975
Views
2
CrossRef citations to date
0
Altmetric
Mathematical Modelling, Symmetry and Topology

A predictive theoretical model for stretch-induced instabilities in liquid crystal elastomers

, , , &
Pages 1426-1438 | Received 21 Nov 2022, Published online: 08 Jan 2023

ABSTRACT

Following the experimental lead, we construct a general mathematical model which, depending on whether the uniaxial scalar order parameter is constant or not, can predict either the classical shear striping instability or the molecular auxetic response and mechanical Fréedericksz transition observed in different liquid crystal elastomers. Our theoretical model can shed some light on the role of nematic order in these stretch-induced mechanical instabilities not observed in conventional rubber-like solids.

Graphical abstract

1. Introduction

In Refs [Citation1–6], experimental observations are reported for a nematic liquid crystal elastomer (LCE) where an auxetic response occurs, such that, when a material sample is stretched longitudinally, its volume remains unchanged, but its thickness, measured in the direction perpendicular to the sample’s plane, first decreases, then increases if the deformation is sufficiently large. This behaviour is accompanied by a mechanical Fréedericksz transition whereby the nematic director, which is initially aligned within the sample’s plane and perpendicular to the longitudinal direction, is mechanically deformed to become parallel to the applied force. Experimentally, this can appear as a sudden rotation of the director, and is different from the gradual rotation occurring in many other LCEs where alternating shear stripes form at very low stress [Citation7–13]. Discontinuous rotation of the nematic director was also reported previously by Mitchell et al. [Citation14] (see [Citation15, Section 5.5]).

While the shear-striping phenomenon in nematic LCEs is well understood and has been modelled extensively [Citation8,Citation16–24], auxetic LCEs require more physical and theoretical investigation.

In Mihai et al. [Citation25], a descriptive three-term Ogden-type model [Citation26,Citation27] is adopted which permits the calibration to available experimental data for the auxetic LCE. However, despite the inherent mathematical simplicity of the Ogden strain-energy function, fewer terms are needed in order to assess the effect of changing parameter values.

In this paper, guided by experimental observations, we propose a two-term Ogden-type constitutive model, introduced in Section 2, which, depending on whether the uniaxial scalar order parameter is constant or not, predicts either the auxetic response and the mechanical Fréedericksz transition (Section 2.1) or the classical shear striping (Section 2.2) presented by various LCEs (see ). These results are illustrated by a set of examples presented in Section 3.

Figure 1. (Colour online) Deformation-dependent director rotation leading to alternating shear stripes or large-strain auxetic response in LCEs [Citation28].

Figure 1. (Colour online) Deformation-dependent director rotation leading to alternating shear stripes or large-strain auxetic response in LCEs [Citation28].

2. Model function

We consider the following two-term strain-energy function describing a nematic LCE (see also [Citation25]),

(1) W(lce)=μ12n2λ12n+λ22n+λ32n+μ22n2α12n+α22n+α32n,(1)

where μ1>0, μ2>0 and n>0 are constant parameters, and μ=μ1+μ2 represents the shear modulus at infinitesimal strain. In the above equation, {λ12,λ22,λ32} denote the eigenvalues of the Cauchy–Green tensor FFT, where F is the deformation gradient from the reference cross-linking state, and {α12,α22,α32} are the eigenvalues of AAT, where A=G1FG0 is the local elastic deformation tensor, with G0 and G the ‘natural’ deformation tensors due to the liquid crystal director in the reference and current configuration, respectively. These tensors satisfy the following relations [Citation29] (see also [Citation15, Chapter 3]):

(2) G02=c0I+2Q0,G2=cI+2Q,(2)

with c0 and c representing the effective step length of the polymeric chain, Q0 and Q the symmetric traceless order parameter tensors [15, pp.48-49], and I=diag(1,1,1) the tensor identity. The macroscopic tensor parameters are known to describe orientational order in nematic liquid crystals [Citation30].

In the following sections, we demonstrate that, for a suitable choice of the constitutive parameters, the model defined by (1) is capable of predicting either a large-strain auxetic response accompanied by mechanical Fréedericksz transition or the shear-striping instability.

2.1. Variable order parameter and mechanical Fréedericksz transition

First, we examine the emergence of large-strain auxetic behaviour. To obtain this, for the LCE sample, in a Cartesian system of coordinates (X1,X2,X3), we designate the plane formed by the first two directions as the sample’s plane and the third direction as its thickness direction. The material sample is elastically deformed by application of a tensile force in the first (longitudinal) direction, while the nematic director is initially along the second (transverse) direction.

Setting the nematic director in the reference and current configuration, respectively, as follows:

(3) n0=010,n=sinθcosθ0,(3)

where θ[0,π/2] is the angle between n and n0, the deformation gradient takes the form

(4) F=diag(λ1,λ2,λ3),(4)

with λ1λ2λ3=1 and diag(,,) denoting a diagonal second-order tensor.

Guided by experimental observations (see also [Citation24]), we assume that the rotation of the nematic director within the plane formed by its initial orientation and the applied force from θ=0 to θ=π/2 happens suddenly [Citation25], and denoted by λcrt the longitudinal stretch ratio λ where the rotation occurs. In practice, although the angle θ will also take intermediate values between 0 and π/2, the deformation interval for which the director rotation is observed is very short, and its separate analysis can be omitted. However, in this case, a jump may be observed in the orthogonal stretch ratios λ2 and λ3 at the critical longitudinal stretch λ1=λcrt.

In the reference configuration, the LCE is uniaxial, and the order parameter tensor is equal to [Citation15, p. 14]

(5) Q0=diagQ02,Q0,Q02,(5)

where Q0 is the scalar order parameter.

In the deformed configuration, biaxiality can also emerge [Citation15, p. 15], and therefore

• If θ=0, then the order parameter tensor takes the form

(6) Q=diagQb2,Q,Q+b2;(6)

• If θ=π/2, then

(7) Q=diagQ,Qb2,Q+b2,(7)

where Q and b are the uniaxial and biaxial scalar order parameters, respectively.

Numerically, we will consider both the cases when biaxiality is neglected, i.e. b=0, and when the magnitude of biaxial parameter b increases (decreases) as the uniaxial order parameter decreases (increases).

For the corresponding elastic Cauchy–Green tensor ATA=diagα12,α22,α32, where T denotes the transpose (see also [Citation25]), we have

• If θ=0, then

(8) α12=detI+2QdetI+2Q01/31Q01(Qb)λ12,(8)
(9) α22=detI+2QdetI+2Q01/31+2Q01+2Qλ22,(9)
(10) α32=detI+2QdetI+2Q01/31Q01(Q+b)λ32;(10)

• If θ=π/2, then

(11) α12=detI+2QdetI+2Q01/31Q01+2Qλ12,(11)
(12) α22=detI+2QdetI+2Q01/31+2Q01(Qb)λ22,(12)
(13) α32=detI+2QdetI+2Q01/31Q01(Q+b)λ32.(13)

Under the uniaxial deformation considered here, the second and third directions must be stress free. Then the corresponding principal Cauchy stresses, defined by

(14) Ti(lce)=W(lce)λiλip,i=1,2,3.(14)

with p denoting the usual Lagrange multiplier for the incompressibility constraint, take the form

(15) Ti(lce)=μ1nλi2nλ22nμ2nαi2nα22n=μ1nλi2nλ32nμ2nαi2nα32n,  \break              i=1,2,3.(15)

The associated first Piola–Kirchhoff stresses are

(16) Pi(lce)=Ti(lce)λi1,i=1,2,3.(16)

Taking λ1=λ, we solve for λ2 the equation

(17) T2(lce)T3(lce)=0,(17)

and obtain

(18) λ2=λ2(λ,Q,b)=1λμ1+μ2λ2ng22nμ1+μ2λ2ng32n1/4n,(18)

where g22=α22/λ22 and g32=α32/λ32, i.e.

• If θ=0, then

(19) g22=detI+2QdetI+2Q01/31+2Q01+2Q,(19)
(20) g32=detI+2QdetI+2Q01/31Q01(Q+b);(20)

• If θ=π/2, then

(21) g22=detI+2QdetI+2Q01/31+2Q01(Qb),(21)
(22) g32=detI+2QdetI+2Q01/31Q01(Q+b).(22)

To capture the auxetic response, while keeping our continuum mechanics formulation tractable, we let the uniaxial scalar parameter Q>0 decrease linearly with respect to the longitudinal stretch ratio λ from a given value Q0(0,1), at λ=1, to a minimum value Qcrt(0,Q0) that occurs at the critical stretch ratio λ=λcrt, then increase again, also linearly, to a value Q1(Qcrt,Q0). That is

(23) Q(λ)=C1λ+D1for1λλcrt,C2λ+D2 forλcrtλλmax,(23)

such that Q(1)=Q0, Q(λcrt)=Qcrt and Q(λmax)=Q1.

Similarly, we let the biaxial scalar parameter b<0 to decrease linearly with respect to the longitudinal stretch ratio λ from a given value b0(1,0), at λ=1, to a minimum value bcrt(1,b1), occurring also at the critical stretch ratio λ=λcrt, then increase linearly to a value b1(bcrt,b0). Thus

(24) b(λ)=c1λ+d1for1λλcrt,c2λ+d2 forλcrtλλmax,(24)

such that b(1)=b0, b(λcrt)=bcrt and b(λmax)=b1. Note that the magnitude (absolute value) of this parameter increases then decreases while its numerical value decreases then increases.

Taking λ1=λ, λ2 given by EquationEquation (18), λ3=1/(λ1λ2), and Q=Q(λ) and b=b(λ) described by Equations (23) and (24), respectively, we define the energy function

(25) w(λ,θ)=W(lce)(λ1,λ2,λ3,Q,b,θ),(25)

with W(lce)(λ1,λ2,λ3,Q,b,θ)=W(lce) given by EquationEquation (1).

To find the critical stretch ratio λ=λcrt, we solve the equation

(26) w(λ,0)=w(λ,π/2).(26)

Once λcrt is obtained, we can also determine the minimum stretch ratio λ=λaux at which the auxetic behaviour begins to be observed, i.e. where λ3 starts to increase.

The purpose of this paper is to investigate numerically how a theoretical model of the form given by EquationEquation (1) can capture either a large strain auxetic effect similar to that shown in Mihai et al. [Citation25] or the classical shear striping. After several numerical tests, the proposed model with n=2 was selected as being sufficiently simple to analyse in some detail and able to present each of those effects under different conditions, as demonstrated in the following sections. Henceforth, we set n=2 in the model strain–energy function.

2.2. Constant order parameter and shear striping

Next, we analyse the deformation of the LCE model defined by EquationEquation (1) under uniaxial tensile force, such that the nematic director is uniformly distributed in the sample’s plane and rotates slowly from the initial position perpendicular to the tensile force to align with this force via an intermediate state where alternating shear stripes occur. For the corresponding deformation, the stretch ratios in the directions perpendicular to the applied force are equal, agreeing with experimental observations recorded, for example, in of [Citation31] and of [Citation32].

In this case, we assume that the scalar order parameter Q remains constant and there is no biaxiality, i.e. b=0. This assumption is based on experimental observations showing that variations in the order parameter are relatively small before, during and after stripe domains form [Citation8,Citation12,Citation13,Citation32]. The natural gradient tensor then takes the form

(27) G=a1/6I+a1/3a1/6nn,(27)

where

(28) a=1+2Q1Q(28)

represents the natural anisotropy parameter for the nematic solid, n is the nematic director, denotes the tensor product of two vectors, and I is the identity tensor.

In particular, when the director for the reference and current configuration are given by (3), the associated natural deformation tensors described by (2) are, respectively,

(29) G0=a1/6000a1/3000a1/6(29)

and

(30) G=a1/6+a1/3a1/6sin2θa1/3a1/6sinθcosθ0a1/3a1/6sinθcosθa1/6+a1/3a1/6cos2θ000a1/6.(30)

To demonstrate shear-striping instability, we consider the following perturbed deformation gradient

(31) F=λε00λ1/2000λ1/2,(31)

where λ is the stretch ratio in the longitudinal direction, which is the direction of the applied tensile force, and 0<ε1 is a small shear parameter.

The elastic deformation tensor A=G1FG0 is then equal to

(32) A=λa1/2sin2θ+cos2θλ1/21a1/2sinθcosθ0λa1/21sinθcosθλ1/2a1/2sin2θ+cos2θ000λ1/2+ε0sin2θ+a1/2cos2θ001a1/2sinθcosθ0000.(32)

Taking n=2 in the strain-energy function described by EquationEquation (1), we obtain

(33) W(lce)=μ18λ14+λ24+λ34+μ28α14+α24+α34,(33)

where the eigenvalues {λ12,λ22,λ32} of the Cauchy–Green tensor FFT and {α12,α22,α32} of the tensor AAT satisfy the following relations, respectively,

(34) λ14+λ24+λ34=trFFT22CofFFT(34)

and

(35) α14+α24+α34=CofAAT22trAAT,(35)

with “tr“denoting the trace and “Cof“the cofactor. Hence,

(36) λ14+λ24+λ34=λ2+λ1+ε222λ+λ2(36)

and

(37) α14+α24+α34=λ2λ2a1/2sin2θ+cos2θ2+a1/212sin2(2θ)4+εsin2θ+a1/2cos2θ+λ1/21a1/2sin(2θ)22+ε1a1/2sin(2θ)2+λ1/2a1/2sin2θ+cos2θ222λ1+λ2.(37)

We further define the following function:

(38) w(λ,ε,θ)=W(lce)(λ1,λ2,λ3,Q,b,θ),(38)

with W(lce)(λ1,λ2,λ3,Q,b,θ)=W(lce) described by EquationEquation (33). Differentiating the above function with respect to ε and θ, respectively, gives

(39) w(λ,ε,θ)ε=μ12ελ2+λ1+ε2+μ22λ2λ2a1sin2θ+a1/2cos2θ2+1a1/22sin2(2θ)4+εsin2θ+a1/2cos2θ+λ1/21a1/2sin(2θ)22+ε1a1/2sin(2θ)2+λ1/2a1/2sin2θ+cos2θ2εsin2θ+a1/2cos2θ+λ1/21a1/2sin(2θ)2sin2θ+a1/2cos2θ+ε1a1/2sin(2θ)2+λ1/2a1/2sin2θ+cos2θ1a1/2sin(2θ)2(39)

and

(40) w(λ,ε,θ)θ=μ22λ2λ2a1sin2θ+a1/2cos2θ2+1a1/22sin2(2θ)4+εsin2θ+a1/2cos2θ+λ1/21a1/2sin(2θ)22+ε1a1/2sin(2θ)2+λ1/2a1/2sin2θ+cos2θ2λ2a1sin2θ+a1/2cos2θ1a1/2+1a1/22cos(2θ)2sin(2θ)+εsin2θ+a1/2cos2θ+λ1/21a1/2sin(2θ)2ε1a1/2sin(2θ)+λ1/21a1/2cos(2θ)+ε1a1/2sin(2θ)2+λ1/2a1/2sin2θ+cos2θε1a1/2cos(2θ)+λ1/2a1/21sin(2θ).(40)

As the equilibrium solution minimises the energy, it must satisfy the simultaneous equations

(41) w(λ,ε,θ)ε=0andw(λ,ε,θ)θ=0.(41)

At ε=0 and θ=0, both the partial derivatives defined by EquationEquations (39) and (Equation40) are equal to zero, i.e.

(42) wε(λ,0,0)=wθ(λ,0,0)=0.(42)

Therefore, this trivial solution is always an equilibrium state, and for sufficiently small values of ε and θ, we have the second-order approximation

(43) w(λ,ε,θ)w(λ,0,0)+12ε22wε2(λ,0,0)+2εθ2wεθ(λ,0,0)+θ22wθ2(λ,0,0),(43)

where

(44) 2wε2(λ,0,0)=μ12λ2+λ1+μ22λ2aλ2+λ1,(44)
(45) 2wεθ(λ,0,0)=μ22λ5/2λ2+λ11a,(45)
(46) 2wθ2(λ,0,0)=μ22λ2λ2+λ1λ2a11+λ1a1).(46)

First, we find the equilibrium value θ0 for θ as a function of ε by solving the second equation in (Equation41). By the approximation (EquationEquation (43)) the respective equation takes the form

(47) ε2wεθ(λ,0,0)+θ2wθ2(λ,0,0)=0,(47)

and implies

(48) θ0(ε)=ε2wεθ(λ,0,0)/2wθ2(λ,0,0).(48)

Next, substituting θ=θ0(ε) in (43) gives the following function of ε,

(49) w(λ,ε,θ0(ε))w(λ,0,0)ε222wε2(λ,0,0)2wεθ(λ,0,0)2/2wθ2(λ,0,0).(49)

Depending on whether the expression on the right-hand side in EquationEquation (49) is positive, zero, or negative, the respective equilibrium state is stable, neutrally stable or unstable [Citation23].

We deduce that the equilibrium state with ε=0 and θ=0 is unstable if

(50) λ<λ<a1/3,(50)

where λ solves the equation

(51) μ1μ2λ2+aaλ3a(a1)=0.(51)

Similarly, at ε=0 and θ=π/2, both the partial derivatives defined by EquationEquations (39) and (Equation40) are equal to zero, and

(52) 2wε2(λ,0,π/2)=μ12λ2+λ1+μ22λ2λ2a1+λ1a,(52)
(53) 2wεθ(λ,0,π/2)=μ22λ5/2λ2a1+λ1aa1,(53)
(54) 2wθ2(λ,0,π/2)=μ22λ2λ2a1+λ1aλ21a1+λ11a).(54)

Thus the equilibrium state with ε=0 and θ=π/2 is unstable if

(55) a1/3<λ<λ,(55)

where λ solves the equation

(56) μ1μ2λ2λ2+λ1λ2a1+λ1a+1λ3aa(a1)=0.(56)

There is also an inhomogeneous equilibrium solution (ε0,θ0) when λ satisfies either (50) or (55). For the equilibrium angle θ0(ε), given by EquationEquation (48), the second equation in (Equation41) is satisfied. To find the equilibrium value ε0 of ε as a function of λ, we solve the simultaneous equations EquationEquation (41) numerically.

3. Results

To illustrate the mechanical behaviour of the model described by EquationEquation (1), first we assume that the nematic director rotates suddenly, as described in Section 2.1, and consider the following four cases:

  1. Variable uniaxial order parameter Q and variable biaxiality parameter b, given by EquationEquations (23) and (Equation24), respectively;

  2. Variable uniaxial order parameter Q given by EquationEquation (23), without biaxiality (b=0);

  3. Constant uniaxial order parameter Q and variable biaxiality parameter b given by EquationEquation (24);

  4. Constant uniaxial order parameter Q, without biaxiality (b=0).

We then turn our attention to case where the director rotates slowly and the uniaxial order parameter Q is constant, as analysed in Section 2.2.

In our numerical examples, we chose the maxima and minima of the order parameters Q and b similar to those reported in previous papers where experimental results were described. From those values, the coefficients of the respective piecewise linear functions are inferred, as follows: C1=(Q0Qcrt)/(1λcrt), D1=(QcrtλcrtQ0)/(1λcrt),C2=(QcrtQ1)\break/(λcrtλmax), D2=(λcrtQ1λmaxQcrt)/(λcrtλmax) and c1=(b0bcrt)/(1λcrt), d1=(bcrtλcrtb0)\break/(1λcrt), c2=(bcrtb1)/(λcrtλmax), d2=(λcrtb1\breakλmaxbcrt)/(λcrtλmax). In each case, we set the same initial model parameters of μ1=5 μ2=1, n=2, and Q0=0.5880, the latter being a typical value for the scalar uniaxial order parameter in a nematic liquid crystalline material.

For cases (I)–(IV), summarises how the scalar order parameters change with longitudinal stretch ratio, while shows the sudden change of orientation in the nematic director. For these cases, plots the energy function described by EquationEquation (25). Solving EquationEquation (26) then gives the critical stretch ratio λcrt. In each case, respectively, we have

Figure 2. (Colour online) Scalar uniaxial and biaxial order parameters for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 2. (Colour online) Scalar uniaxial and biaxial order parameters for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 3. (Colour online) The sine function of the angle θ between the nematic director in the current and reference configuration for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is observed. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 3. (Colour online) The sine function of the angle θ between the nematic director in the current and reference configuration for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is observed. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 4. (Colour online) Energy functions for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 4. (Colour online) Energy functions for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.
  1. (Q0,Qcrt,Q1)=(0.5880,0.0010,0.100), (b0,bcrt,b1)=(0.0102,0.1518,0.1056), λcrt=1.5252, while the minimum stretch ratio where auxeticity occurs is λaux=1.2770. Thus, C1=1.1177, D1=1.7057, C2=1.1177, D2=0.2227, c1=0.2696, d1=0.2594, c2=0.0685, d20.2562.

  2. (Q0,Qcrt,Q1)=(0.5880,0.0010,0.1000), b=0, λcrt=1.5514, and the minimum stretch ratio where auxeticity is observed is λaux=1.3579. Thus, C1=1.0647, D1=1.6527, C2=1.0647, D20.2358 and c1=c2=d1=d2=0.

  3. Q=0.5880, (b0,bcrt,b1)=(0.0102,0.1518,\break0.1056), and λcrt=1.6253. Thus, C1=C2\break=0, D1=D2=0.5880, c1=0.2265, d1=0.2163, c2=0.0804, d2=0.2824.

  4. Q=0.5880, b=0, and λcrt=1.6842. Thus, C1=C2=0, D1=D2=0.5880, c1=c2=d1\break=d2=0.

In , cases (I) and (II), we note that the plots of the energy function with θ=0 and θ=π/2, respectively, appear very close for stretches greater than the critical stretch where the director rotates. However, numerically, and also after a closer inspection of those plots, one can see that the energy with θ=π/2, plotted in red, is numerically slightly below that with θ=0, depicted in blue. Before the critical stretch, the energy with θ=0 is below that with θ=π/2. In cases (III) and (IV), the difference between the two curves is more pronounced.

shows the corresponding transverse stretch ratios λ2 and λ3 as functions of the longitudinal stretch ratio λ1=λ. This figure provides insight into the nature of the auxetic response seen in some LCEs. In case (I), we see how non-constant Q and b lead to a physically realistic deformation. Critically, this case predicts large anisotropy in the transverse stretch ratios λ2 and λ3, and also a physically realisable auxetic response. In case (II), where Q varies and b=0, a similar behaviour is predicted, but with a less pronounced auxetic response, as the interval between λaux and λcrt is smaller than in case (I). In cases (III) and (IV), although an auxetic reponse is predicted at λcrt, the model shows large discontinuity in λ2 and λ3 at λ1=λcrt, which can be indicative of the fact that the director may rotate slowly rather than suddenly. We can also see that, before λcrt is reached, in case (III), the non-zero biaxial order parameter allows the evolution of anisotropy in the transverse stretch ratios, while in case (IV), where the uniaxial order parameters is constant, an isotropic deformation in the transverse directions is predicted. displays the associated first Piola-Kirchhoff stresses.

Figure 5. (Colour online) The transverse stretch ratios for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 5. (Colour online) The transverse stretch ratios for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 6. (Colour online) The first Piola-Kirchhoff stresses for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

Figure 6. (Colour online) The first Piola-Kirchhoff stresses for cases (I)–(IV), respectively. For cases (I) and (II), the vertical lines correspond to the predicted longitudinal stretch ratio λcrt where the director rotates suddenly and the minimum stretch ratio λaux<λcrt where auxeticity is obtained. For cases (III) and (IV), the vertical line corresponds to λcrt.

In , the predicted longitudinal stretch ratio λcrt where the director rotates suddenly is indicated in each case. For cases (I) and (II), the minimum stretch ratio λaux<λcrt where auxeticity is obtained is also presented.

In , the critical longitudinal stretch ratios for the director rotation are plotted vs. the parameter ratio μ1/μ2 when n=2 (Q0,Qcrt,Q1)=(0.5880,0.0010,0.1000)andb=0 or (b0,bcrt,b1)=(0.0102,0.1518,0.1056),respectively. In both cases, the predicted critical stretch ratio λcrt increases as μ1/μ2 increases. In particular, if μ2 is fixed, then λcrt increases as μ1 increases. Since the shear modulus at infinitesimal strain is equal to μ=μ1+μ2, we conclude that λcrt increases with the linear shear modulus μ, i.e. as the elastic stiffness of the LCE sample increases. This figure further suggests that, for the same parameter ratio μ1/μ2, if b=0, the auxetic response is still captured, but the stretch deformation at which director rotation occurs is slightly delayed compared to when b0.

Figure 7. (Colour online) Predicted longitudinal critical stretch ratio λcrt for director rotation vs. parameter ratio μ1/μ2 when (Q0,Qcrt,Q1)=(0.5880,0.0010,0.1000) and (b0,bcrt,b1)=(0.0102,0.1518,0.1056) or b=0, respectively.

Figure 7. (Colour online) Predicted longitudinal critical stretch ratio λcrt for director rotation vs. parameter ratio μ1/μ2 when (Q0,Qcrt,Q1)=(0.5880,0.0010,0.1000) and (b0,bcrt,b1)=(−0.0102,−0.1518,−0.1056) or b=0, respectively.

For the case when the director rotates slowly, in , we show the constant uniaxial and biaxial order parameters Q and b, respectively, and the stretch ratios λ1=λ and λ2=λ3=1/λ. The constant anisotropy parameter given by EquationEquation (28) when Q=0.5880 is a=5.2816.

Figure 8. (Colour online) The two vertical lines correspond to the lower and upper bounds [λ,λ]=[1.5900,1.8548] on the extension ratio λ, between which shear striping occurs.

Figure 8. (Colour online) The two vertical lines correspond to the lower and upper bounds [λ∗,λ∗]=[1.5900,1.8548] on the extension ratio λ, between which shear striping occurs.

depicts the strain-energy function w(λ,ε,θ) defined by EquationEquation (38), for (ε,θ)=(0,0), (ε,θ)=(ε0,θ0), and (ε,θ)=(0,π/2). The predicted longitudinal stretch ratio interval where shear striping occurs is [λ,λ]=[1.5900,1.8548]. In , the sine of the director angle θ0 and the shear parameter ε0 are illustrated for various parameter ratios μ(1)/μ(2). These results are in qualitative agreement with those shown in [Citation15,Citation33]. Similar theoretical results can be found, for example, in Refs [Citation21,Citation23,Citation24] (see also [Citation34, Chapter 6]).

Figure 9. (Colour online) The energy function when the director rotates slowly. The two vertical lines correspond to the lower and upper bounds [λ,λ]=[1.5900,1.8548] on the extension ratio λ, between which the second solution, with (ε,θ)=(ε0,θ0), minimises the energy.

Figure 9. (Colour online) The energy function when the director rotates slowly. The two vertical lines correspond to the lower and upper bounds [λ∗,λ∗]=[1.5900,1.8548] on the extension ratio λ, between which the second solution, with (ε,θ)=(ε0,θ0), minimises the energy.

Figure 10. (Colour online) The sine of the director angle θ0 and the shear parameter ε0 during stripe formation in the case when the director rotates slowly for values of μ1/μ2 from 1 to 10 (as labelled).

Figure 10. (Colour online) The sine of the director angle θ0 and the shear parameter ε0 during stripe formation in the case when the director rotates slowly for values of μ1/μ2 from 1 to 10 (as labelled).

4. Conclusion

LCE stretching ultimately causes a reorientation of the nematic director which tends to become parallel to the applied force. Sometimes a change in order also occurs. This correlation between macroscopic deformation and molecular architecture causes a variety of mechanical instabilities not observed in conventional rubber-like solids.

For example, experimental evidence demonstrates that, for LCEs with planar mesogen alignment, when a uniaxial tensile force is imposed perpendicular to the nematic director, the director rotates either slowly giving rise to a shear stripe pattern or almost suddenly causing auxeticity at a characteristic large strain. Experimentally, for auxetic LCEs, biaxiality has also been found to emerge.

The present theoretical study proposes a general mathematical model capable of predicting either the classical shear-striping instability or the auxetic response observed in various LCEs. From the modelling point of view, these two phenomena are captured by letting the uniaxial scalar order parameter decrease continuously to a sufficiently low value then increase again in the case of auxetic LCEs, and by maintaining it constant for LCEs exhibiting shear stripes. The theoretical model further suggests that biaxiality also plays a role in enhancing the auxetic reponse. These results may shed some light on the role of nematic order in the observed mechanical behaviours.

Further experimental investigation where nematic parameters must be varied systematically is required to fully understand the auxetic phenomenon and will be a subject for our future study.

Disclosure statement

No potential conflict of interest was reported by the authors.

Data availability statement

All supporting data for this research are included in the paper.

Additional information

Funding

We are grateful for the support by the Engineering and Physical Sciences Research Council of Great Britain under research grants EP/R020205/1 to AG, EP/S028870/1 to LAM and EP/M009521/1 to HFG. DM thanks the Leverhulme Trust for an Early Career Fellowship.

References

  • Mistry D, Morgan PB, Clamp JH, et al. New insights into the nature of semi-soft elasticity and “mechanical-Fréedericksz transitions” in liquid crystal elastomers. Soft Matter. 2018;14(8):1301–1310.
  • Mistry D, Connell SD, Mickthwaite SL, et al. Coincident molecular auxeticity and negative order parameter in a liquid crystal elastomer. Nat Commun. 2018;9(1):5095.
  • Mistry D, Gleeson HF. Mechanical deformations of a liquid crystal elastomer at director angles between 0° and 90°: deducing an empirical model encompassing anisotropic nonlinearity. J Polym Sci. 2019;57:1367–1377. DOI:10.1002/polb.24879
  • Mistry D, Nikkhou M, Raistrick T, et al. Isotropic liquid crystal elastomers as exceptional photoelastic strain sensors. Macromolecules. 2020;53(10):3709–3718.
  • Raistrick T, Zhang Z, Mistry D, et al. Understanding the physics of the auxetic response in a liquid crystal elastomer. Phys Rev Res. 2021;3(2):023191.
  • Raistrick T, Reynolds M, Gleeson HF, et al. Influence of liquid crystallinity and mechanical deformation on the molecular relaxations of an auxetic liquid crystal elastomer. Molecules. 2021;26(23):7313.
  • Finkelmann H, Kundler I, Terentjev EM, et al. Critical stripe-domain instability of nematic elastomers. J Phys II. 1997;7(8):1059–1069.
  • Kundler I, Finkelmann H. Strain-induced director reorientation in nematic liquid single crystal elastomers. Macromol Rapid Commun. 1995;16(9):679–686.
  • Kundler I, Finkelmann H. Director reorientation via stripe-domains in nematic elastomers: influence of cross-link density, anisotropy of the network and smectic clusters. Macromol Chem Phys. 1998;199(4):677–686.
  • Petelin A, Čopič M. Observation of a soft mode of elastic instability in liquid crystal elastomers. Phys Rev Lett. 2009;103(7):077801.
  • Petelin A, Čopič M. Strain dependence of the nematic fluctuation relaxation in liquid-crystal elastomerss. Phys Rev E. 2010;82(1):011703.
  • Talroze RV, Zubarev ER, Kuptsov SA, et al. Liquid crystal acrylate-based networks: polymer backbone–LC order interaction. React Funct Polym. 1999;41(1–3):1–11.
  • Zubarev ER, Kuptsov SA, Yuranova TI, et al. Monodomain liquid crystalline networks: reorientation mechanism from uniform to stripe domains. Liq Cryst. 1999;26(10):1531–1540.
  • Mitchell GR, Davis FJ, Guo W. Strain-induced transition in liquid-crystal elastomers. Phys Rev Lett. 1993;71(18):2947.
  • Warner M, Terentjev EM. Liquid Crystal Elastomers. Oxford, UK: Oxford University Press; 2007.
  • Conti S, DeSimone A, Dolzmann G. Soft elastic response of stretched sheets of nematic elastomers: a numerical study. J Mech Phys Solids. 2002;50(7):1431–1451.
  • DeSimone A, Dolzmann G. Material instabilities in nematic elastomers. Physica D. 2000;136(1–2):175–191.
  • DeSimone A, Teresi L. Elastic energies for nematic elastomers. Eur Phys J E. 2009;29(2):191–204.
  • Fried E, Sellers S. Free-energy density functions for nematic elastomers. J Mech Phys Solids. 2004;52(7):1671–1689.
  • Fried E, Sellers S. Orientational order and finite strain in nematic elastomers. J Chem Phys. 2005;123(4):043521.
  • Fried E, Sellers S. Soft elasticity is not necessary for striping in nematic elastomers. J Appl Phys. 2006;100(4):043521.
  • Goriely A, Mihai LA. Liquid crystal elastomers wrinkling. Nonlinearity. 2021;34(8):5599–5629.
  • Mihai LA, Goriely A. Likely striping in stochastic nematic elastomers. Math Mech Solids. 2020;25(10):1851–1872.
  • Mihai LA, Goriely A. Instabilities in liquid crystal elastomers. MRS Bull. 2021;46(9):784–794.
  • Mihai LA, Mistry D, Raistrick T, et al. A mathematical model for the auxetic response of liquid crystal elastomers. Philos Trans R Soc A. 2022;380(2234):20210326.
  • Ogden RW. Large deformation isotropic elasticity - on the correlation of theory and experiment for incompressible rubberlike solids. Proc R Soc Lond A. 1972;326:565–584.
  • Ogden RW. Non-linear elastic deformations. New York (NY): Dover; 1997.
  • Mistry D. The richness of liquid crystal elastomer mechanics keeps growing. Liq Cryst Today. 2021;30(4):59–66.
  • Finkelmann H, Greve A, Warner M. The elastic anisotropy of nematic elastomers. Eur Phys J E. 2001;5(3):281–293.
  • de Gennes PG, Prost J. The physics of liquid crystals. Oxford: Clarendon Press; 1993.
  • Urayama K, Mashita R, Kobayashi I, et al. Stretching-induced director rotation in thin films of liquid crystal elastomers with homeotropic alignment. Macromolecules. 2007;40(21):7665–7670.
  • Higaki H, Takigawa T, Urayama K. Nonuniform and uniform deformations of stretched nematic elastomers. Macromolecules. 2013;46(13):5223–5231.
  • Verwey GC, Warner M, Terentjev EM. Elastic instability and stripe domains in liquid crystalline elastomers. J Phys. 1996;6(9):1273–1290.
  • Mihai LA. Stochastic elasticity: a nondeterministic approach to the nonlinear field theory. Switzerland: Springer; 2022. doi:10.1007/978-3-031-06692-4.