Effect of the liquid crystal solute on the rotator phase transitions of n-alkanes

Prabir K. Mukherjee*
Department of Physics, Government College of Engineering and Textile Technology, 12 William Carey Road, Serampore, Hooghly-712201, India. E-mail: pkmuk1966@gmail.com

Received 8th November 2014 , Accepted 8th January 2015

First published on 9th January 2015


Abstract

Recent experimental studies have shown that the liquid crystal substance plays an important role in determining the structures and the phase transitions of the different rotator phases in binary mixtures of n-alkane and liquid crystals. The mixtures exhibit crystal and four different rotator phases RI, RII, RIII and RV, although the RIII phase is absent in pure alkane. We described these experimental observations within phenomenological theory. The influence of the liquid crystal solute on the rotator phase transitions and the transition temperatures was discussed by varying the coupling between the concentration variable and the order parameters. The theoretical predictions were found to be in good qualitative agreement with the experimental results.


I. Introduction

Normal alkanes, CH3–(CH2)(n−2)CH3, are of physical interest because of their specific rotator phases and their intricate phase transition behavior. The pure n-alkanes have some peculiar features in phase equilibria, such as the odd–even effect, surface freezing effect1 and nucleation kinetics.2,3 Despite their simple molecular structures, long chain n-alkanes are the building blocks of many molecules, including liquid crystals, surfactants, lipids and polymers. Normal alkanes are also the main ingredient in many petroleum products, such as fuels and lubricants, as well as many pharmaceutical and petroderivative products. Special interest has been paid to the physical properties of the rotator phases in normal alkanes, with their specific anisotropic intermolecular and intramolecular interactions, which are of fundamental interest for the understanding of the phenomenological properties of other related macromolecular materials. Over the past few years, the rotator phases of n-alkanes have received considerable attention because of their unique and unusual properties. These include surface crystallization, anomalous heat capacity, negative thermal compressibilities, unusually high thermal expansions, and a vanishing nucleation barrier for the n-alkane rotator to crystal transformation.4 They also induce crystallization even beyond their range of thermal stability.5 It was observed that the rotator to crystal interfacial energy decreases as the entropy decreases. However, when the interfacial tension is extrapolated to zero, it disappears at a finite value of the entropy. This extrapolated disappearance of interfacial energy arises from the loss of the distinction between the rotator and X phases when the rotator phase is forced to a very low temperature.4 Another very interesting feature of the rotator phases is that they are very similar to Langmuir monolayers6,7 in terms of the lattice distortion and the area/molecule. Thus, the rotator phases are interesting in their own right and we also expect that insights into their properties will help us to understand other condensed phases that exist in nature.

Rotator phases consist of layered structures with three-dimensional crystalline order in the positions of the molecular centers of mass, but no long-range orientational order in the rotation of the molecules around their long axes.8–10 In alkanes (CnH2n+2), rotator phases have been observed for carbon numbers 9 ≤ n ≤ 40 interposed, in temperature, between the liquid and fully crystalline phases. Normal alkanes form a variety of different crystalline structures with triclinic, orthorhombic or hexagonal symmetry. The crystal structure of the even n-alkanes (triclinic) is different from that of the odd n-alkanes (orthorhombic). According to the literature,11–15 in the crystalline form, odd n-alkanes exhibit a layered structure consisting of bilayer stacking of the lamellae (ABAB…). The low temperature crystalline phase (X) has orthorhombic (distorted-hexagonal) or triclinic packing within a layer, as well as long-range herringbone order of the rotational degrees of freedom of the backbones. The appearance of the long-range herringbone order in the low temperature X phase was determined by the presence of a Bragg reflection at the {2,1,0} position6,11 (Fig. 3(c)). This peak is absent in the rotator phases and present in the X phase. In principle, herringbone order would appear as a diffusion peak at the {2,1,0} position, whose inverse width is related to the correlation length.6 The herringbone peak occurs at a position of q = 2.098 Å−1 for the lattice parameters reported by Bohanon.16 Although such a diffuse peak is observed in liquid crystals17,18 and in the crystalline phase of Langmuir monolayers,7 no such peak has been identified in alkanes due to the low intensities. In the case of liquid crystals, the centres of mass of the molecules are arranged at the nodes of a periodic rectangular lattice, the orientation of the molecules around their long axes being ensured by a glide-mirror placed in such a way that the molecular sections parallel to the smectic layers are arranged in a herringbone array. In the rigid crystal, most of the molecules contain all-trans configurations. However, when the temperature is increased a finite number of gauche bonds are known to occur near the chain ends. In the rotator phases, non-planar conformers begin to play an important role.19 The main type of such conformers is that containing the end-gauche modification, or gauche-trans-gauche sequence. The number of gauche defects is known to increase with increasing temperature in the rotator phases of longer chain lengths. Thus the gauche defects are known to strongly influence the transitions between rotator phases.8,9

Five different rotator phases have been identified using X-ray techniques.8,9,11,12,15,20–23 These include the rotator-I (RI) phase, rotator-II (RII) phase, rotator-III (RIII) phase, rotator-IV (RIV) phase and rotator-V (RV) phase. These five rotator phases are distinguished by distortion, tilt and azimuthal order. In the rotator-I (RI) phase the molecules are untilted with respect to the layers and there is a rectangularly distorted hexagonal lattice. The layers are stacked in an ABAB… bilayer stacking sequence (Fig. 1(a) and (b)). The rotator-II (RII) phase is usually described as being composed of molecules that are untilted with respect to the layers that are packed in a hexagonal lattice. The layers are stacked in an ABCABC… trilayer stacking sequence (Fig. 1(c) and (d)). The RIII phase is triclinic, with molecules tilted in a non-symmetric azimuthal direction, in between the next nearest-neighbor (NNN) and the nearest-neighbor (NN) directions (Fig. 1(e) and (f)). The RIV phase is monoclinic, with NNN-tilted molecules and end-to-end layer stacking (Fig. 1(g) and (h)). The distortion is small and in the same direction as the tilt. The finite distortion in this case is often considered to be induced by a finite coupling to the tilt. The RV phase is the same as RI phase, except that the molecules are tilted towards their next nearest neighbor (NNN) (Fig. 1(i) and (j)). The RV phase contains a finite molecular tilt toward the NNN, the same direction as the direction of distortion. The transitions between the different rotator phases have been studied extensively experimentally by several authors,8,9,24 using calorimetry and X-ray diffraction. Rotator phases pass through at least one rotator phase between crystalline solid and isotropic liquid. The transition between different phases corresponds to the breaking of some symmetry. Usually, the system does not re-enter the original state, unlike the liquid crystals. There is no pronounced even–odd effect within the rotator phases, unlike the liquid crystals.8 According to the calorimetry and X-ray diffraction measurements,8,9 transitions from the RV phase to the X phase (RV–X) or from the RI phase to the X phase (RI–X) are first order. The RI–X transition is known to involve the onset of long-range herringbone order.11 The RI–X and RV–X transitions exhibit significant hysteresis due to the supercooling of the RI and RV phases. The RII–RI transition is weakly first order in nature, because only the distortion order parameter is lost at TII–I and the heat of transition is only 0.1–1 kJ mol−1.9 Here, TII–I is the RII–RI transition temperature. The jump of the distortion order parameter at TII–I is 0.003–0.03.9 The isotropic liquid to RII (IL–RII) transition is the strongest first order transition, with a supercooling value of less than 0.03 °C.


image file: c4ra14116d-f1.tif
Fig. 1 Molecular arrangement of various rotator phases of n-alkanes. The unfilled circles represent the chain end positions in the second layer of the bilayer structure. The gray circles for the RII phase represent the third layer in a trilayer structure. The arrow represents the tilt (θ) direction in the NNN tilted RV phase.

Recently, investigations on binary mixtures of n-alkanes and nanoparticles, including porous matrices, have attracted increasing attention.25–28 Experimental results25–28 show that these substances can greatly enhance the physical properties of n-alkanes and rotator phases.25–28 Very recently, Kumar et al.29 carried out calorimetric and X-ray investigations on a binary system of n-tetracosane (C24H50) and the liquid crystalline compound butyloxybenzilideneoctylaniine (BBOA). According to the literature,8,11,15 pure C24H50 exhibits the following four phases: an orthorhombic crystalline phase (X), a RV phase, a RI phase and a RII phase. BBOA exhibits two liquid crystalline phases, namely the nematic (N) and smectic A (SmA) phases above the plastic phase referred to as the crystal B (CrB) phase. Owing to the chemical dissimilarity between the alkane and the liquid crystal molecule, the latter could act like an impurity and therefore try to destabilize the system. Kumar et al.29 observed that if the alkane (C24H50) is present in a small concentration, the binary system (C24H50–BBOA) is nanophase segregated, whereas if the liquid crystal molecules are present in a small concentration, the layered structure merely gets roughened without any segregation. The more significant result of their experiments, at low liquid crystal concentration, is the induction of the RIII phase between the RI and RV phases. The phase diagram of the C24H50–BBOA system exhibits the following phase sequence: isotropic liquid (IL)–RII–RI–RIII–RV–X. This is shown in Fig. 2. They pointed out that the appearance of the induced rotator phase RIII between the RI and RV phases could be due to the creation of additional gauche bonds at the termini of the alkane molecules. The addition of the liquid crystal component up to the alkane C24H50 leads to a decrease in the various transition temperatures and lattice distortion.


image file: c4ra14116d-f2.tif
Fig. 2 Partial temperature–concentration phase diagram of C24H48–BBOA, showing the interesting feature that the RIII phase gets induced for a small concentration of BBOA and is bounded for a higher concentration. This results in two three-phase points in the vicinity of the concentration marked A and B.29

The properties of rotator phases have been determined in a number of molecular simulation studies30–36 and phenomenological studies.37–44 Simulation studies showed that the structural properties of these phases can be captured by molecular models of n-alkane chains. Detailed atomistic simulations of shorter chains by Ryckaert et al.30,31 confirmed the importance of translations in the rotator phases. Molecular dynamics simulations carried out by Ryckaert et al.31 showed that in the crystalline, orthorhombic phase, the chains remain all-trans and fully ordered, with a herringbone packing. By contrast, in the pseudohexagonal RI phase, a dramatic increase in longitudinal chain motion is observed. Cao et al.32 studied the solid–fluid and solid–solid phase equilibria for binary mixtures of hard sphere chains, modelling n-alkanes, using Monte Carlo simulations and observed the first order character of the RII–RI transition. Marbeuf et al.34 studied the pre-melting phenomena and molecular motions in the RI and RII phases of n-alkanes below the melting point using molecular dynamics simulations. Wentzel et al.35,36 performed atomistic molecular dynamics simulations of ordered phases in pure C23 alkane and mixed C21C23 alkane systems, using several different all-atom potentials and confirmed the weak first order character of the RII–RI transition. Mukherjee37–44 extensively studied the various rotator–rotator phase transitions within the framework of the Landau phenomenological model and renormalization group theory.

To the best of the author’s knowledge there are so far no detailed theoretical or simulation studies on the binary mixture of C24H50 and liquid crystals (BBOA). The purpose of the present paper is to develop a phenomenological theory to discuss some of the experimental results observed in binary mixtures of C24H50 and liquid crystals (BBOA) and to determine the influence of the concentration of the liquid crystal component on the various rotator phase transitions. The appearance of the induced rotator phase RIII in the binary system is discussed in great detail. It is shown how the coupling between the concentration and the order parameters shifts the transition temperatures.

II. Theory

In this section we apply the Landau theory of phase transitions, which has been quite successful37–41 in explaining many features of rotator phases in n-alkanes. First we discuss the rotator phase transitions in pure alkane (C24H50). For simplicity, we neglect the weak interlayer interactions between the stacking layers in the X and rotator phases so that the problem becomes two-dimensional. The lattice distortion is the same in the X, RI and RV phases. The long-range herringbone order disappears in the rotator phases. Thus we take long-range herringbone order to be an order parameter for the RV–X and RI–X transitions. The herringbone ordering is a type of crystallization. Thus, the herringbone order parameter is described by7,45 Φ = ϕ[thin space (1/6-em)]exp(i2η), where η is the angle between the local herringbone axis and a reference direction (Fig. 3(c)). The RV phase differs from the RI phase by the tilt angle (Fig. 3(b)), and so we take the tilt angle θ as an another order parameter for the RI–RV transition. The RI phase differs from the RII phase only in the distortion of the hexagonal lattice. Sirota et al.8,9 define the lattice distortion parameter D = 1 − a/b based on an ellipse passing through all six nearest neighbors of a given molecule (Fig. 3(a)). a and b are the minor and major axes of an ellipse passing through all six nearest neighbors when viewed along the axis of the chains. However, Kaganer et al.7,46 suggested the symmetrized form of the distortion parameter ξ = (a2b2)/(a2 + b2), where a and b are the major and minor axes of the ellipse. The distortion, D or ξ, must be defined with respect to a plane whose normal is parallel to the long molecular axes.6 We take ξ as the order parameter of the RI–RII transition, as it breaks the 6-fold symmetry of the RII phase. It is zero in the high-symmetry RII phase and finite in the low-symmetry RI phase. Thus we take the herringbone order, tilt angle and lattice distortion as three primary order parameters involved in the various transitions involved in C24H50. Explicitly, we consider the ordering of a 2D hexagonal crystal. Thus the herringbone order, tilt angle, distortion and hexatic order are described by two-component order parameters. The components of the herringbone order can be expressed as amplitude ϕ and azimuth 2η. The tilt components can be expressed through a polar tilt angle θ and the tilt azimuth δ. The distortion components are expressed through the distortion amplitude ξ and the azimuth 2ω. The multiplier 2 comes from the fact that the distortion is a symmetric traceless tensor. Then the Landau free energy should be invariant ηη + π/3, δδ + π/3, ωω + π/3. Since the free energy is a scalar, the expansion only contains terms that are invariant combinations of the order parameters ξ, θ and ϕ. Expanding the total free energy F in terms of the abovementioned order parameters yields38
 
image file: c4ra14116d-t1.tif(2.1)
where FII is free energy of the RII phase. The coefficients p, α and a are temperature dependent. The other coefficients are assumed to be temperature independent. The first three terms in the free energy eqn (2.1) describe the herringbone order variation and first order RV–X transition for θ = 0. The fluctuations and the cubic invariant of ξ in the free energy expansion give rise to the first order RI–X and RV–X transitions. The last two terms in the first line describe the tilt angle variation and the first term in the second line determines the tilt azimuth δ. For γ > 0, the only minimum of F is at δ = nπ/3 (n is an integer), i.e. the tilt occurs in the nearest neighbor (NN) direction. If γ < 0, the minimum is achieved at δ = π/6 + nπ/3, i.e. for tilt in the direction of the next nearest neighbor (NNN). Since tilt occurs in the NNN direction in the RV phase, eqn (2.1) describes a second order RI–RV transition for β > 0, γ < 0 and ϕ = 0. For θ = 0 and ϕ = 0, the second, third and fourth terms in the third line describes a first order RII–RI transition for b > 0 and c > 0. In this case, the minimum free energy occurs at ω = 0 for b > 0 and at ω = π/2 for b < 0. According to the experimental observations, in the RI phase distortion amounts to a compression in the NN bond direction, hence eqn (2.1) describes a first order RII–RI transition for b > 0, c > 0, θ = 0 and ϕ = 0. The ϕθ coupling term possesses a minimum at δ = 2η + π/2 for E > 0 and δ = 2η for E < 0. The coupling term ξϕ gives a minimum at ω = 2η for F > 0 and ω = 2η + π/2 for F < 0. In the X phase, distortion is in the same direction as the herringbone order. Since in the X phase ω = η = 0, we take F > 0 to describe the RV–X transition. The θξ coupling term gives ω = δ for J > 0 and ω = δ + π/2 for G < 0. Since in the RV phase δ = ω = 0, we take J > 0 for the description of the RI–RV transition.

image file: c4ra14116d-f3.tif
Fig. 3 (a) Diagram showing the definition of the distortion order parameter ξ. (b) Diagram showing the tilt angle θ in the RV phase. (c) Diagram showing possible herringbone orientations in the crystal (X) phase and the angle between the axis along which the nearest neighbors have parallel body orientations and the reference direction defining the herringbone angle η.

Now we consider the binary mixture of n-tetracosane (C24H50) and the liquid crystal compound (BBOA). When mixing C24H50 with BBOA the experiment confirmed the existence of the induced RIII phase between the RI and RV phases. To describe the anisotropy of the liquid crystal molecules, we introduced a scalar order parameter ψ, which models the orientation of the liquid crystal. The quantity image file: c4ra14116d-t2.tif defines the strength of the nematic ordering,47 where θ1 is the angle between the director and the long molecular axis. Furthermore, in the case of this mixture, the free energy must be expressed in terms of the symmetry-breaking order parameters and the concentration (weight percent) of the solute. Let x be the concentration (wt%) of BBOA in a mixture of C24H50–BBOA.

Consequently, the free energy for the binary mixture of C24H50–BBOA can be expressed as

 
image file: c4ra14116d-t3.tif(2.2)
where μ is the chemical potential difference between two components. The local coupling terms G1, H1 and K1 guarantee that the anisotropy of the liquid crystal molecules remains small in the rotator phases where θ and ξ are large. The term image file: c4ra14116d-t4.tif in the free energy eqn (2.2) is the entropic cost of imposing anisotropy on liquid crystal molecules. G2, K2, H2 and H3 are coupling constants that describe the strength of the interactions between the liquid crystals and n-alkane. According to the experimental observations, for the RIII phase ω ≠ 0 and δ ≠ 0. The θξ coupling term gives ω = δ for J > 0 and ω = δ + π/2 for J < 0. Since in the RV phase δ = ω = 0, we take J > 0 for the description of the RIII–RV transition. The higher order terms like ξ2θ2[thin space (1/6-em)]cos(2δ + 4ω) and ξθ4[thin space (1/6-em)]cos(4δ + 2ω) can be added, when necessary. We will see later that the higher order coupling terms d and η1 are responsible for the description of the RI–RIII and RIII–RV transitions. Here, we assume a = a0(TT*1), α = α0(TT*2) and p = p0(TT*3). T*1, T*2 and T*3 are the hypothetical second order transition temperatures. a0, p0 and α0 are positive constants. Furthermore, we assume b = b0(xx0) and γ = γ0(xx0). x0 is the value of the concentration of BBOA for which the RIII phase begins to appear. b0 and γ0 are positive constants. In order of increasing concentration of the liquid crystal component, the relevant stable solutions of eqn (2.2) are as follows:

(i) RII phase: ϕ = 0, θ = 0 and ξ = 0.

This phase exists for p > 0, α > 0, and a > 0.

(ii) RI phase: ϕ = 0, θ = 0 and ξ ≠ 0.

This phase exists for p > 0, α > 0 and a < 0.

(iii) RIII phase: ϕ = 0, θ ≠ 0, and ξ ≠ 0 with tilt, and the distortion azimuth is directed along the tilt azimuth.

This phase exists for d ≠ 0, p + 2 > 0, η1 ≠ 0, α < 0 and a < 0.

(iii) RV phase: ϕ = 0, θ ≠ 0, and ξ ≠ 0.

This phase exists for p + 2 > 0, α < 0 and a < 0.

(iv) X phase: ϕ ≠ 0, θ = 0 and ξ ≠ 0.

This phase exists for p < 0, α > 0 and a < 0.

All of these phases and their stabilities are listed in Table 1.

Table 1 Crystalline and rotator phases in the mixture of C24H50–BBOA
Phase type Order parameters Stability conditions
RII phase ϕ = 0, θ = 0, ξ = 0 p > 0, α > 0, a > 0
RI phase ϕ = 0, θ = 0, ξ ≠ 0 p > 0, α > 0, a < 0
RIII phase ϕ = 0, θ ≠ 0, ξ ≠ 0, ω ≠ 0, δ ≠ 0 p + 2 > 0, η ≠ 0, α < 0
RV phase ϕ = 0, θ ≠ 0, ξ ≠ 0 p + 2 > 0, α < 0, a < 0
X Phase ϕ ≠ 0, θ = 0, ξ ≠ 0 p < 0, α > 0, a < 0


Here we must point out that the isotropic liquid phase (IL) must appear above the RII phase. By lowering the temperature from the isotropic liquid phase, the above phases can appear sequentially. Thus it is clear from the solutions that four types of transition are possible: (i) RII–RI; (ii) RI–RIII; (iii) RIII–RV; and (iv) RV–X, in addition to the IL–RII transition. Experimental results29 show that the above transitions are possible.

The elimination of ψ from eqn (2.2) and minimizing eqn (2.2) (after the elimination of ψ) with respect to x yields

 
image file: c4ra14116d-t5.tif(2.3)

Eqn (2.3) does not give ξ, ϕ and θ in the closed form. If image file: c4ra14116d-t6.tif, image file: c4ra14116d-t7.tif, image file: c4ra14116d-t8.tif, etc. are small in the region of interest, the exponential can be expanded, giving

 
image file: c4ra14116d-t9.tif(2.4)
where image file: c4ra14116d-t10.tif, image file: c4ra14116d-t11.tif, image file: c4ra14116d-t12.tif, image file: c4ra14116d-t13.tif and image file: c4ra14116d-t14.tif

From the relation eqn (2.4) it is clear that the additional terms δx1, δx2, δx3 and δx4 at the concentration x are related to the RI, X, and RV order parameters ξ, ϕ and θ and the azimuth ω and δ of the RIII phase. In the case of second order RV–X, RI–RIII, RIII–RV and RII–RI transitions ϕ = 0, θ = 0, ξ = 0, δ = 0, and ω = 0, and δx1 = 0, δx2 = 0, δx3 = 0, and δx4 = 0. Hence only the RII phase appears above the RI phase. In the case of the first order RV–X transition the jump δx2 is proportional to that of ϕ. For the first order RI–RIII and RIII–RV transitions, the jumps δx3 and δx4 are proportional to those of θ, δ and ω. For the first order RII–RI transition the jump δx1 is proportional to that of ξ. Thus, as long as δx1 ≠ 0, δx2 ≠ 0, δx3 ≠ 0 and δx4 ≠ 0 there is a possibility of two phase regions. Then the RV + X and RII + RI phases can coexist, i.e. a two phase region appears. Thus, the addition of the liquid crystal component BBOA to the pure n-alkane C24H50 forms a two phase region. Experimental results29 show the possibility of the formation of the RV + X and RII + RI two phase regions. The above analysis agrees well with experimental phase diagrams (Fig. 2).

We will now discuss the effect of the concentration of BBOA on the different transitions in detail.

A. RV–X transition

According to the experimental observations,6 the RV–X transition is the onset of long-range herringbone order. Thus, to describe the first order RV–X transition we now consider the free energy in terms of the order parameters ϕ, ξ and ψ and the concentration x. Expanding the free energy near the RV–X phase transition to the lowest relevant order yields
 
image file: c4ra14116d-t15.tif(2.5)

FV is the free energy of the RV phase and a1 is a positive constant.

Eliminating the equilibrium values of ξ and ψ from eqn (2.5) leads to the description of the free energy becoming

 
image file: c4ra14116d-t16.tif(2.6)
where
image file: c4ra14116d-t17.tif

image file: c4ra14116d-t18.tif

image file: c4ra14116d-t19.tif

We notice from the renormalized coefficient q* that taking into account the coupling of ϕ, ψ and ξ with the concentration x leads to the renormalization of the coefficient q. Hence the coefficient q changes with the changes in the concentration x. Then the order of the RV–X transition can also change. The quantitative determination of q* will reflect the strength of the first order character of the RV–X transition. For a low concentration x of BBOA, q* > 0. Then the RV–X transition becomes first order, which occurs in the C24H50–BBOA system. In this case, both the RV and X phases can coexist, i.e. a two phase region appears. This two phase region is formed due to the presence of δx2 in x in eqn (2.4). The RV–X transition is accompanied by a jump of the herringbone order parameter ΦV–X.

The renormalized RV–X transition temperature TV–X is given by

 
image file: c4ra14116d-t20.tif(2.7)

Eqn (2.7) can be rewritten in the expansion of x as

 
TV–X(x) = A1B1x + C1x2 (2.8)
where image file: c4ra14116d-t21.tif, image file: c4ra14116d-t22.tif, image file: c4ra14116d-t23.tif, image file: c4ra14116d-t24.tif, image file: c4ra14116d-t25.tif and image file: c4ra14116d-t26.tif

Experimental findings29 have shown that the RV–X transition temperature decreases with the addition of the liquid crystal component BBOA to the pure alkane. The decrease in the RV–X transition temperature is connected with the width of the two phase region and the entropy of the transition.

The jump of the enthalpy density at the transition point is

 
image file: c4ra14116d-t27.tif(2.9)

Eqn (2.9) can be rewritten in the expansion of x as

 
ΔHV–X = A0B0x + C0x2 + D0x3 (2.10)
where image file: c4ra14116d-t28.tif, image file: c4ra14116d-t29.tif, image file: c4ra14116d-t30.tif and image file: c4ra14116d-t31.tif.

B. RI–RIII and RIII–RV transitions

We will now discuss the effect of the liquid crystal component on the RI–RIII and RIII–RV transitions in the mixture of C24H50–BBOA. Then the total free energy can be written as
 
image file: c4ra14116d-t32.tif(2.11)

The RIII phase is an intermediate tilt where the tilt azimuth δ varies from 0° to 30°. Because the distortion itself is relatively small in the RIII phase, it was difficult to uniquely determine δ and ω independently. Hence for the RIII phase, we assume δ = ω ≠ 0. The distortion ξ in the RIII phase is very weak and is induced by a finite coupling to the tilt magnitude θ. Thus the tilt causes an induced distortion ξθ2. As we already pointed out, the lowest order angle dependent term for the ξθ coupling gives ω = δ for J > 0. Therefore J is constant. A positive value of η1 and positive value of γ favors the RIII phase over the RV and RI phases and would describe the RI–RIII and RIII–RV transitions in the mixture of C24H50–BBOA. Thus, a positive value of η1 and negative value of γ favor the RV phase, corresponding to the mixture of n-tetracosane and an anisometric component. For α > 0, η1 = 0 and γ = 0 describe the RI phase.

The second order character of the RIII–RV transition can be explained by taking into account the term ∼−6[thin space (1/6-em)]cos[thin space (1/6-em)]12ω in the free energy expansion eqn (2.11). Since b is a function of the concentration, it can be positive as well as negative. Assuming d is positive, then for b = −23, i.e. for b < 0, the only minimum is at ω = 0°. Then the system is in the RV phase. When b = 23, i.e. b > 0, the minimum at ω = 0° passes through 0° to 30°. We therefore have a second order transition from the RIII phase to the RV phase at b = 23. As b increases from −23 to 23, the minimum in the RIII phase shifts from 0° to 30°.

Again, since γ is also a function of concentration, it can be positive as well as negative. Assuming η1 is positive, then for γ = −2η1θ6, i.e. for γ < 0, the only minimum is at δ = 0°. Then the system is in the RV phase. When γ = 2η1θ6, i.e. γ > 0, the minimum at δ = 0° passes through 0° to 30°. We therefore have a second order transition from the RIII phase to the RV phase at γ = 2η1θ6. As γ increases from −2η1θ6 to 2η1θ6, the minimum in the RIII phase shifts from 0° to 30°.

We will now discuss the RI–RIII transition in the mixture of C24H50–BBOA.

After the substitution of ω = δ into the free energy eqn (2.11) and elimination of ψ from eqn (2.11), the free energy near the RI–RIII transition reads

 
image file: c4ra14116d-t33.tif(2.12)
where FI(ξ) is the free energy of the RI phase. image file: c4ra14116d-t34.tif

The minimization of eqn (2.12) over δ gives

 
image file: c4ra14116d-t35.tif(2.13)

Suppose that η* is fixed at a negative value. γ decreases from a positive to negative value since γ is a function of x. When γ = 2η*θ6, i.e. for γ > 0, the only minimum is at δ = 0°. Then the system is in the RI phase. When γ = −2η*θ6, i.e. for γ < 0, then the minimum at δ = 0° passes through 0° to 30°. Since in the RIII phase δ changes from 0° to 30°, we have the RIII phase for γ < 0. We therefore have a second order transition from the RIII phase to the RI phase at γ = −2η*θ6, as observed in the mixture of C24H50–BBOA. As γ decreases from 2η*θ6 to −2η*θ6, the minimum in the RIII phase shifts from 0° to 30°as observed experimentally. The variation of the tilt azimuth δ with temperature in the RIII phase can be obtained by the expansion of cos[thin space (1/6-em)]6δ and cos[thin space (1/6-em)]12δ over the powers of δ.

Experimental findings29 have also shown that a second order transition takes place between the RI and RIII phases. Near the RI–RIII transition the free energy reads

 
image file: c4ra14116d-t36.tif(2.14)

Here the tilt causes induced distortion ξθ2. Then the renormalized free energy reads

 
image file: c4ra14116d-t37.tif(2.15)
with image file: c4ra14116d-t38.tif and α1 = α0(TTC).

So tilt θ = 0 for T > TC and θ ≠ 0 for T < TC. So a second order transition takes place between the RI and RIII phases. Here tilt is the order parameter for the RI–RIII phase transition.

C. RII–RI transition

The RII–RI transition is accompanied by the finite jump of the distortion order parameter. This indicates the first order character of the RII–RI transition. Now the expansion of the free energy including the distortion order parameter ξ, anisotropic order parameter ψ and concentration x near the RII–RI transition can be written as
 
image file: c4ra14116d-t39.tif(2.16)

We now write down the free energy in terms of ξ only. So the elimination of ψ from the free energy eqn (2.16) yields

 
image file: c4ra14116d-t40.tif(2.17)
where image file: c4ra14116d-t41.tif

The variation of the distortion order parameter ξ(x,T) with x in the RI phase is given by

 
image file: c4ra14116d-t42.tif(2.18)
where ξ+II–I = b/2c. Eqn (2.18) shows that the distortion order parameter decreases with an increase in the concentration of the liquid crystal component in the RI phase. To show more clearly the variation of the order parameter ξ with the concentration of the liquid crystal component as well as temperature in the RI phase, we have plotted the order parameter ξ for three different concentrations (x) in Fig. 4. This is done for a set of phenomenological parameters for which the RII–RI transition is possible. Fig. 4 shows decrease of ξ with concentration x. This analysis clearly supports the experimental results.29


image file: c4ra14116d-f4.tif
Fig. 4 Temperature variation of the order parameter ξ in the RI phase for the three different concentrations x = 5, x = 10 and x = 15.

The RII–RI transition temperature TII–I is given by

 
image file: c4ra14116d-t43.tif(2.19)

Eqn (2.19) can be rewritten in the expansion of x as

 
image file: c4ra14116d-t44.tif(2.20)
where image file: c4ra14116d-t45.tif, image file: c4ra14116d-t46.tif and image file: c4ra14116d-t47.tif

The jump of the enthalpy density at the transition point is

 
image file: c4ra14116d-t48.tif(2.21)

Eqn (2.20) predicts that the RII–RI transition temperature decreases with the concentration x, which also agrees with experimental results. The jump of the order parameter ξII–I also changes with changing concentration. Since the Landau coefficient b changes with the concentration, b can become smaller. The smaller the value of b, the weaker the first order character of the RII–RI transition. Thus in the binary mixture of C24H50–BBOA, one can expect the weak first order character of the RII–RI transition, as confirmed experimentally.29

D. IL–RII transition

Zammit et al.24 studied the IL–RII transition in binary mixtures of alkanes. Over the IL–RII transition, they observed that the single peak in both the specific heat and latent heat in the pure material splits into two features at different temperatures. This indicates the first order character of the IL–RII transition. Kumar et al.29 observed the weakening of the IL–RII transition as the concentration of BBOA increased. The enthalpy jump became half at the IL–RII transition with the addition of BBOA. Now, to describe the IL–RII transition, we have to define the suitable order parameter for the IL–RII transition. The RII phase has a trilayer stacking sequence. The trilayer stacking can be represented as L1L2L3L1L2L3L1L2L3… Following our previous work,41 we represent the probability of the k-th molecule of the system to occupy any of the j-th layer sequence as Pk(Lj), one of which is 1 and the other 0 in perfect layer ordering. Now we define a correlation factor βk for the k-th molecule in a trilayer structure, i.e. in the RII phase.
 
βk = Pk(L1)Pk(L2) + Pk(L2)Pk(L3) + Pk(L3)Pk(L1). (2.22)

Its value is 0 (or almost zero) in the RII phase. βk does have a finite value in the liquid phase which can be calculated. In the IL phase the probability density of the k-th molecule is constant everywhere in space. So, Pk(L1) = Pk(L2) = Pk(L3). In the IL phase the probability density of the k-th molecule is constant everywhere in space. So

 
image file: c4ra14116d-t49.tif(2.23)
where Nlayer is the total number of layers. Hence βk can be expressed as
 
image file: c4ra14116d-t50.tif(2.24)

Thus βk is independent of k. Hence we define the correlation order parameter ζ as41

 
image file: c4ra14116d-t51.tif(2.25)

Here the average correlation factor is calculated for all molecules together.

Hence ζ is 0 in the IL phase and nonzero in the RII phase. Thus, we take ζ and ψ as two order parameters involved in the IL–RII transition in the mixture of C24H50–BBOA. Then the free energy near the IL–RII transition can be expressed as

 
image file: c4ra14116d-t52.tif(2.26)
where F0 is the free energy of the isotropic liquid phase. The material parameters u and v can be assumed as u = u0(TT*4) and v = v0(xx0). T*4 is the virtual transition temperature. u0 and v0 are constant.

Again, we now write down the free energy in terms of ζ only. So the elimination of ψ from the free energy eqn (2.26) yields

 
image file: c4ra14116d-t53.tif(2.27)
where image file: c4ra14116d-t54.tif.

The variation of the correlation order parameter ζ(x,T) with x in the RII phase is given by

 
image file: c4ra14116d-t55.tif(2.28)
where ζ+II–I = v/2w. Eqn (2.28) shows that ζ changes with changes in the concentration x in the RII phase. The IL–RII transition temperature TIL–II is given by
 
image file: c4ra14116d-t56.tif(2.29)

Eqn (2.29) can be rewritten in the expansion of x as

 
TIL–II(x) = A3B3x + C3x2 (2.30)
where image file: c4ra14116d-t57.tif, image file: c4ra14116d-t58.tif and image file: c4ra14116d-t59.tif

The jump of the enthalpy density at the transition point is

 
image file: c4ra14116d-t60.tif(2.31)

Eqn (2.31) can be rewritten in the expansion of x as

 
ΔHIL–II = (A4B4x + C4x3)(A3B3x + C3x2) (2.32)
where image file: c4ra14116d-t61.tif, image file: c4ra14116d-t62.tif and image file: c4ra14116d-t63.tif

Eqn (2.30) predicts that the IL–RII transition temperature decreases with the concentration x which also agrees with experimental results. The jump of the order parameter ζIL–II also changes with changing concentration. Since the Landau coefficients change with the concentration, v can become smaller. The smaller the value of v, the weaker the first order character of the IL–RII transition. Thus, in the binary mixture one can also expect the weak first order character of the IL–RII transition, as confirmed experimentally.29

III. Comparison with experimental findings

In this section we will compare our theoretical results of the effect of the concentration of the liquid crystal solute on the IL–RII, RII–RI and RV–X transitions with the measurements of Kumar et al.29 on the C24H50–BBOA mixture. These are, to the best of our knowledge, the only experimental results on the C24H50–BBOA mixture available in the literature.

For temperature versus concentration (T vs. x) phase diagrams for the IL–RIII, RII–RI and RV–X transitions there exists only one experiment.29 From eqn (2.18) and eqn (2.28), when ξ and ζ are fixed T vs. x should be non-linear, which agrees well with the experiment findings.29 In principle, eqn (2.8) shows that the TV–X vs. x phase diagram is also non-linear, which also agree with the experiment findings. According to eqn (2.8), eqn (2.20) and eqn (2.30), the RV–X, RII–RI and IL–RII transition temperatures decrease with an increasing concentration of the liquid crystal solute, as observed experimentally.29 According to eqn (2.10) and eqn (2.32), the RV–X and IL–RII transition enthalpies also decrease with increasing concentrations of the liquid crystal solute, as observed experimentally.29 In order to check eqn (2.8), eqn (2.20) and eqn (2.30), the measured T vs. x for the RV–X, RII–RI and IL–RIII transitions of the C24H50–BBOA mixture of Kumar et al.29 is plotted in Fig. 5. The fit yields A1 = 42.52 °C, B1 = 0.25 °C wt%−1, C1 = 0.0094 °C wt%−2, A2 = 46.39 °C, B2 = 0.081 °C wt%−1, C2 = 0.0031 °C wt%−2, A3 = 49.65 °C, B3 = 0.14 °C wt%−1 and C3 = 0.0018 °C wt%−2. Furthermore, to check eqn (2.10) and eqn (2.32) the ΔHV–X vs. x and ΔHIL–II vs. x of Kumar et al.29 are plotted in Fig. 6. The lines are the best fits from eqn (2.10) and eqn (2.32) to the data, resulting in the parameter values A0 = 71.22 J g−1, B0 = 5.22 J g−1 wt%−1, C0 = 0.54 J g−1 wt%−2, D0 = −0.015 J g−1 wt%−3, A3 = 49.65 °C, B3 = 4.66 °C wt%−1, C3 = 0.11 °C wt%−2, A4 = 3.23 J g−1 °C−1, B4 = 0.91 J g−1 °C−1 wt%−1 and C4 = 0.061 J g−1 °C−1 wt%−2. The fits to the measured values are good in Fig. 5 and 6. The agreement of the theory with the experiment is very good considering the scattered experimental data.


image file: c4ra14116d-f5.tif
Fig. 5 The concentration dependence of the IL–RII, RII–RI and RV–X transition temperatures in the mixture of C24H50–BBOA. The measured data (points) are from ref. 29 and the lines are the best fits from eqn (2.8), eqn (2.20) and eqn (2.30).

image file: c4ra14116d-f6.tif
Fig. 6 The concentration dependence enthalpy for the IL–RII and RV–X transitions in the mixture of C24H50–BBOA. The measured data (points) are from ref. 29 and the lines are the best fits from eqn (2.10) and eqn (2.32).

IV. Conclusions

We have presented here a mean-field analysis to describe the rotator–crystal and rotator–rotator phase transitions in the mixture of n-alkane and liquid crystals. The theory describes the effect of the liquid crystal component on the IL–RII, RII–RI, RI–RIII, RIII–RV and RV–X transitions. The coupling between the anisotropic order parameter of the liquid crystals and the macroscopic order parameters of the rotator phases of n-alkane is found to play a crucial role in determining the phase behavior and the order of the transitions. We found that although the RIII phase is absent in pure forms of C24H50, the RIII phase is induced with the increase of the concentration of the liquid crystal solute in the mixture of C24H50–BBOA. The RV–X, RII–RI and IL–RII transitions must always be first order, even in the mixture of C24H50–BBOA. The RIII–RV and RI–RIII transitions are second order in the mixture of C24H50–BBOA. Thus, our results are in qualitative agreement with experimental results. The suitable choice of the functional forms of the Landau coefficients, particularly the coefficients of γ and b will provide better ideas for the construction of the free energy and description of the RIII–RV and RI–RIII transitions. The values of the coefficients of γ and b are very sensitive for the description of the RIII phase. In order to gain a better understanding of the effect of the liquid crystal component on the rotator phase transitions in the mixture of C24H50–BBOA, further experimental and theoretical works are needed.

References

  1. X. Z. Wu, B. M. Ocko, E. B. Sirota, S. K. Sinha, M. Deutsch, B. H. Cao and M. W. Kim, Science, 1993, 261, 1018–1021 CAS.
  2. D. Turnbull and R. L. Cormia, J. Chem. Phys., 1961, 34, 820–831 CrossRef CAS PubMed.
  3. D. Uhlmann, G. Kritchevsky, R. Straff and G. Scherer, J. Chem. Phys., 1975, 62, 4896–4903 CrossRef CAS PubMed.
  4. A. B. Herhold, H. E. King Jr and E. B. Sirota, J. Chem. Phys., 2002, 116, 9036–9050 CrossRef CAS PubMed.
  5. E. B. Sirota and A. B. Herhold, Science, 1999, 283, 529–532 CrossRef CAS.
  6. E. B. Sirota, Langmuir, 1997, 13, 3849–3859 CrossRef CAS.
  7. V. M. Kaganer, H. Möhwald and P. Dutta, Rev. Mod. Phys., 1999, 71, 779–819 CrossRef CAS.
  8. E. B. Sirota, H. E. King Jr, D. M. Singer and H. H. Shao, J. Chem. Phys., 1993, 98, 5809–5824 CrossRef CAS PubMed.
  9. E. B. Sirota and D. M. Singer, J. Chem. Phys., 1994, 101, 10873–10882 CrossRef CAS PubMed.
  10. G. R. Strobl, B. Ewen, E. W. Fischer and W. J. Piesczek, J. Chem. Phys., 1974, 61, 5257–5264 CrossRef PubMed.
  11. J. Doucet, I. Denicolo and A. Craievich, J. Chem. Phys., 1981, 75, 1523–1529 CrossRef CAS PubMed.
  12. J. Doucet, I. Denicolo, A. Craievich and A. Collet, J. Chem. Phys., 1981, 75, 5125–5127 CrossRef CAS PubMed.
  13. J. Doucet, I. Denicolo and A. Craievich, J. Chem. Phys., 1983, 78, 1465–1469 CrossRef PubMed.
  14. G. Ungar and N. Masic, J. Phys. Chem., 1983, 89, 1036–1042 CrossRef.
  15. G. Ungar, J. Phys. Chem., 1983, 87, 689–695 CrossRef CAS.
  16. T. M. Bohanon, B. Lin, M. C. Shih, G. E. Ice and P. Dutta, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 4846R–4849R CrossRef.
  17. A. M. Levelut, J. Phys., Colloq., 1976, 37, 51–56 CrossRef.
  18. R. Pindak, D. E. Moncton, S. C. Davey and J. W. Goodby, Phys. Rev. Lett., 1981, 46, 1135–1138 CrossRef CAS.
  19. M. Maroncelli, S. P. Qi, H. L. Strauss and R. G. Snyder, J. Am. Chem. Soc., 1982, 104, 6237–6247 CrossRef CAS.
  20. I. Denicolo, A. F. Craievich and J. Doucet, J. Chem. Phys., 1984, 80, 6200–6203 CrossRef CAS PubMed.
  21. E. B. Sirota Jr, H. E. King Jr, G. J. Hughes and W. K. Wan, Phys. Rev. Lett., 1992, 68, 492–495 CrossRef.
  22. R. G. Snyder, G. Conti, H. L. Strauss and D. L. Dorset, J. Phys. Chem., 1993, 97, 7342–7350 CrossRef CAS.
  23. E. B. Sirota and X. Z. Wu, J. Chem. Phys., 1996, 105, 7763–7773 CrossRef CAS PubMed.
  24. U. Zammit, M. Marinelli and F. J. Mercuri, J. Phys. Chem. B, 2010, 114, 8134–8139 CrossRef CAS PubMed.
  25. H. Rensmo, A. Ongaro, D. Ryam and D. Fitzmaurice, J. Mater. Chem., 2002, 12, 2762–2768 RSC.
  26. V. J. Kumar, S. K. Prasad and D. S. S. Rao, Langmuir, 2010, 26, 18362–18368 CrossRef PubMed.
  27. K. Jiang, B. Xiae, D. Fu, F. Luo, G. Liu, Y. Su and D. Wang, J. Phys. Chem. B, 2010, 114, 1388–1392 CrossRef CAS PubMed.
  28. U. Zammit, M. Marinelli, F. Mercuri and F. Scudieri, J. Phys. Chem. B, 2011, 115, 2331–2337 CrossRef CAS PubMed.
  29. V. J. Kumar, S. K. Prasad, D. S. S. Rao and P. K. Mukherjee, Langmuir, 2014, 30, 4465–4473 CrossRef PubMed.
  30. J. P. Ryckaert and M. L. Klein, J. Chem. Phys., 1986, 85, 1613–1620 CrossRef CAS PubMed.
  31. J.-P. Ryckaert, M. L. Klein and I. R. McDonald, Phys. Rev. Lett., 1987, 58, 698–701 CrossRef.
  32. M. Cao and P. A. Monson, J. Chem. Phys., 2004, 120, 2980–2988 CrossRef CAS PubMed.
  33. M. Cao and P. H. Monson, J. Phys. Chem. B, 2009, 113, 13866–13873 CrossRef CAS PubMed.
  34. A. Marbeuf and R. J. Brown, J. Chem. Phys., 2006, 124, 054901 CrossRef PubMed.
  35. N. Wentzel and S. T. Milner, J. Chem. Phys., 2010, 132, 044901 CrossRef PubMed.
  36. N. Wentzel and S. T. Milner, J. Chem. Phys., 2011, 134, 224504 CrossRef PubMed.
  37. P. K. Mukherjee and M. Deutsch, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 60, 3154–3162 CrossRef CAS.
  38. P. K. Mukherjee, J. Chem. Phys., 2002, 116, 10787–10793 CrossRef CAS PubMed.
  39. P. K. Mukherjee, J. Chem. Phys., 2007, 126, 114501 CrossRef PubMed.
  40. P. K. Mukherjee, J. Chem. Phys., 2008, 129, 021101 CrossRef PubMed.
  41. P. K. Mukherjee, J. Phys. Chem. B, 2010, 114, 5700–5703 CrossRef CAS PubMed.
  42. P. K. Mukherjee, J. Chem. Phys., 2011, 135, 134505 CrossRef PubMed.
  43. P. K. Mukherjee, J. Phys. Chem. B, 2012, 116, 1517–1523 CrossRef CAS PubMed.
  44. P. K. Mukherjee and S. Dey, J. Mod. Phys., 2012, 3, 80–84 CrossRef CAS.
  45. R. Bruinsma and G. A. Aeppli, Phys. Rev. Lett., 1982, 48, 1625–1628 CrossRef CAS.
  46. V. M. Kaganer and E. B. Loginov, Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 1995, 71, 2599–2602 Search PubMed.
  47. P. G. de Gennes and J. Prost, The Physics of Liquid Crystals, Clarendon Press, Oxford, 1993 Search PubMed.

This journal is © The Royal Society of Chemistry 2015