This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Structural properties of additive binary hard-sphere mixtures. II. Asymptotic behavior and structural crossovers

Sławomir Pieprzyk [email protected] Institute of Molecular Physics, Polish Academy of Sciences, M. Smoluchowskiego 17, 60-179 Poznań, Poland    Santos B. Yuste [email protected] http://www.unex.es/eweb/fisteor/santos/    Andrés Santos [email protected] http://www.unex.es/eweb/fisteor/andres/ Departamento de Física and Instituto de Computación Científica Avanzada (ICCAEx), Universidad de Extremadura, Badajoz, E-06006, Spain    Mariano López de Haro [email protected] Instituto de Energías Renovables, Universidad Nacional Autónoma de México (U.N.A.M.), Temixco, Morelos 62580, Mexico    Arkadiusz C. Brańka [email protected] Institute of Molecular Physics, Polish Academy of Sciences, M. Smoluchowskiego 17, 60-179 Poznań, Poland
Abstract

The structural properties of additive binary hard-sphere mixtures are addressed as a follow-up of a previous paper [S. Pieprzyk et al., Phys. Rev. E 101, 012117 (2020)]. The so-called rational-function approximation method and an approach combining accurate molecular dynamics simulation data, the pole structure representation of the total correlation functions, and the Ornstein–Zernike equation are considered. The density, composition, and size-ratio dependencies of the leading poles of the Fourier transforms of the total correlation functions hij(r)h_{ij}(r) of such mixtures are presented, those poles accounting for the asymptotic decay of hij(r)h_{ij}(r) for large rr. Structural crossovers, in which the asymptotic wavelength of the oscillations of the total correlation functions changes discontinuously, are investigated. The behavior of the structural crossover lines as the size ratio and densities of the two species are changed is also discussed.

I Introduction

The close connection between the thermodynamic properties and the structural correlation functions of fluids in the statistical-mechanical formulation of liquid state theory is well known. In particular, for a fluid mixture of NcN_{c} components, the virial route to the equation of state leads to [1, 2, 3]

Z\displaystyle Z\equiv pρkBT\displaystyle\frac{p}{\rho k_{B}T}
=\displaystyle= 1ρ6kBTi,j=1Ncxixj𝑑𝐫gij(r)rϕij(r)r,\displaystyle 1-\frac{\rho}{6k_{B}T}\sum_{i,j=1}^{N_{c}}x_{i}x_{j}\int d\mathbf{r}\,g_{ij}(r)r\frac{\partial\phi_{ij}(r)}{\partial r}, (1)

where ZZ is the compressibility factor, pp is the pressure, ρ\rho is the number density, kBk_{B} is the Boltzmann constant, TT is the absolute temperature, xix_{i} is the mole fraction of molecules of species ii, ϕij(r)\phi_{ij}(r) is the interaction potential (assumed to be spherically symmetric and pairwise additive) between a particle of species ii and a particle of species jj, and gij(r)g_{ij}(r) is the radial distribution function (RDF), which is a measure of the probability of finding a molecule of species ii at a distance rr from another molecule of species jj. In the case of hard spheres, Eq. (I) reduces to

Z=1+23πρi,j=1Ncxixjσijgij(σij),Z=1+\frac{2}{3}\pi\rho\sum_{i,j=1}^{N_{c}}x_{i}x_{j}{\sigma_{ij}}g_{ij}(\sigma_{ij}), (2)

where σij\sigma_{ij} is the hard-core diameter of the interaction between a sphere of species ii and another sphere of species jj, and gij(σij)g_{ij}(\sigma_{ij}) is the contact value of the RDF.

On the other hand, from the compressibility route to the equation of state one gets the relationship [3]

χ1\displaystyle\chi^{-1}\equiv 1kBT(pρ)T\displaystyle\frac{1}{k_{B}T}\left(\frac{\partial p}{\partial\rho}\right)_{T}
=\displaystyle= i,j=1Ncxixj[𝖨+𝗁^(0)]ij1\displaystyle\sum_{i,j=1}^{N_{c}}\sqrt{x_{i}x_{j}}\left[\mathsf{I}+\widehat{\mathsf{h}}(0)\right]^{-1}_{ij}
=\displaystyle= 1ρi,j=1Ncxixjc~ij(0),\displaystyle 1-\rho\sum_{i,j=1}^{N_{c}}{x_{i}x_{j}}\widetilde{c}_{ij}(0), (3)

where 𝖨\mathsf{I} is the Nc×NcN_{c}\times N_{c} identity matrix and the element h^ij(k)ρxixjh~ij(k)\widehat{h}_{ij}(k)\equiv\rho\sqrt{x_{i}x_{j}}\widetilde{h}_{ij}(k) of the matrix 𝗁^(k)\widehat{\mathsf{h}}(k) is proportional to the Fourier transform

h~ij(k)=𝑑𝐫eı𝐤𝐫hij(r)\widetilde{h}_{ij}(k)=\int d\mathbf{r}\,e^{-\imath\mathbf{k}\cdot\mathbf{r}}h_{ij}\left({r}\right) (4)

of the total correlation functions hij(r)gij(r)1h_{ij}({r})\equiv g_{ij}(r)-1, ı\imath being the imaginary unit. Further, in the last equality of Eq. (I), c~ij(0)\widetilde{c}_{ij}(0) is the zero wavenumber limit of the Fourier transform

c~ij(k)=𝑑𝐫eı𝐤𝐫cij(r)\widetilde{c}_{ij}(k)=\int d\mathbf{r}\,e^{-\imath\mathbf{k}\cdot\mathbf{r}}c_{ij}\left({r}\right) (5)

of the direct correlation function cij(r)c_{ij}({r}). The latter is defined through the Ornstein–Zernike (OZ) relation, namely,

hij(r12)=cij(r12)+ρ=1Ncx𝑑𝐫3ci(r13)hj(r23),h_{ij}(r_{12})=c_{ij}(r_{12})+\rho\sum_{\ell=1}^{N_{c}}x_{\ell}\int d\mathbf{r}_{3}\,c_{i\ell}(r_{13})h_{\ell j}(r_{23}), (6a)
h~ij(k)=c~ij(k)+ρ=1Ncxc~i(k)h~j(k),\widetilde{h}_{ij}(k)=\widetilde{c}_{ij}(k)+\rho\sum_{\ell=1}^{N_{c}}x_{\ell}\widetilde{c}_{i\ell}({k})\widetilde{h}_{\ell j}({k}), (6b)

in real and Fourier spaces, respectively.

Apart from this thermodynamic connection, it has been further established that abrupt changes in the structural correlation functions of a fluid may also show up in its phase behavior. This is the case of the Fisher–Widom line of simple fluids [4, 5, 6, 7, 8, 9, 10, 11], which distinguishes between the region where the large-rr behavior of the total correlation function shows damped oscillatory decay (typical of dense and/or high-temperature liquids) and the region where the nature of the decay is monotonic (typical of low-density gases, low-temperature liquids, and near-critical fluids). Structural transitions such as the one related to the Fisher–Widom line, including the physics behind them, are clearly of interest but their study is hampered by the lack of exact results for the structural correlation functions and, despite the availability of interesting work on this subject, additional efforts are clearly required, especially in the case of mixtures.

In a previous paper [12], hereafter referred to as paper I, we presented a method (denoted as the WM scheme) allowing us to obtain an accurate representation of the structural correlation functions of additive binary hard-sphere (BHS) mixtures. The WM method successfully combines molecular dynamics (MD) simulation data, residue-theorem analysis, and the OZ relations, additionally taking into account the tail parts of the structural correlation functions, without using any approximate closures. In particular, by considering a mixture with a fixed diameter ratio of 0.6480.648 and a fixed total packing fraction of 0.50.5 (which was the system analyzed previously theoretically and through experimental data by Statt et al. [13]), we confirmed in paper I the presence of structural crossovers in such a mixture and examined the role played in the crossover by the first two poles of the Fourier transforms of the total correlation functions. We also found very good agreement between the results of the new WM method and those obtained from the use of the rational-function approximation (RFA) [14, 15, 3, 16] to compute analytically the total correlation functions, as well as an improvement of the agreement between the RFA and WM results and the ones derived from experimental data when compared to the analysis performed in Ref.  [13].

The aim of the present paper is two-fold. On the one hand, to consolidate the RFA approach as a valuable tool to investigate asymptotic behavior and structural crossover issues, by testing such approach against the results of the WM method. On the other hand, to carry out a more thorough analysis of the role of the pole structure of h~ij(k)\widetilde{h}_{ij}(k) on the asymptotic behavior of hij(r)h_{ij}(r) and the structural crossovers in these functions by considering various BHS mixtures.

The paper is organized as follows. The system of interest (BHS mixtures) is briefly described in Sec. II, where also succinct accounts of the RFA approach and of the WM scheme developed in paper I, as well as some details of our MD simulations, are presented. In Sec. III we provide the basics of the analysis of the poles of the total correlation functions, while in Sec. IV we provide the results of our calculations and an illustration of our main findings. The paper closes in Sec. V with some concluding remarks.

II Structural properties of a binary hard-sphere mixture

II.1 System

Let us consider a binary (Nc=2N_{c}=2) fluid mixture of “small” (label ss) and “big” (label bb) hard spheres. The additive hard core of the interaction between a sphere of species ii and a sphere of species jj (i,j=s,bi,j=s,b) is σij=12(σi+σj)\sigma_{ij}=\frac{1}{2}(\sigma_{i}+\sigma_{j}), where the diameter of a sphere of species ii is σii=σi\sigma_{ii}=\sigma_{i}. Let the number density of the mixture be ρ\rho, the mole fraction of species ii be xi=ρi/ρx_{i}=\rho_{i}/\rho (where ρi=Ni/V\rho_{i}=N_{i}/V is the partial number density, NiN_{i} and VV being the number of particles of species ii and the volume of the system, respectively), and let the size ratio be q=σs/σb1q=\sigma_{s}/\sigma_{b}\leq 1. From these quantities one can define the partial packing fractions ηi=π6ρiσi3\eta_{i}=\frac{\pi}{6}\rho_{i}\sigma_{i}^{3} and the total packing fraction η=ηs+ηb=π6ρσb3(xsq3+xb)\eta=\eta_{s}+\eta_{b}=\frac{\pi}{6}\rho\sigma_{b}^{3}(x_{s}q^{3}+x_{b}). Note that in this system there are three characteristic separations between particles at contact, namely the small-small particle separation, σs=qσb\sigma_{s}=q\sigma_{b}, the small-big particle separation, σsb=12σb(1+q)\sigma_{sb}=\frac{1}{2}\sigma_{b}(1+q), and the big-big particle separation, σb\sigma_{b}.

II.2 The rational-function approximation method

In order to examine the structural properties, we shall now sketch the RFA approach to obtain the structural properties of additive hard-sphere mixtures. The detailed description of such an approach may be found in Refs.  [14, 15, 3, 16]. First, the Laplace transforms of rgij(r)rg_{ij}(r) are introduced:

Gij(z)0𝑑rezrrgij(r).G_{ij}(z)\equiv\int_{0}^{\infty}dr\,e^{-zr}rg_{ij}(r). (7)

Next, an explicit form for Gij(z)G_{ij}(z) in terms of a free parameter ξ\xi and an Nc×NcN_{c}\times N_{c} matrix 𝖫(z)=𝖫(0)+𝖫(1)z+𝖫(2)z2\mathsf{L}(z)=\mathsf{L}^{(0)}+\mathsf{L}^{(1)}z+\mathsf{L}^{(2)}z^{2} is proposed. Then, by imposing certain consistency conditions, the elements of the matrices 𝖫(0)\mathsf{L}^{(0)}, 𝖫(1)\mathsf{L}^{(1)}, 𝖫(2)\mathsf{L}^{(2)} are expressed as linear functions of ξ\xi. In particular, Lij(2)=2πξσijgij(σij)L_{ij}^{(2)}=2\pi\xi\sigma_{ij}g_{ij}(\sigma_{ij}).

Interestingly, the simple choice ξ=0\xi=0, and hence Lij(2)=0L_{ij}^{(2)}=0, gives the Percus–Yevick (PY) solution [17, 1], which is known to yield different equations of state via the virial [Eq. (2)] and compressibility [Eq. (I)] routes. However, by an appropriate determination of ξ0\xi\neq 0, the RFA can be made thermodynamically consistent and, additionally, allows one to freely choose the contact values gij(σij)g_{ij}(\sigma_{ij}). A convenient choice is provided by the popular BGHLL expression proposed independently by Boublík [18], Grundke and Henderson [19], and Lee and Levesque [20].

Clearly, once Gij(z)G_{ij}(z) has been determined, inverse Laplace transformation directly yields rgij(r)rg_{ij}(r) and hence also hij(r)h_{ij}(r). On the other hand, explicit knowledge of Gij(z)G_{ij}(z) also allows us to determine the Fourier transform h~ij(k)\widetilde{h}_{ij}(k) through the relation

h~ij(k)=2πGij(z)Gij(z)z|z=ık.\widetilde{h}_{ij}(k)=-2\pi\left.\frac{G_{ij}(z)-G_{ij}(-z)}{z}\right|_{z=\imath k}. (8)

In the RFA (as well as in the PY approximation), Gij(z)G_{ij}(z) is obtained from the inner product of the matrix 𝖫(z)\mathsf{L}(z) and the inverse of another related matrix 𝖡(z)\mathsf{B}(z). Therefore, the Laplace transforms Gij(z)G_{ij}(z) for all the pairs ijij share the same poles, namely the zeros of the determinant D(z)D(z) of 𝖡(z)\mathsf{B}(z). In the particular case of a binary mixture, the functional form of D(z)D(z) is

D(z)=\displaystyle D(z)= 𝒫6(0)(z)+𝒫4(s)(z)eσsz+𝒫4(b)(z)eσbz\displaystyle\mathcal{P}_{6}^{(0)}(z)+\mathcal{P}_{4}^{(s)}(z)e^{-\sigma_{s}z}+\mathcal{P}_{4}^{(b)}(z)e^{-\sigma_{b}z}
+𝒫2(sb)(z)e2σsbz,\displaystyle+\mathcal{P}_{2}^{(sb)}(z)e^{-2\sigma_{sb}z}, (9)

where 𝒫6(0)(z)\mathcal{P}_{6}^{(0)}(z), 𝒫4(s)(z)\mathcal{P}_{4}^{(s)}(z), 𝒫4(b)(z)\mathcal{P}_{4}^{(b)}(z), and 𝒫2(sb)(z)\mathcal{P}_{2}^{(sb)}(z) are polynomials of degrees 66, 44, 44, and 22, respectively. In the PY case (ξ=0\xi=0), the degrees of those polynomials decrease in two units, i.e., 𝒫6(0)(z)𝒫4(0)(z)\mathcal{P}_{6}^{(0)}(z)\to\mathcal{P}_{4}^{(0)}(z), 𝒫4(s)(z)𝒫2(s)(z)\mathcal{P}_{4}^{(s)}(z)\to\mathcal{P}_{2}^{(s)}(z), 𝒫4(b)(z)𝒫2(b)(z)\mathcal{P}_{4}^{(b)}(z)\to\mathcal{P}_{2}^{(b)}(z), and 𝒫2(sb)(z)const\mathcal{P}_{2}^{(sb)}(z)\to\text{const}. A basic property of D(z)D(z) is D(0)=0D(0)=0, which is tied to the physical condition limrgij(r)=1\lim_{r\to\infty}g_{ij}(r)=1.

It should be stressed that perhaps the most valuable asset of the RFA approach is that, apart from ensuring thermodynamic consistency, it leads to explicit analytic expressions for all the structural properties of the BHS mixture [21]. Additionally, the asymptotic long-range behavior of hij(r)h_{ij}(r) is directly obtained from the roots of Eq. (II.2).

II.3 The WM method

The WM method proposed in paper I [12] allows us to obtain the structural properties of additive BHS mixtures by combining accurate MD simulation data, the pole structure representation of the total correlation functions, and the OZ equation.

In the method, a semi-empirical approximation for the structural properties of additive BHS mixtures is constructed by considering the following analytic form of hij(r)h_{ij}({r}):

hijWM(r)={1,0<r<σij,n=1Wbij(n)rn1,σij<rrijmin,n=1MAij(n)reαnrsin(ωnr+δij(n)),rrijmin.h^{WM}_{ij}({r})=\begin{cases}-1,\qquad 0<r<\sigma_{ij},\\ \sum\limits_{n=1}^{W}b_{ij}^{(n)}r^{n-1},\qquad\sigma_{ij}<r\leq r_{ij}^{{\min}},\\ \sum\limits_{n=1}^{M}\frac{A_{ij}^{(n)}}{r}e^{-\alpha_{n}r}\sin\left(\omega_{n}r+\delta_{ij}^{(n)}\right),\quad r\geq r_{ij}^{{\min}}.\end{cases} (10)

The parameters {bij(1),bij(2),,bij(W)}\{b_{ij}^{(1)},b_{ij}^{(2)},\ldots,b_{ij}^{(W)}\} and {Aij(1),α1,ω1,δij(1),,Aij(M),αM,ωM,δij(M)}\{A_{ij}^{(1)},\alpha_{1},\omega_{1},\delta_{ij}^{(1)},\ldots,A_{ij}^{(M)},\alpha_{M},\omega_{M},\delta_{ij}^{(M)}\} are obtained by enforcing the BGHLL contact values and the continuity conditions at r=rijminr=r_{ij}^{{\min}}, and by a nonlinear fitting procedure based on the minimization of |hijWM(r)hijMD(r)|\left|h_{ij}^{WM}(r)-h_{ij}^{\text{MD}}(r)\right| for each r/σij(1,rc)r/\sigma_{ij}\in(1,r_{c}^{*}), where hijMD(r)h_{ij}^{\text{MD}}(r) is obtained from our MD simulations, the details of which will be specified below. In our fitting procedure, the values of |hijWM(r)hijMD(r)||h_{ij}^{WM}(r)-h_{ij}^{\text{MD}}(r)| for 1<r/σij<rc1<r/\sigma_{ij}<r_{c}^{*} were required to be smaller than 10310^{-3}. The suitable value of rcr_{c}^{*} depends on the size ratio qq and we took rc=5r_{c}^{*}=5, 44, 33, and 33 for q=0.648q=0.648, 0.40.4, 0.30.3, and 0.20.2, respectively. This value is connected with the half-length of the simulation box. Note that decreasing qq causes a decrease of the available space (smaller simulation box). Therefore, in order to carry out simulations and increase the simulation box with the assumed rcr_{c}^{*} values, it was necessary to add more spheres in the cases q=0.3q=0.3 and 0.20.2. For our calculations an appropriate choice for rijminr_{ij}^{{\min}} was the position of the first minimum of hij(r)h_{ij}({r}). Also, as in paper I, for the subsequent calculations we will usually take W=15W=15 and M=10M=10.

Table 1: BHS mixtures investigated by MD in this work.
qq ηb\eta_{b} ηs\eta_{s}
0.6480.648 0.100.10 0.200.20, 0.250.25, 0.300.30, 0.350.35, 0.400.40
0.200.20 0.150.15, 0.200.20, 0.250.25, 0.300.30, 0.350.35
0.40.4 0.050.05 0.050.05, 0.100.10, 0.150.15, 0.200.20, 0.250.25, 0.300.30, 0.350.35
0.200.20 0.050.05, 0.100.10, 0.150.15, 0.200.20, 0.250.25
0.30.3 0.050.05 0.100.10, 0.150.15, 0.200.20, 0.250.25, 0.300.30, 0.350.35
0.200.20 0.100.10, 0.150.15, 0.200.20
0.20.2 0.020.02 0.250.25
0.200.20 0.100.10

II.4 Details of the molecular dynamics simulations

The computation of hijMD(r)h_{ij}^{\text{MD}}(r) was performed with the DynamO program [22] for the size ratios q=0.648q=0.648, 0.40.4, 0.30.3 and 0.20.2, and different values of the partial packing fractions (ηb,ηs)(\eta_{b},\eta_{s}), which were chosen according to the investigated size ratio to examine a substantial part of the phase diagram (including low, moderate, and high total densities) of the BHS mixture. More specifically, the values of the partial packing fractions studied by MD are given in Table 1 and denoted in Fig. 3 as open yellow circles.

In order to reduce sufficiently finite-size effects and the statistical errors in the simulations, the data of hijMD(r)h_{ij}^{\text{MD}}(r) for r/σij<rcr/\sigma_{ij}<r_{c}^{*} must be obtained from long simulations with a large number of particles (N104)(N\sim 10^{4}). It has been checked that 16 38416\,384 particles were sufficient to obtain reasonably accurate data for the size ratios q=0.648q=0.648 and 0.40.4. In the cases q=0.3q=0.3 and 0.20.2, due to the reduction of the size of the simulation box, the systems were investigated with 48 66848\,668 particles. The histogram grid size of hijMD(r)h_{ij}^{\text{MD}}(r) was set to δr/σij=0.01\delta r/\sigma_{ij}=0.01, which was found to be a suitable choice to balance finite-size effects and statistical errors.

The MD simulations were carried out typically for a total number of 2×1092\times 10^{9} collisions, and the statistical uncertainty of hijMD(r)h_{ij}^{\text{MD}}(r) was obtained with the block averaging method [23]. For each density, and in the whole range r/σij(1,rc)r/\sigma_{ij}\in(1,r_{c}^{*}), the accuracy of hijMD(r)h_{ij}^{\text{MD}}(r) was such that the estimated uncertainty was typically smaller than 10310^{-3}, being up to 0.0050.005 near contact (for the highest densities) and becoming less than 0.00020.0002 at larger particle separations. For large systems, the finite-size effects in the MD calculations of the RDF arise mainly from fixing the particle number, i.e., from the relation between canonical and grand-canonical ensembles. The corrections required to convert data from the MD simulations to the canonical ensemble are of 𝒪(1/N2)\mathcal{O}(1/N^{2}) [24, 25], which are negligible here. Also, it was checked for a few densities that the remaining part of the correction factor involving density derivatives was smaller than the obtained data accuracy and therefore could be neglected.

q=0.648q=0.648                          q=0.625q=0.625                          q=0.600q=0.600                          q=0.500q=0.500             
Refer to caption      Refer to caption      Refer to caption      Refer to caption
    q=0.425q=0.425                          q=0.400q=0.400                          q=0.375q=0.375                          q=0.350q=0.350             
Refer to caption      Refer to caption      Refer to caption      Refer to caption
    q=0.315q=0.315                          q=0.300q=0.300                          q=0.275q=0.275                          q=0.250q=0.250             
Refer to caption      Refer to caption      Refer to caption      Refer to caption
Refer to caption

Figure 1: RFA predictions for the contour plots of the (reduced) oscillation frequency ωσb/2πσb/λ\omega\sigma_{b}/2\pi\equiv\sigma_{b}/\lambda corresponding to the leading pole for decreasing representative values of the size ratio (from q=0.648q=0.648 to q=0.250q=0.250). The colormap in the bottom bar indicates the color code for the values of ωσb/2π\omega\sigma_{b}/2\pi. In each panel, the (black) dashed diagonal line represents the locus ηs+ηb=0.5\eta_{s}+\eta_{b}=0.5. The (colored) stars and circles indicate the end points and the splitting points, respectively. Note that ns=2{n_{s}}=2 for q=0.648q=0.648, 0.6250.625, 0.6000.600, 0.5000.500 and 0.4250.425, ns=3{n_{s}}=3 for q=0.375q=0.375, 0.3500.350, 0.3150.315, and 0.3000.300, and ns=4{n_{s}}=4 for q=0.275q=0.275 and 0.2500.250.
Refer to caption
Figure 2: The qq-evolution of the end point (colored stars) and the splitting point (colored circles) of the generations G0–G3. The arrows signal the evolution as qq decreases. The dashed tie lines connect end and splitting points at the same qq.

q=0.2q=0.2                             q=0.3q=0.3                             q=0.4q=0.4                             q=0.648q=0.648
Refer to caption      Refer to caption      Refer to caption      Refer to caption
Refer to caption      Refer to caption      Refer to caption      Refer to caption
Refer to caption

Figure 3: RFA predictions for the contour plots of the (reduced) damping coefficient ασb\alpha\sigma_{b} (top panels) and the (reduced) oscillation frequency ωσb/2πσb/λ\omega\sigma_{b}/2\pi\equiv\sigma_{b}/\lambda (bottom panels) corresponding to the leading pole for a size ratio (a, e) q=0.2q=0.2, (b, f) q=0.3q=0.3, (c, g) q=0.4q=0.4, and (d, h) q=0.648q=0.648. The colormap in the bottom bar indicates the color code for the values of ασb\alpha\sigma_{b} and ωσb/2π\omega\sigma_{b}/2\pi. In each panel, the (black) dashed diagonal line represents the locus ηs+ηb=0.5\eta_{s}+\eta_{b}=0.5, the (black) solid vertical lines represent the values ηb=0.02\eta_{b}=0.02 (for q=0.2q=0.2) or 0.050.05 (for q=0.3q=0.3, 0.40.4, and 0.6480.648), 0.100.10, and 0.200.20 considered in Figs. 69, and the (yellow) circles denote those cases where MD simulations have been performed. The (colored) thick solid lines represent the crossover lines, while the solid (colored) stars and circles indicate the end points and the splitting points, respectively.
Refer to caption
Figure 4: 3D plots, as predicted by the RFA, for (a) the (reduced) damping coefficient ασb\alpha\sigma_{b} and (b) the (reduced) oscillation frequency ωσb/2πσb/λ\omega\sigma_{b}/2\pi\equiv\sigma_{b}/\lambda, corresponding to the leading pole for a size ratio q=0.4q=0.4. The used colormap is the same as the one in Fig. 3; it indicates the change in ασb\alpha\sigma_{b} in the case of panel (a) and of ωσb/2π\omega\sigma_{b}/2\pi in the case of panel (b). In each panel, the (black) solid lines represent the cuts ηb=0.05\eta_{b}=0.05 and 0.200.20 considered in Fig. 8. The (red) thick lines represent the crossover lines and the solid (red) star indicates the end point.
Refer to caption
Figure 5: RFA (thick lines) and PY (thin lines) predictions for the crossover lines CC and CC^{\prime} in the plane ηs\eta_{s} vs ηb\eta_{b} for q=0.2q=0.2, 0.30.3, 0.40.4, and 0.6480.648. The solid circles represent results from MD computer simulations (via the WM scheme), which confirm the theoretical RFA predictions. The WM points for ηb=0.2\eta_{b}=0.2 correspond, from top to bottom, to q=0.648q=0.648, 0.30.3, and 0.40.4, respectively, whereas for ηb=0.05\eta_{b}=0.05 the points correspond to q=0.3q=0.3 and 0.40.4. The (magenta) diamond represents a result from paper I [12]. The solid (colored) stars indicate the end points (see Figs. 14), while the solid (cyan) circles represent the splitting points of G3 generation (see Figs. 1 and 2). The (black) vertical lines represent the values ηb=0.02\eta_{b}=0.02, 0.050.05, 0.100.10, and 0.200.20 considered in Figs. 69.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Dependence of the first six poles on ηs\eta_{s} for q=0.2q=0.2 and ηb=0.02\eta_{b}=0.02 (top panels), ηb=0.10\eta_{b}=0.10 (middle panels), and ηb=0.20\eta_{b}=0.20 (bottom panels). The thick (colored) and thin (gray) lines correspond to the RFA and PY predictions, respectively, while the solid circles represent the WM values for the cases where MD simulations were performed. The colored squares denote poles for a small sphere packing fraction ηs=0.01\eta_{s}=0.01, and the lines indicate trajectories for increasing values of ηs\eta_{s}. The horizontal lines denote the crossovers as ηs\eta_{s} increases. In the central and right panels, the open circles and crosses represent the leading and subleading poles, respectively.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Dependence of the first six poles on ηs\eta_{s} for q=0.3q=0.3 and ηb=0.05\eta_{b}=0.05 (top panels), ηb=0.10\eta_{b}=0.10 (middle panels), and ηb=0.20\eta_{b}=0.20 (bottom panels). The thick (colored) and thin (gray) lines correspond to the RFA and PY predictions, respectively, while the solid circles represent the WM values for the cases where MD simulations were performed. The colored squares denote poles for a small sphere packing fraction ηs=0.01\eta_{s}=0.01, and the lines indicate trajectories for increasing values of ηs\eta_{s}. The horizontal lines denote the crossovers as ηs\eta_{s} increases. In the central and right panels, the open circles and crosses represents the leading and subleading poles, respectively.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: Same as in Fig. 7, except that in this instance q=0.4q=0.4.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Same as in Fig. 7, except that q=0.648q=0.648.
Refer to caption
Figure 10: Plot of the amplitudes (a) AssA_{ss}, (b) AsbA_{sb}, and (c) AbbA_{bb} as functions of ηs\eta_{s} at ηb=0.05\eta_{b}=0.05 for q=0.3q=0.3. The first three poles (in order of increasing ω\omega) are considered. The vertical thick (orange) line at ηs0.16\eta_{s}\simeq 0.16 represents a structural crossover.
Refer to caption
Figure 11: (a) Locus hssmin=0h_{ss}^{\min}=0 [i.e., line of changing sign in the value of the first minimum of hss(r)h_{ss}(r)], as predicted by the RFA, for a size ratio q=0.3q=0.3. The insets show representative behaviors of hss(r)h_{ss}(r) above and below the locus, and the triangles are MD results. (b) RFA predictions for the loci (from top to bottom for each qq) hssmin=0h_{ss}^{\min}=0 (red), hsbmin=0h_{sb}^{\min}=0 (blue), and hbbmin=0h_{bb}^{\min}=0 (green). The size ratios are q=0.2q=0.2 (solid lines), q=0.3q=0.3 (dashed lines), and q=0.4q=0.4 (dotted lines). In each case, hijmin>0h_{ij}^{\min}>0 in the region below the corresponding curve. Some of the curves end at the points marked with crosses. The triangles represent MD results.
Refer to caption
Figure 12: Big-big correlation function hbb(r)h_{bb}(r) for q=0.3q=0.3, ηb=0.05\eta_{b}=0.05, and (a) ηs=0.1\eta_{s}=0.1, (b) ηs=0.3\eta_{s}=0.3. The (red) open circles are MD data, while the (blue) solid, (gray) dotted, and (green) dashed lines represent the contribution of the first (Π1\Pi_{1}), second (Π2\Pi_{2}), and third (Π3\Pi_{3}) pole, respectively. Note that at ηs=0.1\eta_{s}=0.1 the leading pole describing the asymptotic decay of hbb(r)h_{bb}(r) is Π1\Pi_{1} and at ηs=0.3\eta_{s}=0.3 it is Π3\Pi_{3}.

III Pole analysis and structural crossover

In general, the representation of the total correlation functions of additive BHS mixtures may be expressed as [2, 26, 27, 28]

hij(r)={1,0<r<σij,n=1Aij(n)reαnrsin(ωnr+δij(n)),r>σij.h_{ij}({r})=\begin{cases}-1,\qquad 0<r<\sigma_{ij},\\ \sum\limits_{n=1}^{\infty}\frac{A_{ij}^{(n)}}{r}e^{-\alpha_{n}r}\sin\left(\omega_{n}r+\delta_{ij}^{(n)}\right),\qquad r>\sigma_{ij}.\end{cases} (11)

In fact, the functional form of hijWM(r)h_{ij}^{WM}(r) for r>rijminr>r_{ij}^{\text{min}} in Eq. (10) is based on Eq. (11). While the amplitudes, Aij(n)A_{ij}^{(n)}, and the phase shifts, δij(n)\delta_{ij}^{(n)}, are specific for each hij(r)h_{ij}(r), the damping coefficients, αn\alpha_{n}, and the oscillation (angular) frequencies, ωn2π/λn\omega_{n}\equiv 2\pi/\lambda_{n} (λn\lambda_{n} being the associated wavelengths), are common to all the pairs [27]. Although an infinite number of terms is formally considered in Eq. (11), only a few leading terms (those with the smallest damping coefficients) are needed to characterize the asymptotic behavior of hij(r)h_{ij}(r). Note that, while each ωn\omega_{n} is actually a wavenumber, we will use throughout this paper the nomenclature “frequency” with the proviso that it does not refer here to time but to space.

A successful way to study the asymptotic decay behavior of the total correlation functions is based on the pole analysis of their Laplace or Fourier transforms. In Laplace space, the real and imaginary parts of the complex poles zn=αn±ıωnz_{n}=-\alpha_{n}\pm\imath\omega_{n} provide the damping coefficient and the oscillation frequencies, respectively. Similarly, in Fourier space the poles are kn=ızn=±ωn+ıαnk_{n}=-\imath z_{n}=\pm\omega_{n}+\imath\alpha_{n}. In order to avoid later confusion, it is convenient at this stage to clarify the nomenclature we adopt in this work. We will refer to leading, subleading, sub-subleading, etc. poles to an order in increasing αn\alpha_{n}, while the nomenclature first, second, third, etc. poles will refer to an increasing order in ωn\omega_{n}. Apart from that, it must be remarked that the infinite set of poles includes roots with ωn=0\omega_{n}=0 (thus representing contributions decaying monotonically), although they are generally far from the most dominant ones.

Depending on the values of the parameters of the BHS mixture, the position of the poles in the complex plane varies. In the simplest scenario, given a size ratio qq, the plane ηs\eta_{s} vs ηb\eta_{b} can be split into two main regions such that the leading pole in one of the regions has an angular frequency ω2π/σb\omega\approx 2\pi/\sigma_{b}, which corresponds to a wavelength in the oscillatory decay of the total correlation function comparable to the diameter of the big spheres (i.e., λσb\lambda\approx\sigma_{b}), while in the other region the leading pole has ω2π/σs\omega\approx 2\pi/\sigma_{s} (i.e., λσs\lambda\approx\sigma_{s}) [26, 27, 12]. The line separating both regions signals a structural crossover behavior in which the wavelength λ2π/ω\lambda\equiv 2\pi/\omega of the oscillations in the large-rr asymptotic regime changes discontinuously from approximately σb\sigma_{b} to approximately σs\sigma_{s} as the relative amount of the small spheres is increased. This crossover line in the ηs\eta_{s} vs ηb\eta_{b} phase diagram occurs when the corresponding two pairs of poles have the same α\alpha. As will be seen in Sec. IV, this basic scenario for the structural crossover can become much more complicated as the total packing fraction increases, giving rise to the presence of “harmonics” of the “fundamental” oscillation frequency 1/σb1/\sigma_{b}.

Further, we will talk about a first-order, a second-order, a third-order, etc. crossover to the one involving a change in the leading, subleading, sub-subleading, etc. pole, respectively. In particular, for different values of qq, ηs\eta_{s}, and ηb\eta_{b}, the decay of hij(r)h_{ij}(r) is determined by different combinations of poles (i.e., first and second, first and third, first and fifth, and so on).

As said before, in the RFA the poles are obtained from the zeros of D(z)D(z) in Eq. (II.2) and as follows in the case of the WM scheme. Once the total correlation functions hijWM(r)h_{ij}^{WM}(r) [see Eq. (10)] are known after fitting the parameters to the MD data, the direct correlation functions cijWM(r)c_{ij}^{WM}(r) are determined via Fourier transforms and the OZ relation in Eq. (6b), as described in paper I [12]. This in turn allows one to find the poles by the method of Evans et al[26] [see Eqs. (21) of paper I].

IV Results

Evidence of the crossover behavior in BHS mixtures was first pointed out by Grodon et al. [27, 29], who used two different formulations of Rosenfeld’s fundamental measure theory: the original Rosenfeld functional (which is equivalent to the PY approximation) and the White Bear version. Such crossover behavior was later also mentioned in connection with experiments with colloidal suspensions [30, 13].

Let us now turn to the RFA predictions. We begin with the analysis of the leading pole in the ηs\eta_{s} vs ηb\eta_{b} plane. The detailed landscape as one changes the size ratio is rather complex, so here we provide the most general features [31]. By focusing on the behavior of the oscillation frequency ω\omega associated with the leading pole, one can observe that, given a value of qq, the ηs\eta_{s} vs ηb\eta_{b} plane splits into different regions, in each one of which the (reduced) natural frequency ωσb/2πσb/λ\omega\sigma_{b}/2\pi\equiv\sigma_{b}/\lambda is of order of 11 (region R1R_{1}) or is of order of n=2,3,n=2,3,\ldots (region RnR_{n}). The most relevant regions are R1R_{1} (where λσb\lambda\approx\sigma_{b}) and RnsR_{n_{s}} (where ns{n_{s}} is the integer closest to 1/q1/q, so that λσs\lambda\approx\sigma_{s}). Both regions are separated by a crossover line (hereafter labeled as CC), which is present for any qq. Interestingly, as qq decreases (and thus ns{n_{s}} increases), one can observe a second crossover line (CC^{\prime}) separating region RnsR_{n_{s}} from either region Rns+1R_{{n_{s}}+1} or region Rns1R_{{n_{s}}-1}, and even a third line (C′′C^{\prime\prime}) separating region R1R_{1} from region Rns1R_{{n_{s}}-1}.

The previous scenario can be observed in Fig. 1, which shows the evolution of the different regions and crossover lines as one decreases the size ratio from q=0.648q=0.648 to q=0.250q=0.250. At q=0.648q=0.648 it is quite apparent the existence of the conventional crossover line CC separating the regions R1R_{1} (below the line) and Rns=R2R_{n_{s}}=R_{2} (above the line). Interestingly, the line CC terminates at an “end point,” so that one can move continuously between regions R1R_{1} and Rns=R2R_{n_{s}}=R_{2} by circumventing the end point from the left. Decreasing the size ratio from q=0.648q=0.648 to q=0.500q=0.500 (top row of Fig. 1) produces a downward bending of line CC and a left shift of the end point until it eventually disappears at ηb=0\eta_{b}=0. Let us now analyze the middle row of Fig. 1. At q=0.425q=0.425, region Rns+1=R3R_{{n_{s}}+1}=R_{3} starts to compete with region Rns=R2R_{n_{s}}=R_{2}, thus giving rise to the second crossover line CC^{\prime}, which terminates at a new end point. At the transition value q=0.400q=0.400 (where ns{n_{s}} could equally be taken as 22 or 33) the preceding line CC^{\prime} has started a tendency to bend down and the end point has moved to the left. Next, at q=0.375q=0.375, R3R_{3} has changed from being the old region Rns+1R_{{n_{s}}+1} to being the new region RnsR_{n_{s}}, while R2R_{2} has changed from being the old region RnsR_{{n_{s}}} to being the new region Rns1R_{{n_{s}}-1}. Moreover, the old line CC^{\prime} has merged with line CC producing a “splitting point” (where three different pairs of poles share the same damping coefficient α\alpha); now, the crossover line CC extends to the right of the splitting point, while to the left it experiences a pitchfork bifurcation into a line CC^{\prime} (separating region Rns=R3R_{n_{s}}=R_{3} from region Rns1=R2R_{{n_{s}}-1}=R_{2} and still having an end point) and a line C′′C^{\prime\prime} (separating regions R1R_{1} and Rns1=R2R_{{n_{s}}-1}=R_{2}). At q=0.350q=0.350, the splitting point has moved to the left, the end point of line CC^{\prime} has disappeared, and the residual region Rns1R_{{n_{s}}-1} has significantly shrunk. In the bottom row of Fig. 1, we see that, at q=0.315q=0.315, region R2R_{2} has almost disappeared, region Rns+1=R4R_{{n_{s}}+1}=R_{4} starts to compete with region Rns=R3R_{n_{s}}=R_{3}, and a new line CC^{\prime} with an end point appears, in analogy with what happened at q=0.425q=0.425. The rest of the evolution is analogous to what has just been described in relation with the middle row: line CC^{\prime} moves down, it eventually merges with line CC creating a splitting point, a relay from old region Rns+1R_{{n_{s}}+1} to new region RnsR_{n_{s}} and from old region RnsR_{{n_{s}}} to new region Rns1R_{{n_{s}}-1} takes place, and region Rns1R_{{n_{s}}-1} shrinks until eventually disappearing.

A splitting point is absent in the sequence represented by the top row of Fig. 1. In the second row, however, a splitting point (joining regions R1R_{1}, R2R_{2}, and R3R_{3}) is generated and then it disappears, giving rise to the generation of a new splitting point (joining regions R1R_{1}, R3R_{3}, and R4R_{4}) and its later disappearance along the bottom row. Thus, we will refer to the behavior represented by the top, middle, and bottom rows of Fig. 1 as generations G0, G1, and G2, respectively. As qq keeps decreasing beyond q=0.250q=0.250, a new generation G3 appears. The evolution with qq of the end point (generations G0–G3) and splitting points (generations G1–G3) are shown in Fig. 2.

At this stage we should point out two things. On the one hand, it is worth mentioning that the presence of a second crossover line (i.e., the CC^{\prime} line) was already also reported in Ref. [27], at least for q=0.4q=0.4. Nevertheless, the distinction between the CC, CC^{\prime}, and C′′C^{\prime\prime} lines that we have made here is important for understanding and/or predicting the structural crossover behavior, including the appearance of different branches for different qq systems. For instance, without such assets, neither the reason for the sequence of the crossover lines in Fig. 5 of Ref. [27] nor the reasons for the appearance of the second crossover in the case q=0.4q=0.4 or of the behavior observed in the case q=0.65q=0.65 can be explained. In any case, the overall view of the structural crossovers becomes clearer once one realizes that the scenario depicted by Figs. 1 and 2 takes place. Note in particular that, notwithstanding the theoretical interest of the crossover line CC^{\prime}, it must be remarked that, as seen from Fig. 1, it generally lies (except for sufficiently asymmetric mixtures) above the region ηs+ηb=0.5\eta_{s}+\eta_{b}=0.5, where the fluid phase is expected to coexist with a solid phase [32]. On the other hand, it should also be clear that the present scenario is only a coarse-grained description and, while providing a fair picture of what goes on, is not geared towards addressing all the details pertaining to shrinking regions, the merging of crossover lines, the disappearance and appearance of end points, and the formation of splitting points.

To complete the picture advocated in the present paper, let us now present the results of the RFA and the WM scheme for both the damping coefficient and the oscillation frequency associated with the leading pole for q=0.2q=0.2, 0.30.3, 0.40.4, and 0.6480.648. These results are displayed in Fig. 3. We observe that the leading damping coefficient smoothly changes with ηs\eta_{s}, ηb\eta_{b}, and qq, except for the presence of kinks signaling pole crossings. In general, given a value of ηs\eta_{s}, the reduced damping coefficient ασb\alpha\sigma_{b} decreases with increasing ηb\eta_{b}. On the other hand, ασb\alpha\sigma_{b} exhibits a nonmonotonic dependence on ηs\eta_{s} at fixed ηb\eta_{b}. In what respects qq, its influence on ασb\alpha\sigma_{b} is rather weak, although there is a general tendency for a slight decrease of ασb\alpha\sigma_{b} with increasing qq at fixed (ηb,ηs)(\eta_{b},\eta_{s}). In contrast to the behavior of the damping coefficient, the oscillation frequency can experience discontinuous changes in the ηs\eta_{s} vs ηb\eta_{b} diagram at a given qq, as discussed above in connection with Fig. 1. The complexity of the landscape beyond the description of Fig. 1 is exemplified by Fig. 3(e) for q=0.2q=0.2, where, apart from the pitchfork bifurcation at (ηb,ηs)(0.04,0.26)(\eta_{b},\eta_{s})\simeq(0.04,0.26) (giving rise to an encapsulated region R4R_{4}), a second splitting point is born at (ηb,ηs)(0.29,0.38)(\eta_{b},\eta_{s})\simeq(0.29,0.38).

As a representative example, Fig. 4 shows a 3D view visualizing the overall structure behavior, in particular the crossover lines CC and CC^{\prime} and the specific regions R1R_{1}, R2R_{2}, and R3R_{3} for the size ratio q=0.4q=0.4. This helps the understanding of the different features observed in Figs. 1 and 3, as well as in Figs. 69 below. In particular, Fig. 4(b) shows that, as said before, one can move continuously between regions R2R_{2} and R3R_{3} by circumventing the end point.

For the investigated cases q=0.2q=0.2, 0.30.3, 0.40.4, and 0.6480.648, the crossover lines CC and CC^{\prime} are plotted in Fig. 5, where RFA and PY predictions, as well as a few points obtained via the WM scheme, are represented. As can be observed, the shape of lines CC and CC^{\prime} is qualitatively similar for q=0.3q=0.3 and q=0.4q=0.4, while the cases q=0.2q=0.2 and q=0.648q=0.648 present distinctive features. We observe that the black solid circle representing the result from the WM method for q=0.648q=0.648 and ηb=0.20\eta_{b}=0.20 lies on the line obtained from the RFA prediction better than on the PY line. It is worth noting that the differences between RFA and PY grow with increasing density and decreasing size ratio, being especially apparent for q=0.2q=0.2. In that case, for instance, the second splitting point changes from (ηb,ηs)(0.29,0.38)(\eta_{b},\eta_{s})\simeq(0.29,0.38) in the RFA to (ηb,ηs)(0.46,0.27)(\eta_{b},\eta_{s})\simeq(0.46,0.27) in the PY approximation. In any case, it must be remarked that the main separation between the RFA and PY lines for q=0.2q=0.2 takes place in regions of the plane (ηb,ηs)(\eta_{b},\eta_{s}) where the total packing fraction is rather large (η>0.6\eta>0.6) and hence the stable system is expected to consist of coexisting fluid and solid phases [32].

The information presented in Figs. 3 and 5 is complemented by Figs. 69 for q=0.2q=0.2, 0.30.3, 0.40.4, and 0.6480.648, respectively, where the dependence on ηs\eta_{s} of the first six poles is shown for ηb=0.02\eta_{b}=0.02 [for q=0.2q=0.2] or 0.050.05 [for q=0.3q=0.3, 0.40.4, and 0.6480.648] (top panels), 0.100.10 (middle panels), and 0.200.20 (bottom panels). It can be observed that typically the first poles correspond to (reduced) frequencies σb/λ1\sigma_{b}/\lambda\approx 1, σb/λ1/q\sigma_{b}/\lambda\approx 1/q, and the first few harmonics σb/λ=2,3,4,\sigma_{b}/\lambda=2,3,4,\ldots. Note that a certain overlap between the pole with σb/λ1/q\sigma_{b}/\lambda\approx 1/q and that with the nearest harmonic might exist, in what could be viewed as sort of “resonance”. This happens for σb/λ5\sigma_{b}/\lambda\approx 5, σb/λ3\sigma_{b}/\lambda\approx 3, σb/λ2\sigma_{b}/\lambda\approx 2, and σb/λ2\sigma_{b}/\lambda\approx 2 in the cases q=0.2q=0.2 (see right panels of Fig. 6), q=0.3q=0.3 (see right panels of Fig. 7), q=0.4q=0.4 (see right panels of Fig. 8), and q=0.648q=0.648 (see right panels of Fig. 9), respectively.

As seen from Figs. 7(c) and 8(c), only the conventional crossover CC is present at ηb=0.05\eta_{b}=0.05 for q=0.3q=0.3 and 0.40.4, in agreement with what is observed in Figs.  3(f) and 3(g). On the other hand, Figs.  7(f), 7(i), 8(f), and 8(i) show that the crossover CC is followed by the crossover CC^{\prime} as ηs\eta_{s} increases at ηb=0.10\eta_{b}=0.10 and 0.200.20. In the case q=0.2q=0.2, Fig. 6(c) shows two successive crossovers with ηb=0.02\eta_{b}=0.02 when traversing the two branches stemming from the pitchfork bifurcation at the splitting point (ηb,ηs)(0.04,0.26)(\eta_{b},\eta_{s})\simeq(0.04,0.26). Also for q=0.2q=0.2, the crossovers CC and CC^{\prime} are observed in Fig. 6(i) at ηb=0.20\eta_{b}=0.20, while at η=0.10\eta=0.10 the second crossover CC^{\prime} is beyond the range of ηs\eta_{s} shown [see Figs. 5 and 6(f)]. On the other hand, according to Figs. 9(c), 9(f), and 9(i), a single crossover exists at ηb=0.20\eta_{b}=0.20 and q=0.648q=0.648 since ηb=0.05\eta_{b}=0.05 and ηb=0.10\eta_{b}=0.10 are located to the right of the end point [see Fig. 3(h)].

Figures 69 also show that some quantitative differences between the RFA predictions and those of the PY approximation are present, as already mentioned in connection with Fig. 5. We can observe as well a good agreement of the results of the RFA method with the poles obtained from the WM scheme for those cases where MD simulations were carried out. This shows (via WM) that the RFA can effectively be used to predict the structural properties of BHS mixtures over a wide range of the phase diagram.

It should be noted that, apart from the damping coefficients and the oscillation frequencies, the amplitudes Aij(n)A_{ij}^{(n)} [see Eq. (11)] can be obtained from both the RFA and the WM scheme. As an illustration, Fig. 10 shows the ηs\eta_{s}-dependence of the amplitudes corresponding to the first three poles in the case ηb=0.05\eta_{b}=0.05 and q=0.3q=0.3. For this rather disparate mixture, it can be observed that Abb10Asb100AssA_{bb}\sim 10A_{sb}\sim 100A_{ss}. The general agreement between the RFA and WM results is fair, except for the amplitude AbbA_{bb} associated with the second pole (ωσb/2πσb/λ2\omega\sigma_{b}/2\pi\equiv\sigma_{b}/\lambda\approx 2) when ηs\eta_{s} increases. Note, however, that this second pole is never the leading one. If one focuses on the leading and subleading poles (σb/λ1\sigma_{b}/\lambda\approx 1 or σb/λ3\sigma_{b}/\lambda\approx 3), the agreement is very good.

Obviously, this complex behavior concerning the asymptotic decay of hij(r)h_{ij}(r) and the associated structural crossovers emerge as a consequence of the competition between the three basic length scales of the problem, namely σs\sigma_{s}, σb\sigma_{b}, and σsb\sigma_{sb}. Another manifestation of this competition appears when dealing with the sign of the first (local) minimum of hij(r)h_{ij}(r), here denoted as hijminh_{ij}^{\min}, typically located at rijmin2σijr_{ij}^{\min}\approx 2\sigma_{ij}. The conventional expectation is hijmin<0h_{ij}^{\min}<0, thus signaling the beginning of the oscillations around hij(r)=0h_{ij}(r)=0. In fact, this is what happens in monocomponent fluids. However, given ηb\eta_{b} and qq, it turns out that hijmin>0h_{ij}^{\min}>0 if ηs\eta_{s} is smaller than a certain transition value. The loci hijmin=0h_{ij}^{\min}=0 separating the conventional behavior hijmin<0h_{ij}^{\min}<0 (above the locus) from the peculiar property hijmin>0h_{ij}^{\min}>0 (below the locus) are shown in Fig. 11 for q=0.2q=0.2, 0.30.3, and 0.40.4. Given a value of qq, the locus hssmin=0h_{ss}^{\min}=0 envelops the locus hsbmin=0h_{sb}^{\min}=0, and the latter envelops the locus hbbmin=0h_{bb}^{\min}=0. Additionally, the region with hijmin>0h_{ij}^{\min}>0 shrinks as qq increases. Moreover, the curves hsbmin=0h_{sb}^{\min}=0 and hbbmin=0h_{bb}^{\min}=0 for the cases q=0.3q=0.3 and 0.40.4, as well as the curve hssmin=0h_{ss}^{\min}=0 for q=0.4q=0.4, lose their meaning to the right of the crosses. This is because at a larger value of ηb\eta_{b}, hijmin=hij(r2σij)h_{ij}^{\min}=h_{ij}(r\approx 2\sigma_{ij}) changes from being a negative local minimum to being negative, but not an extremum, as ηs\eta_{s} decreases.

The structural crossover phenomenon is illustrated in Fig. 12, where the decay of hbb(r)h_{bb}(r) at ηb=0.05\eta_{b}=0.05 and (a) ηs=0.1\eta_{s}=0.1 and (b) ηs=0.3\eta_{s}=0.3 for q=0.3q=0.3 is shown. In agreement with the top rightmost panel of Fig. 7, the leading pole changes from being the first one (wavelength λσb\lambda\approx\sigma_{b}) at ηs=0.1\eta_{s}=0.1 to being the third one (wavelength λσs\lambda\approx\sigma_{s}) at ηs=0.3\eta_{s}=0.3, the transition taking place at ηs0.18\eta_{s}\simeq 0.18. Additionally, Fig. 12(a) shows that, at least for intermediate distances (say 1<r/σb<21<r/\sigma_{b}<2), the three leading poles are needed to capture the (large-wavelength) oscillations of hbb(r)h_{bb}(r) at ηs=0.1\eta_{s}=0.1. This situation becomes even more relevant as one approaches the transition value ηs0.18\eta_{s}\simeq 0.18 since the damping coefficients associated with the three first poles almost coincide near ηs0.18\eta_{s}\simeq 0.18 [see Fig. 7(b)]. However, at ηs=0.3\eta_{s}=0.3, Fig. 12(b) shows that the leading pole is enough to account for the (small-wavelength) oscillations, even for intermediate distances.

V Concluding remarks

The results we have presented in this paper deserve more consideration. First of all, it must be emphasized that the good agreement found between the results of the RFA method and those of the WM scheme in paper I [12] for a single value of the total packing fraction (η=0.5\eta=0.5) and size ratio (q=0.648q=0.648) has been hereby confirmed. Therefore, we now have a powerful theoretical (almost completely analytic) tool to examine the complex behavior of the structural properties of BHS mixtures, including their asymptotic decay. In particular, we have found that, in the case of the leading pole of the total correlation functions, given a value of ηs\eta_{s}, the (reduced) damping coefficient ασb\alpha\sigma_{b} generally decreases with increasing ηb\eta_{b}, while it exhibits a nonmonotonic dependence on ηs\eta_{s} at fixed ηb\eta_{b}. Also, the influence of qq on ασb\alpha\sigma_{b} appears to be rather weak, although there is a general tendency for a slight decrease with increasing qq at fixed (ηb,ηs)(\eta_{b},\eta_{s}).

On the other hand, in agreement with the work of Grodon et al. [27, 29] (which considered a different approximation and a particular value of qq), we have confirmed that there exists a crossover line (CC) in the plane ηs\eta_{s} vs ηb\eta_{b} separating a region (R1R_{1}) where the (reduced) natural oscillation frequency ωσb/2πσb/λ\omega\sigma_{b}/2\pi\equiv\sigma_{b}/\lambda is of the order of 11 from another region (RnsR_{n_{s}}, with ns1/q{n_{s}}\approx 1/q) where σb/λ\sigma_{b}/\lambda is of the order of 1/q1/q. The former extends to smaller values of ηs\eta_{s}, while the latter extends to larger values of ηs\eta_{s}. Further, we have also found that, if qq is small enough, there is a second crossover line (CC^{\prime}) separating RnsR_{n_{s}} from a third region (Rns+1R_{{n_{s}}+1}) where σb/λ\sigma_{b}/\lambda is of the order of the first harmonic of σb/λ=1\sigma_{b}/\lambda=1 that turns out to be larger than ns{n_{s}}. This line CC^{\prime} may terminate at an end point located at a large value of ηs\eta_{s} and a small value of ηb\eta_{b}, which implies that one can move continuously between regions RnsR_{n_{s}} and Rns+1R_{{n_{s}}+1} by circumventing the end point from the left. However, except for small qq, this line tends to lie above the region ηs+ηb=0.5\eta_{s}+\eta_{b}=0.5, where the fluid phase is expected to coexist with the solid one. Finally, we have also shown that the above scenario can even be more complex, with additional crossover lines (CC^{\prime} separating RnsR_{n_{s}} from Rns1R_{{n_{s}}-1} and C′′C^{\prime\prime} separating R1R_{1} from Rns1R_{{n_{s}}-1}) and splitting points, as the total packing fraction η=ηs+ηb\eta=\eta_{s}+\eta_{b} increases or the size ratio qq decreases. One important issue, which remains to be assessed, is to understand the behavior of the crossover lines in the limiting region ηb0\eta_{b}\to 0. It should be remarked that this region is hardly accessible by simulations but certainly needs to be studied in more depth.

To close this paper, two other outcomes of our work are worth pointing out. The first one concerns the value of the first local minimum in the total correlation function hij(r)h_{ij}(r), which changes sign as one crosses a certain transition line which depends on qq and the pair under consideration. This point is dealt with by Fig. 11. The second is that the direct correlation function csb(r)c_{sb}(r) is not monotonic in the region 0<r<σsb0<r<\sigma_{sb} and presents a well-defined minimum. To our knowledge, this feature has not been pointed out before. Due to the fact that we are persuaded that this finding is relevant, we will address this point and analyze it in detail in the following (third) paper of this series.

Acknowledgements.
S.B.Y. and A.S. acknowledge financial support from Grant No. PID2020-112936GB-I00/AEI/10.13039/501100011033 and from the Junta de Extremadura (Spain) through Grants No. IB20079 and No. GR18079, all of them partially financed by Fondo Europeo de Desarrollo Regional funds. S.P. is grateful to the Universidad de Extremadura, where most of this work was carried out during his scientific internship, which was supported by Grant No. DEC 2018/02/X/ST3/03122 financed by the National Science Center, Poland. Some of the calculations were performed at the Poznań Supercomputing and Networking Center (PCSS).

References

  • Barker and Henderson [1976] J. A. Barker and D. Henderson, What is “liquid”? Understanding the states of matter, Rev. Mod. Phys. 48, 587 (1976).
  • Hansen and McDonald [2013] J.-P. Hansen and I. R. McDonald, Theory of Simple Liquids, 4th ed. (Academic Press, London, 2013).
  • Santos [2016] A. Santos, A Concise Course on the Theory of Classical Liquids. Basics and Selected Topics, Lecture Notes in Physics, vol. 923 (Springer, New York, 2016).
  • Fisher and Widom [1969] M. E. Fisher and B. Widom, Decay of correlations in linear systems, J. Chem. Phys. 50, 3756 (1969).
  • Evans et al. [1993] R. Evans, J. R. Henderson, D. C. Hoyle, A. O. Parry, and Z. A. Sabeur, Asymptotic decay of liquid structure: oscillatory liquid-vapour density profiles and the Fisher–Widom line, Mol. Phys. 80, 755 (1993).
  • Vega et al. [1995] C. Vega, L. F. Rull, and S. Lago, Location of the Fisher-Widom line for systems interacting through short-ranged potentials, Phys. Rev. E 51, 3146 (1995).
  • Brown [1996] W. E. Brown, The Fisher–Widom line for a hard core attractive Yukawa fluid, Mol. Phys. 88, 579 (1996).
  • Dijkstra and Evans [2000] M. Dijkstra and R. Evans, A simulation study of the decay of the pair correlation function in simple fluids, J. Chem. Phys. 112, 1449 (2000).
  • Tarazona et al. [2003] P. Tarazona, E. Chacón, and E. Velasco, The Fisher–Widom line for systems with low melting temperature, Mol. Phys. 101, 1595 (2003).
  • Škrbić et al. [2016] T. Škrbić, A. Badasyan, T. X. Hoang, R. Podgornik, and A. Giacometti, From polymers to proteins: the effect of side chains and broken symmetry on the formation of secondary structures within a Wang–Landau approach, Soft Matter 12, 4783 (2016).
  • López de Haro et al. [2018] M. López de Haro, A. Rodríguez-Rivas, S. B. Yuste, and A. Santos, Structural properties of the Jagla fluid, Phys. Rev. E 98, 012138 (2018).
  • Pieprzyk et al. [2020] S. Pieprzyk, A. C. Brańka, S. B. Yuste, A. Santos, and M. López de Haro, Structural properties of additive binary hard-sphere mixtures, Phys. Rev. E 101, 012117 (2020).
  • Statt et al. [2016] A. Statt, R. Pinchaipat, F. Turci, R. Evans, and C. P. Royall, Direct observation in 3d of structural crossover in binary hard sphere mixtures, J. Chem. Phys. 144, 144506 (2016).
  • Yuste et al. [1998] S. B. Yuste, A. Santos, and M. López de Haro, Structure of multicomponent hard-sphere mixtures, J. Chem. Phys. 108, 3683 (1998).
  • López de Haro et al. [2008] M. López de Haro, S. B. Yuste, and A. Santos, Alternative Approaches to the Equilibrium Properties of Hard-Sphere Liquids, in Theory and Simulation of Hard-Sphere Fluids and Related Systems, Lecture Notes in Physics, vol. 753, edited by A. Mulero (Springer, Berlin, 2008) pp. 183–245.
  • Santos et al. [2020] A. Santos, S. B. Yuste, and M. López de Haro, Structural and thermodynamic propertiesof hard-sphere fluids, J. Chem. Phys. 153, 120901 (2020).
  • Lebowitz [1964] J. L. Lebowitz, Exact solution of generalized Percus–Yevick equation for a mixture of hard spheres, Phys. Rev. 133, A895 (1964).
  • Boublík [1970] T. Boublík, Hard-sphere equation of state, J. Chem. Phys. 53, 471 (1970).
  • Grundke and Henderson [1972] E. W. Grundke and D. Henderson, Distribution functions of multi-component fluid mixtures of hard spheres, Mol. Phys. 24, 269 (1972).
  • Lee and Levesque [1973] L. L. Lee and D. Levesque, Perturbation theory for mixtures of simple liquids, Mol. Phys. 26, 1351 (1973).
  • not [a] A code using the Mathematica computer algebra system to obtain Gij(z)G_{ij}(z) and gij(r)g_{ij}(r) with the RFA method is available from the web page http://www.unex.es/eweb/fisteor/santos/filesRFA.html.
  • Bannerman et al. [2011] M. N. Bannerman, R. Sargant, and L. Lue, Dynamo: a free 𝒪(n)\mathcal{O}(n) general event-driven molecular dynamics simulator, J. Comput. Chem. 32, 3329 (2011).
  • Allen and Tildesley [2017] M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids (Oxford University Press, Oxford, 2017).
  • Salacuse et al. [1996] J. J. Salacuse, A. R. Denton, and P. A. Egelstaff, Finite-size effects in molecular dynamics simulations: Static structure factor and compressibility. I. Theoretical method, Phys. Rev. E 53, 2382 (1996).
  • Baumketner and Hiwatari [2001] A. Baumketner and Y. Hiwatari, Finite-size dependence of the bridge function extracted from molecular dynamics simulations, Phys. Rev. E 63, 061201 (2001).
  • Evans et al. [1994] R. Evans, R. J. F. Leote de Carvalho, J. R. Henderson, and D. C. Hoyle, Asymptotic decay of correlations in liquids and their mixtures, J. Chem. Phys. 100, 591 (1994).
  • Grodon et al. [2004] C. Grodon, M. Dijkstra, R. Evans, and R. Roth, Decay of correlation functions in hard-sphere mixtures: Structural crossover, J. Chem. Phys. 121, 7869 (2004).
  • Pieprzyk et al. [2017] S. Pieprzyk, A. C. Brańka, and D. M. Heyes, Representation of the direct correlation function of the hard-sphere fluid, Phys. Rev. E 95, 062104 (2017).
  • Grodon et al. [2005] C. Grodon, M. Dijkstra, R. Evans, and R. Roth, Homogeneous and inhomogeneous hard-sphere mixtures: manifestations of structural crossover, Mol. Phys. 103, 3009 (2005).
  • Baumgartl et al. [2007] J. Baumgartl, R. P. A. Dullens, M. Dijkstra, R. Roth, and C. Bechinger, Experimental observation of structural crossover in binary mixtures of colloidal hard spheres, Phys. Rev. Lett. 98, 198303 (2007).
  • not [b] See file movie.mp4 in the Supplemental Material at http://link.aps.org/supplemental/10.1103/PhysRevE.104.024128 for a video with the evolution of ωσb/2π\omega\sigma_{b}/2\pi in the plane ηs\eta_{s} vs ηb\eta_{b} from q=0.65q=0.65 to q=0.2q=0.2.
  • Dijkstra et al. [1999] M. Dijkstra, R. van Roij, and R. Evans, Phase diagram of highly asymmetric binary hard-sphere mixtures, Phys. Rev. E 59, 5744 (1999).