ACS Publications. Most Trusted. Most Cited. Most Read
Efficient Molecular Dynamics Simulations of Deep Eutectic Solvents with First-Principles Accuracy Using Machine Learning Interatomic Potentials
My Activity

Figure 1Loading Img
  • Open Access
Molecular Mechanics

Efficient Molecular Dynamics Simulations of Deep Eutectic Solvents with First-Principles Accuracy Using Machine Learning Interatomic Potentials
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

Journal of Chemical Theory and Computation

Cite this: J. Chem. Theory Comput. 2023, 19, 23, 8732–8742
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jctc.3c00944
Published November 16, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

In recent years, deep eutectic solvents emerged as highly tunable and ecofriendly alternatives to common organic solvents and liquid electrolytes. In the present work, the ability of machine learning (ML) interatomic potentials for molecular dynamics (MD) simulations of these liquids is explored, showcasing a trained neural network potential for a 1:2 ratio mixture of choline chloride and urea (reline). Using the ML potentials trained on density functional theory data, MD simulations for large systems of thousands of atoms and nanosecond-long time scales are feasible at a fraction of the computational cost of the target first-principles simulations. The obtained structural and dynamical properties of reline from MD simulations using our machine learning models are in good agreement with the first-principles MD simulations and experimental results. Running a single MD simulation is highlighted as a general shortcoming of typical first-principles studies if the dynamic properties are investigated. Furthermore, velocity cross-correlation functions are employed to study the collective dynamics of the molecular components in reline.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2023 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Deep eutectic solvents (DESs) are an emerging class of compounds characterized by melting points significantly lower than their individual components. (1−4) They have many desirable characteristics such as low vapor pressure, thermal stability, and tunable properties, which make them suitable candidates as room-temperature solvents (5−7) and electrolytes. (8,9) DESs are generally nontoxic, (10,11) biodegradable, (12) and easy to prepare. Furthermore, the ability to fine-tune their physicochemical properties, through the exchange or modification of their components, makes them a promising class of “designer solvents”.
Preparation of novel DESs from a large number of potential components and a wide range of compositions is still a challenging task. In this context, computer modeling and simulation are valuable complements to experiment, providing insights into the nature of interactions at the atomic scale and the connections between structure and functionality. In particular, molecular dynamics (MD) simulations are an efficient and predictive tool for exploring a large number of potential compounds and their properties for targeted applications. There are a large number of MD simulation studies of DESs, including first-principles MD (FPMD) simulations (13−15) and classical MD simulations using additive (16−20) and polarizable (21−23) interatomic potentials (or force fields). While FPMD simulations based on electronic structure methods such as density functional theory (DFT) (24) can predict the structure and dynamics of the studied systems with a higher accuracy, their computational cost generally scales as the cube of the number of electrons in the system (25) and thus are limited to short length and time scales. In classical force fields (FFs), interatomic interactions are modeled using physically inspired functional forms that can be evaluated efficiently. (26,27) Hence, they are better suited for the size and time scales required to properly reproduce the structural and dynamical correlations in extended systems, such as DESs. MD simulations using FFs based on fixed atomic charges can provide important insights into the structure of ionic systems such as DESs, but their predicted thermodynamic and transport properties are often inconsistent with experiments showing significantly reduced dynamics. (28,29) The use of polarizable FFs can greatly improve the simulated dynamics in these systems since the explicit inclusion of induced dipoles in polarizable models effectively screens the long-range electrostatic interactions. (30) However, a frequent issue of polarizable FFs is the overpolarization of small ionic species, which demand special considerations such as damping of the electrostatic interactions between the selected pair of atoms. (22,23,31) This makes the parametrization of polarizable models for ionic systems such as DESs challenging, and the resulting FFs are not generally transferable. The main drawback of classical FFs is that their accuracy is limited to the underlying functional forms and the quality of their parametrization. Furthermore, classical FFs require predefined bonding patterns and, thus, cannot describe bond breaking or bond formation.
In recent years, with the advancement of machine learning (ML) techniques, a data-driven approach to the development of interatomic potentials has emerged in the context of ML. In this approach, the parametrization of an interatomic potential is regarded as a regression problem, where a set of continuous inputs such as atomic coordinates are mapped to a set of continuous outputs such as energies and atomic forces. Thus, the focus is shifted toward a sufficiently general functional form that can be trained on a set of data from electronic-structure (e.g., DFT) calculations. ML interatomic potentials (MLIPs) hence aim to approximate a set of energy and forces for a given atomic configuration, with the target first-principles accuracy at improved computational efficiency that scales linearly with the number of atoms. A variety of different approaches for training MLIPs based on kernel methods (32−35) and deep neural networks (NNs) (36−43) have been proposed. NN-based models, in particular, gained a lot of attention due to their high flexibility and scalability to larger data sets. A common approach for the construction of MLIPs is the decomposition of the total energy of a chemical system into atomic contributions. (36) To ensure the invariance of the energy under translation, rotation, and permutation of equivalent atoms, the atomic energies depend on the local environment of atoms through atom-centered representations rather than directly on the Cartesian coordinates. This is based on the chemically intuitive assumption that an atom’s contribution to the total energy is local and depends on its environment. In recent years, many different structural representations have been proposed such as ACSFs, (44) SNAP, (45) DeepPot, (42) and ACE, (46) for which the environment of an atom is encoded in a hand-crafted descriptor and SchNet (41) and PhysNet (43) as rotationally invariant and NequIP, (47) PaiNN, (48) and NewtonNet (49) as rotationally equivariant graph-based message-passing neural networks (MPNNs), for which the suitable atomic representations are learned from the reference data. While MPNNs, especially the equivariant types, are generally more data efficient and could achieve higher prediction accuracies, they are more expensive to train and evaluate, which limits their application for larger systems. It should be noted that the more recently proposed class of equivariant networks, with a strictly local representation of the atomic environments, can readily scale to large systems. (50)
In this work, we utilized DeePMD, (51) a descriptor-based NN representation, to train an MLIP for reline DES, a mixture of choline chloride (ChCl) and urea in 1:2 ratio (Figure 1). We employ an iterative active learning scheme based on model ensemble uncertainty estimates to systematically improve our models. We show that our trained MLIP on a set of DFT data points can achieve small prediction errors for energies and atomic forces on a set of unseen atomic configurations. Using our MLIP, we are able to run nanoseconds-long MD simulations for thousands of atoms with the target DFT accuracy. Finally, we show that our MLIP can successfully recover the structural and dynamic properties of reline from MD simulations.

Figure 1

Figure 1. Schematic representation of choline chloride (left) and urea (right) with the corresponding atom labels used throughout this study.

2. Methods

Click to copy section linkSection link copied!

2.1. Machine Learning Model

We utilized the DeePMD (52) framework, an end-to-end deep NN representation of a many-body potential energy surface (PES), (51) to train and evaluate our ML interatomic potentials. In the deep potential (DP) model, like many of the size-extensive MLIPs, (36,37,39) the total energy of the system is decomposed into atomic contributions using a local geometric description of the atomic environment. Initially, the local environment of atom i is represented by matrix RiRNi×3, where Ni is the number of neighboring atoms within a cutoff radius rc. In Ri, each row represents one neighbor j and is given by the relative coordinate rijrjri = (xij, yij, zij), with ri and rj being the Cartesian coordinates of atom i and j, respectively. Next, the relative coordinates Ri are mapped onto generalized coordinates R~iRNi×4. In R~i, each row is given by {s(rij),x^ij,y^ij,z^ij} where x^ij=s(rij)xij/rij, y^ij=s(rij)yij/rij, and z^ij=s(rij)zij/rij, with rij being the Euclidean distance between atoms i and j and s(rij) is a continuous and differentiable weighting function applied to each component
s(rij)={1rijrij<rs1rij[x3(6x2+15x10)+1],rsrij<rc0rijrc
(1)
where x = (rijrs)/(rcrs) switches from 0 at rs to 1 at the cutoff radius rc. The function s(rij) gives greater weight to the nearest neighbors of atom i, and parameter rs defines a smooth cutoff radius that allows the components in R~i to smoothly go to zero at the boundary of the local region rc. (51) In order to include the global symmetries including translation, rotation, and atomic permutation invariance of the energy in the local atomic representations, (34,37) the first column of the generalized coordinates R~i is mapped to an embedding matrix GiRNi×M1 using a local embedding NN. Finally, by taking the first M2 columns of the embedding matrix, as Gi2RNi×M2, an invariant feature matrix DiRM1×M2 is constructed
Di=(Gi)TR~i(R~i)TGi2
(2)
The parameters of the embedding network are optimized simultaneously with the fitting network, which maps the feature matrix Di to the atomic energy Ei. The total energy E is then calculated as the sum of atomic energies Ei. Atomic forces are obtained by taking the derivative of the total energy with respect to the atomic coordinates; this enables the use of the force components in addition to the total energy for the training of the NNs. We used a cutoff value of rc = 6 Å, the smoothing cutoff rs = 5 Å, and the periodic boundary condition to construct the local atomic representations. A three-layer embedding network containing 25, 50, and 100 neurons was used with M2 = 16 axis neurons. We used a three-layer fitting network containing 240 neurons in each layer for mapping the atomic representations Di to the atomic energies. The hyperbolic tangent was used as the activation function in both embedding and fitting networks.
The objective in the training of the networks is to minimize a loss function with respect to NN parameters using an Adam stochastic gradient descent optimizer. (53) For the optimization step n, the loss function is defined as an average over a batch of Nb configurations, denoted as B
L(n)=1NbBNbL(B,n)
(3)
L(B,n)=pε(n)ΔE2+pf(n)3NiN|ΔFi|2
(4)
Here, Δ denotes the difference between the prediction and the reference training data, N is the number of atoms, E is the total potential energy of the configuration, and Fi is the force on atom i. The use of forces in the training process significantly reduces the number of reference data needed to train accurate and robust models. (51) The prefactors pϵ and pf are tunable weights that vary during the training. We progressively increased pϵ and decreased pf, so that the force term dominates at the beginning of the training, while the energy term becomes equally important toward the end. We used a batch size Nb = 1, 0.02 and 1 as starting and final values of pϵ and 1000 and 1 as starting and final values for pf, respectively. The learning rate (i.e., the step size in gradient descent optimization) is decreased during training following the formula α(n) = α(0)λ(n/τ), where λ is the decay rate and τ is the decay steps. We used decay steps of 5000 and set the starting and stopping values of the learning rate to 10–3 and 5 × 10–8, respectively.

2.2. Reference Data Generation

For generation of the training data sets, sampling the configuration space of the system using faster methods such as density functional-based tight-binding (DFTB) (54) or classical FFs is significantly more efficient than performing FPMD simulations using the target DFT method. Our primary data set is built on the basis of configurations sampled from DFTB MD trajectories for a system containing 8 ChCl and 16 urea molecules. The sampled structures are subsequently labeled with energy and atomic forces using the target DFT method. The details of the DFTB-based simulations are provided in the Supporting Information. The DFT calculations on the selected configurations are carried out using the Gaussian and plane-wave (GPW) method as implemented in the CP2K package. (55,56) The Becke–Lee–Yang–Parr (BLYP) functional (57,58) is used along with the empirical dispersion correction (D3) of Grimme. (59) We used a MOLOPT-TZVP basis set and a plane wave cutoff of 700 Ry. The core electrons were described with the Goedecker–Teter–Hutter (GTH) pseudopotential. (60,61) The periodic boundary conditions were applied in all directions. A total of 19,000 configurations were selected and labeled with DFT energies and atomic forces, from which 18,000 are used as our primary training data set and the rest as the validation set.

2.3. Active Learning

In order to reduce the computational cost of the reference data generation and to ensure that the generated data set is representative of the relevant parts of the PES, we utilized an active learning scheme to extend the primary training data set. (62,63) In active learning, which can be seen as the generation of training data on demand, an uncertainty measure is used to validate the model predictions for new configurations. (64) A set of configurations with high prediction error are then selected to be labeled with the reference DFT energies and forces to extend the data set. We used the DP-GEN package as the active learning platform to extend our primary data set. (65) In the DP-GEN workflow, a series of successive iterations are performed, with each iteration composed of three steps of exploration, labeling, and training. (19) Initially, in the training step, a set (i.e., an ensemble) of four models are trained on our primary data set. The trained models differ only in the initialization of NN parameters. Since the loss function in eq 4 is a nonconvex function of the NN parameters, such a difference in initialization leads to different local solutions to the optimization problem and therefore different PES models. Around atomic configurations where there are sufficient training data, the different models should all be reasonably accurate and give predictions that are close to each other. For configurations far from the training data, the predictions from the different models will deviate from each other with considerable variance. Accordingly, the variance of the set of predictions from the model ensemble for a particular configuration can be used as a measure of uncertainty of the prediction and a criterion for whether the configuration should be selected as a candidate to extend the data set. For a configuration Rt, the error indicator ϵt is defined as the maximal standard deviation of the atomic force predicted by the model ensemble
εt=maxiFi(Rt)Fi(Rt)21/2
(5)
where Fi(Rt) denotes the predicted force on atom i and the notations ⟨···⟩ denote the ensemble average of the predictions. Being a derivative of energy, forces typically show higher variability outside the training domain and thus describe uncertainty better than the energy. Furthermore, for a given atomic configuration, the energy is a global quantity and does not provide sufficient resolution. However, force is an atomic property and is sensitive to the local accuracy of the models, especially when a failure happens. (19) Given an upper and lower bound as the trust level, σhi and σlo, those structures whose error indicator εt fall between these bounds will be considered candidates to be labeled with DFT energy and forces and be added to the training data set. In the exploration step, a series of 30 MD simulations were carried out using a trained model, starting from different initial configurations, which were randomly selected from a set of 100 configurations. MD simulations are performed in LAMMPS, (66) using the DeePMD extension, (52) at different temperatures in 360–440 K interval with an integrator time step of 0.5 fs and an increasing time length of 10–200 ps from the starting to the final iteration. The structures along MD trajectories were saved and evaluated by the model ensemble at a time interval of 0.05 ps and those with error indicator εt between 0.15 and 0.30 eV Å–1 were selected as candidates for the labeling step. The upper and lower bounds for εt were determined through experimenting and are selected in a way to minimize the number of candidate structures while maintaining a reasonable accuracy. In the labeling step, a maximum number of 900 structures were randomly selected from the set of candidates collected in the exploration step and labeled with DFT energy and forces as described before. In the next iteration, new approximations of the PES were obtained by retraining the models using the extended training data set. These steps are repeated until the number of candidate structures from the exploration step was less than 100 structures. At the end of the active learning steps, around 10,000 configurations were generated, labeled and added to the primary data set.

2.4. Molecular Dynamics Simulations

From the set of trained models, the one with the lowest root-mean-square error (RMSE) value for the predicted energies and atomic forces on a test set was chosen as our MLIP to run MD simulations. MD simulations were carried out in LAMMPS (66) using the DeePMD extension. (52) The canonical (NVT) ensemble was used with an integrator time step of 0.5 fs, and trajectories were collected at a time interval of 10 fs. A mixture of 64 ChCl and 128 urea molecules in a periodic box was simulated at six different temperatures from 323 to 373 K in 10 K steps. A Nose–Hoover chain thermostat was used with a relaxation time of 0.1 ps. (67) Initial configurations are generated using the selected structures from our DFTB trajectories, which subsequently deformed to the experimental densities. (68) The simulation box lengths are reported in Table S1 of the Supporting Information. The mixtures were equilibrated for 1 ns, followed by a production run of 2 ns for the first three temperatures and 1 ns for the rest. To estimate the liquid densities, additional simulations were performed using the isothermal–isobaric (NPT) ensemble at each temperature for 0.5 ns. The pressure was set to 1 bar using a Nose–Hoover barostat with a relaxation time of 1.0 ps. The error indicator described in the active learning step can also be used to validate the simulation runs. During the simulations, we evaluated the structures at a time interval of 0.1 ps using our set of trained models to ensure that the maximum standard deviation of the ensemble predictions on atomic forces stays below a certain trust value.

3. Results and Discussion

Click to copy section linkSection link copied!

3.1. Model Accuracy

The accuracy of the trained model was evaluated for a set of 100 atomic configurations that were not part of the training data set. Half of these test configurations are selected from a separate DFTB MD trajectory than the ones used for generation of the primary data set and have the same composition (8 ChCl and 16 urea molecules) as the configurations used for the training of the models. The other half are configurations randomly sampled from a FPMD trajectory by Zahn (13) for a larger system of 18 ChCl and 36 urea molecules, which were subsequently labeled with our target DFT energy and forces. Figure 2 shows a comparison of the predicted and reference values of the total energy per atom and force on each atom for all configurations. Our MLIP for reline achieves an RMSE of 0.92 meV/atom (0.02 kcal mol–1) and 5.48 meV/atom (0.13 kcal mol–1) in predicted energies for systems of size 304 and 684 atoms, respectively. Unlike energies, our trained model achieves consistent accuracy in the prediction of forces for both sets of structures, as is evident by an RMSE value of ∼ 0.05 eV Å–1 for both systems. These results indicate that our trained models on configurations from a smaller system can be successfully employed for larger systems without a noticeable loss in accuracy. It should be noted that achieving small prediction errors on a test set by an ML potential does not necessarily guarantee stable MD simulations. (69) It has been shown that sufficiently large training data sets and improved data generation schemes such as active learning are essential for obtaining robust ML potentials. (70)

Figure 2

Figure 2. Reference DFT energies and atomic forces are compared to the predicted values by the ML model for 50 test structures of a system of 304 atoms (left) and 684 atoms (right). The reference (ref) and predicted (pred) energies are standardized by subtracting their mean value from them.

3.2. Density and Structure of the Liquid

To investigate the ability of our ML potential to predict the liquid density of reline, average densities from NPT simulations are presented in Table 1 and compared to the reported experimental densities. (68) The estimated densities from MD simulations using our ML potential are within the acceptable range and underpredict the experimental densities by 3–3.5%. It should be mentioned that training the ML potentials on stress tensor data, together with energy and forces, could further improve the accuracy of the predicted densities from NPT simulations.
Table 1. Comparison of the Liquid Density of Reline (in g cm–3) from MD Simulations Using the MLIP to Experimental Values of Ref (68) at the Range of Studied Temperatures
T [K]323333343353363373
Exp.1.1831.1781.1731.1671.1621.157
MLIP1.1431.1381.1341.1291.1241.118
% error3.433.383.293.313.253.40
The ability of the trained model to reproduce the equilibrium structure of the reline mixture was tested by comparing the pair distribution functions from MD simulations using MLIP to an FPMD reference. A mixture of 18 ChCl and 36 urea molecules was simulated at 375 K for 0.5 ns. As a reference, a short FPMD trajectory was generated for a system of the same size using our target DFT method. The computed radial distribution functions (RDFs) in Figure 3, for different interaction sites on choline and urea, are in excellent agreement with the FPMD reference. Considering that none of the structures in the training data set were generated using our reference DFT potential, reference energies and atomic forces are sufficient to train an accurate MLIP, regardless of the method used to sample or generate the training structures. Additionally, RDFs of chloride (Cl) and urea’s oxygen (OU) as hydrogen-bond acceptors around hydrogen atoms of urea separated for the cis and trans hydrogen atoms, as well as selected intramolecular bond lengths and bond angle distribution functions, are also presented and compared to the FPMD reference in the Supporting Information. In Figure S4, the RDFs of OU and Cl around the hydrogen atom(s) of choline (H4) and urea (HU) are compared to selected classical and polarizable FF simulations from the literature. The RDFs around HU atoms from FF simulations are in good agreement with the MLIP results, but the selected FFs could not accurately predict the RDFs around the hydrogen atom of choline. The polarizable FF overestimates the intensity of the Cl–H4 interaction and underestimates the intensity of the OU–H4 interactions. The classical FF underestimates the intensity of both interactions with H4 atoms, most likely as a result of scaled atomic charges, but the relative intensity of the OU and Cl RDFs around H4 is in qualitative agreement with the MLIP results.

Figure 3

Figure 3. Comparison of the redial distribution functions from the FPMD simulation (dashed lines) and simulations using the MLIP (solid lines) for interactions with H4 and N1 of choline (top) and hydrogen atoms of urea HU (bottom). Atom labels are given in Figure 1.

3.3. Translational Dynamics

In Figure 4, as depicted by the center-of-mass velocity autocorrelation functions (VACFs)
A(t)=vi(0)·vi(t)vi21
(6)
the equilibrium diffusive dynamics of the three components of reline mixture are successfully reproduced compared to the reference FPMD simulations.

Figure 4

Figure 4. Center-of-mass velocity autocorrelation functions A(t) for the three components of the reline mixture from the reference FPMD simulations (dashed lines) are compared to the MLIP simulations (solid lines) for a mixture of 18 ChCl and 36 urea molecules at 375 K.

The self-diffusion coefficients (D) were calculated from the MD trajectories of a mixture of 64 ChCl and 128 urea molecules using the mean-square displacement of the molecular center-of-mass and the Einstein relation
D=16limtddti=1N[ri(0)ri(t)]2
(7)
where ri is the position of the center of mass of particle i at any specific time and the brackets indicate the average over all particles and over all time origins. The mean-square displacements of the center of mass for three components of reline are shown in Figure S7, and the diffusion coefficients are obtained from a linear fit to the second half of the plots. Reline is a highly viscous liquid, and its viscosity is comparable to some room-temperature ionic liquids. This makes the prediction of the diffusion coefficients from MD simulations challenging, and as it has been shown before, long simulation times are needed to converge the dynamical properties at room temperature. (71) To investigate the suitability of single trajectories for computing the dynamical properties such as diffusion coefficients, at a selected temperature of 353 K, 10 separate trajectories were generated and the results are presented in Figure S6 of the Supporting Information. As is evident from the obtained results, at least five trajectories are needed for sufficient sampling and more reliable calculation of diffusion coefficients from our MD simulations. Considering the efficiency of the MD simulations using our ML potential, as 1 ns simulation of a system of ∼ 2500 atoms on a single graphics processing unit (GPU) takes about 9 h, the additional computational cost of generating more trajectories is overall small. The self-diffusion coefficients in Figure 5, for the three components of reline mixture, are obtained from the average of five different trajectories and are corrected for the finite-size effects of the MD simulations using the correction term of Yeh and Hummer (eq S1). (72) The lines in Figure 5 are the best exponential fits to the data used to obtain the activation energies of diffusion using Arrhenius’ law
D=D0exp(EaRT)
(8)
with pre-exponential factor D0 and activation energy of diffusion Ea. The experimental values (73) of self-diffusion coefficients for the cation and urea are also plotted in Figure 5, and the calculated activation energies of diffusion are summarized in Table 2. The self-diffusion coefficients from our MD simulations in Figure 5 underestimate the reported experimental results by a factor of 3–4, which could be attributed to the limited accuracy of the reference DFT method. The obtained self-diffusion coefficients are similar to the reported results from MD simulations using a polarizable FF (22) and are ∼10 times higher than the obtained values from nonpolarizable FFs that do not empirically scale the charge on the ions. Furthermore, the obtained activation energies of diffusion match very well with the experimental values, indicating the ability of the MLIP to accurately predict the temperature dependence of the diffusive dynamics in reline.

Figure 5

Figure 5. Calculated values of self-diffusion coefficients for the three components of reline mixture from MD simulations (filled symbols) are plotted against the inverse of the simulation temperature and compared to the experimental values (empty symbols) of ref (73). Lines are the best exponential fits to the data used to obtain the activation energies listed in Table 2.

Table 2. Activation Energies of Diffusion (in kJ mol–1) for the Three Components of Reline Mixture Calculated from an Arrhenius Exponential Fit to the Data in Figure 5 Are Compared to the Reported Experimental Values of Ref (73)
 Exp.MLIP
cation47.8 ± 0.148.26
anion 46.83
urea45.0 ± 0.146.76
To better understand the collective motions of the different components in reline, the velocity cross-correlation functions were computed. The time cross-correlation function between the initial velocity of a particle of species i and the latter velocities of particles of species j, initially located inside a spherical shell of radius R centered at the particle of species i, are defined as
C(t)=vi(0)·vj(t)r(vi2vj2)1/2
(9)
where vi(t) and vj(t) are the velocity of the particles i and j, ⟨vj2⟩ is the mean-squared velocities of all particles of species i and j, and ⟨ ⟩r is a restricted statistical average
vi(0)·vj(t)r=N1jvi(0)·vj(t)·u(Rrij(0))
(10)
where u(x) is the step function, rij(0) is the initial distance between the central particle i and a particle j, and the brackets indicate the average over all particles i and over all time origins. N is the coordination number or the mean number of j particles in the spherical shell of particle i. The values of R for each cross-correlation function between species i and j is set to the position of the first minima of the corresponding center-of-mass RDFs in Figure S5 and are listed in Table 3. The functions NC(t) in Figure 6, represent the sum of the velocity cross-correlations between a central particle and all the neighboring particles within the first shell and are plotted alongside the VACFs. The values of N for different pair of species are listed in Table 3. It should be noted that due to the finite size of the system, initial values of the velocity cross-correlation functions are less than 0. (74)
Table 3. Mean Number of Particles (Nij) of Species j in a Spherical Shell of Radius R (in Å) Around a Particle of Species ia
 CCCACUAAACAUUUUCUA
Nij4.893.319.542.863.313.634.914.771.81
R8.06.57.47.06.55.36.27.45.3
a

i and j are choline cation (C), chloride anion (A), or urea molecule (U).

Figure 6

Figure 6. Velocity cross-correlation functions NC(t) for the cation as the central particle (top), anion as the central particle (middle), and urea as the central particle (bottom). The dashed lines are the velocity autocorrelation functions A(t) of the cation (gray), anion (green), and urea (red). The cross-correlation functions for longer correlation times are presented in Figures S8–S10.

Figure 6 shows the velocity cross-correlation functions between the different components of reline DES at 363 K. The initial rise of the velocity cross-correlation functions, with the decay of the corresponding VACFs toward the minimum, indicates the transfer of the initial momentum of the central particle to the neighboring particles, which on average tend to move in the same direction. The decay of NC(t) functions after the first peak should be associated with the spread of the momentum to the particles in the outer shells. Furthermore, the cross-correlation functions show a negative region at intermediate times, which could be attributed to the rebound between the central particle and particles within the first shell. (75) With the exception of the cross-correlation function between distinct anions, which does not show a clear peak, the velocity cross-correlation functions between species with a larger mass show a slower increase and reach their first maximum later than those between lighter ones. In the case of correlation between cations and urea molecules (cation-urea and urea-cation) and between urea molecules and anions (urea-anion and anion-urea), the velocity cross-correlation functions reach their maximum at the same time that the particles with larger mass lose their initial momentum and the corresponding VACF becomes negative. These results are in agreement with the findings for dense simple liquids. (76−78) However, the cross-correlation functions between the cations and anions reach their maximum shortly before the VACF of cations becomes negative. This suggests that anions on average lose the momentum, they initially gained from the cations, to their surrounding particles faster than urea molecules. In the case of cross-correlations between distinct like species, the velocity cross-correlation between distinct urea molecules becomes maximum at the same time that the corresponding VACF becomes negative. However, the velocity cross-correlation function between distinct cations reaches its maximum shortly after the corresponding VACF becomes negative, indicating that the heavier cations on average maintain their initially gained momentum for longer times. Furthermore, a clear relationship between the oscillations of the urea-anion (also anion-urea) and urea–urea cross-correlation function with those of the urea’s VACF is visible, suggesting that a part of the initial momentum transferred to the neighboring particles is returned to a urea molecule. (79,80) In the middle panel of Figure 6, even though the anion–cation cross-correlation function does not show clear oscillations after the first maximum, a noticeable phase relationship with the oscillations of the anion-urea velocity cross-correlation function is visible. The maxima of the anion-urea cross-correlation function coincide with the minima of the anion–cation cross-correlation function and vice versa. These features could be attributed to the oscillatory motion of the anions surrounded by the heavier cations and urea molecules. (23,81)

3.4. Ionic Conductivity

The ionic conductivity (σ) of the simulated mixtures was calculated by integrating the charge current autocorrelation function using the following Green–Kubo relation
σ=13VkBT0j(0)·j(t)dt
(11)
where kB is the Boltzmann constant, T is the temperature, V is the volume of the simulation box, and the brackets indicate the average over all time origins. The charge current j(t) is given by
j(t)=i=1Nqivi(t)
(12)
where N is the number of ions and qi and vi(t) represent the charge and the center-of-mass velocity of the ith ion at time t, respectively. The ionic conductivity (σ) is obtained from the long-time limit of an exponential decay function fitted to the running integral of the charge current autocorrelation function. It is difficult to evaluate the true conductivity through this approach since conductivity is a collective dynamical property and, for this reason, suffers from a relatively high statistical uncertainty. (82) However, the ability to calculate its value from a larger number of MD simulations leads to considerably more sampling and more reliable results. The obtained ionic conductivities were corrected for the finite-size effect using the same relative correction factor as for diffusion coefficients of ionic species and are plotted alongside a set of experimental values in Figure 7. The reported experimental values for the ionic conductivity of reline cover a wider range, but the results from more recent measurements are in better agreement and are expected to be more accurate. The estimated values of ionic conductivity from our MD simulations, as is expected from the self-diffusion coefficient results, are closer to the lower end of the range of experimental values and underestimate the more recent results by a factor of 1.5–2. The temperature dependence of ionic conductivity is best described by the Vogel–Fulcher–Tammann (VFT) equation
σ=σ0exp(BTTVF)
(13)
with pre-exponential factor σ0, fitting parameter B, and the Vogel–Fulcher temperature Tvf. The dashed line in Figure 7 is the VFT fit to the data with pre-exponential factor (σ0) of 2585 S m–1, 1941, and 112 K for B and Tvf, respectively.

Figure 7

Figure 7. Calculated values of ionic conductivity from MD simulations using our ML potential are plotted against the inverse of the simulation temperature and compared to the experimental values from refs (1), (68), and (83). Lines are the VFT fits to the data.

The ionic conductivity can also be estimated using the Nernst–Einstein (NE) expression from the self-diffusion coefficients of ions
σNE=NpairVkBTi=1Nqi2Di
(14)
where Npair is the number of ion pairs, D is the diffusion coefficient, and i denotes cations or anions. The NE expression assumes that the diffusion of individual ions contributes entirely to the ionic conductivity of the system. Consequently, the deviations of the Green–Kubo conductivity from the NE conductivity can be interpreted as cross-correlations among ions. The obtained values for ionic conductivity from the Green–Kubo and NE expressions and the corresponding ratio γ = σ/σNE are listed in Table 4. The NE conductivities are calculated using the self-diffusion coefficients obtained from the VACFs. The self-diffusion coefficients are estimated from an exponential fit to the running integral of VACFs and are listed in Table S2 of the Supporting Information. The ratio γ is the percentage of ions that contribute to the ionic conductivity and can be seen as the degree of uncorrelated motion of ions. A γ = 1 indicates that the ion motions are completely uncorrelated, and γ values of less than one point to some degree of correlated motion between oppositely charged ions. Moreover, the obtained results suggest a temperature-dependent trend for the σ/σNE ratio in reline.
Table 4. Estimated Values of Ionic Conductivity from the Green–Kubo (σ) and NE (σNE) Approach (in S m–1) and the Corresponding Ratio (γ = σ/σNE) at the Range of Studied Temperatures
T [K]323333343353363373
σ0.2830.3920.5830.8411.1621.543
σNE0.2450.3610.6100.9411.3901.962
γ1.151.080.960.890.840.79

4. Conclusions

Click to copy section linkSection link copied!

We trained an MLIP for reline DES, on a set of atomic configurations labeled with DFT energies and forces, using the DeePMD framework. To the best of our knowledge, this is the first time that the ability of MLIPs to represent the PES of DESs has been explored and evaluated. Our trained model achieves small prediction errors for energies and atomic forces on a set of unseen atomic configurations. Using our MLIP, we are able to run fast and scalable MD simulations with the target DFT accuracy for thousands of atoms and nanosecond time scales. The obtained structural and dynamical properties of reline from our MD simulations are in good agreement with the results from experiments and first-principles MD simulations. Furthermore, we investigated the collective motions of the three components of reline DES using velocity cross-correlation functions, showing that there are positive correlations at short time scales, mainly due to the transfer of momentum from species with a larger mass to surrounding particles. Additionally, we showed that for a reliable investigation of dynamical properties, such as self-diffusion coefficients, a larger number of simulation runs are required. Considering the size and time scale of the performed simulations, MLIPs enable these types of analyses that otherwise are not accessible by the first-principles methods.
We used an iterative active learning scheme based on uncertainty estimates from an ensemble of models to efficiently explore the underrepresented regions of the configuration space in the training data set and systematically improve our model. The use of an active learning approach is considered an important step in the construction of a robust ML potential. Additionally, the uncertainty estimates of the predicted forces by the ML model are a valuable metric to validate the fidelity of the MD simulation runs. It should be noted that the quality of the uncertainty quantification method plays an important role in active learning. It has been shown that the ensemble-based uncertainty quantification, compared to single-model quantification methods, performs consistently better across different data sets albeit at a higher computational cost. (84)
In the DeePMD model, similar to many other ML potentials, the total energy of the system is predicted as the sum of atomic energies that are a function of local atomic descriptors. Thus, it is considered a short-range potential in a sense that it does not include a separate treatment of long-range interactions. Although, the long-range contributions to the total energy are present in the reference training data and hence are implicitly accounted for in the fitting process. Nevertheless, as it has been shown before and the current results suggest, an accurate approximation of the PES of polarizable bulk liquids (85) and dense ionic systems (86) by ML potentials does not necessarily require a specific treatment of long-range interactions. This can be rationalized in terms of an efficient screening of electrostatic interactions in a bulk system. However, in systems lacking translational invariance such as extended interfaces, truncated interactions result in unbalanced forces, which are difficult to approximate with short-range ML potentials, and specific treatment of long-range interactions becomes important. (85)

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.3c00944.

  • Details of the DFTB-based MD simulations, additional radial and angular distribution functions, simulation box sizes and viscosities, mean-square displacement plots and self-diffusion coefficients from the Green–Kubo approach with details of the correction term for the finite-size effects, and velocity cross-correlation functions for longer correlation times (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

Acknowledgments

Click to copy section linkSection link copied!

The authors thank the German Research Foundation (DFG) for financial support (project ZA 606/8-1). Computational time from the Center for Information Services and High Performance Computing (ZIH) of TU Dresden is gratefully acknowledged.

References

Click to copy section linkSection link copied!

This article references 86 other publications.

  1. 1
    Abbott, A. P.; Capper, G.; Davies, D. L.; Rasheed, R. K.; Tambyrajah, V. Novel solvent properties of choline chloride/urea mixturesElectronic supplementary information (ESI) available: spectroscopic data. See http://www.rsc.org/suppdata/cc/b2/b210714g/. Chem. Commun. 2002, 7071,  DOI: 10.1039/b210714g
  2. 2
    Zhang, Q.; De Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep eutectic solvents: syntheses, properties and applications. Chem. Soc. Rev. 2012, 41, 71087146,  DOI: 10.1039/c2cs35178a
  3. 3
    Smith, E. L.; Abbott, A. P.; Ryder, K. S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 1106011082, PMID: 25300631  DOI: 10.1021/cr300162p
  4. 4
    Hansen, B. B.; Spittle, S.; Chen, B.; Poe, D.; Zhang, Y.; Klein, J. M.; Horton, A.; Adhikari, L.; Zelovich, T.; Doherty, B. W.; Gurkan, B.; Maginn, E. J.; Ragauskas, A.; Dadmun, M.; Zawodzinski, T. A.; Baker, G. A.; Tuckerman, M. E.; Savinell, R. F.; Sangoro, J. R. Deep Eutectic Solvents: A Review of Fundamentals and Applications. Chem. Rev. 2021, 121, 12321285, PMID: 33315380  DOI: 10.1021/acs.chemrev.0c00385
  5. 5
    Abbott, A. P.; Capper, G.; Davies, D. L.; McKenzie, K. J.; Obi, S. U. Solubility of Metal Oxides in Deep Eutectic Solvents Based on Choline Chloride. J. Chem. Eng. Data 2006, 51, 12801282,  DOI: 10.1021/je060038c
  6. 6
    Jenkin, G. R.; Al-Bassam, A. Z.; Harris, R. C.; Abbott, A. P.; Smith, D. J.; Holwell, D. A.; Chapman, R. J.; Stanley, C. J. The application of deep eutectic solvent ionic liquids for environmentally-friendly dissolution and recovery of precious metals. Miner. Eng. 2016, 87, 1824, Processing of Precious Metal Ores  DOI: 10.1016/j.mineng.2015.09.026
  7. 7
    Söldner, A.; Zach, J.; König, B. Deep eutectic solvents as extraction media for metal salts and oxides exemplarily shown for phosphates from incinerated sewage sludge ash. Green Chem. 2019, 21, 321328,  DOI: 10.1039/C8GC02702A
  8. 8
    Chakrabarti, M. H.; Mjalli, F. S.; AlNashef, I. M.; Hashim, M. A.; Hussain, M. A.; Bahadori, L.; Low, C. T. J. Prospects of applying ionic liquids and deep eutectic solvents for renewable energy storage by means of redox flow batteries. Renew. Sustain. Energy Rev. 2014, 30, 254270,  DOI: 10.1016/j.rser.2013.10.004
  9. 9
    Cong, G.; Lu, Y.-C. Organic Eutectic Electrolytes for Future Flow Batteries. Chem 2018, 4, 27322734,  DOI: 10.1016/j.chempr.2018.11.018
  10. 10
    Radošević, K.; Cvjetko Bubalo, M.; Gaurina Srček, V.; Grgas, D.; Landeka Dragičević, T.; Radojčić Redovniković, I. Evaluation of toxicity and biodegradability of choline chloride based deep eutectic solvents. Ecotoxicol. Environ. Saf. 2015, 112, 4653,  DOI: 10.1016/j.ecoenv.2014.09.034
  11. 11
    Halder, A. K.; Cordeiro, M. N. D. S. Probing the Environmental Toxicity of Deep Eutectic Solvents and Their Components: An In Silico Modeling Approach. ACS Sustain. Chem. Eng. 2019, 7, 1064910660,  DOI: 10.1021/acssuschemeng.9b01306
  12. 12
    Khandelwal, S.; Tailor, Y. K.; Kumar, M. Deep eutectic solvents (DESs) as eco-friendly and sustainable solvent/catalyst systems in organic transformations. J. Mol. Liq. 2016, 215, 345386,  DOI: 10.1016/j.molliq.2015.12.015
  13. 13
    Zahn, S. Deep eutectic solvents: similia similibus solvuntur?. Phys. Chem. Chem. Phys. 2017, 19, 40414047,  DOI: 10.1039/C6CP08017K
  14. 14
    Stefanovic, R.; Ludwig, M.; Webber, G. B.; Atkin, R.; Page, A. J. Nanostructure, hydrogen bonding and rheology in choline chloride deep eutectic solvents as a function of the hydrogen bond donor. Phys. Chem. Chem. Phys. 2017, 19, 32973306,  DOI: 10.1039/C6CP07932F
  15. 15
    Fetisov, E. O.; Harwood, D. B.; Kuo, I.-F. W.; Warrag, S. E. E.; Kroon, M. C.; Peters, C. J.; Siepmann, J. I. First-Principles Molecular Dynamics Study of a Deep Eutectic Solvent: Choline Chloride/Urea and Its Mixture with Water. J. Phys. Chem. B 2018, 122, 12451254, PMID: 29200290  DOI: 10.1021/acs.jpcb.7b10422
  16. 16
    Perkins, S. L.; Painter, P.; Colina, C. M. Molecular Dynamic Simulations and Vibrational Analysis of an Ionic Liquid Analogue. J. Phys. Chem. B 2013, 117, 1025010260, PMID: 23915257  DOI: 10.1021/jp404619x
  17. 17
    Doherty, B.; Acevedo, O. OPLS Force Field for Choline Chloride-Based Deep Eutectic Solvents. J. Phys. Chem. B 2018, 122, 99829993, PMID: 30125108  DOI: 10.1021/acs.jpcb.8b06647
  18. 18
    Nandy, A.; Smiatek, J. Mixtures of LiTFSI and urea: ideal thermodynamic behavior as key to the formation of deep eutectic solvents?. Phys. Chem. Chem. Phys. 2019, 21, 1227912287,  DOI: 10.1039/C9CP01440C
  19. 19
    Zhang, Y.; Poe, D.; Heroux, L.; Squire, H.; Doherty, B. W.; Long, Z.; Dadmun, M.; Gurkan, B.; Tuckerman, M. E.; Maginn, E. J. Liquid Structure and Transport Properties of the Deep Eutectic Solvent Ethaline. J. Phys. Chem. B 2020, 124, 52515264, PMID: 32464060  DOI: 10.1021/acs.jpcb.0c04058
  20. 20
    Shayestehpour, O.; Zahn, S. Molecular Features of Reline and Homologous Deep Eutectic Solvents Contributing to Nonideal Mixing Behavior. J. Phys. Chem. B 2020, 124, 75867597, PMID: 32790398  DOI: 10.1021/acs.jpcb.0c03091
  21. 21
    Goloviznina, K.; Gong, Z.; Costa Gomes, M. F.; Pádua, A. A. H. Extension of the CL&Pol Polarizable Force Field to Electrolytes, Protic Ionic Liquids, and Deep Eutectic Solvents. J. Chem. Theory Comput. 2021, 17, 16061617, PMID: 33555860  DOI: 10.1021/acs.jctc.0c01002
  22. 22
    Jeong, K.-j.; McDaniel, J. G.; Yethiraj, A. Deep Eutectic Solvents: Molecular Simulations with a First-Principles Polarizable Force Field. J. Phys. Chem. B 2021, 125, 71777186, PMID: 34181852  DOI: 10.1021/acs.jpcb.1c01692
  23. 23
    Shayestehpour, O.; Zahn, S. Ion Correlation in Choline Chloride–Urea Deep Eutectic Solvent (Reline) from Polarizable Molecular Dynamics Simulations. J. Phys. Chem. B 2022, 126, 34393449, PMID: 35500254  DOI: 10.1021/acs.jpcb.1c10671
  24. 24
    Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133A1138,  DOI: 10.1103/PhysRev.140.A1133
  25. 25
    Goedecker, S. Linear scaling electronic structure methods. Rev. Mod. Phys. 1999, 71, 10851123,  DOI: 10.1103/RevModPhys.71.1085
  26. 26
    Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996, 118, 1122511236,  DOI: 10.1021/ja9621760
  27. 27
    Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 2010, 31, 671690,  DOI: 10.1002/jcc.21367
  28. 28
    Cadena, C.; Zhao, Q.; Snurr, R. Q.; Maginn, E. J. Molecular Modeling and Experimental Studies of the Thermodynamic and Transport Properties of Pyridinium-Based Ionic Liquids. J. Phys. Chem. B 2006, 110, 28212832, PMID: 16471891  DOI: 10.1021/jp056235k
  29. 29
    Rajput, N. N.; Murugesan, V.; Shin, Y.; Han, K. S.; Lau, K. C.; Chen, J.; Liu, J.; Curtiss, L. A.; Mueller, K. T.; Persson, K. A. Elucidating the Solvation Structure and Dynamics of Lithium Polysulfides Resulting from Competitive Salt and Solvent Interactions. Chem. Mater. 2017, 29, 33753379,  DOI: 10.1021/acs.chemmater.7b00068
  30. 30
    Yan, T.; Wang, Y.; Knox, C. On the Dynamics of Ionic Liquids: Comparisons between Electronically Polarizable and Nonpolarizable Models II. J. Phys. Chem. B 2010, 114, 68866904, PMID: 20443608  DOI: 10.1021/jp908914d
  31. 31
    Szabadi, A.; Elfgen, R.; Macchieraldo, R.; Kearns, F. L.; Lee Woodcock, H.; Kirchner, B.; Schröder, C. Comparison between ab initio and polarizable molecular dynamics simulations of 1-butyl-3-methylimidazolium tetrafluoroborate and chloride in water. J. Mol. Liq. 2021, 337, 116521,  DOI: 10.1016/j.molliq.2021.116521
  32. 32
    Glielmo, A.; Sollich, P.; De Vita, A. Accurate interatomic force fields via machine learning with covariant kernels. Phys. Rev. B 2017, 95, 214302,  DOI: 10.1103/PhysRevB.95.214302
  33. 33
    Chmiela, S.; Tkatchenko, A.; Sauceda, H. E.; Poltavsky, I.; Schütt, K. T.; Müller, K. R. Machine learning of accurate energy-conserving molecular force fields. Sci. Adv. 2017, 3, e1603015  DOI: 10.1126/sciadv.1603015
  34. 34
    Chmiela, S.; Sauceda, H. E.; Müller, K. R.; Tkatchenko, A. Towards exact molecular dynamics simulations with machine-learned force fields. Nat. Commun. 2018, 9, 3887,  DOI: 10.1038/s41467-018-06169-2
  35. 35
    Bartók, A. P.; Payne, M. C.; Kondor, R.; Csányi, G. Gaussian Approximation Potentials: The Accuracy of Quantum Mechanics, without the Electrons. Phys. Rev. Lett. 2010, 104, 136403,  DOI: 10.1103/PhysRevLett.104.136403
  36. 36
    Behler, J.; Parrinello, M. Generalized Neural-Network Representation of High-Dimensional Potential-Energy Surfaces. Phys. Rev. Lett. 2007, 98, 146401,  DOI: 10.1103/PhysRevLett.98.146401
  37. 37
    Behler, J. Representing potential energy surfaces by high-dimensional neural network potentials. J. Phys.: Condens. Matter 2014, 26, 183001,  DOI: 10.1088/0953-8984/26/18/183001
  38. 38
    Ghasemi, S. A.; Hofstetter, A.; Saha, S.; Goedecker, S. Interatomic potentials for ionic systems with density functional accuracy based on charge densities obtained by a neural network. Phys. Rev. B 2015, 92, 045131,  DOI: 10.1103/PhysRevB.92.045131
  39. 39
    Smith, J. S.; Isayev, O.; Roitberg, A. E. ANI-1: an extensible neural network potential with DFT accuracy at force field computational cost. Chem. Sci. 2017, 8, 31923203,  DOI: 10.1039/C6SC05720A
  40. 40
    Yao, K.; Herr, J. E.; Toth, D. W.; Mckintyre, R.; Parkhill, J. The TensorMol-0.1 model chemistry: a neural network augmented with long-range physics. Chem. Sci. 2018, 9, 22612269,  DOI: 10.1039/C7SC04934J
  41. 41
    Schütt, K. T.; Sauceda, H. E.; Kindermans, P.-J.; Tkatchenko, A.; Müller, K. R. SchNet – A deep learning architecture for molecules and materials. J. Chem. Phys. 2018, 148, 241722,  DOI: 10.1063/1.5019779
  42. 42
    Han, J.; Zhang, L.; Car, R.; E, W. Deep Potential: A General Representation of a Many-Body Potential Energy Surface. Commun. Comput. Phys. 2018, 23, 629639,  DOI: 10.4208/cicp.oa-2017-0213
  43. 43
    Unke, O. T.; Meuwly, M. PhysNet: A Neural Network for Predicting Energies, Forces, Dipole Moments, and Partial Charges. J. Chem. Theory Comput. 2019, 15, 36783693, PMID: 31042390  DOI: 10.1021/acs.jctc.9b00181
  44. 44
    Behler, J. Atom-centered symmetry functions for constructing high-dimensional neural network potentials. J. Chem. Phys. 2011, 134, 074106,  DOI: 10.1063/1.3553717
  45. 45
    Thompson, A.; Swiler, L.; Trott, C.; Foiles, S.; Tucker, G. Spectral neighbor analysis method for automated generation of quantum-accurate interatomic potentials. J. Comput. Phys. 2015, 285, 316330,  DOI: 10.1016/j.jcp.2014.12.018
  46. 46
    Drautz, R. Atomic cluster expansion for accurate and transferable interatomic potentials. Phys. Rev. B 2019, 99, 014104,  DOI: 10.1103/PhysRevB.99.014104
  47. 47
    Batzner, S.; Musaelian, A.; Sun, L.; Geiger, M.; Mailoa, J. P.; Kornbluth, M.; Molinari, N.; Smidt, T. E.; Kozinsky, B. E(3)-equivariant graph neural networks for data-efficient and accurate interatomic potentials. Nat. Commun. 2022, 13, 2453,  DOI: 10.1038/s41467-022-29939-5
  48. 48
    Schütt, K., Unke, O., Gastegger, M. Equivariant message passing for the prediction of tensorial properties and molecular spectra. International Conference on Machine Learning 2021, 93779388.
  49. 49
    Haghighatlari, M.; Li, J.; Guan, X.; Zhang, O.; Das, A.; Stein, C. J.; Heidar-Zadeh, F.; Liu, M.; Head-Gordon, M.; Bertels, L.; Hao, H.; Leven, I.; Head-Gordon, T. NewtonNet: a Newtonian message passing network for deep learning of interatomic potentials and forces. Digital Discovery 2022, 1, 333343,  DOI: 10.1039/D2DD00008C
  50. 50
    Musaelian, A.; Batzner, S.; Johansson, A.; Sun, L.; Owen, C. J.; Kornbluth, M.; Kozinsky, B. Learning local equivariant representations for large-scale atomistic dynamics. Nat. Commun. 2023, 14, 579,  DOI: 10.1038/s41467-023-36329-y
  51. 51
    Zhang, L.; Han, J.; Wang, H.; Saidi, W. A.; Car, R.; Weinan, E.: Red Hook, NY, USA, 2018, pp 44414451. End-to-End Symmetry Preserving Inter-Atomic Potential Energy Model for Finite and Extended Systems Proceedings of the 32nd International Conference on Neural Information Processing Systems
  52. 52
    Wang, H.; Zhang, L.; Han, J.; E, W. DeePMD-kit: A deep learning package for many-body potential energy representation and molecular dynamics. Comput. Phys. Commun. 2018, 228, 178184,  DOI: 10.1016/j.cpc.2018.03.016
  53. 53
    Kingma, D. P.; Ba, J. Adam: A Method for Stochastic Optimization. 2014, arXiv:1412.6980. https://arxiv.org/abs/1412.698web0.
  54. 54
    Frauenheim, T.; Seifert, G.; Elsterner, M.; Hajnal, Z.; Jungnickel, G.; Porezag, D.; Suhai, S.; Scholz, R. A Self-Consistent Charge Density-Functional Based Tight-Binding Method for Predictive Materials Simulations in Physics, Chemistry and Biology. Phys. Status Solidi B 2000, 217, 4162,  DOI: 10.1002/(SICI)1521-3951(200001)217:1<41::AID-PSSB41>3.0.CO;2-V
  55. 55
    VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. Quickstep: Fast and accurate density functional calculations using a mixed Gaussian and plane waves approach. Comput. Phys. Commun. 2005, 167, 103128,  DOI: 10.1016/j.cpc.2004.12.014
  56. 56
    Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. cp2k: atomistic simulations of condensed matter systems. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2014, 4, 1525,  DOI: 10.1002/wcms.1159
  57. 57
    Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 30983100,  DOI: 10.1103/PhysRevA.38.3098
  58. 58
    Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785789,  DOI: 10.1103/PhysRevB.37.785
  59. 59
    Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104,  DOI: 10.1063/1.3382344
  60. 60
    Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 17031710,  DOI: 10.1103/PhysRevB.54.1703
  61. 61
    Hartwigsen, C.; Goedecker, S.; Hutter, J. Relativistic separable dual-space Gaussian pseudopotentials from H to Rn. Phys. Rev. B 1998, 58, 36413662,  DOI: 10.1103/PhysRevB.58.3641
  62. 62
    Smith, J. S.; Nebgen, B.; Lubbers, N.; Isayev, O.; Roitberg, A. E. Less is more: Sampling chemical space with active learning. J. Chem. Phys. 2018, 148, 241733,  DOI: 10.1063/1.5023802
  63. 63
    Podryabinkin, E. V.; Shapeev, A. V. Active learning of linearly parametrized interatomic potentials. Comput. Mater. Sci. 2017, 140, 171180,  DOI: 10.1016/j.commatsci.2017.08.031
  64. 64
    Peterson, A. A.; Christensen, R.; Khorshidi, A. Addressing uncertainty in atomistic machine learning. Phys. Chem. Chem. Phys. 2017, 19, 1097810985,  DOI: 10.1039/C7CP00375G
  65. 65
    Zhang, Y.; Wang, H.; Chen, W.; Zeng, J.; Zhang, L.; Wang, H. .. DP-GEN: A concurrent learning platform for the generation of reliable deep learning based potential energy models. Comput. Phys. Commun. 2020, 253, 107206,  DOI: 10.1016/j.cpc.2020.107206
  66. 66
    Thompson, A. P.; Aktulga, H. M.; Berger, R.; Bolintineanu, D. S.; Brown, W. M.; Crozier, P. S.; in ’t Veld, P. J.; Kohlmeyer, A.; Moore, S. G.; Nguyen, T. D.; Shan, R.; Stevens, M. J.; Tranchida, J.; Trott, C.; Plimpton, S. J. LAMMPS - a flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comput. Phys. Commun. 2022, 271, 108171,  DOI: 10.1016/j.cpc.2021.108171
  67. 67
    Martyna, G. J.; Tobias, D. J.; Klein, M. L. Constant pressure molecular dynamics algorithms. J. Chem. Phys. 1994, 101, 41774189,  DOI: 10.1063/1.467468
  68. 68
    Agieienko, V.; Buchner, R. Densities, Viscosities, and Electrical Conductivities of Pure Anhydrous Reline and Its Mixtures with Water in the Temperature Range (293.15 to 338.15) K. J. Chem. Eng. Data 2019, 64, 47634774,  DOI: 10.1021/acs.jced.9b00145
  69. 69
    Fu, X.; Wu, Z.; Wang, W.; Xie, T.; Keten, S.; Gomez-Bombarelli, R.; Jaakkola, T. Forces are not Enough: Benchmark and Critical Evaluation for Machine Learning Force Fields with Molecular Simulations. 2022, arXiv:2210.07237. https://arxiv.org/abs/2210.0723web7.
  70. 70
    Stocker, S.; Gasteiger, J.; Becker, F.; Günnemann, S.; Margraf, J. T. How robust are modern graph neural network potentials in long and hot molecular dynamics simulations?. Machine Learning: Science and Technology 2022, 3, 045010,  DOI: 10.1088/2632-2153/ac9955
  71. 71
    McDaniel, J. G.; Choi, E.; Son, C. Y.; Schmidt, J. R.; Yethiraj, A. Ab Initio Force Fields for Imidazolium-Based Ionic Liquids. J. Phys. Chem. B 2016, 120, 70247036, PMID: 27352240  DOI: 10.1021/acs.jpcb.6b05328
  72. 72
    Yeh, I.-C.; Hummer, G. System-Size Dependence of Diffusion Coefficients and Viscosities from Molecular Dynamics Simulations with Periodic Boundary Conditions. J. Phys. Chem. B 2004, 108, 1587315879,  DOI: 10.1021/jp0477147
  73. 73
    D’Agostino, C.; Harris, R. C.; Abbott, A. P.; Gladden, L. F.; Mantle, M. D. Molecular motion and ion diffusion in choline chloride based deep eutectic solvents studied by 1H pulsed field gradient NMR spectroscopy. Phys. Chem. Chem. Phys. 2011, 13, 2138321391,  DOI: 10.1039/c1cp22554e
  74. 74
    Balucani, U.; Vallauri, R.; Murthy, C. S. Interparticle velocity correlations in simple liquids. J. Chem. Phys. 1982, 77, 32333237,  DOI: 10.1063/1.444199
  75. 75
    Verdaguer, A.; Padró, J. A.; Trullàs, J. Molecular dynamics study of the velocity cross-correlations in liquids. J. Chem. Phys. 1998, 109, 228234,  DOI: 10.1063/1.476555
  76. 76
    Balucani, U.; Vallauri, R.; Murthy, C. Momentum transfer analysis in Lennard-Jones fluids. Phys. Lett. A 1981, 84, 133136,  DOI: 10.1016/0375-9601(81)90736-2
  77. 77
    Endo, Y.; Endo, H. Microscopic motions and the local environments of atoms in simple liquids. J. Chem. Phys. 1984, 80, 20872091,  DOI: 10.1063/1.446974
  78. 78
    Flener, P.; Vesely, F. J. Correlated motion of two particles in a fluid. Mol. Phys. 1992, 77, 601615,  DOI: 10.1080/00268979200102651
  79. 79
    Balucani, U.; Zoppi, M. Dynamics of the Liquid State; Clarendon Press, 1995; Vol. 10.
  80. 80
    Verdaguer, A.; Padró, J. A. Computer simulation study of the velocity cross correlations between neighboring atoms in simple liquid binary mixtures. J. Chem. Phys. 2001, 114, 27382744,  DOI: 10.1063/1.1340581
  81. 81
    Hansen, J.-P.; McDonald, I. R. Theory of Simple Liquids, 3rd ed.; Hansen, J.-P., McDonald, I. R., Eds.; Academic Press: Burlington, 2006; pp 291340.
  82. 82
    Rey-Castro, C.; Vega, L. F. Transport Properties of the Ionic Liquid 1-Ethyl-3-Methylimidazolium Chloride from Equilibrium Molecular Dynamics Simulation. The Effect of Temperature. J. Phys. Chem. B 2006, 110, 1442614435, PMID: 16854152  DOI: 10.1021/jp062885s
  83. 83
    Reuter, D.; Binder, C.; Lunkenheimer, P.; Loidl, A. Ionic conductivity of deep eutectic solvents: the role of orientational dynamics and glassy freezing. Phys. Chem. Chem. Phys. 2019, 21, 68016809,  DOI: 10.1039/C9CP00742C
  84. 84
    Tan, A. R.; Urata, S.; Goldman, S.; Dietschreit, J. C. B.; Gómez-Bombarelli, R. Single-model uncertainty quantification in neural network potentials does not consistently outperform model ensembles. 2023, arXiv:2305.01754 https://arxiv.org/abs/2305.0175web4.
  85. 85
    Niblett, S. P.; Galib, M.; Limmer, D. T. Learning intermolecular forces at liquid–vapor interfaces. J. Chem. Phys. 2021, 155, 164101,  DOI: 10.1063/5.0067565
  86. 86
    Montes-Campos, H.; Carrete, J.; Bichelmaier, S.; Varela, L. M.; Madsen, G. K. H. A Differentiable Neural-Network Force Field for Ionic Liquids. J. Chem. Inf. Model. 2022, 62, 88101, PMID: 34941253  DOI: 10.1021/acs.jcim.1c01380

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 14 publications.

  1. Glenn R. Pastel, Travis P. Pollard, Oleg Borodin, Marshall A. Schroeder. From Ab Initio to Instrumentation: A Field Guide to Characterizing Multivalent Liquid Electrolytes. Chemical Reviews 2025, 125 (6) , 3059-3164. https://doi.org/10.1021/acs.chemrev.4c00380
  2. Leon de Villiers Engelbrecht, Narcis Cibotariu, Xiaoyan Ji, Aatto Laaksonen, Francesca Mocci. Deep Eutectic Solvents Meet Non-Aqueous Cosolvents: A Modeling and Simulation Perspective─A Tutorial Review. Journal of Chemical & Engineering Data 2025, 70 (1) , 19-43. https://doi.org/10.1021/acs.jced.4c00505
  3. Srijan Chatterjee, Tubai Chowdhury, Sayan Bagchi. Solvation Dynamics and Microheterogeneity in Deep Eutectic Solvents. The Journal of Physical Chemistry B 2024, 128 (51) , 12669-12684. https://doi.org/10.1021/acs.jpcb.4c06295
  4. Yezhi Jin, Gustavo R. Perez-Lemus, Pablo F. Zubieta Rico, Juan J. de Pablo. Improving Machine Learned Force Fields for Complex Fluids through Enhanced Sampling: A Liquid Crystal Case Study. The Journal of Physical Chemistry A 2024, 128 (34) , 7257-7268. https://doi.org/10.1021/acs.jpca.4c01546
  5. Da Zhu, Li Sheng, Taiping Hu, Sian Chen, Mengchao Shi, Haiming Hua, Kai Yang, Jianlong Wang, Yaping Tang, Xiangming He, Hong Xu. Investigation of the Degradation of LiPF6– in Polar Solvents through Deep Potential Molecular Dynamics. The Journal of Physical Chemistry Letters 2024, 15 (15) , 4024-4030. https://doi.org/10.1021/acs.jpclett.4c00575
  6. Xinyue Pi, Junfeng Lu, Simin Li, Jinliang Zhang, Yanlei Wang, Hongyan He. Computer-aided ionic liquid design for green chemical processes based on molecular simulation and artificial intelligence. Separation and Purification Technology 2025, 361 , 131585. https://doi.org/10.1016/j.seppur.2025.131585
  7. Marcelle B.M. Spera, Samir Darouich, Jürgen Pleiss, Niels Hansen. Influence of water content on thermophysical properties of aqueous glyceline solutions predicted by molecular dynamics simulations. Fluid Phase Equilibria 2025, 592 , 114324. https://doi.org/10.1016/j.fluid.2024.114324
  8. Monika Sobiech. Computer-Assisted Strategies as a Tool for Designing Green Monomer-Based Molecularly Imprinted Materials. International Journal of Molecular Sciences 2024, 25 (23) , 12912. https://doi.org/10.3390/ijms252312912
  9. Chen Fan, Linru Cheng, Wanlin Deng. Design of deep eutectic solvents for multiple perfluoroalkyl substances removal: Energy-based screening and mechanism elucidation. Science of The Total Environment 2024, 949 , 175039. https://doi.org/10.1016/j.scitotenv.2024.175039
  10. Fabian Zills, Moritz René Schäfer, Samuel Tovey, Johannes Kästner, Christian Holm. Machine learning-driven investigation of the structure and dynamics of the BMIM-BF 4 room temperature ionic liquid. Faraday Discussions 2024, 253 , 129-145. https://doi.org/10.1039/D4FD00025K
  11. Omid Shayestehpour, Stefan Zahn. Structure and transport properties of LiTFSI-based deep eutectic electrolytes from machine-learned interatomic potential simulations. The Journal of Chemical Physics 2024, 161 (13) https://doi.org/10.1063/5.0232631
  12. Muzammil Kuddushi, Ben Bin Xu, Naved Malek, Xuehua Zhang. Review of ionic liquid and ionogel-based biomaterials for advanced drug delivery. Advances in Colloid and Interface Science 2024, 331 , 103244. https://doi.org/10.1016/j.cis.2024.103244
  13. Leonid Kahle, Benoit Minisini, Tai Bui, Jeremy T. First, Corneliu Buda, Thomas Goldman, Erich Wimmer. A dual-cutoff machine-learned potential for condensed organic systems obtained via uncertainty-guided active learning. Physical Chemistry Chemical Physics 2024, 26 (34) , 22665-22680. https://doi.org/10.1039/D4CP01980F
  14. J. Betzabe González-Campos, Alejandra Pérez-Nava, Mario Valle-Sánchez, Luis Humberto Delgado-Rangel. Deep eutectic solvents applications aligned to 2030 United Nations Agenda for Sustainable Development. Chemical Engineering and Processing - Process Intensification 2024, 199 , 109751. https://doi.org/10.1016/j.cep.2024.109751

Journal of Chemical Theory and Computation

Cite this: J. Chem. Theory Comput. 2023, 19, 23, 8732–8742
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jctc.3c00944
Published November 16, 2023

Copyright © 2023 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

2037

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Schematic representation of choline chloride (left) and urea (right) with the corresponding atom labels used throughout this study.

    Figure 2

    Figure 2. Reference DFT energies and atomic forces are compared to the predicted values by the ML model for 50 test structures of a system of 304 atoms (left) and 684 atoms (right). The reference (ref) and predicted (pred) energies are standardized by subtracting their mean value from them.

    Figure 3

    Figure 3. Comparison of the redial distribution functions from the FPMD simulation (dashed lines) and simulations using the MLIP (solid lines) for interactions with H4 and N1 of choline (top) and hydrogen atoms of urea HU (bottom). Atom labels are given in Figure 1.

    Figure 4

    Figure 4. Center-of-mass velocity autocorrelation functions A(t) for the three components of the reline mixture from the reference FPMD simulations (dashed lines) are compared to the MLIP simulations (solid lines) for a mixture of 18 ChCl and 36 urea molecules at 375 K.

    Figure 5

    Figure 5. Calculated values of self-diffusion coefficients for the three components of reline mixture from MD simulations (filled symbols) are plotted against the inverse of the simulation temperature and compared to the experimental values (empty symbols) of ref (73). Lines are the best exponential fits to the data used to obtain the activation energies listed in Table 2.

    Figure 6

    Figure 6. Velocity cross-correlation functions NC(t) for the cation as the central particle (top), anion as the central particle (middle), and urea as the central particle (bottom). The dashed lines are the velocity autocorrelation functions A(t) of the cation (gray), anion (green), and urea (red). The cross-correlation functions for longer correlation times are presented in Figures S8–S10.

    Figure 7

    Figure 7. Calculated values of ionic conductivity from MD simulations using our ML potential are plotted against the inverse of the simulation temperature and compared to the experimental values from refs (1), (68), and (83). Lines are the VFT fits to the data.

  • References


    This article references 86 other publications.

    1. 1
      Abbott, A. P.; Capper, G.; Davies, D. L.; Rasheed, R. K.; Tambyrajah, V. Novel solvent properties of choline chloride/urea mixturesElectronic supplementary information (ESI) available: spectroscopic data. See http://www.rsc.org/suppdata/cc/b2/b210714g/. Chem. Commun. 2002, 7071,  DOI: 10.1039/b210714g
    2. 2
      Zhang, Q.; De Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep eutectic solvents: syntheses, properties and applications. Chem. Soc. Rev. 2012, 41, 71087146,  DOI: 10.1039/c2cs35178a
    3. 3
      Smith, E. L.; Abbott, A. P.; Ryder, K. S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 1106011082, PMID: 25300631  DOI: 10.1021/cr300162p
    4. 4
      Hansen, B. B.; Spittle, S.; Chen, B.; Poe, D.; Zhang, Y.; Klein, J. M.; Horton, A.; Adhikari, L.; Zelovich, T.; Doherty, B. W.; Gurkan, B.; Maginn, E. J.; Ragauskas, A.; Dadmun, M.; Zawodzinski, T. A.; Baker, G. A.; Tuckerman, M. E.; Savinell, R. F.; Sangoro, J. R. Deep Eutectic Solvents: A Review of Fundamentals and Applications. Chem. Rev. 2021, 121, 12321285, PMID: 33315380  DOI: 10.1021/acs.chemrev.0c00385
    5. 5
      Abbott, A. P.; Capper, G.; Davies, D. L.; McKenzie, K. J.; Obi, S. U. Solubility of Metal Oxides in Deep Eutectic Solvents Based on Choline Chloride. J. Chem. Eng. Data 2006, 51, 12801282,  DOI: 10.1021/je060038c
    6. 6
      Jenkin, G. R.; Al-Bassam, A. Z.; Harris, R. C.; Abbott, A. P.; Smith, D. J.; Holwell, D. A.; Chapman, R. J.; Stanley, C. J. The application of deep eutectic solvent ionic liquids for environmentally-friendly dissolution and recovery of precious metals. Miner. Eng. 2016, 87, 1824, Processing of Precious Metal Ores  DOI: 10.1016/j.mineng.2015.09.026
    7. 7
      Söldner, A.; Zach, J.; König, B. Deep eutectic solvents as extraction media for metal salts and oxides exemplarily shown for phosphates from incinerated sewage sludge ash. Green Chem. 2019, 21, 321328,  DOI: 10.1039/C8GC02702A
    8. 8
      Chakrabarti, M. H.; Mjalli, F. S.; AlNashef, I. M.; Hashim, M. A.; Hussain, M. A.; Bahadori, L.; Low, C. T. J. Prospects of applying ionic liquids and deep eutectic solvents for renewable energy storage by means of redox flow batteries. Renew. Sustain. Energy Rev. 2014, 30, 254270,  DOI: 10.1016/j.rser.2013.10.004
    9. 9
      Cong, G.; Lu, Y.-C. Organic Eutectic Electrolytes for Future Flow Batteries. Chem 2018, 4, 27322734,  DOI: 10.1016/j.chempr.2018.11.018
    10. 10
      Radošević, K.; Cvjetko Bubalo, M.; Gaurina Srček, V.; Grgas, D.; Landeka Dragičević, T.; Radojčić Redovniković, I. Evaluation of toxicity and biodegradability of choline chloride based deep eutectic solvents. Ecotoxicol. Environ. Saf. 2015, 112, 4653,  DOI: 10.1016/j.ecoenv.2014.09.034
    11. 11
      Halder, A. K.; Cordeiro, M. N. D. S. Probing the Environmental Toxicity of Deep Eutectic Solvents and Their Components: An In Silico Modeling Approach. ACS Sustain. Chem. Eng. 2019, 7, 1064910660,  DOI: 10.1021/acssuschemeng.9b01306
    12. 12
      Khandelwal, S.; Tailor, Y. K.; Kumar, M. Deep eutectic solvents (DESs) as eco-friendly and sustainable solvent/catalyst systems in organic transformations. J. Mol. Liq. 2016, 215, 345386,  DOI: 10.1016/j.molliq.2015.12.015
    13. 13
      Zahn, S. Deep eutectic solvents: similia similibus solvuntur?. Phys. Chem. Chem. Phys. 2017, 19, 40414047,  DOI: 10.1039/C6CP08017K
    14. 14
      Stefanovic, R.; Ludwig, M.; Webber, G. B.; Atkin, R.; Page, A. J. Nanostructure, hydrogen bonding and rheology in choline chloride deep eutectic solvents as a function of the hydrogen bond donor. Phys. Chem. Chem. Phys. 2017, 19, 32973306,  DOI: 10.1039/C6CP07932F
    15. 15
      Fetisov, E. O.; Harwood, D. B.; Kuo, I.-F. W.; Warrag, S. E. E.; Kroon, M. C.; Peters, C. J.; Siepmann, J. I. First-Principles Molecular Dynamics Study of a Deep Eutectic Solvent: Choline Chloride/Urea and Its Mixture with Water. J. Phys. Chem. B 2018, 122, 12451254, PMID: 29200290  DOI: 10.1021/acs.jpcb.7b10422
    16. 16
      Perkins, S. L.; Painter, P.; Colina, C. M. Molecular Dynamic Simulations and Vibrational Analysis of an Ionic Liquid Analogue. J. Phys. Chem. B 2013, 117, 1025010260, PMID: 23915257  DOI: 10.1021/jp404619x
    17. 17
      Doherty, B.; Acevedo, O. OPLS Force Field for Choline Chloride-Based Deep Eutectic Solvents. J. Phys. Chem. B 2018, 122, 99829993, PMID: 30125108  DOI: 10.1021/acs.jpcb.8b06647
    18. 18
      Nandy, A.; Smiatek, J. Mixtures of LiTFSI and urea: ideal thermodynamic behavior as key to the formation of deep eutectic solvents?. Phys. Chem. Chem. Phys. 2019, 21, 1227912287,  DOI: 10.1039/C9CP01440C
    19. 19
      Zhang, Y.; Poe, D.; Heroux, L.; Squire, H.; Doherty, B. W.; Long, Z.; Dadmun, M.; Gurkan, B.; Tuckerman, M. E.; Maginn, E. J. Liquid Structure and Transport Properties of the Deep Eutectic Solvent Ethaline. J. Phys. Chem. B 2020, 124, 52515264, PMID: 32464060  DOI: 10.1021/acs.jpcb.0c04058
    20. 20
      Shayestehpour, O.; Zahn, S. Molecular Features of Reline and Homologous Deep Eutectic Solvents Contributing to Nonideal Mixing Behavior. J. Phys. Chem. B 2020, 124, 75867597, PMID: 32790398  DOI: 10.1021/acs.jpcb.0c03091
    21. 21
      Goloviznina, K.; Gong, Z.; Costa Gomes, M. F.; Pádua, A. A. H. Extension of the CL&Pol Polarizable Force Field to Electrolytes, Protic Ionic Liquids, and Deep Eutectic Solvents. J. Chem. Theory Comput. 2021, 17, 16061617, PMID: 33555860  DOI: 10.1021/acs.jctc.0c01002
    22. 22
      Jeong, K.-j.; McDaniel, J. G.; Yethiraj, A. Deep Eutectic Solvents: Molecular Simulations with a First-Principles Polarizable Force Field. J. Phys. Chem. B 2021, 125, 71777186, PMID: 34181852  DOI: 10.1021/acs.jpcb.1c01692
    23. 23
      Shayestehpour, O.; Zahn, S. Ion Correlation in Choline Chloride–Urea Deep Eutectic Solvent (Reline) from Polarizable Molecular Dynamics Simulations. J. Phys. Chem. B 2022, 126, 34393449, PMID: 35500254  DOI: 10.1021/acs.jpcb.1c10671
    24. 24
      Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133A1138,  DOI: 10.1103/PhysRev.140.A1133
    25. 25
      Goedecker, S. Linear scaling electronic structure methods. Rev. Mod. Phys. 1999, 71, 10851123,  DOI: 10.1103/RevModPhys.71.1085
    26. 26
      Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996, 118, 1122511236,  DOI: 10.1021/ja9621760
    27. 27
      Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 2010, 31, 671690,  DOI: 10.1002/jcc.21367
    28. 28
      Cadena, C.; Zhao, Q.; Snurr, R. Q.; Maginn, E. J. Molecular Modeling and Experimental Studies of the Thermodynamic and Transport Properties of Pyridinium-Based Ionic Liquids. J. Phys. Chem. B 2006, 110, 28212832, PMID: 16471891  DOI: 10.1021/jp056235k
    29. 29
      Rajput, N. N.; Murugesan, V.; Shin, Y.; Han, K. S.; Lau, K. C.; Chen, J.; Liu, J.; Curtiss, L. A.; Mueller, K. T.; Persson, K. A. Elucidating the Solvation Structure and Dynamics of Lithium Polysulfides Resulting from Competitive Salt and Solvent Interactions. Chem. Mater. 2017, 29, 33753379,  DOI: 10.1021/acs.chemmater.7b00068
    30. 30
      Yan, T.; Wang, Y.; Knox, C. On the Dynamics of Ionic Liquids: Comparisons between Electronically Polarizable and Nonpolarizable Models II. J. Phys. Chem. B 2010, 114, 68866904, PMID: 20443608  DOI: 10.1021/jp908914d
    31. 31
      Szabadi, A.; Elfgen, R.; Macchieraldo, R.; Kearns, F. L.; Lee Woodcock, H.; Kirchner, B.; Schröder, C. Comparison between ab initio and polarizable molecular dynamics simulations of 1-butyl-3-methylimidazolium tetrafluoroborate and chloride in water. J. Mol. Liq. 2021, 337, 116521,  DOI: 10.1016/j.molliq.2021.116521
    32. 32
      Glielmo, A.; Sollich, P.; De Vita, A. Accurate interatomic force fields via machine learning with covariant kernels. Phys. Rev. B 2017, 95, 214302,  DOI: 10.1103/PhysRevB.95.214302
    33. 33
      Chmiela, S.; Tkatchenko, A.; Sauceda, H. E.; Poltavsky, I.; Schütt, K. T.; Müller, K. R. Machine learning of accurate energy-conserving molecular force fields. Sci. Adv. 2017, 3, e1603015  DOI: 10.1126/sciadv.1603015
    34. 34
      Chmiela, S.; Sauceda, H. E.; Müller, K. R.; Tkatchenko, A. Towards exact molecular dynamics simulations with machine-learned force fields. Nat. Commun. 2018, 9, 3887,  DOI: 10.1038/s41467-018-06169-2
    35. 35
      Bartók, A. P.; Payne, M. C.; Kondor, R.; Csányi, G. Gaussian Approximation Potentials: The Accuracy of Quantum Mechanics, without the Electrons. Phys. Rev. Lett. 2010, 104, 136403,  DOI: 10.1103/PhysRevLett.104.136403
    36. 36
      Behler, J.; Parrinello, M. Generalized Neural-Network Representation of High-Dimensional Potential-Energy Surfaces. Phys. Rev. Lett. 2007, 98, 146401,  DOI: 10.1103/PhysRevLett.98.146401
    37. 37
      Behler, J. Representing potential energy surfaces by high-dimensional neural network potentials. J. Phys.: Condens. Matter 2014, 26, 183001,  DOI: 10.1088/0953-8984/26/18/183001
    38. 38
      Ghasemi, S. A.; Hofstetter, A.; Saha, S.; Goedecker, S. Interatomic potentials for ionic systems with density functional accuracy based on charge densities obtained by a neural network. Phys. Rev. B 2015, 92, 045131,  DOI: 10.1103/PhysRevB.92.045131
    39. 39
      Smith, J. S.; Isayev, O.; Roitberg, A. E. ANI-1: an extensible neural network potential with DFT accuracy at force field computational cost. Chem. Sci. 2017, 8, 31923203,  DOI: 10.1039/C6SC05720A
    40. 40
      Yao, K.; Herr, J. E.; Toth, D. W.; Mckintyre, R.; Parkhill, J. The TensorMol-0.1 model chemistry: a neural network augmented with long-range physics. Chem. Sci. 2018, 9, 22612269,  DOI: 10.1039/C7SC04934J
    41. 41
      Schütt, K. T.; Sauceda, H. E.; Kindermans, P.-J.; Tkatchenko, A.; Müller, K. R. SchNet – A deep learning architecture for molecules and materials. J. Chem. Phys. 2018, 148, 241722,  DOI: 10.1063/1.5019779
    42. 42
      Han, J.; Zhang, L.; Car, R.; E, W. Deep Potential: A General Representation of a Many-Body Potential Energy Surface. Commun. Comput. Phys. 2018, 23, 629639,  DOI: 10.4208/cicp.oa-2017-0213
    43. 43
      Unke, O. T.; Meuwly, M. PhysNet: A Neural Network for Predicting Energies, Forces, Dipole Moments, and Partial Charges. J. Chem. Theory Comput. 2019, 15, 36783693, PMID: 31042390  DOI: 10.1021/acs.jctc.9b00181
    44. 44
      Behler, J. Atom-centered symmetry functions for constructing high-dimensional neural network potentials. J. Chem. Phys. 2011, 134, 074106,  DOI: 10.1063/1.3553717
    45. 45
      Thompson, A.; Swiler, L.; Trott, C.; Foiles, S.; Tucker, G. Spectral neighbor analysis method for automated generation of quantum-accurate interatomic potentials. J. Comput. Phys. 2015, 285, 316330,  DOI: 10.1016/j.jcp.2014.12.018
    46. 46
      Drautz, R. Atomic cluster expansion for accurate and transferable interatomic potentials. Phys. Rev. B 2019, 99, 014104,  DOI: 10.1103/PhysRevB.99.014104
    47. 47
      Batzner, S.; Musaelian, A.; Sun, L.; Geiger, M.; Mailoa, J. P.; Kornbluth, M.; Molinari, N.; Smidt, T. E.; Kozinsky, B. E(3)-equivariant graph neural networks for data-efficient and accurate interatomic potentials. Nat. Commun. 2022, 13, 2453,  DOI: 10.1038/s41467-022-29939-5
    48. 48
      Schütt, K., Unke, O., Gastegger, M. Equivariant message passing for the prediction of tensorial properties and molecular spectra. International Conference on Machine Learning 2021, 93779388.
    49. 49
      Haghighatlari, M.; Li, J.; Guan, X.; Zhang, O.; Das, A.; Stein, C. J.; Heidar-Zadeh, F.; Liu, M.; Head-Gordon, M.; Bertels, L.; Hao, H.; Leven, I.; Head-Gordon, T. NewtonNet: a Newtonian message passing network for deep learning of interatomic potentials and forces. Digital Discovery 2022, 1, 333343,  DOI: 10.1039/D2DD00008C
    50. 50
      Musaelian, A.; Batzner, S.; Johansson, A.; Sun, L.; Owen, C. J.; Kornbluth, M.; Kozinsky, B. Learning local equivariant representations for large-scale atomistic dynamics. Nat. Commun. 2023, 14, 579,  DOI: 10.1038/s41467-023-36329-y
    51. 51
      Zhang, L.; Han, J.; Wang, H.; Saidi, W. A.; Car, R.; Weinan, E.: Red Hook, NY, USA, 2018, pp 44414451. End-to-End Symmetry Preserving Inter-Atomic Potential Energy Model for Finite and Extended Systems Proceedings of the 32nd International Conference on Neural Information Processing Systems
    52. 52
      Wang, H.; Zhang, L.; Han, J.; E, W. DeePMD-kit: A deep learning package for many-body potential energy representation and molecular dynamics. Comput. Phys. Commun. 2018, 228, 178184,  DOI: 10.1016/j.cpc.2018.03.016
    53. 53
      Kingma, D. P.; Ba, J. Adam: A Method for Stochastic Optimization. 2014, arXiv:1412.6980. https://arxiv.org/abs/1412.698web0.
    54. 54
      Frauenheim, T.; Seifert, G.; Elsterner, M.; Hajnal, Z.; Jungnickel, G.; Porezag, D.; Suhai, S.; Scholz, R. A Self-Consistent Charge Density-Functional Based Tight-Binding Method for Predictive Materials Simulations in Physics, Chemistry and Biology. Phys. Status Solidi B 2000, 217, 4162,  DOI: 10.1002/(SICI)1521-3951(200001)217:1<41::AID-PSSB41>3.0.CO;2-V
    55. 55
      VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. Quickstep: Fast and accurate density functional calculations using a mixed Gaussian and plane waves approach. Comput. Phys. Commun. 2005, 167, 103128,  DOI: 10.1016/j.cpc.2004.12.014
    56. 56
      Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. cp2k: atomistic simulations of condensed matter systems. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2014, 4, 1525,  DOI: 10.1002/wcms.1159
    57. 57
      Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 30983100,  DOI: 10.1103/PhysRevA.38.3098
    58. 58
      Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785789,  DOI: 10.1103/PhysRevB.37.785
    59. 59
      Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104,  DOI: 10.1063/1.3382344
    60. 60
      Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 17031710,  DOI: 10.1103/PhysRevB.54.1703
    61. 61
      Hartwigsen, C.; Goedecker, S.; Hutter, J. Relativistic separable dual-space Gaussian pseudopotentials from H to Rn. Phys. Rev. B 1998, 58, 36413662,  DOI: 10.1103/PhysRevB.58.3641
    62. 62
      Smith, J. S.; Nebgen, B.; Lubbers, N.; Isayev, O.; Roitberg, A. E. Less is more: Sampling chemical space with active learning. J. Chem. Phys. 2018, 148, 241733,  DOI: 10.1063/1.5023802
    63. 63
      Podryabinkin, E. V.; Shapeev, A. V. Active learning of linearly parametrized interatomic potentials. Comput. Mater. Sci. 2017, 140, 171180,  DOI: 10.1016/j.commatsci.2017.08.031
    64. 64
      Peterson, A. A.; Christensen, R.; Khorshidi, A. Addressing uncertainty in atomistic machine learning. Phys. Chem. Chem. Phys. 2017, 19, 1097810985,  DOI: 10.1039/C7CP00375G
    65. 65
      Zhang, Y.; Wang, H.; Chen, W.; Zeng, J.; Zhang, L.; Wang, H. .. DP-GEN: A concurrent learning platform for the generation of reliable deep learning based potential energy models. Comput. Phys. Commun. 2020, 253, 107206,  DOI: 10.1016/j.cpc.2020.107206
    66. 66
      Thompson, A. P.; Aktulga, H. M.; Berger, R.; Bolintineanu, D. S.; Brown, W. M.; Crozier, P. S.; in ’t Veld, P. J.; Kohlmeyer, A.; Moore, S. G.; Nguyen, T. D.; Shan, R.; Stevens, M. J.; Tranchida, J.; Trott, C.; Plimpton, S. J. LAMMPS - a flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comput. Phys. Commun. 2022, 271, 108171,  DOI: 10.1016/j.cpc.2021.108171
    67. 67
      Martyna, G. J.; Tobias, D. J.; Klein, M. L. Constant pressure molecular dynamics algorithms. J. Chem. Phys. 1994, 101, 41774189,  DOI: 10.1063/1.467468
    68. 68
      Agieienko, V.; Buchner, R. Densities, Viscosities, and Electrical Conductivities of Pure Anhydrous Reline and Its Mixtures with Water in the Temperature Range (293.15 to 338.15) K. J. Chem. Eng. Data 2019, 64, 47634774,  DOI: 10.1021/acs.jced.9b00145
    69. 69
      Fu, X.; Wu, Z.; Wang, W.; Xie, T.; Keten, S.; Gomez-Bombarelli, R.; Jaakkola, T. Forces are not Enough: Benchmark and Critical Evaluation for Machine Learning Force Fields with Molecular Simulations. 2022, arXiv:2210.07237. https://arxiv.org/abs/2210.0723web7.
    70. 70
      Stocker, S.; Gasteiger, J.; Becker, F.; Günnemann, S.; Margraf, J. T. How robust are modern graph neural network potentials in long and hot molecular dynamics simulations?. Machine Learning: Science and Technology 2022, 3, 045010,  DOI: 10.1088/2632-2153/ac9955
    71. 71
      McDaniel, J. G.; Choi, E.; Son, C. Y.; Schmidt, J. R.; Yethiraj, A. Ab Initio Force Fields for Imidazolium-Based Ionic Liquids. J. Phys. Chem. B 2016, 120, 70247036, PMID: 27352240  DOI: 10.1021/acs.jpcb.6b05328
    72. 72
      Yeh, I.-C.; Hummer, G. System-Size Dependence of Diffusion Coefficients and Viscosities from Molecular Dynamics Simulations with Periodic Boundary Conditions. J. Phys. Chem. B 2004, 108, 1587315879,  DOI: 10.1021/jp0477147
    73. 73
      D’Agostino, C.; Harris, R. C.; Abbott, A. P.; Gladden, L. F.; Mantle, M. D. Molecular motion and ion diffusion in choline chloride based deep eutectic solvents studied by 1H pulsed field gradient NMR spectroscopy. Phys. Chem. Chem. Phys. 2011, 13, 2138321391,  DOI: 10.1039/c1cp22554e
    74. 74
      Balucani, U.; Vallauri, R.; Murthy, C. S. Interparticle velocity correlations in simple liquids. J. Chem. Phys. 1982, 77, 32333237,  DOI: 10.1063/1.444199
    75. 75
      Verdaguer, A.; Padró, J. A.; Trullàs, J. Molecular dynamics study of the velocity cross-correlations in liquids. J. Chem. Phys. 1998, 109, 228234,  DOI: 10.1063/1.476555
    76. 76
      Balucani, U.; Vallauri, R.; Murthy, C. Momentum transfer analysis in Lennard-Jones fluids. Phys. Lett. A 1981, 84, 133136,  DOI: 10.1016/0375-9601(81)90736-2
    77. 77
      Endo, Y.; Endo, H. Microscopic motions and the local environments of atoms in simple liquids. J. Chem. Phys. 1984, 80, 20872091,  DOI: 10.1063/1.446974
    78. 78
      Flener, P.; Vesely, F. J. Correlated motion of two particles in a fluid. Mol. Phys. 1992, 77, 601615,  DOI: 10.1080/00268979200102651
    79. 79
      Balucani, U.; Zoppi, M. Dynamics of the Liquid State; Clarendon Press, 1995; Vol. 10.
    80. 80
      Verdaguer, A.; Padró, J. A. Computer simulation study of the velocity cross correlations between neighboring atoms in simple liquid binary mixtures. J. Chem. Phys. 2001, 114, 27382744,  DOI: 10.1063/1.1340581
    81. 81
      Hansen, J.-P.; McDonald, I. R. Theory of Simple Liquids, 3rd ed.; Hansen, J.-P., McDonald, I. R., Eds.; Academic Press: Burlington, 2006; pp 291340.
    82. 82
      Rey-Castro, C.; Vega, L. F. Transport Properties of the Ionic Liquid 1-Ethyl-3-Methylimidazolium Chloride from Equilibrium Molecular Dynamics Simulation. The Effect of Temperature. J. Phys. Chem. B 2006, 110, 1442614435, PMID: 16854152  DOI: 10.1021/jp062885s
    83. 83
      Reuter, D.; Binder, C.; Lunkenheimer, P.; Loidl, A. Ionic conductivity of deep eutectic solvents: the role of orientational dynamics and glassy freezing. Phys. Chem. Chem. Phys. 2019, 21, 68016809,  DOI: 10.1039/C9CP00742C
    84. 84
      Tan, A. R.; Urata, S.; Goldman, S.; Dietschreit, J. C. B.; Gómez-Bombarelli, R. Single-model uncertainty quantification in neural network potentials does not consistently outperform model ensembles. 2023, arXiv:2305.01754 https://arxiv.org/abs/2305.0175web4.
    85. 85
      Niblett, S. P.; Galib, M.; Limmer, D. T. Learning intermolecular forces at liquid–vapor interfaces. J. Chem. Phys. 2021, 155, 164101,  DOI: 10.1063/5.0067565
    86. 86
      Montes-Campos, H.; Carrete, J.; Bichelmaier, S.; Varela, L. M.; Madsen, G. K. H. A Differentiable Neural-Network Force Field for Ionic Liquids. J. Chem. Inf. Model. 2022, 62, 88101, PMID: 34941253  DOI: 10.1021/acs.jcim.1c01380
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.3c00944.

    • Details of the DFTB-based MD simulations, additional radial and angular distribution functions, simulation box sizes and viscosities, mean-square displacement plots and self-diffusion coefficients from the Green–Kubo approach with details of the correction term for the finite-size effects, and velocity cross-correlation functions for longer correlation times (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.