Grazing-incidence x-ray diffraction (GIXRD) is a scattering technique that allows one to characterize the structure of fluid interfaces down to the molecular scale, including the measurement of surface tension and interface roughness. However, the corresponding standard data analysis at nonzero wave numbers has been criticized as to be inconclusive because the scattering intensity is polluted by the unavoidable scattering from the bulk. Here, we overcome this ambiguity by proposing a physically consistent model of the bulk contribution based on a minimal set of assumptions of experimental relevance. To this end, we derive an explicit integral expression for the background scattering, which can be determined numerically from the static structure factors of the coexisting bulk phases as independent input. Concerning the interpretation of GIXRD data inferred from computer simulations, we extend the model to account also for the finite sizes of the bulk phases, which are unavoidable in simulations. The corresponding leading-order correction beyond the dominant contribution to the scattered intensity is revealed by asymptotic analysis, which is characterized by the competition between the linear system size and the x-ray penetration depth in the case of simulations. Specifically, we have calculated the expected GIXRD intensity for scattering at the planar liquid–vapor interface of Lennard-Jones fluids with truncated pair interactions via extensive, high-precision computer simulations. The reported data cover interfacial and bulk properties of fluid states along the whole liquid–vapor coexistence line. A sensitivity analysis shows that our findings are robust with respect to the detailed definition of the mean interface position. We conclude that previous claims of an enhanced surface tension at mesoscopic scales are amenable to unambiguous tests via scattering experiments.
I. INTRODUCTION
The liquid–vapor interface is a ubiquitous confining boundary of fluids and has been the subject of enduring experimental, theoretical, and simulational interest. These efforts focus on the properties of adsorbed liquid films,1–4 droplets,5–10 and interfaces out of equilibrium,11–16 as well as on applications such as wetting,17–20 solvation,21–25 and the presence of surfactants.26–32 A considerable number of investigations is motivated by the long-standing issue concerning the influence of van der Waals or dispersion forces on the structure and behavior of the interface.33–48 Not only due to this issue, the liquid–vapor interface has been an important testing ground for the theory of inhomogeneous fluids. Recent advances concerning their theoretical description have been stimulated by simulation data for a range of temperatures49 and led to accurate predictions of interfacial density correlations by combining the concept of a position-dependent, local structure factor50–52 and insight into resonances stemming from the bulk structure.53–55 In view of these predictions, a fully comprehensive interpretation of experiments concerning liquid–vapor interfaces still remains to be formulated.
Experimentally, scattering techniques such as x-ray reflectometry and grazing-incidence x-ray diffraction (GIXRD) provide the most detailed view of fluid interfaces. Whereas reflectometry allows one to infer interfacial density profiles, GIXRD probes density fluctuations56–58 and surface structures.29,30 Such kind of scattering experiments under grazing incidence have been carried out for liquid surfaces of water,36 molecular fluids,37 and liquid metals.38 The analysis of the scattered intensities requires accounting for the inevitable scattering from the fluid bulk phases, which depends on a suitable background model,59–61 nowadays available within GIXRD analysis software;62 yet, such a modeling choice can potentially introduce ambiguities in the interpretation of experimental data.
Focusing on thermal equilibria of coexisting liquid and vapor phases, an increase in the interfacial area by a macroscopic amount ΔA incurs a free energy cost of γ0 ΔA, which defines the (macroscopic) surface tension γ0. At the molecular scale, the interface is roughened by thermal fluctuations, which are tamed by an analogous cost in free energy. This mesoscopic picture leads to the capillary wave (CW) theory,63–65 which assumes the interface to be a smooth, membrane-like surface with a tension modulus governed by an interface Hamiltonian. On the other hand, by adopting the particle perspective, the theory of inhomogeneous fluids64,66,67 considers (two-point) density fluctuations in order to characterize the interfacial region. For a free, planar interface, the classical result by Wertheim68 and Weeks69 states that fluctuations with wavevector q parallel to the surface lead to a scattering intensity proportional to kBT/(γ0q2), which applies for small q = |q|, i.e., for macroscopic wavelengths 2π/q [cf. Fig. 8 and Eq. (65)]; the thermal energy scale is denoted by kBT as usual. This asymptotic result (q → 0) matches with the free energy cost of the excitation of corresponding CWs. It has stimulated elaborate derivations of effective interfacial Hamiltonians33–35,70–72 in order to capture also the (anticipated) structure of the two-point correlation function at higher orders in q. These extended CW theories include also a bending modulus κ of the interface43,44 or, more generally, a wave number-dependent surface tension , which exhibits the relation . The leading correction to this macroscopic limit is either nonanalytic [of the form q2 log(q)] or quadratic (κq2), depending on whether dispersion forces are present or not. Molecular dynamics (MD) simulations testing the behavior of have frequently been analyzed in line with the CW picture by evaluating the fluctuation spectrum of the local interface position.20,40–46,73–75 This requires one to adopt a suitable definition of this position down to the molecular scale. Common choices are either a local Gibbs dividing surface or an elaborate algorithm to identify an intrinsic interface separating CWs from bulk-like density fluctuations. However, in view of the various possible choices, the published data for disagree on even the sign of the bending coefficient κ (in the absence of dispersion forces). Moreover, it was shown within density functional theory (DFT)50 that any effective interface Hamiltonian fails to reproduce the behavior of interfacial density correlations, which are obtained from simulations.
As an alternative that bypasses the ambiguities associated with the definition of a local interface position, we previously introduced an effective surface tension γ(q) as a proxy for interfacial density fluctuations that is entirely based on quantities amenable to scattering experiments49 [see below, Eq. (8)]. Similarly, as for the interpretation of scattering data, this approach hinges on a consistent model for the background scattering, which is the subject of the present study. It is based on a hypothetical liquid–vapor interface with all interface-related correlations switched off (Fig. 1), for which the scattered intensity can be worked out analytically. Somewhat unexpectedly, these open boundary conditions at the interface lead to corrections in the small-wave number density correlations due to bulk fluctuations,76 which, as it will turn out, interfere with the small-q behavior of γ(q); a related boundary effect has been observed for the two-dimensional strip geometry.77,78 Accounting for these corrections removes an inconsistency between the results for γ(q) obtained from the direct calculation of the density correlations and obtained from simulated scattering intensities. Moreover, it renders the bending coefficient in γ(q) positive for all temperatures at liquid–vapor coexistence, between the triple point and the critical point of the fluid.
Top: Cross section of a snapshot of a simulated liquid–vapor interface configuration for a truncated Lennard-Jones (LJ) fluid at liquid–vapor bulk coexistence for the temperature T ≈ 0.8Tc; Tc is the critical temperature of this fluid. Bottom: Interface-related fluctuations are switched off in a gedanken experiment (see the main text). This snapshot depicts the initial configuration of the simulation, composed of two independent bulk configurations, before further equilibration.
Top: Cross section of a snapshot of a simulated liquid–vapor interface configuration for a truncated Lennard-Jones (LJ) fluid at liquid–vapor bulk coexistence for the temperature T ≈ 0.8Tc; Tc is the critical temperature of this fluid. Bottom: Interface-related fluctuations are switched off in a gedanken experiment (see the main text). This snapshot depicts the initial configuration of the simulation, composed of two independent bulk configurations, before further equilibration.
The paper is organized as follows. In Sec. II, we provide a number of relations that will be useful for the theoretical analysis; in particular, we discuss the separation of the GIXRD intensity into interfacial and background parts, provide the definition of γ(q), and connect with Ref. 76. The background scattering for the above-mentioned reference system is analyzed in Sec. III with special emphasis on the non-commuting limits of infinite x-ray penetration depth and infinite sample width. In Sec. IV, these results are applied to the analysis of simulation data for scattering intensities from liquid–vapor interfaces.
II. GENERAL CONSIDERATIONS
A. GIXRD master formula
The function f(z) is determined by the scattering geometry (i.e., the angles αi and αf of incoming and outgoing beams, respectively; see Fig. 2 in Ref. 56) and the mean density profile ϱ(z). It describes the decay of the evanescent wave on the liquid side of the interface, f(z < 0) ∼ exp(−κ|z|), with the penetration depth 1/κ(αi, αf).56,80 For αi and αf approaching the critical angle of total reflection, the penetration depth diverges, and, in the thermodynamic limit, I(q) is dominated by scattering from the liquid bulk. Thus, experimental setups aim at minimizing the penetration depth 1/κ, which in practice can reach down to a few nanometres36,37 or even below.81,82 On the other hand, one has to ensure that the penetration depth 1/κ is considerably larger than the interfacial width ζ [cf. Eq. (50)]; otherwise, the observation of interfacial fluctuations would be incomplete. Therefore, the condition to access interfacial properties fully is κζ ≪ 1, which is met by the typical experimental setups; in analytical and simulation work, it implies that the undamped limit, κ → 0, is considered.
B. Capillary wave divergence
C. Structure factor of an open slab of liquid
III. BULK REFERENCE WITH OPEN BOUNDARIES
A. Density–density correlation function
The bulk scattering follows by inserting this definition of a reference system [i.e., Eq. (15)] into the master formula [Eq. (1)], which is carried out in Sec. III B. Concerning simulations of GIXRD experiments, the same model for the bulk correlations can be used with the only modification that the nonzero width of the liquid film must be taken into account. Accordingly, the scattering depends on both the penetration depth 1/κ and the sample width L with the relevant, non-commuting limits κ → 0 and L → ∞ (see Sec. III C).
B. Scattering from a macroscopic half-space of bulk liquid
In this section, we turn to the specific setup of GIXRD experiments, i.e., a macroscopic liquid sample (L → ∞) and an evanescent wave on the liquid side (κ > 0). According to the master formula in Eq. (1), the GIXRD intensity for the bulk reference model [Eq. (15)] is the sum of two independent contributions stemming from the integrals over Gb(q, z, z′) in the domains z, z′ < 0 (bulk liquid) and z, z′ > 0 (bulk vapor). At low vapor pressure, the scattering from the latter side of the interface is negligible; alternatively, the result of Ref. 76 can be used directly because there is no damping of the propagating beam, i.e., f(z > 0) = 1.
The behavior of the bulk contribution Iℓ(q) is illustrated in Fig. 2 for a Lennard-Jones (LJ) liquid along the liquid–vapor coexistence line at two temperatures: T*≔kBT/ɛ = 0.70 slightly above the triple point temperature and T* = 1.15 in proximity of the critical temperature . The pair potential was truncated at rc = 3.5σ, and ɛ and σ refer to the interaction strength and range of the LJ potential, respectively; the phase diagram is shown in Fig. 3 and further details of the simulations are given in Appendix A. We calculated Iℓ(q) according to Eq. (22) from the bulk structure factors Sℓ(k) as input, using the same simulation data for Sℓ(k) as the ones in Ref. 76. The integration over kz in Eq. (22) was restricted to 0 ⩽ kz ⩽ kmax = 50/σ, and Sℓ(k) was extrapolated to such large wave numbers as described in Ref. 76. The curves obtained for a range of penetration depths (0.01 ⩽ κσ ⩽ 1) convincingly corroborate the convergence of (2κ/ϱℓ) Iℓ(q) to Sℓ(q) as κ → 0. The corrections, however, are significant for κσ ≳ 0.1: From Fig. 2, one infers that the value of Iℓ(q → 0) is increased over the value of Sℓ(q → 0) by a factor of up to for T* = 0.70, which is close to the triple point, whereas it is suppressed by a factor of about 1.8 at the higher temperature (T* = 1.15). Moreover, the minimum in Iℓ(q) near qσ ≈ 2 becomes more shallow for increasing κ and seems to disappear at low temperatures.
Contribution of the bulk liquid to the GIXRD intensity for five penetration depths κ−1 (in units of the LJ diameter σ) as calculated from Eq. (22). As input serve the simulated bulk structure factors Sℓ(k) of the (truncated) LJ liquids at temperatures T* = kBT/ɛ = 0.70 (a) and T* = 1.15 (b) in reduced units with the LJ energy scale ɛ. The normalization is taken such that in the limit κ → 0 the curves stay finite and approach Sℓ(q) [Eq. (24)].
Contribution of the bulk liquid to the GIXRD intensity for five penetration depths κ−1 (in units of the LJ diameter σ) as calculated from Eq. (22). As input serve the simulated bulk structure factors Sℓ(k) of the (truncated) LJ liquids at temperatures T* = kBT/ɛ = 0.70 (a) and T* = 1.15 (b) in reduced units with the LJ energy scale ɛ. The normalization is taken such that in the limit κ → 0 the curves stay finite and approach Sℓ(q) [Eq. (24)].
The liquid–vapor transition within the phase diagram of the truncated LJ fluid with interaction cutoff radius rc = 3.5σ. (a) Binodal line with the coexisting densities ϱℓ (dark blue disks) and ϱv (cyan blue disks) taken from fits to the density profile of the inhomogeneous system (Table II). Solid lines represent the critical law, ϱℓ/v(T) ≃ ϱsym(T) ± |t|β for t≔(T − Tc)/Tc ↑ 0 with the universal critical exponent β = 0.325 of the three-dimensional Ising universality class. The gray data points indicate the symmetry line of the binodal, i.e., ϱsym = (ϱℓ + ϱv)/2; its linear extrapolation to Tc (dashed line) yields the critical density ϱc. (b) Rectification of the critical law by plotting vs T for the same data as in (a); a linear regression to the three data points for T* ⩾ 1.0 yields the critical temperature . (c) Coexistence line in the pressure–temperature plane; symbols are simulation data with the disk marking the critical pressure Pc = (0.110 74 ± 0.000 03)ɛσ−3; the dotted line is a smooth interpolation of the simulation data.
The liquid–vapor transition within the phase diagram of the truncated LJ fluid with interaction cutoff radius rc = 3.5σ. (a) Binodal line with the coexisting densities ϱℓ (dark blue disks) and ϱv (cyan blue disks) taken from fits to the density profile of the inhomogeneous system (Table II). Solid lines represent the critical law, ϱℓ/v(T) ≃ ϱsym(T) ± |t|β for t≔(T − Tc)/Tc ↑ 0 with the universal critical exponent β = 0.325 of the three-dimensional Ising universality class. The gray data points indicate the symmetry line of the binodal, i.e., ϱsym = (ϱℓ + ϱv)/2; its linear extrapolation to Tc (dashed line) yields the critical density ϱc. (b) Rectification of the critical law by plotting vs T for the same data as in (a); a linear regression to the three data points for T* ⩾ 1.0 yields the critical temperature . (c) Coexistence line in the pressure–temperature plane; symbols are simulation data with the disk marking the critical pressure Pc = (0.110 74 ± 0.000 03)ɛσ−3; the dotted line is a smooth interpolation of the simulation data.
C. GIXRD intensity from a liquid slab of finite width
Snapshot of the simulated LJ liquid–vapor coexistence at T ≈ 0.8Tc, comprising 210 000 particles in total; in the vapor phase, only every tenth particle has been drawn for clarity. The yellow frames indicate the mean positions of the two planar interfaces, which delimit the liquid slab of width Ls = 25.0σ [see Eq. (B1)]; the area of each of the mean interfaces is (100σ)2. The lateral scattering vector q lies in the plane parallel to the interfaces (xy-plane). Along the direction normal to the interfaces (z-axis), the length of the simulation box is Lz = 125σ, but only a part of it is shown.
Snapshot of the simulated LJ liquid–vapor coexistence at T ≈ 0.8Tc, comprising 210 000 particles in total; in the vapor phase, only every tenth particle has been drawn for clarity. The yellow frames indicate the mean positions of the two planar interfaces, which delimit the liquid slab of width Ls = 25.0σ [see Eq. (B1)]; the area of each of the mean interfaces is (100σ)2. The lateral scattering vector q lies in the plane parallel to the interfaces (xy-plane). Along the direction normal to the interfaces (z-axis), the length of the simulation box is Lz = 125σ, but only a part of it is shown.
In Fig. 5, we test the expression in Eq. (30) for the GIXRD intensity scattered from finite-sized liquid samples at low (T* = 0.70) and high (T* = 1.15) temperatures. As for Fig. 2, we have employed the bulk structure factors Sℓ(q) obtained from MD simulations. Alternatively, we have calculated Iℓ(q) within the simulations and via Eq. (1), with the integration domain restricted to 0 ⩽ z, z′ ⩽ Lℓ. These simulation data are in excellent agreement with the results from Eq. (30) for all parameter sets. The top row of the panels [(a), (b)] shows the convergence of Iℓ(q) upon increasing the sample width L → ∞ at a fixed, exemplary penetration depth κσ = 0.01, whereas in the bottom row [(c), (d)] the limit κ → 0 is taken at the fixed sample width Lℓ = 20σ. We note that in the latter case (κ → 0, Lℓ fixed), the scattered intensity does not converge to the bulk structure factor Sℓ(q) (black lines), but rather to the slab structure factor S(q; Lℓ) (not shown), as expected from Eq. (37); the difference is and vanishes for macroscopic samples [see Eq. (13)].
Contribution to the GIXRD intensities collected from liquid slabs of finite width Lℓ and with open boundary conditions. Colored solid lines have been obtained from Eq. (33) using the bulk structure factor Sℓ(q) (black line) as input; symbols are results that follow from the direct evaluation of Eq. (1) within MD simulations for bulk liquids. The top panels [(a), (b)] show results for liquid slabs of four widths at fixed inverse penetration depth κσ = 0.01. The bottom panels [(c), (d)] show results for a liquid slab of fixed width Lℓ = 20σ with five penetration depths. The left panels [(a), (c)] refer to a liquid at the temperature T* = 0.70, i.e., close to the triple point, and the right panels [(b), (d)] correspond to T* = 1.15 near the liquid–vapor critical point. The thin horizontal lines at large q indicate the limits given in Eq. (34).
Contribution to the GIXRD intensities collected from liquid slabs of finite width Lℓ and with open boundary conditions. Colored solid lines have been obtained from Eq. (33) using the bulk structure factor Sℓ(q) (black line) as input; symbols are results that follow from the direct evaluation of Eq. (1) within MD simulations for bulk liquids. The top panels [(a), (b)] show results for liquid slabs of four widths at fixed inverse penetration depth κσ = 0.01. The bottom panels [(c), (d)] show results for a liquid slab of fixed width Lℓ = 20σ with five penetration depths. The left panels [(a), (c)] refer to a liquid at the temperature T* = 0.70, i.e., close to the triple point, and the right panels [(b), (d)] correspond to T* = 1.15 near the liquid–vapor critical point. The thin horizontal lines at large q indicate the limits given in Eq. (34).
D. Uniform asymptotic behavior
The term in the last line vanishes exponentially if the limit Lℓ → ∞ is taken first at fixed κ. For finite Lℓ, the remaining integral is bounded for all κ ⩾ 0 [e.g., by ] and it is the Fourier cosine transform of an integrable and continuous function and thus decays as for Lℓ → ∞ [see also Eq. (13)]. The limit κ → 0 can be interchanged with the integration over kz. Therefore, the last line of Eq. (40) behaves as .
IV. GIXRD FROM LIQUID–VAPOR INTERFACES
A. Wave number-dependent surface tension
B. Simple DFT models: Square-gradient approximation
Temperature dependence of the macroscopic surface tension γ0 along the liquid–vapor coexistence line from the triple point temperature to the critical temperature . The data points stem from MD simulations for truncated LJ fluids (rc = 3.5σ, Table I). The solid line depicts the critical scaling law γ0 ≃ Aγ|t|2ν for t≔(T − Tc)/Tc upon t ↑ 0 with the Ising exponent ν = 0.630. The inset shows the same data in a rectification plot of the critical law, yielding the critical temperature and the amplitude Aγ = (2.554 ± 0.008)ɛσ−2 from a linear regression to the three data points for T* ⩾ 1.0.
Temperature dependence of the macroscopic surface tension γ0 along the liquid–vapor coexistence line from the triple point temperature to the critical temperature . The data points stem from MD simulations for truncated LJ fluids (rc = 3.5σ, Table I). The solid line depicts the critical scaling law γ0 ≃ Aγ|t|2ν for t≔(T − Tc)/Tc upon t ↑ 0 with the Ising exponent ν = 0.630. The inset shows the same data in a rectification plot of the critical law, yielding the critical temperature and the amplitude Aγ = (2.554 ± 0.008)ɛσ−2 from a linear regression to the three data points for T* ⩾ 1.0.
Interfacial properties of truncated LJ fluids (rc = 3.5σ) along the liquid–vapor coexistence line. The surface tension γ0 was calculated from the anisotropy of the microscopic stress tensor. The length ℓ quantifies the contribution O(q2) to γ(q) and was obtained from fits of Eq. (65) to the data in Fig. 10(a). The interfacial width ζ was determined from fits of Eq. (B1) to the density profile of inhomogeneous systems with a mean interfacial area of (100σ)2. The numbers in parentheses give the statistical uncertainty in the last digit.
kBT/ɛ . | . | ℓ/σ . | ζ/σ . |
---|---|---|---|
0.70 | 0.868(2) | 0.23(2) | 0.528(2) |
0.80 | 0.664(2) | 0.26(2) | 0.642(2) |
0.90 | 0.473(1) | 0.37(2) | 0.802(2) |
1.00 | 0.294(2) | 0.53(2) | 1.048(3) |
1.10 | 0.137(2) | 0.85(2) | 1.548(4) |
1.15 | 0.069(2) | 1.24(2) | 2.123(5) |
kBT/ɛ . | . | ℓ/σ . | ζ/σ . |
---|---|---|---|
0.70 | 0.868(2) | 0.23(2) | 0.528(2) |
0.80 | 0.664(2) | 0.26(2) | 0.642(2) |
0.90 | 0.473(1) | 0.37(2) | 0.802(2) |
1.00 | 0.294(2) | 0.53(2) | 1.048(3) |
1.10 | 0.137(2) | 0.85(2) | 1.548(4) |
1.15 | 0.069(2) | 1.24(2) | 2.123(5) |
The relevance of the correction in Eq. (55) can be assessed by comparison with the small-q behavior of γ(q). To this end, we consider a simplified DFT treatment of the liquid–vapor interface based on the square-gradient approximation with the symmetric double-parabola potential.50 By defining γ(q) via the scattered intensity, for κ → 0 Parry et al.50 obtained . Accounting for the correction of the bulk scattering due to the open boundary [Eq. (59)] subtracts the contribution , i.e., it decreases the contribution O(q2) by 20%, leading to γ(q → 0)/γ0 ≃ 1 + (qξ)2. We note that adopting a different definition of the q-dependent surface tension solely based on the two-point density correlation function G(q, z, z′) would yield a different prefactor at order O(q2), and such a definition would not require as a prerequisite a model for the bulk scattering as input.50 However, to date, experimentally G(q, z, z′) is not directly accessible.
C. Hard sphere approximation for the bulk liquid
At temperatures close to the triple point, the liquid phase is characterized by a low compressibility and a small correlation length; in particular, the bulk structure is dominated by the interparticle repulsion. In order to estimate the value of in this regime, we approximate the liquid bulk structure factor by that of the Ashcroft–Lekner (AL) model for hard spheres;89 for simplicity, we ignore the contribution due to the attractive part of the pair potential.90 We recall that hard spheres do not form liquid or vapor phases; accordingly, there is no liquid–vapor interface. Nonetheless, one can expect that the AL model for Sℓ(k) renders a useful estimate of the bulk correction for dense and nearly incompressible liquids.
Leading-order correction to the bulk background [Eq. (26)] as obtained from quadratures of Eq. (14) with bulk structure factors Sℓ(k) taken from (i) the square-gradient DFT [Eq. (57)], (ii) the Ashcroft–Lekner model (Sec. IV C), and MD simulations of a LJ liquid at the temperatures (iii) T* = 1.15 (close to the critical one, ) and (iv) T* = 0.70 (close to the triple point ). The parameters of the theoretical models were chosen to correspond to the simulated liquids: (i) S0 = 1.18 and ξ = 1.25σ for the DFT model and (ii) at packing fraction η = 0.45 for the hard sphere model. Parts of the figure are reproduced from Ref. 76.
Leading-order correction to the bulk background [Eq. (26)] as obtained from quadratures of Eq. (14) with bulk structure factors Sℓ(k) taken from (i) the square-gradient DFT [Eq. (57)], (ii) the Ashcroft–Lekner model (Sec. IV C), and MD simulations of a LJ liquid at the temperatures (iii) T* = 1.15 (close to the critical one, ) and (iv) T* = 0.70 (close to the triple point ). The parameters of the theoretical models were chosen to correspond to the simulated liquids: (i) S0 = 1.18 and ξ = 1.25σ for the DFT model and (ii) at packing fraction η = 0.45 for the hard sphere model. Parts of the figure are reproduced from Ref. 76.
D. Molecular dynamics simulations of Lennard-Jones fluids
In Fig. 8, the decomposition of the scattered intensity into interfacial and background contributions [Eq. (6)] is illustrated (a) for the temperature T* = 0.70 close to the triple point and (b) for the temperature T* = 1.15 in proximity of the liquid–vapor critical point , both for a penetration depth 1/κ = 10σ. [The analogous decomposition of Stot(q) for T* = 1.15 is provided by Fig. 2 in Ref. 49.] For T* = 0.70, a liquid slab of width Ls = 25σ was simulated, but only particles with positions 0 ⩽ z ⩽ Lℓ = 20σ were admitted for the calculation of I(q; κ); here, z = 0 denotes the mean position of the interface. For the higher temperature, these values have been doubled in order to accommodate the much lower (macroscopic) surface tension and thus larger fluctuations of the local interface position. On the vapor side, only particles within slabs of Lv = 50σ (T* = 0.70) and Lv = 75σ (T* = 1.15) were taken into account. The background scattering was calculated according to Eq. (33) using the bulk structure factors Sℓ(q) and Sv(q) as input, which were obtained from separate simulations of homogeneous fluids (see Fig. 5). Close to the triple point [Fig. 8(a)], the CW divergence is clearly visible in the scattered intensity, I(q → 0; κ) ∼ q−2, without adjusting any parameter. To this end, the prefactor of the divergence was taken from Eq. (51) using the macroscopic surface tension γ0 (Fig. 6 and Table I) as determined independently from an integral over the stress tensor profile across the interface;64,91 the interfacial width ζ was obtained from the simulated mean density profiles ϱ(z) [see Eq. (B1)]. For the interfacial part of the scattering Iint(q; κ) ≈ H(q), this behavior of the CW divergence extends to a wide range of wave numbers qσ ≲ 2. At high temperature [Fig. 8(b)], the CW divergence is barely visible in I(q; κ) itself, but it can clearly be recognized in Iint(q; κ) for qσ ≲ 0.2. The slight mismatch between the predicted and the actual prefactors of the CW divergence (compare the thin and the thick black lines for qσ ≲ 0.2) disappears for larger penetration of the liquid side, e.g., κσ = 0.01. The mismatch is presumably due to higher-order terms in Eq. (4); the apparently obvious cause, that large-amplitude CWs are not probed properly for insufficiently small κ, is already accounted for in Eq. (51). The sizable deviations of Iint(q; κ) from the asymptotic behavior at large wave numbers give rise to the q-dependent surface tension.
Simulated GIXRD intensity I(q; κ) = Iint(q; κ) + Ib(q; κ) (red line) from scattering off the corresponding liquid–vapor interface and its decomposition into bulk and interface contributions Ib(q; κ) and Iint(q; κ), respectively (gray dashed and thick black lines); all quantities shown are normalized by Ib(q → ∞) [Eq. (48)]. The two panels show results at the reduced temperatures (a) T* = 0.70 and (b) T* = 1.15, both for a penetration depth 1/κ = 10σ. The thin black line indicates the CW divergence [Eq. (51)], I(q → 0; κ) ∼ 1/(γ0q2) [Eq. (51)], with the macroscopic surface tension γ0 obtained independently from the simulated stress tensor (Fig. 6 and Table I); the deviation of Iint(q; κ) from the CW divergence (thick vs thin black lines), which is well developed in panel (b), gives rise to the wave number-dependent surface tension γ(q). The bulk contribution Ib(q; κ) (gray dashed line) was calculated according to Eqs. (33) and (47), using the simulated bulk structure factors of the coexisting liquid (green line) and vapor (blue) phases Sℓ(q) and Sv(q), respectively; the data are extrapolated to small q assuming the Ornstein–Zernike form (green and blue dotted lines). For the calculation of I(q; κ), only particles within a slab of width (a) Lℓ = 20σ and (b) Lℓ = 40σ, respectively, were considered on the liquid side. For the quantities I(q), Sℓ(q), and Sv(q), lines connect actual simulation data (symbols, only shown for the five smallest wave numbers). In panel (b), the tiny wiggles in Iint(q) at qσ ≳ 1 reflect the statistical uncertainty of the simulation data.
Simulated GIXRD intensity I(q; κ) = Iint(q; κ) + Ib(q; κ) (red line) from scattering off the corresponding liquid–vapor interface and its decomposition into bulk and interface contributions Ib(q; κ) and Iint(q; κ), respectively (gray dashed and thick black lines); all quantities shown are normalized by Ib(q → ∞) [Eq. (48)]. The two panels show results at the reduced temperatures (a) T* = 0.70 and (b) T* = 1.15, both for a penetration depth 1/κ = 10σ. The thin black line indicates the CW divergence [Eq. (51)], I(q → 0; κ) ∼ 1/(γ0q2) [Eq. (51)], with the macroscopic surface tension γ0 obtained independently from the simulated stress tensor (Fig. 6 and Table I); the deviation of Iint(q; κ) from the CW divergence (thick vs thin black lines), which is well developed in panel (b), gives rise to the wave number-dependent surface tension γ(q). The bulk contribution Ib(q; κ) (gray dashed line) was calculated according to Eqs. (33) and (47), using the simulated bulk structure factors of the coexisting liquid (green line) and vapor (blue) phases Sℓ(q) and Sv(q), respectively; the data are extrapolated to small q assuming the Ornstein–Zernike form (green and blue dotted lines). For the calculation of I(q; κ), only particles within a slab of width (a) Lℓ = 20σ and (b) Lℓ = 40σ, respectively, were considered on the liquid side. For the quantities I(q), Sℓ(q), and Sv(q), lines connect actual simulation data (symbols, only shown for the five smallest wave numbers). In panel (b), the tiny wiggles in Iint(q) at qσ ≳ 1 reflect the statistical uncertainty of the simulation data.
Convergence of the effective q-dependent surface tension γ(q; κ → 0) as function of the inverse penetration depth κ for three, fixed wave numbers q at the temperature T* = 0.70. The data for γ(q; κ) were obtained via Eq. (52) from MD simulations for the GIXRD intensity, using only a fraction Lℓ = 20σ of the width Ls of the bulk liquid in order to calculate I(q) and carrying out the decomposition shown in Fig. 8(a). Thick solid lines indicate the corresponding limits γ(q) calculated from Stot(q), with setting κ = 0 directly within the simulations. Dashed lines show as obtained from Stot(q) when accounting only for the divergent part of the bulk scattering [see the main text and Eq. (63)]. The thin blue line (qσ/2π = 0.1) represents the κ-dependence of γ(q → 0; κ) as implied by Eq. (51).
Convergence of the effective q-dependent surface tension γ(q; κ → 0) as function of the inverse penetration depth κ for three, fixed wave numbers q at the temperature T* = 0.70. The data for γ(q; κ) were obtained via Eq. (52) from MD simulations for the GIXRD intensity, using only a fraction Lℓ = 20σ of the width Ls of the bulk liquid in order to calculate I(q) and carrying out the decomposition shown in Fig. 8(a). Thick solid lines indicate the corresponding limits γ(q) calculated from Stot(q), with setting κ = 0 directly within the simulations. Dashed lines show as obtained from Stot(q) when accounting only for the divergent part of the bulk scattering [see the main text and Eq. (63)]. The thin blue line (qσ/2π = 0.1) represents the κ-dependence of γ(q → 0; κ) as implied by Eq. (51).
It was this inconsistency between the analysis of scattering data (as obtained from experiments) and the total structure factor (as obtained within MD simulations) that gave rise to the refined treatment of the bulk background as elaborated here. The inconsistency would not be lifted by considering only the divergent part of the background in the analysis of GIXRD intensities I(q; κ), which upon κ → 0 would yield as introduced after Eq. (53). In retrospect, the issue is well understood by comparing Eqs. (44) and (45), which imply that . This clarifies that the O(1)-term must be included in the background contribution for a consistent analysis of GIXRD data. This yields γ(q) irrespective of whether it was calculated along route (i) or route (ii), described at the end of the first paragraph in Sec. IV D.
MD simulation results (symbols) for the wave number-dependent surface tension γ(q) of the truncated LJ liquid as obtained from Stot(q) after subtracting (a) the full background contribution [Eq. (64)] and (b) only its divergent part [Eq. (63)]; the data points in panel (a) are reproduced from Ref. 49. The differences are most apparent at the two lowest temperatures. In order to facilitate the comparison, the solid lines are the same in both panels: for T* ⩾ 0.9, the lines show fits of the small-q behavior as in Eq. (65) to the data for γ(q) [panel (a)]; for T* = 0.7 and 0.8, they represent empirical power law fits to the data for [panel (b)].
MD simulation results (symbols) for the wave number-dependent surface tension γ(q) of the truncated LJ liquid as obtained from Stot(q) after subtracting (a) the full background contribution [Eq. (64)] and (b) only its divergent part [Eq. (63)]; the data points in panel (a) are reproduced from Ref. 49. The differences are most apparent at the two lowest temperatures. In order to facilitate the comparison, the solid lines are the same in both panels: for T* ⩾ 0.9, the lines show fits of the small-q behavior as in Eq. (65) to the data for γ(q) [panel (a)]; for T* = 0.7 and 0.8, they represent empirical power law fits to the data for [panel (b)].
A related observation was made for the curvature-dependence of the macroscopic surface tension, upon replacing 1/q by the radius of a spherical droplet.92 The simulation results for ℓ(T) exhibit a temperature dependence similar to that of the OZ bulk correlation lengths of the coexisting liquid and vapor phases (Fig. 11). Moreover, anticipating the same critical scaling exponent as for the correlation length, ℓ(T ↑ Tc) ∼|T − Tc|−ν, the product γ0ℓ2 is expected to converge to a constant. Interestingly, our data suggest that γ0ℓ2/(kBT) → (4πω)−1 ≈ 0.09, where is a universal amplitude ratio;92–94 its most reliable estimate stems from Monte Carlo simulations of the three-dimensional Ising model:94 ω ≈ 0.87. This would imply that indeed ℓ(T)/ξℓ,v(T) → 1 as T ↑ Tc.
Temperature dependence of the length ℓ governing the small-wave number behavior of γ(q → 0) ≃ γ0[1 + (qℓ)2] in comparison to the bulk correlation lengths, ξℓ and ξv, of the coexisting liquid and vapor phases (Table I). The inset tests the convergence γ0ℓ2/(kBT) → (4πω)−1 ≈ 0.09 as T → Tc, where ω ≈ 0.87 is a universal amplitude ratio.92–94
Temperature dependence of the length ℓ governing the small-wave number behavior of γ(q → 0) ≃ γ0[1 + (qℓ)2] in comparison to the bulk correlation lengths, ξℓ and ξv, of the coexisting liquid and vapor phases (Table I). The inset tests the convergence γ0ℓ2/(kBT) → (4πω)−1 ≈ 0.09 as T → Tc, where ω ≈ 0.87 is a universal amplitude ratio.92–94
The right panel of Fig. 10(b) shows the corresponding results for , which have been obtained from the same input data, but taking into account only the divergent part of the bulk scattering [Eq. (63)] (we recall that changes sign as function of temperature, which is seen in Fig. 7). Whereas the data points shift slightly upward at higher temperatures, as expected from Eq. (55) due to at these temperatures [Eq. (57)], the repercussions are more significant at low temperatures (T* ≲ 0.80): opposed to the almost constant behavior of γ(q) up to qσ ≲ 2, bends downward, which results from in this case and due to the small value of ℓ. Empirically, the data are described by with and exponents α = 4 for T* = 0.80 and α = 2.5 for T* = 0.70. In particular, the data for suggest that there is a distinguished temperature T0 with such that ℓ(T) = 0 for T < T0, which appears to be implausible on physical grounds. This issue is removed by considering the full background scattering [Eq. (64)], which includes the correction given by and which leads to γ(q) as shown in Fig. 10(a).
E. Sensitivity to the mean interface position
So far, for the interpretation of simulation data, we have anticipated that the widths Lℓ and Lv of the coexisting phases [Eq. (27)] are known. Our protocol to construct the equilibrated inhomogeneous samples (see Appendix A) suggests a fixed ratio Lv: Lℓ (e.g., 3:1 or 4:1) of the bulk phases, where we set Lℓ = Ls and have chosen Ls = 25σ or 50σ for the width of the pre-equilibrated slabs, depending on temperature. However, due to interface broadening by capillary waves, these nominal values of Lℓ and Lv are in general (slightly) different from the values that are deduced from the inhomogeneous sample. For the latter step, various definitions of the mean interface position were proposed and are used in the literature.40,42,51,64,65,74
A common choice is based on Gibbs’ dividing surface (GDS), which in integral form is equivalent to ϱℓLℓ + ϱvLv = N/A; further, Lℓ + Lv = Lz is fixed to the extent of the simulation domain along the interface normal (i.e., the z-axis). Combining these two relations yields Lℓ for given N, A, ϱℓ, ϱv, and Lz. On the other hand, counting the particles in the inhomogeneous system, which was assembled from pre-equilibrated bulk slabs ( Appendix A), yields the same expression for N/A as the GDS criterion. Hence, Lℓ according to the GDS definition agrees with the nominal values for Lℓ.
In the present analysis, we followed a different approach and exploited the fact that there are two well-separated interfaces in the simulation setup: We have determined Lℓ from fits to the simulated mean density profile ϱsim(z) using an inflected sigmoidal function [Eq. (B1)]. The obtained values for Lℓ are very close to the nominal values according to the GDS definition; the absolute deviations are less than 0.1σ at all considered temperatures, i.e., a difference of less than 4‰. In addition to the values for Lℓ, the fits produced precise values of the coexisting densities, ϱℓ and ϱv, and the interfacial width ζ (see Tables I and II), which allowed us to obtain an accurate estimate of the liquid–vapor critical point (see Fig. 3 and Appendix B), although it is not in our focus here.
Bulk properties of truncated LJ fluids (rc = 3.5σ) along the liquid–vapor coexistence line. The densities ϱℓ and ϱv of the coexisting liquid and vapor phases were determined from the density profiles obtained in simulations of the inhomogeneous system. The corresponding pressures P, the isothermal compressibilities and , and the correlation lengths ξℓ and ξv stem from separate simulations of the bulk phases; the last four quantities were calculated from OZ fits to the bulk structure factors [Eq. (56)]. The numbers in parentheses give the measurement uncertainty in the last digit(s).
. | . | . | . | . | . | . | . |
---|---|---|---|---|---|---|---|
0.70 | 0.0022(5) | 0.8239(1) | 0.0032(3) | 0.0918(2) | 471(1) | 0.36(2) | 0.17(5) |
0.80 | 0.0063(2) | 0.7769(1) | 0.0085(2) | 0.1343(5) | 152(1) | 0.49(2) | 0.12(10) |
0.90 | 0.0168(1) | 0.7253(1) | 0.0215(1) | 0.208(1) | 69.2(2) | 0.56(2) | 0.39(2) |
1.00 | 0.0342(2) | 0.6658(1) | 0.0436(1) | 0.373(2) | 39.5(2) | 0.72(3) | 0.61(3) |
1.10 | 0.0623(2) | 0.5908(3) | 0.0842(4) | 0.89(2) | 29.1(3) | 0.95(10) | 0.91(7) |
1.15 | 0.0807(1) | 0.5403(1) | 0.1182(1) | 1.92(3) | 31.7(2) | 1.42(8) | 1.51(5) |
. | . | . | . | . | . | . | . |
---|---|---|---|---|---|---|---|
0.70 | 0.0022(5) | 0.8239(1) | 0.0032(3) | 0.0918(2) | 471(1) | 0.36(2) | 0.17(5) |
0.80 | 0.0063(2) | 0.7769(1) | 0.0085(2) | 0.1343(5) | 152(1) | 0.49(2) | 0.12(10) |
0.90 | 0.0168(1) | 0.7253(1) | 0.0215(1) | 0.208(1) | 69.2(2) | 0.56(2) | 0.39(2) |
1.00 | 0.0342(2) | 0.6658(1) | 0.0436(1) | 0.373(2) | 39.5(2) | 0.72(3) | 0.61(3) |
1.10 | 0.0623(2) | 0.5908(3) | 0.0842(4) | 0.89(2) | 29.1(3) | 0.95(10) | 0.91(7) |
1.15 | 0.0807(1) | 0.5403(1) | 0.1182(1) | 1.92(3) | 31.7(2) | 1.42(8) | 1.51(5) |
For γ(q), we have tested this procedure for the simulation results for γ(q) shown in Fig. 10(a). The data appear to follow a decay γ(q) ≃ γ0(bq)−2 for large q, albeit only in a small wave number window; such a decay for large q would correspond to a nonzero limit of the interfacial structure factor, H(q → ∞) > 0 [see Eq. (8) and Fig. 8]. From fits to the data for γ(q) in the range 2.6 ≲ qσ ≲ 3.8, we have obtained b = 0.41σ and 0.21σ at T* = 0.70 and 1.15, respectively. The assumed asymptotic form of γ(q) for large q is equivalent to H(q → ∞) ≃ kBT(Δϱ)2 b2/γ0 ≕ H∞, and setting δH(q → ∞) = −H∞ will remove such a spurious large-q contribution from H(q). The corresponding shift δL = H∞/Δϱ for the two temperatures renders δL = 0.11σ and δL = 0.30σ, respectively. However, shifting the mean interface position by such an amount has only a marginal influence on γ(q) and does not yield the desired qualitative change: that γ(q) bends upward (Fig. 12). Instead, γ(q), over the entire accessible range of wave numbers, is found to be rather robust against variations of Lℓ within a physically meaningful range.
Robustness of the q-dependent surface tension with respect to a variation of the width Lℓ of the liquid slab in the background contribution Sb(q) by an amount δL. Results for this, a posteriori adjusted quantity γadj(q) [Eqs. (66) and (67)], are shown for seven distinct values of δL and for the temperatures T* = 0.70 (disks) and T* = 1.15 (diamonds). These results are based on the data provided for γ(q) in Fig. 10(a).
Robustness of the q-dependent surface tension with respect to a variation of the width Lℓ of the liquid slab in the background contribution Sb(q) by an amount δL. Results for this, a posteriori adjusted quantity γadj(q) [Eqs. (66) and (67)], are shown for seven distinct values of δL and for the temperatures T* = 0.70 (disks) and T* = 1.15 (diamonds). These results are based on the data provided for γ(q) in Fig. 10(a).
The reason that the above choice for δL does not remove the apparent large-q decay of γ(q) can be understood by noting that the wave number window, within which one has γ(q) ≈ γ0(bq)−2, is still to the left of the first peak of the bulk structure (which is near qσ ≈ 6.8, see Fig. 8). There, Sℓ(q) ≈ Sv(q) ≈ 1 does not hold in this regime, which was used to deduce the relation δL = H∞/Δϱ. Therefore, in order to remove an apparent plateau in H(q) around a certain intermediate wave number q*, one has to consider the full q-dependence of δH(q) given in Eq. (66), which suggests to set δL = −H(q*)/[ϱℓSℓ(q*) − ϱvSv(q*)]. At T* = 0.70, reasonable estimates of Sv(q*) and Sℓ(q*) are given by their values for q → 0. Using H(q*) = H∞, one finds δL = −b2/Λ in terms of the length Λ [introduced after Eq. (68)]. This expression leads to δL ≈ 2.3σ, which is more than four times the interfacial width ζ and thus physically not plausible.
Based on recent insight into the resonance structure of interfacial two-point correlations, Parry and Rascón53–55 have proposed that the full wave number dependence of γ0/γ(q) is well approximated by a linear combination of the bulk structure factors Sℓ(q)/Sℓ(q → 0) and Sv(q)/Sv(q → 0) with suitable, weakly q-dependent weights to account for the liquid–vapor asymmetry. Moreover, these DFT studies reveal that H(q → ∞) ∼ q−2 (with the exception of the overly simplified square-gradient models, for which S(q → ∞) ∼ q−2 and thus H(q → ∞) ∼ q−4). Concerning the present MD simulation data, we conclude that a finding of H(q → ∞) = H∞ > 0 would indeed be in conflict with the above prediction. However, the observed decrease of γ(q) corresponds well with the increase of Sℓ(q) in the rising flank of its first peak (near qσ ≈ 4, see, e.g., Fig. 8). Moreover, the actual behavior of γ(q) for large wave number, i.e., qσ ≳ 5 cannot be obtained from the data due to unavoidable statistical noise. Thus, from our data one cannot rule out that the actual γ(q) has a small, positive limit γ∞≔γ(q → ∞) > 0 or, equivalently, that H(q → ∞) ∼ q−2—which would be consistent with the DFT calculations.
V. SUMMARY AND CONCLUSIONS
In sum, we have discussed the wave number dependence of the GIXRD intensity I(q; κ) due to scattering off liquid–vapor interfaces. We have proposed an unambiguous separation I(q; κ) = Iint(q; κ) + Ib(q; κ) into an interface-related contribution Iint(q; κ) and the bulk background Ib(q; κ), as illustrated in Fig. 8; κ is the inverse penetration length. The separation is based on a simple reference system for the coexisting bulk phases that avoids any assumption concerning the interfacial region. The essential ingredients are free boundary conditions for the bulk phases on both sides of the interface. This means that the reference system is composed of independent liquid and vapor phases and that their structures are identical to the respective bulk structures and are unperturbed by the presence of the interface. (Necessarily, such an idealized situation cannot occur in thermal equilibrium, but only on paper, because, capillary waves and other interfacial fluctuations would render any physical quantity to vary smoothly across the interface.) Accepting this simple reference model, it turns out that the background scattering Iℓ(q; κ) from, e.g., the liquid phase is not simply proportional to the structure factor Sℓ(q) of the bulk liquid [Eq. (26)]; rather, it is given as an integral over this function [Eq. (21) and Fig. 2]. This is a consequence of the free boundary conditions and appears likewise in the static structure factor of a liquid slab of finite width76 [Eqs. (12) and (13)]. We note that any “continuous” model for the background scattering, i.e., one that imposes a continuous interpolation of Gb(q, z, z′) across the interface, would require knowledge of the microscopic structure of the interfacial region, at least on the length scale over which the interpolation takes place. Already at the level of the local mean density, the question how to switch between the different correlation lengths on the liquid and the vapor sides has no obvious answer without providing microscopic details. As was shown in Sec. III A, the discontinuity of the background contribution and thus of the interfacial part of the two-point correlation function G(q, z, z′) is compatible with the asymptotically rigorous Wertheim–Weeks result for the CW divergence [Eq. (4)].
The interfacial part of the scattering yields the interfacial structural factor H(q) = Iint(q; κ → 0) for sufficiently deep sample penetration on the liquid side (κ−1 ≫ ζ, which is idealized as κ → 0). This expression of H(q) defines an effective, wave number-dependent surface tension γ(q) that is entirely based on density pair correlations [Eq. (8)]. Only for small wave numbers, qℓ(T) ≪ 1, the classical CW divergence, i.e., H(q) ∼ q−2, is observed in the scattered intensity because in this regime γ(q) ≃ γ0 [see Eq. (65) for the definition of ℓ(T)]. Here, we have shown that considering merely the singular part of the background contribution, Iℓ(q; κ) ≈ (ϱℓ/2κ)Sℓ(q), as usually done in the analysis of GIXRD data, results in a different interfacial structure factor and, correspondingly, in a different surface tension . In particular, the neglected background terms do not drop out in the limit κ → 0 but modify the surface tension at order O(q2) [Eq. (55)]. The magnitude of this difference is controlled by the correction integral , which is determined by the bulk structure factors [Eq. (14)] and which has the dimension of a length. Depending on the temperature, can have a positive (close to the triple point temperature Tt) or a negative sign (close to the critical temperature Tc). It turns out that at low temperatures, γ(q) and exhibit qualitatively different q-dependences (Fig. 10). At higher temperatures, the relative difference is diminished due to the emergence of a contribution of O(q2) in γ(q), which is characterized by a another length ℓ(T) that grows as T is increased [Eq. (65)].
Based on MD simulations for the truncated LJ fluid with cutoff distance rc = 3.5σ, we have presented evidence that ℓ(T) in fact diverges upon T ↑ Tc and that ℓ(T) approaches the bulk correlation lengths near criticality (Fig. 11). We observed further that the macroscopic surface tension γ0(T) of this truncated LJ fluid happens to be described very well by the critical scaling law along the whole coexistence line, from the triple point to the critical point (Fig. 6); the same observation was made earlier95,96 for a cutoff of rc = 2.5σ. Whereas the present study relies on LJ fluids as a generic test bed, analogous large-scale MD simulations could be performed for other substances, which would permit the direct comparison between the existing GIXRD data36–38 and the simulation results; a similar program was carried out successfully for the bulk structure of water.97
At large wave number, γ(q) is only mildly affected by a change in temperature, which together with the increase of ℓ(T) leads to a maximum in γ(q) at a certain wave number. This phenomenon was observed first in MD simulations49 and has since been put on firm theoretical ground by Parry et al.52–55 These theoretical studies use DFT calculations for exactly solvable models, which give insight into the structure of the two-point correlation function of the inhomogeneous fluid and which suggest that G(q, z, z′) can reliably be approximated using solely the bulk structure factors and related bulk properties.53,54 The related expressions can, in principle, be translated into quantitative predictions for the scattered intensity I(q; κ), which could be tested against its small-κ counterpart Stot(q) in the simulations [Eq. (62)]. The corresponding expressions, however, are complicated by the liquid–vapor asymmetry55 and are not yet readily available in an explicit form. Nevertheless, our data for γ(q), covering the full range of the wave number, are qualitatively consistent with the corresponding expectations based on the DFT approximations for G(q, z, z′); this includes the possibility that γ∞≔γ(q → ∞) > 0, which cannot be resolved from the available data. A complementary, first principles route to G(q, z, z′) has come into reach within a novel Barker–Henderson-like DFT treatment of inhomogeneous fluids,98 albeit such an endeavor may be technically challenging.
Experiments with phase-separated colloidal suspensions can, in principle, render knowledge of the three-dimensional positions of all colloids, given the tremendous advances in confocal microscopy during the past two decades, and thus provide experimental data for G(q, z, z′). In previous experiments on polymer–colloid dispersions,5 single scans of the focal plane perpendicular to the interface were used to obtain slices of the microscopic local density (compare Fig. 1). On this basis, capillary wave theory was then tested by assigning local interface positions and by calculating height–height correlation functions, closely resembling the traditional analysis of simulation data. Yet, the reconstruction of all three-dimensional particle positions from sequences of such focal scans appears to be an ambitious task.
Here, differential dynamic microscopy (DDM)99–101 offers an alternative: It is based on the correlation of intensity images and can yield similar information as contained in the interfacial structure factor H(q) discussed in the present study. DDM is also applicable to dense suspensions that scatter multiple times if the confocal mode of the microscope is used.102 In this case, the observation volume along the optical axis is restricted by the confocal depth, which introduces corrections in the obtained correlation functions that are analogous to the finite-κ and finite width effects discussed here and for bulk liquids.76 We expect that a refined interpretation of confocal DDM data, accounting for such corrections, can be developed along the lines presented here.
DDM is also a suitable tool for the characterization of motile suspensions, with micro-organisms or synthetic microswimmers as constituents.102–105 Despite being inherently out of equilibrium, such suspensions exhibit a motility-induced phase separation which shares certain universal features of the liquid–vapor transition.106,107 A surface tension and a surface stiffness have been associated with simulation data for such phenomena,108,109 although a debate about even the sign of the surface tension shows that active flows and mechanical contributions must be distinguished carefully in order to arrive at a consistent physical interpretation (see Ref. 110 and references therein). Similarly as for the equilibrium situation, a microscopic theory for the two-point density correlations in inhomogeneous active matter would be desirable in order to overcome the ambiguities associated with the notion of a fluctuating surface dividing the coexisting phases.
In a recent contribution, Hernández-Muñoz, Tarazona, and Chacón75 discuss the predictions of extended CW theory for surface diffraction at liquid–vapor interfaces with the fluctuating surface obtained from the intrinsic sampling method (ISM).41,73 The latter approach considers the many-particle structure in the interfacial region, which is accessible within simulations, in order to define a local interface position and, in this sense, goes beyond the mere use of pair correlations as considered here. Within both approaches to γ(q) (i.e., via multiparticle and via pair correlations), there is consensus that the “bending” contribution O(q2) to the q-dependent surface tension should be positive; in particular, this should also hold for almost incompressible liquids at temperatures close to the triple point. Yet, the ISM values for the corresponding length ℓ are considerably larger than what we have found here. We have tested whether this difference can be diminished by tuning the slab width Lℓ of the bulk liquid; Lℓ enters the expression for the background contribution in the simulations [Eq. (64)]. However, for the investigated LJ fluid we find that, for all accessible wave numbers, γ(q) responds only marginally to changes of Lℓ within a physically plausible range (Fig. 12). Thus, we can conclude that our findings for γ(q)—as obtained from density pair correlations—are robust with respect to the details of the definition of the mean interface position. Moreover, the resulting shape of γ(q) is consistent with theoretical predictions52–55 for G(q, z, z′). We note that a different, ISM-based definition of γ(q) renders46,74,75 γ(q) divergent for large q. We conjecture that this apparent controversy on the large-q behavior of γ(q) is a consequence of whether the definition of γ(q) contains implicit information about three- and many-body correlations or not. This claim is motivated, first, by noting that the ISM approach relies on this additional information whereas the present analysis of the simulation data is restricted to the use of two-point density correlations. Second, we recall the good agreement between the data for γ(q) as obtained along this route and the above-mentioned DFT calculations for γ(q), both using essentially the same definition of γ(q) in terms of G(q, z, z′). Finally, we note that GIXRD experiments on fluid interfaces merely probe two-point correlations, although confining the fluid in a disordered host lattice provides scope for GIXRD studies of higher-order correlations.111
In the context of extended CW theory, the physical interpretation of γ(q) is broader than serving just as a proxy for the interfacial two-point correlations; rather, γ(q) provides a mesoscopic characterization of liquid interfaces. In this picture,33–35,71 the interface is thought of as a sharp surface that is locally dressed in an intrinsic density profile perpendicular to the surface (interpolating between the coexisting bulk phases as if there are no CWs). The fluctuations of the local surface position are then governed by a corresponding surface Hamiltonian such that γ(q)q2 is the free energy cost associated with surface corrugations of wave number q (“capillary waves”). Naturally, such a mesoscopic description must break down at short distances (large wave numbers), for which the molecular discreteness becomes relevant. It has been demonstrated for solvable toy models50 that one cannot unambiguously single out the naked CW contribution to the local density fluctuations at O(q2) due to a nonlocal entanglement of the two; however, one can try to push the frontier as far as possible.72,74 Overall, we can state that the long and slow-burning controversy on the concept of the wave number-dependent surface tension has been resolved, but care must be taken to respect its limitations and to not compare apples with pears.
The nonanalytic contribution O(q2 log(q)) to γ(q) due to dispersion forces is within the scope of mesoscopic surface Hamiltonians and unambiguously identifiable.33,35 This contribution results in a minimum in γ(q) at mesoscopic wave numbers, corresponding to an enhancement of CW fluctuations. However, the magnitude of this effect is not well understood yet: The minimum was found to be surprisingly shallow in simulations of untruncated LJ fluids46 but sizable in experimental data from GIXRD on various liquid surfaces.36–38 For the latter, the correction discussed here has the potential to reduce the depth of the minimum to some extent, but we do not expect that it would qualitatively change the conclusions drawn from the experiments. For a direct comparison, simulation data for GIXRD on liquid surfaces of other substances than LJ fluids, e.g., water, would be of great value. It would also be of interest to highlight the role of dispersion forces in the interfacial density correlations, exploiting recent insight into their analytic structure.52–55
It is straightforward to extend the concepts developed here to fluid interfaces in phase-separating binary liquid mixtures,70,112–115 which would lay the basis for probing local changes of the composition (and its fluctuations) in the interfacial region. Within the Gaussian theory, the wave number-dependent surface tension γ(q) determines not only the fluctuations of the local interface height but also the fluctuations of the local interface normal.116 Therefore, in addition to GIXRD, the present results for γ(q) are relevant to a variety of further surface-specific experimental techniques, such as fluorescence spectroscopy, infrared spectroscopy and linear dichroism, generation of Maxwell displacement current (MDC), second-harmonic generation (SHG), direct measurement of the tilt angle distribution, and laser scanning confocal miscroscopy.116 Furthermore, it may prove fruitful to investigate the local density correlations under nonequilibrium conditions such as liquid–vapor interfaces in a temperature gradient.117–119 Eventually, the wave number-dependent relaxation dynamics of capillary waves and interfacial fluctuations16 may be probed within the framework put forward here. To this end, one merely needs to replace the static structure factors of the bulk by their corresponding intermediate scattering functions and to introduce a time lag between the factors of the two-point density correlation function [see Eq. (A1)].
SUPPLEMENTARY MATERIAL
See the supplementary material for the data needed to generate the published figures. See the README file within the archive.
ACKNOWLEDGMENTS
We thank Robert Evans, Andrew Parry, and Pedro Tarazona for helpful discussions and useful correspondence.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
F. Höfling: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Software (equal); Writing – original draft (equal); Writing – review & editing (equal). S. Dietrich: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article and its supplementary material.
APPENDIX A: MOLECULAR DYNAMICS SIMULATIONS OF LIQUID–VAPOR INTERFACES
Our present analytic results have been tested against large-scale MD simulations of LJ fluids with the pair potential truncated at the cutoff radius rc = 3.5σ, so that U(r > rc) = 0; the parameters ɛ and σ serve as units of energy and length, respectively. Two temperatures have been investigated in detail: T* = kBT/ɛ = 1.15, which is close to the liquid–vapor critical point , and T* = 0.70, slightly above the triple point temperature. Periodic boundary conditions were applied at all faces of the cuboid simulation box with its edge lengths chosen as Lx = Ly < Lz, so that stable, planar interfaces occur perpendicular to the z-axis. We used Lx = 100σ in order to obtain a large area of the mean interface and in order to access small wave numbers q, which must be integer multiples of 2π/Lx. The MD simulations were carried out in the canonical ensemble with the software “HAL’s MD package,”120 which exploits the massively parallel architecture of high-end graphics processors and which is well suited for the study of long-wavelength and low-frequency phenomena in liquids.49,76,114,121 (Concerning the relationship between the canonical and the grand canonical description of finite-size systems, see Ref. 122.) Initial and final particle configurations, time series of observables, as well as correlation functions such as I(q; κ) were stored efficiently in the H5MD file format.123 Further details on the simulations can be found in Ref. 49.
For the simulation of free liquid–vapor interfaces, a sufficiently thick film of bulk liquid is placed within the simulation domain, and the remaining space is filled with the coexisting vapor phase as to form two parallel, planar liquid–vapor interfaces (Fig. 4). The detailed protocol was as follows:
Determine the coexisting liquid and vapor densities at the prescribed temperature.
Equilibrate the bulk phases of liquid and vapor independently, using slab-like, periodic boxes of width Ls = 25σ (T* ⩽ 1.0) or 50σ (T* ⩾ 1.1).
Assemble the two phases, after squeezing the different configurations slightly in order to avoid particle overlaps at the boundaries between the phases (e.g., by an amount of 0.5σ along the z-direction via rescaling of the positions).
Equip the assembled system with periodic boundaries and let it relax to form an inhomogeneous fluid in equilibrium.
For steps (i) and (ii), only relatively short simulation runs are needed, whereas much longer simulations are required for step (iv) in order to ensure equilibration of the capillary waves, especially at small wave numbers. In step (iii), we combined several replicas of the vapor phase with one slab of liquid so that the overall box size was Lz = 5 × 25σ at low temperatures and Lz = 4 × 50σ at high temperatures, yielding a total number of particles of N = 209 300 and 447 000, respectively.
APPENDIX B: COEXISTENCE LINE AND LIQUID–VAPOR CRITICAL POINT
The latter observation is phenomenologically known as the “law of rectilinear diameter.”124,125 From the linear extrapolation of ϱsym(T) to Tc, the critical density was found to be ϱc = (0.318 ± 0.001)σ−3. In a more refined analysis of the critical behavior of the coexisting densities, it is argued that the slope of the curve ϱsym(T) vs T is proportional to the isochoric specific heat cV, so that one expects a singular dependence:124,126,127 as T ↑ Tc, where α ≈ 0.110 is the Ising universal exponent of the specific heat. However, only the two data points for ϱsym(T) closest to Tc (i.e., for T* ⩾ 1.10) are compatible with this singular scaling law—the linear law (i.e., mean-field like with α = 0) provides a better description of the data. A similar observation was made before for other simple fluids (see, e.g., Refs. 124 and 125); yet, it is particularly surprising here, given that the T-dependences of Δϱ and γ0 were very well captured by their corresponding critical laws over a wide range of temperatures [Figs. 3(b) and 6]. These findings underscore that the true scaling behavior sets in only asymptotically for T ↑ Tc. In particular, the temperature dependence of cV is found to be non-monotonic along the liquid branch of the coexistence line and has its minimum near T* ≈1.05 (data not shown; this calls for future research).
Along the transition line, we have also computed the pressures of the coexisting liquid and vapor phases from the bulk simulations [Table II and Fig. 3(c)], which served as a consistency check. Eventually, the critical pressure was obtained from a separate simulation of the bulk fluid at the quoted critical point (Tc, ϱc), yielding Pc = (0.110 74 ± 0.000 03)ɛσ−3.
REFERENCES
On the vapor side, the integral in Eq. (1) is divergent. However, this issue is present neither in the simulations, due to the finite size of the simulation box, nor in experiments, due to the (weak) absorption of x rays in the vapor phase.