The formally exact framework of equilibrium Density Functional Theory (DFT) is capable of simultaneously and consistently describing thermodynamic and structural properties of interacting many-body systems in arbitrary external potentials. In practice, however, DFT hinges on approximate (free-)energy functionals from which density profiles (and hence the thermodynamic potential) follow via an Euler–Lagrange equation. Here, we explore a relatively simple Machine-Learning (ML) approach to improve the standard mean-field approximation of the excess Helmholtz free-energy functional of a 3D Lennard-Jones system at a supercritical temperature. The learning set consists of density profiles from grand-canonical Monte Carlo simulations of this system at varying chemical potentials and external potentials in a planar geometry only. Using the DFT formalism, we nevertheless can extract not only very accurate 3D bulk equations of state but also radial distribution functions using the Percus test-particle method. Unfortunately, our ML approach did not provide very reliable Ornstein–Zernike direct correlation functions for small distances.

Given the massive present-day availability of computer power and data, the grown general interest in machine learning (ML) should not come as a big surprise. This interest also extends to physics whose community excels at gathering, organizing, and analyzing data in order to predict and model the behavior of systems with many degrees of freedom, which is also one of the strengths of ML. An important distinction between the field of physics and ML is that physicists tend to understand, model, and predict the systems of their interest via a stepwise chain of reasoning from cause to effect, whereas ML algorithms tend to “only” directly relate cause to effect without necessarily understanding (in the traditional “human” sense) the steps in between. In other words, ML can often be regarded as a black box that is as incomprehensible as the initial raw data itself.

Here, we will also suffer, at least to some extent, from this black-box character of ML applied to a problem in classical Density Functional Theory (DFT).1,2 However, only in a limited way because we can build on the foundations of physics to exploit, in this case, a few ingredients of the DFT formalism. As we will explain in full detail below, DFT is an exact framework to describe thermodynamic and structural properties of interacting many-body systems. This involves the solution of Euler–Lagrange equations for the equilibrium density profile for a particle system in an external potential. Now, DFT hinges for given particle–particle interactions on approximate free-energy density functionals. By comparison with Monte Carlo (MC) simulations of density profiles in a learning set of external potentials, a free-energy functional can be constructed during an ML process. The additional physics that can be extracted beyond the learning set not only includes density profiles for external potentials outside the learning set but also (i) thermodynamic bulk quantities (because the minimal value of the functional corresponds to the thermodynamic potential at equilibrium from which, for instance, the bulk pressure follows) and (ii) the two-body direct correlation function (because it is related to the second functional derivative of the functional) from which the radial distribution function follows. Moreover, thermodynamic surface properties such as the adsorption and the interfacial tension can be extracted from the functional. Our work is strongly inspired by recent ML work to construct a classical DFT for the Lennard-Jones (LJ) fluid in one spatial dimension,3,4 which we here extend to the three-dimensional LJ fluid. Similar to Refs. 3 and 4, we use grand-canonical Monte Carlo (MC) simulations at a learning set of chemical potentials and external potentials, however, in a planar geometry. We stress that the planar geometry yields an effective 1D problem embedded in 3D, not to be confused with an actual 1D problem. We will show the ability not only to “learn” a free-energy functional that predicts density profiles of this system at chemical and external potentials outside the training set but also to extract several system properties that were not at all present in the data of the training set or at least not explicitly. In particular, we will show that from a learning set in a planar geometry, a machine-learned functional can be constructed that is capable of predicting the 3D mechanical bulk equation of state of the homogeneous fluid (the pressure–density–chemical potential relations), the 3D radially symmetric direct correlation function and the radial distribution function at any density, and (in principle) Lennard-Jones density profiles in an arbitrary external potential in 3D. The agreement of these predictions against simulations varies from very good (the equation of state, radial distributions from the Percus test-particle method, and density profiles outside the learning set) to, admittedly, rather poor (direct correlation function). The poor prediction for the latter is probably due to the rather simple form (and in retrospect perhaps an overly simple form compared to Ref. 4) for the free-energy functional and due to the treatment of the repulsive part of the LJ interaction. The main thrust of our findings at this point, therefore, is not the construction of the Lennard-Jones free-energy functional that compares “best” with MC simulations, but rather the notion that free-energy functionals for 3D systems can be constructed from relatively simple geometries (here planar) in the learning set. Extensions to other systems, for instance electrolytes and ionic liquids forming an electric double layer in contact with planar electrodes, could be a next step with actual applications in modeling the osmotic equation of state, the differential capacitance, and the adsorption in porous geometries.

This paper is organized as follows. We start in Sec. II with an extensive introduction for classical DFT—which can easily be skipped by readers familiar with this framework. In Secs. III and IV, we discuss the system and the simulation and machine-learning methods that we use, and in Sec. V, we discuss the resulting kernels, density profiles, equations of state, and pair correlation functions. We end in Sec. VI with a discussion and outlook.

We consider a classical one-component system of N spherical particles with linear momenta pi and center-of-mass positions ri, with i = 1, …, N being the particle label. The particles interact with each other via an isotropic pair potential ϕ(rij), where rij = |rirj| is the distance between particles i and j. All particles are subject to a static external potential Vext(ri) such that the Hamiltonian of the system reads

(1)

where m denotes the mass of the particles. Here, we note that Eq. (1) can also describe macroscopic bulk systems by considering the external potential to be zero in a box of volume V at a homogeneous density ρ = N/V and temperature T. For these homogeneous systems, typical thermodynamic equilibrium quantities of interest include the caloric and mechanical equations of state u(ρ, T) and p(ρ, T) for the internal energy and pressure, respectively. In addition, structural quantities such as the radial distribution function g(r) (at particle–particle separation r) and the structure factor S(q) (at wavenumber q) are of interest for homogeneous systems.2 Equilibrium statistical mechanics offers a variety of techniques to calculate (approximations to) these quantities, for instance, systematic low-ρ or high-T expansions, integral equations based on the Ornstein–Zernike equation, or computer simulations. However, the situation is more complicated in the case of a nontrivial external potential due to, for instance, the Earth’s gravity, an attractive or repulsive substrate, or a porous matrix that may confine the particles of interest. In this case, the system described by Eq. (1) becomes heterogeneous in thermodynamic equilibrium such that the local density ρ(r) varies in space. Consequently, the energy density u and the pressure p become ill-defined (except, of course, within a local density approximation), and the broken translation invariance causes the radial distribution function to be of the form g(r, r′) rather than g(|rr′|). Nevertheless, the formalism of Density Functional Theory (DFT) can provide a consistent picture of the thermodynamic and structural properties of inhomogeneous fluids in an external potential. Although DFT finds its roots in the quantum-many-body description of electrons, it has also found many applications in the (essentially) classical context of soft-matter systems to describe molecular liquids, electrolytes, colloidal dispersions, etc.1,5–12

DFT is essentially a grand-canonical framework in which the temperature T and the chemical potential μ of the particles are fixed to characterize the heat bath and the particle bath with which the system is in thermal and diffusive equilibrium. The corresponding thermodynamic potential is the grand potential Ω0 defined by βΩ0=lnN=0dpNdrNexp[β(μNHN)]/N!h3N, where β−1 = kBT, kB is the Boltzmann constant, and h is an arbitrary constant with the same dimension as the Planck constant. From Ω0, essentially all thermodynamic properties would follow, for instance, the pressure of the homogeneous system equals −Ω0/V, and the internal energy is ∂βΩ0/∂β. Of course, this involves the immense problem of evaluating the 6N-dimensional phase-space integral in the definition of Ω0. The key of classical DFT is that it circumvents this high-dimensional phase-space integral by a proof1 of the existence of a grand potential functional Ω[ρ] of the variational one-body density profile ρ(r), with the properties that (i) the equilibrium density profile ρ0(r) minimizes the functional Ω[ρ], and (ii) this minimum equals the equilibrium grand potential Ω0. This implies that

(2)

The problem is thus reduced to finding the functional Ω[ρ] and, after that, to solving the 3D Euler–Lagrange equation (2), which amounts to a huge reduction in the problem compared to the high-dimensional phase-space integral.

One can also prove rigorously1,2,13 that the grand potential functional Ω[ρ] can always be written as

(3)

where F[ρ] is the intrinsic Helmholtz free-energy functional that, and this is crucial for our machine-learning approach, only and uniquely depends on the particle–particle interactions [here, the pair potential ϕ(r)] and on the temperature, and not on μ and Vext(r). In other words, the same and unique functional F[ρ] for a given ϕ(r) applies at any chemical and external potential. That F[ρ] is a Helmholtz free-energy functional follows straightforwardly from the thermodynamic relation Ω0 = F0μN0, with N0 = ∫drρ0(r) being the equilibrium number of particles and F0 being the equilibrium Helmholtz free energy, which can be decomposed into the sum of the potential energy drρ0(r)Vext(r) due to the external field and the remaining intrinsic free energy F[ρ0].

Unfortunately, F[ρ] is not explicitly known in most cases. An exception is the ideal-gas case of ϕ(r) ≡ 0, where it is possible to construct the intrinsic free-energy functional as Fid[ρ]=kBTdrρ(r)lnρ(r)Λ31, with Λ=h/2πmkBT being the thermal wavelength. The common practice in DFT is now to split the intrinsic free energy into the ideal and the excess-over-ideal part, F[ρ]=Fid[ρ]+Fexc[ρ], and to find an explicit (usually approximate) expression for Fexc[ρ]. Once such an explicit expression has been found, we can cast the minimum condition for ρ0(r) of Eq. (2) in the explicit form

(4)

Note that Eq. (4) is a self-consistency relation for interacting systems, which usually takes the form of a nonlinear integro-differential equation that needs to be solved numerically for given μ and Vext(r) for a system of interest with pair potential ϕ(r) at temperature T and, hence, with a given excess functional Fexc[ρ]. In relatively simple geometries, for instance, with planar or radial symmetry, a numerical solution of Eq. (4) can be found at relatively low computational cost by means of a Picard iteration scheme.

Thus, the remaining problem of DFT lies in constructing an explicit form for Fexc[ρ], for which no universal recipe is available —not unlike the case of partition functions of interacting systems. There is, however, one more exact relation that can be and has been exploited and involves the second functional derivative βδ2Fexc[ρ]/δρ(r)δρr, which equals by definition the Ornstein–Zernike direct correlation function c(r, r′) and is hence directly related to the two-body structure of the system. In particular, in a homogeneous bulk system, the direct correlation function is of the form cb(|rr′|), and its Fourier transform ĉb(q) yields the structure factor S(q)=(1ρĉb(q))1, from which the radial distribution function g(r) follows by an inverse Fourier transformation.

In this manuscript, we will focus on a Lennard-Jones fluid. In the DFT treatment, we split the pair potential ϕ(r) = ϕ0(r) + ϕ1(r) into a steep repulsion ϕ0(r) at short distances and an attractive tail ϕ1(r) of well depth −ɛ < 0 in accordance with Barker–Henderson theory, as further detailed in Sec. III. On the basis of the vast body of knowledge on the thermodynamics and the two-body structure of the hard-sphere system, extremely accurate approximations have been constructed for its intrinsic excess Helmholtz free-energy functional FHSexc[ρ], for which we will use the White-Bear mark II12 version of the fundamental measure theory6,11 throughout this paper. It is common practice in liquid-state theory to treat the attractions as a perturbation on the hard-sphere system, and a popular version results in the van der Waals-like mean-field (MF) approximation,

(5)

The high-temperature limit of Eq. (5) returns the hard-sphere limit, but the MF approximation fails to give accurate results for lower temperatures where the attractions play a more prominent role.2,3,14 We therefore seek improved excess free-energy functionals in terms of corrections to the mean-field functional of Eq. (5) of the quadratic and cubic forms,

(6)
(7)

where the labels ML2 and ML3 refer to the fact that we will use machine learning (ML) to find the optimal form of the kernels Ω2(r) and Ω3(r). We note that ML2 reduces to the mean-field form for Ω2(r) ≡ 0 and that ML3 reduces to ML2 for Ω3(r) ≡ 0. We also emphasize that the ML3 form of the functional is not compatible with the third-order virial-type expansion, which would have entailed an additional spatial integration (say over r′) and a kernel of the triple product form f(|rr|)f|rr|f|rr|; finding the optimal kernel f(r) proved to be computationally too demanding and inconvenient at the exploration phase of this project, and hence, we settled for the simpler form of the cubic term ML3.

The building of ML functionals upon the MF functional is important. For long-ranged potentials, the mean-field functional retrieves the correct asymptotic decay of the direct correlation function such that the range of the ML corrections can conveniently be limited. For short-ranged potentials, the direct correlations are short-ranged anyway, and the inclusion of the mean-field term puts no constraint on the resulting ML functional.

Interestingly, we can exploit the fact that the optimal kernels Ω2(r) and Ω3(r) that we seek must be independent of μ and Vext(r) to determine them in systems with a planar geometry, i.e., systems that have translation invariance in the x- and y-directions with external potentials Vext(z) and density profiles ρ(z) that only depend on the normal coordinate z. One easily checks that the non-HS part of the MF functional of Eq. (5) then reduces to

(8)

where A = dxdy is the (macroscopically large) area of the planar surface and ϕ1,z(|zz|)=dxdyϕ1((zz)2+x2+y2) is the laterally integrated pair potential ϕ1(r). Likewise, the non-HS contributions to the functionals of Eq. (6) can be cast in the form

(9)
(10)

where the laterally integrated kernels ω2 and ω3 can be written as

(11)

Interestingly, Eq. (11) can be inverted such that we find

(12)

In other words, once we find ω2(z) and ω3(z) from calculations in planar geometry, we can determine Ω2(r) and Ω3(r) from Eq. (12) such that the direct correlation function follows by taking second functional derivatives of Eqs. (6) and (7). Hence, we have access to thermodynamic and structural properties in bulk, in any geometric confinement, at any external potential, at any chemical potential, and solely on the basis of input in a planar geometry.

In this paper, we consider a 3D fluid in which the particles interact with a truncated and shifted Lennard-Jones (LJ) interaction given by

(13)

where ɛ > 0 denotes the well depth and σ is the LJ particle diameter. The full LJ potential is truncated at rcut = 4σ and shifted upward by εcut = 0.98 · 10−3ɛ such that ϕLJ(rcut) = 0. The splitting of ϕLJ(r) into a hard-sphere reference and an attractive tail in the DFT treatment is performed on the basis of the well-known Barker–Henderson theory15,16 that leads to an effective and temperature-dependent hard-core diameter 0 < dσ that does not depend on the bulk density, as explained in Refs. 15 and 17. At the temperature kBT/ɛ = 2 of our main interest, the effective diameter is given by d = 0.9568σ. The resulting expression for ϕ1(r) then reads

(14)

We stress that the value inside the core, i.e., ϕ1(r < d), is not uniquely defined,18,19 and its value can be used as a fit parameter for better agreement between simulations and DFT. However, we choose here to set it to 0 in line with previous studies on the LJ system.20–22 

The external potentials Vext(z) that we consider in this manuscript all mimic a planar slit geometry. The slit is translationally invariant in the lateral xy plane and is mirror-symmetric in the midplane z = 0 such that Vext(z) = Vext(−z). We employ a family of external wall-particle potentials that is repulsive and parameterized by

(15)

where the dimensionless strength s = βVext(L/2) ≥ 40 characterizes the potential at |z| = L/2, w ∈ [0, 1] denotes the width of the central part of the slit, βVext(z) = 0, and p > 0 is the power that characterizes the steepness of the potential. Figure 1 illustrates the external potential for general s, w, and L and for steepness parameters p = 2 (black) and p = 8 (blue).

FIG. 1.

A general visualization of the external potential described in Eq. (15). This external potential is applied in the training data and is given by the parameters w, s, p and L. Two values of p are considered, namely, p = 2 (black) and p = 8 (blue).

FIG. 1.

A general visualization of the external potential described in Eq. (15). This external potential is applied in the training data and is given by the parameters w, s, p and L. Two values of p are considered, namely, p = 2 (black) and p = 8 (blue).

Close modal

To generate the training and validation datasets to “learn” the density functional, we perform grand-canonical Monte Carlo (MC) simulations of the 3D truncated and shifted Lennard-Jones (LJ) fluid confined between two planar soft-repulsive walls described by the external potential Vext(z) of Eq. (15). Here, we only consider highly repulsive walls with s ≥ 40 to ensure that the density reduces to essentially 0 at |z| = L/2. We measure the equilibrium density profile ρMC(z) in a cubic simulation box of volume V = L3, with L = 10σ. We impose periodic boundary conditions in the x- and y-directions and equilibrate the system for at least 105 MC cycles before the measurements start. Each MC cycle consists of N trial moves, with N denoting the instantaneous number of particles. Each trial move can either displace a particle or insert/remove a particle. The probability of selecting a trial move to displace a particle instead of an insertion/deletion move is set to 90%. The sampling of the density profile ρMC(z) is performed by dividing the volume in 320 equidistant bins that represent planar slices normal to the z axis, each of width σ/32 such that the interval z ∈ [−5σ, 5σ] is fully covered. The density in each bin is measured and stored after every fourth MC cycle.

In order to avoid (interesting but at this stage undesired) complications due to possible phase transitions (condensation, pre-wetting, capillary evaporation, etc.), we consider only a supercritical temperature kBT/ɛ = 2. Eight different chemical potentials μ are imposed in the grand-canonical MC simulations of the LJ system, given by βμ ∈ {−3.0, −2.5, …, 0.0, 0.5}. Here, the arbitrary offset of μ is chosen such that the thermal wavelength equals the particle diameter, Λ = σ; it implies that βμ →  log ρbσ3 in the dilute (ideal-gas) limit ρbσ3 ≪ 1. A total of 24 different external potentials are considered as training sets, all with total slit length L = 10σ and strengths s ∈ {40, 60}, widths w ∈ {0.4, 0.65, 0.9}, and steepness parameters p ∈ {2, 4, 8, 10}.

As an illustration, we show in Fig. 2 the simulated density profiles ρMC(z) of a LJ fluid at kBT/ɛ = 2 and chemical potentials βμ = {−3.0, −2.5, −2, −1.5, −1.0, −0.5, 0, 0.5} (symbols) corresponding to (separately simulated) bulk densities ρbσ3 ≈ {0.056, 0.10, 0.19, 0.33, 0.47, 0.56, 0.62, 0.67} in the external potential Vext(z) characterized by s = 60, w = 0.4, and p = 4, as denoted by the red solid line. We observe monotonous density profiles at the lowest μ’s, the development of density oscillations at higher μ′s, and a fairly well-defined “bulk” density in the vicinity of z = 0 (except at the highest μ′s, where the profiles of the two walls show some overlap due to the limited system size).

FIG. 2.

Equilibrium density profiles ρ(z) (symbols) of a LJ fluid at temperature kBT/ɛ = 2 and chemical potentials βμ = −3.0, −2.5, −2, −1.5, −1.0, −0.5, 0, 0.5 from bottom to top, in an external potential Vext(z) (red solid line, right vertical axis) characterized by a strength s = 60, a width w = 0.4, and steepness parameter p = 4 as obtained from Monte Carlo simulations.

FIG. 2.

Equilibrium density profiles ρ(z) (symbols) of a LJ fluid at temperature kBT/ɛ = 2 and chemical potentials βμ = −3.0, −2.5, −2, −1.5, −1.0, −0.5, 0, 0.5 from bottom to top, in an external potential Vext(z) (red solid line, right vertical axis) characterized by a strength s = 60, a width w = 0.4, and steepness parameter p = 4 as obtained from Monte Carlo simulations.

Close modal

With an optimization process that uses several techniques from the field of ML, we will construct intrinsic free-energy functionals of the form of Eqs. (9) and (10) such that the density profiles ρML(z) that follow from this machine-learned functional are an “optimal” approximation to the corresponding MC densities ρMC(z). We recall that ρML(z) is to be determined as a solution of the Euler–Lagrange equation (4), not only for Λ = σ at a given temperature, chemical potential, and external potential but also for a given excess functional Fexc[ρ]. In other words, we are interested in optimal kernels ω2(z) and ω3(z) [where ω3(z) ≡ 0 for ML2].

In order to quantify “optimal,” we define the so-called loss function L that characterizes the difference between ML and MC profiles and that we will minimize with respect to ω2(z) and ω3(z). Here, we define L=L1+L2 to consist of a dominant contribution L1 and a regularization term L2. The dominant loss term is defined by the mean-square error23 between the MC and ML profiles,

(16)

where j = 1, …, n labels the n = 24 × 8 = 192 combinations of 24 external potentials and the eight chemical potentials of the training set as identified above. We normalize the difference between the MC and ML profiles by the MC bulk density at the chemical potential μj of training set j, for which we performed separate bulk simulations. This scaling promotes equal weights to high- and low-density states during the learning process. The regularization term L2 is independent of the MC and ML profiles and defined by

(17)

where f(z) is given by

(18)

It accounts for the constraint that ω2(z) and ω3(z) must decay smoothly to 0 for zσ, where our statistics is poor. Moreover, L2 also suppresses undue high-wavenumber undulations that tend to develop at |z| < σ. We tune the (positive) regularization parameter λ by trial and error such that it contributes less to the minimization procedure than L1, while not being too small to be irrelevant. Note that L2 effectively reduces the range of ωi(z) by suppressing it exponentially for |z| > σ.

The minimization of the total loss function L is performed with the stochastic and iterative optimization method Adam as proposed by Kingma and Ba in Ref. 24. We use their suggested default step size α = 0.001 and exponential decay rates β1 = 0.9 and β2 = 0.999 and refer the reader to their work for a full description of the method and its parameters. During each iteration of the minimization process, the gradient of the loss function with respect to the kernels ωi(z) is required, which are straightforwardly derived for L2 to be the functional derivatives,

(19)

The functional derivatives of L1 with respect to ωi(z) for i = 2, 3 are more involved and stem from the dependence of L1 on the ML density profiles ρjML(z) for j = 1, …, n such that the functional chain-rule yields

(20)

where we replaced the dummy integration variable z of Eq. (16) by z′. From the Euler–Lagrange equation (4) for the DFT equilibrium profiles ρ0—which are represented by ρjML in Eq. (20)—one checks that δρ0(z)/δωi(z)=ρ0(z)×δ2βFexc/δρ(z)δωi(z). Upon considering ωi and ρ independent variables in Eqs. (9) and (10) for the ML2 and ML3 excess functionals, respectively, the second (cross) derivative for i = 2 equals ρ0(z + z′), and for ML3 and i = 3, it equals 13[ρ02(z+z)+2ρ0(z)ρ0(z+z)]. Hence, within this approximation, a numerical integration of z′ suffices to evaluate Eq. (20), and in combination with Eq. (19), we can numerically calculate δL/δωi(z) for the grid points z of our system. Thus, we have all ingredients to minimize L by means of Adam.24 

The training process starts with an initial guess for the two kernels, for which we take the MF approximation ω20(z)=ω30(z)0, where the superscript 0 denotes the 0-th iteration in the training process. Next, we use these kernels to calculate the n density profiles ρj,kML(z) for learning sample j = 1, …, n and iteration label k = 0 by solving the Euler–Lagrange equation [Eq. (4)] using a Picard iteration scheme with the MC profile ρjMC(z) as the initial guess.

On the basis of Eqs. (16)--(20), we can then evaluate L and δL/δωi(z) for i = 1, 2, from which improved kernels ωik(z) are constructed for k = 1 by employing Adam,24 which will give rise to improved density profiles ρj,1ML(z), etc. For k ≥ 2, we take ρj,k1ML(z) as the initial guess in the Picard-iteration of ρj,kML. The iteration process is repeated until the loss function has converged.

Although Adam is already an efficient algorithm for the learning process, its computational cost can be significantly reduced by making use of stochastic optimization. Rather than using all n elements of the training set in every iteration, which involves the addition of all n terms in Eq. (20) at every iteration level k, we consider mini batches with only 20 randomly selected elements of the training set during each Picard iteration k. The gradient of the loss function L1 is computed by only taking into account this mini batch; thus, the summation over the n density profiles of Eq. (20) changes to a summation over 20 randomly selected density profiles, and the normalization factor 1/n becomes 1/20. A new mini batch is randomly selected during every iteration in the ML process.

We perform MC simulations of the LJ system to generate MC density profiles with 24 different external potentials, described in Sec. III, and eight equi-distant different chemical potentials, βμ ∈ {−3.0, −2.5, …, 0.0, 0.5}, for the temperature kBT/ɛ = 2. We describe the kernels, the resulting density profiles, the mechanical equations of state of the bulk fluid, and the radial distribution functions that follow from the functionals ML2 and ML3 using two different routes.

For several choices of the regularization parameter λ in Eq. (17), we determined the kernel ω2(z) for ML2 and ω2(z) and ω3(z) for ML3. Without a significant L2 contribution, λ ≤ 10−3, we found spurious peaks in both ω2(z) and ω3(z) for |z| > 8σ, i.e., at the largest separations (with the poorest statistics) we considered in the learning set; these spurious peaks disappeared, and ω2(z) smoothly decayed to 0 for λ ≥ 10−2, and throughout, we settle for λ = 10−2 as a reasonable compromise between error-correction and minimization of the actual loss function of interest L1.

In Fig. 3, we present the evolution of the loss functions L1 (blue) and L2 (red) during the training process with iteration label k in (a) for ML2 and in (b) for ML3; the gray curves in (a) and (b) represent a moving average of L1 over 15 iterations. We observe good convergence after, say, k = 5000 iterations, respectively. We note that the minimized loss function L1 of ML2 is as small as 5 × 10−4, and for ML3, it is even about four times smaller. We also note that L2<L1 for ML2, as desired for a regularization term that is (naively) supposed to be a small correction to the total loss function. However, for ML3, we find L1 to be so small that it has dropped below the regularization term L2, which in retrospect should be seen as a consequence of the good accuracy of the ML3 functional rather than as a problem for the relative magnitude of the two contributions to the loss function.

FIG. 3.

The loss function contributions L1 (blue) and L2 (red) as a function of the iteration label k in (a) for ML2 and in (b) for ML3. The gray traces in (a) and (b) represent a moving average of L1, and the insets show the zoomed-in view.

FIG. 3.

The loss function contributions L1 (blue) and L2 (red) as a function of the iteration label k in (a) for ML2 and in (b) for ML3. The gray traces in (a) and (b) represent a moving average of L1, and the insets show the zoomed-in view.

Close modal

In Fig. 4(a), we show the MF (scaled) kernel βϕ1,z(z) (blue dashed line) and its ML2 correction βϕ1,z(z) + ω2(z) (black solid line), as obtained after 5000 iterations. For all z, the ML2 kernel is more negative than the MF kernel, as if there is actually more cohesive energy in the system than predicted by MF. We see that ω2(z) develops a peculiar and unexpected small “bump” close to z = 0. For ML3, a similar feature occurs close to z = 0 in both ω2(z) and ω3(z), as can be seen in Fig. 4(b), where we plot ω2(z) (green solid line) and ω3(z) (green dotted line) for the ML3 case as obtained after 5000 iterations, together with the ML2 kernel ω2(z) (black solid line) for comparison. We see that ω2 from ML3 is again essentially negative (except for a tiny positive feature at z = 0 and |z| ≃ 2σ) and contains a “bump” similar to the ML2 case. We also see that ω3(z) has a structure that is quite similar to ω2(z), however more pronounced with higher peaks and lower valleys. Below, we will investigate the thermodynamic and structural properties that follow from DFT based on these kernels.

FIG. 4.

(a) The mean-field (MF) kernel βϕ1,z(z)/σ2 (blue dashed line) and its quadratic Machine-Learning (ML2) improvement (βϕ1,z(z) + ω2(z))/σ2 (black solid line), as obtained for the LJ system at temperature kBT/ɛ = 2. (b) The cubic Machine-Learning (ML3) kernels ω2(z)/σ2 (green solid line) and ω3(z)/σ5 (green dotted line), also at kBT/ɛ = 2, for comparison together with the ML2 kernel ω2(z)/σ2 (black solid line).

FIG. 4.

(a) The mean-field (MF) kernel βϕ1,z(z)/σ2 (blue dashed line) and its quadratic Machine-Learning (ML2) improvement (βϕ1,z(z) + ω2(z))/σ2 (black solid line), as obtained for the LJ system at temperature kBT/ɛ = 2. (b) The cubic Machine-Learning (ML3) kernels ω2(z)/σ2 (green solid line) and ω3(z)/σ5 (green dotted line), also at kBT/ɛ = 2, for comparison together with the ML2 kernel ω2(z)/σ2 (black solid line).

Close modal

The first test of the quality of the ML functionals is a comparison of their resulting density profiles with the simulated ones from the training set. In Fig. 5, this comparison is illustrated for the external potential parameterized by w = 0.65, p = 2.0, and s = 40 (shown in red) and the four chemical potentials βμ ∈ {−2.5, −1.5, −0.5, 0.5}; for symmetry reasons, we only plot the regime 0 < z < L/2, and for comparison, we also show the MF profiles. Clearly, the MF predictions are substantially worse than ML2 and ML3, except at the lowest μ, and ML3 constitutes a small improvement over ML2, especially at the peaks of the profiles at intermediate to high μ. In fact, we can also conclude from Fig. 5 that the main improvement of ML2 and ML3 over MF compared to the simulations concerns the bulk density ρb that is approached in the center of the slit at z = 0, as will be made more explicit below.

FIG. 5.

Density profiles of a (truncated) Lennard-Jones fluid confined in a planar slit characterized by a repulsive external potential given by Eq. (15) with parameters w = 0.65, p = 2.0, s = 40 at temperature kBT/ɛ = 2 and at four chemical potentials βμ ∈ {−2.5, −1.5, −0.5, 0.5} from bottom to top. Symbols stem from the grand-canonical MC simulations, and curves stem from the MF (blue dashed), ML2 (black solid), and ML3 (green solid) functionals; all four state points are part of the training set.

FIG. 5.

Density profiles of a (truncated) Lennard-Jones fluid confined in a planar slit characterized by a repulsive external potential given by Eq. (15) with parameters w = 0.65, p = 2.0, s = 40 at temperature kBT/ɛ = 2 and at four chemical potentials βμ ∈ {−2.5, −1.5, −0.5, 0.5} from bottom to top. Symbols stem from the grand-canonical MC simulations, and curves stem from the MF (blue dashed), ML2 (black solid), and ML3 (green solid) functionals; all four state points are part of the training set.

Close modal

In Fig. 6, we consider a comparison of MC simulations with MF, ML2, and ML3 density profiles in a particular external potential outside the training set, at βμ = −1. The external potential consists of hard walls at z = 0 and z = 20σ, and for z ∈ [0, 20σ], the potential varies irregularly with wells and barriers between ±kBT, as shown by the red solid curve in Fig. 6. We see again that both ML2 and ML3 are largely of comparable quality and substantially more accurate than MF.

FIG. 6.

The equilibrium density profile for a LJ fluid at chemical potential βμ = −1 and temperature kBT/ɛ = 2 in the external potential Vext(z) (red solid line) outside the ML training set. Symbols stem from grand-canonical Monte Carlo (MC) simulations, and lines represent DFT predictions based on the mean-field (MF) approximation (blue dashed) and on the quadratic (ML2, black solid) and cubic (ML3, green solid) corrections with machine-learned kernels.

FIG. 6.

The equilibrium density profile for a LJ fluid at chemical potential βμ = −1 and temperature kBT/ɛ = 2 in the external potential Vext(z) (red solid line) outside the ML training set. Symbols stem from grand-canonical Monte Carlo (MC) simulations, and lines represent DFT predictions based on the mean-field (MF) approximation (blue dashed) and on the quadratic (ML2, black solid) and cubic (ML3, green solid) corrections with machine-learned kernels.

Close modal

The (isothermal) mechanical bulk equation of state provides relations between the bulk density ρb, the pressure p, and the chemical potential μ, satisfying the constraint of the Gibbs–Duhem equation dp = ρb such that we can equivalently consider ρb(μ), p(μ), or p(ρb). Within DFT, the bulk density ρb(μ) that follows from a particular free-energy excess functional follows from the solution of the Euler–Lagrange equation (4) for the homogeneous bulk case Vext ≡ 0, which reduces for ML2 and ML3 to a nonlinear algebraic equation with coefficients that depend on ∫dz ωi(z). Hence, ρb(μ) is straightforwardly solved numerically for the three functionals ML2, ML3, and MF of our interest here. The bulk pressure follows as p(μ) = −Ω[ρb]/V, from which p(ρb) follows upon inversion of ρb(μ). For the temperature of interest, kBT/ɛ = 2, these three representations of the equation of state are shown in Figs. 7(a)7(c) for the three functionals MF (blue dashed line), ML2 (black solid line), and ML3 (green solid line) together with the MC data (purple symbols). The regime of the training set is hatched gray. In the μ-dependent curves of (a) ρb(μ) and (b) p(μ), we find agreement in the low-density limit βμ < −3, as expected, since all functionals include the ideal-gas limit properly. In the regime of the training set, we also see ML2 and ML3 outperforming MF by a large margin in (a) and (b), with a small but hardly noticeable improvement of ML3 compared to ML2, as we could have expected on the basis of the density profiles of Fig. 6 and the loss functions of Fig. 3. At the high-μ side outside the training set, Fig. 7(a) shows an increasingly deteriorating quality of the ML2 and, especially, the ML3 prediction, which are systematically higher than the MC data, although they are still much more accurate than the predictions based on the MF functional. Interestingly, however, the picture that emerges from the p(ρb) representation shown in Fig. 7(c) is much more forgiving for the MF functional, which is now of comparable good agreement in the complete regime of the training set and deviates as much as ML2 (and even less than ML3) from the MC data. Clearly, this relatively good MF and ML2 performance is due to a fortunate cancellation of errors occurring in the process of eliminating the dependence on the chemical potential.

FIG. 7.

The relations between (a) the bulk density ρb and (b) the bulk pressure p as a function of the chemical potential μ for the (truncated) Lennard-Jones fluid at temperature kBT/ɛ = 2 as obtained from grand-canonical Monte Carlo simulations (MC, symbols) and the machine-learning functionals ML2 (black) and ML3 (green), and the mean-field function (MF, blue dashed). In (c), the corresponding pressure–density relation p(ρb) is shown, obtained by the elimination of μ from ρb(μ) of (a) and p(μ) of (b).

FIG. 7.

The relations between (a) the bulk density ρb and (b) the bulk pressure p as a function of the chemical potential μ for the (truncated) Lennard-Jones fluid at temperature kBT/ɛ = 2 as obtained from grand-canonical Monte Carlo simulations (MC, symbols) and the machine-learning functionals ML2 (black) and ML3 (green), and the mean-field function (MF, blue dashed). In (c), the corresponding pressure–density relation p(ρb) is shown, obtained by the elimination of μ from ρb(μ) of (a) and p(μ) of (b).

Close modal

It is, perhaps, remarkable that rather accurate bulk equations of state in a complete density interval can be obtained from MC simulations at only a few chemical potentials in only a few external potentials. Here, it is crucial to appreciate the DFT formalism, which includes the statement that the intrinsic excess free-energy functional Fexc[ρ] that we construct by the ML ansatz of Eqs. (9) and (10) is independent of the external and chemical potentials and, hence, can also be applied at any μ in the homogeneous bulk where Vext ≡ 0.

A key feature of DFT is that it provides not only thermodynamic but also structural information, where we have seen that the first functional derivative δFexc[ρ]/δρ(r) plays a key role in the Euler–Lagrange equation (4) for the equilibrium one-body distribution function. We have also seen already that the second functional derivative βδ2Fexc[ρ]/δρ(r)δρrcr,r equals the direct correlation function and governs the two-body distribution function.1,2 In a homogeneous and isotropic bulk fluid, symmetry dictates that the direct correlation takes the bulk form cb(|rr′|), and the radial distribution function g(r) follows from the Ornstein–Zernike equation g(r)1=cb(r)+ρbdr(g(r)1)cb(|rr)|. Since the ML2 and ML3 functionals have been fully determined in terms of ωi(z) in planar geometry and its conversion to Ωi(r) according to Eq. (12), we can write from Eqs. (9) and (10) that

(21)

where for cHS(r), we used the White-Bear mark I direct correlation function reported in Ref. 25 (and where Ω3 ≡ 0 for ML2). Consequently, our ML2 and ML3 functionals (and likewise also the MF functional) give direct access to the two-body structure encoded in cb(r) and g(r).

In Fig. 8, we plot the resulting cb(r) for bulk density ρbσ3 = 0.39 and find fairly good agreement between MF, ML2, and ML3, except close to r = 0 where cb(r) from ML2 and, especially, ML3 become deeply negative. This can be traced back directly to Fig. 4, which reveals that (i) there is close structural similarity between the functionals outside the hard core and (ii) that the “bumps” of ω2(z) and ω3(z) close to z = 0 give rise to a substantial slope 3(z)/dz for z/σ ∈ [0, 0.05] that translates via Eq. (12) in a relatively large effect in cb(r) in the vicinity of r = 0. Note also that cb(r) vanishes for d < rσ due to the Barker–Henderson splitting and that all three versions of cb(r) agree pretty accurately outside the hard core, at least on the scale of the plot.

FIG. 8.

The Ornstein–Zernike direct correlation function cb(r) of the bulk Lennard-Jones system at temperature kBT/ɛ = 2 and bulk density ρbσ3 = 0.39, as predicted by the second functional derivative of the excess free-energy functional within the mean-field (MF) approximation and its quadratic (ML2) and cubic (ML3) Machine-Learning corrections.

FIG. 8.

The Ornstein–Zernike direct correlation function cb(r) of the bulk Lennard-Jones system at temperature kBT/ɛ = 2 and bulk density ρbσ3 = 0.39, as predicted by the second functional derivative of the excess free-energy functional within the mean-field (MF) approximation and its quadratic (ML2) and cubic (ML3) Machine-Learning corrections.

Close modal

Upon insertion of cb(r) into the (Fourier transform of the) Ornstein–Zernike equation, we find (after an inverse Fourier transform) the radial distribution functions g(r) that we compare with canonical MC simulations at a given density ρb (at the fixed temperature of interest kBT/ɛ = 2). The three lines in Fig. 9 show these radial distribution functions for MF, ML2, and ML3 at bulk densities (a) ρbσ3 = 0.39 and (b) ρbσ3 = 0.837, together with the MC simulation results (symbols). For both the lower density in (a) and the higher one in (b), we find reasonably good overall agreement outside the hard core (r > d), with MF and ML2 actually outperforming ML3 close to contact. Inside the hard core, our prediction for g(r) is poor in all cases, which is not surprising, given that the underlying cHS(r) is constructed such as to cause vanishing g(r) inside the hard core; any tampering of the direct correlation [such as adding terms as we do in Eq. (21)] will give rise to spurious nonzero contributions to g(r) for r < d.26 

FIG. 9.

The radial distribution function g(r) of a truncated Lennard-Jones fluid at bulk density (a) ρbσ3 = 0.39 and (b) ρbσ3 = 0.837, as obtained from the Ornstein–Zernike equation with a direct correlation function cb(r) that follows from the free-energy functionals ML2 (black), ML3 (green), and MF (blue dashed). The symbols denote g(r) as obtained from canonical Monte Carlo simulations at the same bulk density and temperature.

FIG. 9.

The radial distribution function g(r) of a truncated Lennard-Jones fluid at bulk density (a) ρbσ3 = 0.39 and (b) ρbσ3 = 0.837, as obtained from the Ornstein–Zernike equation with a direct correlation function cb(r) that follows from the free-energy functionals ML2 (black), ML3 (green), and MF (blue dashed). The symbols denote g(r) as obtained from canonical Monte Carlo simulations at the same bulk density and temperature.

Close modal

Interestingly, DFT provides another procedure to calculate the radial distribution function of a bulk fluid. This so-called “Percus test-particle method”27 is based on the identification of ρbg(r) with the equilibrium density profile ρ0(r) that surrounds a given (test) particle that is fixed in the origin of an otherwise homogeneous fluid at bulk density ρbρ0(). In other words, g(r) = ρ0(r)/ρ0(), with ρ0(r) being the spherically symmetric density profile of the fluid in an external potential that equals the pair potential, Vext(r) = ϕ(r), scaled such that g() = 1. For a given chemical potential μ and a given functional Fexc[ρ], one thus obtains g(r) through the solution ρ0(r) of the Euler–Lagrange equation (4). For the same two state points as used in Fig. 9, we present the resulting radial distributions in Figs. 10(a) and 10(b). For both densities, the agreement between simulation and all three DFTs is substantially better than obtained from the Ornstein–Zernike route shown in Fig. 9, not only for r < σ where the Boltzmann factor of the external potential in Eq. (4) ensures a vanishingly small contribution to g(r) but also at larger distances where the oscillations in the MC data are rather accurately captured by all three DFTs. Interestingly, however, the peaks of the oscillations in (b) are actually better accounted for by MF and ML2 than by ML3, which underestimates them especially at close contact. The relatively good performance of MF in predicting g(r) via the Percus test-particle method compared to its relatively poor prediction of the equation of state ρb(μ) is due to the scaling-out of ρb in the density profile ρ0(r) such that g() = 1 by construction. Clearly, an overall comparison of Figs. 9 and 10 shows that g(r) based on the test-particle method is much more accurate compared to the MC simulations than those based on the Ornstein–Zernike equation. This is not surprising, given that we constructed the direct correlation function in Eq. (21) based on a modification of that of a reference hard-sphere system, which yields non-vanishing radial distributions inside the hard core if the OZ route is used. A more careful discussion on the different radial distribution functions from the two routes can be found in Ref. 26.

FIG. 10.

The radial distribution function g(r) for the Lennard-Jones system as obtained from the Percus test-particle method for bulk densities (a) ρbσ3 = 0.39 and (b) ρbσ3 = 0.837. System, legends, and the MC data are identical to those in Fig. 9.

FIG. 10.

The radial distribution function g(r) for the Lennard-Jones system as obtained from the Percus test-particle method for bulk densities (a) ρbσ3 = 0.39 and (b) ρbσ3 = 0.837. System, legends, and the MC data are identical to those in Fig. 9.

Close modal

In this article, we combine the formalism of classical density functional theory (DFT) with machine-learning (ML) density profiles from Monte Carlo (MC) simulations to construct approximations to the excess intrinsic Helmholtz free-energy functional Fexc[ρ] of a (truncated and shifted) Lennard-Jones fluid at the supercritical temperature kBT/ɛ = 2. This functional consists of a well-known and accurate hard-sphere contribution, a standard van der Waals-type mean-field account of the attractions, and new machine-learned corrections that are, for simplicity, either of a quadratic (ML2) or an additional cubic (ML3) form in the density. The kernels of ML2 and ML3 are radially symmetric and translation-invariant two-point functions of the form Ωi(|rr′|) for i = 2 and 3; see Eqs. (9) and (10). By comparing DFT predictions of the equilibrium density profiles ρ0(z) with grand-canonical MC simulations at a learning set of chemical potentials μ and external potentials Vext(z) in a 3D planar geometry, we can construct the optimal planar kernels ωi(|zz′|) using Adam to minimize a suitable loss function, from which we can reconstruct the full radially symmetric kernels Ωi(|rr′|). Given that Fexc[ρ] is independent of the external potential and the chemical potential, the functional and its Euler–Lagrange equation (4) for ρ0(r) can be applied to any μ and any Vext(r). By comparisons with density profiles obtained from grand-canonical MC simulations, for conditions within and outside of the learning set, we find that the ML2 and ML3 functionals generally outperform MF by far because the latter predicts densities that are systematically too low; ML3 improves ML2 somewhat on some of the details at higher μ, at least within the training set. A similar picture emerges from the resulting representations of the mechanical equations of state, viz., the bulk density ρb(μ) and the pressure p(μ), where MF is too low by a large margin and ML3 performs only slightly better than ML2 within the training set, while showing a slightly poorer performance outside. The functional Fexc[ρ] can also be used to calculate the direct pair correlation function, from which the radial distribution g(r) of a bulk fluid follows via the Ornstein–Zernike equation. At the relatively low bulk density ρbσ3 = 0.39, this yields, outside the hard core, an (almost) equally satisfying result for MF, ML2, and ML3; see Fig. 9. Inside the hard core and also close to contact, say, σ < r < 1.3σ, the prediction for g(r) is poor in all cases. However, all three functionals give a rather good account of the simulated g(r) at these two state points if the Percus test-particle method is employed, although here ML3 overestimates the peaks at the higher density slightly. The reason for the relatively good MF performance for g(r) compared to the equation of state ρb(μ) and density profiles stems from the imposed asymptotic normalization g(r) = 1.

A disadvantage of the ML approach is its black-box character and the associated difficulty in interpreting the outcome. In particular, the “hump” in the ML2 and ML3 kernels ωi(z) close to z = 0 shown in Fig. 4 and the associated deeply negative direct correlation cb(r) close to r = 0 for ML3 in Fig. 8 are actually rather suspicious. In retrospect, we expect these features to be the result of some degree of overfitting the data in the learning process. This is also borne out by closer inspection of the bulk equations of state ρb(μ) and p(μ) in Figs. 7(a) and 7(b), respectively, where ML3 hardly improves upon ML2 in the (hatched) regime of the learning set while performing even poorer outside and, likewise, for g(r) of Figs. 9(a) and 10(a) at the density ρbσ3 = 0.39 that lies comfortably in the middle of the training set. Of course, ML3 does outperform ML2 somewhat for the density profiles of Fig. 5. Nevertheless, some more caution could or should have been exercised in the diversity of the training set of external potentials, perhaps with attractive components and discontinuities. We leave studies along these lines for future work.

Although there is room for improvement and extensions, we have shown here anyway that it is, in principle, possible to construct a free-energy functional for an atomic fluid by an ML process that takes data from grand-canonical MC simulations at a variety of chemical and external potentials, from which further predictions outside the training set can be made. Interestingly, even data taken in a planar geometry can suffice to construct the full functional, at least for the (relatively simple) functional forms that we considered here, which are linear in the kernels Ωi(|rr′|); nonlinear forms probably require a different treatment. It is important to realize that we fixed the temperature, and although Fexc[ρ] is independent of μ and Vext(r), it is dependent on T, so strictly speaking, a new functional is to be constructed at every temperature of interest. We leave the T-dependence of the functional to future work. Another rather straightforward extension is to use the newly constructed functional to calculate the Gibbs adsorption and the tensions of wall–fluid interfaces, for which we expect good accuracy on the basis of the good agreement of the density profiles. We also expect that it is possible to extend studies of this type to other systems with spherically symmetric particles, also to mixtures such as electrolytes. Systems of particles with orientation degrees of freedom are probably challenging, in practice, because of their larger number of variables, although one could imagine first attempts based on truncated expansions in spherical harmonics or an initial focus on homogeneous bulk states (nematics). We hope that this paper will stimulate further explorations of the combination of DFT, ML, and MC simulation.

This work was part of the D-ITP consortium, a program of The Netherlands Organisation for Scientific Research (NWO) that is funded by the Dutch Ministry of Education, Culture and Science (OCW). It also forms part of the NWO program “Data-driven science for smart and sustainable energy research,” with Project Nos. 16DDS003 and 16DDS014.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
R.
Evans
, “
The nature of the liquid-vapour interface and other topics in the statistical mechanics of non-uniform, classical fluids
,”
Adv. Phys.
28
,
143
200
(
1979
).
2.
J.-P.
Hansen
and
I. R.
McDonald
,
Theory of Simple Liquids: With Applications to Soft Matter
(
Academic Press
,
2013
).
3.
S.-C.
Lin
and
M.
Oettel
, “
A classical density functional from machine learning and a convolutional neural network
,”
SciPost Phys.
6
,
25
(
2019
).
4.
S.-C.
Lin
,
G.
Martius
, and
M.
Oettel
, “
Analytical classical density functionals from an equation learning network
,”
J. Chem. Phys.
152
,
021102
(
2020
).
5.
R.
Evans
, in
Fundamentals of Inhomogeneous Fluids
, edited by
D.
Henderson
(
CRC Press
,
1992
).
6.
Y.
Rosenfeld
, “
Free-energy model for the inhomogeneous hard-sphere fluid mixture and density-functional theory of freezing
,”
Phys. Rev. Lett.
63
,
980
(
1989
).
7.
A.
Härtel
,
M.
Oettel
,
R. E.
Rozas
,
S. U.
Egelhaaf
,
J.
Horbach
, and
H.
Löwen
, “
Tension and stiffness of the hard sphere crystal-fluid interface
,”
Phys. Rev. Lett.
108
,
226101
(
2012
).
8.
A.
Härtel
,
M.
Janssen
,
S.
Samin
, and
R.
van Roij
, “
Fundamental measure theory for the electric double layer: Implications for blue-energy harvesting and water desalination
,”
J. Phys.: Condens. Matter
27
,
194129
(
2015
).
9.
M.
Marechal
and
H.
Löwen
, “
Density functional theory for hard polyhedra
,”
Phys. Rev. Lett.
110
,
137801
(
2013
).
10.
M.
Schmidt
,
H.
Löwen
,
J. M.
Brader
, and
R.
Evans
, “
Density functional for a model colloid-polymer mixture
,”
Phys. Rev. Lett.
85
,
1934
(
2000
).
11.
R.
Roth
, “
Fundamental measure theory for hard-sphere mixtures: A review
,”
J. Phys.: Condens. Matter
22
,
063102
(
2010
).
12.
H.
Hansen-Goos
and
R.
Roth
, “
Density functional theory for hard-sphere mixtures: The White Bear version mark II
,”
J. Phys.: Condens. Matter
18
,
8413
(
2006
).
13.
N. D.
Mermin
, “
Thermal properties of the inhomogeneous electron gas
,”
Phys. Rev.
137
,
A1441
(
1965
).
14.
S.
de Wind
, “
Constructing the excess Helmholtz free-energy functional of a supercritical Lennard-Jones fluid with machine learning
,” Bachelor’s thesis,
University Utrecht
,
2019
.
15.
J. A.
Barker
and
D.
Henderson
, “
Perturbation theory and equation of state for fluids. II. A successful theory of liquids
,”
J. Chem. Phys.
47
,
4714
4721
(
1967
).
16.
Y.-X.
Yu
, “
A novel weighted density functional theory for adsorption, fluid-solid interfacial tension, and disjoining properties of simple liquid films on planar solid surfaces
,”
J. Chem. Phys.
131
,
024704
(
2009
).
17.
R. L.
Cotterman
,
B. J.
Schwarz
, and
J. M.
Prausnitz
, “
Molecular thermodynamics for fluids at low and high densities. Part I: Pure fluids containing small or large molecules
,”
AIChE J.
32
,
1787
1798
(
1986
).
18.
A. J.
Archer
,
D.
Pini
,
R.
Evans
, and
L.
Reatto
, “
Model colloidal fluid with competing interactions: Bulk and interfacial properties
,”
J. Chem. Phys.
126
,
014104
(
2007
).
19.
P. I.
Ravikovitch
,
A.
Vishnyakov
, and
A. V.
Neimark
, “
Density functional theories and molecular simulations of adsorption and phase transitions in nanopores
,”
Phys. Rev. E
64
,
011602
(
2001
).
20.
Y.
Tang
,
Z.
Tong
, and
B. C.-Y.
Lu
, “
Analytical equation of state based on the Ornstein-Zernike equation
,”
Fluid Phase Equilib.
134
,
21
42
(
1997
).
21.
Y.
Tang
, “
Role of the Barker–Henderson diameter in thermodynamics
,”
J. Chem. Phys.
116
,
6694
6700
(
2002
).
22.
Y.
Tang
and
J.
Wu
, “
A density-functional theory for bulk and inhomogeneous Lennard-Jones fluids from the energy route
,”
J. Chem. Phys.
119
,
7388
7397
(
2003
).
23.
F.
Ruehle
, “
Data science applications to string theory
,”
Phys. Rep.
839
,
1
117
(
2020
).
24.
D. P.
Kingma
and
J.
Ba
, “
Adam: A method for stochastic optimization
,” arXiv:1412.6980 (
2014
).
25.
R.
Roth
,
R.
Evans
,
A.
Lang
, and
G.
Kahl
, “
Fundamental measure theory for hard-sphere mixtures revisited: The White Bear version
,”
J. Phys.: Condens. Matter
14
,
12063
12078
(
2002
).
26.
A. J.
Archer
,
B.
Chacko
, and
R.
Evans
, “
The standard mean-field treatment of inter-particle attraction in classical DFT is better than one might expect
,”
J. Chem. Phys.
147
,
034501
(
2017
).
27.
J. K.
Percus
, “
Approximation methods in classical statistical mechanics
,”
Phys. Rev. Lett.
8
,
462
(
1962
).