A Density Functional Theory - American Chemical Society

Jun 27, 2012 - Department of Physics, Canisius College, Buffalo, New York 14208, United States. Eli Ruckenstein*. Department of Chemical and Biologica...
0 downloads 0 Views 901KB Size
Article pubs.acs.org/Langmuir

Adsorption Isoterms and Capillary Condensation in a Nanoslit with Rough Walls: A Density Functional Theory Gersh O. Berim Department of Physics, Canisius College, Buffalo, New York 14208, United States

Eli Ruckenstein* Department of Chemical and Biological Engineering, State University of New York at Buffalo, Buffalo, New York 14260, United States ABSTRACT: Adsorption isoterms and capillary condensation in an open slit with walls decorated with arrays of pillars are examined using the density functional theory. Compared with the main substrate, the pillars can have the same or different parameters in the Lennard-Jones interaction potential between them and the fluid in the slit. The roughness of the solid surface, defined as the ratio between the area of the actual surface and the area of the surface free of pillars, is controlled by the height of the pillars. It is shown that the capillary condensation pressure first increases with increasing roughness, passes through a maximum, and then decreases. The amount of adsorbed fluid at constant volume of the slit has, in general, a nonmonotonic dependence on roughness. These features of adsorption and capillary condensation are results of increased surface area and changes in the fluid−solid potential energy due to changes in roughness.

1. INTRODUCTION Experimental and theoretical studies have shown that the properties of a fluid are affected by changes in spatial confinements. Such changes affect the phase behavior of fluids resulting in phenomena such as capillary condensation, capillary freezing, and layering transition. The behavior of confined fluids is well understood for one-component fluids as well as binary mixtures and simple models of the pores (long slits, spherical or cylindrical pore shapes, etc.) with smooth uniform walls. However, the surfaces of real pores usually possess either physical roughness (grooves, asperities, etc.) or chemical roughness (nonuniform chemical composition of the substrate), being geometrically or chemically heterogeneous. To estimate the role of such heterogeneities, several recent studies employed the quenched−annealed model of the fluid−solid system and developed a suitable version of the density functional theory (DFT).1−3 One of the main features of that approach was the modeling of roughness as a layer (matrix) of quenched molecules on the solid surface, with nonuniform density along the normal to the solid surface. In lateral directions the layer was considered uniform. The adsorbate molecules were allowed to penetrate into this layer.3 The density distribution of the matrix molecules across the layer was selected in a simple form and used as an input into the theory.1−3 A different procedure was used in ref 4 where the density distribution of the surface layer was calculated by minimizing the free energy of the system using the density © 2012 American Chemical Society

functional theory. The approach used in refs 1−4 allowed reduction of the Euler−Lagrange equation for the fluid density distribution to a one-dimensional form, thus providing a major simplification of the calculations. Among the results obtained in refs 1−3, one can mention the damping of the oscillations of the adsorbate density which occurs in the absence of the layer, the change in the capillary condensation pressure, and the shift in the hysteresis loops of the adsorption isoterms. The capillary condensation pressure increases (decreases) and the hysteresis loops are shifted to higher (lower) pressures when the adsorbate−matrix interaction is weaker (stronger) than the adsorbate−substrate one. The increase in capillary condensation pressure can become so large that the capillary condensation can be replaced by capillary evaporation.2 The latter phenomenon occurs typically when the fluid is confined between hydrophobic walls.5 One of the main features of the model developed in refs 1−4 is that the surface area of the slit does not depend on the properties of the surface layer and remains equal to that of the smooth surface. Obviously, this is not true for physically rough surfaces for which the surface area increases with increasing roughness, the latter being defined as the ratio between the actual surface area and the area of the smooth surface. Received: April 9, 2012 Revised: June 26, 2012 Published: June 27, 2012 11384

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390

Langmuir

Article

Nonetheless, the progress in modern nanotechnologies allows one to fabricate physically rough surfaces with prescribed structures, for instance, surfaces decorated with regular arrays of asperities of various shapes, sizes, and compositions.6−11 It is well-known that such modifications of the surface lead to considerable changes in its wetting properties. In particular, if the fluid penetrates between macroscopic asperities (the socalled Wenzel regime), the contact angle θW, which a macroscopic liquid drop makes with a rough surface, is provided by the Wenzel equation12 cos θ W = r cos θ

(1)

where r is the roughness and θ is the contact angle on a smooth surface. Consequently, a hydrophobic surface (θ > 90°) becomes more hydrophobic and a hydrophilic one (θ < 90°) more hydrophilic with increasing roughness. For nanosize asperities, the hydrophilicity of a hydrophilic surface decreases with increasing roughness and the surface can become even hydrophobic.13,14 Along with changes in wetting, one also expects roughness to affect adsorption. Because the approaches developed in refs 1−4 are not applicable to the systems possessing the above geometrical roughness, in the present paper we examine adsorption and capillary condensation in a planar slit with physically rough surfaces using the traditional DFT, by assuming that the adsorbate does not penetrate into the substrate. The roles of the size of the pillars, which provide the roughness, as well as the roles of the strength of the pillar−fluid interactions are examined in some detail. Because the potential generated by a real heterogeneous solid surface is a complicated function of three spacial coordinates, the numerical solution of the Euler− Lagrange equation providing the fluid density distribution in the system is practically impossible. To simplify the problem, a two-dimensional geometry of the system will be employed. The Euler−Lagrange equation in this case involves two independent spacial variables and is solved by iteration using the procedure developed in ref 14. It is worth noting that the two-dimensional versions of DFT were applyed previously (see refs 15−18) to examine the fluids confined between smooth solid surfaces.

Figure 1. Schematic representation of the considered system. σfs is the hard core diameter of the fluid−solid interactions. All distances between surfaces are measured between the centers of the molecules forming the first layers.

constant. The latter volume can be controlled by changing the distance Lh between the walls W1 and W2. For the smooth slit walls (no pillars), Lh was selected 10σff + 2σfs. Hence, when hp is changed, the distance Lh has to be changed according to the formula Lh = 10σff + 2σfs + hp. Note that pillars with height hp > 10σff obstruct the fluid passage in the x-direction because the distance between the pillars located in front of each other on opposite solid walls becomes too small to allow the fluid molecules to move between them. For this reason, only pillars with heights hp < 10σff are considered below. The fluid density distribution (FDD) ρ(r) in this system is considered uniform in the y-direction and nonuniform in the xand h-directions; hence, ρ(r) ≡ ρ(x, h). Because of the regular geometrical location of identical pillars, the rough surface generates a potential which is periodic in the x-direction. If there is no symmetry breaking in the x-direction, the FDD ρ(x, h) has the period Lx = Δp + dp and can be calculated in the finite range −Lx/2 ≤ x ≤ Lx/2. This reduces considerably the computational time. The fluid is exposed to an external potential due to the fluid−substrate and fluid−pillars interactions. The temperature was fixed at T = 87.3 K. The interaction potentials between the spherical molecules of fluid and between the molecules of fluid and those of the solid are selected in the Lennard-Jones form with hard core repulsion

2. THEORY 2.1. The System. The considered system involves a onecomponent fluid (argon) confined between two semiinfinite uniform solid walls W1 and W2 decorated with regular arrays of pillars of height hp, width dp, and distance between pillars Δp (see Figure 1). The pillars are infinite in the y-direction (perpendicular to the plane of the figure) and are composed of a material which can differ from the material of the substrate. The system is in contact with a reservoir of fluid (argon) at the chemical potential μ < μ0 (at pressure P < P0) where μ0 and P0 are the chemical potential and the pressure at the saturation point of the bulk fluid, respectively. The slit has infinite dimensions in the x and y directions, while the distance Lh between walls is finite. The origin of the coordinate system in Figure 1 is located at a distance from the solid substrate equal to one hard core diameter σfs of the fluid− solid interaction. All distances between the surfaces of substrate and pillars are measured between the centers of the molecules forming the first layers of the corresponding surfaces. All pillars have the same width dp = 2.410σff, where σff is the hard core diameter of the fluid−fluid interactions, and the distance between them Δp = dp. In what follows, the volume of the system accessible to the fluid molecules will be considered

⎧ ⎡⎛ σ ⎞12 ⎛ σ ⎞6 ⎤ ⎪ ⎪ 4εα⎢⎜ α ⎟ − ⎜ α ⎟ ⎥ , Δr ≥ σα ⎝ Δr ⎠ ⎦ ϕα(|r − r′|) = ⎨ ⎣⎝ Δr ⎠ ⎪ ⎪ ∞, Δr < σα ⎩

(2)

where the subscript α is “ff”, “fs”, and “fp” for fluid−fluid, fluid− substrate, and fluid−pillars interactions, respectively, εff, εfs, and εfp are energy parameters, σff, σfs, and σfp are hard core diameters, r and r′ provide the locations of the interacting molecules, and Δr = |r − r′| . For argon, εff/kB = 119.76 K, and σff = 3.405 Å, where kB is the Boltzmann constant. For the substrate, the cases εfs = ε0fs and εfs = 2ε0fs will be considered where ε0fs/kB = 153.0 K corresponds to the solid carbon dioxide. For σfs and ρs the values specific to solid carbon dioxide where employed, namely, σfs = 3.727 Å and ρs = 1.91 × 1028 m−3. In real cases, the parameters σfp, εfp, and ρp (density of pillars) can differ from those of the substrate. However, for simplicity, σfp 11385

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390

Langmuir

Article

and ρp were selected equal to σfs and ρs, respectively, the difference between pillars and substrate being provided only by the parameter εfp. The latter parameter was selected ε0fs or 2ε0fs, and three kinds of slits with εfs = εfp = ε0fs (case 1), εfs = εfp = 2ε0fs (case 2), and εfs = ε0fs, εfp = 2ε0fs (case 3) will be considered. Because of the geometry of the system, the external potential Usf generated by the solid (substrate plus pillars) at a point M(x, y, h) depends on the distance h + σfs of this point to the surface of the solid and the position x of point M along this surface, and is independent of y, i.e., Usf ≡ Usf(x, h). The potential Usf can be obtained by integrating the Lennard-Jones potential (eq 2) for the fluid−substrate and fluid−pillars interactions over the entire volume of the solid and can be written as Ufs(r) =

reservoir and F[ρ(r)] is the Helmholtz free energy. The latter can be expressed as the sum of an ideal gas free energy, Fid[ρ(r)], an excess free energy Fhs[ρ(r)] (with respect to the ideal gas) of a reference system of hard spheres, a free energy Fattr[ρ(r)] due to the attractive interactions between the fluid molecules (in the mean field approximation), and a free energy Ffs[ρ(r)] due to the interactions between the fluid and solid (substrate plus pillars). Explicit expressions for these contributions, which can be obtained in standard ways,21 are provided in the Appendix. The Euler−Lagrange equation for ρ(x, h) (−Lx/2 ≤ x ≤ Lx/ 2) can be written in the following form22 μ log[Λ3ρ(x , h)] − Q (x , h) = kBT (7)

∫V ρs (r′)ϕfs(|r − r′|) dr′ + ∫V ρp(r′)ϕfp(|r − r′|) s

where the function Q(x, h), which is a functional of ρ(x, h), is given in the Appendix, Λ = hP/(2πmkBT)1/2 is the thermal de Broglie wavelength, hP is the Planck constant, T is the absolute temperature, and m is the mass of a fluid molecule.

p

(3)

dr′

where Vs is the volume of the substrate, Vp is the volume occupied by pillars, ρs(r′) and ρp(r′) are the densities of the substrate and pillars, respectively, which, in general, depend on the coordinates. For a uniform substrate, ρs(r′) ≡ ρs, and the first integral in eq 3 can be calculated analytically:

3. RESULTS The aim of the present calculations is to examine the changes in adsorption isoterms and features of capillary condensation due to the physical roughness of the surface, r, and the nature of pillars. The first parameter (r) is controlled by the height of the pillars, hp, and for the considered geometry of pillars is given by the equation

∫V ρs (r′)ϕfs(|r − r′|) dr′ s

=

9 ⎡ ⎛ ⎛ σfs ⎞3⎤ σfs ⎞ 2π 3⎢ 2 εfsρσ − ⎜ ⎟ ⎜ ⎟⎥ s fs ⎢ σ σ 3 15 h h + + ⎝ ⎠ ⎝ ⎠ ⎥⎦ fs fs ⎣

r=1+

(4)

The contribution ΔUfp,i(r) of a pillar to the potential Ufs(r) can be represented in the form ΔUfp, i(r) = ρp

∫A

p, i

∫−∞ dy′ϕfp(|r − r′|)

(5)

where Ap,i is the area of the cross section of the pillar in the x−h plane. After integration over y′, eq 5 acquires the form ΔUfp, i(r) =

3π εfpρp σfp3 dx′ dh′ A p, i 2 ⎡ σfp9 ⎢ 31 ⎢⎣ 32 [(x − x′)2 + (h − h′)2 ]11/2

ρav* =





⎤ ⎥ [(x − x′)2 + (h − h′)2 ]5/2 ⎥⎦

d p + Δp

(8)

The second parameter (nature of the pillars) is reflected in the changes of the energy parameter εfp in the pillar−fluid interactions. First, let us consider the adsorption isoterms obtained for various roughnesses and all considered parameters of fluid− solid interactions. In Figure 2 the dimensionless average density, ρav * ≡ ρavσff3, of the fluid in the slit



dx′ dh′

2hp

Lx /2

∫−L /2 dx ∫0

Lh

dhρ*(x , h)/(LxL h − 2hp d p)

x

(9)

where ρ*(x, h) = ρ(x, is plotted versus the pressure in the reservoir for selected roughnesses (indicated on each curve) and for a smooth surface (r = 1). Note that the denominator in eq 9 represents the total volume of the system filled with the fluid. The data in Figure 2 are presented only for thermodynamically stable states. The metastable states which exhibit hysteresis loops and correspond to local minima of the free energy are not plotted to avoid too many curves on the figure. They will be discussed below with regard to Figure 7. The pressure P in the reservoir as a function of chemical potential μ was calculated as in ref 23. Because the volume of the slit accessible to the fluid is constant (does not depend on roughness), the average density provides the amount of fluid adsorbed in the slit as a function of roughness. As expected, the equilibrium capillary condensation pressure Pcond, defined as the pressure at which the vapor and liquid branches of the adsorption isoterm have equal free energies (point of equilibrium phase transition), depends on the roughness of the surface. In Figure 3 its behavior as a function of roughness is presented for all considered surfaces. For case 1, Pcond monotonically increases with increasing roughness and at r ≃ 1.65 becomes larger than P0; that is, the equilibrium h)σff3

σfp3

(6)

The integral on the right-hand side of eq 6 was calculated numerically. To calculate the fluid−solid potential generated by all pillars in the range −Lx/2 ≤ x ≤ Lx/2, the contributions of the upper and lower pillars from that interval as well as the contributions of their six nearest neighbors along the corresponding surface were taken into account. The contributions of all other pillars were negligibly small. Note that more general but more complicated analytical calculations of the potential for parallelepipedal and cylindrical pillars are provided in ref 19 and for stepped surfaces in ref 20. 2.2. Euler−Lagrange Equation for the Fluid Density Distribution. The equation for the fluid density distribution (FDD) ρ(x, h) is obtained through the minimization of the grand canonical potential of the system Ω[ρ(r)] = F[ρ(r)] − μ∫ ρ(r)dr where μ is the chemical potential of the fluid in the 11386

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390

Langmuir

Article

Figure 3. Capillary condensation pressure Pcond as function of roughness for case 1 (εfs = εfp = ε0fs) (solid line), case 2 (εfs = εfp = 2ε0fs) (dashed line), and case 3 (εfp = 2ε0fs, εfs = ε0fs) (dashed-dotted solid line).

above the pillars decreases rapidly with increasing pillar height, and the contact angle becomes dominated by the attractive interaction between fluid and pillars. The latter attraction increases rapidly with increasing hp, approaching an asymptotic value which remains smaller than the fluid−substrate one. In this case, the equilibrium capillary condensation transition occurs at bulk pressures larger than P0. In cases 2 and 3, the fluid−pillars interaction constant is stronger than that in case 1 (εfp = 2ε0fs) and the increase in pillar height is accompained by two competitive tendencies. One of them is a decrease in hydrophilicity due to the decrease of the fluid−substrate attraction above the pillars, which results in increasing Pcond. The other one is an increase in hydrophilicity due to increasing attraction between fluid and pillars and consequently Pcond decreases. The first tendency dominates at small pillar heights and the second dominates at sufficiently large pillar heights. As a result, the dependence of Pcond on hp passes through a maximum (see Figure 3). In Figure 4, the dependence of the average fluid density, ρav *, in the slit on roughness is presented for several values of the pressure in the reservoir P < Pcond for all considered cases of the fluid−solid interaction parameters. For cases 1 and 2 where the pillars are made of the same material as the substrate, the average density as a function of roughness has a minimum at r ≃ 1.6 for all considered values of P/P0 (see Figure 4a,b). In case 1, the dependence has a maximum at r ≃ 3.5 (Figure 4a). For case 2, such a maximum is most likely exhibited at a higher roughness not considered in the present paper (Figure 4b). In case 3 where the substrate and pillars are made from different materials, the average density increases monotonically with increasing roughness (see Figure 4c). For Pcond < P < P0, the average density in the slit for all cases monotonically increases with increasing roughness. In Figure 5, examples of such a dependence for case 2 and case 3 are presented for P/P0 = 0.88. For the same roughness, the difference between liquid densities in those cases is very small and the corresponding curves almost coincide. To explain qualitatively the behavior of ρav * under different conditions, one has to take into account the two factors which affect adsorption in the slit. The first one is the area of the surface of the slit which increases with increasing roughness. Because of large fluid densities near the solid, the increase of that area increases the amount of adsorbed fluid. The second

Figure 2. Adsorption isoterms as functions of the pressure in the reservoir for various roughnesses indicated on each curve. (a) Case 1 (εfs = εfp = ε0fs), (b) case 2 (εfs = εfp = 2ε0fs), and (c) case 3 (εfs = ε0fs, εfp = 2ε0fs). Solid lines are for r = 1, long-dashed line is for r = 1.41, dashed lines are for r = 1.83, dot-dashed line is for r = 2.66, and dotted lines are for r = 3.49.

capillary condensation transition becomes a capillary evaporation one. In the two other cases, Pcond slightly increases with increasing roughness, passes through a maximum at r ≃ 1.6 for case 3 and r ≃ 2.25 for case 2, and then monotonically decreases. Qualitatively, the behavior of Pcond can be associated with changes in the wetting properties of the surface with increasing roughness. In all considered cases, the surface without pillars is hydrophilic (θ < 90°).14,22 It was shown in ref 14 that for fluid−substrate and fluid−pillar interaction constants corresponding to case 1 (εfs = ε0fs, εfp = εfs), the hydrophilicity of the surface decreases with increasing pillar height and the surface becomes hydrophobic for sufficiently large hp. This occurs because the fluid−substrate attraction 11387

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390

Langmuir

Article

Figure 5. Average density ρav * as function of roughness for P > Pcond for case 2 (εfs = εfp = 2ε0fs, solid line) and case 3 (εfp = 2ε0fs, εfs = ε0fs, dashed line).

Figure 6. Potential of fluid−solid interaction at distance σfs above the middle of a pillar as a function of pillar height for case 1 (εfs = εfp = ε0fs) (solid line) and case 3 (εfp = 2ε0fs, εfs = ε0fs) (dashed line).

monotonic increase of ρav * with increasing hp. For this reason, there is no extremum in the ρav * vs r dependence in Figure 4c. In Figure 7 the hysteresis loops in the dependence of the average density of adsorbed fluid on the pressure in the reservoir is presented at various roughnesses for all considered cases. The dotted vertical lines indicate the location of the bulk saturation pressure (P = P0). In case 1, the adsorption branches of the hysteresis loops are located in the range P > P0 for all considered roughnesses. Because we consider only the range P < P0, those corresponding to P > P0 are not displayed in the figure. In case 2, the hysteresis loops are located completely in the range P < P0. However, their location depends on roughness (see Figure 7b). In case 3, for small roughnesses (r ≤ 3.0) the adsorption branches of the hysteresis loops are located in the range P > P0. For r > 3.0 the entire hysteresis loops are located in the vapor-phase range of the bulk fluid. This means that the increase in roughness in the last case favors capillary condensation.

Figure 4. Average density ρav * as a function of roughness for various values of the pressure in the reservoir (indicated on each curve) for (a) case 1 (εfs = εfp = ε0fs), (b) case 2 (εfs = εfp = 2ε0fs), and (c) case 3 (εfp = 2ε0fs, εfs = ε0fs). In all cases, P is smaller than Pcond.

factor is the attractive potential near the solid surface, which depends in the considered cases on the height and nature of the pillars. If this potential becomes less attractive, the fluid densities close to the surface decrease, thus decreasing ρav * . Vice versa, if this potential becomes more attractive, ρ*av increases. As an illustration, in Figure 6 the fluid−solid potential is presented at distance σfs above the center of a pillar for cases 1 and 3. In the former case, the potential increases for 0 ≤ hp/σff ≤ 1.2 (becomes less attractive) with increasing pillar height (roughness) and then remains almost constant. However, the area of the solid surface continuously increases with increasing pillar height (roughness). The competition between these two factors is resposible for the behaviors of ρ*av presented in Figure 4a,b. In case 3, the potential decreases for 0 ≤ hp/σff ≤ 1.2 (becomes more attractive) with increasing pillar height (roughness). Along with the increasing area of the surface, this provides a

4. CONCLUSION It was shown that the presence of pillars changes the phase behavior of confined fluids. First, the roughness due to pillars changes the equilibrium capillary condensation pressure, Pcond, in nontrivial ways, resulting in some cases in maxima in the r dependence of Pcond (see Figure 3). Second, the r dependence of ρav * for P < Pcond can exhibit nonmonotonic behavior resulting in the existence of extrema (see Figure 4). The reason for such 11388

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390

Langmuir

Article

where Λ = hP/(2πmkBT)1/2 is the thermal de Broglie wavelength, hP and kB are the Planck and Boltzmann constants, respectively, T is the absolute temperature, m is the mass of a fluid molecule, ΔΨhs(r) = kBTηρ

4 − 3ηρ ̅ (1

̅

− ηρ )2 ̅

(A.3)

ηρ̅ = is the packing fraction of the fluid molecules, σff is the fluid hard core diameter, and ρ̅(r) is the smoothed density defined as (1/6)πρ̅(r)σff3

ρ ̅ (r) =

∫ dr′ρ(r′) W (|r − r′|)

(A.4)

The weighting function W(|r − r′|) is selected in the form25 ⎧ 3 ⎛ r ⎞ ⎪ 1 − ⎟ , r ≤ σff 3⎜ σff ⎠ W (|r − r′|) = ⎨ πσff ⎝ ⎪ 0, r > σff ⎩

where r = |r − r′|. The contribution to the excess free energy due to the attraction between the fluid−fluid molecules is calculated in the mean-field approximation Fattr[ρ(r)] =

1 2

∫ ∫ dr dr′ ρ(r) ρ(r′) ϕff (|r − r′|)

(A.5)

where ϕff(|r − r′|) is the Lennard-Jones potential of the fluid− fluid interactions. The last contribution, Ffs[ρ(r)], is given by the expression Ffs[ρ(r)] =

∫V dr ρ(r) Ufs(r)

(A.6)

where V is the volume occupied by the fluid and Ufs(r) is the net potential generated by the solid (substrate plus pillars). This potential is provided by eq 3. The Euler−Lagrange equation for the fluid density distribution ρ(x, h), obtained by minimizing the grand canonical potential, can be written in the following general form μ log[Λ3ρ(x , h)] − Q (x , h) = kBT (A.7) where μ is the chemical potential in the reservoir and the function Q(x,h) is given by

Figure 7. Hysteresis loops in adsorption isoterm for various roughnesses as a function of the pressure in the reservoir. (a) Case 1 (εfs = εfp = ε0fs), (b) case 2 (εfs = εfp = 2ε0fs), and (c) case 3 (εfp = 2ε0fs, εfs = ε0fs). Solid line is for r = 1, dashed lines are for r = 1.83, dot-dashed line is for r = 2.66, and dotted lines are for r = 3.49.

Q (x , h) = −

1 [ΔΨhs(x , h) + ΔΨ′hs (x , h) + Uff (x , h) kBT

+ Ufs(x , h)]

features is the competition between the increased area of the rough surface and the decrease in the net fluid−solid interaction potential.

where Uff (x , h) =

5. APPENDIX: EULER−LAGRANGE EQUATION To derive the Euler−Lagrange equation, one has to consider the contributions to the free energy mentioned in section 2.2, which can be represented as follows:21,24 Fid[ρ(r)] = kBT Fhs[ρ(r)] =

∫ dr ρ(r){log[Λ3ρ(r)] − 1}

∫ dr ρ(r) ΔΨhs(r)

(A.8)

∫ ∫ dx′ dh′ ρ(x′, h′) ϕff,y(|x − x′|, |h − h′|) (A.9)

ΔΨ′hs (x , h) =

∫∫

dx′ dh′ ρ(x′, h′)Wy(|x − x′| ,

|h − h′|)

(A.1)

∂ ΔΨhs(ρ ̅ )| ρ ̅ = ρ ̅ (x , h ) ′ ′ ∂ρ ̅

(A.10)

ϕff,y(|x − x′|,|h − h′|) and Wy(|x − x′|,|h − h′|) are obtained by integrating the potential ϕff(|r − r′|) and the weighted function W(|r − r′|) with respect to y from −∞ to + ∞.

(A.2) 11389

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390

Langmuir

Article

(19) Wu, H.; Borhan, A.; Fichthorn, K. A. Interaction of fluids with physically patterned solid surfaces. J. Chem. Phys. 2010, 133 (5), 054704. (20) Bojan, M. J.; Steele, W. A. Computer-Simulation of Physical Adsorption on Stepped Surfaces. Langmuir 1993, 9 (10), 2569−2575. (21) Tarazona, P. Free-energy density functional for hard-spheres. Phys. Rev. A 1985, 31 (4), 2672−2679. (22) Berim, G.; Ruckenstein, E. Nanodrop on a nanorough solid surface: Density functional theory considerations. J. Chem. Phys. 2008, 129 (1), 014708. (23) Ancilotto, F.; Faccin, F.; Toigo, F. Wetting transitions of He-4 on alkali-metal surfaces from density-functional calculations. Phys. Rev. B 2000, 62 (24), 17035−17042. (24) Tarazona, P.; Marconi, U. M. B.; Evans, R. Phase-equilibria of fluid interfaces and confined fluids - nonlocal versus local density functionals. Mol. Phys. 1987, 60 (3), 573−595. (25) Nilson, R. H.; Griffiths, S. K. Condensation pressures in small pores: An analytical model based on density functional theory. J. Chem. Phys. 1999, 111 (9), 4281−4290.

When calculating the term Uff(x,h) of the Euler−Lagrange equation arising due to the long-range fluid−fluid interactions, a cutoff at a distance equal to four molecular diameters σff for the range of Lennard-Jones attraction was employed.



AUTHOR INFORMATION

Corresponding Author

*E-mail: feaeliru@buffalo.edu. Phone: (716) 645-1179. Fax: (716) 645-3822. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Bryk, P.; Henderson, D.; Sokolowski, S. The wetting transition associated with the adsorption of a gas on a rough surface. Langmuir 1999, 15 (18), 6026−6034. (2) Bryk, P.; Rzysko, W.; Malijevsky, A.; Sokolowski, S. Capillary condensation in pores with rough walls: A density functional approach. J. Colloid Interface Sci. 2007, 313 (1), 41−52. (3) Ravikovitch, P. I.; Neimark, A. V. Density functional theory model of adsorption on amorphous and microporous silica materials. Langmuir 2006, 22 (26), 11171−11179. (4) Berim, G. O.; Ruckenstein, E. Sorption on deformable dolids. Density functional theory approach. J. Phys. Chem. B 2011, 115 (45), 13271−13274. (5) Roth, R.; Kroll, K. M. Capillary evaporation in pores. J.Phys.: Condens. Matter 2006, 18 (28), 6517−6530. (6) Shibuichi, S.; Onda, T.; Satoh, N.; Tsujii, K. Super water-repellent surfaces resulting from fractal structure. J. Phys. Chem. 1996, 100 (50), 19512−19517. (7) Onda, T.; Shibuichi, S.; Satoh, N.; Tsujii, K. Super-waterrepellent fractal surfaces. Langmuir 1996, 12 (9), 2125−2127. (8) Song, Y.; Nair, R. P.; Zou, M.; Wang, Y. Q. Superhydrophobic Surfaces Produced by Applying a Self-Assembled Monolayer to Silicon Micro/Nano-Textured Surfaces. Nano Res. 2009, 2 (2), 143−150. (9) Wang, H.; Zou, M.; Wei, R. Superhydrophilic textured-surfaces on stainless steel substrates. Thin Solid Films 2009, 518 (5), 1571− 1574. (10) Reyssat, M.; Pardo, F.; Quere, D. Drops onto gradients of texture. EPL 2009, 87, (3). (11) Kwon, Y.; Choi, S.; Anantharaju, N.; Lee, J.; Panchagnula, M. V.; Patankar, N. A. Is the Cassie-Baxter Formula Relevant? Langmuir 2010, 26 (22), 17528−17531. (12) Wenzel, R. N. Resistance of Solid Surfaces to Wetting by Water. Ind. Eng. Chem. 1936, 28 (8), 988−994. (13) Daub, C. D.; Wang, J. H.; Kudesia, S.; Bratko, D.; Luzar, A. The influence of molecular-scale roughness on the surface spreading of an aqueous nanodrop. Faraday Discuss. 2010, 146, 67−77. (14) Berim, G. O.; Ruckenstein, E. Nanodrop on a nanorough hydrophilic solid surface: Contact angle dependence on the size, arrangement, and composition of the pillars. J. Colloid Interface Sci. 2011, 359 (1), 304−310. (15) Ustinov, E. A.; Do, D. D. Application of density functional theory to capillary phenomena in cylindrical mesopores with radial and longitudinal density distributions. J. Chem. Phys. 2004, 120 (20), 9769−9781. (16) Siderius, D. W.; Gelb, L. D. Predicting Gas Adsorption in Complex Microporous and Mesoporous Materials Using a New Density Functional Theory of Finely Discretized Lattice Fluids. Langmuir 2009, 25 (3), 1296−1299. (17) Jagiello, J.; Olivier, J. P. A Simple Two-Dimensional NLDFT Model of Gas Adsorption in Finite Carbon Pores.Application to Pore Structure Analysis. J. Phys. Chem. C 2009, 113 (45), 19382−19385. (18) Jagiello, J.; Kenvin, J.; Olivier, J. P.; Lupini, A. R.; Contescu, C. I. Using a New Finite Slit Pore Model for NLDFT Analysis of Carbon Pore Structure. Adsorp. Sci. Technol. 2011, 29 (8), 769−780. 11390

dx.doi.org/10.1021/la3014464 | Langmuir 2012, 28, 11384−11390