The following article is Open access

Computational Studies of Solubilities of LiO2 and Li2O2 in Aprotic Solvents

, , , , and

Published 12 August 2017 © The Author(s) 2017. Published by ECS.
, , Citation Lei Cheng et al 2017 J. Electrochem. Soc. 164 E3696 DOI 10.1149/2.0721711jes

1945-7111/164/11/E3696

Abstract

Knowledge of the solubilities of Li2O2 and LiO2 in aprotic solvents is important for insight into the discharge and charge processes of Li-O2 batteries, but these quantities are not well known. In this contribution, the solvation free energies of molecular LiO2 and Li2O2 in various organic solvents were calculated using various explicit and implicit solvent models, as well as ab initio molecular dynamics (AIMD) methods. The solvation energies from these calculations along with calculated lattice energies of Li2O2 and LiO2 were used to determine the solubility of bulk LiO2 and Li2O2. The computed solubility of LiO2 (1.8 × 10−2 M) is about 15 orders higher than that of Li2O2 (2.0 × 10−17 M) due to a much less negative lattice energy of bulk LiO2 compared to that of Li2O2. The difference in solubilities between LiO2 and Li2O2 likely will affect the nucleation and growth mechanisms and resulting morphologies of the products formed during battery discharge, influencing the performance of the battery cell. The calculated LiO2 and Li2O2 solubilities provide important information for fundamental studies of discharge and charge chemistries in Li-O2 batteries.

Export citation and abstract BibTeX RIS

This is an open access article distributed under the terms of the Creative Commons Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is not changed in any way and is properly cited. For permission for commercial reuse, please email: oa@electrochem.org.

Lithium-O2 battery technology has received much research interest due to its high theoretical gravimetric energy density compared to conventional Li-ion batteries.1 The system uses oxidation of lithium at the anode and reduction of oxygen at the cathode to induce a current flow. Depending on the reaction kinetics of different discharge mechanisms, the discharge products are generally composed of Li2O2 with a low solubility and in some cases different ratios of LiO2 and Li2O2.26 Much of research focus has been on the cathode. One of the challenges of Li-O2 batteries is the incomplete utilization of the active materials due to the insulating nature of the discharge products and the clogging the porous carbon cathode. The formation of these products may involve a nucleation and growth mechanism from the solution phase.2,713 Nazar reported a Li2O2 formation mechanism by disproportionation reaction of a limited amount of LiO2 in solution.2 A through-solution mechanism was also proposed involving heterogeneous nucleation from a supersaturated solution of LiO2.10 There is other evidence that LiO2 is soluble in the electrolyte prior to deposition on the surface based on a quartz microbalance study.14 Results from several research groups suggest that the composition and morphology of the discharge products have a significant effect on charging overpotential and rechargeability of the Li-O2 battery.10,15,16 Therefore, understanding the chemistry and growth mechanism of discharge products in the cell is crucial for improving Li-O2 battery. The study and modeling of nucleation and crystal growth requires knowledge of bulk solubility.17,18 However, there are limited experimental solubility measurements reported in the literature on discharge products LiO2 and Li2O2. Solubility also highly depends on the size and morphologies of the solid as well as impurities so the values will depend on the experimental conditions.

The solubility of Li2O2 has been reported in various studies; however, the values cover a very wide range. Lodge et al.19 reported a solubility product constant of Li2O2 of 10−51 in a pyrrolidinium-based ionic liquid, which corresponds to a solubility of 10−17 M. This low solubility of Li2O2 is consistent with the value found in the Handbook of Inorganic Compounds20 where Li2O2 is reported to be insoluble in ethanol. In contrast, there have been other reports of much larger solubilities for Li2O2. The Li2O2 solubility in 0.1 M LiTFSI/DMSO solution was measured to be 2.5 mM using ionic-coupled plasma mass spectroscopy,21 <0.2 mM in dimethyl formamide,22 1 mM in acetronitrile/propylene carbonate mixture23 and 0.02 mM in propylene carbonate and dimethyl carbonate mixture.24 The reason for the wide range of values for the solubility is unclear. The solubility of LiO2 has not been reported in literature since it is inherently unstable and quickly disproportionates under ambient condition. However, Lodge et al.19 estimated a solubility product of greater than 10−15 based on electrochemical measurements, which would correspond to a solubility of greater than 10−7 M. Also, a recent mesoscale model study of Li-O2 discharge product growth used a LiO2 solubility of 1 mM to successfully simulate particle size evolution observed experimentally.18

Predicting solubility using ab initio computational methods is challenging due to the difficulties of modeling both crystalline and solution phases. In this paper we report on density functional calculations of the solvation free energies of Li2O2 and LiO2 to provide theoretical predictions of solubility of these species, for which there is conflicting experimental data. The solvation energies were calculated in dimethoxyethane (DME), a common solvent used in Li-O2 battery that has been shown with improved performance, as a model solvent using implicit and a mixed explicit solvent and implicit model, as well as benchmark results from ab initio molecular dynamics (AIMD) simulations. These values are then used along with calculated bulk lattice energies to obtain solubilities of bulk LiO2 and Li2O2 using a thermodynamic cycle. We also make predictions of solubilities for a large number of solvent molecules with differing dielectric constants. In section Computational methods we present the quantum mechanical methods used for energy calculations. In section Calculation of solubility we discuss the method used to determine solubility, including the models used for the calculation of the solvation energy and the derivation of sublimation energies. In section Results and discussion the results for the solvation energies and solubilities are given along with an assessment of the different solvation models used. Finally, comparison with experiment is discussed.

Computational Methods

All density functional theory (DFT) and MP2 calculations on solvated Li2O2 and LiO2 clusters were done with Gaussian 09.25 Structures were optimized at the B3LYP/6-31G(2df,p) level and frequencies were scaled by 0.9854 as in G4(MP2) theory. Single-point calculations with PBE/6-31G(2df,p) were performed at the B3LYP geometries for comparison. The PCM,26 CPCM,27,28 and SMD29 continuum solvation models implemented in Gaussian09 were used to calculate solvation energies implicitly. Single-point MP2 calculations were done in addition to the DFT calculations. The MP2 calculations were done with frozen core according to the G4 convention30 using the G3MP2LargeXP basis set as in G4(MP2) theory.30

AIMD calculations were done using DFT methods with plane wave basis sets as implemented in the VASP code.3134 All the calculations were spin-polarized and carried out using the gradient corrected exchange-correlation functional of Perdew, Burke and Ernzerhof (PBE)35 under the projector augmented wave (PAW)36,37 method, with plane wave basis sets up to a kinetic energy cutoff of 300 eV to accommodate the large number of atoms (up to 324 in the simulation). A comparison of total radial distribution function vs radial distance calculated using 300 eV and 400 eV shows that the lower cutoff energy is sufficient for the system (Figure S4). The PAW method was used to represent the interaction between the core and valence electrons, and the Kohn-Sham valence states (i.e. 1s for H, 2s for Li, 2s 2p for C and O) are expanded in plane wave basis sets. The convergence criterion of the total energy was set to be within 1 × 10−4 eV within a single K-point grid (i.e. Γ-point). For the simulations, supercells of 12.0 Å × 12.0 Å × 12.0 Å containing 10 DME molecules and 15.0 Å × 15.0 Å × 15.0 Å containing 20 dimethyoxyethane (DME) molecules are adopted throughout the work. The DME densities corresponding to such a model are ∼0.88 and 0.90 g/cm3, respectively, close to the reported DME solvent density, i.e. 0.87 g/cm3. To investigate the thermodynamic stability of the system at room temperature, the 20 DME solvent molecules were initially randomly arranged inside the simulation box and were then thermally equilibrated at T = 300 K for about 2 ps using AIMD simulations based on an NVT-ensemble with a time step of 1 fs. For the production run, the sampling is obtained from a 4 ps AIMD simulation run.

Calculations of Solubility

Solubility is determined by the intrinsic stability of the species in its solid form and solvation by the particular solution. The solubility of substance A (SA) can be calculated using the following equation:38

where the dissolution free energy ΔG*dissolution, A is the summation of sublimation free energy ΔG*sub, A and solvation free energy ΔG*solv, A. This is a thermodynamic cycle in which molecules of the crystal are first sublimed into the gas phase and then dissolved in the solvent. The ΔG*sub, A is the free energy change of one unit of A dissociating from a bulk entity and it is typically endothermic. The ΔG*solv, A is the solvation free energy of one unit of A and can be calculated using either an implicit continuum model, or an explicit solvent model where the solvent-solute interaction is calculated using ab initio methods. The computational details of these two terms are discussed further below. R and T are the gas constant and the temperature, respectively. SA and VA are the solubility (mol/L) and the molar volume (L/mol) of A in its solid structure, respectively. The molar volume VA of LiO2 in this work was obtained from a theoretical crystal structure39 since there is no available experimental data in literature due to the instability of the structure under normal condition, and that of Li2O2 from the density and molecular weight.

Solvation energy

In this work, we use both implicit and explicit models to calculate solvation free energies of LiO2 and Li2O2 in DME. With implicit solvation models, the free energy of the molecule in solution can be calculated directly from a dielectric continuum model. The solvation energy is calculated by subtracting the gas phase electronic energy from the energy in solution. In the continuum models, because DME is not a standard solvent for the method, we specified the solvent as dichloroethane which has the same shape as DME,40 and the dielectric constant is the literature value of 7.2.

With the explicit models including 1∼4 DME solvent molecules, several different cluster structures constructed based on chemical intuition of the molecular interactions were tried and the lowest energy ones were used. It has been shown before that the complete of the first solvation shell is satisfactory for calculating solvation free energies of lithium-oxygen species.41 Since 4 DME molecules complete the first solvation, we did not calculate structures with more than 4 solvent molecules. These structures are shown in Figure 1. The first solvation shell of the Li2O2 from a snapshot from AIMD is also shown in Figure 1. Two different thermodynamic cycles, the monomer cycle and the cluster cycle,41,42 were used to derive the solvation free energy. The two models use different reference energies. In the monomer cycle the solute is treated as binding to n solvent molecules while in the cluster cycle the solute binds to a cluster of n solvent molecules. Correction factors are added so that all reactants and products are in the same standard states, namely 1 M ideal gas and 1 M solution. For the monomer cycle, the solvation free energy is calculated as:

where ΔGobind, g, A is the gas phase binding free energy between the solute and n DME molecules. Solvation free energies of the solute A with n surrounding explicit solvent DME molecules ΔG*solv, [A(DME)n] and the solvation energy of DME molecule ΔG*solv, DME were calculated with PCM by subtracting the gas phase electronic energy from the corresponding SCRF energy.CDME is the concentration of 1 mol of DME and it is 9.6 mol/L. ΔGo − >* is the free energy change to convert one mol of ideal gas from 1 atmosphere (24.46 L/mol) to 1M (1mol/L), which is 1.9 kcal/mol at 298 K.42 For all continuum models used (PCM, CPCM and SMD), this ΔGo − >* correction is added.

Figure 1.

Figure 1. Optimized structures of LiO2 and Li2O2 interacting with 1∼4 dimethoxy ethane (DME) molecules DMEs. A structure of Li2O2 interacting with 6 DME molecules (Li2O2...6DME) obtained from a snapshop of AIMD simulation with 20 DMEs in a box is also shown.

For the cluster cycle, the solvation free energy is:

where ΔG*solv, (DME)n is the solvation energy of a cluster of DME surrounding the solute calculated using a continuum model and CDME/n is the concentration of the solvent cluster.

We also calculated the solvation energy using the AIMD method. The computed solvation energy is approximated as the formation energy ΔE*solv, A(T) from direct AIMD sampling at a given temperature (T) defined as the following:

where Etotsol(DME, A)(avg, T) is the statistical average of the total energy of a solution from AIMD sampling of the solute A with 10 or 20 surrounding DME solvent molecules thermally equilibrated at T = 298 K. Similarly, the Etotsol(DME)(avg, T) and EtotA(avg, T) are the total energies of a solution from AIMD sampling of the 10 or 20 DME solvent molecules and one A solute molecule thermally equilibrated at T = 298 K, respectively. More statistical information on the AIMD simulations can be found in the Supplemental Materials. In this work, we compared all other methods against the AIMD results as the model captures more complete solvation shell information and thus is the most physical and likely the most accurate. It has also been shown in literature that the solvation energies calculated using PBE AIMD are in good agreement with the experimental values.43

Sublimation energy

The standard sublimation free energy ΔGosub, A must be corrected for isothermal expansion of an ideal gas:42

where Po is the standard condition pressure. Molar volumes VA for LiO2 and Li2O2 are 0.0173 and 0.0199 L/mol, respectively.

The ΔGosub, A of bulk is the summation of ΔHosub, A and ΔSosub, A. The ΔHosub, A can be calculated according to:

where Ulatt is the lattice energy calculated from crystal structures defined as Ulatt = (Etotcrystal, AnEmol, Atot)/n with Etotcrystal, A being the total energy of the A crystal that consists of n basic A molecular units in a crystal cell, and Etotmol, A being the corresponding total energy of a basic A molecular unit from a DFT calculation. Based on this relation, we obtained −1.85 and −3.26 eV for Ulatt for LiO2 and Li2O2, respectively. The 2RT term arises by including lattice vibrational energy and energy of the vapor.38 The ΔSosub, A is approximated by the difference between the rotation and translation contributions to the entropy of the gas phase at 298 K and the intermolecular vibrational contribution to the entropy of the crystal at 298 K. Details of these calculations were described in a previous publication.39 With these entropy terms, ΔGosub, A of LiO2 and Li2O2, are calculated to be 1.47 and 2.83 eV, respectively.

Results and Discussion

Calculated free energies of solvation for the various methods, including the AIMD results, are reported in Table I with the optimized structures of Li+-DMEn cluster binding shown in Figure 1. The B3LYP and PBE energies for the cluster cycles (Table I, Rows 5–7) and the continuum models (Row 8–10) are very comparable (∼1 kcal/mol). With a monomer cycles (Table I, Row 1–4), the differences between ΔG*solv, A with different numbers of DME molecules are noticeable (6–12 kcal/mol for B3LYP). When clusters of DME are used instead (cluster cycle), ΔG*solv, A (Table I, Row 5–7) calculated using different cluster sizes are more similar. However, ΔG*solv, A calculated using monomer cycles (Table I, Row 1–4) are in better agreement with the AIMD results (Row 11–12) with the 4DME monomer cycle (Row 4) giving the best agreement. For cluster and monomer cycles (Table I, row 1–7), MP2 solvation energies for LiO2 are slightly more negative than the B3LYP results except for the 1DME monomer model (Row 1), while for Li2O2 the MP2 energies are less negative and the deviation is larger.

Table I. Solvation free energies (kcal/mol) of molecular LixO2 calculated using different models. For DFT calculations, 6–31G(2df,p) basis sets were used. For MP2 calculations, the GTMP2LargeXP basis sets were used.

    B3LYP/6-31(2df,p)e PBE/6-31(2df,p)f MP2/GTMP2LargeXPg
  Method x = 1 x = 2 x = 1 x = 2 x = 1 x = 2
1 aDME + LixO2 → DMELixO2 −35.1 −42.4 −34.0 −42.1 −30.4 −37.0
2 a2DME + LixO2 → DME2LixO2 −32.9 −46.4 −34.7 −47.4 −34.2 −37.5
3 a3DME + LixO2 → DME3LixO2 −28.1 −40.2 −31.5 −44.8 −31.1 −36.5
4 a4DME + LixO2 → DME4LixO2 −23.5 −36.8 −27.6 −43.4 −29.5 −35.8
5 bDME2 + LixO2 → DME2LixO2 −36.6 −50.1 −37.1 −49.9 −37.3 −40.5
6 bDME3 + LixO2 → DME3LixO2 −36.5 −48.7 −36.6 −49.9 −38.4 −43.9
7 bDME4 + LixO2 → DME4LixO2 −35.2 −48.4 −35.8 −50.2 −37.7 −43.9
8 cLixO2 PCM −25.2 −34.7 −24.4 −34.2 −26.7 −34.6
9 cLixO2 CPCM −26.4 −36.6 −25.5 −36.0 −28.1 −36.5
10 cLixO2 SMD −13.9 −16.4 −13.3 −16.0 −15.2  
11 d10DME + LixO2 → DME10LixO2     −29.4 −38.4    
12 d20DME + LixO2 → DME20LixO2       −40.3    

aMonomer model. bCluster model. cContinuum model. dAIMD. eAt B3LYP/6-31G(2df,p) optimized geometries. fAll PBE results are with the 6–31G(2df,p) basis set at the B3LYP/6-31G(2df,p) geometries, except for the AIMD results (see text for details of those calculations). gAt B3LYP/6-31G(2df,p) optimized geometries.

For the implicit solvation models PCM and CPCM, the results with both DFT and MP2 methods (Table I, Row 8 and 9) are similar to the AIMD results (Row 11 and 12), while for the SMD model (Row 10) results deviate more. This is because SMD is not parameterized for Li. The AIMD simulations using a larger box of 20 DME (Table I, Row 12) gives a slightly more negative ΔG*solv, A for Li2O2 than the 10 DME (Row 11) box (−40.3 vs −38.4 kcal/mol). Li2O2 has a higher (more negative) ΔG*solv, A than LiO2 due to its stronger binding to DME. Considering the computational cost, the PCM and CPCM methods are preferable to cluster or monomer cycles for solvation energy calculation. In addition, it is not always clear if the lowest energy structure has been found.

Solubilities were calculated and are reported in Table II by combining theoretical ΔG*sub, A and ΔG*solv, A as outlined in the methods section. The AIMD solubility (Table II, Row 11 and 12) of LiO2 is 1.8 × 10−2 mol/L and that of Li2O2 is 9.1 × 10−19 mol/L (2.0 × 10−17 mol/L with the larger 20 DME box), indicating that bulk LiO2 is more soluble than Li2O2 in DME. The PCM and CPCM values as well as the monomer cycle in Table II based on 4MDE are in agreement with these values. We note that because the solubility is an exponential function of solvation energies, small errors in solvation energy calculation will lead to a significant difference in the calculated absolute solubility. Our calculations showing that Li2O2 is very insoluble in DME and other solvents is consistent with some studies19,20 and in disagreement with others2124 which are many orders of magnitudes larger (details on these experimental measurements were discussed earlier). For example, the solubility of 10−17 M reported by Lodge et al.19 and the insolubility of Li2O2 in ethanol reported in the Handbook of Inorganic Compouds20 are consistent with our calculated results. Our calculations indicate that although the solvation energy of Li2O2 is quite large, it is the stability of Li2O2 (large lattice energy) that makes it insoluble. Although the solvation energies of the differing levels of theory differ, they all are in agreement on this qualitative result with values of 10−21 to 10−17 mol/L. The reports of large solubility of Li2O2 in the literature2124 could be for several reasons. One possibility is that since Li2O2 is known to react with H2O, trace amounts of water contamination in solution might result in erroneous solubility measurement and in one case the solubility closely matched the ppm of water reported in the solvent.22 Another possibility is that solubility measurements may depend on particle size or surface effects that increase solubility.

Table II. Solubilities (mol/L) of bulk LixO2 calculated using different solvation models. For DFT calculations, 6–31G(2df,p) basis sets were used. For MP2 calculations, the GTMP2LargeXP basis sets were used. For easy comparison, the row labels shown in the first column are consistent with the label numbers for the same methods in Table I.

        MP2/GTMP2Large
    B3LYPa AIMD XP//B3LYPa
  Method x = 1 x = 2 x = 1 x = 2 x = 1 x = 2
4 4DME + LixO2 → DME4LixO2 8.8E-07 5.7E-20     2.2E-02 1.0E-20
8 LixO2 PCM 1.5E-05 1.8E-21     2.0E-04 1.6E-21
9 LixO2 CPCM 1.1E-04 4.2E-20     1.9E-03 3.8E-20
11 10DME + LixO2 → DME10LixO2     1.8E-02 9.1E-19    
12 20DME + LixO2 → DME20LixO2       2.0E-17    

aUsing the optimized geometries at the B3LYP/6-31G(2df,p) level of theory.

As mentioned earlier, the experimental solubility of LiO2 is not available. However, a recent investigation that uses LiO2 solubility of 1 mM to model particle growth successfully simulated the particle size evolution observed experimentally.18 The solubility used for modeling is close to our AIMD result of 1.8 × 10−2 mol/L, supporting our calculation result.

Solvation free energies of LixO2 in some other common organic solvents with dielectrics ranging from 2 to 109 were also calculated using the CPCM model and are reported in Table III. Also included in Table III are the computed solubilities of LiO2 and Li2O2 in these solvents. The solubilities in various other solvents are found to vary greatly (up to 15 orders of magnitude) depending on the dielectric constant. The solubility of Li2O2 ranges from 10−16 to 10−32 and that of LiO2 ranges from 10−1 to 10−13.

Table III. Solvation free energies (kcal/mol) and solubilities (mol/L) of LixO2 in different organic solvents calculated using B3LYP/6-31G(2df,p) and CPCM model.

    Solvation energy Solubility
Solvent Dielectric constant x = 1 x = 2 x = 1 x = 2
1,4-Dioxane 2.2 −14.9 −20.3 4.2E-13 5.0E-32
DiButylEther 3.0 −18.7 −25.6 2.4E-10 3.7E-28
Anisole 4.2 −21.7 −29.8 4.0E-08 4.7E-25
Dimethoxy ethane(DME)a 7.2 −26.4 −36.6 1.1E-04 4.2E-20
TetraHydroFuran 7.4 −25.4 −34.9 2.0E-05 2.6E-21
2,4-DiMethylPyridine 9.4 −26.5 −36.4 1.3E-04 3.3E-20
3-MethylPyridine 11.6 −27.3 −37.5 4.9E-04 2.1E-19
Pyridine 13.0 −27.6 −38.0 8.8E-04 4.7E-19
ButanoNitrile 24.3 −29.1 −40.0 1.0E-02 1.4E-17
BenzoNitrile 25.6 −29.2 −40.1 1.2E-02 1.7E-17
PropanoNitrile 29.3 −29.4 −40.4 1.7E-02 2.7E-17
Acetonitrile 35.7 −29.6 −40.8 2.6E-02 4.9E-17
n,n-DiMethylAcetamide 37.8 −29.7 −40.9 2.9E-02 5.7E-17
DiMethylSulfoxide 46.8 −29.9 −41.1 4.2E-02 9.3E-17
Formamide 108.9 −30.4 −41.9 1.0E-01 3.1E-16

aDME is not one of the standard solvents in the solvation model of Gaussian 09 software. A dielectric constant of DME was specified with other parameters set to dichloroethane (similar shape as DME) in the calculation.

Conclusions

The solvation energies and solubilities of LiO2 and Li2O2 in aprotic solvents were calculated using density functional theory and various solvent models for dimethoxy ethane (DME). The following conclusions are drawn:

  • (1)  
    The AIMD simulations give free energies of solvation of −29.4 and −40.3 kcal/mol for LiO2 and Li2O2, respectively. Solvation energies calculated using the monomer cycles with 4 DME solvent molecules and continuum models PCM and CPCM were found to be comparable with the AIMD results. The PCM and CPCM methods are recommended to be used for solvation energy calculation due to their relatively lower computational cost.
  • (2)  
    The solvation energies were used with calculated free energies of sublimation to determine solubilities. The calculated solubilities of LiO2 (1.8 × 10−2 mol/L) and Li2O2 (2.0 × 10−17 mol/L) in DME indicates that bulk LiO2 is slightly soluble and Li2O2 is very insoluble.
  • (3)  
    The computed solubilities of LixO2 (x = 1 or 2) in various organic solvents vary significantly depending on the dielectric constant of the solvent. For example, the computed solubility of LiO2 in 1,4-dioxane (ɛ = 2.2), DME (ɛ = 7.2) and DMSO (ɛ = 46.5) is 10−13, 10−4 and 10−2 mol/L, respectively. The solubilities of Li2O2 in the three solvents above are 10−32, 10−20 and 10−17 mol/L, respectively.
  • (4)  
    Experimental solubilities for Li2O2 range from 10−17 mol/L or "insoluble" to 10−3 mol/L. Our calculations agree with the former. The results giving larger solubilities may be due to impurities or surface effects. The results for LiO2 agree with a mesoscale model for particle growth that successfully simulated particle size evolution observed experimentally.

The calculated LiO2 and Li2O2 solubilities provide important information for fundamental studies of discharge and charge chemistries in Li-O2 batteries. Since solubility is an important parameter that determines nucleation and growth of the discharge product,18 tuning the solubility of the LiO2 and Li2O2 using different electrolytes by design can help achieving desired product morphology.

Acknowledgment

This work was supported by the Joint Center for Energy Storage Research (JCESR), an Energy Innovation Hub funded by the U. S. Department of Energy, Office of Science, Basic Energy Sciences. We gratefully acknowledge the computing resources provided on Blues and Fusion, a high-performance computing cluster operated by the Laboratory Computing Resource Center at Argonne National Laboratory. This research also used resources of the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility supported by the Office of Science of the U. S. Department of Energy.

Please wait… references are loading.
10.1149/2.0721711jes