was then possible to examine the ion pairing of REFERENCES
calcium sulfate in aqueous solution. Based on this force field, as (1) Azimi, G.; Papangelakis, V. G.; Dutrizac, J. E. Modelling of
well as AMOEBA and BLYP-D3, there is a consistent picture of Calcium Sulphate Solubility in Concentrated Multi-Component
the ion pairing process. Consistent with previous estimates of Sulphate Solutions. Fluid Phase Equilib. 2007, 260 (2), 300−315.
the ion pairing free energy, we find exothermic binding under (2) Van Driessche, A. E. S.; Benning, L. G.; Rodriguez-Blanco, J. D.;
standard conditions of approximately −13 kJ/mol. By Ossorio, M.; Bots, P.; García-Ruiz, J. M. The Role and Implications of
decomposing the free energy into the enthalpy and entropy, Bassanite as a Stable Precursor Phase to Gypsum Precipitation. Science
based on the temperature dependence, it is found that ion 2012, 336 (6077), 69−72.
pairing is entropically driven, which is consistent with the (3) Pina, C. M. Nanoscale Dissolution and Growth on Anhydrite
release of an average of three water molecules observed during Cleavage Faces. Geochim. Cosmochim. Acta 2009, 73 (23), 7034−7044.
binding. The contact ion pair is found to be between 2 and 8 (4) Demichelis, R.; Raiteri, P.; Gale, J. D.; Quigley, D.; Gebauer, D.
Stable Prenucleation Mineral Clusters Are Liquid-Like Ionic Polymers.
kJ/mol less stable than the solvent-shared state, with a barrier
Nat. Commun. 2011, 2, 590.
of at least 14 kJ/mol to reach a distance at which direct (5) Gebauer, D.; Voelkel, A.; Coelfen, H. Stable Prenucleation
coordination between the ions occurs. Therefore, it appears Calcium Carbonate Clusters. Science 2008, 322 (5909), 1819−1822.
unlikely that the contact state plays a substantial role in the (6) Gebauer, D.; Kellermeier, M.; Gale, J. D.; Bergström, L.; Cölfen,
initial ion pairing. However, the balance between these two H. Pre-Nucleation Clusters as Solute Precursors in Crystallisation.
states may change as further ion association occurs, and so Chem. Soc. Rev. 2014, 43 (7), 2348.
I DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article
(7) Navrotsky, A. Energetic Clues to Pathways to Biomineralization: (29) Allan, N. L.; Rohl, A.; Gay, D.; Catlow, C. R. A.; Davey, R.;
Precursors, Clusters, and Nanoparticles. Proc. Natl. Acad. Sci. U. S. A. Mackrodt, W. Calculater Bulk and Surface Properties of Sulfates.
2004, 101 (33), 12096−12101. Faraday Discuss. 1993, 95, 273−280.
(8) Raiteri, P.; Gale, J. D. Water Is the Key to Nonclassical (30) VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.;
Nucleation of Amorphous Calcium Carbonate. J. Am. Chem. Soc. 2010, Chassaing, T.; Hutter, J. Quickstep: Fast and Accurate Density
132 (49), 17623−17634. Functional Calculations Using a Mixed Gaussian and Plane Waves
(9) De Yoreo, J. J.; Vekilov, P. G. Principles of Crystal Nucleation Approach. Comput. Phys. Commun. 2005, 167 (2), 103−128.
and Growth. Rev. Mineral. Geochem. 2003, 54, 57−93. (31) Krack, M. Pseudopotentials for H to Kr Optimized for Gradient-
(10) Erdemir, D.; Lee, A. Y.; Myerson, A. S. Nucleation of Crystals Corrected Exchange-Correlation Functionals. Theor. Chem. Acc. 2005,
From Solution: Classical and Two-Step Models. Acc. Chem. Res. 2009, 114 (1−3), 145−152.
42 (5), 621−629. (32) Becke, A. Density-Functional Exchange-Energy Approximation
(11) Raiteri, P.; Gale, J. D.; Quigley, D.; Rodger, P. M. Derivation of with Correct Asymptotic Behavior. Phys. Rev. A: At., Mol., Opt. Phys.
an Accurate Force-Field for Simulating the Growth of Calcium 1988, 38 (6), 3098−3100.
Carbonate From Aqueous Solution: a New Model for the Calcite− (33) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and
Water Interface. J. Phys. Chem. C 2010, 114 (13), 5997−6010. Accurate Ab Initio Parametrization of Density Functional Dispersion
(12) Raiteri, P.; Demichelis, R.; Gale, J. D. Thermodynamically Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010,
Consistent Force Field for Molecular Dynamics Simulations of 132 (15), 154104−154120.
Alkaline-Earth Carbonates and Their Aqueous Speciation. J. Phys. (34) Bankura, A.; Karmakar, A.; Carnevale, V.; Chandra, A.; Klein, M.
Chem. C 2015, 119 (43), 24447−24458. L. Structure, Dynamics, and Spectral Diffusion of Water From First-
(13) Laio, A.; Parrinello, M. Escaping Free-Energy Minima. Proc. Principles Molecular Dynamics. J. Phys. Chem. C 2014, 118 (50),
Natl. Acad. Sci. U. S. A. 2002, 99 (20), 12562−12566. 29401−29411.
(14) Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular (35) VandeVondele, J.; Hutter, J. An Efficient Orbital Trans-
Dynamics. J. Comput. Phys. 1995, 117 (1), 1−19. formation Method for Electronic Structure Calculations. J. Chem. Phys.
(15) Martyna, G.; Tobias, D.; Klein, M. L. Constant Pressure 2003, 118 (10), 4365−4369.
Molecular Dynamics Algorithms. J. Chem. Phys. 1994, 101, 4177. (36) Shao, Y.; Molnar, L. F.; Jung, Y.; Kussmann, J.; Ochsenfeld, C.;
(16) Wu, Y.; Tepper, H.; Voth, G. A. Flexible Simple Point-Charge Brown, S. T.; Gilbert, A. T. B.; Slipchenko, L. V.; Levchenko, S. V.;
Water Model with Improved Liquid-State Properties. J. Chem. Phys. O’Neill, D. P.; et al. Advances in Methods and Algorithms in a Modern
2006, 124 (2), 024503. Quantum Chemistry Program Package. Phys. Chem. Chem. Phys. 2006,
(17) Kirkwood, J. G. Statistical Mechanics of Fluid Mixtures. J. Chem. 8 (27), 3172−3191.
Phys. 1935, 3 (5), 300. (37) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Perspective on
(18) Zwanzig, R. High-Temperature Equation of State by a Foundations of Solvation Modeling: the Electrostatic Contribution to
Perturbation Method. I. Nonpolar Gases. J. Chem. Phys. 1954, 22 the Free Energy of Solvation. J. Chem. Theory Comput. 2008, 4 (6),
(8), 1420. 877−887.
(19) Hummer, G.; Pratt, L. R.; Garcia, A. Ion Sizes and Finite-Size (38) Wu, Y.; Chen, H.; Wang, F.; Paesani, F.; Voth, G. A. An
Corrections for Ionic-Solvation Free Energies. J. Chem. Phys. 1997, 107 Improved Multistate Empirical Valence Bond Model for Aqueous
(21), 9275−9277. Proton Solvation and Transport. J. Phys. Chem. B 2008, 112 (2), 467−
(20) Kastenholz, M. A.; Hünenberger, P. H. Computation of 482.
(39) Marcus, Y. Ion Solvation; Wiley, 1985.
Methodology-Independent Ionic Solvation Free Energies From
(40) Marcus, Y. A Simple Empirical Model Describing the
Molecular Simulations. II. the Hydration Free Energy of the Sodium
Thermodynamics of Hydration of Ions of Widely Varying Charges,
Cation. J. Chem. Phys. 2006, 124 (22), 224501.
Sizes, and Shapes. Biophys. Chem. 1994, 51 (2), 111−127.
(21) Barducci, A.; Bussi, G.; Parrinello, M. Well-Tempered
(41) Marcus, Y. Ions in Solution and Their Solvation; John Wiley &
Metadynamics: a Smoothly Converging and Tunable Free-Energy
Sons, Inc., 2016.
Method. Phys. Rev. Lett. 2008, 100 (2), 20603. (42) Cramer, C. J.; Truhlar, D. G. A Universal Approach to Solvation
(22) Raiteri, P.; Laio, A.; Gervasio, F. L.; Micheletti, C.; Parrinello, M. Modeling. Acc. Chem. Res. 2008, 41 (6), 760−768.
Efficient Reconstruction of Complex Free Energy Landscapes by (43) Ohtaki, H.; Radnai, T. Structure and Dynamics of Hydrated
Multiple Walkers Metadynamics. J. Phys. Chem. B 2006, 110 (8), Ions. Chem. Rev. 1993, 93 (3), 1157−1204.
3533−3539. (44) Cannon, W.; Pettitt, B.; McCammon, J. A. Sulfate Anion in
(23) Tribello, G. A.; Bonomi, M.; Branduardi, D.; Camilloni, C.; Water - Model Structural, Thermodynamic, and Dynamic Properties. J.
Bussi, G. Computer Physics Communications. Comput. Phys. Commun. Phys. Chem. 1994, 98 (24), 6225−6230.
2014, 185 (2), 604−613. (45) Pegado, L.; Marsalek, O.; Jungwirth, P.; Wernersson, E.
(24) Ponder, J. W.; Wu, C.; Ren, P.; Pande, V. S.; Chodera, J. D.; Solvation and Ion-Pairing Properties of the Aqueous Sulfate Anion:
Schnieders, M. J.; Haque, I.; Mobley, D. L.; Lambrecht, D. S.; DiStasio, Explicit Versus Effective Electronic Polarization. Phys. Chem. Chem.
R. A., Jr.; et al. Current Status of the AMOEBA Polarizable Force Phys. 2012, 14 (29), 10248.
Field. J. Phys. Chem. B 2010, 114 (8), 2549−2564. (46) Williams, C. D.; Carbone, P. A Classical Force Field for
(25) Gale, J. D. GULP: Capabilities and Prospects. Z. Kristallogr. - Tetrahedral Oxyanions Developed Using Hydration Properties: the
Cryst. Mater. 2005, 220 (5−6), 552−554. Examples of Pertechnetate (TcO4−) And Sulfate (SO42−). J. Chem.
(26) Ponder, J. W.; Richards, F. M. An Efficient Newton-Like Phys. 2015, 143 (17), 174502−174509.
Method for Molecular Mechanics Energy Minimization of Large (47) Vchirawongkwin, V.; Rode, B. M.; Persson, I. Structure and
Molecules. J. Comput. Chem. 1987, 8, 1016. Dynamics of Sulfate Ion in Aqueous Solution-an Ab Initio QMCF MD
(27) Piquemal, J.-P.; Perera, L.; Cisneros, G. A. S.; Ren, P.; Pedersen, Simulation and Large Angle X-Ray Scattering Study. J. Phys. Chem. B
L. G.; Darden, T. A. Towards Accurate Solvation Dynamics of 2007, 111 (16), 4150−4155.
Divalent Cations in Water Using the Polarizable Amoeba Force Field: (48) Yuan-Hui, L.; Gregory, S. Diffusion of Ions in Sea Water and in
From Energetics to Structure. J. Chem. Phys. 2006, 125 (5), 054511− Deep-Sea Sediments. Geochim. Cosmochim. Acta 1974, 38 (5), 703−
054517. 714.
(28) Lambrecht, D. S.; Clark, G. N. I.; Head-Gordon, T.; Head- (49) Mohamed, N. A.; Bradshaw, R. T.; Essex, J. W. Evaluation of
Gordon, M. Exploring the Rich Energy Landscape of Sulfate. J. Phys. Solvation Free Energies for Small Molecules with the AMOEBA
Chem. A 2011, 115, 11438−11454. Polarizable Force Field. J. Comput. Chem. 2016, 37 (32), 2749−2758.
J DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article
(50) Bell, D. R.; Qi, R.; Jing, Z.; Xiang, J. Y.; Mejias, C.; Schnieders, (70) Schwerdtner, W. M.; Tou, J.; Hertz, P. B. Elastic Properties of
M. J.; Ponder, J. W.; Ren, P. Calculating Binding Free Energies of Single Crystals of Anhydrite. Can. J. Earth Sci. 1965, 2 (6), 673−683.
Host-Guest Systems Using the AMOEBA Polarizable Force Field.
Phys. Chem. Chem. Phys. 2016, 18 (44), 30261−30269.
(51) Demichelis, R.; Raiteri, P.; Gale, J. D.; Dovesi, R. Examining the
Accuracy of Density Functional Theory for Predicting the Thermody-
namics of Water Incorporation Into Minerals: the Hydrates of Calcium
Carbonate. J. Phys. Chem. C 2013, 117 (34), 17814−17823.
(52) Baer, M. D.; Mundy, C. J. Local Aqueous Solvation Structure
Around Ca 2+During Ca 2+···Cl − Pair Formation. J. Phys. Chem. B
2016, 120 (8), 1885−1893.
(53) Fuoss, R. M.; Kraus, C. A. Properties of Electrolytic Solutions.
III. the Dissociation Constant. J. Am. Chem. Soc. 1933, 55 (3), 1019−
1028.
(54) Ainsworth, R. G. Dissociation Constant of Calcium Sulphate
From 25 to 50°C. J. Chem. Soc., Faraday Trans. 1 1973, 69 (0), 1028.
(55) Wachter, R.; Riederer, K. Properties of Dilute Electrolyte-
Solutions From Calorimetric Measurements. Pure Appl. Chem. 1981,
53 (7), 1301−1312.
(56) Bester-Rogac, M. Determination of the Limiting Conductances
and the Ion-Association Constants of Calcium and Manganese Sulfates
in Water From Electrical Conductivity Measurements. Acta Chim. Slov.
2008, 55 (1), 201−208.
(57) Nordstrom, D. K.; Plummer, L. N.; Langmuir, D.; Busenberg,
E.; May, H. M.; Jones, B. F.; Parkhurst, D. L. Revised Chemical
Equilibrium Data for Major Water−Mineral Reactions and Their
Limitations. In Chemical Modeling of Aqueous Systems II; ACS
Symposium Series; American Chemical Society: Washington, DC,
2009; Vol. 416, pp 398−413.
(58) Kellermeier, M.; Raiteri, P.; Berg, J. K.; Kempter, A.; Gale, J. D.;
Gebauer, D. Entropy Drives Calcium Carbonate Ion Association.
ChemPhysChem 2016, 17 (21), 3535−3541.
(59) Dunitz, J. D. The Entropic Cost of Bound Water in Crystals and
Biomolecules. Science 1994, 264 (5159), 670−670.
(60) Konigsberger, E.; Konigsberger, L.; Gamsjager, H. Low-
Temperature Thermodynamic Model for the System Na2Co3-
MgCO3-CaCO3-H2O. Geochim. Cosmochim. Acta 1999, 63 (19−20),
3105−3119.
(61) De Visscher, A.; Vanderdeelen, J.; Königsberger, E.; Churagulov,
B. R.; Ichikuni, M.; Tsurumi, M. IUPAC-NIST Solubility Data Series.
95. Alkaline Earth Carbonates in Aqueous Systems. Part 1.
Introduction, Be and Mg. J. Phys. Chem. Ref. Data 2012, 41 (1),
013105.
(62) Ballard, A. J.; Dellago, C. Toward the Mechanism of Ionic
Dissociation in Water. J. Phys. Chem. B 2012, 116 (45), 13490−13497.
(63) Shi, Y.; Beck, T. Deconstructing Free Energies in the Law of
Matching Water Affinities. J. Phys. Chem. B 2017, 121 (9), 2189−2201.
(64) Krynicki, K.; Green, C. D.; Sawyer, D. W. Pressure and
Temperature Dependence of Self-Diffusion in Water. Faraday Discuss.
Chem. Soc. 1978, 66, 199−208.
(65) Schofield, P. F.; Knight, K. S.; Stretton, I. C. Thermal Expansion
of Gypsum Investigated by Neutron Powder Diffraction. Am. Mineral.
1996, 81 (7−8), 847−851.
(66) Bezou, C.; Nonat, A.; Mutin, J. C.; Christensen, A. N.;
Lehmann, M. S. Investigation of the Crystal Structure of Γ-CaSO4,
CaSO4· 0.5 H2O, and CaSO4· 0.6 H2O by Powder Diffraction
Methods. J. Solid State Chem. 1995, 117, 165−176.
(67) Hawthorne, F. C.; Ferguson, R. B. Anhydrous Sulphates; II,
Refinement of the Crystal Structure of Anhydrite. Can. Mineral. 1975,
13, 289−292.
(68) Lager, G. A.; Armbruster, T.; Rotella, F. J.; Jorgensen, J. D.;
Hinks, D. G. A Crystallographic Study of the Low-Temperature
Dehydration Products of Gypsum, CaSO4· 2h2O - Hemihydrate
CaSO4· 0.50h2O, and Γ-CaSO4. Am. Mineral. 1984, 69 (9−10), 910−
919.
(69) Comodi, P.; Nazzareni, S.; Zanazzi, P. F.; Speziale, S. High-
Pressure Behavior of Gypsum: a Single-Crystal X-Ray Study. Am.
Mineral. 2008, 93 (10), 1530−1537.
K DOI: 10.1021/acs.jpcc.7b09820
J. Phys. Chem. C XXXX, XXX, XXX−XXX