Estimating Fluid-solid Interfacial Free Energies for Wettabilities: A Review of Molecular Simulation Methods
Abstract
Fluid-solid interfacial free energy (IFE) is a fundamental parameter influencing wetting behaviors, which play a crucial role across a broad range of industrial applications. Obtaining reliable data for fluid-solid IFE remains challenging with experimental and semi-empirical methods, and the applicability of first-principle theoretical methods is constrained by a lack of accessible computational tools. In recent years, a variety of molecular simulation methods have been developed for determining the fluid-solid IFE. This review provides a comprehensive summary and critical evaluation of these techniques. The developments, fundamental principles, and implementations of various simulation methods are presented from mechanical routes, such as the contact angle approach, the technique using Bakker’s equation, and the Wilhelmy simulation method, as well as thermodynamic routes, including the cleaving wall method, the Frenkel-Ladd technique, and the test-volume/area methods. These approaches can be applied to compute various fluid-solid interfacial properties, including IFE, relative IFE, surface stress, and superficial tension, although these properties are often used without differentiation in the literature. Additionally, selected applications of these methods are reviewed to provide insight into the behavior of fluid-solid interfacial energies in diverse systems. We also illustrate two interpretations of the fluid-solid IFE based on the theory of Navascués and Berry and Bakker’s equation. It is shown that the simulation methods developed from these two interpretations are identical. This review advocates for the broader adoption of molecular simulation methods in estimating fluid-solid IFE, which is essential for advancing our understanding of wetting behaviors in various chemical systems.
KEYWORDS: Interfacial Free Energy; Interfacial Tension; Contact Angle; Fluid-solid Interface; Molecular Simulation.
1 Introduction
Accurately characterizing wetting interactions at liquid-solid interfaces is essential for optimizing processes in industries such as oil recovery, food production, pharmaceuticals, and coatings, where interfacial phenomena directly influence material performance and process efficiency.1, 2 Historically, determining the interfacial properties of materials has advanced in parallel with the development of thermodynamic principles, with foundational work by Gibbs3 setting the stage for modern investigations. In particular, interfacial free energy (IFE) plays a key role in interface science.4, 5, 6, 7 Understanding IFE is essential for the theoretical explanation of the interfacial phenomenon and for accurate predictions of material behavior in different industrial and engineering applications.
In the three-phase contact region of fluid-fluid-solid systems, IFEs are balanced as described by Young’s equation:8
(1) |
where is the wettability of fluid phase 1 () on solid phase () surrounded by fluid phase 2 (), , , and are the IFEs between phases denoted in the subscript.
In experiments, although fluid-fluid IFE and wettability can be directly obtained,9, 10 the measurement of fluid-solid IFE is difficult and usually inaccurate.11 Many techniques12, 13, 14 exist for ascertaining the vacuum-solid IFE and a substantial body of experimental findings has been systematically correlated by Kumikov and Khokonov.15 Meanwhile, due to the limitations of experimental techniques when approaching the nanoscale interfaces, indirect methods have been developed for the fluid-solid IFE of various systems, including cleavage test,16 solubility test,17 adhesion force measurement,18 deformation analysis on the solid film,19 contact angle/contact line curvature measurement,20 and contact angle experiment combined with Makkonen hypothesis.21 A recent review comprehensively discusses both indirect and direct techniques for determining the IFE of fluid-solid interfaces in experiments, highlighting advancements and ongoing challenges.11 However, precisely determining fluid-solid IFE experimentally remains challenging due to factors such as the size effects, contaminations, and irregularities of surfaces.22, 23, 11
Semi-empirical theories have been extensively employed to estimate fluid-solid IFE using measured contact angle data. These theories encompass approaches like Neumann’s equation of state approach,24, 25 Zisman method,26 Fowkes method,27 geometric-mean approach,28 harmonic-mean approach,29 and van Oss-Good method.30 A comprehensive summary of these semi-empirical theories, along with their assumptions, is provided by Żenkiewicz.31 However, limited research has focused on providing a microscopic foundation for these semi-empirical theories, and significant controversies remain over the validity of several ad hoc assumptions inherent in their formulations.32, 33, 34, 11
First-principle theories have also been developed for determining the interfacial energies. In the statistical mechanical theory of Navascués and Berry35, the Kirkwood-Buff method36 was extended to the case where fluids are in contact with a rigid solid phase, and the fluid-solid IFE was split into one solid and two fluid-solid contributions. Note that this theory was later combined with molecular simulation37 for estimating the fluid-solid relative IFE, which will be discussed in detail in the next section. The square gradient theory, which was first introduced by Rayleigh38 and van der Waals39 and later rediscovered by Cahn and Hilliard,40 was applied to study wetting problems.41, 42 In square gradient theory, the fluid-solid relative IFE is derived from the surface excess transverse stress, where principal stress profiles are computed based on density distribution across the interface.42, 43 Meanwhile, classical density functional theory (cDFT) has become a pivotal tool for analyzing wettability phenomena at the molecular level.44, 32, 45 Within the cDFT framework, the fluid-solid relative IFE is typically calculated using excess grand potential: , where , , , and are the grand potential, bulk pressure, volume of the fluid, and interfacial area, respectively. The cDFT is a robust and versatile tool grounded in thermodynamic principles, offering exceptional accuracy and efficiency.46, 47 The cDFT has been used for computing the fluid-solid IFE for a broad range of systems including confined fluid,48 chemically patterned wall,49 and heterogeneous surface,50 while the application of other first-principle theories has been relatively rare nowadays. Despite the demonstrated precision and efficacy of cDFT, it remains underutilized across experimental, theoretical, and computational communities. Major obstacles to its wider adoption include the theoretical complexity and a lack of accessible, user-friendly software for fluid-solid IFE computations.51, 52, 53
Molecular simulation is a powerful tool for investigating fluid-solid IFE due to several distinct advantages. Firstly, there is a wealth of open-source software available, offering a wide array of simulation tools to choose from.54, 55, 56 Moreover, molecular simulations are based on precise atomic-level representations, offering a robust physical foundation for IFE calculations.57, 58 Furthermore, molecular simulations provide flexibility, enabling the study of diverse fluid-solid interfaces across a wide range of chemical systems. This versatility arises from the extensive library of force field parameters available in the literature.59, 60, 61, 62, 63, 64 Although certain acceleration techniques, such as coarse-grained modeling,65, 66, 67, 68, 69 can reduce computational time, molecular simulation methods remain more computationally intensive than semi-empirical and first-principle approaches.
Despite the advancement of various molecular simulation methods for estimating fluid-solid IFE, there remains a scarcity of comprehensive reviews. Jiang and Patel70 published a review paper focusing on molecular simulation techniques for estimating contact angles. Those methods are useful for calculating the differences of IFE. Nevertheless, several essential methodologies for the direct estimation of fluid-solid IFE remain uncovered in the previous review, and numerous innovative techniques have emerged over the past few years. This review aims to bridge these gaps by providing a comprehensive analysis of state-of-the-art molecular simulation methodologies for calculating fluid-solid IFE. The molecular simulation methods for estimating IFE can be broadly classified into two primary categories: mechanical and thermodynamic approaches. The mechanical approach encompasses methods such as the contact angle approach, the method using Bakker’s equation, and the Wilhelmy simulation method, each of which will be discussed in detail. For the thermodynamic approach, key methods include the cleaving wall technique, the Frenkel-Ladd technique, and the test-volume/area methods.
Note that while there are various experimental71, 72, 73, 74, 75, 76 and theoretical77, 78, 79 methods available for studying fluid-solid IFE in the context of crystallization/nucleation, the focus of this review is on wetting problems and methods that are not suitable for studying common wetting problems were not included in the above discussion. Therefore, this review will not cover molecular simulation techniques such as the capillary fluctuation technique,80 umbrella sampling,81 extrapolation method,82, 83 metadynamics,84, 85, 86, 87 the superheating and undercooling method,88, 89, 90 the seeding technique,91, 92, 93 tethered Monte Carlo,94 and the mold integration method95.
2 Mechanical Routes
The mechanical route is based on the direct computation of forces at the interface within molecular simulations. This route includes the contact angle approach,96 the technique based on Bakker’s equation,37 and the Wilhelmy simulation method.97
2.1 Contact Angle Approach
While the contact angle approach does not provide direct access to the absolute values of fluid-solid IFE , the method is still noteworthy considering its capability to estimate the differences of fluid-solid IFE ( in Eq. 1). This property, termed “superficial tension” by Gibbs98, 99 (sometimes also noted as “adhesion tension”100), describes the difference in fluid-solid IFE across different phases. Notably, when the fluid phase 2 is vapor with low density and the interfacial adsorption of the vapor is little, can be used to approximate the vacuum-solid IFE. In this case, the relative IFE between fluid phase 2 and the solid ( defined in the next section) is close to zero, and the relative IFE between fluid phase 1 and the solid can be estimated as the negative of the superficial tension.101, 102 The method involves calculation of fluid-fluid IFE and contact angle in separate simulations, and superficial tension is calculated as the product of and , based on Eq. 1.103, 34
The sketches of molecular system setups for estimating and are shown in Fig. 1a and b, respectively. To calculate fluid-fluid IFE in molecular simulation, the simulation box usually contains slabs of bulk liquid phases to form interfaces (see Fig. 1a).104, 105 In an orthogonal system with z-direction normal to the flat fluid interface, the principle components of the pressure tensor (, , and ) are used to calculate IFE according to Bakker’s equation:106, 107
(2) |
To estimate the contact angle, the molecular system containing a sessile droplet on a solid substrate is simulated, and the contact angle can be determined from the average shape of the droplet (see Fig. 1b).96 It is important to note that the direct simulation of such a three-phase system may be subject to errors including finite size effects108, 109 and hysteresis.110, 111 To circumvent these issues, a large cylindrical shaped droplet is suggested109 and enhanced sampling method112 can be combined. Jiang and Patel70 summarized those issues and provided a comprehensive review on methods for contact angle calculation in molecular simulations. Additionally, the contact angle values predicted using molecular simulations are dependent on the choice of the force field. For example, using two different force fields, the water contact angle on calcite is predicted to be 0° and about 38° at 323 K and 20 MPa.109, 113 It also is worth mentioning that a freely available tool named “ContactAngleCalculator” has been recently developed to rapidly and accurately estimate contact angles from molecular simulations.114
The molecular simulation data on or of different systems is extensive.104, 115, 116, 117, 118, 119 Although superficial tension can be estimated based on those data, direct analysis of its behaviors remains relatively scarce. Few authors have investigated superficial tensions in water/rock, water/gas/rock, water/oil/rock, and water/gas/oil/rock systems. 120, 121, 122, 103, 123, 103, 124, 125 Such systems are important for applications including gas storage and enhanced gas/oil recovery.126, 127, 128, 129 Remarkably, superficial tension appears as the denominator of the capillary number :
(3) |
where and are the viscosity and characteristic velocity, respectively. The capillary number, representing the ratio of viscous to capillary force, is crucial in enhanced gas/oil recovery.130, 131 Note that superficial tension also appears in the Young-Laplace equation for the capillary pressure. Many authors have reported the values of capillary pressure from molecular simulations.132, 133, 134, 135, 120
Here, we provide a summary of the reported superficial tensions in water/gas/silica systems,120, 122, 121, 123, 103 as depicted in Fig. 2. This serves as an illustrative example of the behavior of superficial tension under varying conditions such as temperature, pressure, gas type, surface wettability, and salinity. In general, an increase in pressure results in a decrease in superficial tension, while higher temperatures weaken the magnitude of superficial tension. Notably, the reduction of superficial tension induced by high pressure is significantly more pronounced in systems with CO2 compared to those with N2 or H2 (see Fig. 2a). This phenomenon can be ascribed to the strong interactions of CO2-H2O and CO2-silica pairs. Interfacial density distribution results indicate that CO2 exhibits greater adsorption in the gas-H2O and gas-silica interfaces compared to N2 or H2. This results in a more pronounced reduction of and .120, 121, 122 Adjusting the density of surface silanol groups can modify silica’s wettability from hydrophilic to hydrophobic.136, 119 Experimental modifications to the silica surface, such as the ionization of surface silanol groups, are achievable, particularly through controlled temperature adjustments.137 Superficial tensions are positive in systems with hydrophilic silica, while negative superficial tensions have been reported in systems with hydrophobic silica surfaces due to the change in the sign of (see Fig. 2b). Additionally, an increase in salinity leads to a reduction in superficial tension (see Fig. 2c). This reduction comes mainly from the decrease in , given that increases with salinity.
2.2 Technique using Bakker’s Equation
While Bakker’s original equation (Eq. 2) was designed specifically for fluid-fluid interfaces, it has been adapted from fluid-fluid to fluid-solid interfaces to quantify surface stress in systems with flexible solid substrates using the same methods for the fluid-fluid interface (see section 2.1).138, 139, 140, 141, 142 Note that the calculated in this direct approach can be separated into two terms according to the Shuttleworth equation:143
(4) |
Here, the fluid-solid IFE is represented by the first term, while the second term quantifies strain-related energy contributions from the solid substrate.144 The second term separates the IFE of solid interfaces from that of liquid interfaces and can be estimated by a numerical derivative using IFE data under slightly varied strains.139 Note that “surface tension” could be used synonymically with “surface stress”, although a subtle difference exists.139
In the rest of this section, we present the technique using Bakker’s equation for estimating the fluid-solid relative IFE:145
(5) |
where is the IFE of the solid substrate in the absence of any fluid. Note that both superficial tension and relative IFE represent differences in interfacial energies. In this work, relative IFE is considered as a specific case of negative superficial tension. This distinction is based on the reference state: relative IFE is defined with respect to the vacuum-solid interface, whereas superficial tension may be defined with either vacuum-solid or fluid-solid states. The technique using Bakker’s equation usually deals with molecular systems containing rigid solids or solids that can be treated as external potential. Flexible solids can in principle be handled. Note that cannot be determined within this method. While is not directly computable in this method, it can be estimated using complementary techniques including the Frenkel-Ladd techniques,146, 147, 148, 149 the cleaving wall methods,150, 151, 152, 23 -integration techniques,153, 154, 155 and approach based on the generalized Gibbs adsorption equation.138 The fluid-solid relative IFE can be used to calculate wettability and adhesion work since could be canceled out when subtracting from .156, 34
Note that for an interface in equilibrium, the value of IFE is always positive according to the principles of thermodynamic stability.157, 158, 159 However, negative values of relative IFE are frequently reported.160, 34 Negative does not violate the thermodynamic stability as the is not included. The sign of indicates the direction of tension. Positive arises from contracting tension to reduce the contact area (similar to fluid-fluid interfaces), while negative comes from the spreading tension to increase the contact area. This can be understood from mechanical balance across phase boundaries in the vacuum/fluid/solid systems by assuming the existence of a vacuum/solid interface ( the relative IFE of the vacuum/solid interface is 0 mN/m). If the contact angle is greater than 90, has to be contracting force (positive) to balance the fluid-vacuum IFE in the tangential direction of the surface (see Fig. 3a). While spreading force (negative ) is present when the contact angle is less than 90 (see Fig. 3b).
Navascués and Berry35 extended the statistical mechanical theory of Kirkwood and Buff36 to the case where fluids are in contact with a rigid solid phase. Within this theory, the fluid-solid IFE can be understood through a process that combines a vacuum-solid interface and a vacuum-liquid interface into a liquid-solid interface. The fluid-solid IFE was split into one solid and two fluid-solid contributions:35
(6) |
Here is the IFE of the fluid phase when the solid phase is removed without relaxing the structure of , which can be calculated using Eq. 2. And denotes the free energy change, interpretable as the work required to separate and without relaxing :35
(7) |
where and are the force exerted on fluid by solid in the z-direction ( direction normal to the surface) and number density, respectively. The origin point is defined at the furthest location where the fluid density is zero from the solid surface and is the point in the bulk region of . Then the expression for is as follows:
(8) |
The physical meaning of each term in Eq. 6 is shown in Fig. 4a and similar figures were given in Refs.35 and 102. The theory of Navascués and Berry35 was later combined with molecular dynamics simulation for estimating .102, 161, 34, 162
Nijmeijer and Leeuwen163 also derived an expression for the fluid-solid relative IFE based on Bakker’s equation:
(9) |
where and are the principal components of the pressure tensor of the fluid considering the solid substrate as external potential.
This equation can be understood by a thought experiment designed by Yamaguchi et al.145 as shown in Fig. 4b. This thought experiment is an extension of the thought experiment for Bakker’s description of the relationship between the IFE and the fluid stress anisotropy.106, 107 In this thought experiment, one piston is positioned perpendicular to the fluid-solid interface, covering the area where the fluid is present. This ranges from (where the fluid density is 0 g/cm3) to (a location within the bulk). Another piston is placed parallel to the fluid-solid interface, far from the interface, to regulate the bulk pressure . By making simultaneous, infinitesimal virtual displacements of the pistons, we can alter only the interface area without affecting the fluid volumes. Let’s denote the depth normal to the -plane as . If represents the infinitesimal volume change caused by the downward displacement of the top piston, and represents the corresponding displacement of the side piston, then we can deduce the following:145
(10) |
Assuming that the displacement occurs in a quasi-static manner under a constant temperature, the minimum mechanical work, denoted as , needed for this change is linked to the alteration in the Helmholtz energy , as expressed by the following equation:145
(11) |
The outcome of this change is that the solid-vacuum interface is supplanted by the solid-fluid interface. Consequently, Eq. 9 can be reestablished by taking the partial derivative of the Helmholtz energy with respect to the fluid-solid area , where :145
(12) |
The Eq. 9 was implemented in molecular dynamics simulations for computing .37, 145, 164 Importantly, the expression of from the theory of Navascués and Berry35 and the derivation of Nijmeijer and Leeuwen163 are identical. The in Eq. 9 is the same as the pressure component tangential to the interface in , and the in Eq. 9 is the summation of the negative of the integrand of and the pressure component normal to the interface in . However, the agreement was rarely noted.165, 166 For instance, the work of Nijmeijer and Leeuwen163 was overlooked by several authors.102, 161, 34, 162
We now outline the computation details required for estimating in molecular simulations. The setup of the simulation box resembles the top figure in Fig. 4a. An atomistic piston could be implemented to regulate the bulk pressure.145, 34 The number density is readily accessible in simulations and can be calculated on the fly or in post-processing. Once converges, the locations of the origin point and can be identified. Without modifying the simulation code, the is calculated by rerunning the saved trajectories with fluid self-interactions disabled. With the above properties calculated, can be obtained by numerical integration using Eq. 7.
The can be calculated using Eq. 2 by rerunning the saved trajectories with solid substrate deleted from the simulation box. This task usually involves the calculation of the microscopic profile of the pressure tensor. The LAMMPS package offers built-in features to determine local pressure tensors.54 There are several other analysis tools for this task including the GROMACS-LS code prepared by Vanegas et al.,167 the MDStressLab tool developed by Admal et al.,168 and the LAMMPS patch developed by Nakamura et al.169 It is important to note that the definition of microscopic pressure tensor remains under debate, particularly in terms of its contour formalism.170 Different definitions were used in previous studies for estimating fluid-solid IFE.171, 145, 164, 34 However, further investigation is required to determine whether different methods yield consistent . Another issue is related to the challenges for systems with many-body and long-range potentials for the computation of local pressure tensors.169 Additionally, the calculation of pressure tensor is a computationally intensive task given that it involves a second-order property with a slow convergence rate. The readers are referred to the perspective paper by Shi et al.170 for more information regarding the calculation of microscopic pressure tensors using molecular simulations.
Bakker’s equation has been widely utilized to estimate the relative IFE at fluid-solid interfaces in many studies.163, 172, 171, 173, 174, 145, 164, 175, 34 For example, Yang et al.34 applied the method to study the relative IFE in water/gas/solid three-phase systems at various temperatures, pressures, and wettabilities. Note that the water wettability on a solid surface depends on the water-solid interaction energy.176, 177 A binary interaction parameter was introduced to adjust the interaction energy between water particle and solid particle: , where is the potential well depth obtained from the mixing rule.178
The influence of on interfacial energies and wettabilities in the water (vapor/liquid)/solid system is depicted in Fig. 5a-c. It is observed that the relative IFE of the liquid H2O-solid interface escalates from -102.0 to 65.0 mN/m, as varies from 0.0 to 0.9. Hydrophilic surfaces correspond to negative , while hydrophobic surfaces correspond to positive (c.f. Fig. 5a and c). Simultaneously, the magnitude of the relative IFE of the vapor H2O-solid interface is typically less than that of . ascends with and rapidly converges to 0 mN/m. Furthermore, wettability can be determined based on and using Young’s equation. The wettabilities derived from Young’s equation using relative IFE values from Bakker’s equation align well with those obtained from the contact angle approach, as illustrated in Fig. 5c.
The relative IFEs and wettabilities in the water/CO2/solid system are depicted in Fig. 5d-f. Three distinct values of for the H2O-solid pair are selected to exemplify hydrophilic (), neutral (), and hydrophobic () surfaces. The impact of temperature and pressure on the relative IFE of the water-solid interface is complex (see Fig. 5d). Generally, the influence of pressure on is less noticeable at 298 K. However, of water+CO2+hydrophobic solid (, ) system initially decreases and then increases as pressure varies from approximately 5 to 100 MPa. Conversely, at a higher temperature of 403 K, diminishes with increasing pressure, and the effect of pressure is more pronounced with more hydrophobic solids. In addition, the relative IFE for the CO2-solid interface is negative (see Fig. 5e). The magnitude of intensifies as pressure increases and the solid surface becomes more hydrophilic. High temperatures generally reduce the magnitude of . Furthermore, the contact angles enlarge as pressure rises and temperature decreases for all three surface types (see Fig. 5e).
These simulation results provide a basis to evaluate the assumptions in semi-empirical theories like the Neumann equation of state, which is commonly used to infer fluid-solid IFEs from fluid-fluid IFE and wettability data.25 In H2O/gas/solid systems, the Neumann equation of state is typically combined with the presumption that the H2O-solid IFE remains unchanged across different pressure levels.179, 180, 181, 182 However, simulation computed values in water/gas/solid systems suggest that this presumption may not hold true in cases that feature non-hydrophilic surfaces with strong gas-solid interactions due to the adsorption of gases in the water-solid interface.34
2.3 Wilhelmy Simulation Method
Based on the experimental Wilhelmy method183 for measuring the interfacial tension of fluid-fluid interface, Imaizumi et al.97 developed the Wilhelmy simulation method for estimating interfacial properties in fluid-fluid-solid three-phase systems using a single molecular dynamics simulation.
As shown in Fig. 6a, the molecular arrangement simulates the experimental configuration, wherein a quasi-2D meniscus emerges upon a hollow rectangular solid plate, submerged partially in a liquid reservoir. The simulation system applies periodic boundary conditions in x- and y-directions. In equilibrium, the system can be approximated as homogeneous in the y-direction (not shown in Fig. 6a). Two repulsive potential walls were fixed at the top and bottom of the simulation box making the boundary non-periodic in the z-direction. A graphene sheet was bent into a rectangular shape to form the solid plate, which was then positioned in contact with the fluid, and the solid particles were anchored at specific coordinates on this structure. The right and left surfaces of the solid plate were aligned parallel to the yz-plane, while the upper and lower surfaces were oriented parallel to the xy-plane.
Several properties are measured in such molecular systems at an equilibrium state. Most of them are marked in Fig. 6b. Those properties include: Bulk pressures in the liquid and vapor phases ( and ); Contact angle of the meniscus ; Downward forces , , and experienced by the top, middle (contact line), and bottom parts of the solid plate from the fluid, respectively; Solid-fluid potential energy densities with unit of energy per area and at SV and SL interfacial regions, separately. and are derived by dividing averaged forces on the top/bottom potential walls by surface areas. can be calculated based on equilibrium density distributions (see section 2.1). () can be computed by summarizing the pair-wise force in z-direction (potential energy) between solid particles per unit area and fluid particles within the cutoff distance. Fig. 6c show typical distributions of and , which can be used to extract , , , and as noted in the figure.
The interfacial properties of interest (including , and the pinning force at three-phase contact line ) can be calculated based on the measured properties using the following relations:97
(13) |
(14) |
(15) |
(16) |
where is the position of the solid-fluid interface ( the nearest location where the fluid can reach near the solid), which can be determined from the density profile. is half the size of the simulation box in the x-direction (see Fig. 6b).
Here, we briefly introduce the relations (Eq. 13-16) between interested interfacial properties and the measured properties (full derivations are available in Ref.97). The Eq. 13 is the Young-Laplace equation.157 As given in APPENDIX A of the Ref.97, the Eq. 14 is derived by a mean-field approach through the analysis of forces around the contact line ( the blue-dotted control volume (CV) shown in Fig. 6b). In a similar method, the following relation can be derived for the red-dotted CV:97
(17) |
where is the upward forces experienced by the fluid in the red-dotted CV in Fig. 6b from the solid. In equilibrium, the force balance on the red-dotted CV can be described by:97
(18) |
where is the principle component of 2D fluid stress tensor (averaged in the y-direction) in the z-direction, and the locations of and are marked in Fig. 6b. The fluid bulk pressure is related to the fluid stress tensor and the first term in the left-hand side of Eq. 18 can be expressed as:97
(19) |
Applying Bakker’s equation for the SL relative interfacial tension ( Eq. 9 with shifted origin), we can express the second term in the left-hand side of Eq. 18 as:97
(20) |
Substituting Eq. 17, 19 and 20 into Eq. 18, Eq. 15 is recovered. By doing the same procedures for the brown-dotted CV, Eq. 16 can be derived.
To optimize simulation accuracy with the Wilhelmy simulation method, the dimensions of the simulation box and the solid plate need to be carefully calibrated. Firstly, the contact angle was calculated using the same one in the Contact Angle Approach in section 2.1. Therefore, a large size of box length in the x-direction should be used so that the radius of the meniscus is large enough to mitigate the system size effects.70 Note that the method may also suffer from the hysteresis issue discussed above.110, 111 Meanwhile, it is important to ensure that the size of the simulation box in the z-direction is sufficiently large. This allows for the existence of a bulk liquid phase between the solid plate and the potential wall at the bottom. Similarly, it also accommodates the presence of a bulk vapor phase between the solid plate and the potential wall at the top. Additionally, the width of the solid plate should be large enough ( larger than the cutoff distance) to prevent interactions between fluids adjacent to the edges of the plate. Similarly, the length of the plate should be large enough to ensure that the bulk regions of both the liquid and vapor phases are present at the simulation box in the x-directional boundary. This is related to the position of the three-phase contact line which is better positioned near the center of the plate in equilibrium. It is also important to note that the derivation assumes that particle interaction potentials are truncated using a cutoff distance. However, studies show that long-range pairwise interactions significantly impact fluid-fluid IFEs.184 Therefore, a large cutoff distance is recommended, necessitating a solid plate with a large width. Consequently, these constraints mandate a large simulation system, entailing considerable computational resources.
Using the Wilhelmy simulation method, one can obtain the fluid-solid relative IFEs from just one equilibrium molecular simulation, thereby avoiding the complex and often disputed process of calculating local pressure profiles.170 An additional important advantage of this method is the ability to compute the pinning force. This is particularly useful for contact lines that experience pinning as a result of irregularities, such as surface roughness, impurities, or deformations of the surface.145 In the example illustrated above, the solid plate was made using graphene, and the derivations employ a mean-field approach, assuming zero-thickness for the solid plate. However, the approach can be extended to other solid materials with non-zero thickness. Note that the material for the top and bottom surfaces of the solid plate can be different from that for the left and right surfaces. For example, potential walls could be used for the top/bottom surfaces to reduce the computation cost.
Remarkably, the method has been extended to deal with cylindrical-shaped solid plates to understand the effects of curvature on fluid-solid interfaces.185 The interfacial characteristics of a Lennard-Jones fluid around solid cylinders with different radii are displayed in Fig. 7. Fig. 7a displays the density distributions surrounding the plate and the solid cylinder with the smallest radius, = 0.777 nm. It is intriguing to note that the cylinder and plate appear to have substantially distinct meniscus forms. This implies that in order to accurately estimate the wettabilities from the meniscus shape, different force balances should be used.185 The wettability was determined by the fluid-solid interaction parameter . is implemented in the pair-wise potential energy between the fluid particle and the solid particle: , where is the potential well depth obtained from the mixing rule.185 The contact angle derived from the meniscus forms decreases as increases (see Figs. 7a and b). Furthermore, a decreased radius is observed to result in a slight rise in the contact angle. The relative IFEs of the liquid-solid and vapor-solid interfaces are displayed in Fig. 7c. The effects of fluid-solid interaction on relative IFE are similar to those in water/solid system34 discussed above (c.f. Figs. 5a,b and Fig. 7c). It is important to note that reducing the radius leads to an increase in the relative IFE values. In order to validate the calculated relative IFEs, those data were further substituted into Young’s equation to estimate the contact angle . As seen in Fig. 7d, this contact angle matches perfectly with obtained from fitting the meniscus shape.185
A potential extension of the Wilhelmy method includes handling the solid in a spherical shape. On the other hand, advancing this method to incorporate more complex systems represents a pivotal aim for forthcoming research. For example, the method could be applied to the system in the presence of electrostatic interactions within solid and fluid molecules.
3 Thermodynamic Routes
The thermodynamic route refers to a series of methods based on thermodynamic principles to calculate IFE in fluid-solid systems. These methods involve monitoring changes in free energy with respect to interface area in molecular simulations. In the review paper of Jiang and Patel70, several methods of the thermodynamic route that can be used to calculate the superficial tension and the fluid-solid relative IFE are discussed in detail. Here, we briefly discuss those methods with recent advances for completeness. Errington and co-workers186, 187 developed the interface potential methods. By systematically changing the chemical potential of the fluid in a grand canonical Monte Carlo simulation, a fluid-solid interface is gradually changed from a liquid-solid to a vapor-solid interface during a drying process or in a reverse direction during a wetting process. A sketch of the method is given in Fig. 8a. This method calculates superficial tension based on the thickness of the vapor/liquid film, using the grand canonical transition matrix Monte Carlo algorithms.188, 189, 190 From this, the fluid-solid relative IFE can also be derived. Recent developments of this method involve the implementations within an isothermal-isobaric ensemble using Monte Carlo simulation methods191, 192 and canonical molecular dynamics simulations.193
The relative IFEs of the fluid-solid interface could also be estimated through a thermodynamic integration (TI) scheme using molecular dynamics simulations. Leroy et al.194 developed the phantom-wall method. In this method, a repulsive wall with interactions only with the fluid (by turning off the wall-solid interactions) is reversibly shifted away from the surface. This converts the fluid-solid interface into a vacuum-solid interface and a repulsive wall-fluid interface. The sketch of such a process has been presented in Fig. 8b. The relative IFE could be calculated by subtracting the relative IFE of the repulsive wall-fluid interface from the free energy change of the separation process. It is important to note that the relative IFE of the repulsive wall-fluid interface is typically approximated by the fluid interfacial tension. This approximation has been demonstrated to be sufficiently accurate in scenarios involving water near purely repulsive walls at room temperature.195, 196 However, the error of this approximation could be significant if the effects of the vapor phase on the interfacial tension cannot be disregarded, such as in the case of water under high temperatures.162 In such instances, the technique utilizing Bakker’s equation may be a more suitable choice for estimating the relative IFE. Recently, Uranagase et al.197 developed a scheme for calculating the work of adhesion between a liquid and complex surface, and the scheme could also be used to estimate the fluid-solid relative IFE. Such a scheme has been implemented in a freely available code named “FE-CLIP” by Uranagase and Ogata.198
The dry-surface method, developed by Leroy and Müller-Plathe, utilizes a procedure in which the attractions between the surface and the fluid are progressively turned off (see sketch shown in Fig. 8c).162 Recently, Surblys et al.199 addressed the challenge of long-range Coulombic interactions at interfaces in the dry-surface method by substituting them with damped Coulomb interactions and investigating various thermal integration paths. It is worth mentioning that the dry-surface method has been used as the reference method for validating the Wilhelmy simulation method for surfaces with various wettabilities.97 However, similar to the potential issue faced in the phantom-wall method, estimating the relative IFE within the dry-surface method also operates under the assumption that the relative IFE of the fluid-repulsive wall interface can be approximated as the fluid interfacial tension. Additionally, Kanduč and Netz200, 201 have introduced a TI scheme that is anticipated to be more effective in analyzing the wetting properties of highly hydrophilic surfaces.
In the next sections, we discuss other important methods of the thermodynamic route for estimating fluid-solid IFE, which are classified into two groups, namely, TI methods and free energy perturbation (FEP) methods.
3.1 Thermodynamic Integration Methods
As noted above, TI methods in molecular simulations are used to calculate free energy differences between two states by systematically integrating energy changes during state transitions.57 The literature on TI methods developed for computing the fluid-solid IFE is numerous. In addition to the methods mentioned above, here, we list several other methods including the Gibbs-Cahn TI techniques,202, 203, 204 the ensemble mixing/switch methods,205, 206, 207, 208, 209 the Schilling-Schmid method,210, 211 the method using the Gibbs adsorption equation,212 the method based on the excess grand potential of confined fluids213, and the method of Das and Binder for superficial tensions.214, 215 In the following parts, we focus on two typical TI methods for computing fluid-solid IFEs, namely, the cleaving wall method and the Frenkel-Ladd technique.
3.1.1 Cleaving Wall Method
In this section, we begin with a concise overview of developments in cleaving wall methods. We provide a detailed discussion of two distinct versions of these methods: one for calculating the fluid-solid IFE, , and the other for the relative IFE, . Finally, a practical application example will be presented. Note that the cleaving wall method is also capable of calculating the vacuum-solid IFE directly using a similar procedure.151, 152
The cleaving wall method employs a reversible process to separate a fluid-solid system into distinct bulk fluid and solid phases using external potential walls. The methods share common features with the phantom-wall method mentioned above, and the latter can be considered as a special case of the former. Eriksson216 introduced the concept of “cleaving”, which has since evolved,217, 218 despite early debates about the reversibility of the process.219, 220 The rigorous definition of the concept has been built based on statistical mechanics.221 The use of reversible cleaving in molecular simulations began with Miyazaki et al.222 Then, Broughton and Gilmer developed the cleaving wall method in molecular dynamics for creating an interface in a liquid-solid system.151 Their approach employed uniquely designed “cleaving potentials” to divide bulk liquid and solid phases into two segments separated by a cleaving plane, followed by merging these segments and ultimately eliminating the “cleaving potential”. However, the method, which involves creating specific cleaving-potential functions for each crystal facet, faces challenges due to the high uncertainty from hysteresis and the complexity of designing these potentials for various orientations and conditions. The approach was further developed by Davidchack and Laird.223, 224, 225, 226, 227 To resolve the anisotropy in IFE, they adopted flat cleaving potentials constructed from particles similar to those in the system, comprising several crystalline layers that mirror the structure of the actual crystal phase. However, eliminating the hysteresis caused by the movement of the crystal-liquid interface appears to be a challenging issue to fully resolve. Nevertheless, attempts have been made to address this problem by carrying out several independent TI runs and choosing the one that demonstrated the least amount of hysteresis.224, 227
Benjamin and Horbach228 proposed a version of the cleaving wall method to address the hysteresis issue by utilizing a planar wall with an extremely short interaction range (modeled using a Gaussian potential) to distinctly separate the liquid and solid phases. Here we briefly illustrate the TI path for the calculation of the fluid-solid IFE within this method. The TI scheme starts with two separate simulations: one of a bulk crystal and the other of a bulk liquid, both conducted under coexistence conditions and ends with a combined crystal-liquid two-phase system containing two fluid-solid interfaces.
The TI scheme, which includes 6 steps as depicted in Fig. 9a, should be carried out progressively through each step. Step 1: A flat potential wall, with a short interaction range modeled by a Gaussian function, is placed at the end of the simulation box along the z-axis. Step 2: Repeat step 1 for the simulation box of the bulk crystal. Step 3: Create two solid walls (shown as A and B inside grey dotted boxes in sketch (2)). The solid walls consist of 2 to 3 layers of crystalline particles adjacent to the Gaussian wall that has been inserted in step 2. Subsequently, these walls are connected to the correct ends of the liquid simulation box from step 1. Meanwhile, the periodic boundary conditions in the z-direction of the liquid simulation cell are deactivated. Step 4: Repeat step 3 for the crystal simulation box from step 2. Step 5: Merge the liquid and solid systems obtained from Steps 3 and 4. Activate the interactions between two phases at the contact, and also across the box boundary by enabling the periodic boundary condition in the z-direction. Step 6: Remove the very short-ranged Gaussian walls.
It is crucial to mark that the Gaussian wall serves solely to stop the liquid and crystalline particles from passing the border without altering the bulk properties. The usage of the very short-range Gaussian wall necessitates a very small timestep in molecular dynamics simulations. However, this difficulty can be effectively addressed by employing a multiple time step algorithm.57 Moreover, the incorporation of structured solid walls in steps 3 and 4 aims to minimize the disruption of the crystal during the merge of the two phases. This is vital to ensure that the liquid organizes into layers in the interfacial region, aligning with the interactions from the crystal. Remarkably, as the barrier of the Gaussian wall weakens in step 6, particles gain the ability to traverse the boundaries, potentially causing the interfaces to shift and resulting in hysteresis along the TI path. Nevertheless, the approach effectively addresses this hysteresis due to the minimal impact of the final step on fluid-solid IFE as a result of the extremely short-ranged flat walls.
The above-mentioned cleaving wall methods involve manipulating the periodic boundary conditions ( steps 3-5 in Fig. 9a). Such a function requires specialized implementation and is not available in most of the open-source codes. Moreover, analytical differentiation of the potential energy can be challenging to calculate. To generalize the application of the Benjamin and Horbach approach, Qi et al.229 proposed a multi-scheme TI method combined with numerical approximation of thermodynamic integrands to circumvent these problems. The scheme has been utilized for determining the IFEs of the Ag-ethylene glycol interface and holds potential for application across a wide variety of systems of interest. It is noteworthy that a LAMMPS package “CLEAVING” has been recently released for the calculation of IFE of solid-fluid and solid-solid interfaces using the cleaving wall method.230 This package enables the reproduction of results from Refs. 23,139, and 224.
Building on prior methods, Addula and Punnathanam231 developed a cleaving wall technique that exclusively cleaves the fluid phase. This method integrates the Frenkel-Ladd approach (outlined in section 3.1.2) to calculate the vacuum-solid IFE. The overall IFE can be estimated by combining these two components. As shown in Fig. 9b, this version of the cleaving wall method includes 3 steps. In the first step, the bulk liquid is split by applying a cleaving potential to create a vacuum region. The initial and final snapshots of the first step are shown in Fig. 9b(1i) and (1f), respectively. Note that a cleaving potential is composed of atoms that exclusively interact with the fluid, incorporating both repulsive and attractive terms. In the second step, the solid slab is placed into the created region (see Fig. 9b(2)). In the final step, the cleaving potential is withdrawn, leading to the development of two solid-liquid interfaces (see Fig. 9b(3)). All procedures are performed within the NPzT ensemble.232 The free energy changes in step 1 () and step 3 () are determined through thermodynamic integration. For step 2, the free energy change () is estimated using the thermodynamic perturbation method.
The method has an advantage over the phantom wall method in that it eliminates the need to estimate the fluid-wall IFE. Nevertheless, it is more complex, as it requires three thermodynamic integration steps, whereas the phantom wall method necessitates only one. Additionally, it is important to note that a strong cleaving potential can induce phase transitions, leading to an overly organized fluid structure near the wall.233, 231 Therefore, calibrating the interaction strength of the cleaving potential is crucial to maintaining a reversible path and preventing such phase transitions that violate the reversibility condition.
The cleaving wall method of Addula and Punnathanam231 has been applied to investigate the wettabilities of water/oil/rock systems.234 Understanding the wettability of water/oil/rock systems is crucial for optimizing enhanced oil recovery processes.235, 236 Fig. 10a and b present relative IFEs and their contributions from each TI step of the water/solid and water/oil interfaces, respectively. Three silica surfaces with different hydrophilic silanol densities are studied, namely, Q2 with Si-OH density of 9.4 nm2 ,Q3 with Si-OH density of 4.7 nm2, and Q4 with Si-OH density of 0 nm2 (see snapshots in Fig. 10a).136 Notably, the free energy change from the insertion of the solid slab (, step 2) is little in contrast to those in other steps. This is attributed to the strong attraction between the cleaving wall atoms and the fluid, which preserves the structure of the fluid phase during insertion.234
As the silanol density decreases (, from Q2 to Q4), the relative IFEs change from negative to positive, which is consistent with the results given in Fig. 5 and Fig. 7. The magnitudes of relative IFEs of water-silica interfaces are much larger than those of heptane-silica interfaces. For the water-silica interface, both Q2 and Q3 are water-wet surfaces, while Q4 is hydrophobic in water/vacuum/silica systems by comparing Fig. 10a and Fig. 3. For the heptane-silica interface, only Q2 surface is oil-wet, while oil contact angles in the water/vacuum/silica (Q3 and Q4) systems are greater than 90 by comparing Fig. 10b and Fig. 3. Since the miscibility of water and oil is low,237 those relative IFEs can be substituted into Young’s equation to estimate the wettability. The water contact angles are 0, 0, and 128 for Q2, Q3, and Q4 surfaces, respectively.234 Those values are in excellent agreement with results obtained from the contact angle approach.234 Furthermore, the directions of interfacial energies during the wetting and dewetting processes are illustrated in Figs. 10c-e. Notably, the relative IFEs for both water-solid and oil-solid interfaces align in the same direction on the Q3 surface, promoting the spreading of the water droplet.
3.1.2 Frenkel-Ladd Technique
The Frenkel-Ladd technique, introduced in 1984, provides a method for calculating the absolute Helmholtz free energy of periodic crystals.146 The method establishes a reversible thermodynamic pathway connecting the solid with an Einstein crystal, the free energy of which can be analytically determined. Note that in an Einstein crystal, each atom is confined within its own harmonic potential well, which isolates them from interacting with one another.238 In other words, the potential energy is solely determined by the positions of the atoms within their respective harmonic potential wells, and the potential energy of the system is independent of the relative positions of the atoms. The developments and applications of the Frenkel-Ladd technique for free energies of bulk solids have been summarized in several review papers.239, 240, 241
Pretti and Mittal242 extended the Frenkel-Ladd technique to non-periodic and semi-periodic systems. This advancement enables the determination of absolute free energies in finite-sized crystals characterized by distinct shapes and surface structures in contact with a vacuum environment ( ). Addula and Punnathanam231 proposed two TI methods for computing the fluid-solid IFE . The first TI method is based on the cleaving wall method discussed in section 3.1.1 (also see Fig. 9b). The second approach, known as the “adsorption method,” involves the integration of an adsorption isotherm. Both of these methods calculate the fluid-solid relative IFE . And the solid-vacuum IFE ( ) is handled by a modified Frenkel-Ladd technique.149 The fluid-solid IFE can then be derived by combining and (see Eq. 5).
Recently, Yeandel et al.243 presented a similar but more general method for computing . The method also uses the Einstein crystal as the reference state. The following describes the four-step process in the TI path, illustrated in Fig 11:
Step (1) transforms the bulk solid into an Einstein crystal through two stages: 1. Turn on the harmonic wells for every atom; 2. Turn off all interactions both between and within molecules. The corresponding free energy change is denoted as .
Step (2) creates a vacuum space inside a liquid film. The free energy change of this step is simply , here is the vacuum-liquid IFE. This term can be approximated as the vapor-liquid IFE and calculated using the relatively efficient method based on Bakker’s equation (see Section 2.1).
Step (3) inserts the Einstein crystal from step (1) into the middle vacuum space of the split liquid film from step (2). Since there are no interactions between the Einstein crystal and the liquid, the free energy change in this step is zero.
Step (4) turns the liquid-Einstein crystal system from step (3) into the liquid-slab solid system. The corresponding free energy change is denoted as . is calculated reversely noting . Two TI stages are used starting from the liquid-slab solid system: 1. Turn on the harmonic wells for every atom in the solid slab; 2. Turn off all the solid-liquid and solid-solid interactions at the same time.
Finally, the fluid-solid IFE can be calculated as follows:243
(21) |
It is important to note that an assumption of zero vacuum-vapor IFE is made in step (2). This approximation reduces computation cost because the vapor-liquid IFE can be calculated by the efficient method based on Bakker’s equation. However, when the vacuum-vapor IFE significantly deviates from zero ( cases with high vapor density), the cleaving wall method of Addula and Punnathanam231 could be applied for computing instead of the method based on Bakker’s equation. Moreover, liquid molecules may move through the solid part when transforming an immersed slab into an Einstein crystal in step (4), potentially causing high forces and instabilities that may crash the simulation. To avoid this situation, two extra TI stages can be introduced to add and remove potential walls, protecting the slab during its transformation, as implemented by Benjamin and Horbach in their cleaving wall method.228
One key advantage of this method in contrast to methods of Addula and Punnathanam231 is that it incorporates correction for miscible species, which is particularly useful for some solid substances containing miscible components that are loosely attached to the surface and can move into the liquid layer. The procedures for calculation of this correction are detailed in the original reference243 and are not included here. In addition, the inclusion of vacuum regions on both sides of the liquid-solid-liquid setup (see step (4) of Fig. 11) enables the method to study surfaces with dipole moments by applying the dipole correction method of Ballenegger et al.244, 245. The LAMMPS scripts necessary for implementing this approach are provided in the supplementary material of the Ref. 243. These features make the method highly promising for applications across a wide range of realistic systems.
The approach has been utilized to examine IFEs of the calcium sulfate hydrate systems to demonstrate the advantages of the method.243 Calcium sulfate hydrate exists in three distinct hydration phases: anhydrous anhydrite (CaSO0H2O), hemihydrate bassanite (CaSO0.5H2O), and dihydrate gypsum (CaSO42H2O). Due to the presence of strongly binding Ca2+ ions, significant ordering of the water layer on the interfaces is anticipated. Additionally, the structure contains stoichiometric water, indicating that interfacial water molecules are formally part of the solid, yet exhibit liquid-like behavior. Therefore, when estimating the IFE, it is important to include the correction term for the miscible species.243
Fig. 12a illustrates the IFEs of nine bassanite interfaces with varying Miller indices in contact with water. Both the enthalpic and entropic contributions were found to be positive. The enthalpic contribution is typically minimal due to the strong interaction between water and Ca2+ ions, which lowers the energy expanse for creating the interface.243 Conversely, the strong water binding causes significant ordering of interfacial water, leading to an entropy loss compared to the bulk water (, negative S) and a substantial entropic destabilization.243 Entropic contributions to the IFE range from around 40% to 90%. Additionally, two variants of the bassanite 1 1 0 interface were analyzed. The first variant is highly rough, featuring interface crenellations of CaSO4 chains. The second variant, indicated by “F” in the subscript, has these crenellations removed. It was found that the crenelated interface is significantly more stable, mainly due to smaller enthalpic contributions. This increased stability is due to the enhanced surface area provided by CaSO4 crenellations, which strengthens water binding.243 Notably, the entropy contributions to the free energies are nearly identical for both interfaces, implying similar water ordering on each. This similarity in water ordering was explained by considering the arrangement of water molecules beyond the initial adsorbed layer.243
The IFEs and IFE contributions of the gypsum interfaces are shown in Fig. 12b. Generally, the IFEs of gypsum interfaces are quite similar to those of bassanite. But the 0 1 0 interface exhibits a negative entropy contribution to the IFE (, positive S). This was explained by the highly structured water layers within the gypsum structure, which become exposed when the [0 1 0] plane is cleaved.243 As these water molecules gain disorder after cleaving, their entropy rises, resulting in a negative contribution.243 The other gypsum interfaces display positive entropic contributions to the IFE, ranging from approximately 20% to 60%, which is much smaller than those observed for bassanite.
3.2 Free Energy Perturbation Methods
Free Energy Perturbation methods in molecular simulations estimate the free energy differences between two states by sampling the energy differences as the system is slightly perturbed from one state to the other.246 This section introduces two techniques, the test-volume and test-area methods, tailored for calculating relative IFE in fluid-solid systems.
3.2.1 Test-volume Method
The test-volume method was initially employed by Eppenga and Frenkel247 to obtain the bulk pressure of isotropic and nematic phases in systems containing repulsive hard discs. The method has since evolved to estimate fluid-solid relative IFE in complex molecular systems with various types of interactions,248, 249, 250 and applied for estimating the fluid-solid relative IFE in molecular simulations.251, 252, 253, 173, 254 Remarkably, Fujiwara and Shibahara173 extended the method for subsystems, and local pressure profiles can be obtained through their approach. They also introduced an instantaneous expression for local pressure components and relative IFE using the test-volume method.255 This approach is particularly useful for studying time-dependent fluid-solid interfacial energies in non-equilibrium molecular dynamics simulations.
The test-volume method estimates the principle components of the pressure tensor of the fluid directly by slightly perturbing the fluid volume in directions normal and tangential to the interface. A sketch of the method for the fluid-solid system is shown in Fig. 13a. A rigorous derivation of working equations has been given for both canonical and grand canonical ensembles based on statistical mechanics by Fujiwara and Shibahara.173 While full derivation details are given in Ref. 173, a brief outline is presented here to give the working equation in the canonical ensemble as an example for calculating principle components of the pressure tensor:173
(22) |
where with the Boltzmann constant , denotes the ensemble average in unperturbed system, and indicates the change of certain property after perturbation. , , , , and are fluid particle number, the volume of the fluid, temperature, size of the fluid in -direction, and interfacial area, respectively. , , and are the Helmholtz free energy, the potential energy between fluid particles, and the potential energy between fluid particles and the solid (treated as an external field), separately.
The volume perturbation changes the distance between particles in the system leading to variations in and . The equations for transformations of the distances between fluid particles and between fluid particles and solid atoms can be found elsewhere.256, 173 It is also essential to note that the fluid volume is determined by the region occupied by its particles. To ensure accuracy, the magnitude of volume perturbation must be optimized through convergence testing.173 The method usually combines the central finite difference method to enhance the accuracy (see Fig. 13a).256, 173
After getting the principle components of the stress tensor of the inhomogeneous fluid, the fluid-solid relative IFE can then be calculated after applying the sum rule to Bakker’s equation (Eq. 2):173
(23) |
where is the size of the fluid in the z-direction ( direction normal to the interface).
The test-volume method has been recognized for its effectiveness in determining the solid-fluid interfacial tension, especially in scenarios involving high-density fluid that interacts through discontinuous potentials. Brumby et al.254 applied the test-volume method to understand the relative IFE of a hard-rod fluid made of hard-spherocylinder particles in planar confinement. Such a study could offer insights into how confinement affects a dense nematic liquid-crystalline fluid, which would be valuable for applications such as liquid-crystal displays.257 Part of their results are summarized in Fig. 14. Fig. 14a presents four representative snapshots of configurations at different bulk concentrations. As the bulk concentration increases, the bulk phase transitions from an isotropic phase to a nematic phase. At small bulk concentrations, dewetting of particles at the interface of the wall occurs due to the depletion of spherocylinder particles from the impermeable wall. With an increase in bulk density, a notable change is observed in the structure of the interface.257
Fig. 14b presents the relative IFE values. The trend of the relative IFEs as a function of concentration in the isotropic bulk phase exhibits a non-monotonic behavior. The maximum in relative IFEs is observed at a bulk concentration of around 2, indicating the initiation of the wetting transition and surface biaxial order.254 As the density progresses further into the nematic region, the relative IFE gradually decreases and stabilizes at a constant value at high concentrations near the bulk nematic-smectic transition, estimated to occur at a bulk concentration of approximately 5.254 The comparison of relative IFEs obtained using the test-volume method was conducted against values at low bulk densities established by Mao et al.,212 where the relative IFEs are obtained from integrating the Gibbs adsorption equation with data from grand canonical simulations. However, employing Monte Carlo insertions in grand canonical simulations becomes increasingly challenging as density increases. It is noteworthy that the relative IFEs obtained from the test-volume method are qualitatively consistent with those from cDFT.254
The surface adsorption can be calculated by subtracting the bulk contribution from the density distributions. The relationship between surface adsorption and bulk concentrations is illustrated in Fig. 14c. Remarkably, the decrease in relative IFEs in bulk isotropic states correlates with a sharp increase in surface adsorption.254 It is important to note that for the fluid-fluid interface, the non-monotonic behavior of IFE with varying density (or pressure) has been linked to the reversal of the sign of relative surface adsorption, according to the Gibbs adsorption equation.258, 259, 121
3.2.2 Test-area Method
The test-area method, developed by Gloor et al.,256 was originally applied to determine IFE in planar fluid-fluid interfaces within molecular simulations conducted in the canonical ensemble. The method was extended to address the relative solid-fluid IFE of inhomogeneous fluids inside slit-like pores utilizing the grand canonical260, 173 and canonical173 ensembles based on statistical mechanics. The technique was also adapted to address fluids confined within cylindrical pores.261 Note that Ghoufi and Malfreyt introduced a different approach to the test-area method by directly deriving the partition function. This method, known as TA2, avoids challenges related to exponential averages and the division of the surface into local elements along the normal direction.262, 263 Fujiwara and Shibahara173 also extended the test-area method to investigate subsystems, enabling the extraction of local pressure profiles through this refined approach. Excellent agreement of local profiles of the fluid-solid relative IFE has been reported between predictions from the perturbation method and method based on Bakker’s equation using the Irving-Kirkwood contour.173
The test-area method calculates the variation in free energy resulting from an infinitesimally small alteration in the interfacial area while maintaining the volume of the fluid. A sketch of the method for the slit pore system is shown in Fig. 13b. Under area perturbations in the canonical ensemble, the fluid-solid relative IFE can be calculated using the following equation:173
(24) |
Similar to the test-volume method, implementation of the area perturbation involves transformations of the distance between particles in the system.256, 173 The suitable area perturbation for the method can be ascertained through a convergence test,173 and the method typically integrates the central finite difference method,256, 173 as illustrated in Fig. 13b.
It is important to note that the formulation of the test-volume and test-area methods described above treats the solids as external fields. In other words, the positions of solid particles remain unchanged during the perturbation. Several authors have explicitly considered the solid phase during area perturbation for determining the surface stress.264, 265 The applied approach is analogous to the conventional test-area method used for fluid-fluid interfaces,256 and it involves perturbing the positions of both fluid and solid particles. However, Wu and Firoozabadi pointed out that this method overlooks the deformation work of the bulk solid, potentially resulting in unphysical negative values of surface stress.140 They proposed a correction term that compensates for the work from the solid deformation to address the issue:
(25) |
where denotes the surface stress (referred to as surface tension in “Supporting Information” of Ref. 140), is the interfacial energy obtained from the conventional test-area method with solid particles perturbed, and the last term is the correction from solid deformation work. Factor 2 accounts for two interfaces. is the thickness of the solid slab, and is the deviatoric stress of the bulk solid. Note that and are determined by fitting Eq. 25 with various from conventional test-area method and determined from separate simulations of bulk solid under various stress conditions. With this approach, the contributions from variations in surface area and solid bulk deformations to the free energy change during test-area perturbation can be estimated and analyzed separately.266
Figs. 15a and b show that significant differences exist between the conventional test-area method and the one that considers solid deformation work. The interfacial energy from the conventional test-are method ranges from negative to positive. The negative values violate thermodynamic stability.140 The corrected surface stresses consistently yield positive values, contrasting with the conventional method’s tendency to yield both negative and positive values. Remarkably, the pressure and temperature effects are opposite for those two methods.140 Fig. 15c displays the contributions of each term in Eq. 25 under various deviatoric stresses. It is observed that with higher , deviates more from . Under high-compression conditions, values are found to be negative, and the contribution from solid deformation is positive. Opposite signs are observed under high-tension condtions.140
Very recently, Ghoufi proposed a semi-empirical method for calculating the excess free energy per unit interfacial area:267
(26) |
where is the excess helmholz free energy, is the temperature, is the excess internal energy per unit interfacial area (), and is the excess entropy per unit interfacial area (). is readily available from molecular simulations. However, it is challenging to estimate . Ghoufi derived an expression that relates with :267
(27) |
where is the excess number of molecules of component , is the parallel strain, and is the normal stress. Eq. 27 allows to calculate numerically with data of under various . Empirical fitting the with a second-order polynomial without a linear term is proposed to solve for .267 Moreover, Ghoufi rediscovered the Shuttleworth equation143 and Dong’s relation,268 specifically applied to , from the first law of thermodynamics.267
Note that the method of Ghoufi does not calculate IFE directly. However, is related to the IFE through the following relation:267
(28) |
where is the surface excess of species which can be estimated based on density distributions, and is the chemical potential of component . Therefore, can be accessed with chemical potential being estimated elsewhere.267
The semi-empirical method offers valuable insights into by integrating its energetic () and entropic () contributions. This approach can be applied to systems with curved interfaces without the need to calculate the local pressure tensor or introduce perturbations to the system. Additionally, the method only requires simulations at various temperatures, making it relatively more computationally efficient than other computation-intensive methods.267
The method of Ghoufi does not belong to the class of free energy perturbation methods. Nevertheless, Ghoufi compared the obtained from the semi-empirical approach with those derived from the TA2 method.267 It was concluded that the test-area method is unsuitable for computing the interfacial energies of flexible solids. Furthermore, there is an ongoing debate regarding the formulation of the free energy of fluid-solid system considering the solid deformation.99, 266, 202, 203 Therefore, it is essential to provide a more detailed clarification of the formulation and to rigorously validate the method by conducting comparisons with other well-established techniques.
4 Conclusion
This review examines methodologies for estimating the interfacial energies of fluid-solid interfaces, a critical factor for understanding wettability via Young’s equation. An overview of experimental, semi-empirical, and first-principle theoretical methods was presented at the beginning. Obtaining reliable fluid-solid IFE data using experimental and semi-empirical methods remains challenging. Furthermore, first-principle theoretical methods are limited by the intricate nature of the underlying theory and the scarcity of robust computational tools specifically designed for fluid-solid IFE calculations. We highlight the advantages of methods based on molecular simulations from various perspectives. The primary aim of this review is to deliver an in-depth analysis of recent advancements in molecular simulation techniques for estimating fluid-solid IFE. We discuss the fundamental principles, methodological developments, and practical implementations of various molecular simulation techniques, categorized into mechanical approaches—such as the contact angle method, Bakker’s equation-based technique, and the Wilhelmy simulation—and thermodynamic approaches, including the cleaving wall method, the Frenkel-Ladd technique, and the test-volume/area methods. Notably, both the mechanical and thermodynamic approaches yield consistent values for fluid-solid IFE.145, 97, 173 The listed methods are suitable for calculating a range of interfacial energies of fluid-solid interfaces, including IFE, relative IFE, surface stress, and superficial tension, terms that have often been used in previous studies without clear differentiation. Meanwhile, selected applications of these methods are presented to gain insights into the behaviors of fluid-solid interfacial energies. In addition, the fluid-solid IFE is analyzed within the theoretical framework of Navascués and Berry35 and the thought experiment based on Bakker’s equation.145 It is shown that the simulation methods based on the theory of Navascués and Berry35 and the extended Bakker’s equation derived by Nijmeijer and Leeuwen163 are identical.
A key limitation in applying molecular simulation methods for fluid-solid interfacial energy estimation is the lack of widely accessible software tools and standardized benchmark examples. Although certain tools exist, such as the “CLEAVING” package230 and the LAMMPS scripts for the Frenkel-Ladd technique243), numerous other methods lack accessible, published code for replication and implementation. Implementing these methods may require a significant coding effort, particularly for non-expert users. Therefore, developing user-friendly software packages and comprehensive benchmark examples is essential to enable wider adoption of these simulation techniques and to promote reproducibility across the research community.
Another challenge arises from the computational complexity associated with calculating interfacial energies. Determining the pressure tensor is computationally demanding due to its second-order nature and slow convergence rate. Methods like the contact angle and Wilhelmy simulation approaches require large-scale simulation systems and prolonged equilibrium times to account for size effects and hysteresis,70 significantly raising computational costs. Additionally, thermodynamic integration methods typically involve multiple intricate steps that require numerous simulations at various states. Overcoming these computational challenges necessitates advanced numerical strategies, optimized sampling algorithms, and careful design of thermodynamic paths.
There are ongoing debates rooted in certain methodologies. For example, the historical controversies surrounding the non-unique definition of the microscopic pressure tensor in Bakker’s method underscore the need for clarity and consensus.170 Further clarification is required in the formulation of test-area methods when applied to flexible solids to improve both their conceptual understanding and practical applicability.267, 99, 266 Moreover, while the available methods primarily focus on classical force fields, this limitation underscores the critical need for the development of methods based on quantum mechanics.269, 270 Advancements in quantum mechanics-based methodologies hold the potential to broaden the scope of applications and achieve more precise prediction in complex chemical systems.
The combination of experiment, simulation, and theory holds great promise for advancing the knowledge of fluid-solid interfacial energies. Accurate measurement of fluid-solid IFEs at the nanoscale remains challenging, primarily due to the technical limitations of current experimental methods. A breakthrough in measurement techniques would greatly benefit simulation- and theory-based methods. Meanwhile, cDFT is a promising tool for estimating interfacial energies. Qualitative agreement in relative IFEs between simulations and cDFT has been reported for the nematic liquid-crystalline fluid under confinement.254 However, cDFT’s application is currently restricted by a lack of accessible, dedicated software tools, a limitation even more significant than that seen in molecular simulation methods. This highlights an urgent need for software development to fully harness cDFT’s potential.
The fluid-solid IFE data is scarce in the literature in contrast to the fluid-fluid IFE data.271, 272, 273, 121 The insights provided in this review can broaden the application of molecular simulation methods across diverse chemical systems. These systems include but are not limited to, spherical and cylindrical surfaces, confined systems, charged surfaces, surfaces grafted with complex functional groups, and fluids exhibiting intricate phase behaviors. Interfacial energy data from molecular simulations can provide valuable reference values that improve theoretical methods31, 46, 47 and data-driven approaches.274, 275 These advances will support more accurate and efficient predictions of fluid-solid IFEs in a range of engineering and scientific applications.
Acknowledgments
The authors acknowledge the support from the National Natural Science Foundation of China under Grant No. 42203041, the Natural Science Foundation of Jiangsu Province under Grant No. BK20221132, and the Hong Kong Scholars Program (XJ2023042).
The authors also acknowledge the support from King Abdullah University of Science and Technology (KAUST) through the grants BAS/1/1351-01 and URF/1/5028-01.
The authors sincerely thank the anonymous reviewers and the editor for their valuable and constructive feedback.
References
- De Gennes 1985 De Gennes, P.-G. Wetting: statics and dynamics. Reviews of Modern Physics 1985, 57, 827
- Drelich et al. 2019 Drelich, J. W.; Boinovich, L.; Chibowski, E.; Della Volpe, C.; Hołysz, L.; Marmur, A.; Siboni, S. Contact angles: History of over 200 years of open questions. Surface Innovations 2019, 8, 3–27
- Gibbs 1906 Gibbs, J. W. The Scientific Papers of J. Willard Gibbs; Longmans, Green and Company, 1906; Vol. 1
- Israelachvili 2022 Israelachvili, J. N. The Handbook of Surface Imaging and Visualization; CRC Press, 2022; pp 793–816
- Jian et al. 2018 Jian, W.; Tam, L.-h.; Lau, D. Atomistic study of interfacial creep behavior in epoxy-silica bilayer system. Composites Part B: Engineering 2018, 132, 229–236
- Tam et al. 2017 Tam, L.-h.; Chow, C. L.; Lau, D. Moisture effect on interfacial integrity of epoxy-bonded system: a hierarchical approach. Nanotechnology 2017, 29, 024001
- Yu and Lau 2015 Yu, Z.; Lau, D. Molecular dynamics study on stiffness and ductility in chitin–protein composite. Journal of materials science 2015, 50, 7149–7157
- Young 1805 Young, T. III. An essay on the cohesion of fluids. Philosophical Transactions of the Royal Society of London 1805, 65–87
- Drelich et al. 2002 Drelich, J.; Fang, C.; White, C. Measurement of interfacial tension in fluid-fluid systems. Encyclopedia of Surface and Colloid Science 2002, 3, 3158–3163
- Lander et al. 1993 Lander, L. M.; Siewierski, L. M.; Brittain, W. J.; Vogler, E. A. A systematic comparison of contact angle methods. Langmuir 1993, 9, 2237–2239
- Zhu et al. 2023 Zhu, S.; Xie, K.; Lin, Q.; Cao, R.; Qiu, F. Experimental determination of surface energy for high-energy surface: A review. Advances in Colloid and Interface Science 2023, 315, 102905
- Geguzin and Ovcharenko 1962 Geguzin, Y. E.; Ovcharenko, N. Surface energy and surface processes in solids. Soviet Physics Uspekhi 1962, 5, 129
- Linford 1972 Linford, R. Surface energy of solids. Chemical Society Reviews 1972, 1, 445–464
- Brown and March 1976 Brown, R. C.; March, N. H. Structure and excitations in liquid and solid surfaces. Physics Reports 1976, 24, 77–169
- Kumikov and Khokonov 1983 Kumikov, V.; Khokonov, K. B. On the measurement of surface free energy and surface tension of solid metals. Journal of Applied Physics 1983, 54, 1346–1350
- Bailey and Kay 1967 Bailey, A. I.; Kay, S. M. A direct measurement of the influence of vapour, of liquid and of oriented monolayers on the interfacial energy of mica. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 1967, 301, 47–56
- Parks 1984 Parks, G. A. Surface and interfacial free energies of quartz. Journal of Geophysical Research: Solid Earth 1984, 89, 3997–4008
- Claesson et al. 1986 Claesson, P. M.; Blom, C. E.; Herder, P. C.; Ninham, B. W. Interactions between water—stable hydrophobic Langmuir—Blodgett monolayers on mica. Journal of Colloid and Interface Science 1986, 114, 234–242
- Nadermann et al. 2013 Nadermann, N.; Hui, C.-Y.; Jagota, A. Solid surface tension measured by a liquid drop under a solid film. Proceedings of the National Academy of Sciences 2013, 110, 10541–10545
- Ward and Wu 2008 Ward, C.; Wu, J. Effect of contact line curvature on solid-fluid surface tensions without line tension. Physical Review Letters 2008, 100, 256103
- Sarkar et al. 2023 Sarkar, S.; Gukeh, M. J.; Roy, T.; Gaikwad, H.; Bellussi, F. M.; Moitra, S.; Megaridis, C. M. A new methodology for measuring solid/liquid interfacial energy. Journal of Colloid and Interface Science 2023, 633, 800–807
- Tyson and Miller 1977 Tyson, W.; Miller, W. Surface free energies of solid metals: Estimation from liquid surface tension measurements. Surface Science 1977, 62, 267–276
- Di Pasquale and Davidchack 2022 Di Pasquale, N.; Davidchack, R. L. Cleaving method for molecular crystals and its application to calculation of the surface free energy of crystalline -d-mannitol at room temperature. The Journal of Physical Chemistry A 2022, 126, 2134–2141
- Neumann et al. 1974 Neumann, A.; Good, R.; Hope, C.; Sejpal, M. An equation-of-state approach to determine surface tensions of low-energy solids from contact angles. Journal of Colloid and Interface Science 1974, 49, 291–304
- Li and Neumann 1992 Li, D.; Neumann, A. Equation of state for interfacial tensions of solid-liquid systems. Advances in Colloid and Interface Science 1992, 39, 299–345
- Fox and Zisman 1952 Fox, H.; Zisman, W. The spreading of liquids on low-energy surfaces. III. Hydrocarbon surfaces. Journal of Colloid Science 1952, 7, 428–442
- Fowkes 1964 Fowkes, F. M. Attractive forces at interfaces. Industrial & Engineering Chemistry 1964, 56, 40–52
- Owens and Wendt 1969 Owens, D. K.; Wendt, R. Estimation of the surface free energy of polymers. Journal of Applied Polymer Science 1969, 13, 1741–1747
- Wu 1971 Wu, S. Calculation of interfacial tension in polymer systems. Journal of Polymer Science Part C: Polymer Symposia. 1971; pp 19–30
- Van Oss et al. 1986 Van Oss, C.; Good, R.; Chaudhury, M. The role of van der Waals forces and hydrogen bonds in “hydrophobic interactions” between biopolymers and low energy surfaces. Journal of Colloid and Interface Science 1986, 111, 378–390
- Żenkiewicz 2007 Żenkiewicz, M. Methods for the calculation of surface free energy of solids. Journal of Achievements in Materials and Manufacturing Engineering 2007, 24, 137–145
- Sullivan 1981 Sullivan, D. Surface tension and contact angle of a liquid–solid interface. The Journal of Chemical Physics 1981, 74, 2604–2615
- Good 1975 Good, R. J. Spreading pressure and contact angle. Journal of Colloid and Interface Science 1975, 52, 308–313
- Yang et al. 2023 Yang, Y.; Wan, J.; Shang, X.; Sun, S. Molecular insights into fluid-solid interfacial tensions in water+ gas+ solid systems at various temperatures and pressures. The Journal of Chemical Physics 2023, 159
- Navascués and Berry 1977 Navascués, G.; Berry, M. The statistical mechanics of wetting. Molecular Physics 1977, 34, 649–664
- Kirkwood and Buff 1949 Kirkwood, J. G.; Buff, F. P. The statistical mechanical theory of surface tension. The Journal of Chemical Physics 1949, 17, 338–343
- Nijmeijer et al. 1990 Nijmeijer, M.; Bruin, C.; Bakker, A.; Van Leeuwen, J. Wetting and drying of an inert wall by a fluid in a molecular-dynamics simulation. Physical Review A 1990, 42, 6052
- Rayleigh 1892 Rayleigh, L. On the theory of surface forces.—II. Compressible fluids. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 1892, 33, 209–220
- van der Waals 2012 van der Waals, J. D. Lehrbuch der Thermodynamik; BoD–Books on Demand, 2012; Vol. 7
- Cahn and Hilliard 1958 Cahn, J. W.; Hilliard, J. E. Free energy of a nonuniform system. I. Interfacial free energy. The Journal of Chemical Physics 1958, 28, 258–267
- Cahn 1977 Cahn, J. W. Critical point wetting. The Journal of Chemical Physics 1977, 66, 3667–3672
- Benner et al. 1981 Benner, R. E.; Scriven, L.; Davis, H. T. Structure and stress in the gas–liquid–solid contact region. Faraday Symposia of the Chemical Society. 1981; pp 169–190
- Davis and Scriven 1982 Davis, H.; Scriven, L. Stress and structure in fluid interfaces. Advances in The Journal of Chemical Physics 1982, 49, 357–454
- Ebner and Saam 1977 Ebner, C.; Saam, W. New phase-transition phenomena in thin argon films. Physical Review Letters 1977, 38, 1486
- Evans and Marini Bettolo Marconi 1987 Evans, R.; Marini Bettolo Marconi, U. Phase equilibria and solvation forces for fluids confined between parallel walls. The Journal of Chemical Physics 1987, 86, 7138–7148
- Evans 1979 Evans, R. The nature of the liquid-vapour interface and other topics in the statistical mechanics of non-uniform, classical fluids. Advances in Physics 1979, 28, 143–200
- Wu and Li 2007 Wu, J.; Li, Z. Density-functional theory for complex fluids. Annual Review of Physical Chemistry 2007, 58, 85–112
- Yu 2009 Yu, Y.-X. A novel weighted density functional theory for adsorption, fluid-solid interfacial tension, and disjoining properties of simple liquid films on planar solid surfaces. The Journal of Chemical Physics 2009, 131
- Yatsyshin et al. 2018 Yatsyshin, P.; Durán-Olivencia, M.; Kalliadasis, S. Microscopic aspects of wetting using classical density functional theory. Journal of Physics: Condensed Matter 2018, 30, 274003
- Dabrowska et al. 2022 Dabrowska, K.; Pizio, O.; Sokołowski, S. Contact angle of water on a model heterogeneous surface. A density functional approach. arXiv preprint arXiv:2210.01022 2022,
- Salinger et al. 2008 Salinger, A. G.; Frischknecht, A. L.; Frink, L. J. Tramonto: Software for nanostructured fluids in materials and biology.; 2008
- Jiang 2022 Jiang, J. Software package: An advanced theoretical tool for inhomogeneous fluids (Atif). Chinese Journal of Polymer Science 2022, 1–11
- Rehner et al. 2023 Rehner, P.; Bauer, G.; Gross, J. FeOs: An open-source framework for equations of state and classical density functional theory. Industrial & Engineering Chemistry Research 2023, 62, 5347–5357
- Thompson et al. 2022 Thompson, A. P.; Aktulga, H. M.; Berger, R.; Bolintineanu, D. S.; Brown, W. M.; Crozier, P. S.; in’t Veld, P. J.; Kohlmeyer, A.; Moore, S. G.; Nguyen, T. D.; others LAMMPS-a flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Computer Physics Communications 2022, 271, 108171
- Abraham et al. 2015 Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1, 19–25
- Hens et al. 2020 Hens, R.; Rahbari, A.; Caro-Ortiz, S.; Dawass, N.; Erdős, M.; Poursaeidesfahani, A.; Salehi, H. S.; Celebi, A. T.; Ramdin, M.; Moultos, O. A.; others Brick-CFCMC: Open source software for Monte Carlo simulations of phase and reaction equilibria using the continuous fractional component method. Journal of Chemical Information and Modeling 2020, 60, 2678–2682
- Frenkel and Smit 2023 Frenkel, D.; Smit, B. Understanding molecular simulation: from algorithms to applications; Elsevier, 2023
- Allen and Tildesley 2017 Allen, M. P.; Tildesley, D. J. Computer simulation of liquids; Oxford university press, 2017
- Heinz et al. 2013 Heinz, H.; Lin, T.-J.; Kishore Mishra, R.; Emami, F. S. Thermodynamically consistent force fields for the assembly of inorganic, organic, and biological nanostructures: the INTERFACE force field. Langmuir 2013, 29, 1754–1765
- Cygan et al. 2004 Cygan, R. T.; Liang, J.-J.; Kalinichev, A. G. Molecular models of hydroxide, oxyhydroxide, and clay phases and the development of a general force field. The Journal of Physical Chemistry B 2004, 108, 1255–1266
- Schmitt et al. 2023 Schmitt, S.; Kanagalingam, G.; Fleckenstein, F.; Froescher, D.; Hasse, H.; Stephan, S. Extension of the MolMod database to transferable force fields. Journal of Chemical Information and Modeling 2023,
- Senftle et al. 2016 Senftle, T. P.; Hong, S.; Islam, M. M.; Kylasa, S. B.; Zheng, Y.; Shin, Y. K.; Junkermeier, C.; Engel-Herbert, R.; Janik, M. J.; Aktulga, H. M.; others The ReaxFF reactive force-field: development, applications and future directions. npj Computational Materials 2016, 2, 1–14
- Hao et al. 2020 Hao, H.; Chow, C. L.; Lau, D. Effect of heat flux on combustion of different wood species. Fuel 2020, 278, 118325
- Hao et al. 2020 Hao, H.; Chow, C. L.; Lau, D. Carbon monoxide release mechanism in cellulose combustion using reactive forcefield. Fuel 2020, 269, 117422
- Avendano et al. 2011 Avendano, C.; Lafitte, T.; Galindo, A.; Adjiman, C. S.; Jackson, G.; Müller, E. A. SAFT- force field for the simulation of molecular fluids. 1. A single-site coarse grained model of carbon dioxide. The Journal of Physical Chemistry B 2011, 115, 11154–11169
- Nie et al. 2023 Nie, F.; Jian, W.; Yu, Z.; Chow, C. L.; Lau, D. Mesoscale modeling to study isolated asphaltene agglomerates. Construction and Building Materials 2023, 379, 131249
- Qin et al. 2021 Qin, R.; Zhou, A.; Yu, Z.; Wang, Q.; Lau, D. Role of carbon nanotube in reinforcing cementitious materials: An experimental and coarse-grained molecular dynamics study. Cement and Concrete Research 2021, 147, 106517
- Yu et al. 2016 Yu, Z.; Zhou, A.; Lau, D. Mesoscopic packing of disk-like building blocks in calcium silicate hydrate. Scientific Reports 2016, 6, 36967
- Yu and Lau 2015 Yu, Z.; Lau, D. Development of a coarse-grained -chitin model on the basis of MARTINI forcefield. Journal of Molecular Modeling 2015, 21, 1–9
- Jiang and Patel 2019 Jiang, H.; Patel, A. J. Recent advances in estimating contact angles using molecular simulations and enhanced sampling methods. Current Opinion in Chemical Engineering 2019, 23, 130–137
- Bokeloh et al. 2011 Bokeloh, J.; Rozas, R. E.; Horbach, J.; Wilde, G. Nucleation barriers for the liquid-to-crystal transition in Ni: experiment and simulation. Physical Review Letters 2011, 107, 145701
- Aman and Koh 2016 Aman, Z. M.; Koh, C. A. Interfacial phenomena in gas hydrate systems. Chemical Society Reviews 2016, 45, 1678–1690
- Gránásy et al. 2002 Gránásy, L.; Pusztai, T.; James, P. F. Interfacial properties deduced from nucleation experiments: A Cahn–Hilliard analysis. The Journal of Chemical Physics 2002, 117, 6157–6168
- Hu et al. 2011 Hu, Z.; Schall, P.; others Single crystal growth and anisotropic crystal-fluid interfacial free energy in soft colloidal systems. Physical Review E 2011, 84, 011607
- Louhichi et al. 2013 Louhichi, A.; Tamborini, E.; Ghofraniha, N.; Caton, F.; Roux, D.; Oberdisse, J.; Cipelletti, L.; Ramos, L. Nucleation and growth of micellar polycrystals under time-dependent volume fraction conditions. Physical Review E 2013, 87, 032306
- Djikaev and Ruckenstein 2017 Djikaev, Y.; Ruckenstein, E. Self-consistent determination of the ice–air interfacial tension and ice–water–air line tension from experiments on the freezing of water droplets. The Journal of Physical Chemistry C 2017, 121, 16432–16439
- Härtel et al. 2012 Härtel, A.; Oettel, M.; Rozas, R. E.; Egelhaaf, S. U.; Horbach, J.; Löwen, H. Tension and stiffness of the hard sphere crystal-fluid interface. Physical Review Letters 2012, 108, 226101
- Wang et al. 2013 Wang, X.; Mi, J.; Zhong, C. Density functional theory for crystal-liquid interfaces of Lennard-Jones fluid. The Journal of Chemical Physics 2013, 138
- Schoonen and Lutsko 2022 Schoonen, C.; Lutsko, J. F. Crystal Polymorphism Induced by Surface Tension. Physical Review Letters 2022, 129, 246101
- Hoyt et al. 2001 Hoyt, J.; Asta, M.; Karma, A. Method for computing the anisotropy of the solid-liquid interfacial free energy. Physical Review Letters 2001, 86, 5530
- Auer and Frenkel 2001 Auer, S.; Frenkel, D. Prediction of absolute crystal-nucleation rate in hard-sphere colloids. Nature 2001, 409, 1020–1023
- Ferreira and Barroso 2000 Ferreira, A.; Barroso, M. Temperature and density extrapolations in canonical ensemble Monte Carlo simulations. Physical Review E 2000, 61, 1195
- Barroso and Ferreira 2002 Barroso, M.; Ferreira, A. Solid–fluid coexistence of the Lennard-Jones system from absolute free energy calculations. The Journal of Chemical Physics 2002, 116, 7145–7150
- Laio and Parrinello 2002 Laio, A.; Parrinello, M. Escaping free-energy minima. Proceedings of the National Academy of Sciences 2002, 99, 12562–12566
- Angioletti-Uberti et al. 2010 Angioletti-Uberti, S.; Ceriotti, M.; Lee, P. D.; Finnis, M. W. Solid-liquid interface free energy through metadynamics simulations. Physical Review B 2010, 81, 125416
- Lau et al. 2014 Lau, D.; Broderick, K.; Buehler, M. J.; Büyüköztürk, O. A robust nanoscale experimental quantification of fracture energy in a bilayer material system. Proceedings of the National Academy of Sciences 2014, 111, 11990–11995
- Lau et al. 2012 Lau, D.; Büyüköztürk, O.; Buehler, M. J. Characterization of the intrinsic strength between epoxy and silica using a multiscale approach. Journal of Materials Research 2012, 27, 1787–1796
- Luo et al. 2003 Luo, S.-N.; Ahrens, T. J.; Çağın, T.; Strachan, A.; Goddard III, W. A.; Swift, D. C. Maximum superheating and undercooling: Systematics, molecular dynamics simulations, and dynamic experiments. Physical Review B 2003, 68, 134206
- Luo et al. 2004 Luo, S.-N.; Strachan, A.; Swift, D. C. Nonequilibrium melting and crystallization of a model Lennard-Jones system. The Journal of Chemical Physics 2004, 120, 11640–11649
- Luo et al. 2005 Luo, S.-N.; Strachan, A.; Swift, D. C. Deducing solid–liquid interfacial energy from superheating or supercooling: application to H2O at high pressures. Modelling and Simulation in Materials Science and Engineering 2005, 13, 321
- Bai and Li 2005 Bai, X.-M.; Li, M. Differences between solid superheating and liquid supercooling. The Journal of Chemical Physics 2005, 123
- Knott et al. 2012 Knott, B. C.; Molinero, V.; Doherty, M. F.; Peters, B. Homogeneous nucleation of methane hydrates: Unrealistic under realistic conditions. Journal of the American Chemical Society 2012, 134, 19544–19547
- Sanz et al. 2013 Sanz, E.; Vega, C.; Espinosa, J.; Caballero-Bernal, R.; Abascal, J.; Valeriani, C. Homogeneous ice nucleation at moderate supercooling from molecular simulation. Journal of the American Chemical Society 2013, 135, 15008–15017
- Fernández et al. 2012 Fernández, L.; Martín-Mayor, V.; Seoane, B.; Verrocchio, P. Equilibrium fluid-solid coexistence of hard spheres. Physical Review Letters 2012, 108, 165701
- Espinosa et al. 2014 Espinosa, J.; Vega, C.; Sanz, E. The mold integration method for the calculation of the crystal-fluid interfacial free energy from simulations. The Journal of Chemical Physics 2014, 141
- Maruyama et al. 2002 Maruyama, S.; Kimura, T.; Lu, M.-C.; others Molecular scale aspects of liquid contact on a solid surface. National Heat Transfer Symposium of Japan. 2002; pp 577–578
- Imaizumi et al. 2020 Imaizumi, Y.; Omori, T.; Kusudo, H.; Bistafa, C.; Yamaguchi, Y. Wilhelmy equation revisited: A lightweight method to measure liquid–vapor, solid–liquid, and solid–vapor interfacial tensions from a single molecular dynamics simulation. The Journal of Chemical Physics 2020, 153
- Toshev and Platikanov 2006 Toshev, B. V.; Platikanov, D. Wetting: Gibbs’ superficial tension revisited. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2006, 291, 177–180
- Valiya Parambathu et al. 2022 Valiya Parambathu, A.; Pinheiro dos Santos, T. J.; Chapman, W. G.; Asthagiri, D. N. Comment on “Calculation of Solid–Fluid Interfacial Free Energy with Consideration of Solid Deformation by Molecular Dynamics”. The Journal of Physical Chemistry A 2022, 126, 1782–1783
- Defay et al. 1966 Defay, R.; Prigogine, I.; Bellemans, A.; Everett, D. H. Surface tension and adsorption. Materials Science 1966,
- Taherian et al. 2013 Taherian, F.; Marcon, V.; van der Vegt, N. F.; Leroy, F. What is the contact angle of water on graphene? Langmuir 2013, 29, 1457–1465
- Fan et al. 2020 Fan, J.; De Coninck, J.; Wu, H.; Wang, F. Microscopic origin of capillary force balance at contact line. Physical Review Letters 2020, 124, 125502
- Yang et al. 2022 Yang, Y.; Nair, A. K. N.; Ruslan, M. F. A. C.; Sun, S. Interfacial properties of the alkane+ water system in the presence of carbon dioxide and hydrophobic silica. Fuel 2022, 310, 122332
- Ghoufi et al. 2016 Ghoufi, A.; Malfreyt, P.; Tildesley, D. J. Computer modelling of the surface tension of the gas–liquid and liquid–liquid interface. Chemical Society Reviews 2016, 45, 1387–1409
- Müller et al. 2020 Müller, E. A.; Ervik, Å.; Mejía, A. A guide to computing interfacial properties of fluids from molecular simulations [Article v1. 0]. Living Journal of Computational Molecular Science 2020, 2, 21385–21385
- Green et al. 1960 Green, H.; Ono, S.; Kondo, S.; Buff, F. P.; Ono, S.; Kondo, S. Molecular theory of surface tension in liquids. Structure of Liquids/Struktur der Flüssigkeiten 1960, 134–280
- Bakker 1928 Bakker, G. Kapillarität und oberflächenspannung; Wien-Harms, 1928; Vol. 6
- Kanduč 2017 Kanduč, M. Going beyond the standard line tension: Size-dependent contact angles of water nanodroplets. The Journal of Chemical Physics 2017, 147
- Silvestri et al. 2019 Silvestri, A.; Ataman, E.; Budi, A.; Stipp, S.; Gale, J. D.; Raiteri, P. Wetting properties of the CO2–water–calcite system via molecular simulations: shape and size effects. Langmuir 2019, 35, 16669–16678
- Godawat et al. 2008 Godawat, R.; Jamadagni, S. N.; Errington, J. R.; Garde, S. Structure, stability, and rupture of free and supported liquid films and assemblies in molecular simulations. Industrial & Engineering Chemistry Research 2008, 47, 3582–3590
- MacDowell et al. 2006 MacDowell, L. G.; Shen, V. K.; Errington, J. R. Nucleation and cavitation of spherical, cylindrical, and slablike droplets and bubbles in small systems. The Journal of Chemical Physics 2006, 125
- Jiang et al. 2019 Jiang, H.; Fialoke, S.; Vicars, Z.; Patel, A. J. Characterizing surface wetting and interfacial properties using enhanced sampling (SWIPES). Soft Matter 2019, 15, 860–869
- Le et al. 2021 Le, T. T. B.; Divine-Ayela, C.; Striolo, A.; Cole, D. R. Effects of surface contamination on the interfacial properties of CO 2/water/calcite systems. Physical Chemistry Chemical Physics 2021, 23, 18885–18892
- Wang et al. 2022 Wang, Y.; Kiziltas, A.; Blanchard, P.; Walsh, T. R. ContactAngleCalculator: An automated, parametrized, and flexible code for contact angle estimation in visual molecular dynamics. Journal of Chemical Information and Modeling 2022, 62, 6302–6308
- Stephan et al. 2019 Stephan, S.; Thol, M.; Vrabec, J.; Hasse, H. Thermophysical properties of the Lennard-Jones fluid: Database and data assessment. Journal of Chemical Information and Modeling 2019, 59, 4248–4265
- Pan et al. 2020 Pan, B.; Yin, X.; Iglauer, S. A review on clay wettability: From experimental investigations to molecular dynamics simulations. Advances in Colloid and Interface Science 2020, 285, 102266
- Shi and Dhir 2009 Shi, B.; Dhir, V. K. Molecular dynamics simulation of the contact angle of liquids on solid surfaces. The Journal of Chemical Physics 2009, 130
- Tenney and Cygan 2014 Tenney, C. M.; Cygan, R. T. Molecular simulation of carbon dioxide, brine, and clay mineral interactions and determination of contact angles. Environmental Science & Technology 2014, 48, 2035–2042
- Chen et al. 2015 Chen, C.; Zhang, N.; Li, W.; Song, Y. Water contact angle dependence with hydroxyl functional groups on silica surfaces under CO2 sequestration conditions. Environmental Science & Technology 2015, 49, 14680–14687
- Yang et al. 2022 Yang, Y.; Che Ruslan, M. F. A.; Narayanan Nair, A. K.; Qiao, R.; Sun, S. Interfacial properties of the hexane+ carbon dioxide+ water system in the presence of hydrophilic silica. The Journal of Chemical Physics 2022, 157
- Yang et al. 2023 Yang, Y.; Nair, A. K. N.; Zhu, W.; Sang, S.; Sun, S. Molecular perspectives of interfacial properties of the hydrogen+ water mixture in contact with silica or kerogen. Journal of Molecular Liquids 2023, 122337
- Yao et al. 2024 Yao, X.; Narayanan Nair, A. K.; Che Ruslan, M. F. A.; Yang, Y.; Yan, B.; Lau, D.; Sun, S. Interfacial Properties of the Nitrogen+ Water System in the Presence of Hydrophilic Silica. Industrial & Engineering Chemistry Research 2024, 63, 5765–5772
- Cui et al. 2023 Cui, R.; Narayanan Nair, A. K.; Che Ruslan, M. F. A.; Yang, Y.; Sun, S. Interfacial properties of the hexane+ carbon dioxide+ brine system in the presence of hydrophilic silica. Industrial & Engineering Chemistry Research 2023, 62, 13470–13478
- Yang et al. 2022 Yang, Y.; Nair, A. K. N.; Ruslan, M. F. A. C.; Sun, S. Interfacial properties of the aromatic hydrocarbon+ water system in the presence of hydrophilic silica. Journal of Molecular Liquids 2022, 346, 118272
- Yang et al. 2024 Yang, Y.; Narayanan Nair, A. K.; Lau, D.; Sun, S. Interfacial properties of the brine+ carbon dioxide+ oil+ silica system. The Journal of Chemical Physics 2024, 160
- Pan et al. 2021 Pan, B.; Yin, X.; Ju, Y.; Iglauer, S. Underground hydrogen storage: Influencing parameters and future outlook. Advances in Colloid and Interface Science 2021, 294, 102473
- Arif et al. 2019 Arif, M.; Abu-Khamsin, S. A.; Iglauer, S. Wettability of rock/CO2/brine and rock/oil/CO2-enriched-brine systems: Critical parametric analysis and future outlook. Advances in Colloid and Interface Science 2019, 268, 91–113
- Wang et al. 2023 Wang, J.; Wang, H.; Wang, X.; Yang, S.; Wu, H.; Leung, C.; Tian, J. A multiphysical-geochemical coupling model for caprock sealing efficiency in CO2 geosequestration. Deep Underground Science and Engineering 2023, 2, 188–203
- Chen et al. 2023 Chen, S.; Zhang, J.; Yin, D.; Li, F.; Lu, J.; Zhu, P. Visualizing experimental investigation on gas–liquid replacements in a microcleat model using the reconstruction method. Deep Underground Science and Engineering 2023, 2, 295–303
- Chatzis and Morrow 1984 Chatzis, I.; Morrow, N. R. Correlation of capillary number relationships for sandstone. Society of Petroleum Engineers Journal 1984, 24, 555–562
- Guo et al. 2020 Guo, H.; Song, K.; Hilfer, R. A critical review of capillary number and its application in enhanced oil recovery. SPE Improved Oil Recovery Conference. 2020; p D031S046R001
- Amarasinghe et al. 2014 Amarasinghe, P. M.; Anandarajah, A.; Ghosh, P. Molecular dynamic study of capillary forces on clay particles. Applied Clay Science 2014, 88, 170–177
- Chen et al. 2017 Chen, C.; Chai, Z.; Shen, W.; Li, W.; Song, Y. Wettability of supercritical CO2–brine–mineral: The effects of ion type and salinity. Energy & Fuels 2017, 31, 7317–7324
- Chen et al. 2019 Chen, C.; Hu, W.; Li, W.; Song, Y. Model comparison of the CH4/CO2/water system in predicting dynamic and interfacial properties. Journal of Chemical & Engineering Data 2019, 64, 2464–2474
- Sun and Bourg 2020 Sun, E. W.-H.; Bourg, I. C. Molecular dynamics simulations of mineral surface wettability by water versus CO2: Thin films, contact angles, and capillary pressure in a silica nanopore. The Journal of Physical Chemistry C 2020, 124, 25382–25395
- Emami et al. 2014 Emami, F. S.; Puddu, V.; Berry, R. J.; Varshney, V.; Patwardhan, S. V.; Perry, C. C.; Heinz, H. Force field and a surface model database for silica to simulate interfacial properties in atomic resolution. Chemistry of Materials 2014, 26, 2647–2658
- Lamb and Furlong 1982 Lamb, R. N.; Furlong, D. N. Controlled wettability of quartz surfaces. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases 1982, 78, 61–73
- Baidakov et al. 2017 Baidakov, V.; Tipeev, A.; Protsenko, K. Surface free energy and some other properties of a crystal–vapor interface: Molecular dynamics simulation of a Lennard-Jones system. The Journal of Chemical Physics Letters 2017, 680, 10–16
- Di Pasquale and Davidchack 2020 Di Pasquale, N.; Davidchack, R. L. Shuttleworth equation: A molecular simulations perspective. The Journal of Chemical Physics 2020, 153
- Wu and Firoozabadi 2021 Wu, T.; Firoozabadi, A. Calculation of solid–fluid interfacial free energy with consideration of solid deformation by molecular dynamics simulations. The Journal of Physical Chemistry A 2021, 125, 5841–5848
- Dreher et al. 2019 Dreher, T.; Pineau, N.; Bourasseau, E.; Malfreyt, P.; Soulard, L.; Lemarchand, C. Anisotropic surface stresses of a solid/fluid interface: Molecular dynamics calculations for the copper/methane interface. The Journal of Chemical Physics 2019, 151
- Velázquez et al. 2023 Velázquez, J. H.; Sánchez-Balderas, G.; Goicochea, A. G.; Pérez, E. The effective interfacial tensions between pure liquids and rough solids: a coarse-grained simulation study. Physical Chemistry Chemical Physics 2023, 25, 10325–10334
- Shuttleworth 1950 Shuttleworth, R. The surface tension of solids. Proceedings of the Physical Society. Section A 1950, 63, 444
- Makkonen 2012 Makkonen, L. Misinterpretation of the Shuttleworth equation. Scripta Materialia 2012, 66, 627–629
- Yamaguchi et al. 2019 Yamaguchi, Y.; Kusudo, H.; Surblys, D.; Omori, T.; Kikugawa, G. Interpretation of Young’s equation for a liquid droplet on a flat and smooth solid surface: Mechanical and thermodynamic routes with a simple Lennard-Jones liquid. The Journal of Chemical Physics 2019, 150
- Frenkel and Ladd 1984 Frenkel, D.; Ladd, A. J. New Monte Carlo method to compute the free energy of arbitrary solids. Application to the fcc and hcp phases of hard spheres. The Journal of Chemical Physics 1984, 81, 3188–3193
- Polson et al. 2000 Polson, J. M.; Trizac, E.; Pronk, S.; Frenkel, D. Finite-size corrections to the free energies of crystalline solids. The Journal of Chemical Physics 2000, 112, 5339–5342
- Vega and Noya 2007 Vega, C.; Noya, E. G. Revisiting the Frenkel-Ladd method to compute the free energy of solids: The Einstein molecule approach. The Journal of Chemical Physics 2007, 127
- Reddy A and Punnathanam 2018 Reddy A, R. K.; Punnathanam, S. N. Calculation of excess free energy of molecular solids comprised of flexible molecules using Einstein molecule method. Molecular Simulation 2018, 44, 781–788
- Broughton and Gilmer 1983 Broughton, J.; Gilmer, G. Surface free energy and stress of a Lennard-Jones crystal. Acta Metallurgica 1983, 31, 845–851
- Broughton and Gilmer 1986 Broughton, J. Q.; Gilmer, G. H. Molecular dynamics investigation of the crystal–fluid interface. IV. Free energies of crystal–vapor systems. The Journal of Chemical Physics 1986, 84, 5741–5748
- Tipeev et al. 2021 Tipeev, A. O.; Rino, J. P.; Zanotto, E. D. Direct determination of Lennard-Jones crystal surface free energy by a computational cleavage method. The Journal of Chemical Physics 2021, 155
- Smith and Lynden-Bell 1999 Smith, P.; Lynden-Bell, R. Determining surface free energies of crystals with highly disordered surfaces from simulation. Molecular Physics 1999, 96, 1027–1032
- Grochola et al. 2002 Grochola, G.; Russo, S. P.; Snook, I. K.; Yarovsky, I. On simulation methods to compute surface and interfacial free energies of disordered solids. The Journal of Chemical Physics 2002, 116, 8547–8555
- Modak et al. 2016 Modak, V. P.; Wyslouzil, B. E.; Singer, S. J. On the determination of the crystal-vapor surface free energy, and why a Gaussian expression can be accurate for a system far from Gaussian. The Journal of Chemical Physics 2016, 145
- Fan et al. 2021 Fan, J.; De Coninck, J.; Wu, H.; Wang, F. A generalized examination of capillary force balance at contact line: On rough surfaces or in two-liquid systems. Journal of Colloid and Interface Science 2021, 585, 320–327
- Firoozabadi 2016 Firoozabadi, A. Thermodynamics and applications in hydrocarbon energy production; McGraw-Hill Education, 2016
- Ruckenstein 1978 Ruckenstein, E. The origin of thermodynamic stability of microemulsions. Chemical Physics Letters 1978, 57, 517–521
- Okazawa and Bron 1979 Okazawa, T.; Bron, J. On thermodynamically stable emulsions: I. Thermodynamic Background. Journal of Colloid and Interface Science 1979, 69, 86–96
- Surblys et al. 2014 Surblys, D.; Yamaguchi, Y.; Kuroda, K.; Kagawa, M.; Nakajima, T.; Fujimura, H. Molecular dynamics analysis on wetting and interfacial properties of water-alcohol mixture droplets on a solid surface. The Journal of Chemical Physics 2014, 140
- Xu et al. 2022 Xu, J.; Jia, L.; Dang, C.; Liu, X.; Ding, Y. Effects of solid–liquid interaction and mixture concentration on wettability of nano-droplets: Molecular dynamics simulations. AIP Advances 2022, 12
- Leroy and Müller-Plathe 2015 Leroy, F.; Müller-Plathe, F. Dry-surface simulation method for the determination of the work of adhesion of solid–liquid interfaces. Langmuir 2015, 31, 8335–8345
- Nijmeijer and Leeuwen 1990 Nijmeijer, M.; Leeuwen, J. v. Microscopic expressions for the surface and line tension. Journal of Physics A: Mathematical and General 1990, 23, 4211
- Yen 2020 Yen, T.-H. Influence of gas aggregation on water-solid interface: molecular simulation. Molecular Simulation 2020, 46, 1373–1382
- Tang and Harris 1995 Tang, J. Z.; Harris, J. G. Fluid wetting on molecularly rough surfaces. The Journal of chemical physics 1995, 103, 8201–8208
- Nijmeijer and Bruin 1996 Nijmeijer, M.; Bruin, C. Comment on “Fluid wetting on molecularly rough surfaces”[J. Chem. Phys. 103, 8201 (1995)]. The Journal of Chemical Physics 1996, 105, 4889–4890
- Torres-Sánchez et al. 2016 Torres-Sánchez, A.; Vanegas, J. M.; Arroyo, M. Geometric derivation of the microscopic stress: A covariant central force decomposition. Journal of the Mechanics and Physics of Solids 2016, 93, 224–239
- Admal and Tadmor 2010 Admal, N. C.; Tadmor, E. B. A unified interpretation of stress in molecular systems. Journal of Elasticity 2010, 100, 63–143
- Nakamura et al. 2015 Nakamura, T.; Kawamoto, S.; Shinoda, W. Precise calculation of the local pressure tensor in Cartesian and spherical coordinates in LAMMPS. Computer Physics Communications 2015, 190, 120–128
- Shi et al. 2023 Shi, K.; Smith, E. R.; Santiso, E. E.; Gubbins, K. E. A perspective on the microscopic pressure (stress) tensor: History, current understanding, and future challenges. The Journal of Chemical Physics 2023, 158
- Nishida et al. 2014 Nishida, S.; Surblys, D.; Yamaguchi, Y.; Kuroda, K.; Kagawa, M.; Nakajima, T.; Fujimura, H. Molecular dynamics analysis of multiphase interfaces based on in situ extraction of the pressure distribution of a liquid droplet on a solid surface. The Journal of Chemical Physics 2014, 140
- Hamada et al. 2009 Hamada, Y.; Koga, K.; Tanaka, H. Phase behavior and fluid–solid surface tension of argon in slit pores and carbon nanotubes. Physica A: Statistical Mechanics and its Applications 2009, 388, 2289–2298
- Fujiwara and Shibahara 2014 Fujiwara, K.; Shibahara, M. Local pressure components and interfacial tension at a liquid-solid interface obtained by the perturbative method in the Lennard-Jones system. The Journal of Chemical Physics 2014, 141
- Dreher et al. 2018 Dreher, T.; Lemarchand, C.; Soulard, L.; Bourasseau, E.; Malfreyt, P.; Pineau, N. Calculation of a solid/liquid surface tension: A methodological study. The Journal of Chemical Physics 2018, 148
- Kaur and Deb 2021 Kaur, J.; Deb, D. Pressure-tensor method evaluation of the interfacial tension between Gay–Berne isotropic fluid and a smooth repulsive wall. Soft Matter 2021, 17, 10566–10579
- Theodorakis et al. 2015 Theodorakis, P. E.; Müller, E. A.; Craster, R. V.; Matar, O. K. Modelling the superspreading of surfactant-laden droplets with computer simulation. Soft Matter 2015, 11, 9254–9261
- Herdes et al. 2018 Herdes, C.; Petit, C.; Mejía, A.; Muller, E. A. Combined experimental, theoretical, and molecular simulation approach for the description of the fluid-phase behavior of hydrocarbon mixtures within shale rocks. Energy & Fuels 2018, 32, 5750–5762
- Lafitte et al. 2013 Lafitte, T.; Apostolakou, A.; Avendaño, C.; Galindo, A.; Adjiman, C. S.; Müller, E. A.; Jackson, G. Accurate statistical associating fluid theory for chain molecules formed from Mie segments. The Journal of Chemical Physics 2013, 139
- Ameri et al. 2013 Ameri, A.; Kaveh, N. S.; Rudolph, E.; Wolf, K.-H.; Farajzadeh, R.; Bruining, J. Investigation on interfacial interactions among crude oil–brine–sandstone rock–CO2 by contact angle measurements. Energy & Fuels 2013, 27, 1015–1025
- Abdulelah et al. 2021 Abdulelah, H.; Al-Yaseri, A.; Ali, M.; Giwelli, A.; Negash, B. M.; Sarmadivaleh, M. CO2/Basalt’s interfacial tension and wettability directly from gas density: Implications for Carbon Geo-sequestration. Journal of Petroleum Science and Engineering 2021, 204, 108683
- Pan et al. 2020 Pan, B.; Gong, C.; Wang, X.; Li, Y.; Iglauer, S. The interfacial properties of clay-coated quartz at reservoir conditions. Fuel 2020, 262, 116461
- Hosseini et al. 2022 Hosseini, M.; Ali, M.; Fahimpour, J.; Keshavarz, A.; Iglauer, S. Assessment of rock-hydrogen and rock-water interfacial tension in shale, evaporite and basaltic rocks. Journal of Natural Gas Science and Engineering 2022, 106, 104743
- Wilhelmy 1863 Wilhelmy, L. Ueber die Abhängigkeit der Capillaritäts-Constanten des Alkohols von Substanz und Gestalt des benetzten festen Körpers. Annalen der Physik 1863, 195, 177–217
- Stephan and Hasse 2020 Stephan, S.; Hasse, H. Influence of dispersive long-range interactions on properties of vapour–liquid equilibria and interfaces of binary Lennard-Jones mixtures. Molecular Physics 2020, 118, e1699185
- Watanabe et al. 2022 Watanabe, K.; Kusudo, H.; Bistafa, C.; Omori, T.; Yamaguchi, Y. Quantifying the solid–fluid interfacial tensions depending on the substrate curvature: Young’s equation holds for wetting around nanoscale cylinder. The Journal of Chemical Physics 2022, 156
- Rane et al. 2011 Rane, K. S.; Kumar, V.; Errington, J. R. Monte Carlo simulation methods for computing the wetting and drying properties of model systems. The Journal of Chemical Physics 2011, 135
- Kumar and Errington 2014 Kumar, V.; Errington, J. R. The use of Monte Carlo simulation to obtain the wetting properties of water. Physics Procedia 2014, 53, 44–49
- Errington 2003 Errington, J. R. Direct calculation of liquid–vapor phase equilibria from transition matrix Monte Carlo simulation. The Journal of Chemical Physics 2003, 118, 9915–9925
- Grzelak and Errington 2008 Grzelak, E. M.; Errington, J. R. Computation of interfacial properties via grand canonical transition matrix Monte Carlo simulation. The Journal of Chemical Physics 2008, 128
- Grzelak and Errington 2010 Grzelak, E. M.; Errington, J. R. Calculation of interfacial properties via free-energy-based molecular simulation: The influence of system size. The Journal of Chemical Physics 2010, 132
- Jain et al. 2019 Jain, K.; Rane, K. S.; Errington, J. R. Using isothermal-isobaric Monte Carlo simulation to study the wetting behavior of model systems. The Journal of Chemical Physics 2019, 150
- Jain et al. 2019 Jain, K.; Schultz, A. J.; Errington, J. R. Application of the interface potential approach for studying wetting behavior within a molecular dynamics framework. The Journal of Chemical Physics 2019, 150
- Jain et al. 2019 Jain, K.; Schultz, A. J.; Errington, J. R. Construction of the interface potential from a series of canonical ensemble simulations. The Journal of Chemical Physics 2019, 151
- Leroy et al. 2009 Leroy, F.; Dos Santos, D. J.; Müller-Plathe, F. Interfacial excess free energies of solid–liquid interfaces by molecular dynamics simulation and thermodynamic integration. Macromolecular Rapid Communications 2009, 30, 864–870
- Leroy and Müller-Plathe 2011 Leroy, F.; Müller-Plathe, F. Rationalization of the behavior of solid- liquid surface free energy of water in Cassie and Wenzel wetting states on rugged solid surfaces at the nanometer scale. Langmuir 2011, 27, 637–645
- Huang et al. 2001 Huang, D. M.; Geissler, P. L.; Chandler, D. Scaling of hydrophobic solvation free energies. The Journal of Physical Chemistry B 2001, 105, 6704–6709
- Uranagase et al. 2018 Uranagase, M.; Ogata, S.; Tanaka, K.; Mori, H.; Tajima, S. Efficient scheme for calculating work of adhesion between a liquid and polymer-grafted substrate. The Journal of Chemical Physics 2018, 149
- Uranagase and Ogata 2020 Uranagase, M.; Ogata, S. FE-CLIP: A tool for the calculation of the solid–liquid interfacial free energy. Computer Physics Communications 2020, 254, 107252
- Surblys et al. 2022 Surblys, D.; Müller-Plathe, F.; Ohara, T. Computing the work of solid–liquid adhesion in systems with damped Coulomb interactions via molecular dynamics: Approaches and insights. The Journal of Physical Chemistry A 2022, 126, 5506–5516
- Kanduč and Netz 2015 Kanduč, M.; Netz, R. R. From hydration repulsion to dry adhesion between asymmetric hydrophilic and hydrophobic surfaces. Proceedings of the National Academy of Sciences 2015, 112, 12338–12343
- Kanduč and Netz 2017 Kanduč, M.; Netz, R. R. Atomistic simulations of wetting properties and water films on hydrophilic surfaces. The Journal of Chemical Physics 2017, 146
- Frolov and Mishin 2009 Frolov, T.; Mishin, Y. Temperature dependence of the surface free energy and surface stress: An atomistic calculation for Cu (110). Physical Review B 2009, 79, 045430
- Laird et al. 2009 Laird, B. B.; Davidchack, R. L.; Yang, Y.; Asta, M. Determination of the solid-liquid interfacial free energy along a coexistence line by Gibbs–Cahn integration. The Journal of Chemical Physics 2009, 131
- Frolov and Mishin 2009 Frolov, T.; Mishin, Y. Solid-liquid interface free energy in binary systems: Theory and atomistic calculations for the (110) Cu–Ag interface. The Journal of Chemical Physics 2009, 131
- Heni and Löwen 1999 Heni, M.; Löwen, H. Interfacial free energy of hard-sphere fluids and solids near a hard wall. Physical Review E 1999, 60, 7057
- Deb et al. 2011 Deb, D.; Winkler, A.; Yamani, M. H.; Oettel, M.; Virnau, P.; Binder, K. Hard sphere fluids at a soft repulsive wall: A comparative study using Monte Carlo and density functional methods. The Journal of Chemical Physics 2011, 134
- Deb et al. 2012 Deb, D.; Wilms, D.; Winkler, A.; Virnau, P.; Binder, K. Methods to compute pressure and wall tension in fluids containing hard particles. International Journal of Modern Physics C 2012, 23, 1240011
- Schmitz and Virnau 2015 Schmitz, F.; Virnau, P. The ensemble switch method for computing interfacial tensions. The Journal of Chemical Physics 2015, 142
- Virnau et al. 2016 Virnau, P.; Schmitz, F.; Binder, K. The ensemble switch method and related approaches to obtain interfacial free energies between coexisting phases from simulations: a brief review. Molecular Simulation 2016, 42, 549–562
- Schilling and Schmid 2009 Schilling, T.; Schmid, F. Computing absolute free energies of disordered structures by molecular simulation. The Journal of Chemical Physics 2009, 131
- Schmid and Schilling 2010 Schmid, F.; Schilling, T. A method to compute absolute free energies or enthalpies of fluids. Physics Procedia 2010, 4, 131–143
- Mao et al. 1997 Mao, Y.; Bladon, P.; Lekkerkerker, H.; Cates, M. Density profiles and thermodynamics of rod-like particles between parallel walls. Molecular Physics 1997, 92, 151–159
- Hamada et al. 2007 Hamada, Y.; Koga, K.; Tanaka, H. Phase equilibria and interfacial tension of fluids confined in narrow pores. The Journal of Chemical Physics 2007, 127
- Das and Binder 2010 Das, S. K.; Binder, K. Does Young’s equation hold on the nanoscale? A Monte Carlo test for the binary Lennard-Jones fluid. Europhysics Letters 2010, 92, 26006
- Das and Binder 2011 Das, S. K.; Binder, K. Simulation of binary fluids exposed to selectively adsorbing walls: a method to estimate contact angles and line tensions. Molecular Physics 2011, 109, 1043–1056
- Eriksson 1969 Eriksson, J. C. Thermodynamics of surface phase systems: V. Contribution to the thermodynamics of the solid-gas interface. Surface Science 1969, 14, 221–246
- Cammarata 1994 Cammarata, R. C. Surface and interface stress effects in thin films. Progress in Surface Science 1994, 46, 1–38
- Haiss 2001 Haiss, W. Surface stress of clean and adsorbate-covered solids. Reports on Progress in Physics 2001, 64, 591
- Marichev 2009 Marichev, V. Current state and problems in the surface tension of solids. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2009, 345, 1–12
- Marichev 2011 Marichev, V. The Shuttleworth equation: Its modifications and current state. Protection of Metals and Physical Chemistry of Surfaces 2011, 47, 25–30
- Chipot and Pohorille 2007 Chipot, C.; Pohorille, A. Free energy calculations; Springer, 2007; Vol. 86
- Miyazaki et al. 1976 Miyazaki, J.; Barker, J.; Pound, G. A new Monte Carlo method for calculating surface tension. The Journal of Chemical Physics 1976, 64, 3364–3369
- Davidchack and Laird 2000 Davidchack, R. L.; Laird, B. B. Direct calculation of the hard-sphere crystal/melt interfacial free energy. Physical Review Letters 2000, 85, 4751
- Davidchack and Laird 2003 Davidchack, R. L.; Laird, B. B. Direct calculation of the crystal–melt interfacial free energies for continuous potentials: Application to the Lennard-Jones system. The Journal of Chemical Physics 2003, 118, 7651–7657
- Davidchack and Laird 2005 Davidchack, R. L.; Laird, B. B. Crystal structure and interaction dependence of the crystal-melt interfacial free energy. Physical Review Letters 2005, 94, 086102
- Laird and Davidchack 2005 Laird, B. B.; Davidchack, R. L. Direct calculation of the crystal-melt interfacial free energy via molecular dynamics computer simulation. 2005
- Davidchack 2010 Davidchack, R. L. Hard spheres revisited: Accurate calculation of the solid–liquid interfacial free energy. The Journal of Chemical Physics 2010, 133
- Benjamin and Horbach 2014 Benjamin, R.; Horbach, J. Crystal-liquid interfacial free energy via thermodynamic integration. The Journal of Chemical Physics 2014, 141
- Qi et al. 2016 Qi, X.; Zhou, Y.; Fichthorn, K. A. Obtaining the solid-liquid interfacial free energy via multi-scheme thermodynamic integration: Ag-ethylene glycol interfaces. The Journal of Chemical Physics 2016, 145
- Di Pasquale et al. 2024 Di Pasquale, N.; Davidchack, R.; Rovigatti, L. CLEAVING: a LAMMPS package to compute surface free energies. Journal of Open Source Software 2024, 9, 5886
- Addula and Punnathanam 2020 Addula, R. K. R.; Punnathanam, S. N. Computation of solid–fluid interfacial free energy in molecular systems using thermodynamic integration. The Journal of Chemical Physics 2020, 153
- Pedersen et al. 2013 Pedersen, U. R.; Hummel, F.; Kresse, G.; Kahl, G.; Dellago, C. Computing Gibbs free energy differences by interface pinning. Physical Review B 2013, 88, 094101
- Leroy and Müller-Plathe 2010 Leroy, F.; Müller-Plathe, F. Solid-liquid surface free energy of Lennard-Jones liquid on smooth and rough surfaces computed by molecular dynamics using the phantom-wall method. The Journal of chemical physics 2010, 133
- Patel et al. 2022 Patel, R.; Addula, R. K. R.; Shaik, M.; Punnathanam, S. N. Computing contact angles for oil–water–rock systems via thermodynamic integration. The Journal of Chemical Physics 2022, 157
- Morrow 1990 Morrow, N. R. Wettability and its effect on oil recovery. Journal of Petroleum Technology 1990, 42, 1476–1484
- Ding and Gao 2021 Ding, F.; Gao, M. Pore wettability for enhanced oil recovery, contaminant adsorption and oil/water separation: A review. Advances in Colloid and Interface Science 2021, 289, 102377
- Ma̧czyński et al. 2004 Ma̧czyński, A.; Wiśniewska-Gocłowska, B.; Góral, M. Recommended liquid–liquid equilibrium data. Part 1. Binary alkane–water systems. Journal of Physical and Chemical Reference Data 2004, 33, 549–577
- Einstein 1907 Einstein, A. Die plancksche theorie der strahlung und die theorie der spezifischen wärme. Annalen der Physik 1907, 327, 180–190
- Monson and Kofke 2000 Monson, P.; Kofke, D. Solid-fluid equilibrium: Insights from simple molecular models. Advances in The Journal of Chemical Physics 2000, 115, 113–179
- Vega et al. 2008 Vega, C.; Sanz, E.; Abascal, J.; Noya, E. Determination of phase diagrams via computer simulation: methodology and applications to water, electrolytes and proteins. Journal of Physics: Condensed Matter 2008, 20, 153101
- Reddy Addula et al. 2021 Reddy Addula, R. K.; Veesam, S. K.; Punnathanam, S. N. Review of the Frenkel–Ladd technique for computing free energies of crystalline solids. Molecular Simulation 2021, 47, 824–830
- Pretti and Mittal 2019 Pretti, E.; Mittal, J. Extension of the Einstein molecule method for solid free energy calculation to non-periodic and semi-periodic systems. The Journal of Chemical Physics 2019, 151
- Yeandel et al. 2022 Yeandel, S. R.; Freeman, C. L.; Harding, J. H. A general method for calculating solid/liquid interfacial free energies from atomistic simulations: Application to CaSO4· xH2O. The Journal of Chemical Physics 2022, 157
- Ballenegger 2014 Ballenegger, V. Communication: On the origin of the surface term in the Ewald formula. The Journal of Chemical Physics 2014, 140
- Ballenegger et al. 2009 Ballenegger, V.; Arnold, A.; Cerda, J. Simulations of non-neutral slab systems with long-range electrostatic interactions in two-dimensional periodic boundary conditions. The Journal of Chemical Physics 2009, 131
- Chipot and Pohorille 2007 Chipot, C.; Pohorille, A. Free energy calculations: theory and applications in chemistry and biology; Springer, 2007; pp 33–75
- Eppenga and Frenkel 1984 Eppenga, R.; Frenkel, D. Monte Carlo study of the isotropic and nematic phases of infinitely thin hard platelets. Molecular Physics 1984, 52, 1303–1334
- Harismiadis et al. 1996 Harismiadis, V.; Vorholz, J.; Panagiotopoulos, A. Efficient pressure estimation in molecular simulations without evaluating the virial. The Journal of Chemical Physics 1996, 105, 8469–8470
- Vörtler and Smith 2000 Vörtler, H. L.; Smith, W. R. Computer simulation studies of a square-well fluid in a slit pore. Spreading pressure and vapor–liquid phase equilibria using the virtual-parameter-variation method. The Journal of Chemical Physics 2000, 112, 5168–5174
- de Miguel and Jackson 2006 de Miguel, E.; Jackson, G. The nature of the calculation of the pressure in molecular simulations of continuous models from volume perturbations. The Journal of Chemical Physics 2006, 125
- De Miguel and Jackson 2006 De Miguel, E.; Jackson, G. Detailed examination of the calculation of the pressure in simulations of systems with discontinuous interactions from the mechanical and thermodynamic perspectives. Molecular Physics 2006, 104, 3717–3734
- Brumby et al. 2011 Brumby, P. E.; Haslam, A. J.; De Miguel, E.; Jackson, G. Subtleties in the calculation of the pressure and pressure tensor of anisotropic particles from volume-perturbation methods and the apparent asymmetry of the compressive and expansive contributions. Molecular Physics 2011, 109, 169–189
- Jiménez-Serratos et al. 2012 Jiménez-Serratos, G.; Vega, C.; Gil-Villegas, A. Evaluation of the pressure tensor and surface tension for molecular fluids with discontinuous potentials using the volume perturbation method. The Journal of Chemical Physics 2012, 137
- Brumby et al. 2017 Brumby, P. E.; Wensink, H. H.; Haslam, A. J.; Jackson, G. Structure and interfacial tension of a hard-rod fluid in planar confinement. Langmuir 2017, 33, 11754–11770
- Fujiwara and Shibahara 2015 Fujiwara, K.; Shibahara, M. Local pressure components and interfacial tensions of a liquid film in the vicinity of a solid surface with a nanometer-scale slit pore obtained by the perturbative method. The Journal of Chemical Physics 2015, 142
- Gloor et al. 2005 Gloor, G. J.; Jackson, G.; Blas, F. J.; de Miguel, E. Test-area simulation method for the direct determination of the interfacial tension of systems with continuous or discontinuous potentials. The Journal of Chemical Physics 2005, 123
- Rasing and Musevic 2013 Rasing, T.; Musevic, I. Surfaces and interfaces of liquid crystals; Springer Science & Business Media, 2013
- Miqueu et al. 2011 Miqueu, C.; Miguez, J. M.; Pineiro, M. M.; Lafitte, T.; Mendiboure, B. Simultaneous application of the gradient theory and Monte Carlo molecular simulation for the investigation of methane/water interfacial properties. The Journal of Physical Chemistry B 2011, 115, 9618–9625
- Staubach and Stephan 2022 Staubach, J.; Stephan, S. Interfacial properties of binary azeotropic mixtures of simple fluids: Molecular dynamics simulation and density gradient theory. The Journal of Chemical Physics 2022, 157
- Míguez et al. 2012 Míguez, J.; Piñeiro, M.; Moreno-Ventas Bravo, A.; Blas, F. On interfacial tension calculation from the test-area methodology in the grand canonical ensemble. The Journal of Chemical Physics 2012, 136
- Blas and Mendiboure 2013 Blas, F. J.; Mendiboure, B. Extension of the Test-Area methodology for calculating solid-fluid interfacial tensions in cylindrical geometry. The Journal of Chemical Physics 2013, 138
- Ghoufi et al. 2008 Ghoufi, A.; Goujon, F.; Lachet, V.; Malfreyt, P. Expressions for local contributions to the surface tension from the virial route. Physical Review E 2008, 77, 031601
- Ghoufi and Malfreyt 2012 Ghoufi, A.; Malfreyt, P. Calculation of the surface tension and pressure components from a non-exponential perturbation method of the thermodynamic route. The Journal of Chemical Physics 2012, 136
- Nair and Sathian 2012 Nair, A. R.; Sathian, S. P. A molecular dynamics study to determine the solid-liquid interfacial tension using test area simulation method (TASM). The Journal of Chemical Physics 2012, 137
- d’Oliveira et al. 2017 d’Oliveira, H.; Davoy, X.; Arche, E.; Malfreyt, P.; Ghoufi, A. Test-area surface tension calculation of the graphene-methane interface: Fluctuations and commensurability. The Journal of Chemical Physics 2017, 146
- Wu and Firoozabadi 2022 Wu, T.; Firoozabadi, A. Reply to “Comment on ‘Calculation of Solid–Fluid Interfacial Free Energy with Consideration of Solid Deformation by Molecular Dynamics Simulations’”. The Journal of Physical Chemistry A 2022, 126, 1784–1786
- Ghoufi 2024 Ghoufi, A. Surface free energy calculation of the solid–fluid interfaces from molecular simulation. AIP Advances 2024, 14
- Dong 2021 Dong, W. Thermodynamics of interfaces extended to nanoscales by introducing integral and differential surface tensions. Proceedings of the National Academy of Sciences 2021, 118, e2019873118
- Gim et al. 2018 Gim, S.; Lim, H.-K.; Kim, H. Multiscale simulation method for quantitative prediction of surface wettability at the atomistic level. The Journal of Physical Chemistry Letters 2018, 9, 1750–1758
- Gim et al. 2019 Gim, S.; Cho, K. J.; Lim, H.-K.; Kim, H. Structure, dynamics, and wettability of water at metal interfaces. Scientific Reports 2019, 9, 14805
- Müller and Mejía 2014 Müller, E. A.; Mejía, A. Resolving discrepancies in the measurements of the interfacial tension for the CO2+ H2O mixture by computer simulation. The Journal of Physical Chemistry Letters 2014, 5, 1267–1271
- Stephan et al. 2023 Stephan, S.; Fleckenstein, F.; Hasse, H. Vapor–Liquid Interfacial Properties of the Systems (Toluene+ CO2) and (Toluene+ N2): Experiments, Molecular Simulation, and Density Gradient Theory. Journal of Chemical & Engineering Data 2023, 69, 590–607
- Yang et al. 2023 Yang, Y.; Wan, J.; Li, J.; Zhao, G.; Shang, X. Molecular modeling of interfacial properties of the hydrogen+ water+ decane mixture in three-phase equilibrium. Physical Chemistry Chemical Physics 2023, 25, 29641–29655
- Pollice et al. 2021 Pollice, R.; dos Passos Gomes, G.; Aldeghi, M.; Hickman, R. J.; Krenn, M.; Lavigne, C.; Lindner-D’Addario, M.; Nigam, A.; Ser, C. T.; Yao, Z.; others Data-driven strategies for accelerated materials design. Accounts of Chemical Research 2021, 54, 849–860
- Wang et al. 2023 Wang, X. Q.; Chen, P.; Chow, C. L.; Lau, D. Artificial-intelligence-led revolution of construction materials: From molecules to Industry 4.0. Matter 2023, 6, 1831–1859















TOC Graphic