Yu
Shi
a,
Stephen T.
Lam
b and
Thomas L.
Beck
*c
aDepartment of Chemistry, University of Cincinnati, Cincinnati, OH 45221, USA
bDepartment of Chemical Engineering, University of Massachusetts Lowell, MA 01854, USA
cNational Center for Computational Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA. E-mail: becktl@ornl.gov
First published on 15th June 2022
With dual goals of efficient and accurate modeling of solvation thermodynamics in molten salt liquids, we employ ab initio molecular dynamics (AIMD) simulations, deep neural network interatomic potentials (NNIP), and quasichemical theory (QCT) to calculate the excess chemical potentials for the solute ions Na+ and Cl− in the molten NaCl liquid. NNIP-based molecular dynamics simulations accelerate the calculations by 3 orders of magnitude and reduce the uncertainty to 1 kcal mol−1. Using the Density Functional Theory (DFT) level of theory, the predicted excess chemical potential for the solute ion pair is −178.5 ± 1.1 kcal mol−1. A quantum correction of 13.7 ± 1.9 kcal mol−1 is estimated via higher-level quantum chemistry calculations, leading to a final predicted ion pair excess chemical potential of −164.8 ± 2.2 kcal mol−1. The result is in good agreement with a value of −163.5 kcal mol−1 obtained from thermo-chemical tables. This study validates the application of QCT and NNIP simulations to the molten salt liquids, allowing for significant insights into the solvation thermodynamics crucial for numerous molten salt applications.
Therefore molecular dynamics (MD) simulations have been exploited to calculate molten salt properties. Classical simulations are efficient for modelling systems over timescales on the order of nanoseconds in order to make viable predictions. Empirical force-field molecular dynamics (FFMD) simulations with classical potentials such as the Born–Mayer–Huggins–Tosi–Fumi (BMHTF) rigid-ion potential12–15 have been demonstrated to lack full predictive capabilities due to the lack of many-body interactions that influence local structure. Polarizable ion models (PIM) developed in conjunction with quantum mechanical calculations have led to significant improvements in modeling structure, thermodynamics, and dynamic properties.16,17
Ab initio molecular dynamics (AIMD) simulations have been shown to be capable of accurately capturing local structure and solute chemistry in comparison with experiments.18–26 AIMD is computationally expensive, however, and therefore it is difficult to access long time and large length scales to predict the dynamic and thermal properties. Recently, state-of-the-art artificial neural network-based interatomic potential molecular dynamics simulations (NNIP-MD) have been demonstrated as a promising computational tool to explore the underlying physics of the high-dimensional molten salt compositions by simulating 104 atoms on the timescale of nanoseconds with an accuracy at the level of density functional theory (DFT). It has been demonstrated that the NNIP-MD studies27–30 are capable of accurately predicting molten salt structure, heat capacity, self-diffusion coefficients, thermal conductivity, electrical conductivity, viscosity and the melting/freezing point in comparison with experimental measurements.
Understanding phase behavior remains a great challenge at the forefront of molten salt research. In addition, during reactor operation, there is continuous evolution due to processes such as transmutations, fission/corrosion product generation, gas bubble formation, precipitation of insoluble species and other reactions.31,32 A challenge facing the molten salt research community is to directly model these properties starting from phase diagrams, but many of the phase diagrams have never been constructed.
The CALPHAD method is a principal method for phase diagram and molten salt database development.32–37 Experimental thermodynamic data (from phase equilibria measurements and activities of solution species) and/or AIMD simulations are used to empirically fit models and then predict stable phases at different temperatures or compositions. As mentioned above it would be difficult, expensive, and time-consuming to obtain data for high-dimensional salt systems through either experiments or AIMD simulations.19,38
The chemical potential is central for understanding molten salt thermodynamic properties and predicting phase behavior. Studies of the chemical potential are limited due to the above-mentioned difficulties, however. The classical PIM potential force field has been used to calculate the activity coefficient ratios for lanthanide cations (e.g. U3+ to Y3+ in the LiCl/KCl eutectic mixture39) and the free energy change for the reaction of the Eu3+/Eu2+ redox couple in molten KCl.40,41 The Widom particle insertion method has been employed with AIMD simulations to calculate the chemical potential and the solubility of the sodium atom in molten NaCl.42
The cutting-edge NNIP-MD methods hold significant promise to play a vital role in exploring the thermodynamic properties since they provide quantum-level accuracy with efficiency similar to classical simulations. In the present work, molten NaCl is chosen as a prototype system to demonstrate the validity of the NNIP-based quasichemical theory (QCT) calculations, which is shown to be an efficient approach for directly calculating the solute ion solvation free energy via molecular dynamics simulations.43–45 To our knowledge, this is the first application of deep learning methods to the calculation of excess thermodynamic properties of molten salts.
μX = μidX + μexX | (1) |
The solute(X)-solvent interaction energy is defined as εX = Usolution − Usolvent − Usolute, where U is the system total potential energy. Then the excess chemical potential is given by
μexX = −kTln〈e−εX/kT〉0 | (2) |
μexX = −kTln〈e−Mλ/kT〉0+kTln〈e−Mλ/kT〉εX−kTln〈e−εX/kT〉Mλ | (3) |
The LR contribution can also be written as43,44
μexLR = kTln〈eεX/kT〉Mλ+εX | (4) |
fλ(γ)=γ3Mλ, | (5) |
The repulsive potential Mλ(r) is a half-harmonic potential
(6) |
The PK contribution is then expressed as
(7) |
(8) |
For the long-ranged contribution, a cumulant expansion to second order yields the (uncoupled) expression
(9) |
(10) |
The regularization47 due to the repulsive potential Mλ (that pushes the solvent molecules/ions away from the solute) produces near-Gaussian statistics for εX, and thus the two fluctuation terms of the uncoupled (eqn (9)) and the coupled (eqn (10)) samplings44 are of close magnitude. Consequently, the average of the two mean terms gives a relatively accurate approximation for the LR contribution.
Alternatively, at a value for which the two interaction energy distribution functions are equal,43
(11) |
The initial configurations were generated during classical molecular dynamics simulations using the GROMACS 4.5.5 package,54 and the ions were modeled with the OPLS-AA force-field.55 With 1 fs as the time step, all classical force-field based simulations were run for 1.5 ns in the NVT ensemble after 1 ns of equilibration in the NPT ensemble. The temperature was set at 1150 K for the classical simulations. For the AIMD simulations, parameters were chosen based on previous studies that accurately predicted local structure, standard reduction potential and sodium solubility in molten NaCl.42 This includes the dual basis sets of Gaussian-type orbitals (double zeta bases, DZVP-GTH) and plane waves with a 600 Ry cutoff.56 Atomic cores were modeled with the Goedecker-Teter-Hutter pseudopotentials (GTH).57 The Perdew–Burke–Ernzerhof (PBE)58 functionals were used for all atoms in the system, and the D2 dispersion corrections59,60 were utilized. D2 dispersion corrections were also employed in a recent NNIP study on molten NaCl liquid.28 While more advanced dispersion methods such as D3 could be used,61 systematic improvement in predictions across different properties is difficult to achieve, due to the semi-empirical nature of most computationally efficient DFT-based dispersion methods. In any case, from density predictions in molten NaCl,42 we find that additional corrections based on higher-level theory are likely necessary to achieve chemical accuracy.
The Ewald potential45,62,63 was applied for the electrostatic interactions under periodic boundary conditions (PBC), and a Nosé-Hoover thermostat chain64 of length 3 was coupled to each ion to maintain a temperature of 1150 K for all the NVT ensemble simulations. Due to the requirement of many simulations along the thermodynamic integration paths to grow the nano-scale cavities, we were not able to employ the path integral formalism for incorporating quantum effects.65 These corrections are expected to be small at such a high temperature, however.
The time step was taken as 0.5 fs. For the LR contribution, two 25 ps simulations (50000 configurations, coupled and uncoupled) were implemented. This is found to be sufficient for calculating ensemble-average energies based on previous AIMD studies with molten salts.26–28,42 We performed 9-step integrations for the PK and IS calculations, and we ran simulations for 10 ps (20000 configurations) for each step.
(12) |
From the AIMD simulations, we generated 20000 configurations for each of the first 9 steps of the thermodynamic integration. For the last step, where γ = 1, 50000 configurations were collected from the two coupled simulations (IS of Na+ and Cl−) and one uncoupled simulation (PK), respectively. We trained a model for each step of the TI process (IS and PK, 10 models in total) and another 2 models for the interaction energy calculations over the uncoupled and coupled configurations of the LR simulations, respectively. We also tried to obtain a single NNIP model by using all of the collected AIMD data. This all-in-one model produced an error of magnitude about 3 kcal mol−1 for the solute–solvent interaction energy compared with the AIMD calculation, and a similar error in the free energy. Due to this relatively large error, we did not utilize the all-in-one method further.
Next, as shown in Panel (b) and Panel (c) of Fig. 1, the overlapping of radial distribution functions (RDF or g(r)) indicates that NNIP sufficiently reproduces the local structure predicted by the AIMD simulations and uncovers the oscillating structures at larger distances as well. Additionally, the RDF of the Na+–Cl− pair in Panel (b) exhibits a first maximum position at 2.77 Å and a first minimum position at 4.27 Å, which are close to those reported recently27,28 using NNIP training employing a different protocol. This indicates that the presence of a cavity with a radius of 4.0 Å does not significantly affect the average liquid structure in the nearby bulk.
Just outside of the cavity, it is observed that there is a slightly higher density of the Cl− ions than the Na+ cations. Integrating to a distance of 5.5 Å the coordination number of Cl− is 0.7 higher than that of Na+. This leads to a dipole layer in the vicinity of a cavity of 4 Å size. When the solute ion is present at the center of the cavity as shown in Panel (c), however, both solvent ions exhibit charge-symmetrical behavior, as evidenced by the overlaps in the IS RDFs.
The LR contribution is estimated from the two distributions of solute–solvent ion interaction energies, which are calculated over configurations from uncoupled and coupled simulations. As shown in Panel (a) of Fig. 2 for the solute Na+, the mean uncoupled interaction energy is −125.62 kcal mol−1 with a fluctuation contribution (eqn (9)) of 28.2 kcal mol−1, while the mean coupled interaction energy is −180.81 kcal mol−1 with a fluctuation contribution of 21.3 kcal mol−1. The difference between the two fluctuation terms leads to an error of 3.5 kcal mol−1 for the LR free energy contribution, indicating that the Gaussian approximation is inappropriate and a larger cavity is required to observe Gaussian behaviour as was previously seen in the calculation of the hydration free energy of the Na+ ion in water.44,45
Fig. 2 The process of excess chemical potential calculation for the systems of solute Na+ and Cl− ions with 256 solvent ion pairs. Panel (a) and Panel (b) are for the long-ranged (LR) contributions to the solvation free energy of the solute ions Na+ and Cl−, respectively. The logarithms of the distribution of interaction energies from uncoupled (right, blue) and coupled (left, black) configurations are presented with the mean value ε and standard deviation σ. The right dashed curves and the left dash-dotted curves are the Gaussian fit to both distributions. Panel (c) is for the packing (PK) and minus the Inner-Shell (IS) cumulative contributions for both solute ions. Along with the expansion of the cavity with(IS)/without(PK) the solute ions at the center, the coupling parameter γ increases from 0 to 1. The IS and PK contributions are listed in Table 1. Panel (d) is the distributions of interaction energy corrections εDFT − εMP2 for the solute ion with 7 and 17 solvent ion pairs. The sampling is over 400 cluster configurations carved from DFT simulation trajectories. The DFT calculation is under Periodic Boundary Conditions (PBC) in a cell of size 25.4 Å. |
As discussed in the theoretical methods section above, however, the long-range contribution is equal to the interaction energy at the intersection point of the two distributions. Thus we estimate the long-ranged contribution for Na+ as −156.0 kcal mol−1. In Panel (b) of Fig. 2, the distributions for the Cl− ion exhibit more closely Gaussian behavior since the fluctuation terms are closer: 25.2 kcal mol−1 for uncoupled simulation with a mean value of 84.05 kcal mol−1 and 27.6 kcal mol−1 for coupled simulation with mean value of 21.40 kcal mol−1.
Based on the estimation of the intersection point of the two energy distributions, we estimate the LR contribution for Cl− ion as 54.0 kcal mol−1. The dramatic increase of the LR contribution (including the sign) for the anion is attributed to the electrostatic potential at the center of the cavity.45 Assuming that the electrostatic interaction dominates the interactions between the center solute ion and solvent ions outside the cavity of 4 Å, the average energies of both Na+ and Cl− ions over the uncoupled configurations gives an estimate of the electrostatic potential at the center of the cavity as −4.55 Volt, relative to the average potential in the bulk region of molten salt liquids.
Multipole electrostatic moment analysis of the cavity-water interfacial potential69 reveals that there is a potential shift of −3.96 Volt from the liquid water bulk phase to the cavity center, which is primarily attributed to the water molecular Bethe potential (quadrupole) contribution. Since the current work focuses on the excess free energy for the solute ion pair (in which case the interfacial potential contributions cancel exactly), the Bethe potential for liquid NaCl is not discussed further here.
The numerical results for the IS and PK contributions are shown in Panel (c) of Fig. 2 by presenting the cumulative work computed during the NNIP-MD simulations. Both Na+ and Cl− ions share the same PK contribution (24.2 kcal mol−1), while the IS contribution to the Na+ solvation free energy is −44.7 kcal mol−1 and that for the Cl− ion is −43.6 kcal mol−1.
The numerical results for both the Na+ and Cl− ions are listed in Table 1. The finite-size correction term is due to the artificial effect of the Ewald potential on the free energy. It is shown to be , where Q is the net charge in unit of elementary charge e, L is the box size, ξ = −2.837297 for a cubic lattice and = 332.063301 kcal mol−1 Å−1 × 102.43 The summation of the excess free energy for both solute ions with 256 solvent ion pairs is −178.5 kcal mol−1, which is converged to within the standard error of the mean (SEM) of ±1.1 kcal mol−1 as the system size increases to 512 ion pairs. The SEM in parentheses is calculated using the block-average method (10 blocks) over the production configurations. To the best of our knowledge, this is the first prediction of the absolute solvation free energy for ions in a molten salt using deep learning techniques.
Solute | μ PKX | μ ISX | μ LRX | FS | μ exX | AIMD | NNIP-MD |
---|---|---|---|---|---|---|---|
Na+ | 25.4(0.2) | −44.8(0.2) | −142.8(0.7) | −28.8 | −191.0(0.8) | ||
Cl− | 25.4(0.2) | −43.8(0.2) | 63.6(0.7) | −28.8 | 16.4(0.7) | ||
NaCl-64 | −174.6(1.1) | 1.5 × 10−3 | 5.5 × 10−7 | ||||
Na+ | 24.5(0.3) | −44.6(0.2) | −149.2(0.7) | −23.1 | −192.4(0.8) | ||
Cl− | 24.5(0.3) | −43.7(0.2) | 58.0(0.7) | −23.1 | 15.7(0.8) | ||
NaCl-128 | −176.7(1.1) | 3.6 × 10−3 | 2.9 × 10−7 | ||||
Na+ | 24.2(0.2) | −44.7(0.2) | −156.0(0.7) | −18.4 | −194.9(0.8) | ||
Cl− | 24.2(0.2) | −43.4(0.2) | 54.0(0.7) | −18.4 | 16.4(0.8) | ||
NaCl-256 | −178.5(1.1) | 4.0 × 10−3 | 2.7 × 10−7 | ||||
Na+ | 24.3(0.2) | −44.4(0.2) | −158.8(0.7) | −14.7 | −193.6(0.8) | ||
Cl− | 24.3(0.2) | −43.6(0.2) | 48.4(0.7) | −14.7 | 14.4(0.8) | ||
NaCl-512 | −179.1(1.1) | 6.6 × 10−3 | 2.5 × 10−7 |
The computational cost by AIMD and NNIP-MD are shown in the last columns of Table 1, where it is shown that the efficiency of the NNIP-MD simulation is about 3 orders greater than that of AIMD for all the molten salt systems and the SEMs are reduced by about 1 order of magnitude compared to the calculation reported.42 The AIMD cost for 128, 256 and 512 systems are estimated over 100 MD-steps.
The experimental data are analyzed in terms of a Born–Haber cycle as illustrated in Fig. 3. Assuming that all the gas states are well-approximated by the ideal gas, the sum of the solvation free energies of both solute ions is calculated from thermochemical tables67,68 to be −163.5 kcal mol−1. Consequently our QCT calculation at the DFT level over-estimates the excess Gibbs binding free energy by roughly −15 kcal mol−1.
Fig. 3 The Born–Haber cycle scheme for Na+ and Cl− ion solvation. The ions are assumed to be solvated from the ideal gas state (g) to liquid state (l) with number density ρo = 15.61(nm)−3 and T1 = 1150 K (NVT ensemble). The ideal gas of ions are changed into the NPT ensemble with pressure ρo = 1 bar and temperature of 1150 K. Then the temperature of the ideal gases is reduced to T0 = 298.15 K isobarically. At room temperature the sodium ion is changed to the sodium atom and chloride ion to the chlorine radical. Both atoms are then changed into elements, respectively. The solid state (s) NaCl is formed from component elements at room temperature and 1 bar pressure. Then the solid salt are heated isobarically into the liquid phase (l) at T1 = 1150 K. The change of free energy from the NVT ensemble (ρo,T1) to the NPT ensemble (po,T1) of the liquid molten NaCl(l) is neglected.43 All of the Gibbs free energy changes are from thermodynamic tables,67,68 and the total change of the solvation free energy is −163.5 kcal mol−1. |
In a previous study by Gray-Weale et al.,42 it was noted that the DFT calculation over-estimates the Na atom solvation free energy in liquid Na by −19 kcal mol−1 and the total Gibbs free energy change for the redox reaction Na(l)+ Cl2(g) ⇌ NaCl(l) by −15 kcal mol−1. The molten salt density is over-estimated by up to 10% when using the PBE DFT functional with the D2 dispersion correction, while the system appears to be unstable without the D2 correction. These results indicate both the importance and the subtlety of dispersion forces in the molten salt liquids. It is not surprising that the over-estimation of the density correlates with the over-estimation of the magnitude of the excess free energy. Some over-binding is also observed in ion solvation for the aqueous solution.45
Herein to estimate the correction for the free energy based on the underlying DFT simulations, we implement higher-level density-fitted MP2 theory70 calculations over 400 cluster configurations with the basis set aug-cc-pvdz71,72 in the Psi4 package.73 The cluster configurations with 7 and 17 solvent ion pairs are carved from a trajectory generated by the DFT simulation.
The correction (in the direction DFT → MP2) to the solvation free energy of each ion Δμex is given by
(13) |
Inclusion of the D2 correction over-estimates the interaction energy by −6.4 kcal mol−1 on average. As mentioned above, the over-binding between ions is also indicated by the 7% over-estimation of the molten NaCl density at 1150 K.42 The PBC interactions (related to the Ewald potential45,62,63) for the simulation cell of size 16.0 Å contributes about 10.0 kcal mol−1 to the interaction energy correction for the solute ion pair, while in the cell of size 25.4 Å it contributes about 2.0 kcal mol−1. The distributions of the interaction energy corrections under PBC in the cell of size 25.4 Å are presented in Panel (d) of Fig. 2.
To estimate the total free energy correction, including contributions from outside the clusters, we extrapolate the cluster interaction energy correction to 256 ion pairs. Considering the cancellation of Ewald potential contributions for the solute ion pair, we assume that it is convergent to 1.6 (1.4) kcal mol−1. Assuming the dispersion interaction energy is proportional to the solvent ion density (which is approximately a constant outside the first solvation shell), the spherical integration leads to the expression of dispersion energy correction as a + b/N, where N is the solvent ion pair number. Using the dispersion energy corrections presented in row number 2 of Table 2, we calculate the parameters a and b as (−11.59, 20.23) for Na+ and (−8.88, 16.66) for Cl−, which leads to the extrapolation for the dispersion energy correction as −11.5(1.1) kcal mol−1 for Na+ and −8.8(0.8) kcal mol−1 for Cl− (in the MP2 → DFT direction). The total dispersion energy correction for the solute ion pair is then −20.3(1.4) kcal mol−1 or +20.3 kcal mol−1 in the desired DFT → MP2 direction. The Ewald correction in this direction is −1.6 kcal mol−1.
Na+ 07 | Cl− 07 | Na+ 17 | Cl− 17 | |
---|---|---|---|---|
0) MP2 | 0.0 | 0.0 | 0.0 | 0.0 |
1) DFT | −1.9(0.5)[2.3] | −0.7(0.4)[3.5] | −3.1(0.4)[4.7] | −1.9(0.4)[3.9] |
2) DFT + D2 | −8.7(0.6)[2.8] | −6.5(0.5)[3.5] | −10.4(0.5)[5.3] | −7.9(0.3)[4.8] |
3) DFT + D2 PBC 16.0 | −15.9(0.6)[4.2] | 10.8(0.5)[4.4] | −31.3(0.6)[9.6] | 23.6(0.7)[7.0] |
4) DFT + D2 PBC 25.4 | −10.5(0.6)[2.8] | −2.5(0.4)[3.5] | −16.7(0.5)[3.4] | 0.0(0.4)[3.4] |
5) D2, (2)-(1) | −6.8(0.8) | −5.8(0.6) | −7.3(0.6) | −6.0(0.5) |
6) PBC 16.0,(3)-(2) | −7.2(0.8) | 17.3(0.7) | −20.9(0.8) | 31.5(0.8) |
7) PBC 25.4,(4)-(2) | −1.8(0.8) | 4.0(0.6) | −6.3(0.7) | 7.9(0.5) |
Summing up the total correction (PBC/Ewald and dispersion) of the solute ions, we see that the correction from the cluster calculations with 17 solvent ion pairs dominates 89.3% of that with 256 solvent ion pairs. As a result, the fluctuation term in eqn (13) is assumed to be primarily captured by the interaction energy correction of 17 solvent ion pair cluster. This leads to a −2.5(0.1) kcal mol−1 correction for each of the Na+ and Cl− solute ions. Inserting these results into eqn (13), we obtain the total correction (in the DFT → MP2 direction) as +13.7(1.9) kcal mol−1 for the solute ions. Compared to the experimental reference value of −163.5 kcal mol−1, the calculated total excess chemical potential of −164.8(2.2) kcal mol−1 provides strong initial validation of the methodology.
The methodology sets the stage for larger-scale simulations of molten salt mixtures at a quantum mechanical level of accuracy, allowing for quantitative investigations of the activities, solubilities, and redox potentials of ionic species (including the corrosion and fission products) in the liquid phase.40–42 The methodology holds the potential to provide essential data for molten salt applications in both concentrated solar energy storage materials and molten salt reactors.
This journal is © The Royal Society of Chemistry 2022 |