7
Eur. Phys. J. E 30, 371-377 (2009) DOI: 10.1140/epje/i2009-10535-2 Response of prestretched nematic elastomers to external fields Andreas M. Menzel 1, a , Harald Pleiner 2, b , and Helmut R. Brand 1,2, c 1 Theoretische Physik III, Universit¨ at Bayreuth, 95440 Bayreuth, Germany 2 Max Planck Institute for Polymer Research, P.O. Box 3148, 55021 Mainz, Germany Received 21 July 2009 and Received in final form 20 October 2009 Published online: 5 December 2009 - c EDP Sciences / Societ` a Italiana di Fisica / Springer-Verlag 2009 Abstract. We investigate the response of prestretched nematic side-chain liquid single crystal elastomers to superimposed external shear, electric, and magnetic fields of small amplitude. The prestretching direction is oriented perpendicular to the initial nematic director orientation, which enforces director reorientation. Furthermore, the shear plane contains the direction of prestretch. In this case, we obtain a strongly de- creased effective shear modulus in the vicinity of the onset and the completion of the enforced director rotation. For the same regions, we find that it becomes comparatively easy to reorient the director by external electric and magnetic fields. These results were derived using conventional elasticity theory and its coupling to relative director-network rotations. PACS. 61.30.Vx Polymer liquid crystals – 61.30.Dk Continuum models and theories of liquid crystal structure – 61.30.Gd Orientational order of liquid crystals; electric and magnetic field effects on order 1 Introduction Side-chain liquid single crystal elastomers (SCLSCEs) con- sist of crosslinked polymer backbones, to which mesogenic molecules are attached as side-groups via flexible spacer units [1]. Information about the state of liquid crystalline order at the time of crosslinking is imprinted into the ma- terials [2,3]. In this way, through specific routes of synthe- sis, liquid crystalline order in a macroscopic monodomain is induced in SCLSCEs [4–6]. Since the first synthesis of nematic SCLSCEs was re- ported [4], many experiments have been performed to study their behavior in external fields. They revealed a unique material-specific property: a coupling of mechan- ical strain deformations to liquid crystalline order. For example, strain deformations of nematic SCLSCEs can induce reorientations of the director [4, 5, 7–9], and, vice versa, reorientations of the director can lead to elastic dis- tortions [3,6,10,11]. Experiments performed include inves- tigations on the stress-strain behavior [4,5,7–9], piezorheo- metric measurements of the linear shear response [12– 14], as well as investigations of the electro-optical [6] and electro-mechanical [3, 10, 11] properties. In the latter case, the response to external electric fields has been studied mainly for nematic SCLSCEs swollen with low molecular weight liquid crystals. As a recent development in this area, the materials are exposed to an external field of large amplitude to drive a e-mail: [email protected] b e-mail: [email protected] c e-mail: [email protected] them into the nonlinear regime. Then, a second exter- nal field is superimposed. The small-amplitude response to this second external field is recorded under the influ- ence of the first, large-amplitude external field. Current experimental work includes small-amplitude shear mea- surements [15] as well as dynamic light scattering mea- surements [16] on prestretched nematic SCLSCEs. In the latter case, essentially the response to externally imposed electromagnetic fields is recorded. 2 Geometries under investigation Here, we investigate two geometries of this kind (fig. 1). In both cases, a nematic SCLSCE is considerably pre- stretched by an external strain field of amplitude A. The latter is applied in a direction ˆ z perpendicular to the initial director orientation ˆ n 0 (kˆ x). Simultaneously, the sample contracts laterally by amplitudes B (kˆ x) and C (kˆ y). To respect an approximately constant volume of the sample, C is expressed by A and B [17]. We noted above the material-specific coupling of me- chanical deformations and director orientation. As a con- sequence, above a certain threshold strain A c , the increas- ing stretching field A leads to a continuous reorientation of the director ˆ n towards the stretching direction [4, 5, 7, 8]. We introduce the angle of director reorientation ϑ = ](ˆ n 0 , ˆ n). This process of director reorientation is con- nected to a decrease in the slope of the corresponding stress-strain curve [4, 5, 9]. In addition, the reorientation process induces shear deformations of amplitude S in the

Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

Eur. Phys. J. E 30, 371-377 (2009) DOI: 10.1140/epje/i2009-10535-2

Response of prestretched nematic elastomers to external fields

Andreas M. Menzel1,a, Harald Pleiner2,b, and Helmut R. Brand1,2,c

1 Theoretische Physik III, Universitat Bayreuth, 95440 Bayreuth, Germany2 Max Planck Institute for Polymer Research, P.O. Box 3148, 55021 Mainz, Germany

Received 21 July 2009 and Received in final form 20 October 2009Published online: 5 December 2009 - c©EDP Sciences / Societa Italiana di Fisica / Springer-Verlag 2009

Abstract. We investigate the response of prestretched nematic side-chain liquid single crystal elastomers tosuperimposed external shear, electric, and magnetic fields of small amplitude. The prestretching directionis oriented perpendicular to the initial nematic director orientation, which enforces director reorientation.Furthermore, the shear plane contains the direction of prestretch. In this case, we obtain a strongly de-creased effective shear modulus in the vicinity of the onset and the completion of the enforced directorrotation. For the same regions, we find that it becomes comparatively easy to reorient the director byexternal electric and magnetic fields. These results were derived using conventional elasticity theory andits coupling to relative director-network rotations.

PACS. 61.30.Vx Polymer liquid crystals – 61.30.Dk Continuum models and theories of liquid crystalstructure – 61.30.Gd Orientational order of liquid crystals; electric and magnetic field effects on order

1 Introduction

Side-chain liquid single crystal elastomers (SCLSCEs) con-sist of crosslinked polymer backbones, to which mesogenicmolecules are attached as side-groups via flexible spacerunits [1]. Information about the state of liquid crystallineorder at the time of crosslinking is imprinted into the ma-terials [2,3]. In this way, through specific routes of synthe-sis, liquid crystalline order in a macroscopic monodomainis induced in SCLSCEs [4–6].

Since the first synthesis of nematic SCLSCEs was re-ported [4], many experiments have been performed tostudy their behavior in external fields. They revealed aunique material-specific property: a coupling of mechan-ical strain deformations to liquid crystalline order. Forexample, strain deformations of nematic SCLSCEs caninduce reorientations of the director [4, 5, 7–9], and, viceversa, reorientations of the director can lead to elastic dis-tortions [3,6,10,11]. Experiments performed include inves-tigations on the stress-strain behavior [4,5,7–9], piezorheo-metric measurements of the linear shear response [12–14], as well as investigations of the electro-optical [6] andelectro-mechanical [3,10,11] properties. In the latter case,the response to external electric fields has been studiedmainly for nematic SCLSCEs swollen with low molecularweight liquid crystals.

As a recent development in this area, the materials areexposed to an external field of large amplitude to drive

a e-mail: [email protected] e-mail: [email protected] e-mail: [email protected]

them into the nonlinear regime. Then, a second exter-nal field is superimposed. The small-amplitude responseto this second external field is recorded under the influ-ence of the first, large-amplitude external field. Currentexperimental work includes small-amplitude shear mea-surements [15] as well as dynamic light scattering mea-surements [16] on prestretched nematic SCLSCEs. In thelatter case, essentially the response to externally imposedelectromagnetic fields is recorded.

2 Geometries under investigation

Here, we investigate two geometries of this kind (fig. 1).In both cases, a nematic SCLSCE is considerably pre-stretched by an external strain field of amplitude A. Thelatter is applied in a direction z perpendicular to the initialdirector orientation n0 (‖x). Simultaneously, the samplecontracts laterally by amplitudes B (‖x) and C (‖y). Torespect an approximately constant volume of the sample,C is expressed by A and B [17].

We noted above the material-specific coupling of me-chanical deformations and director orientation. As a con-sequence, above a certain threshold strain Ac, the increas-ing stretching field A leads to a continuous reorientationof the director n towards the stretching direction [4, 5, 7,8]. We introduce the angle of director reorientation ϑ =](n0, n). This process of director reorientation is con-nected to a decrease in the slope of the correspondingstress-strain curve [4, 5, 9]. In addition, the reorientationprocess induces shear deformations of amplitude S in the

Page 2: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

2 Andreas M. Menzel et al.: Response of prestretched nematic elastomers to external fields

(a)

S

n0

z

x

y

ABδS

(b)

S

n0

z

x

y

E‖x

E‖z

AB

Fig. 1. Geometries of the set-ups under investigation. Here,A, B, and S denote the amplitudes of the corresponding elas-tic deformations. In the first case (a), an additional externalshear deformation of small amplitude δS is imposed onto theprestretched material. In the second case (b), an additional ex-ternal electric field E is applied either in x- or in z-direction.

plane of director rotation. We refer to our previous investi-gation on this kind of set-up [17] for a detailed discussion.

Here, we study the influence of an additional externalfield superimposed onto the external strain field A (fig. 1).First, we consider the effect of an additional shear defor-mation of small amplitude δS (fig. 1 (a)). The importantpoint is that the corresponding shear plane coincides withthe plane of director reorientation. In other words, δS actsto increase or decrease S. Secondly (fig. 1 (b)), an externalelectric field of magnitude E is applied either parallel ton0 (‖x), or along the stretching direction (‖z).

3 Macroscopic description

To investigate the respective set-up, we use the nonlinearcontinuum model introduced in ref. [17]. This model isbased on the variables of relative rotations between thedirector n and the network of crosslinked polymer chains[18,19]. In the nonlinear regime, two variables Ω and Ωnw

are necessary to characterize these relative rotations [17].(More precisely, we describe the relative rotations betweenn and a second preferred direction nnw imprinted intothe elastic network of crosslinked polymer chains duringthe synthesis [17].) Elastic deformations are described byclassical elasticity theory using the Euler strain tensor ε(compare, for example, ref. [20]). Here, εij = [∂iuj+∂jui−(∂iuk)(∂juk)]/2, u being the displacement field.

As an expression for the generalized energy density F ,we use [17]

F = c1 εijεij +1

2c2 εii εjj

+1

2D1 ΩiΩi +D

(2)1 (ΩiΩi)

2 +D(3)1 (ΩiΩi)

3

+D2 niεijΩj +Dnw2 nnwi εijΩ

nwj

+D(2)2 niεijεjkΩk +D

nw,(2)2 nnwi εijεjkΩ

nwk

− 1

2εa (niEi)

2. (1)

Numerical values of the material parameters c1, D1, D(2)1 ,

D(3)1 , D2, Dnw

2 , D(2)2 , and D

nw,(2)2 were determined in

ref. [17] for a specific homeotropic sample, and we willuse these values for our numerical calculations. The lastterm accounts for the dielectric effects via the dielectricanisotropy εa [21].

In expression (1) we only included nonlinearities con-nected with relative rotations. This procedure explicitlyexcludes intrinsic nonlinear behavior of the network ofcrosslinked polymer backbones in the absence of directorreorientations and relative rotations (there are no termsof cubic or higher order in ε).

4 Shear deformation of prestretched nematicSCLSCEs

We now determine numerically the response of aprestretched nematic SCLSCE to an additional shear ofsmall amplitude δS. For this purpose, we set E = 0 ineq. (1).

According to the procedure of corresponding exper-iments, we perform the calculation in two steps. First,δS = 0 and the stretching amplitude A is increased step-wise. For every value of A, we minimize F in eq. (1). Thatis, we solve the system of equations ∂F/∂ϑ = ∂F/∂B =∂F/∂S = 0. In this way, we determine ϑ, B, and S asa function of A without an additional external shear δSapplied. We include the calculated values for ϑ with in-creasing prestretching amplitude A in fig. 2. The resultscoincide with those presented in ref. [17].

Then, for each value of A, the corresponding values ofB and S are kept fixed during a second step. We vary, foreach value of A, the small amplitude δS of the additionalexternally imposed shear. For each pair (δS,A), we mini-mize F w.r.t. ϑ only, solving ∂F/∂ϑ = 0. Experimentally,this procedure corresponds to confining the prestretchedsample between the plates of a shear rheometer [15]. Slip-page along the plate surfaces is inhibited while the exter-nal shear is imposed.

Inserting the results into expression (1), we obtainF (δS,A). For each value of A, we find ∂F/∂(δS)|δS=0 =0, and we can calculate an effective shear modulus forthe prestretched sample, ∂2F/∂(δS)2|δS=0. This effectivemodulus as a function of the prestretching amplitude A isplotted in fig. 3.

Page 3: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

Andreas M. Menzel et al.: Response of prestretched nematic elastomers to external fields 3

0

0.1

0.2

0.3

0.4

0.5

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

numerical data

A

ϑ/π

[rad]

Fig. 2. Angle of director reorientation ϑ as a function of theprestretching amplitude A, without applying an additional ex-ternal field. In this case, the system is prestretched in a direc-tion perfectly perpendicular to the initial director orientationn0.

0

10

20

30

40

50

60

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

numerical data

A

∂2 F/∂

(δS)2| δS

=0[kPa]

Fig. 3. Effective shear modulus ∂2F/∂(δS)2|δS=0 as a func-tion of the prestretching amplitude A. Here, the system is pre-stretched in a direction perfectly perpendicular to the initialdirector orientation n0.

At two different stretching amplitudes, the effectivemodulus tends to zero in fig. 3. These points correspondto the starting point and endpoint of the director reorien-tation, respectively. If we do not start with a shear am-plitude S equilibrated for a certain prestretch A, e.g. bysetting S = 0, the system exerts shear forces. This shiftsand modifies the effective modulus in fig. 3. In an experi-ment (e.g. [15]), a similar situation may for example ariseif the sample is not completely detached from the corre-sponding shear plates, while the prestretching amplitudeA is being changed.

As a technical check of correctness, we verified theinvariance of our numerical results under the symmetrytransformation (ϑ, δS)→ (−ϑ,−δS).

0

0.5

1

1.5

2

2.5

3

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

numerical data

A

(1/ǫ

0)(∂2ϑ/∂

E2)| E

=0[1/kPa]

Fig. 4. Reorientability ∂2ϑ/∂E2|E=0 as a function of the pre-stretching amplitude A, where E‖z and ϑ > 0. In this example,we set εa/ε0 = 2. The system is prestretched in a direction per-fectly perpendicular to the initial director orientation n0.

5 Electric and magnetic field induced directorreorientation in prestretched nematicSCLSCEs

Although we only consider electric fields E in the follow-ing, the reasoning is the same for external magnetic fieldsH. In the latter case, we must make the replacements“E → H” and “εa → χa”, where χa is the diamagneticanisotropy [21].

We now perform a one-step minimization of F . That is,for each given pair (E, A), we solve the system of equations∂F/∂ϑ = ∂F/∂B = ∂F/∂S = 0. Thus we obtain F (E, A).

For each value of A, we find ∂ϑ/∂E|E=0 = 0. There-fore, we take the second derivative ∂2ϑ/∂E2|E=0 to mea-sure the reorientability of the director n for a given stretch-ing amplitude A. An example of ∂2ϑ/∂E2|E=0 as a func-tion of A for E‖z and ϑ > 0 is plotted in fig. 4. We ob-tain the same curve for a reorientation ϑ < 0, with thesign of ∂2ϑ/∂E2|E=0 reversed. Furthermore, E‖x resultsin identical curves, with reversed signs of ∂2ϑ/∂E2|E=0,respectively.

Strikingly, in fig. 4 we find apparent divergences of thereorientability at the critical values of A. The first appar-ent divergence takes place at A = Ac, where the directorreorientation starts (ϑ = 0+). The second one coincideswith the completion of director reorientation (ϑ = π−/2).

We show a fit of a curve ∝ (A − Ac)x to the re-

gion ϑ & 0+ in fig. 5. The best result was obtained forx = −0.474 and Ac = 0.1031, which reproduces the curvein an interval of remarkable width. An exponent x = −0.5still leads to considerable agreement. In order to better un-derstand this behavior, we can reproduce it from expres-sion (1) in a more analytical approach. For this purpose,we solve the equations ∂F/∂B = ∂F/∂S = 0 w.r.t. B andS and introduce the resulting expressions into ∂F/∂ϑ = 0.In the case that either E‖x or E‖z, we find an equation

Page 4: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

4 Andreas M. Menzel et al.: Response of prestretched nematic elastomers to external fields

0

0.5

1

1.5

2

2.5

3

0.1 0.11 0.12 0.13 0.14 0.15 0.16

numerical data

fitting curve: x = -0.474

analytical result: x = -0.5

A

(1/ǫ

0)(∂2ϑ/∂

E2)| E

=0[1/kPa]

Fig. 5. Reorientability ∂2ϑ/∂E2|E=0 fitted in the region ϑ &0+ by a curve ∝ (A − Ac)

x. Again, εa/ε0 = 2 and the pre-stretching direction is oriented perfectly perpendicular to n0.

for ϑ of the form

0 = ϑa(A− Ac) + gϑ2

+O(ϑ5). (2)

Here, higher orders of A have been neglected. a, Ac, andg are combinations of the material parameters containedin expression (1). Besides, Ac and g are functions of E2

x−E2z , where we call Ac = Ac(E = 0). For E = 0, this

leads to the very simple picture of a strain induced forwardbifurcation for the reorientation angle ϑ. The branches ofthe bifurcation are described by ϑ = 0 and ϑ ∝ ±(A −Ac)

0.5.

As a further result, we obtain ∂2ϑ/∂E2|E=0 ∝ ±(A−Ac)

−0.5, where an additional term ∝ (A− Ac)0.5 was ne-glected. This behavior is reflected by the fitting curves infig. 5. Furthermore, we find ∂2ϑ/∂E2

x|E=0 =−∂2ϑ/∂E2

z |E=0, in accordance with our numerical results.

We performed the same fitting procedure also for theregion ϑ . π−/2. There, we found a behavior character-ized by an exponent x = −0.449. The corresponding fit-ting curve represents the numerical data points in a con-siderably smaller interval. Besides, the width of this in-terval becomes rather limited for an exponent x = −0.5.This reflects the fact that the approximation of neglectinghigher order terms in A(≈ 0.3) becomes less reliable forϑ . π−/2. Therefore, the range of validity of an equationanalogous to eq. (2) is more limited in the region ϑ . π−/2and higher order terms in A would have to be included fora better fit.

The reorientability of the director ∂2ϑ/∂E2|E=0 di-verges at the same points where the effective shear mod-ulus ∂2F/∂(δS)2|δS=0 tends to zero. This agrees with thefact that shear deformations and director reorientationsare linearly related in nematic SCLSCEs [13,22] (see sect. 7for further discussion).

0

0.1

0.2

0.3

0.4

0.5

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

numerical data

A

ϑ/π

[rad]

Fig. 6. Angle of director reorientation ϑ as a function of theprestretching amplitude A, without applying an additional ex-ternal field. Here, the initial director orientation n0 slightlydeviates from the perfectly perpendicular orientation ‖x by anangle of 0.01 rad (0.57).

6 Oblique stretching

In the previous sections we considered the nematicSCLSCE to be stretched exactly perpendicularly to theinitial director orientation n0. However, slight deviationsfrom this condition already change dramatically the cor-responding results. As an example, we investigate a verysmall deviation of the initial director orientation n0 fromthe perfectly perpendicular orientation ‖x by an angle ofonly 0.01 rad (0.57). For this case, the angle of directorreorientation ϑ is plotted as a function of A in fig. 6.

Figs. 7 and 8 show the corresponding effective shearmodulus ∂2F/∂(δS)2|δS=0 and the director reorientability∂2ϑ/∂E2|E=0, respectively. Both quantities now remainnonzero and finite. This scenario follows from an imper-fect bifurcation for the angle of director reorientation inthe oblique case (fig. 6). In contrast, a perfect forward bi-furcation underlies the results of figs. 2, 3, and 4 (note thedifference in scales between the ordinates of figs. 4 and 8;see also [17]). An experimental curve similar to the curvein fig. 7 has been presented before [16], where a slightlyoblique stretching was imposed for intrinsic technical rea-sons.

7 Discussion

Above, for the case of exactly perpendicular stretching, weobserved that the effective shear modulus ∂2F/∂(δS)2|δS=0

tends to zero at the same points where the reorientabilityof the director ∂2ϑ/∂E2|E=0 diverges. The physical back-ground of this scenario is connected with the bifurcationbehavior of the system at the respective stretching ampli-tudes. At A = Ac, the initial director orientation (ϑ = 0)becomes unstable w.r.t. a solution of finite director re-orientation 0 < |ϑ| < π/2 (see fig. 2). Likewise, when westart at high stretching amplitudes A and then decrease A,

Page 5: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

Andreas M. Menzel et al.: Response of prestretched nematic elastomers to external fields 5

0

10

20

30

40

50

60

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

numerical data

A

∂2 F/∂

(δS)2| δS

=0[kPa]

Fig. 7. Effective shear modulus ∂2F/∂(δS)2|δS=0 as a func-tion of the prestretching amplitude A. The initial director ori-entation n0 slightly deviates from the perfectly perpendicularorientation ‖x by an angle of 0.01 rad (0.57).

0

0.05

0.1

0.15

0.2

0.25

0.3

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

numerical data

A

(1/ǫ

0)(∂2ϑ/∂

E2)| E

=0[1/kPa]

Fig. 8. Reorientability ∂2ϑ/∂E2|E=0 as a function of the pre-stretching amplitude A (εa/ε0 = 2). Again, the initial directororientation n0 slightly deviates from the perfectly perpendic-ular orientation ‖x by an angle of 0.01 rad (0.57).

the reoriented state |ϑ| = π/2 becomes unstable w.r.t. theintermediate solution 0 < |ϑ| < π/2. This leads to thesecond bifurcation behavior at |ϑ| = π−/2 (see fig. 2).

The reorientation of the director is initiated by thecoupling of the director orientation to the external stretch-ing deformation. This coupling is mediated by relative ro-

tations (the D2-, Dnw2 -, D

(2)2 -, D

nw,(2)2 -terms) in eq. (1).

Due to the process of synthesizing the materials, however,the reorientation of the director n w.r.t. the network ofcrosslinked polymer chains is hindered (the D1-, D

(2)1 -,

D(3)1 -terms in eq. (1)) [2–5, 17]. This competition leads

to the described bifurcation behavior. It also induces theshear deformation S, which in this context acts to reducethe magnitude of energetically expensive director rota-tions w.r.t. the polymer network [17].

In the vicinity of the bifurcation points the directorcan easily be reoriented by an external field. This featurecan be inferred from the diverging slopes of ϑ(A) in fig. 2.It is further reflected by the diverging curves in figs. 4 and5, where the external electric field is used to probe the re-orientability. Since director reorientations couple linearlyto shear deformations in nematic SCLSCEs [13, 22], theeffective shear modulus in fig. 3 decreases to zero whenthe director reorientability diverges. In contrast, the slopeof ϑ(A) remains finite in the remaining regime, distantfrom the bifurcation points (see fig. 2). This correspond-ingly implies a finite reorientability of the director and anonzero effective shear modulus in this regime.

Next, we note that the shape of the curve for the ef-fective shear modulus ∂2F/∂(δS)2|δS=0 in fig. 3 has al-ready been predicted before [23] on the basis of a dif-ferent model [24]. Furthermore, a similar curve has beenobtained using a semi-microscopic model based on rubberelasticity [25]. Very often, the nonlinear behavior of ne-matic SCLSCEs, i.e. a plateau-like stress-strain relationat finite prestretch with a vanishing effective elastic mod-ulus at the beginning and end of the plateau, is called“semi-softness” [24, 26]. It refers to the notion of “idealsoftness”, where the stress-strain plateau is of zero heightand starts at zero prestretch with the effective elastic mod-ulus being exactly zero everywhere along the plateau (andeven in the linear regime at zero prestretch). This idealNambu-Goldstone scenario, however, does not take placein real nematic SCLSCEs. The smallness of Ac ≈ 0.1, theactual prestretch amplitude for the onset of the plateau,is then used to view the real behavior as almost ideal or“semisoft”.

We have demonstrated that corresponding results arealso derived in a general nonlinear elastic framework. Thelatter includes nonlinear relative rotations without any as-sumption or reference to a small parameter in terms of thelinear elastic modulus nor in terms of Ac. In this frame-work, a similar scenario (plateau-like stress-strain relationat finite prestretch with a vanishing effective elastic shearmodulus at the beginning and end of the plateau) is ob-tained even if Ac approaches unity. An example of thiskind is displayed in fig. 9, where for demonstration wechose different numerical values for the material param-

eters (c1 = 121.0 kPa, D1 = 30.0 kPa, D(2)1 = 3.5 kPa,

D(3)1 = 2.9 kPa, D2 = 0.0 kPa, Dnw

2 = 0.0 kPa, D(2)2 =

−55.5 kPa, and D(2),nw2 = −32.0 kPa). We note that

A = 1 corresponds to an infinite stretching deformation.Therefore we view the pronounced elastic properties ofreal nematic SCLSCEs as a manifestation of the profoundlynonlinear elastic properties of the materials, rather thandue to the vicinity of a hypothetical ideal state.

For a more detailed comparison with other approaches,we note the following. Concentrating only on the quadraticterms in the generalized energy density, we obtain as acondition of thermodynamic stability that 4c1D1− (D2 +Dnw

2 )2 > 0 [17]. As explained in ref. [26], the transi-tion from “semi-softness” to “ideal softness” in the linearregime corresponds to the case where the latter expressionvanishes, 4c1D1−(D2+Dnw

2 )2 → 0. On the one hand, this

Page 6: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

6 Andreas M. Menzel et al.: Response of prestretched nematic elastomers to external fields

0

50

100

150

200

250

300

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

numerical data:

numerical data:

A

dF/dA,∂

2 F/∂

(δS)2

| δS=0[kPa]

dF/dA|δS=0

∂2F/∂(δS)2|δS=0

Fig. 9. Example of a stress-strain curve dF/dA|δS=0(A), cal-culated by our model for the geometry depicted in fig. 1(a) fora different set of material parameters. In this case, the processof director reorientation and the corresponding decrease in theslope of the stress-strain curve occur at comparatively largeprestretching amplitudes A. Again, the effective elastic shearmodulus ∂2F/∂(δS)2|δS=0 tends to zero in the vicinity of thepoints where the director reorientation starts and ends.

may occur by D1 → 0, D2 → 0, and Dnw2 → 0. This, how-

ever, implies that relative rotations have no meaning, or,in other words, that the director has no memory of its ini-tial orientation. Definitely, this is not true for real nematicSCLSCEs. On the other hand, the condition may be sat-isfied by 4c1D1 = (D2 +Dnw

2 )2 without D1 vanishing. Asa result, one recovers in the linear regime the vanishingshear modulus predicted by the theory of “soft elastic-ity” when the shear plane contains the director [26, 27].However, the statement becomes problematic when defor-mations and reorientations are not correspondingly small(compare also ref. [28]).

In our approach, a scenario of “softness” can be recov-ered in the following way for the complete reorientationprocess. One neglects the memory of the initial directororientation. Therefore the variables of relative rotation be-come meaningless and one sets D1 = D

(2)1 = D

(3)1 = D2 =

Dnw2 = D

(2)2 = D

nw,(2)2 = 0. A supplement to the elastic

part of the energy leads to F = c1εijεij + c2εiiεjj/2 +c5niεijnj . If we choose c5 < 0, we recover for our geome-try the above-mentioned scenario of “ideal softness”: zeroheight of the stress-strain plateau during the reorientationprocess, starting at zero prestretch. Here, the additionalterm c5niεijnj acts as an intrinsic force of deformationthat arises from the nematic state of the material (com-pare refs. [28, 29] for related approaches). Real nematicSCLSCEs, however, do not show the scenario of “idealsoftness”, therefore we did not include the correspondingterm with the coefficient c5.

We would like to include three further remarks. First,we have demonstrated in ref. [17] that the materials showan additional intrinsic nonlinear stress-strain behavior dueto nonlinear elasticity. This kind of nonlinear behavior is

not connected to director reorientations. It arises simplyfrom straining the crosslinked polymer network. Never-theless, it influences significantly the overall stress-strainbehavior of the elastomers [17]. These nonlinear proper-ties, which we have neglected above, could force the ef-fective shear modulus ∂2F/∂(δS)2|δS=0 to remain at afinite, nonzero level. Secondly, the finiteness of real sam-ples renders the lateral boundaries to move and changeshape during stretching. That leads to deviations fromthe bulk director orientation close to the boundaries. Theconditions for the effective shear modulus to vanish mighttherefore not be met close to the boundaries. Furthermore,within the sample, the critical stretching amplitude Acmay vary on a macroscopic length scale due to spatialheterogeneities. If, in an experiment, one simultaneouslyprobes such macroscopic regions of different Ac, the ob-served effective shear modulus will not vanish.

Our results are interesting from an experimental andan applied point of view. Since director reorientations innematic elastomers are connected to elastic deformations,they are candidates for the potential use as actuators[6,10,30–32]. So far, only swollen nematic elastomers showa considerable director reorientation at reasonably low ex-ternal electric field amplitudes [3,6,10,11]. However, fine-tuned prestretching of nematic elastomers may allow aconsiderable electric field induced director reorientationand resulting elastic deformations also for dry materials.

The authors thank Philippe Martinoty and Kenji Urayama forstimulating discussions. A.M.M. and H.R.B. thank theDeutsche Forschungsgemeinschaft through the ForschergruppeFOR608 ‘Nichtlineare Dynamik komplexer Kontinua’ and theDeutscher Akademischer Austauschdienst (312/pro-ms)through PROCOPE for partial support of this work. A.M.M.acknowledges partial support by the ENB ‘MacromolecularScience’.

References

1. H. Finkelmann, H.-J. Kock, G. Rehage, Makromol. Chem.,Rapid Commun. 2, 317 (1981).

2. H.R. Brand, K. Kawasaki, Macromol. Rapid Commun. 15,251 (1994).

3. D.-U. Cho, Y. Yusuf, P.E. Cladis, H.R. Brand, H. Finkel-mann, S. Kai, Jap. J. Appl. Phys. 46, 1106 (2007).

4. J. Kupfer, H. Finkelmann, Makromol. Chem., Rapid Com-mun. 12, 717 (1991).

5. J. Kupfer, H. Finkelmann, Macromol. Chem. Phys. 195,1353 (1994).

6. K. Urayama, S. Honda, T. Takigawa, Macromol. 38, 3574(2005).

7. I. Kundler, H. Finkelmann, Macromol. Rapid Com-mun. 16, 679 (1995).

8. I. Kundler, H. Finkelmann, Macromol. Chem. Phys. 199,677 (1998).

9. K. Urayama, R. Mashita, I. Kobayashi, T. Takigawa,Macromol. 40, 7665 (2007).

10. Y. Yusuf, J.H. Huh, P.E. Cladis, H.R. Brand, H. Finkel-mann, S. Kai, Phys. Rev. E 71, 061702 (2005).

Page 7: Response of prestretched nematic elastomers to external eldspleiner/papers/prestretchedSCLS… · Response of prestretched nematic elastomers to external elds Andreas M. Menzel1 ;a,

Andreas M. Menzel et al.: Response of prestretched nematic elastomers to external fields 7

11. K. Urayama, S. Honda, T. Takigawa, Macromol. 39, 1943(2006).

12. P. Stein, N. Aßfalg, H. Finkelmann, P. Martinoty,Eur. Phys. J. E 4, 255 (2001).

13. P. Martinoty, P. Stein, H. Finkelmann, H. Pleiner,H.R. Brand, Eur. Phys. J. E 14, 311 (2004).

14. D. Rogez, G. Francius, H. Finkelmann, P. Martinoty,Eur. Phys. J. E 20, 369 (2006).

15. P. Martinoty, presentation at the International LiquidCrystal Elastomer Conference, Ljubljana, 2007.

16. A. Petelin, M. Copic, presentations at the European Con-ference on Liquid Crystals, Colmar, 2009.

17. A.M. Menzel, H. Pleiner, H.R. Brand, J. Appl. Phys. 105,013503 (2009).

18. P.G. de Gennes, in Liquid Crystals of One- and Two-Dimensional Order, edited by W. Helfrich, G. Heppke(Springer, Berlin, 1980), pp. 231 ff.

19. H.R. Brand, H. Pleiner, Physica A 208, 359 (1994).20. H. Temmen, H. Pleiner, M. Liu, H.R. Brand, Phys.

Rev. Lett. 84, 3228 (2000).21. P.G. de Gennes, J. Prost, The Physics of Liquid Crystals

(Clarendon Press, Oxford, 1993).

22. A.M. Menzel, H. Pleiner, H.R. Brand, J. Chem. Phys. 126,234901 (2007).

23. T.C. Lubensky, remarks at the International Liquid Crys-tal Elastomer Conference, Ljubljana, 2007.

24. F.F. Ye, T.C. Lubensky, J. Phys. Chem. B 113, 3853(2009).

25. J.S. Biggins, E.M. Terentjev, M. Warner, Phys. Rev. E 78,041704 (2008).

26. M. Warner, E.M. Terentjev, Liquid Crystal Elastomers(Clarendon Press, Oxford, 2003); and references therein.

27. L. Golubovic, T.C. Lubensky, Phys. Rev. Lett. 63, 1082(1989).

28. T.C. Lubensky, R. Mukhopadhyay, L. Radzihovsky,X. Xing, Phys. Rev. E 66, 011702 (2002).

29. A. Fukunaga, K. Urayama, T. Takigawa, A. DeSimone,L. Teresi, Macromol. 41, 9389 (2008).

30. R. Zentel, Liq. Cryst. 1, 589 (1986).31. Y. Yusuf, Y. Sumisaki, S. Kai, Chem. Phys. Lett. 382, 198

(2003).32. Y. Yu, T. Ikeda, Angew. Chem. Int. Ed. 45, 5416 (2006).