1. Introduction
In order to rationally design adsorption, catalytic, or separation processes using porous materials it is necessary to understand and predict the rates of mass transfer fluxes within the porous network. The specific surface areas of typical mesoporous adsorbent or catalyst pellets are generally of the order of hundreds of metres squared per gram, while microporous materials, such as zeolites and carbons, can be of the order of ~1000 m
2 g
−1. Given the large surface areas available for adsorption, in many adsorbent pellets, heterogeneous catalysts, and separation membranes the mass transport flux is dominated by surface diffusion. For example, it has been observed that, for ethane, propane, and butane adsorbed on a porous silica with a pore size of ~2 nm, the contribution to the flux from surface diffusion was from 60 to 90%, while for a similarly pore-sized carbon it was 70–75% [
1,
2,
3]. Simulations have also shown that the interaction between surface and (bulk) pore diffusion affects both diffusivities [
4], since the interchange between motion on the surface and within the bulk of the pore core affects the pathway taken between any two points.
While the influence of surface coverage or occupancy on the surface diffusion flux has been much studied [
5,
6,
7], the impact of the pore space and surface geometry on surface diffusion has received relatively little attention in the literature [
8]. However, many of the materials, such as silica, alumina, and carbons, used in industrial processes where surface diffusion is important are amorphous, and thus the interior pore networks are highly disordered. The surfaces of amorphous materials are also highly complex, and, thence, difficult to model. The atoms of surfaces are not located at random positions but tend to take up positions of minimum energy [
9]. Further, these positions can relax following the first formation of the surface by a process like fracturing, the abrupt termination of particle growth, or cooling from melts, but also during the adsorption and diffusion of adsorbates. The most detailed models of surfaces are those that utilise potential energy functions to obtain radial structure functions and bond angle distributions for surface atoms. However, relatively few models are constructed with this high, “maximally realistic” level of detail because, first, it is hard to characterise real amorphous materials in sufficient detail, and, second, the computing power required to represent a substantial area of surface in such detail may be limiting. Hence, some level of simplification and/or idealisation is necessary. The process known as “Galilean idealisation” involves introducing omissions from, or distortions of, the original structure of the surface in order to achieve sufficient simplification such that the resulting model obtained is then mathematically or computationally tractable, both for aligning with characterisation data and using it to make predictions of physical processes occurring on the surface [
10]. As mathematical and computational techniques have become more sophisticated over time, the level of simplification required in Galilean idealisation has diminished. However, it is typically not possible to know in advance which features of the real material to leave out and still keep a model predictive of the physical process of interest, and an empirical approach is, thus, needed. The underlying energetics and connectivity of a potential model surface can be tested indirectly via processes that involve the chemical interactions of the surface with specific molecules, such as predictions of the heat of adsorption, and the activation energy for surface diffusion. A different approach to Galilean idealisation is to actively only include in the model those features of the surface structure that are known to ‘make a difference to the occurrence or essential character of the phenomenon in question’ (which in this case is surface diffusion) [
10]. This approach is known as minimalist modelling, and requires some sort of theoretical justification, prior experimentation, or trial and error testing, to determine these essential characteristics of the surface.
Heterogeneities in disordered porous materials can be of many different kinds, and each may have a characteristic length scale associated with it called the correlation length, such that above that size the structure is statistically homogeneous and the property concerned is constant [
11]. For surface diffusion in such a heterogeneous system, at the very smallest length scales, where a molecule hops to a location in close vicinity to its initial location, it will obey the classical law that the mean square displacement scales with time as <
r2> ∝
t. It will also obey this law for root-mean-square (rms) displacements much larger than the correlation length. However, at intermediate length scales, the surface diffusion may be anomalous, and rms displacement may scale with a power of time different to the classical law. The importance of key length scales within amorphous porous structures to surface diffusion will become apparent below.
The diffusion of molecules across disordered, heterogeneous surfaces is also a complex process, and several theories for the fundamental nature of the motion occurring have been presented. Theories of diffusion for molecules physisorbed onto surfaces generally fall into one of three classes, namely [
12]: (1) mechanistic models in which the migrating molecule is described by the trajectory of a “hopping” molecule; (2) a two-dimensional Fick’s law with modifications, in which the surface flow is caused by a gradient in the surface concentration; and (3) hydrodynamic models in which the surface flow is treated as a two-dimensional fluid slipping along the walls of the pores.
Mechanistic models [
13] involve modelling the movement of adsorbed molecules from site to site as a random walk. Small hopping movements of the molecules on the surface lead to their migration, and the hopping is assumed to be an activated process. The hops can take place in random directions but the presence of an overall gradient in concentration in the adsorbed phase in one direction will lead to a net flux in that direction. This type of model requires assumptions to be made concerning the nature of the hopping motion, such as whether it is to a nearest-neighbour site only or involves a longer ‘flight’. In addition, considerations concerning the impact of surface coverage on the blocking of available sites and the hopping motion must also be included. For example, the jump lengths are often assumed to decrease with increased surface coverage due to inter-molecular collisions becoming more common [
14]. Approximate non-equilibrium statistical mechanical approaches have been used wherein the adsorbed phase is treated as a two-dimensional fluid [
15,
16]. However, MacElry and Raghavan [
17] suggested that this approach is probably most suitable for sub-monolayer coverages of macroporous materials at low temperature, but not for transport in micropores since the pore sizes and topologies in microporous materials mean that the two-dimensional fluid is not a good representation therein.
Increased computing power has permitted molecular dynamics simulations of the coupled migration and adsorption processes within models of porous media. In molecular dynamics, Newton’s Laws of Motion are explicitly solved for individual molecules. This permits the tracking of the pathways of individual molecules, as well as the whole collective. It will be seen that this possibility has provided great insights into the interaction between the nature of the pore structure and the surface diffusion processes therein.
The nature of the mathematical method for describing surface diffusion can determine the type of representation of a surface used for modelling. One way to improve the tractability of the model surface is to reduce the degree of randomness in the model as it reduces the number of parameter fields necessary to store. Further, the upper limit placed on the number of adsorbate molecules that can be handled by the available computing power also puts a limit on the size of the unit cell of the model that can be simulated, and thus the size of the region of surface that can be considered. This limitation can be partially overcome by averaging the results of simulations of adsorbate behaviour over many realisations of the solid model. The atomistic simulation of amorphous metal oxide, such as alumina, surfaces presents significant problems because (a) the charges of the individual aluminium and oxygen atoms depend upon their specific location and surrounding atoms, (b) there are variations in the local atomic structure between the interface and the bulk, and (c) there are difficulties in the atomistic representation at the buried interface [
18].
The complexity of the structure of amorphous surfaces can be represented in a variety of ways and characterised by a similar variety of parameters or descriptors, which can then be correlated with surface diffusivity. It will be seen that surfaces can be defined by parameters such as porosities (voidage fractions), pore sizes (determining surface curvature and separation distances of pore walls), surface roughness factors, surface fractal dimensions, and energy distributions of barriers and traps. However, due to the difficulty in characterising amorphous materials in detail, an alternative, to structural characterisation and then the building of a matching model, is to attempt to simulate the formation process of the surface to create a resultant model to perform the surface transport simulations upon. The surface formation models also demonstrate a range of different levels of complexity in describing the physical processes involved, which, naturally, also determines the level of complexity of the description of the resultant surface. It is the aim of this work to survey the various ways in which descriptions of surface geometric and/or chemical heterogeneity have been incorporated into models of surface diffusion.
2. Random Packing and Surface Reconstruction Models
Surface and (bulk) pore diffusion can be considered to occur in parallel, and thus mathematical modelling schemes then omit any interaction between the two. In much past work, it has been tacitly assumed that the mean path length through a porous solid is the same for a bulk gas molecule in the pore void and for a surface diffusing molecule (as would be the case for straight, cylindrical pores). However, Horiguchi et al. [
19] realised that an amorphous solid surface will possess irregularities of molecular dimensions and above that would not be perceived by a gas molecule diffusing in the Knudsen regime, and, thus, the path followed by a surface diffusing molecule will be potentially different and longer than for a bulk molecule. Hence, Horiguchi et al. [
19] proposed that the pore tortuosity τ must be augmented by a surface roughness factor
K to give a separate surface tortuosity:
They suggested that the pore roughness factor ought to be larger for, say, Vycor glass surfaces compared with, say, Graphon carbon, due to the chemical and energetic heterogeneity of the former compared to the relatively smooth and uniform surface of the latter [
19]. However, this roughness factor was purely empirical and not very informative.
In the case of parallel bulk pore and surface diffusion, the impact of the statistical properties of the porous medium, namely the voidage fraction
ϕ and average pore radius
re, on the surface diffusion and pore diffusion can be incorporated into separate effective surface tortuosity
τs and effective pore (void) tortuosity
τp. Hence, in this context, the mass transport flux
J per unit total area is given by the following [
20]:
where the reference surface diffusivity
Ds is that of a single molecule diffusing on a well-defined flat surface, and the bulk diffusivity
D applies in the void of the pore. It is also further typically assumed that a local equilibrium is rapidly reached between the bulk fluid concentration
c and the concentration of adsorbed phase on the surface
cs, which at low coverages would be characterised by a Henry’s law relation. Ho and Strieder [
20] defined the effective diffusivity as the ratio of the flux to the concentration gradient, and derived an expression for the effective diffusivity for a porous material with random, tortuous paths. In order to do this, they formulated the processes of void and surface diffusion locally for each infinitesimal void volume and pore wall element. The total flux, and effective diffusivity, was calculated for a model pore structure consisting of randomly placed, freely overlapping solid spheres. By employing variational principles, the calculations for the model were said to properly contain the tortuosity of the surface and void space pathways by including the perturbation in the flux due to the presence of the solid spheres. Hence, Ho and Strieder [
20] obtained an algebraic expression for the surface tortuosity
τs of a random packing of overlapping spheres:
where it is a function of the voidage fraction
ϕ. The necessity for only the single structural parameter
ϕ to be specified in this expression probably reflects the relative simplicity of the structure of a random packing of well-defined spheres compared to a more complex model surface (as will be seen below).
The assumption of parallel in-pore and surface diffusion mentioned above has been questioned [
21]. Wernert et al. [
21] proposed an alternative approach considering the Brownian motion of molecules in the pore space. These authors also highlighted the issue of whether the surface diffusion is considered to occur on a locally flat surface such that the overall tortuosity is shared with the pore diffusion as arising purely from the pore network interconnectivity, or whether the local surface curvature contribution is also considered. In the former case, this approximation may hold if a surface diffusion stage is short relative to the adjacent pore diffusion stages, as would likely arise with relatively little adsorption.
Based upon a suggestion by Havlin and Ben Avraham [
22], that the geometrical part (the nature of the lattice) of the disorder of random materials is irrelevant to surface diffusion, and only energetic heterogeneity matters, Limoge and Bocquet [
23] constructed a surface model consisting of a regular lattice where the energetic properties were distributed at random. The lattice both the sites and saddle points between sites were assigned characteristic energies. Hence, the model had two extreme types of potential disorder. The first is disorder in the site energy distribution, which corresponds to the so-called random-trapping model, and the second is the saddle disorder, which corresponds to the (activated) random hopping model. Further, Limoge and Bocquet [
23] suggested that, in real amorphous materials, there may be some correlations between values of the energies of adjacent sites and saddle points. The simulation of surface diffusion on the model consisted of random walks of tagged particles performing thermally activated jumping between the lattice sites. The Einstein relation was used to determine the surface diffusivity. Based upon the analysis, using the continuous-time random walk (CTRW) approach, Limoge and Bocquet [
23] found that the disorder in the site and saddle energies could compensate for each other. Complementary Monte Carlo simulations found that the presence of both types of disorder, in lattices with a range of coordination numbers from 4–12, could produce a compensation effect with the site disorder decreasing, through trapping, the infinite temperature limit of the surface diffusivity, and the saddle disorder increasing it. However, Limoge and Bocquet [
23] suggested that, in real amorphous materials, there is probably some correlation between the disorder of the sites and saddles due to the particular structure of the material, and they cannot be taken as independent variables, and thereby capable of varying freely in such a way as to compensate for each other as in the model.
A surface model that takes into account the potential correlation between bond and site disorder is the Dual Site-Bond Model (DSBM) [
24,
25]. This consists of a two-dimensional lattice of given connectivity, such as a square lattice, composed of adsorption sites, that are local potential wells, joined by bonds, which are saddle points in energy between adjoining wells, as seen in
Figure 1. Both the sites and the bonds have their own respective energy density functions. However, these two energy distribution functions for bonds and sites are not completely independent because the absolute energy of a site must be equal to, or higher than, all of its neighbouring bonds. This additional constraint is known as the Construction Principle [
25]. Hence, the bond energy distribution curve must always lie to lower energy than the site distribution curve. If there is no overlap between the two distributions, such that all bonds are lower in absolute energy than all of the sites, then a completely random surface results. Progressively increasing the overlap of the energy distribution necessitates there is some weak spatial correlation between the distribution of bond and site energies, and more homogeneous patches develop. Surface diffusion in such a model can be simulated using Monte Carlo techniques. The surface diffusivity has been found to increase as the degree of overlap of the two energy distributions increases. When the impact of varying temperature on surface diffusivity was tested, and the diffusivities correlated via Arrhenius expressions, it was found that deviations from the Arrhenius law occurred at lower temperatures due to adsorbate molecules becoming trapped in deep sites, and grew with an increased overlap in the energy distributions due to the increased spatial correlation of these deep sites into extensive patches [
24].
Argyrakis et al. [
26] compared the predictions, for the long time and high-temperature limit of the diffusivity for surface-hopping transport on random barrier models, with different distributions of barrier energies, made using critical path theory [
27], the effective medium approximation (EMA), and Monte Carlo methods. At longer times, the diffusivity had an Arrhenius temperature dependence with a high temperature-limiting value. For symmetrical energy distributions on 2D square grids, it was found that, for both the critical path theory and the EMA, the limiting diffusion coefficient just depended upon the mean energy barrier, and not the spread in the distribution. This was because the percolation threshold for 2D square grids is 0.5 (i.e., corresponding then to the mean for a symmetrical distribution). However, for 3D grids, the limiting diffusivity did depend upon the spread of barrier energies. The theoretical results were confirmed using the numerical Monte Carlo methods.
MacElroy and Raghavan [
28] suggested that the use of models of solids with smooth surfaces leads to uncertainties in the comparative analysis between theory and experiment. More realistic models can be constructed using molecular modelling. Stallons and Iglesia [
9] proposed that it is necessary to use molecular modelling in order to be able to include the detailed topological connectivity between surface atoms, or between surface atoms and the bulk atoms immediately below, to better capture the heterogeneity of amorphous surfaces.
Electron microscopy and atomic force microscopy (AFM) studies (see
Figure 2) show that silicas made by the sol–gel method, from colloidal suspensions, consist of packings of small spheres [
29]. Random packings of spheres models come in a variety of types. The simplest consists of a random assembly of spheres where they are placed at random in space, either allowing overlapping (as mentioned above) or not. Overlapping sphere models can represent the results of processes such as sintering. Other types of assemblage include where the spheres do not overlap but are all constrained to be connected to (via touching) each other [
17,
28]. This system is generated from randomly placed spheres by an algorithm akin to diffusion-limited aggregation [
17]. The size of the individual spheres used, and the overall density of spheres in the assemblage, are typically chosen such that the porosity and specific surface area of the model matches that for real silicas. For example, MacElry and Raghavan [
17] had silica spheres of diameter 2.75 nm. However, the silica spheres in a random, non-overlapping system often do not form a percolating network with a continuous surface for diffusion, but, instead, have many high-energy barriers to surface diffusion across empty gaps between spheres. Hence, the topology of the model makes a big difference to surface mobility. For example, Olivares and Aarão Reis [
30] conducted simulations of tracer diffusion in a model porous medium consisting of a packing of spheres situated on a simple cubic lattice. Since the adjoining spheres in the model were considered to have a small contact area, these workers found that the extensive connectivity of the surface of their model solid enabled the tracers to migrate long distances across the surface. This contrasts with models described above where spheres were isolated or poorly connected.
Zalc et al. [
29,
31] used Monte Carlo simulations to study the impact of pore-space structure on surface diffusion on a model consisting of a partially-overlapped random-loose sphere packing with a voidage fraction less than 0.5. In the case of no exchange with the bulk, they found large surface tortuosities at the void- and solid-percolation thresholds. In comparison, Wolf and Streider [
32] derived formulae for how surface tortuosity τ
s varied with voidage fraction ϕ for a model of two-dimensional random fibres consisting of overlapping parallel circular cylinders with different radii (aligned perpendicular to the average flux), such that:
They assumed that pore and surface diffusion were independent and in parallel. These workers found that, relative to random beds of overlapping spheres, the surface flux was higher for the dilute fibre packings compared to the spheres.
However, experimental results suggest that the coupling of surface and bulk (pore) diffusion can significantly change the overall diffusion characteristics [
33,
34]. As a result, Mirbagheri and Hill [
35] applied the model of Albaalbaki and Hill [
36] that includes this coupling to study the effects of voidage fraction and pore size on surface diffusion within models consisting of arrays of solid spheres and spherical cavities. They found that, for solid sphere arrays, the permeability was controlled by surface diffusion for voidage fractions below the solid-percolation threshold, while for spherical cavities, permeability was limited by surface diffusion at voidage fractions between the void- and solid-percolation thresholds. For similar studies of potentially coupled surface and pore diffusion in models consisting of parallel pore and fibre models, Mirbagheri and Hill [
37] found that surface diffusion was mainly controlled by specific surface area, especially in the parallel.
Molecular modelling of energetically and geometrically heterogeneous silica surfaces can be performed by a number of methods. MacElroy and Raghavan [
28] created a molecular model of the bulk vitreous state by starting with crystalline β-cristobalite, and melting it at 10,000 K, and then quenching at a series of lower temperatures back down to 300 K using a Monte Carlo based methodology. A given sphere from a packing model was then cut from the bulk silica model. Any silicon atoms on the surface were discarded. The cutting-out also created numerous non-bridging oxygen atoms at the surface of the sphere which were then considered to form surface hydroxyl groups. However, any trisilanol groups formed were subsequently removed. The concentration of hydroxyl groups formed in this way was equivalent to 6.3 oxygens per nm
2, which is similar to that found experimentally for untreated silica [
28]. The simulated silica surface contained both singlet hydroxyls and twin (doublet) non-bridging hydroxyl (known as geminal) groups. The motion of dilute (structureless) methane molecules across the surface of the silica spheres was modelled using a Monte Carlo method. The interactions between the methane and individual oxygen atoms in the silica framework were modelled using a Lennard-Jones (12-6) potential. Apart from considering the oxygen atoms of hydroxyl groups as having a slightly larger size, in the Lennard-Jones interactions, than bridging (between silicon) oxygen atoms, the hydrogen atoms of the hydroxyl groups were neglected. Microcanonical (NVE) ensemble molecular dynamics was used to calculate the trajectories of adsorbed molecules within the model. These can then be used to calculate the mean square displacements of the molecules, and, thence, diffusivity. However, MacElroy and Raghavan [
17,
28] did not explicitly differentiate between the surface and pore diffusion. When comparing molecular behaviour upon assemblies of spheres where the surface potential was smoothed versus spheres with inhomogeneous potential fields, MacElroy and Raghavan [
28] found the latter led to more backscattering and localisation of adsorbed molecules. At low loadings, for simulations of three-component (one of which did not adsorb) mixtures on silica surfaces, MacElroy and Raghavan [
28] found surface blocking behaviour by adsorbed molecules not unlike that postulated by surface hopping kinetic models. At higher coverages, the surface transport mechanism tended more toward a hydrodynamic regime, but small pore sizes limit the applicability of a continuum fluid model. MacElroy and Raghavan [
28] found that, even with a molecular model surface, while it did not possess long-range order, it did have some short-range order associated with the triangle of surface oxygen atoms in an exposed SiO
4 tetrahedron. Further, adsorbate molecules experienced a potential energy minimum at the centre of such triangles, and these minima tended to have similar spacings from each other. Hence, the molecular model exhibited a geometric feature similar to the lattices used in barrier-hopping models.
Models of amorphous surfaces can also be formed from the exposed face of a dense packing of hard spheres, known as the Bernal surface model [
38]. The hard spheres are taken to represent the oxygen atoms in an oxide surface, such as titania, and are replaced by sites having Lennard-Jones interactions with adsorbate molecules [
38]. Cations are not included as they have relatively weak interactions with adsorbates. The surface can be further roughened by the random deletion of surface ions, to form Corrugated Bernal Surfaces (CBSs). Surfaces of ~1520 atoms were constructed but periodic boundary conditions were also used to attempt to reduce the limitation caused by a small unit cell. The overall size of the simulated unit cell was (57 × 54) Å
2. Riccardo and Steele [
38] suggested that such surfaces, as seen in
Figure 3, are similar to those seen in scanning tunnelling microscopy images. They also evaluated the surface fractal dimension
Ds of the model surfaces by measuring the apparent monolayer (ML) capacity
nm for adsorbate molecules with different sizes
σa. For fractal surfaces, the apparent monolayer capacity will scale with the adsorbate size according to the following [
39]:
Since a monolayer is not strictly defined for this system, for the purposes of the simulation, an adsorbate molecule was considered to be in the monolayer if it directly touched a surface atom. It was found that the BS led to a surface fractal dimension of ~2.2, and the progressive removal of surface atoms for the CBS led to surfaces with fractal dimensions up to ~2.5, which are in the range found for many types of amorphous materials such as carbons, silica, and alumina [
40].
The surface dynamics of 260 adsorbed argon atoms were simulated using an isokinetic molecular dynamics (MD) algorithm where the kinetic energy of the adsorbed molecules was kept constant, equivalent to being isothermal. The surface itself was considered rigid. The intermolecular interactions between the adsorbate molecules were considered negligible, to approximate the dilute (~zero coverage) limit. The surface diffusivity was calculated from the simulated trajectories of the adsorbed molecules using the Einstein relation. The trajectories of the argon atoms were also used to sample the distribution of maxima, minima, and barriers in the potential energy surface, and compare it to that of the Dual Site-Bond Model (DSBM) described above. Simulations were performed at several different temperatures and the surface diffusivities obtained correlated with the Arrhenius expression. The reduced activation energies for surface diffusion (from 3.5 to 6) thereby obtained for the BS, and the CBS after deleting 0.5 ML, were both similar to that (~3.15) assumed by Drain and Morrison [
41] for barriers hindering the surface migration of Ar on powdered titania. Hence, this finding also suggested that relatively low levels of roughness, as for the BS and 0.5 ML CBS, did not affect the activation energy. However, the reduced activation energy (~5.3) was much higher for much rougher surfaces, like the 3 ML CBS. This corresponded to an increase in diffusion activation energy from ~4.7 to 7.1 kJ/mol for rougher surfaces. This finding is more in line with experimental results by other workers [
42]. Further, deviations from pure Arrhenius behaviour were observed. Riccardo and Steele [
38] suggested this result was similar to previous findings for activated diffusion on energetically heterogeneous lattices modelled as sets of adsorption sites separated by saddle points, as in the DSBM [
43,
44,
45]. A common descriptor of surfaces is the energy distribution of local minima in the surface potential. Bakaev and Steele [
46] obtained this distribution for a BSM by sampling the adsorbate–solid interaction energy of argon atoms simulated in a full monolayer. However, in this method, atoms were forced onto sites that they would not necessarily occupy if isolated because the ML was found to be almost close-packed. Riccardo and Steele [
38] went further and obtained additional descriptors of the surface. They obtained the frequency distribution of adsorption energies sampled along MD trajectories, and also the frequency distributions of the maxima, minima, and activation barriers of the argon–solid interaction encountered along the MD trajectories. It was found that the rougher surfaces had broader distributions, which were also shifted towards higher energies, than smoother surfaces. These higher energies with larger fractal dimension are in line with the theory by Rigby [
47,
48,
49] described in
Section 3.2. Further, Riccardo and Steele [
38] found that the distributions of local maxima and minima overlapped, suggesting some local maxima along a given trajectory had a lower energy than some of the minima, but, clearly, could not arise in succession. Instead, there must have been a high degree of spatial correlation between the energies of successive extremes. This feature is similar to the scenario proposed by the DSBM, described above, and suggests that this model would be representative of the surfaces of amorphous materials.
Since, as mentioned above, it is not possible to know in advance the required level of idealisation for a model necessary to both achieve computational tractability and retain empirical adequacy, it is necessary to actively test models with different degrees of simplification. Stallons and Iglesia [
9] compared the structural and transport properties of four models for the surface of amorphous silica: (1) an ordered surface created by cutting the structure of bulk α-cristobalite to the desired geometry and making dangling oxygen bonds into hydroxyl groups; (2) an unrelaxed surface obtained by cutting a bulk silica structure created via molecular dynamics; (3) a relaxed surface obtained using the same molecular dynamics simulation used to create a bulk structure; and (4) a surface created by fracturing a random sphere packing, generated by Monte Carlo methods, where the spheres represented oxygen atoms of the silica. The various surface hydroxyl concentrations (~6–8 per nm
2) for the first three types of surface were at the high end of experimentally measured values, while the fourth type of model does not represent the silicon atoms needed for the calculation of surface hydroxyl concentration.
Predictions of the heat of adsorption for different adsorbates can be used to test the energetics of a surface model. However, potential energy functions that accurately describe the specific surface–adsorbate interactions are not typically available. Hence, it is necessary to use an interaction potential involving experimentally accessible parameters, such as atomic size and polarizability, and Henry’s law constants. Stallons and Iglesia [
9] used a Lennard-Jones potential to describe interactions between adsorbates (dinitrogen, argon, and methane) and the two types of surface oxygen atom (hydroxyl and bridging oxygens). The interactions between adsorbates and silicon atoms were left out, since these were considered negligibly small relative to those of oxygen atoms. Potential energy surfaces, based upon a 0.2 Å pitch grid, were constructed for the four types of models and the adsorbate nitrogen, and those for the relaxed and unrelaxed surfaces are shown in
Figure 4. From
Figure 4, it can be seen that the relaxed silica surface had better defined peaks and valleys than the unrelaxed surface. The random surface had more localized and better-defined adsorption sites than the unrelaxed surface, but the potential energy surface was flatter than for the relaxed surface. The ordered surface had only one type of adsorption site with regular connectivity between these identical sites.
The distribution of adsorption energies, for each of the adsorbates dinitrogen, argon, and methane, were obtained from the potential energy surface by determining the energy values at the bottom of each potential energy well. It was found that, compared with experimentally obtained values, the number-weighted average heats of adsorption were within 0.2 kJ/mol for methane and 2 kJ/mol for dinitrogen and argon. The adsorption energy distributions for dinitrogen on the four types of model surfaces were obtained and are compared in
Figure 5. While the ordered surface had just a single heat of adsorption of 14.3 kJ/mol, the heats of adsorption for the other three surfaces had a distribution with averages of 11.9 kJ/mol for the relaxed surface, 11.4 kJ/mol for the unrelaxed surface, and 12.8 kJ/mol for the random surface. The rougher relaxed surfaces had a broader distribution than the unrelaxed surface, which is similar to the findings of Riccardo and Steele [
38].
The surface diffusion of adsorbates, interacting weakly via a Lennard-Jones potential, was simulated, using molecular dynamics methods, in order to provide information on the connectivity of adsorption sites [
9]. The surface was considered to be rigid, even though in reality the surface would be vibrating due to random thermal motion. Stallon and Iglesia [
9] suggested this simplification was reasonable for weakly interacting adsorbates like dinitrogen. Further, they suggested that the surface diffusivity estimates for static surfaces are within ~5–10% of those where the top layers of the surface are allowed to relax during the migration of adsorbed species. While the potential and kinetic energy of migrating adsorbates was allowed to fluctuate, the temperature was kept constant. The surface diffusivity was estimated using the Einstein relation of the velocity auto-correlation function, and correlated for different temperatures using the Arrhenius expression. Overall, the simulated activation energies were similar in value to those found by experiment by Barrer et al. [
50]. The trajectories of adsorbate molecules typically followed the lowest energy path across the surface. The surface diffusivity was lowest on the relaxed surface, and highest on the unrelaxed surface, which had a similar average heat of adsorption but a narrower range. The surface diffusivity was also higher on the random and ordered surfaces than on the relaxed surface. This is because the latter had some stronger binding sites (it had the largest diffusion activation energy of 7.8 kJ/mol) than the former two (activation energies ~7 kJ/mol), even though the overall average was lower. However, Stallon and Iglesia [
9] acknowledged that the individual activation energies obtained were too close together, and, thus, insufficiently discriminatory to assess the relative empirical adequacy of the different model surfaces against the limited experimental data available. They also suggested that the connectivity of binding sites on a given surface was critical since they found that the relaxed surface also had a higher density of unconnected sites when compared with the random and ordered surfaces. In contrast, the unrelaxed surface had extended “channels” between energy sites, as can be seen in
Figure 4, which provided routes for enhanced diffusion. Overall, Stallon and Iglesia [
9] suggested that the relaxed surface model was probably the best representation of a real amorphous silica surface, but acknowledged that the possible level of discrimination between models provided by the experimental tests available was low.
Molecular dynamics has also been used to construct relaxed, flat alumina surfaces, of simulated size 30 nm
2, and simulate xenon surface diffusion upon them [
51]. Bläckberg et al. [
51] also created a ridged alumina surface by introducing valleys by removing atoms on the surface and then performing an energy minimisation. The potential energy distribution of each type of surface was mapped by creating a 3D mesh of non-interacting xenon atoms with a spacing of 0.5 Å, and then calculating the potential energy of the system for every xenon atom. Examples of the minimum potential energy are shown in
Figure 6. From
Figure 6, it can be seen that, for the ridged valleys, the potential energy minima are located in the valleys, where adsorbed molecules also tend to localise, and it was found to have a wider distribution of minimum energy values than the flat surface. The surface diffusivity was calculated from the xenon trajectories using the Einstein relation. It was found that the surface diffusivity for the flat surface was 2.3 times higher than for the ridged surface. Surface diffusion on the ridged surface was dominated by movement along the valleys, with little movement perpendicular to them. Hence, the difference in diffusivity with the flat surface may be the result of the restriction in the dimensionality of the molecular migrations.
Sonwane and Li [
18] used the approach of Streitz and Mintmire [
52] to construct atomistic models of hollow nanospheres and nanotubes. They also constructed models of porous media consisting of a packing of spheres. The porosity of the model was varied using densification (allowing overlapping of the spheres) or coarsening (the random removal of spheres from a dense packing) algorithms. These authors used a Monte Carlo model to study surface diffusion on the sphere packing model and obtained the diffusivity from the Einstein relation. Sonwane and Li [
18] defined the tortuosity of the surface phase in the porous solid as follows:
where
ϕ is the voidage fraction and where
Es is as follows:
Es is the normalised surface diffusion coefficient, <
r2> is the mean square displacement, the subscript “
pack” is the packing, and “
ref” is the reference material, which was a cylinder. It was found that, for a densified packing of unimodal sized spheres, the variation in surface tortuosity with voidage fraction had a “U” shape, with the relatively horizontal portion corresponding to voidages of ~0.2–0.35. This behaviour was considered due to the fact that, at one extreme, as the degree of overlap, and hence contact area, between individual spheres decreased, then the chance of a tracer molecule finding a path between adjacent spheres reduced. At the other extreme, highly densified spheres created many relatively inaccessible ‘dead-end’ pore features that were hard for tracers to escape. In contrast, the surface diffusivity for the coarsened spherical packing decreased with increasing voidage fraction up to values of ~0.3, and thereafter was fairly constant up to high porosities. This was said to be because the coarsened packings retained many aggregates with good connectivity. For low porosities, the surface tortuosity was lower than the pore tortuosity, but at porosities higher than ~0.28 the reverse was the case. For models of bimodal solids constructed of packings of spheres that were themselves porous, it was found that the surface tortuosity increased with increased macroporosity around the spheres, due to less contact between them.
Aldo Ledesma-Durán et al. [
53] identified the requirement for an adequate definition of the surface diffusion coefficient that can be experimentally measured for surface diffusion within chemical reactors, where the individual molecular trajectories are not of interest but simply quantities reflecting the average flow rate. They initially computed a position-dependent diffusivity, namely the Fick-Jacobs coefficient, which contains information about the variations in the surface and the internal geometry of the boundary. Subsequently, they obtained an effective diffusivity given in terms of the parameters typically utilised in the description of pores, such as their concavity, tortuosity, and degree of constriction. This second diffusivity permitted the use of a 1D diffusion equation along an effective longitudinal coordinate, whose solution closely approximates the numerically calculated surface diffusion on model surfaces with constrictions that restricted molecular trajectories.
Song et al. [
54] constructed 3D pore structure models of coal using a noise algorithm that altered the random distribution of holes in a lattice by adjusting the scale, intensity, and type of noise such that the resulting model matched the pore size and pore growth ratio from experimental characterization results. The model pore structure was then smoothed, by operations such as the removal of disconnected voids. Numerical simulations of the kinetic uptake of oxygen were then performed on these models. This type of model naturally incorporated the impacts of the convoluted surface of the void space on surface diffusion.
In contrast, some authors have considered surface diffusion in more slit-shaped pores. For example, Zhao et al. [
55] conducted molecular dynamics simulations of methane diffusion in graphene slit-shaped pores, while Zhou et al. [
56] conducted molecular dynamics simulations of the diffusion of mixtures of carbon dioxide and methane within atomistic models of slit-shaped kaolinite pores in shales. Both sets of authors found that the influence of surface diffusion diminished as the pore size increased.
The issue with just considering single slit-shaped pores is that, even in disordered materials with such pores, the void space actually consists of an interconnected network of such pores with different sizes, and the surface diffusion molecular motion occurs continuously between the various pores [
57]. Hence, a full representation of this interconnectivity is required, so Di Pino et al. [
57] conducted molecular dynamics simulations of the diffusion of oxygen, carbon dioxide, and methane in models of carbon slit-shaped micropores, both smooth-walled and rough-walled carbon nano-tubes (CNTs), and amorphous, microporous carbon structures. The rough-walled CNTs were created by randomly adding atoms to the surface of smooth-walled CNTs. The amorphous materials were created with a linear quenched molecular dynamics method. It was found that surface diffusion was the dominant migration mechanism in all cases except for the amorphous model. Further, while for smooth-walled structures the self-diffusivity decreased with increased pore size, for rough-walled structures the opposite trend with pore size held. This effect was attributed to the enhanced adsorption of molecules in the ruts present on rough surfaces. However, for the amorphous structural model, these roughness features were so prevalent that they added so much additional tortuosity to surface-bound molecular migration that gas phase transfer was the dominant mechanism. These diverse findings for variously roughened models showed that it is not currently possible to know in advance what level of this type of additional complexification is necessary for a minimalist idealization model to be predictive.
Shan et al. [
58] compared the surface diffusion of methane in smooth, organic pores with that in rough, inorganic (calcite) pores in shale using molecular dynamics. They found that the roughness made a difference to the relative importance of surface diffusion to the overall mass transfer. It was found that surface diffusion in inorganic pores could be ignored when the characteristic length of the flow domain exceeded 8 nm, while enhanced surface diffusion could only be ignored in organic pores when this length exceeded 30 nm. Hou et al. [
59] conducted a similar study of surface diffusion in shale organic and inorganic pores using the lattice Boltzmann technique. They found that gas transport was dominated by the surface diffusion in the organic pores when their width was less than 5 nm owing to the greater gas adsorption effect therein.
Yu et al. [
60] constructed models of amorphous organic nanopores in shales using realistic kerogen molecules. They then conducted molecular dynamics simulations of gas transport through these pore models. They found that the gas-transport velocity of methane in the amorphous organic nanopores dropped substantially (40, 70, and 90%) even with only very tiny roughness factors (0.3, 0.6, and 1.2%) when compared with the corresponding ideally smooth nanochannels.
Molecular dynamics has permitted the following of individual trajectories of surface diffusing molecules within hierarchical porous carbon (HPC) structures [
61]. These models were constructed by first creating carbon nanotubes of pore diameters of 5, 3, and 1.9 nm using computer modelling software. The two larger-sized nanotubes formed mesopores, while the 1.9 nm CNTs formed micropores. The CNTs were connected up by symmetrically excavating eight circular holes of diameter 1.9 nm in the walls of mesopores (with a certain size) of length 6 nm, followed by the docking of eight micropores of length 3 nm with these holes. The liquid phase diffusion of nalidixic acid, as a solution in water, was simulated using molecular dynamics. By following the individual trajectories of molecules it was found that the majority of solute molecules first adsorbed on the surface of mesopores (of whichever size) and then migrated to the micropores by surface diffusion, rather than by entering the latter directly. This process allowed a more efficient penetration of the micropores (by solute) that would otherwise be blocked by the very stable configurations of solvent molecules that typically exist therein [
61]. These findings suggested the advantage of HPC structures, coupled with surface diffusion, for enhancing the adsorption of solutes.
Zheng et al. [
62] conducted simulations of flow including surface diffusion through self-affine fractures using lattice Boltzmann techniques. They used the simulations to confirm that, for self-affine fractures, the scaling relation for the surface tortuosity τ
s is of the following form:
where
<δ> is the mean aperture size,
lc is characteristic length of the self-affine surface, and
H is the scaling law exponent.