Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description

The scaled particle theory is developed for the description of thermodynamical properties of a mixture of hard spheres and hard spherocylinders. Analytical expressions for free energy, pressure and chemical potentials are derived. From the minimization of free energy, a nonlinear integral equation...

Ausführliche Beschreibung

Gespeichert in:
Bibliographische Detailangaben
Datum:2017
Hauptverfasser: Holovko, M.F., Hvozd, M.V.
Format: Artikel
Sprache:English
Veröffentlicht: Інститут фізики конденсованих систем НАН України 2017
Schriftenreihe:Condensed Matter Physics
Online Zugang:https://nasplib.isofts.kiev.ua/handle/123456789/157023
Tags: Tag hinzufügen
Keine Tags, Fügen Sie den ersten Tag hinzu!
Назва журналу:Digital Library of Periodicals of National Academy of Sciences of Ukraine
Zitieren:Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description / M.F. Holovko, M.V. Hvozd // Condensed Matter Physics. — 2017. — Т. 20, № 4. — С. 43501: 1–11. — Бібліогр.: 46 назв. — англ.

Institution

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id nasplib_isofts_kiev_ua-123456789-157023
record_format dspace
spelling nasplib_isofts_kiev_ua-123456789-1570232025-02-09T14:30:23Z Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description Iзотропно-нематичний перехiд в сумiшi твердих сфер та твердих сфероцилiндрiв: застосування теорiї масштабної частинки Holovko, M.F. Hvozd, M.V. The scaled particle theory is developed for the description of thermodynamical properties of a mixture of hard spheres and hard spherocylinders. Analytical expressions for free energy, pressure and chemical potentials are derived. From the minimization of free energy, a nonlinear integral equation for the orientational singlet distribution function is formulated. An isotropic-nematic phase transition in this mixture is investigated from the bifurcation analysis of this equation. It is shown that with an increase of concentration of hard spheres, the total packing fraction of a mixture on phase boundaries slightly increases. The obtained results are compared with computer simulations data. Для опису термодинамiчних властивостей сумiшi твердих сфер та твердих сфероцилiндрiв розвинуто теорiю масштабної частинки. Отримано аналiтичнi вирази для вiльної енергiї, тиску та хiмiчних потенцiалiв. Мiнiмiзацiєю вiльної енергiї сформульовано нелiнiйне iнтегральне рiвняння для орiєнтацiйної унарної функцiї розподiлу. З бiфуркацiйного аналiзу цього рiвняння дослiджено iзотропно-нематичний фазовий перехiд в данiй сумiшi. Показано, що при збiльшеннi концентрацiї твердих сфер загальний коефiцiєнт упаковки сумiшi на границях фаз злегка зростає. Представлено порiвняння отриманих результатiв з даними комп’ютерного моделювання. This project has received funding from the European Unions Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 734276, and from the State Fund For Fundamental Research (project N F73/26-2017). 2017 Article Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description / M.F. Holovko, M.V. Hvozd // Condensed Matter Physics. — 2017. — Т. 20, № 4. — С. 43501: 1–11. — Бібліогр.: 46 назв. — англ. 1607-324X PACS: 51.30.+i, 64.70.Md DOI:10.5488/CMP.20.43501 arXiv:1712.05330 https://nasplib.isofts.kiev.ua/handle/123456789/157023 en Condensed Matter Physics application/pdf Інститут фізики конденсованих систем НАН України
institution Digital Library of Periodicals of National Academy of Sciences of Ukraine
collection DSpace DC
language English
description The scaled particle theory is developed for the description of thermodynamical properties of a mixture of hard spheres and hard spherocylinders. Analytical expressions for free energy, pressure and chemical potentials are derived. From the minimization of free energy, a nonlinear integral equation for the orientational singlet distribution function is formulated. An isotropic-nematic phase transition in this mixture is investigated from the bifurcation analysis of this equation. It is shown that with an increase of concentration of hard spheres, the total packing fraction of a mixture on phase boundaries slightly increases. The obtained results are compared with computer simulations data.
format Article
author Holovko, M.F.
Hvozd, M.V.
spellingShingle Holovko, M.F.
Hvozd, M.V.
Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
Condensed Matter Physics
author_facet Holovko, M.F.
Hvozd, M.V.
author_sort Holovko, M.F.
title Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
title_short Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
title_full Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
title_fullStr Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
title_full_unstemmed Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
title_sort isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description
publisher Інститут фізики конденсованих систем НАН України
publishDate 2017
url https://nasplib.isofts.kiev.ua/handle/123456789/157023
citation_txt Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description / M.F. Holovko, M.V. Hvozd // Condensed Matter Physics. — 2017. — Т. 20, № 4. — С. 43501: 1–11. — Бібліогр.: 46 назв. — англ.
series Condensed Matter Physics
work_keys_str_mv AT holovkomf isotropicnematictransitioninamixtureofhardspheresandhardspherocylindersscaledparticletheorydescription
AT hvozdmv isotropicnematictransitioninamixtureofhardspheresandhardspherocylindersscaledparticletheorydescription
AT holovkomf izotropnonematičnijperehidvsumišitverdihsfertatverdihsferocilindrivzastosuvannâteoriímasštabnoíčastinki
AT hvozdmv izotropnonematičnijperehidvsumišitverdihsfertatverdihsferocilindrivzastosuvannâteoriímasštabnoíčastinki
first_indexed 2025-11-26T20:28:50Z
last_indexed 2025-11-26T20:28:50Z
_version_ 1849886176131940352
fulltext Condensed Matter Physics, 2017, Vol. 20, No 4, 43501: 1–11 DOI: 10.5488/CMP.20.43501 http://www.icmp.lviv.ua/journal Isotropic-nematic transition in a mixture of hard spheres and hard spherocylinders: scaled particle theory description M.F. Holovko, M.V. Hvozd Institute for Condensed Matter Physics of the National Academy of Sciences of Ukraine, 1 Svientsitskii St., 79011 Lviv, Ukraine Received July 27, 2017 The scaled particle theory is developed for the description of thermodynamical properties of a mixture of hard spheres and hard spherocylinders. Analytical expressions for free energy, pressure and chemical potentials are derived. From the minimization of free energy, a nonlinear integral equation for the orientational singlet distribution function is formulated. An isotropic-nematic phase transition in this mixture is investigated from the bifurcation analysis of this equation. It is shown that with an increase of concentration of hard spheres, the total packing fraction of a mixture on phase boundaries slightly increases. The obtained results are compared with computer simulations data. Key words: hard sphere/hard spherocylinder mixture, isotropic-nematic transition, scaled particle theory PACS: 51.30.+i, 64.70.Md 1. Introduction A hard spherocylinder fluid is one of the simplest models widely used for the description of an isotropic-nematic phase transition in liquid crystals [1, 2]. The first treatment of isotropic-nematic transition in a hard-spherocylinder fluid was performed by Onsager about seventy years ago [3]. The Onsager theory can be considered as the density functional theory in which a low-density expansion of the free energy functional is truncated at the level of second virial coefficient. The equilibrium state is determined by the functional variation of free energy with respect to the orientational distribution function. It was shown [3] that such a treatment is exact in the limit of infinitely thin rodswhen L →∞ and D→ 0, but L2D is fixed, where L and D are the length and the diameter of spherocylinders, respectively. It was shown that besides an isotropic-nematic transition, the Onsager theory describes a nematic-smectic transition in a hard-spherocylinder fluid, which appears at higher densities [4]. The application of the scaled particle theory (SPT) [5–9] provides an efficient approximate way to incorporate the higher-order contributions neglected in the Onsager theory. An alternative way of improving the Onsager theory is the Parsons-Lee (PL) approach [10–12], which is based on the mapping of the properties of a spherocylinder fluid to those of the hard-sphere model. The SPT theory was also extended for the description of a hard-spherocylinder fluid in random porous media [13, 14]. During the last decades the approaches developed for a hard-spherocylinder fluid have been gener- alized for the description of mixtures of hard anisotropic particles. In such systems, new phases were observed and their properties were richer and more complicated than those for the one-component case [4, 15–25]. The simplest example of such multi-component systems of hard anisotropic particles is a binary mixture of hard spheres and hard spherocylinders, for the description of which the corresponding approaches have been proposed. Among them there are the Onsager theory [12, 16, 18, 26] and the Parsons-Lee approach [22, 24, 25] in the one-fluid and many-fluid approximations for a hard-spheres mixture and computer simulations [17, 20, 22, 24, 25]. This work is licensed under a Creative Commons Attribution 4.0 International License . Further distribution of this work must maintain attribution to the author(s) and the published article’s title, journal citation, and DOI. 43501-1 https://doi.org/10.5488/CMP.20.43501 http://www.icmp.lviv.ua/journal http://creativecommons.org/licenses/by/4.0/ M.F. Holovko, M.V. Hvozd In this paper we present a development of the scaled particle theory for the description of a binary mixture of hard spheres and hard spherocylinders. We derive expressions for the chemical potentials of hard sphere and hard spherocylinder components. From the minimization of free energy, a non-linear integral equation for the orientational distribution function is obtained. From the bifurcation analysis of this integral equation, an isotropic-nematic phase diagram of a binary mixture of hard spheres and hard spherocylinders is analysed and discussed. The results of the presented approach are numerically compared with some computer simulation data. The paper is arranged as follows. The theoretical part is presented in section 2. The discussion of the obtained results and the comparison to computer simulations data are given in section 3. Finally, we draw some conclusions in the last section. 2. Theory We consider a two-component hard convex body (HCB) fluid consisting of hard spheres (HS) and hard spherocylinders (HSC). In order to characterize HCB particles we use three functionals: the volume V , the surface area S and the mean curvature r taken with the factor 1/4π. For hard spheres of the radius R1, these functionals are as follows: V1 = 4 3 πR3 1 , S1 = 4πR2 1 , r1 = R1 (2.1) and for the hard spherocylinders of the radius R2 and of the length L2 V2 = πR2 2 L2 + 4 3 πR3 2 , S2 = 2πR2L2 + 4πR2 2 , r2 = 1 4 L2 + R2. (2.2) A basic idea of the SPT approach is an insertion of an additional particle of a variable size, e.g., a scaled particle, into a fluid. Adding a scaled hard-sphere particle into our system we use the scaling parameter λs. Therefore, the volume V1s, the surface area S1s and the mean curvature r1s are modified according to the following relations V1s = λ 3 sV1 , S1s = λ2 s S1 , r1s = λsr1. (2.3) When we add a scaled hard-spherocylinder particle into a fluid with the scaling radius R2s and the scaling length L2s, in addition to the scaling parameter λs, we introduce the scaling parameter αs. Therefore, the radius R2s and the length L2s are defined as [5, 6] R2s = λsR2 , L2s = αsL2. (2.4) As a result, the volume V2s, the surface area S2s and the mean curvature r2s of the scaled spherocylinder are equal to V2s = πR2 2 L2αsλ 2 s + 4 3 πR3 2λ 3 s , S2s = 2πR2L2αsλs + 4πR2 2λ 2 s , r2s = 1 4 L2αs + R2λs. (2.5) The procedure of insertion of a scaled particle into a HCB fluid is equivalent to the creation of a corresponding cavity. This cavity is free of centers of any other fluid particles. The key point of the SPT theory is a consideration of the excess chemical potential of a scaled particle µexs [27–33]. The work needed to create such a cavity is equal to µexs . For a small scaledHS andHSC particles, the excess chemical potentials can bewritten in the following form [29] βµex1s(λs) = − ln [ 1 − η1(1 + λs)3 − η2 ( 1 + r1sS2 V2 + r2S1s V2 + V1s V2 )] , (2.6) βµex2s(αs, λs) = − ln [ 1 − η1 ( 1 + r2sS1 V1 + r1S2s V1 + V2s V1 ) − η2 ( 1 + r2sS2 V2 + r2S2s V2 + V2s V2 )] , (2.7) 43501-2 Isotropic-nematic transition in a HS/HSC mixture where β = 1/kBT , kB is the Boltzmann constant, T is the temperature, η1 = ρ1V1 is the HS fluid packing fraction, ρ1 is the HS fluid density, V1 is the volume of a HS particle; η2 = ρ2V2 is the HSC fluid packing fraction, ρ2 is the HSC fluid density, V2 is the volume of a HSC particle. It should be noted that equation (2.7) is written for an isotropic case. After substituting equations (2.1)–(2.5) into equations (2.6)–(2.7) and having generalized equa- tion (2.7) for the anisotropic case, the chemical potentials of the HS and HSC scaled particles can be presented as follows: βµex1s(λs) = − ln { 1 − η1(1 + λs)3 − η2 [ 1 + 1 k1 6γ2 3γ2 − 1 λs + 1 k2 1 3(γ2 + 1) 3γ2 − 1 λ2 s + 1 k3 1 2 3γ2 − 1 λ3 s ]} , (2.8) βµex2s(αs, λs) = − ln ( 1 − η1 [ 3 4 s1αs (1 + k1λs) 2 + (1 + k1λs) 3 ] − η2 { 1 + 3(γ2 − 1) 3γ2 − 1 [1 + (γ2 − 1)τ( f )]αs + 6γ2 3γ2 − 1 λs + 6(γ2 − 1) 3γ2 − 1 [ 1 + 1 2 (γ2 − 1)τ( f ) ] αsλs + 3(γ2 + 1) 3γ2 − 1 λ2 s + 3(γ2 − 1) 3γ2 − 1 αsλ 2 s + 2 3γ2 − 1 λ3 s }) , (2.9) where k1, s1 and γ2 are equal to k1 = R2 R1 , s1 = L2 R1 , γ2 = 1 + L2 2R2 , (2.10) and τ( f ) = 4 π ∫ f (Ω1) f (Ω2) sin γ(Ω1,Ω2)dΩ1dΩ2. (2.11) Here, Ω = (ϑ, ϕ) denotes the orientation of HSC particles and it is defined by the angles ϑ and ϕ, where dΩ = 1 4π sin ϑdϑdϕ is the normalized angle element, γ(Ω1,Ω2) is the angle between orientational vectors of two molecules, f (Ω) is the singlet orientation distribution function normalized in such a way that∫ f (Ω)dΩ = 1. (2.12) f (Ω) is defined herein below from the minimization of the free energy of a considered mixture. For a large scaled particle, the excess chemical potential is equal to the work needed to create a macroscopic cavity within the fluid and it is given by a thermodynamic expression. For a scaled hard sphere particle, it can be presented as follows: βµex1s = w(λs) + βPV1s , (2.13) where P is the pressure of a fluid and V1s is the volume of a scaled HS particle. Similarly, for a scaled HSC particle, we have βµex2s = w(αs, λs) + βPV2s , (2.14) where P and V2s are the pressure of a fluid and the volume of a scaled HSC particle, respectively. According to the ansatz of SPT [14, 27–33] w(λs) and w(αs, λs) can be written in the form of expansions: w(λs) = w0 + w1λs + 1 2 w2λ 2 s , (2.15) w(αs, λs) = w00 + w01αs + w10λs + w11αsλs + 1 2 w20λ 2 s . (2.16) We can derive the coefficients of these expansions from the continuity of µex1s, µ ex 2s and their corresponding derivatives ∂µex1s/∂λs, ∂ 2µex1s/∂λ 2 s at λs = 0 for a scaled HS particle; ∂µex2s/∂αs, ∂µ ex 2s/∂λs, ∂ 2µex2s/∂αs∂λs, ∂2µex2s/∂λ 2 s at αs = λs = 0 for a scaled HSC particle. 43501-3 M.F. Holovko, M.V. Hvozd Therefore, for a scaled HS particle, we obtain: w0 = − ln(1 − η), w1 = 3 η1 1 − η + 1 k1 6γ2 3γ2 − 1 η2 1 − η , w2 = 6 η1 1 − η + 1 k2 1 6(γ2 + 1) 3γ2 − 1 η2 1 − η + 1 (1 − η)2 ( 3η1 + 1 k1 6γ2 3γ2 − 1 η2 )2 , (2.17) where η = η1 + η2 is the total packing fraction of the mixture considered. For a scaled HSC particle, we find: w00 = − ln(1 − η), w01 = 3 4 s1 η1 1 − η + [ 3(γ2 − 1) 3γ2 − 1 + 3(γ2 − 1)2τ( f ) 3γ2 − 1 ] η2 1 − η , w10 = 3k1 η1 1 − η + 6γ2 3γ1 − 1 η2 1 − η , w11 = 3 2 k1s1 η1 1 − η + [ 6(γ2 − 1) 3γ2 − 1 + 3(γ2 − 1)2τ( f ) 3γ2 − 1 ] η2 1 − η + { 3 4 s1 η1 1 − η + [ 3(γ2 − 1) 3γ2 − 1 + 3(γ2 − 1)2τ( f ) 3γ2 − 1 ] η2 1 − η } ( 3k1 η1 1 − η + 6γ2 3γ2 − 1 η2 1 − η ) , w20 = 6k2 1 η1 1 − η + 6(γ2 + 1) 3γ1 − 1 η2 1 − η + 1 (1 − η)2 ( 3k1η1 + 6γ2 3γ2 − 1 η2 )2 . (2.18) After setting λs = 1 in equation (2.15) and αs = λs = 1 in equation (2.16), the HS and HSC scaled particles become of the same sizes as HS and HSC particles of a fluid, respectively. It makes it possible to find the relation between the pressure and the excess chemical potentials µex1 and µex2 of a fluid. The total chemical potentials for HS and HSC particles in a HS/HSC mixture are as follows: βµ1 = ln(ρ1Λ 3 1) + βµ ex 1 , (2.19) βµ2 = ln(ρ2Λ 3 2Λ2R) + βµ ex 2 , (2.20) where Λ1 and Λ2 are the fluid thermal wavelengths of the HS and HSC components, respectively; Λ−1 2R is the rotational partition function of a single HSC molecule [34]. Then, we can write expressions for the total chemical potentials as follows: βµ1 = ln(ρ1Λ 3 1) − ln(1 − η) + a1 η 1 − η + b1 η2 (1 − η)2 + βP η1 ρ1 , (2.21) βµ2 = ln(ρ2Λ 3 2Λ2R) − ln(1 − η) + a2 η 1 − η + b2 η2 (1 − η)2 + βP η2 ρ2 , (2.22) where the coefficients a1, a2, b1 and b2 are: a1 = 6 η1 η + [ 1 k1 6γ2 3γ2 − 1 + 1 k2 1 3(γ2 + 1) 3γ2 − 1 ] η2 η , b1 = 1 2 ( 3 η1 η + 1 k1 6γ2 3γ2 − 1 η2 η )2 (2.23) and a2 ( τ( f ) ) = [ 3 4 s1(1 + 2k1) + 3k1(1 + k1) ] η1 η + [ 6 + 6(γ2 − 1)2τ( f ) 3γ2 − 1 ] η2 η , 43501-4 Isotropic-nematic transition in a HS/HSC mixture b2 ( τ( f ) ) = {( 3 4 s1 + 3 2 k1 ) η1 η + [ 3(2γ2 − 1) 3γ2 − 1 + 3(γ2 − 1)2τ( f ) 3γ2 − 1 ] η2 η } ( 3k1 η1 η + 6γ2 3γ1 − 1 η2 η ) . (2.24) Therefore, we have two equations (2.21) and (2.22), and each of them contains two unknown quantities: one of the chemical potentials and the pressure. In the case of an one-component fluid, we can eliminate one of these unknowns, βµ1 (βµ2) or P, using equation (2.21) or equation (2.22) and the Gibbs-Duhem relation. In our case, the Gibbs-Duhem equation has the form ∂(βP) ∂ρ = 2∑ α=1 ρα ∂(βµα) ∂ρ . (2.25) Expressions for the chemical potentials can be obtained according to the recent paper [35], in which an expression for the pressure is obtained from equation (2.25), and an expression for the free energy is obtained from an integration over the total density ρ. From the differentiation of the free energy with respect to ρ1 and ρ2, we derive expressions for the chemical potentials µ1 and µ2. In order to use equation (2.25) and to get one equation containing only one unknown instead of equations (2.21) and (2.22), we take the derivatives with respect to the total fluid density ρ = ∑2 α=1 ρα on the both sides of equations (2.21) and (2.22) keeping the fluid composition constant: xα = ρα/ρ, α = 1, 2. Hence, we derive ∂(βµ1) ∂ρ = 1 ρ [ 1 + η 1 − η + a1 η (1 − η)2 + 2b1 η2 (1 − η)3 ] + 4 3 πR3 1 ∂(βP) ∂ρ , (2.26) ∂(βµ2) ∂ρ = 1 ρ [ 1 + η 1 − η + a2 η (1 − η)2 + 2b2 η2 (1 − η)3 ] + ( πR2 2 L2 + 4 3 πR3 2 ) ∂(βP) ∂ρ . (2.27) The combination of equations (2.25) and (2.26)–(2.27) makes it possible to write an expression for the fluid compressibility. Taking into account that ∑ α xα = 1, we obtain ∂(βP) ∂ρ = 1 1 − η + (1 + A) η (1 − η)2 + (A + 2B) η2 (1 − η)3 + 2B η3 (1 − η)4 , (2.28) where A = 2∑ α=1 xαaα , B = 2∑ α=1 xαbα . (2.29) From the integration of equation (2.28) over the total density ρ at a constant concentration, we find βP ρ = 1 + η 1 − η + A 2 η (1 − η)2 + 2B 3 η2 (1 − η)3 . (2.30) Now, we calculate the Helmholtz free energy, which is related to the pressure as βF V = ρ ρ∫ 0 dρ′ 1 ρ′ ( βP ρ′ ) . (2.31) We carry out this integration at fixed concentrations xα, where α = 1, 2. Thus, the final expression for the free energy is βF V = βFid V + ρ [ − ln(1 − η) + A 2 η 1 − η + B 3 η2 (1 − η)2 ] , (2.32) where Fid is the ideal gas contribution to the Helmholtz free energy of a mixture: βFid V = ρ1 [ ln(Λ3 1ρ1) − 1 ] + ρ2 [ ln(Λ3 2ρ2) − 1 ] + ρ2σ( f ). (2.33) 43501-5 M.F. Holovko, M.V. Hvozd Here, σ( f ) is the entropic term defined as σ( f ) = ∫ f (Ω) ln f (Ω)dΩ. (2.34) The singlet orientational distribution function f (Ω) can be obtained from the minimization of free energy with respect to variations of this distribution. This procedure leads to a nonlinear integral equation ln f (Ω1) + λ + C ∫ f (Ω′) sin γ(Ω1Ω ′)dΩ′ = 0, (2.35) where C = η2 1 − η [ 3(γ2 − 1)2 3γ2 − 1 + 1 1 − η (γ2 − 1)2 3γ2 − 1 ( 3k1η1 + 6γ2 3γ2 − 1 η2 )] . (2.36) The constant λ is defined from the normalization condition equation (2.12). Using the expression for the Helmholtz free energy, we calculate the total chemical potentials for the components of HS and HSC in a mixture. From the relationship βµα = ∂ ∂ρα ( βF V ) , (2.37) we derive βµ1 = lnΛ3 1ρ1 − ln(1 − η) + 1 2 η 1 − η { a1 + 6 ρ1V1 η + ρ2V1 η [ 3 4 s1(1 + 2k1) + 3k1(1 + k1) ] } + 1 3 η2 (1 − η)2 [ b1 + 3 ρ1V1 η2 ( 3η1 + 1 k1 6γ2 3γ2 − 1 η2 ) + ρ2V1 η2 ( 9k1 ( 1 2 s1 + k1 ) η1 + { 3 4 6γ2 3γ2 − 1 s1 + 3k1 [ 3 + 3(γ2 − 1)2τ( f ) 3γ2 − 1 ]} η2 )] + βPV1 (2.38) for the chemical potential of HS and βµ2 = lnΛ3 2ρ2 + σ( f ) − ln(1 − η) + 1 2 η 1 − η { a2 + ρ1V2 η [ 1 k1 6γ2 3γ2 − 1 + 1 2 1 k2 1 6(γ2 + 1) 3γ2 − 1 ] + ρ2V2 η [ 6 + 6(γ2 − 1)2τ( f ) 3γ2 − 1 ] } + 1 3 η2 (1 − η)2 [ b2 + ρ1V2 η2 1 k1 6γ2 3γ2 − 1 ( 3η1 + 1 k1 6γ2 3γ2 − 1 η2 ) + ρ2V2 η2 ( { 3 4 6γ2 3γ2 − 1 s1 + 3k1 [ 3 + 3(γ2 − 1)2τ( f ) 3γ2 − 1 ]} η1 + 6γ2 3γ2 − 1 [ 6(2γ2 − 1) 3γ2 − 1 + 6(γ2 − 1)2τ( f ) 3γ2 − 1 ] η2 )] + βPV2 (2.39) for the chemical potential of HSC. 3. Results and discussions We use the theory presented in the previous section to study the effect of hard spheres on the isotropic- nematic phase transition in a binary mixture of hard spheres and hard spherocylinders. This investigation is done within the framework of bifurcation analysis of the integral equation equation (2.35) for the singlet distribution function f (Ω). It is worth noting that for the first time this equation was obtained by Onsager [3] for a hard-spherocylinder fluid in the limit L2 → ∞ and R2 → 0, when the dimensionless density of a spherocylinder fluid c2 = 1 2πρ2L2 2 R2 was fixed. Therefore, in the Onsager limit we have C → c2 = 1 2 πρ2L2 2 R2. (3.1) 43501-6 Isotropic-nematic transition in a HS/HSC mixture The result, equation (2.36), for C is the generalization of the SPT result for a HSC fluid for the finite values of L2 and R2 [9, 36]. In this case, C → η2 1 − η2 [ 3 (γ2 − 1)2 3γ2 − 1 + η2 1 − η2 6γ2 (γ2 − 1)2 (3γ2 − 1)2 ] . (3.2) From the bifurcation analysis of the integral equation equation (2.35) for the singlet distribution function f (Ω), it was found that this equation has two characteristic points Ci and Cn [37], which defined the range of stability of a considered mixture. The first point Ci corresponds to the highest possible density of the stable isotropic state and the second point Cn corresponds the lowest possible density of a stable nematic state. For the Onsager model, from the minimization of the free energy with respect to the singlet distribution function f (Ω), and subsequently from the solution of the coexistence equations, the following values of density of coexisting isotropic and nematic phases were obtained [38–40]: ci = 3.289, cn = 4.192. (3.3) In the presence of hard spheres for the Onsager model, we have C = c2 1 − η1 . (3.4) It means that the isotropic-nematic transition in the presence of hard spheres shifts to lower densities of spherocylinders. For the binary mixture of hard spheres and hard spherocylinders at the finite value of L2/2R2, we can put Ci = 3.289, Cn = 4.192, (3.5) where Ci and Cn are determined from equation (2.36). The values of Ci and Cn in equation (3.5) define the isotropic-nematic phase diagram for a HS/HSC mixture depending on the ratios L2/R2 = 2 (γ2 − 1) and k1 = R2/R1, as well as on the densities of HS and HSC particles, η1 and η2, respectively. We note that s1 defined by equation (2.10) is not an independent parameter, since s1 = 2 (γ2 − 1) k1. (3.6) The packing fraction of hard spheres η1 as a function of the packing fraction of hard spherocylinders η2 for γ2 = 21 along the boundaries of isotropic-nematic phase transition is estimated from equation (3.5) for a HS/HSC mixture at a fixed ratio k1. As it is seen from figure 1, the presence of hard spheres shifts the phase transition to lower densities of hard spherocylinders. Moreover, the interfacial region becomes broader if the size of hard spheres increases (k1 decreases). The effect of the size of hard spheres on the isotropic-nematic phase boundaries of the same HS/HCS mixture, but at a fixed packing fraction η1, is demonstrated in figure 2. One can observe that an increase of the packing fraction of hard spheres leads to a contraction of interfacial region. The boundaries of an isotropic-nematic phase transition for the same model are presented in figure 3 using the coordinates η = η1 + η2 and x1 = ρ1/(ρ1 + ρ2). It is seen that an increase of the composition of spherical particles x1 makes the total packing fraction η slightly higher in comparison with the results obtained in the coordinates η1 and η2 (see figure 1). In the case of a pure HSC fluid (x1 = 0), the SPT results are rather close to the computer simulations data taken from [41]. It is noticed that the SPT theory underestimates the values of ηi and ηn, and the difference from the computer simulations increases with a decrease of the parameter L2/2R2. The isotropic-nematic phase transition boundaries for a HS/HSC mixture when γ2 = 6 and k1 = 1 are presented in figure 4 (the left-hand panel) using the coordinates η and x1. A comparison with the computer simulations data taken from [17] and [25] are also shown in this figure. It is found that the theoretical prediction of the isotropic line is about ∆η = 0.05 lower than the simulations results, while for the nematic line it is approximately ∆η = 0.023 lower. On the other hand, qualitatively the effect of the composition x1 predicted by the theory and the one obtained from the computer simulations are similar. Hence, in figure 4 (the right-hand panel) we present the modified results of a theoretical prediction, 43501-7 M.F. Holovko, M.V. Hvozd 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Cn κ1=0.1 Ciη 2 η1 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Cn κ1=0.4 Ciη 2 η1 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Cn κ1=0.7 Ci η 2 η1 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Cn κ1=1 Ci η 2 η1 Figure 1. (Color online) Coexistence lines of isotropic-nematic phases of a HS/HSCmixture for L2/2R2 = 20 presented as a dependence of the packing fraction of HSC particles η2 = ρ2V2 on the packing fraction of HS particles η1 = ρ1V1 at fixed k1 = R2/R1. The black line below denoted by Ci corresponds to the isotropic phase, the red line above denoted by Cn corresponds to the nematic phase. The area between the solid black and dotted red lines corresponds to the region of the coexistence of isotropic and nematic phases. 0.105 0.11 0.115 0.12 0.125 0.13 0.135 0.14 0.145 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Cn η1=0.1 Ci η 2 κ1 0.065 0.07 0.075 0.08 0.085 0.09 0.095 0.1 0.105 0.11 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Cn η1=0.3 Ci η 2 κ1 0.035 0.04 0.045 0.05 0.055 0.06 0.065 0.07 0.075 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Cn η1=0.5 Ci η 2 κ1 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Cn η1=0.7 Ci η 2 κ1 Figure 2. (Color online) Coexistence lines of isotropic-nematic phases of a HS/HSCmixture for L2/2R2 = 20 presented as a dependence of the packing fraction of HSC particles η2 = ρ2V2 on the ratio k1 = R2/R1 at the fixed packing fraction of HS particles η1 = ρ1V1. The black line below denoted by Ci corresponds to the isotropic phase, the red line above denoted by Cn corresponds to the nematic phase. The area between the solid black and dotted red lines corresponds to the region of the coexistence of isotropic and nematic phases. 43501-8 Isotropic-nematic transition in a HS/HSC mixture 0.13 0.135 0.14 0.145 0.15 0.155 0.16 0.165 0.17 0.175 0 0.05 0.1 0.15 0.2 0.25 η x 1 Figure 3. (Color online) Coexistence lines of isotropic-nematic phases of a HS/HSCmixture for L2/2R2 = 20 and k1 = 1 presented as a dependence of the packing fraction of HS/HSC mixture η = η1 + η2 on the composition of spherical particles x1 = ρ1/(ρ1 + ρ2). For the case of pure HSC system (x1 = 0) the SPT results are close to those obtained in computer simulations with a use of the modified Gibbs-Duhem integration procedure (filled black triangle for the isotropic branch and open red triangle for the nematic branch) [41]. 0.35 0.36 0.37 0.38 0.39 0.4 0.41 0.42 0.43 0.44 0 0.05 0.1 0.15 0.2 0.25 η x 1 0.35 0.36 0.37 0.38 0.39 0.4 0.41 0.42 0.43 0.44 0 0.05 0.1 0.15 0.2 0.25 η x 1 Figure 4. (Color online) Coexistence lines of the isotropic and nematic phases of a HS/HSC mixture for L2/2R2 = 5 and k1 = 1 presented as a dependence of the packing fraction of HS/HSCmixture η = η1+η2 on the composition of spherical particles x1 = ρ1/(ρ1 + ρ2). The computer simulations data taken from [17] are denoted by filled circles for the isotropic branch and by open circles for the nematic branch. The computer simulations data taken from [25] are shown as filled squares for the isotropic branch and open squares for the nematic branch. where the coexistence lines are shifted up by ∆η = 0.05 for the isotropic phase and by ∆η = 0.023 for the nematic phase. As one can see, in this case an agreement between theoretical and simulation results is rather satisfactory. It is worth noting that the isotropic-nematic coexistence lines can be also obtained from the condition of thermodynamic equilibrium. According to this condition, the both phases should have the same pressure and the same chemical potentials: Pi(ηi, xi) = Pn(ηn, xn), µi,1(ηi, xi) = µn,1(ηn, xn), µi,2(ηi, xi) = µn,2(ηn, xn), (3.7) where µi,1 (or µi,2) and µn,1 (or µn,2) are the chemical potentials of HS (or HSC) particles in the isotropic and nematic phases, respectively. In [37] it was shown that for the Onsager model the results obtained from the bifurcation analysis and from the thermodynamic calculations coincide exactly. We also observe the same for the mixture of Onsager spherocylinders and hard spheres. On the other hand, it was found in [14] that for the finite values of L2/2R2, there is some difference between the results obtained from these two different approaches, and the difference slightly increases with a decrease of the ratio L2/2R2. 43501-9 M.F. Holovko, M.V. Hvozd 4. Conclusions In this paper we have generalized the scaled particle theory for the investigation of thermodynamic properties of a mixture of hard spheres and hard spherocylinders. The expressions for the chemical potentials of hard spheres and hard spherocylinders are derived from the consideration of a scaled hard sphere and a scaled hard spherocylinder inserted into a system under study. Analytical expressions for the free energy and for the pressure of the considered mixture are also obtained. From the minimization of the free energy, a nonlinear integral equation for the orientational distribution function is obtained. From the bifurcation analysis of this integral equation, an isotropic-nematic phase transition in a mixture of hard spheres and hard spherocylinders is investigated. It is shown that the presence of hard spheres shifts the phase transition to the lower densities of hard spherocylinders. With an increase of the sizes of hard spheres, the interfacial region is expanded and with an increase of the packing fraction of hard spheres, the interfacial region decreases. It is also shown that with an increase of concentration of hard spheres, the total packing fraction of a mixture on the phase boundaries slightly increases in comparison with phase boundaries in the coordinates of the packing fraction of hard spheres η1 and the packing fraction of hard spherocylinders. The obtained results are qualitatively in agreement with computer simulations data. The present work can be extended directly to the presence of disordered porous media. For the present time, the scaled particle theory for a hard sphere fluid in disordered porous media is quite well developed [27–29, 33, 42, 43] and has found applications in describing a reference system within the perturbation theory for fluids with different types of attraction, such as associative [44] and ionic fluids [45, 46]. A generalization of SPT theory for the description of a hard spherocylinder fluid in disordered porous media is presented in [13, 14]. Using the obtained results in our separate paper we are going to consider the generalization of SPT theory for the mixture of hard spheres and hard spherocylinders in random porous media. Acknowledgements This project has received funding from the European Unions Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 734276, and from the State Fund For Fundamental Research (project N F73/26-2017). References 1. Vroege G.J., Lekkerkerker H.N.W., Rep. Prog. Phys., 1992, 55, 1241, doi:10.1088/0034-4885/55/8/003. 2. Franco-Melgar M., Haslam A.J., Jackson G., Mol. Phys., 2008, 106, 649, doi:10.1080/00268970801926958. 3. Onsager L., Ann. N.Y. Acad. Sci., 1949, 51, 627, doi:10.1111/j.1749-6632.1949.tb27296.x. 4. Koda T., Numajiri M., Ikeda S., J. Phys. Soc. Jpn., 1996, 65, 3551, doi:10.1143/JPSJ.65.3551 5. Cotter M.A., Martire D.E., J. Chem. Phys., 1970, 52, 1909, doi:10.1063/1.1673232. 6. Cotter M.A., Phys. Rev. A, 1974, 10, 625, doi:10.1103/PhysRevA.10.625. 7. Lasher G., J. Chem. Phys., 1970, 53, 4141, doi:10.1063/1.1673914. 8. Cotter M.A., In: The Molecular Physics of Liquid Crystals, Luckhurst G.R., Gray G.W. (Eds.), Academic Press, London, 1979, 169–189. 9. Cotter M.A., Wacker D.C., Phys. Rev. A, 1978, 18, 2669, doi:10.1103/PhysRevA.18.2669. 10. Parsons J.D., Phys. Rev. A, 1979, 19, 1225, doi:10.1103/PhysRevA.19.1225. 11. Lee S.D., J. Chem. Phys., 1987, 87, 4972, doi:10.1063/1.452811. 12. Lee S.D., J. Chem. Phys., 1988, 89, 7036, doi:10.1063/1.455332. 13. Holovko M., Shmotolokha V., Patsahan T., J. Mol. Liq., 2014, 189, 30, doi:10.1016/j.molliq.2013.05.030. 14. Holovko M., Shmotolokha V., Patsahan T., In: Physics of Liquid Matter: Modern Problems, Bulavin L., Lebovka N. (Eds.), Springer, Heidelberg, 2015, 3–30. 15. Galindo A., Haslam A.J., Varga S., Jackson G., Vanakaras A.G., Photinos D.J., Dunmur D.A., J. Chem. Phys., 2003, 119, 5216, doi:10.1063/1.1598432. 16. Cinacchi G., Mederos L., Velasco E., J. Chem. Phys., 2004, 121, 3854, doi:10.1063/1.1774153. 17. Lago S., Cuetos A., Martínez-Haya B., Rull L.F., J. Mol. Recognit., 2004, 17, 417, doi:10.1002/jmr.704. 18. Vesely F.J., Mol. Phys., 2005, 103, 679, doi:10.1080/00268970512331328686. 43501-10 https://doi.org/10.1088/0034-4885/55/8/003 https://doi.org/10.1080/00268970801926958 https://doi.org/10.1111/j.1749-6632.1949.tb27296.x https://doi.org/10.1143/JPSJ.65.3551 https://doi.org/10.1063/1.1673232 https://doi.org/10.1103/PhysRevA.10.625 https://doi.org/10.1063/1.1673914 https://doi.org/10.1103/PhysRevA.18.2669 https://doi.org/10.1103/PhysRevA.19.1225 https://doi.org/10.1063/1.452811 https://doi.org/10.1063/1.455332 https://doi.org/10.1016/j.molliq.2013.05.030 https://doi.org/10.1063/1.1598432 https://doi.org/10.1063/1.1774153 https://doi.org/10.1002/jmr.704 https://doi.org/10.1080/00268970512331328686 Isotropic-nematic transition in a HS/HSC mixture 19. Martinez-Ratón Y., Cinacchi G., Velasco E., Mederos L., Eur. Phys. J. E, 2006, 21, 175, doi:10.1140/epje/i2006-10058-4. 20. Cuetos A., Martínez-Haya B., Lago S., Rull L.F., Phys. Rev. E, 2007, 75, 061701, doi:10.1103/PhysRevE.75.061701. 21. Cuetos A., Galindo A., Jackson G., Phys. Rev. Lett., 2008, 101, 237802, doi:10.1103/PhysRevLett.101.237802. 22. Malijevský A., Jackson G., Varga S., J. Chem. Phys., 2008, 129, 144504, doi:10.1063/1.2982501. 23. Belli S., Dijkstra M., van Roij R., J. Phys.: Condens. Matter, 2012, 24, 284128, doi:10.1088/0953-8984/24/28/284128. 24. Gámez F., Acemel R.D., Cuetos A., Mol. Phys., 2013, 111, 3136, doi:10.1080/00268976.2013.771802. 25. Wu L., Malijevský A., Jackson G., Müller E.A., Avendaño C., J. Chem. Phys., 2015, 143, 044906, doi:10.1063/1.4923291. 26. Ågren G., Phys. Rev. A, 1975, 11, 1040, doi:10.1103/PhysRevA.11.1040. 27. Holovko M., Dong W., J. Phys. Chem. B, 2009, 113, 6360, doi:10.1021/jp809706n. 28. Chen W., Dong W., Holovko M., Chen X.S., J. Phys. Chem. B, 2010, 114, 1225, doi:10.1021/jp9106603. 29. Patsahan T., Holovko M., Dong W., J. Chem. Phys., 2011, 134, 074503, doi:10.1063/1.3532546. 30. Reiss H., Frisch H.L., Lebowitz J.L., J. Chem. Phys., 1959, 31, 369, doi:10.1063/1.1730361. 31. Reiss H., Frisch H.L., Helfand E., Lebowitz J.L., J. Chem. Phys., 1960, 32, 119, doi:10.1063/1.1700883. 32. Lebowitz J.L., Helfand E., Praestgaard E., J. Chem. Phys., 1965, 43, 774, doi:10.1063/1.1696842. 33. Holovko M., Shmotolokha V., Dong W., Condens. Matter Phys., 2010, 13, 23607, doi:10.5488/CMP.13.23607. 34. Gray C.G., Gubbins K.E., Theory of Molecular Fluids, Claredon Press, Oxford, 1984. 35. Chen W., Zhao S.L., Holovko M., Chen X.S., Dong W., J. Phys. Chem. B, 2016, 120, 5491, doi:10.1021/acs.jpcb.6b02957. 36. Tuinier R., Taniguchi T., Wensink H.H., Eur. Phys. J. E, 2007, 23, 355, doi:10.1140/epje/i2007-10197-0. 37. Kayser R.F. (Jr.), Raveché H.J., Phys. Rev. A, 1978, 17, 2067, doi:10.1103/PhysRevA.17.2067. 38. Herzfeld J., Berger A.E., Wingate J.W., Macromolecules, 1984, 17, 1718, doi:10.1021/ma00139a014. 39. Lekkerkerker H.N.W., Coulon Ph., Van Der Haegen R., Deblieck R., J. Chem. Phys., 1984, 80, 3427, doi:10.1063/1.447098. 40. Chen Z.Y., Macromolecules, 1993, 26, 3419, doi:10.1021/ma00065a027. 41. Bolhuis P., Frenkel D., J. Chem. Phys., 1997, 106, 666, doi:10.1063/1.473404. 42. Holovko M., Patsahan T., Dong W., Condens. Matter. Phys., 2012, 15, 23607, doi:10.5488/CMP.15.23607. 43. Holovko M., Patsahan T., Dong W., Pure Appl. Chem., 2012, 85, 115, doi:10.1351/PAC-CON-12-05-06. 44. Kalyuzhnyi Y.V., Holovko M., Patsahan T., Cummings P.T., J. Phys. Chem. Lett., 2014, 5, 4260, doi:10.1021/jz502135f. 45. Holovko M.F., Patsahan O., Patsahan T., J. Phys.: Condens. Matter, 2016, 28, 414003, doi:10.1088/0953-8984/28/41/414003. 46. Holovko M.F., Patsahan T.M., Patsahan O.V., J. Mol. Liq., 2017, 235, 53, doi:10.1016/j.molliq.2016.11.030. Iзотропно-нематичний перехiд в сумiшi твердих сфер та твердих сфероцилiндрiв: застосування теорiї масштабної частинки М.Ф. Головко,М.В. Гвоздь Iнститут фiзики конденсованих систем НАН України, вул. Свєнцiцького, 1, 79011 Львiв, Україна Для опису термодинамiчних властивостей сумiшi твердих сфер та твердих сфероцилiндрiв розвинуто тео- рiю масштабної частинки. Отримано аналiтичнi вирази для вiльної енергiї, тиску та хiмiчних потенцiалiв. Мiнiмiзацiєю вiльної енергiї сформульовано нелiнiйне iнтегральне рiвняння для орiєнтацiйної унарної функцiї розподiлу. З бiфуркацiйного аналiзу цього рiвняння дослiджено iзотропно-нематичний фазовий перехiд в данiй сумiшi. Показано, що при збiльшеннi концентрацiї твердих сфер загальний коефiцiєнт упаковки сумiшi на границях фаз злегка зростає. Представлено порiвняння отриманих результатiв з да- ними комп’ютерного моделювання. Ключовi слова: сумiш твердих сфер та твердих сфероцилiндрiв, iзотропно-нематичний перехiд, метод масштабної частинки 43501-11 https://doi.org/10.1140/epje/i2006-10058-4 https://doi.org/10.1103/PhysRevE.75.061701 https://doi.org/10.1103/PhysRevLett.101.237802 https://doi.org/10.1063/1.2982501 https://doi.org/10.1088/0953-8984/24/28/284128 https://doi.org/10.1080/00268976.2013.771802 https://doi.org/10.1063/1.4923291 https://doi.org/10.1103/PhysRevA.11.1040 https://doi.org/10.1021/jp809706n https://doi.org/10.1021/jp9106603 https://doi.org/10.1063/1.3532546 https://doi.org/10.1063/1.1730361 https://doi.org/10.1063/1.1700883 https://doi.org/10.1063/1.1696842 https://doi.org/10.5488/CMP.13.23607 https://doi.org/10.1021/acs.jpcb.6b02957 https://doi.org/10.1140/epje/i2007-10197-0 https://doi.org/10.1103/PhysRevA.17.2067 https://doi.org/10.1021/ma00139a014 https://doi.org/10.1063/1.447098 https://doi.org/10.1021/ma00065a027 https://doi.org/10.1063/1.473404 https://doi.org/10.5488/CMP.15.23607 https://doi.org/10.1351/PAC-CON-12-05-06 https://doi.org/10.1021/jz502135f https://doi.org/10.1088/0953-8984/28/41/414003 https://doi.org/10.1016/j.molliq.2016.11.030 Introduction Theory Results and discussions Conclusions