You are on page 1of 17

Vol.

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

251

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


Introduction
Ferroelectric materials are a subclass of pyro- and piezoelectric materials (Fig. 1) (see PIEZOELECTRIC POLYMERS). They are very rarely found in crystalline organic or polymeric materials because ferroelectric hysteresis requires enough molecular mobility to reorient molecular dipoles in space. So semicrystalline poly(vinylidene uoride) (PVDF) is nearly the only known compound (1). On the contrary, ferroelectric behavior is very often observed in chiral liquid crystalline materials, both low molar mass and polymeric. For an overview of ferroelectric liquid crystals, see Reference 2. Tilted smectic liquid crystals that are made from chiral molecules lack the symmetry plane perpendicular to the smectic layer structure (Fig. 2). Therefore, they develop a spontaneous electric polarization, which is oriented perpendicular to the layer normal and perpendicular to the tilt direction. Because of the liquid-like structure inside the smectic layers, the direction of the tilt and thus the polar axis can be easily switched in external electric elds (see Figs. 2 and 3). Here, we discuss materials (LC-elastomers) that combine a liquid crystalline phase and ferroelectric properties (preferably the chiral smectic C phase) in a polymer network structure (see Fig. 4). The coupling of the liquid crystalline director to the network or the softness of the network is chosen so that reorientation of the polar axis is still possible. Thus, densely cross-linked systems that possess a polar axis but cannot be switched (11) will be excluded. It is the role of the network (1) to form a rubbery matrix for the liquid crystalline phase and (2) to stabilize a director conguration. LC-materials that have these properties can be made either (see Fig. 4) by covalently linking the mesogenic groups to a slightly

Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.

252

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

Vol. 6

Piezoelectric Pyroelectric Pyroelectric Ferroelectric Ferroelectric

Ps

Fig. 1. Ferroelectric hysteresis that shows the spontaneous polarization PS of a ferroelectric material reversed by an applied electric eld E.
Electrodes
+

z n

Fig. 2. Schematic drawing of the bistable switching of a ferroelectric liquid crystal in the surface stabilized FLC conguration.

cross-linked rubbery polymer network structure (see Fig. 4a) (46) or by dispersing a phase-separated polymer network structure within a low molar mass liquid crystal (see Fig. 4b) (9,12). Both systems possess locally a very different structure. They may show, however, macroscopically similar properties. LC-elastomers (see Fig. 4a) have been investigated in detail (48). Although the liquid crystalline phase transitions are nearly unaffected by the network, the network retains the memory of the phase and director pattern during cross-linking

Vol. 6
(a)

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


(b)

253

Fig. 3. Network: soft, can be transformed like rubber band, but retains its shape and couples to director orientation because (a) director is preferably parallel (or perpendicular) to polymer chains (LC-elastomer) (49); (b) director aligns (parallel) to chains in oriented phase-separated polymer network structure (low molar mass LC in LC-thermoset) (9,10).

(7,8). In addition, it freezes uctuations of the smectic layers and leads to a real long range order in one dimension (13). An attempt to change the director pattern by electric or magnetic elds in LC-elastomers leads to a deformation of the network and to an elastic response (see Fig. 3). As a consequence of this, nematic LC-elastomers could never be switched in electric elds, if the shape of the elastomer was kept xed. For freely suspended pieces of nematic LC-elastomers, shape variations in electric elds have been observed sometimes (14,15). In ferroelectric liquid crystals, the interaction with the electric eld is, however, much larger. Thus, it has been possible to prepare real ferroelectric LC-elastomers (see Fig. 3) (3,16). In these systems, the polymer network stabilizes one switching state like a soft spring. It is, however, soft enough to allow ferroelectric switching. Therefore the ferroelectric hysteresis can therefore be measured in these systems. It is, however, shifted away from zero voltage (see Fig. 3).

Synthesis of Ferroelectric LC-Elastomers


The ferroelectric LC-elastomers described so far (3,1621) are mostly prepared from cross-linkable ferroelectric polysiloxanes (see Fig. 5), which are prepared by hydrosilylation of precursor polysiloxanes (22) (see SILICONES). The cross-linking is nally initiated by irradiating a photoradical generator, which leads to oligomerization of acrylamide or acrylate groups (see Fig. 5). The functionality of the net points is thus high (equal to the degree of polymerization) and varies with the cross-linking conditions.

254

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


(a)

Vol. 6

(b)

(c) Optical response

Uncross-linked

Cross-linked

80

40 0 40 Voltage, V

80

200 100 0 100 Voltage, V

200

Fig. 4. Schematic drawing of the ferroelectric LC-elastomer and its two switching states (3): (a) A polymer chain acts as cross-linking point by connecting different mesogenic groups attached to the main polymer chains. A ferroelectric switching in this elastomer extends polymer chains. (b) The entropy elasticity arising from this acts like a spring that stabilizes one state. (c) For the uncross-linked system (left), the hysteresis is symmetrical to zero voltage and both states are equal. After cross-linking in one polar state (right), only that state is stable with no electric eld, and the hysteresis is no longer symmetrical to zero voltage.

The advantage of this photochemical-initiated cross-linking is that the crosslinking can be startedat will after the liquid crystalline polymer is oriented and sufciently characterized in the uncross-linked state (see Fig. 6). The advantage of using polymerizable groups (acrylates) for cross-linking is that small amounts of these groups are sufcient to transform a soluble polymer into a polymer gel and that the chemical reaction happens far away from the mesogen. Cinnamoyl moieties, on the other hand (23), require a high concentration of these groups for cross-linking. The dimers thus formed are, in addition, nonmesogenic. Figure 7 summarizes the ferroelectric LC-elastomers discussed in this article. Two different positions of cross-linkable groups are used. In polymer P1, the cross-linking group is close to the siloxane chains, which are known to microphase-separate from the mesogenic groups (22,23). Therefore, the cross-linking should proceed mostly within the siloxane sublayers. In polymers P2 and P3, the cross-linking group is located at the end of mesogens. Therefore, the cross-linking should proceed mostly

Vol. 6

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


H3C Si O H3C Si O
2.7n

255

H
1n

O + (CH2)9 O
C10H12PtCl2 toluene

O C CH3

CH3

O H3C Si O H3C Si O
2.7n

(CH2)11
1n

C CH3

CH3
H2N NH2 THE

H3C

Si O

(CH2)11
1n

OH

H3C

Si O

CH3
O Cl CH CH3 CH C2 H5 H N C O

2.7n

+ HO + HO

C O C

(CH2)5

DCC, DMAP, THF, CH2Cl 2

O H3C Si CH2
0.9n

(CH2)10

O C

Cl CH

CH3 CH C2 H 5

O H3C Si CH2
0.1n

O (CH2)10 O O C (CH2)5

H N C O

O H3C Si CH3
2.7n
O
= 365 nm in smectic C*

OCH 3 OCH 3

LC network

Fig. 5. Synthetic route to the cross-linkable polysiloxane P2 and the following preparation of the oriented smectic C network using uv light in the presence of a photoinitiator (1,1dimethoxy-1-phenyl-acetophenone) (3).

256

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

Vol. 6

ITO

Polydomain

Electrical field

Monodomain

Initiator hv electrical field

Oriented sc-network

Fig. 6. Preparation of polar smectic C monodomains (3,16) (ITO: indium tin oxide).

between different siloxane layers (see Fig. 7). A comparison of these elastomers allows evaluating structureproperty relationships (18,24,25).

Properties and Characterization Ferroelectric Characterization (Uncross-linked Systems). Before cross-linking, polarization, tilt angle, and switching times can be determined in the usual way (18,22,26). Figure 8 shows the temperature dependence of the spontaneous polarization for polymers P1P3. For the homopolymer related to polymer P2, all relevant parameters were determined in a careful study (27). It seems that the electroclinic effect is especially strong in these polysiloxanes (16). This has implications for the freezing of a memory of the tilt angle present during

Vol. 6

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

257

cross-linking. Therefore, ferroelectric elastomers, which have been crosslinked in the smectic A phase while applying an electric eld, produce a stable macroscopic polarization (tilt) after cooling into the smectic C phase (18). Mechanical Properties of Ferroelectric LC-Elastomers. The crosslinking reactions of a series of copolymers analogous to polymer P2, but differing in the amount of cross-linkable groups, were studied by ftir spectroscopy (17). These measurements show a decrease of the acrylamide double bond on irradiation. Conversions between 60 and 84% were observed. The uncertainty of the conversion, however, is high because only very few double bonds are present in polymer P2 and they are visible in the infrared spectrum at rather low intensity. Mechanical measurements which show how this photo-chemical crosslinking (conversion of double bonds) leads to an elastic response of the network are, however, still at the beginning because photocross-linking can be performed only in thin layers of a few micrometers. It is best performed between two glass slides to exclude oxygen. AFM measurements of photocross-linked free standing lms show changes in topology during stretching (24). They, however, do not allow measuring elastic moduli. The most promising approach to obtaining elastic data for these ferroelectric elastomers is investigation of LC-elastomer balloons (25,28). For this purpose, an experimental setup was developed on the basis of an apparatus designed to study smectic bubbles (28). Freely suspended lms of the uncross-linked material behave like ordinary smectic lms. They can be inated to spherical bubbles several millimeters in diameter (the thickness of a smectic-layer skin is about 50 nm). These bubbles are stabilized by the smectic-layer structure and their inner pressure p is related to the surface tension and the bubble radius R by the LaplaceYoung equation, p 1/R. After exposure to uv light, the material is crosslinked, and an anisotropic elastomer is formed. When the cross-linked bubbles are inated/deated, the radiuspressure curve reverses its slope and gives direct access to the elastic moduli of the material (25). Because the deformation during ination of the balloon is isotropic in the layer plane, the material should contract in the direction of the layer normal. Mechanical measurements of chemically crosslinked LC-elastomers have been made extensively (4,5,2934). For these systems, it can be shown that stretching allows orientation of the liquid crystalline phase. In ideal situations, it is thus possible to prepare a ferroelectric monodomain by stretching (30,31,35). This result can be rationalized as a two-stage deformation process (see Fig. 9) (31). This possibility of orienting or reorienting the polar axis mechanically is the basis for the piezoelectric properties to be discussed later. Ferroelectric switching could not be observed for any of the chemically crosslinked systems. This may occur because chemically cross-linked lms are too thick (several 100 m compared to about 10 m for photochemically cross-linked systems) and the electric eld applied is therefore too small. In addition, the cross-linking density in chemically cross-linked systems is presumably higher. Ferroelectric Properties (Cross-linked Systems). The ferroelectric properties of the photochemically crosslinked elastomers E1E3 differ significantly and depend on the topology of the network formed. For the systems that have interlayer cross-linking (see Fig. 7, E2 and E3), the switching time

H3C

O Si O Si O

Cl CH2
0.9 n

(CH2)10

O O

H3C

O Si O Si O

Cl CH2
0.9n

(CH2)10

O O O N H

H3C

CH2
0.1n

CH2

CH2

H3C

CH2
0.1n

(CH2)10

H3C

Si

CH3
2.7n

O P1 Phase transitions [C]: sX 32 sC 60 sA 92 i n = 30

H3C

Si

CH3
2.7n

P2 O Phase transitions [C]: sX 29 sC 53 sA 89 i n = 30

O H3C Si O Si O H3C Si

O CH2
0.85n

NO2 CH3 C* C6H13 H O (CH2)6 O

(CH2)10

C O

H3C

CH2
0.15n

(CH2)10

CH3
1.5n

P3 Phase transitions [C]: sX 32 sC117 sA 152 i n = 15

258
E Photoinitiator

P1

P2, P3

Photoinitiator

E1

E2, E3

Fig. 7. Chemical structure and phase-transition temperatures of polymers P1P3 (18). (a) P1 is designed to favor intralayer cross-linking; (b) P2 and P3 form an interlayer network.

Vol. 6

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


160 140 120

259

Ps, nC/cm2

100 80 60 40 20 0 50 45 40 35 30 25 20 15 10 5 T Tc, C 0 5

Fig. 8. Temperature dependence of the spontaneous polarization PS for the polymers P1P3 measured by the triangular wave method (18). p1 : , p2 : , p3 : .

First deformation

Second deformation

Fig. 9. Two-step deformation process of a chiral smectic C elastomer that displays macroscopic polarization at the end (31).

is increased greatly. Therefore, spontaneous polarization can no longer be determined by the triangular wave method. Slow switching is, however, still possible and therefore ferroelectric hysteresis can be measured optically (see Figs. 3c and 10) (3,16). After photochemical cross-linking in a ferroelectric monodomain, the ferroelectric hysteresis shows stabilization of the orientation present during cross-linking. At zero external voltage, only this state is stable. The second switching state can, however, be reached. Therefore, the network acts like a spring that stabilized one state because switching to the other state leads to a deviation from the most probable conformation of the polymer chain (36) (see Fig. 10). Then, the shift of the center of the hysteretic loop away from zero voltage gives the magnitude of the electric eld necessary to balance the mechanical eld of the network. The asymmetry of the hysteresis increases with the cross-linking density (18). For

260

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


Temperature, C 42 Optical response 44 46 48

Vol. 6

200

100

0 Voltage, V

100

200

Temperature, C 50 Optical response 54 56

200

100

0 Voltage, V

100

200

Fig. 10. Temperature dependence of the optical hysteresis of elastomer E2 (SC 49 C SA ) (15). (a) Ferroelectric behavior of the SC phase (42, 44, 46, and 48 C, respectively). (b) Electroclinic behavior of the SA phase (50, 54, and 56 C, respectively).

high cross-linking densities, switching remains possible only if the spontaneous polarization is rather high (18). Otherwise, the network prohibits switching. The asymmetry of ferroelectric switching could also be proven by polarized ftir spectroscopy (37). Increasing the temperature of this ferroelectric elastomer leads to narrowing of the hysteretic loop, which is lost at the transition to the smectic A phase (see Fig. 10). This behavior is best interpreted by plotting the liquid crystalline potential, the elastic potential of the network, and their superposition in one graph (16) (see Fig. 11). As the network is formed in the smectic C phase, an internal elastic eld is created, which has its minimum value for the tilt angle and tilt direction during cross-linking. Other tilt angles are destabilized. The elastomer that has preferable intralayer cross-linking (E1, see Fig. 7) shows completely different behavior (see Fig. 12) (18,38). In this case, the switching time increases by less than a factor of 2, the polarization can still be determined, and measurement of the ferroelectric hysteresis shows no stabilization of the switching state present during cross-linking. Then, the coupling between the orientation of the mesogens and the network conformation is obviously very

Vol. 6

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


10000

261

8000

6000 g J/m

4000

2000

2000 30 20 10 0 10 Tilt angle , deg 20 30

Fig. 11. Effect of network force on the free energy density (16) 2 K below the phase transition, SC phase: () calculated potential of the SC phase, ( ) force due to the network, and ( ) superposition of both.
50

40

30
, ms

20

10

0 25 20 15 10 TTc C 5 0

Fig. 12. Temperature dependence of the switching time (dened as 0100% change in transmission) for P1 and E1 [see Ref. 18 for comparison].

weak. The network stabilizes the smectic layer structure, but it does not stabilize the tilt direction. Therefore, the polar axis can be switched easily. This is the result of the network topology (see Fig. 7) in which interlayer cross-linking is rare.

262

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

Vol. 6

AFM Imaging of Thin Films. Freestanding lms can be prepared from uncross-linked polymers (see also the smectic balloons in this context). They can be photocross-linked and transferred to a solid substrate. Thereafter, the topology of the lms can be imaged by afm, which gives direct visualization of the smectic layer structure at low temperatures. The uncross-linked polymers can only be imaged at low temperatures, deep inside the smectic phase and in the tapping mode, which does not induce strong lateral forces. At higher temperatures, the sample is too soft and mobile to allow imaging. Cross-linked elastomers, on the other hand, are mechanically stable, and lms sustain the tapping mode and also the contact mode of the atomic force microscope (39). This holds both for intraand interlayer cross-linked systems. Because measurements can be done in all phases, it is also possible to determine the change of the smectic layer thickness at the phase transitions directly. For elastomer E1, for example, the smectic layer thickness is 4.2 nm in the smectic C phase (36 C, tilt angle about 30 ). It increases to 4.4 nm at 50 C in the smectic A phase (39). This corresponds to x-ray measurements. To analyze the impact of the molecular structure on network properties, elastomers are compared, which are identical except for the molecular position of the cross-linkable group: (1) elastomer E1 that has cross-linkable groups attached to the backbone via a short spacer (intralayer cross-linking) and (2) elastomer E2 where the cross-linkable group is in the terminal position of a mesogenic side group (interlayer cross-linking) (24,40). When mechanical stress (stretching) is imposed on thin lms in homeotropic orientation, the two elastomers react differently to the deformation (24,40), as seen by afm imaging of the surface topology (see Fig. 13). For elastomer E1, intralayer cross-linking results in two-dimensional networks in the backbone layers, separated by liquid-like FLC side-group layers. Because there are practically no vertical connections in this intralayer network, no vertical distortions occur. Therefore, this elastomer can be stretched up to 100%, the surface remains smooth, and the layers deform afnely. In elastomer E2, a threedimensional interlayer network is formed; the system reacts by distorting the smectic layering. Therefore, only smaller stretching ratios are accessible and the surface roughens and buckles during stretching. The distortion strength increases with a higher cross-linking density. Piezoelectric Properties of Ferroelectric LC-Elastomers. Because a ferroelectric material has to be piezoelectric (see Fig. 1), observation of a piezoresponse is natural. It has been observed for ferroelectric LC-elastomers (3,16,19 21) and also for more densely cross-linked systems (10,31,41,42) for which no ferroelectric switching could be observed. For the elastomers described here it is, however, possible to change the piezoresponse (3) by reorienting the polar axis in an external eld (see E2 in Fig. 14). For this experiment, the polar axis was kept in one orientation during cross-linking. This resulted in a positive piezoresponse (see Fig. 14). Thereafter, the direction of the polar axis was inverted by applying an external eld of opposite direction. Then, the external eld was removed and the piezocoefcient was measured. At rst, a piezoresponse of opposite sign (negative) but identical value is determined (see Fig. 14). In the eld-free state, this piezoresponse continuously decreases, it goes through zero, increases again, and nally reaches the original positive value. This experiment is comparable to the hysteresis measurements of Figs. 3 and 10 because it shows

Vol. 6
(a)

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


(b)

263

(c)

(d)

(e)

(f)

Fig. 13. Surface topography of prepared and stretched transferred lms of elastomer E1x = 0.1 (a: = 0, b: = 10%), E2x = 0.07 (c: = 0, d: = 12%), and E2x = 0.25 (e: = 0, f: = 2%,). Scale bars 1 m, height scale 25 nm valid for all images. The surface of all polymers show plateau patterns. In the E2 samples, the lateral strain leads to surface deformation (40).

that two polar states are accessible, but the one present during cross-linking is stabilized. The shape variation under application of an external electric eld was most intensively studied for microtomized pieces of ferroelectric elastomers, which had been oriented by drawing (1921). These experiments show only a small shape variation if the eld is applied parallel to the polar axis of the monodomain. The effects become, on the other hand, rather large if the smectic layer structure (chevron texture) rearranges (19,21). To get a large electrostrictive response, which can be understood on a molecular level, the geometry presented in Fig. 15 was chosen (43). The application of an electric eld parallel to a smectic layer leads to a tilt of the mesogens (electroclinic effect). Thereby, the thickness of the layer decreases. For a stack of layers, the effect sums up over all layers. As a result, the thickness perpendicular to the smectic layers decreases if a eld is applied parallel to the layers (see Fig. 15). Because the electroclinic effect is relatively large in these polymers (16,37), a large variation in thickness is expected. X-ray diffraction measurements prove the electrically induced shrinkage of single smectic layers. A freestanding ultrathin (75 nm) lm of a ferroelectric liquid

264

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS


Charge det. Poling 200 V. 65 Fast cooling to room temperature 80 min.

Vol. 6

d33
0.8 0.6 0.4

d33

d33

d33, pC/N

0.2 0 0.2 0.4 0.6 0.8 0 20 40 60 Time, s 80 100

Fig. 14. Relaxation of the piezocoefcient d33 of elastomer E2 at room temperature after reversal poling at 65 C (SA phase) (3).

Single smectic layer h/z

for E < 0

h0h z

for E = 0

for E > 0

z smectic layers h0 E=0


Electric field E

z smectic layers h0h E0

Fig. 15. In the SA phase, the mesogenic parts (depicted as ellipsoids) of the elastomeric macromolecule stand upright ( = 0 ) inside the single smectic layers. By applying a lateral electric eld (perpendicular to the plane of the paper), a tilt angle that is proportional to the electric eld E can be induced (electroclinic effect). The sign of the tilt depends on the sign of the electric eld E. Hence, each smectic layer shrinks by h/z twice during one period of the electric eld. The shrinkage h of the whole lm is measured by the interferometer as an optical phase shift between the sample beam and the reference beam.

Vol. 6
4

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

265

a (10,000 nm2/V2)

6 5 4 3 2 1 0 55 60 65 70 Temperature, C 75 2nd harmonic SC SA

h, nm

1 1st harmonic

0 0 250 500 750 Uac, V 1000 1250 1500

Fig. 16. Electrostriction of a ferroelectric LC-elastomer (43). Big diagram: Thickness variation h as a function of the applied ac voltage U ac . Interferometric data were obtained at the fundamental frequency of the electric eld (piezoelectricity, rst harmonic: +) and at twice the frequency (electrostriction, second harmonic: ). Sample temperature: 60 C. Inset: Electrostrictive coefcient a (+) versus temperature. At the temperature where the non-cross-linked polymer would have its phase transition SC SA (about 62.5 C), the tilt angle of 0 is unstable. That is why the electroclinic effect is most effective at this temperature. An electric eld of only 1.5 MV/m is sufcient to induce lateral strains of more than 4%.

crystalline elastomer (similiar to E2) was used to measure the shape variation (electrostrictive response) associated with this (43). It was measured by a high precision (3 pm at 133 Hz) Michelson interferometer. The measurements (see Fig. 16) exhibit extremely high electrostrictive strains of 4% in an electric eld of only 1.5 MV/m, which is, to our knowledge, a new world record for the corresponding electrostrictive coefcient a. The effect exhibits typical electroclinic behavior, which means that it is caused by an electrically induced tilt of the chiral LC molecules. As a consequence of chirality, the primary strain is perpendicular to the applied eld. Hence, a new material that has a giant electrostriction effect is introduced, where the effect can be fully understood on a molecular level.

BIBLIOGRAPHY
1. M. G. Broadhurst and G. T. Davis, in P. M. Galetti, D. E. De Rossi, and A. S. De Reggi, eds., Medical Applications of Piezoelectric Polymers, Gordon and Breach, Amsterdam, 1988, pp. 314.

266

FERROELECTRIC LIQUID CRYSTALLINE ELASTOMERS

Vol. 6

2. J. W. Goodby and co-workers, Ferroelectric Liquid Crystals, Principles, Properties and Applications, Ferroelectricity and Related Phenomena, Gordon and Breach, Philadelphia, 1991. 3. M. Brehmer and co-workers, Macromol. Chem. Phys. 195, 18911904 (1994). 4. R. Zentel, Angew. Chem. 101, 14371445 (1989). 5. M. Brehmer and R. Zentel, Mol. Cryst. Liq. Cryst. 243, 353376 (1994). 6. F. J. Davis, J. Mater. Chem. 3, 551562 (1993). 7. W. Gleim and H. Finkelmann, in C. B. McArdle, ed., Side Chain Liquid Crystalline Polymers, Blackie and Son, Glasgow, 1989. 8. H. Finkelmann, in A. Ciferri, ed., Liquid Crystallinity in Polymers, VCH, Weinheim, 1991. 9. R. A. M. Hikmet, H. M. Boots, and M. Michielsen, Liq. Cryst. 19, 6576 (1995). 10. R. A. M. Hikmet and J. Lub, Prog. Polym. Sci. 21, 11651209 (1996). 11. A. Hult and co-workers, Liq. Cryst. 20, 2328 (1996). 12. R. A. M. Hikmet and H. Kempermann, Nature 392, 476478 (1998). 13. G. C. L. Wong and co-workers, Nature 389, 576579 (1997). 14. R. Zentel, Liq. Cryst. 1, 589 (1986). 15. N. R. Barnes, F. J. Davis, and G. R. Mitchell, Mol. Cryst. Liq. Cryst. 13, 168 (1989). 16. M. Brehmer and co-workers, Liq. Cryst. 21, 589596 (1996). 17. E. Gebhard and co-workers, in K. te Nijenhuis and W. J. Mijs, eds., The Wiley Polymer Networks Group Review Series, Vol. 1, Chemical and Physical Networks, Wiley, Chichester, 1998, pp. 387397. 18. E. Gebhard and R. Zentel, Macromol. Chem. Phys. 201, 902922 (2000). 19. F. Kremer and co-workers, Polym. Adv. Technol. 9, 672676 (1998). 20. W. Lehmann and co-workers, Ferroelectrics 208, 373383 (1998). 21. W. Lehmann and co-workers, J. Appl. Phys. 86, 16471652 (1999). 22. H. Poths and R. Zentel, Liq. Cryst. 16, 749767 (1994). 23. L.-C. Chien and L. G. Cada, Macromolecules 27, 3721 (1994). 24. H. M. Brodowsky and co-workers, Langmuir 15, 274278 (1999). 25. H. Schuring and co-workers, Macromolecules 34, 39623972 (2001). 26. H. Poths and co-workers, Liq. Cryst. 18, 811818 (1995). 27. A. Kocot and co-workers, Phys. Rev. B 50, 1634616357 (1994). 28. R. Stannarius and C. Cramer, Europhys. Lett. 42, 43 (1998). 29. J. Kupfer and H. Finkelmann, Macromol. Chem. Phys. 195, 13531367 (1994). 30. K.-H. Hanus and co-workers, Polym. Prepr. 30, 493 (1989). 31. I. Benn , K. Semmler, and H. Finkelmann, Macromolecules 28, 18541858 e (1995). 32. E. Nishikawa, H. Finkelmann, and H.-R. Brand, Macromol. Rapid Commun. 18, 65 (1997). 33. J. Kundler and H. Finkelmann, Macromol. Chem. Phys. 199, 677686 (1998). 34. E. Nishikawa and H. Finkelmann, Macromol. Chem. Phys. 200, 312322 (1999). 35. R. Zentel and co-workers, Makromol. Chem. 190, 2869 (1989). 36. L. R. G. Treloar, The Physics of Rubber Elasticity, Clarendon, Oxford, 1975. 37. S. Shilov and co-workers, Macromolecules 32, 15701575 (1999). 38. M. Brehmer and R. Zentel, Macromol. Chem., Rapid Commun. 16, 659662 (1995). 39. H. M. Brodowsky and co-workers, Langmuir 13, 53785382 (1997). 40. H. M. Brodowsky and co-workers, Ferroelectrics 243, 115123 (2000).

Vol. 6
41. 42. 43. 44. 45.

FIBERS, ELASTOMERIC

267

M. Mauzac and co-workers, Chem. Phys. Lett. 240, 461 (1995). S. U. Vallerien and co-workers, Makromol. Chem., Rapid Commun. 11, 593 (1990). W. Lehmann and co-workers, Nature 410, 447450 (2001). J. Zhao and co-workers, Jpn. J. Appl. Phys. 34, 5658 (1995). T. Jaworek and co-workers, Science 279, 57 (1998).

RUDOLF ZENTEL University of Mainz

You might also like