To explore the curvature dependence of solid–fluid interfacial thermodynamics, we calculate, using Grand Canonical Monte Carlo simulation, the surface free energy for a 2d hard-disk fluid confined in a circular hard container of radius R as a function of the bulk packing fraction η and wall curvature C̄=1/R. (The curvature is negative because the surface is concave.) Combining this with our previous data [Martin et al., J. Phys. Chem. B 124, 7938–7947 (2020)] for the positive curvature case (a hard-disk fluid at a circular wall, C̄=+1/R), we obtain a complete picture of surface thermodynamics in this system over the full range of positive and negative wall curvatures. Our results show that γ is linear in C̄ with a slope that is the same for both positive and negative wall curvatures, with deviations seen only at high negative curvatures (strong confinement) and high density. This observation indicates that the surface thermodynamics of this system is consistent with the predictions of so-called morphometric thermodynamics at both positive and negative curvatures. In addition, we show that classical density functional theory and a generalized scaled particle theory can be constructed that give excellent agreement with the simulation data over most of the range of curvatures and densities. For extremely high curvatures, where only one or two disks can occupy the container at maximum packing, it is possible to calculate γ exactly. In this limit, the simulations and density functional theory calculations are in remarkable agreement with the exact results.

On first glance, a nanoparticle solvated by a fluid and the same fluid confined in a nanopore would seem to be very different physical systems; however, they both can be described as a fluid at a curved surface—a convex surface in the case of the nanoparticle and concave one for the nanopore. This raises the question as to whether these two situations can be understood within a unified thermodynamic description. The central quantity governing the thermodynamics of surfaces is the surface (or interfacial) free energy, γ, which plays a significant role in a number of phenomena of scientific and technological import, including wetting, dendritic growth and morphology, nucleation, and nanoparticle stability.1–3 Although most experimental and computational studies of solid–liquid interfaces focus on planar interfaces, the role of curvature (or shape) on interface thermodynamics has long been of interest since the pioneering work of Gibbs;4 however, besides recognition of the fact that the local curvature of the interface plays a role in determining the work required to deform the interface, Gibbs restricted further study to planar interfaces.

In the years since Gibbs, significant effort has been focused on the development of quantitative descriptions of the curvature dependence of γ. Perhaps the best known of these comes from the work of Tolman5 for a fluid at a spherical surface of radius R, in which

γ=γ01+2δ/Rγ0(12δ/R+),
(1)

where γ0 is the surface free energy of a planar interface and δ is the so-called Tolman length. A number of approaches have been since put forward to generalize the ideas of Tolman to more complex surfaces. For a 3d surface, Helfrich postulated that γ can be expanded in a power series in the average mean (H̄) and average Gaussian (K̄) curvatures of the surface,6 

γ=i=0j=0γijH̄iK̄j,
(2)

where the sum is usually truncated at low order. For a spherical wall of radius R, where H̄=1/R and K̄=1/R2, this would give

γ=γ00+γ10/R+(γ01+γ20)/R2+.
(3)

For a 2d surface, the Helfrich expression would be a Taylor series expansion for γ in powers of the average curvature of the surface C̄, which for a circular wall (disk) would be equal to 1/R.

A more recent development is the so-called Morphometric Thermodynamics (MT),7–9 based on a theorem from integral geometry due to Hadwiger10 that states that a valuation defined on a convex set S in d dimensions can be completely described as a linear combination of d + 1 geometric measures (volume, surface area and d − 1 measures of curvature) as long as the valuation is continuous, motion invariant, and additive. In MT, the valuation is the solvation free energy, Ω, of a nanoparticle with a surface (represented by a convex set S) embedded in a fluid medium.8 

For a 3d system, the MT solvation free energy of the nanoparticle is then given by

Ω(MT)=pV+γ0A+γ1H+γ2K,
(4)

where V, A, H, and K are the surface volume, area, integrated mean curvature, and integrated Gaussian curvature, respectively. Defining the surface free energy γ as γ ≡ (Ω − pV)/A gives

γ3d(MT)=γ0+γ1H̄+γ2K̄,
(5)

where H̄S and K̄S are the average mean and Gaussian curvatures of S, respectively. For a spherical wall particle of radius R(C̄=1/R), the MT expression is

γsph=γ0+γ1/R+γ2/R2.
(6)

Here, γ1 is equal to −2δ, where δ is the Tolman coefficient. For the 2d hard-disk fluid at a circular wall system, there is only one curvature, C̄, giving

γ2d(MT)=γ0+γ1C̄=γ0+γ1/R.
(7)

For a 3d fluid of hard spheres of diameter σ at a spherical hard wall, molecular-dynamics (MD) simulations11,12 have shown that the MT description [Eq. (6)] works extremely well at low to intermediate packing fractions with deviations (beyond quadratic in 1/R) outside of the statistical error appearing only at high packing fractions (ηπσ3ρ/6 > 0.42) approaching the melting density. These simulation results are in agreement with previous calculations using classical Density Functional Theory (cDFT)13 on the same system,14 which showed no deviations from MT below η = 0.42, with no solutions to the cDFT equations above this value. Similar agreement was seen in cDFT calculations for the hard-sphere fluid at a variety of polyhedral walls.3 For the same fluid at a cylindrical hard wall (3d), deviations from MT at low packing fractions were observed in the simulations, but they required very high-precision calculations to resolve.12 In the case of a 2d fluid of hard disks with diameter σ at a circular hard wall of radius R, previous Monte Carlo simulations by the authors showed that the MT expression [Eq. (7)] with C̄=1/R is also very accurate over a wide range of fluid packing fractions and wall curvatures.15 (Note that, in what follows, we assume, without loss of generality, that all lengths are measured in units of σ and all energies are in units of kBT, where kB is the Boltzmann constant.)

A central question is whether the MT formalism can be extended to fluids confined inside a container with curved boundary—for example, within a spherical (3d) or circular (2d) container. Relative to the solvated particle case, where the fluid is outside the curved boundary, this corresponds to a wall curvature that is negative. (Note that, Hadwiger’s theorem does not apply for such concave systems, as it is limited to compact convex sets.) For example, a hard-disk fluid (2d) contained within a circle of radius R, the average curvature C̄=1/R. In the limit that R (zero curvature), both the solvated particle and the confined system have the same limit, namely the fluid in contact with a flat wall and we expect that the surface free energy will be continuous in the zero curvature limit. Sitta et al.16 have examined this problem using a hard-rectangle fluid both at a circular wall and in confinement within a circular container, showing continuity in the slope of γ vs C̄ through the transition (C̄=0) between the confined and the open systems, although only a very limited range of curvatures near zero was studied.

Here, we extend our earlier work to a hard-disk fluid confined within a hard circular container of radius R. The MT equations above will be identical, except the surface free energy is defined slightly differently as γS(ΩS+pVS)/AS. The motivation for examining this in this 2d system is that the MT expression for γ is simpler (linear in 1/R) and relatively large-size systems can be simulated with modest computational cost. This system has been studied in the past to examine the density variation of the fluid as a function of confining radius,17 but this is, to our knowledge, the first systematic study of the thermodynamics of hard disks in circular confinement (negative curvature) over a wide range of confining radii. Combined with our previous study for the hard-disk fluid at a circular wall (positive curvature), this study allows us to examine the effects of curvature on the surface thermodynamics in a unified way for both bounding wall and confinement geometries.

In our previous work on the hard-sphere and hard-disk fluids at curved convex walls, we calculated the surface free energy using the Gibbs–Cahn integration method, in which we use the adsorption relation

γp=vex=VN/ρbA,
(8)

where ρb is the single particle density of the bulk fluid and vex is the excess volume, defined as the difference (per unit area) between the volume (V) of a region completely encompassing the interface and the volume of the bulk fluid containing the same number of particles (N) as the interfacial region. The value of γ at a particular pressure (or density) is then found by determining vex as a function of pressure (or density) and integrating Eq. (8). For the hard-sphere system, this is simplified because one can assume that γ(P = 0) = 0.

Because of the confinement, the pressure of a hard-disk fluid encapsulated within a hard circle of radius R will differ from that of a bulk fluid with the same chemical potential as a consequence of the Young–Laplace equation and the small system size, where the curvature dependence of the surface free energy can be significant.18 Because of this, the Gibbs–Cahn method must be modified. Consider a fluid inside a container in equilibrium with a bulk fluid at chemical potential μ, we can define the surface free energy of the interface by

γA=ΩΩb=Ω+pbV,
(9)

where Ω and Ωb = −pbV are the grand potential of the fluid inside the container and that of a bulk fluid with the same volume, chemical potential, and temperature, respectively. Ω is related to the internal energy U of the fluid by the Legendre transform

Ω=UTSμN,
(10)

with differential

dΩ=dUTdSSdTμdNNdμ.
(11)

The differential for U for this system (including the surface work term) is given by

dU=TdSpdV+μdN+γdA,
(12)

where p is the pressure inside the pore, which is not the same as the pressure of the bulk fluid at the same μ and T. From Eq. (9), we have

d(γA)=dΩ+pbdV+Vdpb.
(13)

Substituting Eqs. (11) and (12) into Eq. (13) yields

Adγ=SdTpdVNdμ+pbdV+Vdpb.
(14)

The differentials in Eq. (14) are not all independent, but they are related to one another through the Gibbs–Duhem equation for the bulk fluid,

0=VbdpbNbdμSbdT,
(15)

where the subscript b indicates the value of the variable in the bulk fluid reservoir. The two simultaneous equations [Eqs. (14) and (15)] can be solved using Cramer’s rule by choosing one of the extensive variables X = N, V, or S, giving

dγ=[S/X]dT[N/X]dμ+[V/X]dpb[Δp/1]dV,
(16)

where Δp = ppb and [Z/X] is defined as

[Z/X]=zex=1AZXZbXbXb,
(17)

with zex defined as the interfacial excess of quantity Z. The choice of X (S, N, or V) eliminates the differential of the intensive variable conjugate to X. In this work, we choose X = N, eliminating the dμ term. We also keep T and V constant, yielding

γpbVT=vex=1AVNVbNb=VA1ρpρb,
(18)

where ρp = N/V and ρb = Nb/Vb are the average densities of the fluid in the pore and bulk reservoir, respectively.

Equation (18) can be integrated over pb to determine the surface free energy as a function of the bulk fluid pressure, or given an equation of state, the packing fraction ηb = πρb/4 of the bulk fluid,

γ(η)=0pbvex(pb)dpb=0ηbvex(ηb)pηbTdηb.
(19)

Here, the dependence of the bulk pressure on packing fraction can be determined to sufficient accuracy from the equation of state (EOS) of Kolafa and Rottner.19 

To determine vex [and, thus, γ through Eq. (19)] for the hard-disk fluid confined within a hard circular wall of radius R, we perform a series of Grand Canonical Monte Carlo (GCMC) simulations (constant μ, V, and T) using a modified version of the HOOMD-blue simulation software.20,21 The software was modified to allow the grand canonical integration scheme to be paired with external potentials. The GCMC simulations require as input the chemical potential (or fugacity) of the bulk fluid, from which the bulk packing fraction can be obtained either through simulation of the bulk fluid or through the application of a high accuracy EOS. To perform simulations of the confined fluid with the same bulk packing fraction as our previous work on positive curvature walls, we apply the Kolafa and Rottner EOS to determine the fugacity by numerical integration

lnfp=0pZ(p)1pdp=0ηZ(η)1ηZ(η)ηZ(η)ηTdη,
(20)

where the compression factor Z = βP/ρ.

The excess volume for the hard-disk fluid/hard circular wall system was calculated directly from average density in the pore using Eq. (18) using GCMC for a range of R from 19/20 = 0.95 to 30 (corresponding to curvatures ranging from −20/19 to −1/30) and bulk packing fractions ranging from 0 to 0.6049. The specific packing fractions were chosen to coincide with those studied for our earlier simulations on the hard-disk fluid/hard circular convex wall system15 for ease of comparison. (For reference, our highest densities are below the first-order fluid/hexatic phase transition for the hard-disk fluid at about η = 0.7.22–24) Data for R = (planar wall) are taken from Ref. 25. The number of GCMC cycles for each simulation was chosen to give similar levels of statistical precision and ranged from 50 × 109 for the smallest cavities at low density to 1.2 × 109 for the largest cavities at high density. The surface free energy γ as a function of η was then obtained through numerical integration of Eq. (19) using the Kolafa–Rottner EOS for the pressure. Simpson’s rule was applied for the numerical integration for all of the even numbered bins in η. For the odd numbered bins (2n + 1), the trapezoid rule was applied to the final bin. For all γ data points, the estimated numerical integration error was smaller than the statistical error.

The simulation results are useful for the validation of theoretical approaches to surface thermodynamics. In this work, we examine two approaches. The first uses a formulation of classical Density Functional Theory (cDFT)13 for 2d hard particles by Roth, Mecke, and Oettel (RMO).26 Within cDFT, the grand potential functional Ω[ρ(r)] is a functional of the single particle density ρ(r) with two important properties: (i) The equilibrium density profile ρ0(r) minimizes the functional; and (ii) at its minimum, the functional reduces to the grand potential of system Ω = Ω[ρ0(r)]. By minimizing the functional via the variational principle

δΩ[ρ]δρ(r)ρ(r)=ρ0(r)=0,
(21)

one obtains the equilibrium structure and thermodynamics at the same time. Hence, one can directly calculate the surface free energy γ and the excess volume vex. The shape of the cavity enters the calculation via the external potential and can be arbitrary. This method makes no assumptions as to the functional form of the curvature dependence of vex and γ, and, therefore, it can predict potential deviations from the MT form.

Here, we employ a fine 2d Cartesian grid, which allows us to make use of efficient convolutions with the help of fast Fourier transforms (FFT)27 on central processing units (CPUs) and graphics processing units (GPUs)28 for the computation of weighted densities. This also makes it possible to treat the circular cavity and the planar geometry on equal footing. Clearly, the Cartesian coordinates introduce small numerical artifacts in the case of circular cavities. In order to minimize artifacts, we employ a square grid of 2048 × 2048 points, and we have verified that the Gibbs adsorption theorem is satisfied.26 For the circular cavities, considered here, it would be possible to make use of the polar symmetry of the problem and reduce the cDFT problem to an effective 1d one, however, at the cost of losing the possibility to employ standard FFT methods for convolutions. In the effective 1d case, one ends up with a Hankel transform, which can be evaluated using FFT methods when a logarithmic grid is employed.29,30

The generalized scaled particle theory (gSPT) method used here is similar to that used in our previous paper on the hard-disk fluid at a planar wall,25 modified to apply to a curved confining wall. While strictly speaking, this approach applies directly to a hard-disk fluid outside a hard circular wall (R > 0), we will be comparing it to both the positive and negative R cases. As in the previous paper, the theory begins with the original scaled particle theory (SPT) expression for the excess free energy density Φ, which reads

ΦSPT=ξ0ln(1ξ2)+ξ124π(1ξ2),
(22)

where ξ2 = πσ2ρ/4 = η, ξ1 = πσρ, and ξ0 = ρ. Here, the scaled particle variable ξ2 is equivalent to the packing fraction η. This free energy is then generalized as follows:

Φ=ξ0ln(1ξ2)f(ξ2)+ξ124π(1ξ2)g(ξ2),
(23)

where f and g are functions of the dimensionless variable ξ2.

The functions f and g can be specified by requiring agreement with a imposed equation of state via the compressibility factor

Z(η)=βpρ,
(24)

which amounts to requiring that

ξ0Z(η)=ξ0Φ+Φξ2ξ2+Φξ1ξ1+Φξ0ξ0.
(25)

This can be then combined with the usual scaled particle relation

ξ0Z(η)=Φξ2.
(26)

One can readily solve Eqs. (25) and (26) for the functions f and g, leading to

f(ξ2)ln(1ξ2)=0ξ21ug(u)1udu
(27)

and

ξ2g(ξ2)1ξ2=(1ξ2)Z(ξ2).
(28)

To extend this to curved confining walls, we use a binary mixture approach; the excess chemical potential of a single hard disk of radius R added to the system can be obtained as

μex=Φξ2πR2+Φξ12πR+Φξ0,
(29)

from which the interfacial free energy for the interface with curvature 1/R is obtained as

γ=Φξ1+Φξ012πR.
(30)

Using the functions f and g, this results in

γσ=2ηg(η)π(1η)ln(1η)f(η)2πR.
(31)

Compressibility factors for the hard-disk fluid include the result from scaled particle theory

ZSPT=1(1η)2,
(32)

which corresponds to f(η) = g(η) = 1 and, hence,

γSPTσ=2ηπ(1η)ln(1η)2πR.
(33)

The more accurate Solana EOS31 reads

ZSol=1+18η2110η4(1η)2
(34)

and leads to

γSolσ=2η+14η215η4π(1η)+γ1Sol(η)R,
(35)

with

γ1Sol(η)=4η4+2η3+η27η80π(1η)4780πln(1η).
(36)

At very large densities, the SMC2 EOS31 with ZSMC2=ZSol50000η40(1η)3 is even more accurate. The resulting functions f and g, and hence γ, can be calculated analytically, but the expressions are lengthy.

The results for the excess volume vex and the surface free energy γ for the confined system as functions of η are plotted in Fig. 1 over the full range of R. To better show the dependence on curvature, we combine the data from the current calculations with those from our previous work for the fluid outside a circular wall15 to show in Fig. 2 the excess surface volume vex a function of the curvature Ĉ over the full range of positive and negative curvatures for selected packing fractions. (Data for all packing fractions and wall curvatures studied can be found in the supplementary material—for clarity, not all packing fractions studied are included in the plots.) Morphometric thermodynamics predicts that vex is linear in the curvature. This is seen in Fig. 2 to hold except for high negative curvature at the highest packing fractions.

FIG. 1.

The excess volume (Top Panel) and surface free energy (Bottom Panel) from GCMC simulation for the hard-disk fluid confined in a hard circular container as functions of packing fraction η for a range of container radii R. Error bars denote the 95% confidence interval.

FIG. 1.

The excess volume (Top Panel) and surface free energy (Bottom Panel) from GCMC simulation for the hard-disk fluid confined in a hard circular container as functions of packing fraction η for a range of container radii R. Error bars denote the 95% confidence interval.

Close modal
FIG. 2.

The excess volume, vex, as a function of curvature, C̄, for selected packing fractions. Filled symbols represent results from the current simulations and open symbols show results from our previous Monte Carlo simulations from Ref. 15 and 25 . Solid lines represent the predictions from classical DFT.

FIG. 2.

The excess volume, vex, as a function of curvature, C̄, for selected packing fractions. Filled symbols represent results from the current simulations and open symbols show results from our previous Monte Carlo simulations from Ref. 15 and 25 . Solid lines represent the predictions from classical DFT.

Close modal

At this point, it is useful to note that, in this work, we define R to represent the actual radius of the circular wall. An alternative definition that is frequently seen uses a radius R′, defined by the point of closest contact of the hard particle to the circular wall.32 For the 2d system here, this gives R′ = R ± 1/2, with the “−” sign for the confined system and the “+” for solvated disk. The data from this work can be converted to the alternate reference using the definition of vex. If we denote the excess volume in the alternate reference as vex, we have

vex=22+C̄vex12C̄8.
(37)

Because the grand potential Ω is a well-defined property of the system, independent of the definition of the radius, we obtain

γ0=γ0pb/2,
(38)
γ1=γ1γ0/2+pb/8,
(39)

where the symbol “′” denotes surface thermodynamic quantities measured using the R′ radius definition. The first non-Hadwiger coefficient γ2, however, is invariant to the choice of radius definition. We have chosen the radius convention used here and in our earlier work because it gives a value for γ0 that is positive and is easier to extend to hard-disk fluid mixtures of different-sized disks.

To see the deviations from MT at high η more clearly, Fig. 3 shows the excess volume vs curvature for the highest packing fractions studied. At the very highest values of η, the slope of the vex vs C̄ curve changes sign for negative curvatures and vex exhibits a minimum at C̄=0. These significant deviations from the MT ansatz at high η are due to the presence of packing constraints in the confined systems that lead to a density maximum far smaller than close packing in the bulk hard-disk fluid. In other words, at high densities the ratio ρp/ρb in Eq. (18) changes from an increasing function of η to a decreasing function of η due to a plateau in ηp as maximum packing in the pore is approached.

FIG. 3.

The excess volume, vex, as a function of curvature, C̄, at the four highest packing fractions studied. Symbols and solid lines are defined as in Fig. 2. For reference, the insets show snapshots from concave and convex systems with R = 3.

FIG. 3.

The excess volume, vex, as a function of curvature, C̄, at the four highest packing fractions studied. Symbols and solid lines are defined as in Fig. 2. For reference, the insets show snapshots from concave and convex systems with R = 3.

Close modal

Figures 2 and 3 also show the predictions from cDFT in comparison to the simulation data. The agreement at low to intermediate packing fractions is excellent. At the higher packing fractions (Fig. 3), the Density Functional Theory (DFT) results for vvex show some deviation from the simulation results for the positive curvature systems; however, the agreement for negative curvatures is remarkably good, with DFT successfully reproducing the nonlinear deviations from the MT predictions.

The results for the surface free energy, γ, as a function of curvature for various packing fractions, are shown in Fig. 4, combined with the data from our previous work for positive curvature.15 For clarity, the same quantities for the highest packing fractions studied are shown in Fig. 5. Moreover, the solid lines in Figs. 4 and 5 show the results for γ from cDFT, which agree extremely well with the simulation results, even up to high negative curvature and high packing fractions. The better agreement between DFT and simulation for γ vs that for vex can be understood by noting that γ is obtained by integration from the excess volume, which decreases the relative differences at high η. The surface free energy follows the MT prediction at low η to intermediate η; however, at the higher packing fractions studies, there is significant deviation from linearity, especially at high negative curvature, due to the effects of confinement. For R = −3 (i.e. C̄=1/3), the pore is six fluid particle diameters across and can accommodate fewer than 32 particles at maximum packing. The additivity condition of Hadwiger’s theorem implies that when correlations in the fluid are on the same length scale as the wall, the geometric foundation of MT no longer holds. However, Fig. 5 shows that even for a highly confined fluid at a high density, where correlations with the wall extended throughout the cavity, MT has strong predictive power.

FIG. 4.

Results for the surface free energy, γ, of the hard-disk fluid at circular walls as a function of C̄. Entries in the legend denote the approximate bulk packing fraction η. Symbols and solid lines are defined as in Fig. 2.

FIG. 4.

Results for the surface free energy, γ, of the hard-disk fluid at circular walls as a function of C̄. Entries in the legend denote the approximate bulk packing fraction η. Symbols and solid lines are defined as in Fig. 2.

Close modal
FIG. 5.

Results for γ as a function of curvature, C̄, for the four highest packing fractions studied. Symbols and solid lines are defined as in Fig. 2.

FIG. 5.

Results for γ as a function of curvature, C̄, for the four highest packing fractions studied. Symbols and solid lines are defined as in Fig. 2.

Close modal

At high packing fractions, the DFT results show an unusual maximum in between the highest two negative curvatures studied by the simulation: R = 0.95 and 1.07735;C̄1.0526 and 0.9282, respectively. These two curvatures correspond to system sizes in which only one and two hard disks, respectively, can fit within the circle at maximum packing. For these systems, the grand partition function can be calculated exactly,33 and, therefore, we have exact results for γ for the two largest negative curvatures,

γ(1disk)=pR212πRlnπfR2+1,
(40)
γ(2disks)=pR212πRlnf2πRR1R2+1/2+2πR2R21arccos(R1/2)/2+πfR2+1,
(41)

where f = eβμ is the fugacity and R′ = R − 1/2. For these two highest negative curvatures, the simulation results agree with the exact expressions above within the error bars (see the comparison table included in the supplementary material). In addition, the agreement with the DFT results, while not perfect, is quite good and shows that the maximum predicted by the DFT is real and not an artifact (Fig. 6).

FIG. 6.

Comparison of simulation results (filled symbols), DFT results (open symbols) and the exact expressions [Eqs. (40) and (41)] for γ for the four highest packing fractions for curvatures corresponding to circular pores containing at most one or two disks.

FIG. 6.

Comparison of simulation results (filled symbols), DFT results (open symbols) and the exact expressions [Eqs. (40) and (41)] for γ for the four highest packing fractions for curvatures corresponding to circular pores containing at most one or two disks.

Close modal

To examine the deviations from MT, we have performed a least-squares fit of the values of γ at each packing fraction to a quadratic in C̄=1/R separately for positive (convex) and negative (concave) curvatures. For C̄0, we use the entire range of curvatures in the fit; however, because of the strong deviations at high curvature for the concave systems, we use only those curvatures in the interval [−1/3, 0] (five data points). The results for γ0 are presented in Fig. 7 along with the predictions from both SPT and gSPT [Eqs. (33) and (36), respectively]. The results for γ0 for the convex and concave systems agree within the statistical error of the calculation, so only the data for the convex system are shown. The SPT expression underestimates the value of γ0, especially at high packing fractions; however, the agreement for gSPT with the simulations is excellent over the entire range. The results for the linear and quadratic coefficients—γ1 and γ2, respectively—are shown in Fig. 8. Here, we see that the predictions of MT, i.e., γ2 = 0 and γ1(R < 0) = γ1(R > 0) hold up to packing fractions of about 0.45, showing a surprisingly large range of validity of MT even for moderate negative curvatures and moderate packing fractions.

FIG. 7.

Comparison of the simulation planar wall value γ0 with the SPT and gSPT predictions.

FIG. 7.

Comparison of the simulation planar wall value γ0 with the SPT and gSPT predictions.

Close modal
FIG. 8.

Values of the coefficients γ1 and γ2 as functions of the packing fraction η for both convex and concave circular walls. Error bars represent the 95% confidence interval. The data for the convex system are taken from the simulation results of Ref. 15. The solid lines represent the values predicted by SPT [Eq. (33)] and gSPT [Eq. (36)].

FIG. 8.

Values of the coefficients γ1 and γ2 as functions of the packing fraction η for both convex and concave circular walls. Error bars represent the 95% confidence interval. The data for the convex system are taken from the simulation results of Ref. 15. The solid lines represent the values predicted by SPT [Eq. (33)] and gSPT [Eq. (36)].

Close modal

The work presented here presents the most comprehensive study of the surface free energy γ of curved surfaces through the zero-total-curvature limit to the knowledge of the authors. Using Monte Carlo simulation, classical Density Functional Theory (cDFT), and a generalized scaled particle theory (gSPT), we find that in the hard-disk fluid at circular hard walls, the predictions of morphometric thermodynamics are highly accurate over a large range of fluid conditions and wall sizes, extending from highly confined fluids to solvated walls on the same size scale as the fluid particles. This indicates that even under conditions where the concavity and additivity requirements of Hadwiger’s theorem are clearly violated, the MT framework has the potential to provide strong predictive power in assessing surface free energies. At high densities and high negative curvatures, both the excess volume (vex) and γ exhibit significant deviations from the predictions of MT due to strong correlations in these dense highly confined systems. The results for γ calculated using cDFT show excellent agreement with the simulation results at all curvatures and packing fractions and both the simulation and cDFT show agreement with the exact values of γ for highly confined systems that can contain at most one or two particles. In addition, gSPT is successful at accurately predicting γ for packing fractions and curvatures where MT works well.

Included in the supplementary material is a table of the raw data for the excess volume vex and the surface free energy γ from the GCMC simulations for all pore sizes R and packing fractions studied. In addition, for the two smallest pores containing a maximum of one or two disks, respectively, a table is included that directly compares the simulation and cDFT values of the surface free energy γ with the results for γ from the exact partition functions.

B.B.L. and S.C.M. acknowledge support from the National Science Foundation for partial funding of this work under Grant No. CHE-1465226.

The authors have no conflicts to disclose.

The data that support the findings of this study are available within the article and its supplementary material.

1.
K.
Lum
,
D.
Chandler
, and
J. D.
Weeks
, “
Hydrophobicity at small and large length scales
,”
J. Phys. Chem.
103
,
4570
4577
(
1999
).
2.
J. R.
Henderson
, “
Solvation of a solvophobic sphere
,”
J. Chem. Phys.
116
,
5039
5045
(
2002
).
3.
Z.
Jin
,
J.
Kim
, and
J.
Wu
, “
Shape effect on nanoparticle solvation: A comparison of morphometric thermodynamics and microscopic theories
,”
Langmuir
28
,
6997
7006
(
2012
).
4.
J. W.
Gibbs
,
The Collected Works of J. Willard Gibbs
, Thermodynamics Vol. I (
Yale University Press
,
New Haven
,
1948
).
5.
R. C.
Tolman
, “
The effect of droplet size on surface tension
,”
J. Chem. Phys.
17
,
333
337
(
1949
).
6.
W.
Helfrich
, “
Elastic properties of lipid bilayers: Theory and possible experiments
,”
Z. Naturforsch., C: J. Biosci.
28
,
693
703
(
1973
).
7.
C. N.
Likos
,
K. R.
Mecke
, and
H.
Wagner
, “
Statistical morphology of random interfaces in microemulsions
,”
J. Chem. Phys.
102
,
9350
9361
(
1995
).
8.
P. M.
König
,
R.
Roth
, and
K. R.
Mecke
, “
Morphological thermodynamics of fluids: Shape dependence of free energies
,”
Phys. Rev. Lett.
93
,
160601
(
2004
).
9.
H.
Hansen-Goos
,
R.
Roth
,
K.
Mecke
, and
S.
Dietrich
, “
Solvation of proteins: Linking thermodynamics to geometry
,”
Phys. Rev. Lett.
99
,
128101
(
2007
).
10.
H.
Hadwiger
,
Vorlesungen Über Inhalt, Oberfläche und Isoperimetrie
(
Springer
,
Berlin, Heidelberg
,
1957
).
11.
B. B.
Laird
,
A.
Hunter
, and
R. L.
Davidchack
, “
Interfacial free energy of a hard-sphere fluid in contact with curved hard surfaces
,”
Phys. Rev. E
86
,
060602
(
2012
).
12.
R. L.
Davidchack
and
B. B.
Laird
, “
Surface free energy of a hard-sphere fluid at curved walls: Deviations from morphometric thermodynamics
,”
J. Chem. Phys.
149
,
174706
(
2018
).
13.
R.
Evans
, “
The nature of liquid-vapour interface and other topics in the statistical mechanics of non-uniform, classical fluids
,”
Adv. Phys.
28
,
143
(
1979
).
14.
P.
Bryk
,
R.
Roth
,
K. R.
Mecke
, and
S.
Dietrich
, “
Hard-sphere fluids in contact with curved substrates
,”
Phys. Rev. E
68
,
031602
(
2003
).
15.
S. C.
Martin
,
H.
Hansen-Goos
, and
B. B.
Laird
, “
Surface free energy of a hard-disk fluid at curved hard walls: Theory and simulation
,”
J. Phys. Chem. B
124
,
7938
7947
(
2020
).
16.
C. E.
Sitta
,
F.
Smallenburg
,
R.
Wittkowski
, and
H.
Löwen
, “
Hard rectangles near curved walls: Tuning the sign of the Tolman length
,”
J. Chem. Phys.
145
,
204508
(
2016
).
17.
S.-C.
Kim
,
Z. T.
Németh
,
M.
Heni
, and
H.
Löwen
, “
Hard-disks in circular cavities: Density functional theory versus simulation
,”
Mol. Phys.
99
,
1875
1881
(
2001
).
18.
I.
Urrutia
, “
Two hard spheres in a pore: Exact statistical mechanics for different shaped cavities
,”
J. Chem. Phys.
133
,
104503
(
2010
).
19.
J.
Kolafa
and
M.
Rottner
, “
Simulation-based equation of state of the hard disk fluid and prediction of higher-order virial coefficients
,”
Mol. Phys.
104
,
3435
3441
(
2006
).
20.
J. A.
Anderson
,
M.
Eric Irrgang
, and
S. C.
Glotzer
, “
Scalable Metropolis Monte Carlo for simulation of hard shapes
,”
Comput. Phys. Commun.
204
,
21
30
(
2016
).
21.
J. A.
Anderson
,
J.
Glaser
, and
S. C.
Glotzer
, “
HOOMD-blue: A Python package for high-performance molecular dynamics and hard particle Monte Carlo simulations
,”
Comput. Mater. Sci.
173
,
109363
(
2020
).
22.
A.
Jaster
, “
The hexatic phase of the two-dimensional hard disk system
,”
Phys. Lett. A
330
,
120
125
(
2004
).
23.
E. P.
Bernard
and
W.
Krauth
, “
Two-step melting in two dimensions: First-order liquid-hexatic transition
,”
Phys. Rev. Lett.
107
,
155704
(
2011
).
24.
M.
Engel
,
J. A.
Anderson
,
S. C.
Glotzer
,
M.
Isobe
,
E. P.
Bernard
, and
W.
Krauth
, “
Hard-disk equation of state: First-order liquid-hexatic transition in two dimensions with three simulation methods
,”
Phys. Rev.E
87
,
042134
(
2013
).
25.
S. C.
Martin
,
B. B.
Laird
,
R.
Roth
, and
H.
Hansen-Goos
, “
Thermodynamics of the hard-disk fluid at a planar hard wall: Generalized scaled-particle theory and Monte Carlo simulation
,”
J. Chem. Phys.
149
,
084701
(
2018
).
26.
R.
Roth
,
K.
Mecke
, and
M.
Oettel
, “
Communication: Fundamental measure theory for hard disks: Fluid and solid
,”
J. Chem. Phys.
136
,
081101
(
2012
).
27.
R.
Roth
, “
Fundamental measure theory of hard-sphere mixtures: A review
,”
J. Phys.: Condens. Matter
22
,
063102
(
2010
).
28.
D.
Stopper
and
R.
Roth
, “
Massively parallel GPU-accelerated minimization of classical density functional theory
,”
J. Chem. Phys.
147
,
064508
(
2017
).
29.
A. J. S.
Hamilton
, “
Uncorrelated modes of the non-linear power spectrum
,”
Mon. Not. R. Astron. Soc.
312
,
257
284
(
2000
).
30.
V.
Botan
,
F.
Pesth
,
T.
Schilling
, and
M.
Oettel
, “
Hard sphere fluids in annular wedges: Density distributions and depletion potentials
,”
Phys. Rev. E
79
,
061402
(
2009
).
31.
A.
Mulero
,
I.
Cachadiña
, and
J. R.
Solana
, “
The equation of state of the hard-disk fluid revisited
,”
Mol. Phys.
107
,
1457
1465
(
2009
).
32.
A.
Reindl
,
M.
Bier
, and
S.
Dietrich
, “
Implications of interface conventions for morphometric thermodynamics
,”
Phys. Rev. E
91
,
022406
(
2015
).
33.
I.
Urrutia
, “
Two hard spheres in a spherical pore: Exact analytic results in two and three dimensions
,”
J. Stat. Phys.
131
,
597
611
(
2008
).

Supplementary Material