Tim J White Review Paper

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

REVIEW ARTICLE

PUBLISHED ONLINE: 22 OCTOBER 2015 | DOI: 10.1038/NMAT4433

Programmable and adaptive mechanics with liquid


crystal polymer networks and elastomers
Timothy J. White1* and Dirk J. Broer2*
Liquid crystals are the basis of a pervasive technology of the modern era. Yet, as the display market becomes commoditized,
researchers in industry, government and academia are increasingly examining liquid crystalline materials in a variety of poly-
meric forms and discovering their fascinating and useful properties. In this Review, we detail the historical development of liquid
crystalline polymeric materials, with emphasis on the thermally and photogenerated macroscale mechanical responses — such
as bending, twisting and buckling — and on local-feature development (primarily related to topographical control). Within this
framework, we elucidate the benefits of liquid crystallinity and contrast them with other stimuli-induced mechanical responses
reported for other materials. We end with an outlook of existing challenges and near-term application opportunities.

L
iquid crystals (LCs) were identified as a state of matter near the how these materials are prepared before discussing the response of
end of the nineteenth century 1, and have since then remained LCNs and LCEs to temperature. We then review recent results on
a topic of intense scientific curiosity. LCs self-organize at the the response of these materials to heat, light and magnetic fields.
molecular level (Box 1), and are classified into the subcategories Notably, we highlight efforts on the programming of LCN and
of thermotropic (order depends on temperature), lyotropic (order LCE materials to localize their mechanical response so as to gen-
depends on the concentration of material in solution) and photo­ erate surface features or shape change. We close with a forward-
tropic (order depends on the presence of light)2,3. Spurred by the looking overview of the implications of these materials for a range
employment of low-molar-mass LCs in display technologies, of applications.
research of these materials has rapidly grown in recent decades, and
is now extending into areas beyond displays, including solar-energy Preparation and properties of LCEs and LCNs
harvesting 4,5, optics and photonics6, mechanics7 and biomedicine8. The preparation of polymeric materials that exhibit liquid
Polymeric materials exhibiting liquid crystallinity have been crystallinity was initially pursued by Vorländer 14, and first realized
referred to by a variety of names, including liquid crystal polymers by Jackson and Kuhfuss15. Concurrent to advances in the develop-
(LCPs), polymeric LCs, liquid crystal elastomers (LCEs), and liquid ment and application of low-molar-mass LCs in display applica-
crystal polymer networks (LCNs). The differences in the chemi- tions, researchers achieved the preparation of well-ordered LCEs by
cal composition, crosslinking and thermo­mechanical properties crosslinking side-chain polymers16–19, where the mesogenic groups
of these materials are illustrated in Fig. 1. A liquid crystal main- are attached to siloxane or acrylate polymer main chains (Fig. 2a).
chain polymer (LCP; Fig. 1a) is a high-performance and typically A two-step crosslinking technique was developed in which the pen-
uncrosslinked macromolecule (such as Vectran) that can organ- dant mesogenic groups were oriented by mechanical stretching of
ize into liquid crystalline phases through stiff rod-like molecular the polymer during or shortly after the first-stage reaction, after
conformation and intramolecular interactions (most commonly, which the alignment is fixed by a second-stage crosslinking reac-
hydrogen bonding). LCNs (Fig. 1b) maintain some of the high- tion to form highly aligned LCEs (so-called single-crystal or mono­
performance properties of LCPs, but notably contain a moderate domain LCEs). The molecular structure of a LCE corresponds to
to densely crosslinked network architecture associated with their that of a traditional rubber: it consists of long chains of molecules
preparation from primarily (meth)acrylate-based multifunctional that can easily slip past one another and thus enable the material to
monomers. Whereas LCPs exhibit almost no change in order be expanded with very little force. Attached to the elastomer chains
(described by the order parameter, S), the order of LCNs can are the smaller rod-like molecular entities similar to those usually
decrease by as much as 5% when subjected to appropriate stimuli. found in low-molar-mass LC molecules. The weak crosslinking
LCEs (Fig. 1c) also consist of crosslinked liquid-crystal side-chain in the LCE allows spontaneous shape changes (strains) of several
and/or main-chain mesogenic units9,10, but the polymer backbone is hundred per cent (Fig. 2b,c) under load, and by the application
typically flexible (commonly a polysiloxane) and the overall cross- of stress or strain they can exhibit some unusual mechano-optical
link density is low. Unlike LCPs or LCNs, LCEs can exhibit large effects20. The mechanical responses depend on the direction of the
changes in order when subject to a stimulus. Swelling both LCNs applied stress relative to the material’s alignment direction and its
or LCEs to form liquid crystalline gels (LCGs) can further sensitize phase (smectic, nematic, cholesteric or isotropic). The stress–strain
the response of the polymeric materials to stimuli11, most notably to response of these materials has been described as ‘soft elastic-
electric fields12,13. ity’ (Fig. 2d), a term that refers to the strain that these materials
In this Review, we discuss the remarkable properties of LCNs exhibit at near-zero stress while the director is reorienting to the
and LCEs, and focus specifically on the burgeoning area of stimuli- strain direction (the physics of this process has been described in
initiated actuation and shape change. We start by briefly introducing detail in ref. 21). The optical properties of LCEs are comparable to

Air Force Research Laboratory, Materials and Manufacturing Directorate, Wright-Patterson Air Force Base, Ohio 45433, USA. 2Eindhoven University
1

of Technology, Institute for Complex Molecular Systems, Department of Chemical Engineering and Chemistry, Helix Building STO 0.34, PO Box 513,
5600 MB Eindhoven, The Netherlands. *e-mail: [email protected]; [email protected]

NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials 1087

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEW ARTICLE NATURE MATERIALS DOI: 10.1038/NMAT4433
a transition from a liquid crystalline phase (typically nematic) to
Box 1 | Order and orientation of liquid crystals. the isotropic (or paranematic) state, a substantial shrinkage strain
is observed parallel to the director vector 10. The light crosslinking
The term ‘liquid crystal’ refers to materials that exhibit long- and elastomeric nature of the materials allow for length changes
range orientational or positional organization at the molecular of as much as a factor of four (plotted as L/Liso in Fig. 2c, where
level. In low-molar-mass LCs of the calamitic subclass (the basis Liso is the length in the isotropic state and L the actual length at
of displays), the molecules self-organize to form a variety of the measuring temperature in the nematic state). The large macro-
mesophases, including nematic or chiral nematic (one dimen- scopic mechanical response is the result of disorganization of the
sion of order, in this case orientational), numerous variants of rod-like mesogenic moieties attached to the polymer main chain
smectic (two dimensions of order, orientational and positional), and to the coupling of the conformation of the macromolecular
and more complex geo­metries, such as those of the blue or bent- backbone with the orientational nematic order S (Fig. 2b)33,34. The
core phases. The order of low-molar-mass LCs can be affected polymer chains elongate when the mesogens orient in the nematic
by temperature or light, which can result in transitions between phase (Fig. 2b), whereas in the isotropic phase they recover, driven
these phases or to an isotropic state. by entropy, a random-coil conformation (Fig. 2b). As predicted
Similarly, in polymeric variants, the term ‘liquid crystalline’ by de Gennes, the individual polymer-chain shape-changes then
indicates that the material has either orientational or positional translate to a macroscopic shape-change of the elastomer sample
order of mesogenic units in the polymer backbone (main chain) at the nematic–isotropic transition35. The so-called genesis of the
or pendant groups (side chain). Figure 1 illustrates select chemi- LCE is known to strongly influence the resulting thermomechanical
cal structures of different subsets of liquid crystalline polymers. response36. These and other theoretical descriptions of contraction
The sensitivity of the order to stimuli is strongly dependent on and expansion at the nematic-to-isotropic transition stimulated a
the composition of the polymer. Inducing chirality through broader interest in artificial muscles based on thermoresponsive
chiral inclusions or surface-alignment treatments generates a LCEs10,25,37–39. Recent work in this topic has focused on exploit-
hierarchical variation in the profile of the common orientation ing the large, thermally induced mechanical response of LCEs in
vector (director) through the sample thickness, as illustrated a variety of applications40 (see ‘Remote heating’). Highlights from
in Fig. 4a–d. the recent literature include the preparation of elastomeric colloidal
materials41,42, the use of LCEs as shape-memory polymers43–45 and
surface-feature patterning46–49.
those of low-molar-mass LCs — a high birefringence and selective The substantial increase in crosslink density in the formation
reflection of polarized light in the case of the cholesteric phase. of LCNs from multifunctional liquid-crystalline monomers pre-
Recent examinations of LCEs have demonstrated potential utility cludes the material from undergoing thermotropic phase transi-
as artificial muscles (robotics)22–26, deformable lasers27 and sensors28. tions before the decomposition temperature of the materials. LCNs
Moderately to densely crosslinked glassy LCNs, which were devel- exhibit a glass transition temperature (Tg) that is typically in the
oped in the 1980s at Philips Research, are obtained from the poly­ range of 40–120 °C (at room temperature, the modulus is in the
merization of multifunctional mesogenic monomers (Fig. 3a). These order of 0.8–2 GPa; ref. 50). Owing to the anisotropy of the system,
monomers exhibit a liquid crystalline phase that can be retained after the compliance of the polymer network perpendicular to the direc-
polymerization (in most cases by photoinitiated polymerization)29–32. tor is roughly three times higher than parallel to it. The mechani-
Copolymerization of monoacrylates, such as RM23, with diacrylates, cal properties of LCNs are strongly influenced by the composition,
such as RM82 (Fig. 3a), generate LCNs with side-chain (pendant) namely the length of the aliphatic spacer of the crosslinking dia-
and main-chain mesogenic units. The advantages of this approach crylate monomer50,51 (Fig. 3a) and the phase behaviour of the mono-
are numerous. Within reasonable limits, the polymerization tem- mer or mixture31,52,53. Given the ease of preparation and availability
perature can be freely chosen, which enables the desired phase of materials, glassy LCN materials prepared from monomers, such
to be retained. Similar to low-molar-mass LCs, the order can be as those illustrated in Fig. 3a, are now widely studied, in particular
manipulated by external boundary conditions and stimuli (includ- because of their potential (see ‘Outlook’).
ing surface-alignment materials, surfactants, shear forces, and elec- As with other polymers, the volume of LCNs increases with tem-
tric, magnetic or optical fields) to prepare engineered materials with perature, and this can be described by the coefficient of thermal
complex properties and alignments that are retained indefinitely expansion α (Fig. 3b). In aligned LCNs, the sign of α is strongly
after polymerization. The ability to arrest the three-dimensional dependent on alignment 52. Figure 3b illustrates the influence of
structure of the liquid crystalline phase in polymeric form has cre- starting materials and preparation conditions on the thermal
ated a number of compelling application possibilities (see ‘Outlook’). expansion of LCN materials. Below Tg, α in the direction parallel to
In conventional polymeric materials the methods to produce such the director is close to zero. Heating above Tg causes α in this axis to
structures are limited in number, and no process that we are aware become negative. Orthogonal to the director, the thermal expansion
of enables such modularity and programmability. Exhaustive reviews rapidly increases above Tg. LCN materials prepared with a longer
of the materials chemistry and processing methods to prepare LCNs aliphatic spacer experience a larger volume increase with tempera-
can be found in refs 7, 20 and 28. ture. Furthermore, the temperature at which the LCN was prepared
can also have a profound impact on the thermal response of these
Thermomechanical responses materials52,54. LCNs prepared close to the nematic-to-­ isotropic
Thermally induced mechanical responses have been widely transition (Tp/Tc = 0.96, where Tp and Tc are the polymerization
observed in both LCEs and LCNs7,20. These thermomechanical and clearing temperatures, respectively) of the monomer show a
responses are ultimately distinguished from those of other materials smaller response than systems cured further below this LC tran-
by the possibility to programme the material’s anisotropy, orienta- sition55 (Tp/Tc = 0.86). A decrease in the curing temperature only
tion and alignment to dictate responses that are inherently driven slightly affects the order of the polymer network, but leads to a sig-
by the heterogeneity in the local orientation of LCN or LCEs rather nificant increase in the thermomechanical response of the system.
than by heterogeneities in material composition or sensitivity to For example, for C6M (Fig. 3b), decreasing the Tc from 0.96 to 0.86
stimuli (see ‘Outlook’). increases the order parameter of the network from 0.71 to 0.76, but
Loosely crosslinked LCEs exhibit thermotropism, similar to that changes the strain parallel to the director from –1.3% to –1.7% on
of low-molar-mass LCs. Accordingly, on heating LCEs through heating from –50 °C to 150 °C.

1088 NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT4433 REVIEW ARTICLE
a Liquid crystal main-chain polymers (LCPs)
Rigid main chain
O O
Tm ~>300 °C
H H E ~>100 GPa
C C N N
Flexible main chain ∆S ~0%
n

b Glassy liquid crystal polymer networks (LCNs)


Tg ~40–120 °C
E ~0.8–2 GPa
∆S ~5%

O O O O
O(CH2)xO O(CH2)xO
O O

R
n
c Liquid crystal elastomers (LCEs)
O O

H3C Si CH2 6
C O C2H5

Side-chain mesogen
Tg ~<20 °C
E ~0.1–5 MPa
n ∆S ~>90%

O O

H3C Si CH2 11
O O CH2 Si CH3
n
Crosslinker

Figure 1 | Liquid crystal polymers, polymer networks and elastomers. a, Liquid crystal polymer (LCP) is a term historically used to refer to high-
performance polymeric materials, such as Vectra (chemical structure, right), that form liquid crystalline phases. These materials are typically linear
polymers, with melting temperatures (Tm) around or exceeding 300 °C and moduli (E) that can exceed 100 GPa, that undergo minimal change in order
(ΔS) on heating. b, ‘Glassy’ liquid crystal polymer networks (LCNs) are moderately to densely crosslinked, and most often formed from the polymerization
of mesogenic (meth)acrylate monomers, which yields glassy polymers that can retain distinctive optical or mechanical properties. These materials have
glass transition temperatures (Tg) in the interval 40–120 °C and moduli of approximately 1–2 GPa, and on heating can exhibit moderate changes in order.
c, Liquid crystal elastomers (LCEs) are a subclass of LCNs for which the polymer backbone is commonly polysiloxane and the crosslink density is low.
Accordingly, subjecting these materials to appropriate stimuli can generate large changes in the order parameter that yield large strains. Red dots indicate
crosslinks, and blue rectangles indicate either main-chain or side-chain mesogenic moieties. Representative chemical formulas are shown (right).

The reduction of order leads to an increasing average tilt of the are the same but the orientation of the strain varies, thus resulting
mesogenic units that decreases the projection of the end-to-end in deflection. Further, if the orientation of the LCN is offset to the
length of the monomeric units. Accordingly, at temperatures below principal axes of the mechanical specimen, the thermal response
Tg, the system expands with temperature due to increasing molar can generate shear, yielding helicoidal and spiral ribbons on heat-
volume. The preferential expansion direction is perpendicular to the ing 57–59 (Fig. 4e,f). The generation of a helicoidal or spiral shape is
long axis of the molecule, as the expansion is dominated by increas- dictated by the aspect ratio of the sample. Moreover, the thermo-
ing intramolecular distances. Around and above Tg there is a small mechanical response can also be harnessed in other geometries41,60
and reversible loss of molecular order, which causes additional (fibres, particles) to extend, and in some ways amplify, the nascent
deformation. The reduction of order is favourable for entropic rea- thermomechanical response to generate actuation.
sons, but is limited by the polymer network. The measured change in
order parameter is small (of the order of a few per cent according to Photomechanical effects
birefringence measurements), yet geometrical arguments show that A variety of approaches have been used to generate large-scale and
this change correlates to the mechanical responses evident in Fig. 3b. efficient transduction of light into work61–66. Yet despite consider-
In practice, the temperature sensitivity of LCNs and LCEs has able effort, the direct conversion of light into large-scale mechani-
been employed to generate shape-adaptive responses. The com- cal output (typically measured as strain) was limited to less than
paratively limited strain of glassy LCNs has not hindered their 1% in photoresponsive amorphous or semicrystalline polymeric
examination for potential utility in actuation. Because these films materials67,68. In 2001, large-scale optically generated and revers-
tend to be thin, the temperature distribution across an LCN is uni- ible strain of as much as 20%69 was achieved in an LCE functional-
form. Accordingly, heating LCNs in which the alignment is uniform ized with azobenzene (azo-LCE; Fig. 4g). Subsequent examinations
(Fig. 4a) across the thickness does not induce motion. However, have reported increases in strain to 100%, and explored variations
LCNs prepared with hierarchical variation in the director (splay, to mesogen connectivity 23,70, the inclusion of guest dopants71, cor-
twisted nematic; Fig. 4c,d) have been shown to bend and coil simi- relations to phototropic phase behaviour 72,73, and theoretical treat-
larly to a bimetallic strip55,56. The thermally induced deflection of ments74–78. It should be noted that this body of work initiated a
hierarchically oriented LCNs is simply related to the anisotropy in α, renaissance in the pursuit of light-to-work transduction not only in
as is evident in Fig. 3b. As the orientation of the director rotates LCEs and LCNs, but also in crystalline solids79,80 and conventional
across the sample thickness, the relative magnitudes of the strain polymeric materials66.

NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials 1089

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEW ARTICLE NATURE MATERIALS DOI: 10.1038/NMAT4433
a
CH3 CH3 CH3
H3C Si O Si O Si CH3
CH3 H CH3
n

O COO OCH3 Pt catalyst


60 °C

O
O

(CH2)10

O
x O

b c
115 °C 90 °C 20 °C
n 3.0

2.5
L/Liso
Heat Lper/Liso
2.0
R|| (L × Lper2)/Liso)

L/Liso
Cool 1.5

1.0

R
0.5
0.80 0.85 0.90 0.95 1.00 1.05
Temperature Tred

d
I II III
I II III
σn

σthreshold

εn σthreshold σn

Figure 2 | Chemical composition, thermomechanical actuation and soft elasticity in liquid crystal elastomers. a, Synthesis of LCEs with polysiloxane
backbone. b, The change in order at or above the nematic-to-isotropic phase transition results in anisotropic deformation with shrinkage of the sample
parallel to the orientation direction. This is illustrated in the conversion of the average chain orientation in a unit (prolate conformation, where radii
R|| is greater than R⊥ in relation to the orientation of the nematic director, n) to an average chain orientation represented by a spherical conformation.
Accordingly, the contractile strain can lift a 10 g weight28. c, The strain (original length, L divided by the length in the isotropic state, Liso; L/Liso) is plotted
against reduced temperature (Tred) parallel (black squares) and perpendicular (green triangles) to the orientation direction. The relative volume, equal to
(L × Lper2)/Liso) (red circles) is constant. The large-magnitude thermomechanical responses result from thermally induced reduction in order10 (S) . d, A
polydomain LCE transitions from scattering (left) to transparent (right) as the orientation of the nematic domains align under stretching20. Concurrent to
the optical changes, LCEs exhibit a soft–elastic plateau, evident in the stress (σ)–strain (ε) curve (region II), which deviates from classical or semi-classical
elasticity (regions I and III). The changes in transparency and the soft–elastic plateaus have been related to director reorientation depicted in the plot of
order (S) versus stress. Figure reproduced with permission from: a,c, ref. 10, © 2013 Walter De Gruyter; b, ref. 28, Wiley; d, ref. 20, Oxford Univ. Press.

The directionality of the bend of a cantilever made of a glassy Oscillatory responses87,88 have been realized in these materials as well
LCN can be regulated by orienting the linear polarization of an (Fig. 4i), and the dependence of the oscillation on the exposure con-
ultraviolet light source to the axes of the film81 (Fig. 4h). Because ditions, polarization and sample thickness has also been described88.
of the large concentration of azobenzene in most LCN composi- Moreover, the frequency of the observed oscillation matches the
tions examined to date, the strong absorbance of the material local- expected resonant frequency of the cantilever 87. Key to the genera-
ized the response to the surface, resulting in bending. Subsequent tion of oscillatory responses is the employment of focused irradia-
examinations of LCN materials have explored mechanical control tion, which can allow the front and back surface of the cantilever to
with the intensity 82–84, polarization81,85 and wavelength82,83,86 of light. deflect into and out of the light. The contribution of photothermal

1090 NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT4433 REVIEW ARTICLE
a
O O O O
O(CH2)6O O(CH2)5O
O O
C6M, RM82 Cr-86-N-116-I

O O O O
O(CH2)3O O(CH2)3O
O O
C3M, RM257 Cr-73-N-129-I

O
O(CH2)5O CN

RM23 Cr-(N-44)-76-I

O O
O O O
O
O O
O O

JL129, SLO4151 Cr-69-CH-97-I (HTP ~6 µm–1)

b R
O
O CH2 x
O
O O O O
O CH2 x
O
x R Tp /Tc O

6 CH3 0.86
6 CH3 0.96
6 H 0.86 1 дL
500 α=
6 H 0.96 L0 дT P
400 3 CH3 0.86
3 CH3 0.96
300
α
200
α (ppm K–1)

100 α
α
0
α||
–100 α||

–200

–300
200 250 300 350 400 450
T (K)

Figure 3 | Liquid crystal monomers and polymer networks. a, Chemical structures of common liquid crystal monomers. The efficiency of the chiral
monomer to induce twist is referred to as helical twisting power, or HTP. Labels below the structures correspond to trade names (left) and denote phases
and transition temperatures (right; Cr, crystalline; N, nematic; I, isotropic; CH, cholesteric). b, The thermal expansion coefficient, α, of liquid crystal
networks is anisotropic in sign. The magnitude of α depends not only on the length of the aliphatic spacer unit, x, but also on the preparation conditions
(polymerization temperature, Tp) and the thermotropism of the mixture (clearing temperature, Tc). The thermal expansion coefficients correspond to
a series of LCNs prepared from the chemical structure shown, and with variations in the aliphatic spacer length (x) as well the presence of a methyl
substituent to the mesogenic core (R). The dashed lines indicate a fit of this data. Panel b adapted with permission from ref. 55, Wiley.

heating was alluded to in these works, and then further clarified in both static83,86 and oscillatory 96 deflections, has been demonstrated
subsequent studies using thermal imaging 83. Notable recent efforts as well. Yet, the magnitude of the twisting is limited in conventional
in the general area of photomechanical responses in glassy LCN domain orientations.
materials include the preparation and photomechanical charac- Taking advantage of the ability to spatially and hierarchically
terization of fibres89, the inclusion of upconverting nanoparticles to manipulate the orientation of anisotropy in LCN materials to gen-
allow for infrared-triggered responses90, bidirectional actuation85,91, erate desired effects, a five-order-of-magnitude enhancement in
shape memory 92, and the systematic examination of the role of work generation has been achieved for azo-LCNs prepared in the
crosslinkers on the generation of strain86,93–95. planar orientation with respect to azo-LCNs prepared with hier-
In addition to in-plane bending, ‘flexural–torsional’ (that is, archical structures (such as splay or twisted nematic)84,97. In addi-
bending and twisting) deflections have also been examined81. tion to enhancing the magnitude of planar deflections, hierarchical
Polarization-controlled twisting 85, also in conditions that induce LCN structures also enhance the magnitude of flexural–torsional

NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials 1091

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEW ARTICLE NATURE MATERIALS DOI: 10.1038/NMAT4433
a b c d

e g h
0.22

0.20
0.18
0.16 >540 nm 365 nm >540 nm
Contraction fraction

0.14
0.22
0.12 0.20
f 0.18 hν
0.10 0.16 –45°
0.08 0.14 –135° 365 nm 365 nm
0.12
0.10
0.06 0.08
0.06
0.04 0.04
0.02 –90°
0.02 0.00
0 50 100 150 200 250 >540 nm 365 nm >540 nm
0.00
0 10 20 30 40 50 60 70 80 90
Time (min)

i j
dL dR
Start position

Azo-LCN
UV (λ = 365 nm)
Light beam

Visible light
Oscillation angle

Holder

Figure 4 | Bends, twists and turns in liquid crystal polymer networks. Representation of changes in LCN or LCE dimensions on exposure to an order-
disrupting stimuli. a–d, Planar uniaxial (a), cholesteric (b), twisted nematic (c) and splay (d) director profiles and their deformations corresponding to
a decrease in the order parameter. e,f, Offsetting the nematic director to the principal axes of the sample can generate shear, which has been observed
on heating in LCN materials. The handedness of the twisting (left or right) of the material is dictated by the material’s chirality across the sample
thickness. The films are approximately 6–10 mm in length and 0.5–2 mm in width. g, Photomechanical effects for LCEs (25 °C, asterisks; 30 °C, circles;
35 °C, triangles; 40 °C, squares) and LCNs. Inset: The relaxation of the photogenerated strain in the dark at 25 °C. h, Photodirected bending of a LCN
film on irradiation with linearly polarized 365-nm light. The orientation of linearly polarized light (0°, −45°, −90°, −135°) dictates the directionality of
the deflection of the samples. Subsequent irradiation with light of wavelength greater than 540 nm flattens the film. i, In appropriate optical conditions,
irradiation with blue–green light can initiate oscillations87. The cantilever length is 5 mm. j, Light can also be used to introduce both left (dL)- and right (dR)-
handed spirals in a photoresponsive LCN99. The films are more than 10 mm in length. Figure reproduced with permission from: e, ref. 58, NAS; f, ref. 57,
Wiley; g, ref. 69, APS; h, ref. 81, Nature Publishing Group; i, ref. 87, RSC; j, ref. 99, Nature Publishing Group.

responses, as exemplified by reports of photoinduced twisting and However, a number of recent reports intentionally hybridized
shape formation in twisted nematic films98,99 (Fig. 4j). photo­mechanical and thermomechanical effects through the addi-
tion of guest materials that are efficient heat-transfer agents. In this
Remote heating regard, the preparation of an LCE composed with carbon nanotube
In addition to employing photochemistry, contactless actuation (CNT) additives initially focused on enabling electro­mechanical
of LCN materials can be also be accomplished directly or indi- effects in these systems105. Building on work in non-liquid-­
rectly through absorptive heating with either optical or magnetic crystalline systems106–109, photomechanical effects in composites
stimuli. Distinguishing photochemical mechanisms from photo- of CNTs and LCEs were later reported110–117. Although the energy
thermal contributions (if any) is a consistent endeavour 86,72,73,100–104. transfer is indirect (that is, light to heat to work), the effects can be

1092 NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT4433 REVIEW ARTICLE
a b
Planar Homeotropic
200 µm

λ=
45
0n
m

O O N
N
O

O O UV

Chiral nematic Homeotropic

c d

140 °C

RT

t=0s t = 10 s

140 °C

RT

t = 17 s t = 38 s

Figure 5 | Reconfigurable topography. a, Monolithic polymer coatings prepared from LCNs change topography with irradiation by light123–125. The polarization
micrograph shows alternating stripes of a perpendicularly oriented LCN next to an area with a planar chiral-nematic order. On actuation with ultraviolet light
the planar chiral areas expand, whereas the perpendicularly oriented area contracts, resulting in a regular surface profile. The process is reversible. b, When
the axis of the helices of the chiral-nematic LCN are oriented parallel to the surface, a fingerprint pattern that switches from planar to homeotropic is formed,
which when subjected to a stimulus, induces a corrugated surface roughness. c,d, Spatial variation in the local orientation of twisted nematic domains within
LCN materials has been shown to generate complex mechanical responses on heating, including ripples (c, top right)57, localized curling (c, bottom right)57,
and localized ridges (d)128. In c, the blue squares denote a planar nematic region and the red squares denote a twisted nematic region in the LCN films. RT,
room temperature. Figure reproduced with permission from: a, ref. 124, Wiley; b, ref. 123, Wiley; c, ref. 57, Wiley; d, ref. 128, Wiley.

triggered with white light and infrared sources. The absorption of LC materials can spontaneously organize, or can be forced to
photons by CNTs or other broadband absorbers, such as graphene, align. Polymerization of mixtures composed of liquid crystalline
is radiated as heat, and heat transfer triggers thermomechanical monomers has been shown to lock this complex orientation in
effects (local strain) in the polymer network. Methods to homo- monolithic form. This is most plainly evident in LCNs that retain
geneously disperse CNTs into LCEs and other matrices have been the twisted nematic and splay orientations or the cholesteric
shown, and the benefits of adding CNTs, for example, increased LC phase (Fig. 4b–d). In each of these geometries, the director
toughness and conductivity 105,110,114,115, have also been documented. rotates across the sample thickness. Because of the association
Cantilever bending has been used to visualize the response of of the thermomechanical or photomechanical responses and the
these materials111. anisotropy of LCN materials, these geometries can be thought of
Magnetic actuation has also been harnessed in elastomeric as monolithic analogues to those of functionally graded compos-
and glassy LCNs with magnetic nanoparticles to allow for remote ites. In combination with lateral (x–y axis) alignment techniques,
actuation118–121. For a review, see ref. 122. such as rubbing, electric field or photoalignment, polymeric
films can be prepared with domain variations across the sam-
Stimuli-responsive topographical effects ple thickness (z axis) so as to achieve complex topographical
One of the distinguishing features of LCN materials, when com- surface deformations123–127.
pared with the large number of functional polymers developed One approach for the generation of spatially resolved variations
to date, is the ability to prepare films that are homogeneous in in the order and orientation of LCNs employs the electro-optic
composition but heterogeneous in mechanical response. We response of materials (Fig. 3a) and electrode patterning 123–125. The
conclude this review of stimuli-responsive LCN materials with a generation of topographical surface features in LCN films through
summary of recent work on the programming of local aniso­tropy this method has been recently reviewed127. The use of light to
in LCN materials to yield spatially complex shape-changes or remotely trigger topographical features in LCN materials is illus-
surface-variations. trated in Fig. 5a,b. Using a two-step polymerization process and

NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials 1093

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEW ARTICLE NATURE MATERIALS DOI: 10.1038/NMAT4433

a b

r0
ϕ

r’sinϕ

c
–5/2 –2 –3/2 –1 –1/2 +1/2 +1 +3/2 +2 +5/2

d e

5 mm 25 °C 55mm
mm 175 °C

Figure 6 | Topography from topological defects. a, Theoretical examination129 of LCNs predict conical and anticonical deformation of topologically
imprinted defects of +1 strength. b, Schematics illustrating the nematic-director orientation of +1 radial and azimuthal defects, and experimental
realization134 of the predicted deformations in LCNs (film approximately 10 mm in diameter). c, Preparation and characterization of films subsumed with
topological defects ranging from –5/2 to +5/2 in strength. Top row: Polarized optical micrographs confirming the charge strength as well as centre of
the point singularity. Middle row: Illustration of the photoinduced mechanical response of the films on irradiation with UV light. Bottom row: Imaged
photoinduced mechanical response135. The diameter of the films is 1 cm. d, Arrays of 41 +2 or +4 topological defects can be actuated to generate periodic
topographical surfaces135. e, Enabled by the formulation of chemistry conducive to photoalignment techniques employed in b–d, LCE films were prepared
with a 3 × 3 array of +1 radial defects137. The increase in strain to 60% substantially increases the deflection of the tip of the conical deformations to as
much as 5 mm. Figure reproduced with permission from: a, ref. 129, APS; b, ref. 134, Wiley; c,d, ref. 135, Wiley; e, ref. 137, AAAS.

electric fields, the LCN material maintains alternating regions of been shown to exhibit large-scale, accordion-like actuation as
homeotropic and planar boundary conditions. Owing to the peri- well128 (Fig. 5d).
odic variation of the director profile in the chiral nematic regions, Motivated by theoretical work129–133 (Fig. 6a), distinctive shape
large surface features have been reported with heat, light and and surface features in LCN films with topological defects have
chemical stimuli. Photoalignment patterning can also be used to been achieved. In an initial demonstration (Fig. 6b), LCN films
prepare LCN films with alternating regions of monodomain and containing a heat-transfer dye were prepared with central-point
twisted nematic orientation (Fig. 5c,d). On heating through Tg, defects of charge +1/2, +1, −1/2 and −1, as well as combinations
the films exhibit complex deformation, in which the monodomain thereof 134. Defects (Fig. 6c) and arrays of defects (Fig. 6d) with
region remains flat and the twisted nematic regions ripple and curl strengths ranging from ±0.5 to 10 have demonstrated rich and
(Fig. 5c)57. Patterned twisted nematic LCN materials have recently diversified photo­induced topographical features135.

1094 NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT4433 REVIEW ARTICLE
a b

Punch
Liquid crystalline
blister
LCE–CNT

Die

Rigid
skeleton

c
Light source
off 3.0 mm

LCE–CNT

Light source
on

Figure 7 | Communicating through touch. Haptic-display principle based on the selective actuation of an elastomeric LCN prepared with carbon nanotube
(CNT) additives148. a, The elastomeric LCN contracts when a light source switches on, leading to a flat surface. b,c, Preparation of the perforations by
stretching the LCE–CNT composite across a die and punching it to generate blisters illustrated in a and depicted in c. In c, Perforations apparent in the
blisters (inset) could be used in applications, such as a dynamic Braille display. Figure reproduced with permission: a,b, ref. 148, IOP.

Recent efforts have extended preparation methods to programme Photoalignment methods offer the potential for elaborate spatial
LCEs136. Very recently, LCE materials were prepared in cells that are control to form volume elements (voxels), analogous to the pixela-
sensitive to photoalignment 137, and very large conical deformations tion of LC displays137. The generation of shape-changes or dynamic
were shown to be capable of generating as much as 2.5 J kg−1 in work topographical features has potential uses in micro­ fluidics143,144,
(Fig. 6e)137. Furthermore, the ability to spatially pattern (voxelate) flow control144, solar-energy harvesting 87,145–147 and haptic dis-
local regions of the material was employed to prepare a self-folding plays112–114,148,149 (Fig. 7).
Miura-ori origami pattern137. Many reports to date have discussed the potential use of these
materials in actuation. Yet it is critical to define what is meant by
Outlook actuation. Sometimes actuation is meant to simply imply motion.
To project the future opportunities for stimuli-responsive LCNs Others define an actuator as a system in which the stimuli-­responsive
and LCEs, it is important to distinguish the materials’ novel fea- element is but a small part of a larger system composed of amplify-
tures with respect to the broader literature on stimuli-responsive ing elements and other mechanisms. Regardless of the definition,
materials and on active mechanisms in conventional materials the term actuation implies purpose. One significant challenge is the
and actuators. Compared with peer material technologies, such extension of the basic understanding of the chemistry and physics
as shape-memory polymers, electroactive polymers and other of these materials, and their responses to stimuli, to what and how
responsive materials, stimuli-responsive LCNs and LCEs have these can enable distinctive performance when in the hands of a
many similarities as well as some advantages and disadvantages. mechanical designer. Towards this end, it is important for individu-
For instance, both LCNs and LCEs are capable of exhibiting either als in the community to engage and partner with peers in mechanics
shape-fixing (shape memory)43–45,57,92,138–141 or shape-restoring to frame and articulate potential end uses that will ultimately guide
(artificial muscle)10,23–25,37–39 responses. Within the larger stimuli- the materials development and characterization processes. In this
responsive polymeric literature, both of these properties have respect, an excellent resource is the Ashby plots150.
been identified as potential enablers to the realization of novel Stimuli-responsive liquid-crystalline polymer networks offer a
biomedical devices, soft robotics and morphing structures. The promising means to generate useful functional devices. As detailed
ability of LCNs or LCEs to self-organize to form materials of in this Review, a diverse range of responses have already been
homogeneous composition with spatial variation of the mechani- reported with a number of stimuli. Building on the foundational
cal response (evident in the localization of planar or hierarchical knowledge of the response of these films, future work exploiting the
domain orientations in Fig. 5c,d, or topological defect structures ability to pattern the director profile of these materials to generate
in Fig. 6a–e) is not simple to emulate in other stimuli-responsive engineered materials without creases or interconnections shows
materials. Key to enabling the distinctive ability to generate spatial promise in a range of applications in haptic displays, lab-on-chip,
variation in the directionality and hierarchical orientation in LCN aerospace and optics.
materials are surface-alignment methods, which can involve rub-
bing, magnetic fields and light (photoalignment). Magnetic-field Received 24 July 2014; accepted 26 August 2015;
alignment has recently been used to prepare a dynamic aperture142. published online 22 October 2015

NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials 1095

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEW ARTICLE NATURE MATERIALS DOI: 10.1038/NMAT4433

References 34. Cotton, J. P. & Hardouin, F. Chain conformation of liquid-crystalline


1. Reinitzer, F. Beiträge zur kenntniss des cholesterins. Monatsh. Chem. polymers studied by small-angle neutron scattering. Prog. Polym. Sci.
9, 421–441 (1888). 22, 795–828 (1997).
2. Collings, P. J. & Hird, M. An Introduction to Liquid Crystals: Chemistry and 35. De Gennes, P. G. Réflexions sur un type de polymères nématiques.
Physics (CRC Press, 1997). C. R. Acad. Sci. B281, 101–103 (1975).
3. Prasad, S. K. Photostimulated and photosuppressed phase transitions in liquid 36. Urayama, K., Kohmon, E., Kojima, M. & Takigawa, T. Polydomain−
crystals. Angew. Chem. Int. Ed. 51, 10708–10710 (2012). monodomain transition of randomly disordered nematic elastomers with
4. Schmidt-Mende, L. et al. Self-organized discotic liquid crystals for high- different cross-linking histories. Macromolecules 42, 4084–4089 (2009).
efficiency organic photovoltaics. Science 293, 1119–1122 (2001). 37. De Gennes, P. G., Hebert, M. & Kant, R. Artificial muscles based on nematic
5. Verbunt, P. P. C. et al. Increased efficiency of luminescent solar concentrators gels. Macromol. Symp. 113, 39–49 (1997).
after application of organic wavelength selective mirrors. Opt. Express 38. Hebert, M., Kant, R. & De Gennes, P. G. Dynamics and thermodynamics of
20, A655–A668 (2012). artificial muscles based on nematic gels. J. Phys. I 7, 909–919 (1997).
6. Li, Q. Liquid Crystals Beyond Displays: Chemistry, Physics, and Applications 39. Thomsen, D. L. III. et al. Liquid crystal elastomers with mechanical
(John Wiley & Sons, 2012). properties of a muscle. Macromolecules 34, 5868–5875 (2001).
7. Broer, D. J., Crawford, G. P. & Zumer, S. Cross-Linked Liquid Crystalline 40. de Jeu, W. H. Liquid Crystal Elastomers: Materials and Applications
Systems: From Rigid Polymer Networks to Elastomers (CRC Press, 2011). (Springer, 2012).
8. Woltman, S. J., Jay, G. D. & Crawford, G. P. Liquid-crystal materials find a 41. Fleischmann, E-K., Ohm, C., Serra, C. & Zentel, R. Preparation of soft
new order in biomedical applications. Nature Mater. 6, 929–938 (2007). microactuators in a continuous flow synthesis using a liquid-crystalline
9. Beyer, P., Terentjev, E. M. & Zentel, R. Monodomain liquid crystal main polymer crosslinker. Macromol. Chem. Phys. 213, 1871–1878 (2012).
chain elastomers by photocrosslinking. Macromol. Rapid Commun. 42. Evans, J. S. et al. Active shape-morphing elastomeric colloids in short-pitch
28, 1485–1490 (2007). cholesteric liquid crystals. Phys. Rev. Lett. 110, 187802 (2013).
10. Wermter, H. & Finkelmann, H. Liquid crystalline elastomers as artificial 43. Rousseau, I. A. & Mather, P. T. Shape memory effect exhibited by smectic-C
muscles. e-Polymers 1, 111–123 (2001). liquid crystalline elastomers. J. Am. Chem. Soc. 125, 15300–15301 (2003).
11. Urayama, K. Selected issues in liquid crystal elastomers and gels. 44. Burke, K. A. & Mather, P. T. Soft shape memory in main-chain liquid
Macromolecules 40, 2277–2288 (2007). crystalline elastomers. J. Mater. Chem. 20, 3449–3457 (2010).
12. Urayama, K., Honda, S. & Takigawa, T. Electrically driven deformations of 45. Burke, K. A. & Mather, P. T. Evolution of microstructure during shape
nematic gels. Phys. Rev. E 71, 051713 (2005). memory cycling of a main-chain liquid crystalline elastomer. Polymer
13. Urayama, K., Honda, S. & Takigawa, T. Deformation coupled to director 54, 2808–2820 (2013).
rotation in swollen nematic elastomers under electric fields. Macromolecules 46. Zupancic, B., Zalar, B., Remskar, M. & Domenici, V. Actuation of gold-coated
39, 1943–1949 (2006). liquid crystal elastomers. Appl. Phys. Express 6, 021701 (2013).
14. Vorlander, D. Investigation of the molecular form by means of crystalline 47. Wu, Z. L. et al. Microstructured nematic liquid crystalline elastomer
liquids. Z. Phys. Chem. 105, 211–254 (1923). surfaces with switchable wetting properties. Adv. Funct. Mater.
15. Jackson, W. J. & Kuhfuss, H. F. Liquid crystal polymers. I. Preparation and 23, 3070–3076 (2013).
properties of p-hydroxybenzoic acid copolyesters. J. Polym. Sci. Polym. Chem. 48. Wei, R., He, Y., Wang, X. & Keller, P. Nematic liquid crystalline elastomer
14, 2043–2058 (1976). grating and microwire fabricated by micro-molding in capillaries.
16. Finkelmann, H., Kock, H-J. & Rehage, G. Investigations on liquid crystalline Macromol. Rapid Commun. 34, 330–334 (2013).
polysiloxanes. 3. Liquid crystalline elastomers — a new type of liquid 49. Corbett, D. R. & Adams, J. M. Tack energy and switchable adhesion of liquid
crystalline material. Macromol. Rapid Commun. 2, 317–322 (1981). crystal elastomers. Soft Matter 9, 1151–1163 (2013).
17. Portugall, M., Ringsdorf, H. & Zentel, R. Synthesis and phase behavior of 50. Hikmet, R. A. M. & Broer, D. J. Dynamic mechanical properties of
liquid crystalline polyacrylates. Makromol. Chem. 183, 2311–2321 (1982). anisotropic networks formed by liquid-crystalline acrylates. Polymer
18. Ringsdorf, H. & Zentel, R. Liquid crystalline side chain polymers and their 32, 1627–1632 (1991).
behavior in the electric field. Makromol. Chem. 183, 1245–1256 (1982). 51. Broer, D. J., Mol, G. N. & Challa, G. In-situ photopolymerization of oriented
19. Küpfer, J. & Finkelmann, H. Nematic liquid single crystal elastomers. liquid-crystalline acrylates. 5. Influence of the alkylene spacer on the
Makromol. Chem. Rapid Commun. 12, 717–726 (1991). properties of the mesogenic monomers and the formation and properties of
20. Terentjev, E. M. & Warner, M. Liquid Crystal Elastomers (Oxford Univ. oriented polymer networks. Makromol. Chem. 192, 59–74 (1991).
Press, 2009). 52. Broer, D. J. & Mol, G. N. Anisotropic thermal expansion of densely
21. Warner, M., Bladon, P. & Terentjev, E. “Soft elasticity” — deformation without crosslinked oriented polymer networks. Polym. Eng. Sci. 31, 625–631 (1991).
resistance in liquid crystal elastomers. J. Phys. II 4, 93–102 (1994). 53. Hikmet, R. A. M., Zwerver, B. H. & Broer, D. J. Anisotropic polymerization
22. De Gennes, P. G. Possibilites offertes par la reticulation de polymeres en presence shrinkage behavior of liquid-crystalline diacrylates. Polymer
d'un cristal liquide. Phys. Lett. 28A, 725–726 (1969). 33, 89–95 (1992).
23. Li, M-H., Keller, P., Li, B., Wang, X. & Brunet, M. Light-driven side-on 54. Wie, J. J., Lee, K. M., Ware, T. H. & White, T. J. Twists and turns in glassy,
nematic elastomer actuators. Adv. Mater. 15, 569–572 (2003). liquid crystalline polymer networks. Macromolecules 48, 1087–1092 (2015).
24. Buguin, A., Li, M-H., Silberzan, P., Ladoux, B. & Keller, P. Micro-actuators: 55. Mol, G. N., Harris, K. D., Bastiaansen, C. W. M. & Broer, D. J. Thermo-
when artificial muscles made of nematic liquid crystal elastomers meet soft mechanical responses of liquid-crystal networks with a splayed molecular
lithography. J. Am. Chem. Soc. 128, 1088–1089 (2006). organization. Adv. Funct. Mater. 15, 1155–1159 (2005).
25. Li, M-H. & Keller, P. Artificial muscles based on liquid crystal elastomers. 56. Harris, K. D., Bastiaansen, C. W. M., Lub, J. & Broer, D. J. Self-assembled
Phil. Trans. R. Soc. A 364, 2763–2777 (2006). polymer films for controlled agent-driven motion. Nano Lett.
26. Tajbakhsh, A. R. & Terentjev, E. M. Spontaneous thermal expansion of 5, 1857–1860 (2005).
nematic elastomers. Eur. Phys. J. E 6, 181–188 (2001). 57. Lee, K. M., Bunning, T. J. & White, T. J. Autonomous, hands-free shape
27. Finkelmann, H., Kim, S. T., Muñoz, A., Palffy-Muhoray, P. & Taheri, B. memory in glassy, liquid crystalline polymer networks. Adv. Mater.
Tunable mirrorless lasing in cholesteric liquid crystalline elastomers. 24, 2839–2843 (2012).
Adv. Mater. 13, 1069–1072 (2001). 58. Sawa, Y. et al. Shape selection of twist-nematic-elastomer ribbons.
28. Ohm, C., Brehmer, M. & Zentel, R. Liquid crystalline elastomers as actuators Proc. Natl Acad. Sci. USA 108, 6364–6368 (2011).
and sensors. Adv. Mater. 22, 3366–3387 (2010). 59. Sawa, Y. et al. Shape and chirality transitions in off-axis twist nematic
29. Broer, D. J., Boven, J., Mol, G. N. & Challa, G. In-situ photopolymerization elastomer ribbons. Phys. Rev. E 88, 022502 (2013).
of oriented liquid-crystalline acrylates. 3. Oriented polymer networks from a 60. Ohm, C. et al. Preparation of actuating fibers of oriented main-chain
mesogenic diacrylate. Makromol. Chem. 190, 2255–2268 (1989). liquid crystalline elastomers by a wet spinning process. Soft Matter
30. Broer, D. J., Finkelmann, H. & Kondo, K. In-situ photopolymerization of an 7, 3730–3734 (2011).
oriented liquid-crystalline acrylate. Makromol. Chem. 189, 185–194 (1988). 61. Ikeda, T. & Zhao, Y. Smart Light-Responsive Materials: Azobenzene-
31. Broer, D. J., Hikmet, R. A. M. & Challa, G. In-situ photopolymerization of Containing Polymers and Liquid Crystals (Wiley, 2009).
oriented liquid-crystalline acrylates. 4. Influence of a lateral methyl substituent 62. Corbett, D. & Warner, M. Changing liquid crystal elastomer ordering
on monomer and oriented polymer network properties of a mesogenic with light — a route to opto-mechanically responsive materials. Liq. Cryst.
diacrylate. Makromol. Chem. 190, 3201–3215 (1989). 36, 1263–1280 (2009).
32. Broer, D. J., Mol, G. N. & Challa, G. In situ photopolymerization of an 63. Ikeda, T., Mamiya, J-I. & Yu, Y. Photomechanics of liquid-crystalline
oriented liquid-crystalline acrylate. 2. Makromol. Chem. 190, 19–30 (1989). elastomers and other polymers. Angew. Chem. Int. Ed. 46, 506–528 (2007).
33. De Gennes, P. G. in Polymer Liquid Crystals (eds Cifferi, A. et al.) 115–131 64. Ikeda, T. & Ube, T. Photomobile polymer materials: from nano to macro.
(Academic, 1982). Mater. Today 14, 480–487 (October, 2011).

1096 NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials

© 2015 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT4433 REVIEW ARTICLE
65. Koerner, H., White, T. J., Tabiryan, N. V., Bunning, T. J. & 95. Lee, K. M., Koerner, H., Vaia, R. A., Bunning, T. J. & White, T. J. Relationship
Vaia, R. A. Photogenerating work from polymers. Mater. Today 11, 34–42 between the photomechanical response and the thermomechanical properties
(July–August, 2008). of azobenzene liquid crystalline polymer networks. Macromolecules
66. White, T. J. Light to work transduction and shape memory in glassy, 43, 8185–8190 (2010).
photoresponsive macromolecular systems: trends and opportunities. 96. Lee, K. M. et al. Photodriven, flexural-torsional oscillations of glassy azobenzene
J. Polym. Sci. B 50, 877–880 (2012). liquid crystal polymer networks. Adv. Funct. Mater. 15, 2913–2918 (2011).
67. Eisenbach, C. D. Isomerization of aromatic azo chromophores in 97. van Oosten, C. L., Harris, K. D., Bastiaansen, C. W. M. & Broer, D. J. Glassy
poly(ethyl acrylate) networks and photomechanical effect. Polymer photomechanical liquid-crystal network actuators for microscale devices.
21, 1175–1179 (1980). Eur. Phys. J. E 23, 329–336 (2007).
68. Agolini, F. & Gay, F. P. Synthesis and properties of azoaromatic polymers. 98. Wie, J. J., Lee, K. M., Smith, M. L., Vaia, R. A. & White, T. J. Torsional
Macromolecules 3, 349–351 (1970). mechanical responses in azobenzene functionalized liquid crystalline polymer
69. Finkelmann, H., Nishikawa, E., Pereira, G. G. & Warner, M. A new opto- networks. Soft Matter 9, 9303–9310 (2013).
mechanical effect in solids. Phys. Rev. Lett. 87, 015501 (2001). 99. Iamsaard, S. et al. Conversion of light into macroscopic helical motion.
70. Sanchez-Ferrer, A., Merekalov, A. & Finkelmann, H. Opto-mechanical Nature Chem. 6, 229–235 (2014).
effect in photoactive nematic side-chain liquid-crystalline elastomers. 100. Warner, M. & Terentjev, E. Thermal and photo-actuation in nematic elastomers.
Macromol. Rapid Commun. 32, 671–678 (2011). Macromol. Symp. 200, 81–92 (2003).
71. Camacho-Lopez, M., Finkelmann, H., Palffy-Muhoray, P. & Shelley, M. Fast 101. Harvey, C. L. M. & Terentjev, E. M. Role of polarization and alignment in
liquid crystal elastomer swims in the dark. Nature Mater. 3, 307–310 (2004). photoactuation of nematic elastomers. Eur. Phys. J. E 23, 185–189 (2007).
72. Cviklinski, J., Tajbakhsh, A. R. & Terentjev, E. M. UV isomerisation in 102. Hon, K. K., Corbett, D. & Terentjev, E. M. Thermal diffusion and bending
nematic elastomers as a route to photo-mechanical transducer. Eur. Phys. J. E kinetics in nematic elastomer cantilever. Eur. Phys. J. E 25, 83–89 (2008).
9, 427–434 (2002). 103. Dawson, N. J., Kuzyk, M. G., Neal, J., Luchette, P. & Palffy-Muhoray, P. Modeling
73. Hogan, P. M., Tajbakhsh, A. R. & Terentjev, E. M. UV manipulation the mechanisms of the photomechanical response of a nematic liquid crystal
of order and macroscopic shape in nematic elastomers. Phys. Rev. E elastomer. J. Opt. Soc. Am. B 28, 2134–2141 (2011).
65, 041720 (2002). 104. Dawson, N. J., Kuzyk, M. G., Neal, J., Luchette, P. & Palffy-Muhoray, P.
74. Warner, M. & Mahadevan, L. Photoinduced deformations of beams, plates, Experimental studies of the mechanisms of photomechanical effects in a
and films. Phys. Rev. Lett. 92, 134302 (2004). nematic liquid crystal elastomer. J. Opt. Soc. Am. B 28, 1916–1921 (2011).
75. Corbett, D. & Warner, M. Nonlinear photoresponse of disordered elastomers. 105. Courty, S., Mine, J., Tajbakhsh, A. R. & Terentjev, E. M. Nematic elastomers with
Phys. Rev. Lett. 96, 237802 (2006). aligned carbon nanotubes: new electromechanical actuators. Europhys. Lett.
76. Corbett, D. & Warner, M. Linear and nonlinear photoinduced deformations 64, 654–660 (2003).
of cantilevers. Phys. Rev. Lett. 99, 174302 (2007). 106. Koerner, H., Price, G., Pearce, N. A., Alexander, M. & Vaia, R. A. Remotely
77. Corbett, D. & Warner, M. Bleaching and stimulated recovery of dyes and of actuated polymer nanocomposites — stress-recovery of carbon-nanotube-filled
photocantilevers. Phys. Rev. E 77, 051710 (2008). thermoplastic elastomers. Nature Mater. 3, 115–120 (2004).
78. Corbett, D. & Warner, M. Polarization dependence of optically driven 107. Ahir, S. V. & Terentjev, E. M. Photomechanical actuation in polymer–nanotube
polydomain elastomer mechanics. Phys. Rev. E 78, 061701 (2008). composites. Nature Mater. 4, 491–495 (2005).
79. Nath, N. K., Panda, M. K., Sahoo, S. C. & Naumov, P. Thermally induced and 108. Ahir, S. V. & Terentjev, E. M. Fast relaxation of carbon nanotubes in polymer
photoinduced mechanical effects in molecular single crystals — a revival. composite actuators. Phys. Rev. Lett. 96, 133902 (2006).
CrystEngComm 16, 1850–1858 (2014). 109. Ahir, S. V., Squires, A. M., Tajbakhsh, A. R. & Terentjev, E. M. Infrared actuation
80. Kim, T., Zhu, L., Al-Kaysi, R. O. & Bardeen, C. J. Organic photomechanical in aligned polymer–nanotube composites. Phys. Rev. B 73, 085420 (2006).
materials. ChemPhysChem 15, 400–414 (2014). 110. Marshall, J. E., Ji, Y., Torras, N., Zinoviev, K. & Terentjev, E. M. Carbon-
81. Yu, Y., Nakano, M. & Ikeda, T. Photomechanics: directed bending of a nanotube sensitized nematic elastomer composites for IR-visible photo-
polymer film by light. Nature 425, 145 (2003). actuation. Soft Matter 8, 1570–1574 (2012).
82. Yamada, M. et al. Photomobile polymer materials: towards light-driven 111. Torras, N., Zinoviev, K. E., Marshall, J. E., Terentjev, E. M. & Esteve, J. Bending
plastic motors. Angew. Chem. Int. Ed. 47, 4986–4988 (2008). kinetics of a photo-actuating nematic elastomer cantilever. Appl. Phys. Lett.
83. Lee, K. M. & White, T. J. Photochemical mechanism and photothermal 99, 254102 (2011).
considerations in the mechanical response of monodomain, azobenzene- 112. Campo, E. M. et al. Nano opto-mechanical systems (NOMS) as a proposal for
functionalized liquid crystal polymer networks. Macromolecules tactile displays. Proc. SPIE 8107, 81070H (2011).
45, 7163–7170 (2012). 113. Camargo, C. J. et al. Microstamped opto-mechanical actuator for tactile displays.
84. Harris, K. D. et al. Large amplitude light-induced motion in high elastic Proc. SPIE 8107, 810709 (2011).
modulus polymer actuators. J. Mater. Chem. 15, 5043–5048 (2005). 114. Camargo, C. J. et al. Localised actuation in composites containing carbon
85. Tabiryan, N., Serak, S., Dai, X-M. & Bunning, T. Polymer film with optically nanotubes and liquid crystalline elastomers. Macromol. Rapid Commun.
controlled form and actuation. Opt. Express 13, 7442–7448 (2005). 32, 1953–1959 (2011).
86. Lee, K. M., Tabiryan, N. V., Bunning, T. J. & White, T. J. Photomechanical 115. Ji, Y., Huang, Y. Y., Rungsawang, R. & Terentjev, E. M. Dispersion and alignment
mechanism and structure–property considerations in the generation of of carbon nanotubes in liquid crystalline polymers and elastomers. Adv. Mater.
photomechanical work in glassy, azobenzene liquid crystal polymer networks. 22, 3436–3440 (2010).
J. Mater. Chem. 22, 691–698 (2012). 116. Li, C., Liu, Y., Lo, C-w. & Jiang, H. Reversible white-light actuation of carbon
87. Serak, S., Tabiryan, N., White, T. J., Vaia, R. A. & Bunning, T. J. Liquid nanotube incorporated liquid crystalline elastomer nanocomposites. Soft Matter
crystalline polymer cantilever oscillators fueled by light. Soft Matter 7, 7511–7516 (2011).
6, 779–783 (2010). 117. Yang, L., Setyowati, K., Li, A., Gong, S. & Chen, J. Reversible infrared actuation
88. White, T. J. et al. High frequency photodriven polymer oscillator. Soft Matter of carbon nanotube-liquid crystalline elastomer nanocomposites. Adv. Mater.
4, 1796–1798 (2008). 20, 2271–2275 (2008).
89. Yoshino, T. et al. Three-dimensional photomobility of crosslinked azobenzene 118. Kaiser, A., Winkler, M., Krause, S., Finkelmann, H. & Schmidt, A. M.
liquid-crystalline polymer fibers. Adv. Mater. 22, 1361–1363 (2010). Magnetoactive liquid crystal elastomer nanocomposites. J. Mater. Chem.
90. Wu, W. et al. NIR-light-induced deformation of cross-linked liquid- 19, 538–543 (2009).
crystal polymers using upconversion nanophosphors. J. Am. Chem. Soc. 119. Winkler, M., Kaiser, A., Krause, S., Finkelmann, H. & Schmidt, A. M.
133, 15810–15813 (2011). Liquid crystal elastomers with magnetic actuation. Macromol. Symp.
91. White, T. J., Serak, S. V., Tabiryan, N. V., Vaia, R. A. & Bunning, T. J. 291–292, 186–192 (2010).
Polarization-controlled, photodriven bending in monodomain liquid crystal 120. Zhou, Y. et al. Hierarchically structured free-standing hydrogels with liquid
elastomer cantilevers. J. Mater. Chem. 19, 1080–1085 (2009). crystalline domains and magnetic nanoparticles as dual physical cross-linkers.
92. Lee, K. M., Koerner, H., Vaia, R. A., Bunning, T. J. & White, T. J. Light- J. Am. Chem. Soc. 134, 1630–1641 (2012).
activated shape memory of glassy azobenzene liquid crystal polymer 121. Petsch, S. et al. A thermotropic liquid crystal elastomer micro-actuator
networks. Soft Matter 7, 4318–4324 (2011). with integrated deformable micro-heater. IEEE 27th Int. Conf. Micro Electro
93. Mamiya, J-i., Yoshitake, A., Kondo, M., Yu, Y. & Ikeda, T. Is chemical Mechanical Sys. 905–908 (2014).
crosslinking necessary for the photoinduced bending of polymer films? 122. Chambers, M. et al. Liquid crystal elastomer–nanoparticle systems for actuation.
J. Mater. Chem. 18, 63–65 (2008). J. Mater. Chem. 19, 1524–1531 (2009).
94. Kondo, M. et al. Effect of concentration of photoactive chromophores on 123. Liu, D. & Broer, D. J. Self-assembled dynamic 3D fingerprints in liquid-crystal
photomechanical properties of crosslinked azobenzene liquid-crystalline coatings towards controllable friction and adhesion. Angew. Chem. Int. Ed.
polymers. J. Mater. Chem. 20, 117–122 (2010). 53, 4542–4546 (2014).

NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials 1097

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEW ARTICLE NATURE MATERIALS DOI: 10.1038/NMAT4433
124. Liu, D., Bastiaansen, C. W. M., den Toonder, J. M. J. & Broer, D. J. Photo- 140. Qin, H. & Mather, P. T. Combined one-way and two-way shape memory in a
switchable surface topologies in chiral nematic coatings. Angew. Chem. Int. Ed. glass-forming nematic network. Macromolecules 42, 273–280 (2009).
51, 892–896 (2012). 141. Ahn, S-k. & Kasi, R. M. Exploiting microphase-separated morphologies
125. Liu, D., Bastiaansen, C. W. M., den Toonder, J. M. J. & Broer, D. J. (Photo-) of side-chain liquid crystalline polymer networks for triple shape memory
thermally induced formation of dynamic surface topographies in polymer properties. Adv. Funct. Mater. 21, 4543–4549 (2011).
hydrogel networks. Langmuir 29, 5622–5629 (2013). 142. Schuhladen, S. et al. Iris-like tunable aperture employing liquid-crystal
126. Stumpel, J. E., Broer, D. J. & Schenning, A. P. H. J. Stimuli-responsive photonic elastomers. Adv. Mater. 26, 7247–7251 (2014).
polymer coatings. Chem. Commun. 50, 15839–15848 (2014). 143. Chen, M. et al. Photodeformable CLCP material: study on photo-activated
127. Liu, D. & Broer, D. J. Liquid crystal polymer networks: preparation, properties, microvalve applications. Appl. Phys. A 102, 667–672 (2011).
and applications of films with patterned molecular alignment. Langmuir 144. van Oosten, C. L., Bastiaansen, C. W. M. & Broer, D. J. Printed artificial cilia
30, 13499–13509 (2014). from liquid-crystal network actuators modularly driven by light. Nature Mater.
128. de Haan, L. T. et al. Accordion-like actuators of multiple 3D patterned liquid 8, 677–682 (2009).
crystal polymer films. Adv. Funct. Mater. 24, 1251–1258 (2014). 145. Hiscock, T., Warner, M. & Palffy-Muhoray, P. Solar to electrical conversion via
129. Modes, C. D., Bhattacharya, K. & Warner, M. Disclination-mediated thermo- liquid crystal elastomers. J. Appl. Phys. 109, 104506 (2011).
optical response in nematic glass sheets. Phys. Rev. E 81, 060701 (2010). 146. Li, C., Liu, Y., Huang, X. & Jiang, H. Direct sun-driven artificial heliotropism
130. Modes, C. D. & Warner, M. Blueprinting nematic glass: systematically for solar energy harvesting based on a photo-thermomechanical liquid-crystal
constructing and combining active points of curvature for emergent elastomer nanocomposite. Adv. Funct. Mater. 22, 5166–5174 (2012).
morphology. Phys. Rev. E 84, 021711 (2011). 147. Yin, R. et al. Can sunlight drive the photoinduced bending of polymer films?
131. Modes, C. D. & Warner, M. Responsive nematic solid shells: topology, J. Mater. Chem. 19, 3141–3143 (2009).
compatibility, and shape. EPL 97, 36007 (2012). 148. Camargo, C. J. et al. Batch fabrication of optical actuators using
132. Modes, C. D. & Warner, M. The activated morphology of grain boundaries in nanotube-elastomer composites towards refreshable Braille displays.
nematic solid sheets. Proc. SPIE 8279, 82790Q (2012). J. Micromech. Microeng. 22, 075009 (2012).
133. Modes, C. D., Warner, M., Sanchez-Somolinos, C., de Haan, L. T. & Broer, D. 149. Torras, N. et al. Tactile device based on opto-mechanical actuation of liquid
Mechanical frustration and spontaneous polygonal folding in active nematic crystal elastomers. Sensor. Actuat. A 208, 104–112 (2014).
sheets. Phys. Rev. E 86, 060701 (2013). 150. Zupan, M., Ashby, M. F. & Fleck, N. A. Actuator classification and selection —
134. de Haan, L. T., Sanchez-Somolinos, C., Bastiaansen, C. M. W., the development of a database. Adv. Eng. Mater. 4, 933–940 (2002).
Schenning, A. P. H. J. & Broer, D. J. Engineering of complex order
and the macroscopic deformation of liquid crystal polymer networks.
Angew. Chem. Int. Ed. 51, 12469–12472 (2012). Acknowledgements
135. McConney, M. E. et al. Topography from topology: photoinduced T.J.W. acknowledges the support of the Air Force Office of Scientific Research and of the
surface features generated in liquid crystal polymer networks. Adv. Mater. Materials and Manufacturing Directorate of the Air Force Research Laboratory.
25, 5880–5885 (2013).
136. Pei, Z. et al. Mouldable liquid-crystalline elastomer actuators with exchangeable
covalent bonds. Nature Mater. 13, 36–41 (2014).
137. Ware, T. H., McConney, M. E., Wie, J. J., Tondiglia, V. P. & White, T. J. Voxelated Additional information
liquid crystal elastomers. Science 347, 982–984 (2015). Reprints and permissions information is available online at www.nature.com/reprints.
138. Yang, Z., Huck, W. T. S., Clarke, S. M., Tajbakhsh, A. R. & Terentjev, E. M. Correspondence and requests for materials should be addressed to T.J.W and D.J.B.
Shape-memory nanoparticles from inherently non-spherical polymer colloids.
Nature Mater. 4, 486–490 (2005).
139. Ahir, S. V., Tajbakhsh, A. R. & Terentjev, E. M. Self-assembled shape-memory Competing financial interests
fibers of triblock liquid-crystal polymers. Adv. Funct. Mater. 16, 556–560 (2006). The authors declare no competing financial interests.

1098 NATURE MATERIALS | VOL 14 | NOVEMBER 2015 | www.nature.com/naturematerials

© 2015 Macmillan Publishers Limited. All rights reserved

You might also like