You are on page 1of 98

3 Photo-Stimulated Phase

Transformations of LCs &


Photocontrollable LC Actuators II
1. Introduction
2. Photomodulation of LCs
3. LMWLC Actuators
4. LCP Actuators
5. LCE Actuators
7. Outlook

References
* Photocontrollable Liquid-Crystalline
Actuators, Adv. Mater. 2011, 23,
21492180
Haifeng Yu and Tomiki Ikeda
* Q. Li, Liquid Crystal Beyond Display
By Prof. Chia-Rong Lee
Liquid crystal Photonics lab.

4. LCP Actuators
With birefringent mesogens in side chain or main chain,
LC polymers (LCPs) integrate LC properties with highperformance polymer materials possessing a filmforming () nature, high processability (), easy
fabrication characteristics, high corrosion () resistance,
low manufacturing costs, and high stability.
Introduction of photoresponsive mesogens into LCPs
(photoresponsive LCPs) provides the designed materials
with two most important photocontrolled features
photochemical phase transition
photocontrolled alignment

Photoresponsive LCPs have been widely explored as


promising photonic materials due to their controllable
properties in response to light.
For photocontrollable actuators, side-chain LCPs are
usually superior to main-chain ones because the latter
usually exhibits lack of molecular mobility by confining
mesogens in macromolecular main chain.

4.1. Photochemical Phase Transition


4.1.1. Copolymers and Polymer Composites
Wendorff et al. (1987) have first light-induced a change
in LC copolymers containing AZ moieties () and
mesogenic groups for photonic applications.

liquid-crystalline polymer containing


mesogenic
groups
of
the
azobenzene type in the side chain

30o

103o

Ikeda et al. (1988) reported the photochemical phase


transition in LMW AZ/LCP guest-host composites. They
demonstrated that the photochromic composites
underwent a N-I phase transition upon photoirradiation to
cause trans-cis isomerization of the AZs, and the isotropic
mixture reverted to the initial NLC phase with cis-trans
back isomerization.

LMW AZ

LCP

BMAB

Although the first example of the photochemical phase


transition was demonstrated in the LMW AZ/LCP guesthost systems, it became obvious that LC copolymers
with AZ moieties are more advantageous than composites of
LMW AZ/LCP gust-host mixtures
because of the low solubility of the guest dye molecules
(< a few mole%) in polymer matrix.
in some cases phase separation occurs when the
concentration of the guest molecules is high, which
decreases the optical performances of materials.

Photonic applications of LCPs originate from their photomanipulation of optical features by actinic light. There
are many factors influencing the properties of LCPs, and
the responsive speed in the optical processing is one of
the most important parameters among them.
In the AZ/LMWLC mixtures, the N-I phase transition
occurred in 100 ms as verified by the loss of
birefringence of the sample.
The NLC-isotropic phase transitions in AZ/LCP
composites were also found to take place in 50200 ms
in AZ-doped LCPs and AZ-containing LC copolymers.

4.1.2. Homopolymers
In the process of photochemical phase transition, LC
copolymers need a relatively longer time for orientational
relaxation of mesogens because of their higher viscosity
comparing with LMWLCs.
To overcome this difficulty, a new system has been
developed, in which each mesogen in LCPs is provided
with a photoresponsive moiety.
<ex> In AZ-containing LCPs, the AZ moiety could play
both roles as a mesogen and as a photoresponsive
group when it is connected with polymer main with a
soft spacer (Fig. 9A)

Fig. 9A LCPs with AZ moieties in the side chain shows a


very stable N phase between the glass transition
temperature (Tg) of the LCP, and the clearing point (~150
oC). These AZs show an LC phase only when they are in
the trans form, and they never show any LC phases in the
cis form.

Figure 9A
Photoinduced phase transition of LCPs in
which AZ plays both roles of a
photoresponsive group and a mesogen. A)
Chemical structures of typical AZcontaining LCPs.

Science

Optical Switching
and Image Storage
by
Means
of
Azobenzene iquidCrystal
Films,
Science
(1995)

268,

1873

Rapid response
200s

Optical image storage and data processing in PA6AB2


> several
years
24hr

T>Tg

T<Tg

T<Tg

Polymeric azobenzene LC films have two advantages:


show a much wider temperature range for N phase,
and so a wide temperature range is available for optical
switching.
have a Tg, so that they can function as image storage
materials when they are operated at T below Tg.

Fig. 9B Since the photoresponsive AZ group is the only


mesogen in the LCP, it was observed that photochemical
phase transition was induced on the same timescale as
photochemical reactions of the photoresponsive moiety in each
mesogen, when the photochemical reactions of a large number
of mesogens are induced simultaneously by means of a short
laser pulse.
It was found that PA6AB2 undergo N-I phase transition
in 200 ns upon irradiation with a laser pulse (20 ps).
This fast response
of LCPs is an
Figure 9B
Photoinduced phase transition
encouraging result
of LCPs in which AZ plays
both
roles
of
a
from the viewpoint
photoresponsive group and a
mesogen. B) Time-resolved
of applications in
measurement
of
quick
response by means of
photonic devices.
photoinduced
phase transition.

In some photonic applications as all-optical switching and


dynamic holography, it is necessary to rapidly induce not
only LCI phase transition, but also ILC phase transition
(recovery of the initial phase).
* When photoirradiation is stopped after the LCI phase
transition is induced, the initial LC phase is restored after
some time because of thermal cistrans backisomerization due to the thermal instability of cis-AZs.
* Generally, the relaxation time of LCP is strongly dependent
on some parameters such as
1 experimental temperature
2 kinds of AZs
3 structure of polymer main chain

Reflection-Mode Analysis
In the measurement of photoresponse
transmission mode is generally used.

of

LCPs,

Similarly, a novel reflection mode was developed, which


also shows a fast photochemical phase transition in AZcontaining LC materials.
* The optical switching in reflection mode was found to be
repeatable over 15000 cycles, which is 10 times more
fatigue-resistant than that in transmission mode.

Rapid Optical Switching by


Means of Photoinduced
Change in Refractive Index
of
Azobenzene
Liquid
Crystals
Detected
by
Reflection-Mode Analysis
(J. Am. Chem. Soc., Vol. 119, 1997, 7791)

AZ-LC

Figure 2. Schematic
illustration of principle
of
reflection-mode
analysis and optical
setup
for
the
measurement. na & nb:
indices of the quartz
substrate and the
sample, respectively.

Figure 3. Birefringence
azobenzene LCs

of

Principle of Reflection-Mode Analysis

Reflectivity, which is a fraction of light reflected at the interface, changes


as the change in the refractive index of the sample,
The author measured the
reflectivity at a fixed incident
angle of 71, at which the
reflectivity and its change on the
photo-induced isothermal phase
transition were significantly large.

Figure 4. Schematic illustration of various


conditions for orientations of LC molecules with
respect to the incident probe light, either spolarization or p-polarization:

Change in Reflectivity

It was found that the reflectivity


changed on pulse irradiation at
355 nm under all conditions (I~VI).

nenng, R

nonng, R

rise~s
decay~ms

Figure 5. Time-resolved
measurements of change in
the reflectivity for BMAB on
pulse irradiation (355 nm, 10
ns, 20 mJ/cm2) at 35oC: (A)
the s-polarized probe light; (B)
the p-polarized probe light.
The measurement conditions
for A and B correspond to
those of I and II shown in
Figure 4, respectively.

Figure 6. Time-resolved measurements of change in the


reflectivity for 8AB8 on pulse irradiation (355 nm, 10 ns, 20
mJ/cm2): (A) 94oC (nematic phase in super cooled state); (B) 90oC
(nematic phase in super cooled state). The measurements of A
and B were performed under the conditions of V and VI in Figure 4,
respectively.

Rise in Optical Switching

To discuss the optical switching


behavior quantitatively, we
defined the rise time (r) as the
time required to raise the
intensity of the reflected light to
90% of the maximum value.

The values of r in these


measurements are smaller by 1~2
orders of magnitude than those of
NLCs to the change in the electric
field. Such an advantageous feature
is quite favorable for application to
the optical devices.

10 mJ/cm2

5 mJ/cm2
3 mJ/cm2

Figure 8. Temperature dependence of r for BMAB


(A) and 8AB8 (B) at various laser powers: () 3
mJ/cm2; ( ) 5 mJ/cm2; ( ) 10 mJ/cm2.

Temp.-dependences of r on T & IUV


Although the difference in the chemical
structure and TNI is large between these two
samples (BMAB and 8AB8), no difference was
observed in the response time (100~200s).
This implies that the response time does not
depend on the absolute temperature (T) but on
the reduced temperature (defined as T/TNI)
because the property of LCs is significantly
influenced by the reduced temperature rather
than the absolute temperature in general.
The decrease of r at lower temperatures has
been observed only in the reflection-mode
analysis and has never been observed in the
transmission-mode analysis (because of the
difference of mechanism of optical switching
between the two modes).

10 mJ/cm2

5 mJ/cm2
3 mJ/cm2

TNI ~ 45oC

TNI ~ 110oC
Figure 8. Temperature dependence of
r for BMAB (A) and 8AB8 (B) at
various laser powers: (4) 3 mJ/cm2; (0)
5 mJ/cm2; (O) 10 mJ/cm2.

N-I phase transition


Mechanism of the N-I phase transition:
production of the cis form on photoirradiation
disorganization of the LC phase by the curve
cis form.
We can explain the decrease in r with a
decrease in the temperature at the same IUV in
terms of the local phase transition.
The decrease in the degree of the N-I phase
transition may be due to a decrease in the
domains in which the N-I phase transition was
induced. It is considered that, in this state of the
local phase transition, change in the refractive
index is small, and the apparent decrease of r
was observed at lower T.
Similarly, when less cis forms are produced
at lower laser powers at constant T, only the
local N-I phase transition is induced at lower IUV.

10 mJ/cm2

5 mJ/cm2
3 mJ/cm2

Figure 8. Temperature dependence of


r for BMAB (A) and 8AB8 (B) at
various laser powers: (4) 3 mJ/cm2; (0)
5 mJ/cm2; (O) 10 mJ/cm2.

Local phase transition

It is clearly that the degree of the N-I


phase transition is equal to 100% at high
temperatures, while it decreases with a
decrease in temperature at any laser
power. This behavior indicates that the N-I
phase transition is induced incompletely at
low temperatures. We suggest a local N-I
phase transition at low temperatures.

Figure 9. Temperature dependence of degree of the


N-I phase transition for BMAB at various laser powers
under the condition of III shown in Figure 4. As shown
in A, we defined the degree of the N-I phase
transition by the ratio of the maximum reflectivity after
pulse irradiation to the reflectivity at the I phase.
Panel B shows the degree of the N-I phase transition
at various laser powers: () 3 mJ/cm2; ( ) 5 mJ/cm2;
() 10 mJ/cm2.

Figure 10. Schematic illustration for the state of the local N-I phase
transition: filled area, I phase; unfilled area, N phase. In the state of
the local phase transition, we obtain the refractive index between
the N phase and the I phase.

Decay in Optical Switching

To discuss the optical switching behavior, it is


necessary to evaluate not only the rise time,
which means the N-I phase transition, but also
the decay time, which corresponds to I-N phase
transition. We defined the decay time (d) as the
time required to decrease the intensity to 10% of
maximum value.
It was found that the I-N phase transition,
namely the recovery of the initial N phase,
completed in only 1 ms in BMAB and 8AB8. This
value is smaller by 6 orders of magnitude than
that obtained (?hrs for BMAB, ~30s for 8AB8) in
the transmission-mode analysis. The value of d
is much different between the reflection-mode
measurement and the transmission-mode
measurement, while no difference in r has been
observed between these two methods.

Figure
11.
Temperature
dependence of d in BMAB (A)
and 8AB8 (B) at various laser
powers: () 3 mJ/cm2; ( ) 5
mJ/cm2; (O) 10 mJ/cm2.

Switching Mechanism

The molar extinction coefficients of the


azobenzene moieties at 355 nm are so large
(~104) that the pumping light at 355 nm is
absorbed entirely in the surface. Thus, the
trans-cis photoisomerization is induced near
the surface and the N-I phase transition
occurs only in the surface, leaving the bulk
area intact as an N phase. In the reflectionmode analysis, the probe light can penetrate
only in the surface region.
A pulse irradiation at 355 nm brings about
the N-I phase transition only in the surface
with the rest of the sample remaining in the
N phase. The molecules of the cis form
produced in the surface instantly diffuse into
the bulk region due to concentration gradient
of the molecules of the cis form between the
surface region and the bulk region.

Figure 12. Switching mechanism in the


reflection-mode analysis: (A) an initial N
phase before pulse irradiation; (B) an
induction of the N-I phase transition by the
laser pulse; (C) recovery of the initial N
phase only in the surface region through
the diffusion and reorientation processes.

Similarly the molecules of the trans form in


the bulk region diffuse into the surface
region and reorient on the substrate in the
initial manner. Since the diffusion and the
reorientation processes are much quicker
than the cis-trans back-isomerization
process, the rapid optical switching has
been achieved successfully in the
reflection-mode measurements.

Figure 12. Switching mechanism in the


reflection-mode analysis: (A) an initial N
phase before pulse irradiation; (B) an
induction of the N-I phase transition by the
laser pulse; (C) recovery of the initial N
phase only in the surface region through
the diffusion and reorientation processes.

4.2. Photoalignment
Weigert effect

Figure 10A
Photoinduced alignment of LCPs. A)
Photoalignment of AZ-containing LCPs
with linearly polarized light.

Trans-AZs exhibit transition


moments approximately parallel to the
molecular long axis; thereby the trans-AZs
show angular-dependent absorption of the
actinic light.
Accoring to Weigert effect, AZs can be
3D photoalignment. Trans-AZ molecules
with their transition moments parallel to the
polarization direction of LPL are effectively
activated to their excited states, followed by
trans-cis isomerization, while molecules with
their transition moments perpendicular to the
polarization direction of actinic light are
inactive towards isomerization.

4.2. Photoalignment

Figure 10A
Photoinduced alignment of LCPs. A)
Photoalignment of AZ-containing LCPs
with linearly polarized light.

After repetition of trans-cis-trans


isomerization cycles, once trans-AZ
molecules are perpendicularly to the
polarization direction of the actinic light, they
become inert towards the incident radiation.
This means that at the end of the multicycles of isomerization, there will be a net
population
of
trans-AZs
aligned
perpendicularly to the light polarization,
which is well known as the Weigert effect.
Their photoinduced alignment can be
erased thermally and photochemically, and
re-induced by irradiation with appropriate
LPL.

Figure 10A
Photoinduced alignment of LCPs. A)
Photoalignment of AZ-containing LCPs
with linearly polarized light.

Such a photoinduced molecular


alignment has been extended to LC
copolymer systems with both mesogens
and AZ moieties in the side chains.
Several studies clearly showed that nonphotoactive mesogens could undergo
reorientation concomitantly with AZ
moieties above Tg owing to their MCM.
Recent works on the photoalignment
behavior of side-chain LCPs have revealed
that dipole-dipole interaction affecting
MCM is crucial in the reorientation process.

Photoinduced optic axis

Figure 10B

When the AZ moieties are aligned


with the molecular long axis along the
propagation direction of the actinic
light,
photoisomerization
occurs
difficultly, because the propagation
direction of light is always
perpendicular to its electric field vector.
In the case of unpolarized light, only
the
propagation
direction
is
perpendicular to the electric vector of
the light. Thus, when unpolarized light
is employed, it is expected that the AZ
moieties become aligned only in the
propagation direction of the actinic
light. In fact, several works have been
reported concerning the out-of-plane
alignment
behavior
by
using
unpolarized light.

Photoinduced optic axis

Figure 10B

From the viewpoint of 3D


manipulation of molecules by light,
promising results have been obtained
for LCPs: by changing the incident
angle of the actinic unpolarized
light, one can control the alignment of
AZ and mesogenic moieties in a 3D
fashion, precisely along the
propagation direction of the actinic
light at room temperature.
It has been also confirmed that the
induced 3D alignment of molecules is
very stable below Tg of the LCPs.
These results are expected to open a
new methodology to photomanipulate molecules in a 3D
manner.

4.3. Photoinduced Large Change in Birefringence


Optical birefringence can be produced by change in
transition moments of chromophores by photoalignment.
Early work in this field was carried out on LPL-induced
change in alignment of AZs embedded in amorphous
polymers. The first report about the generation of
birefringence was employed with AZ-doped PVA films (1).
Although many research focused on the AZ-doped
polymers, the stability of the induced anisotropy was low.
In amorphous polymers covalently attached with an AZ
moiety (2), optical anisotropy was also observed upon
LPL irradiation. These polymers showed stable
photoinduced birefringence at T<Tg. The induced
birefringence could be erased thermally and photochemically,
which can be regenerated upon LPL irradiation. But the
formed birefringence is small due to the amorphous state,
which cannot meet the requirement of photonic
applications.

A large and stable (T<Tg) photo-induced birefringence


can be induced in AZ-containing LCPs (3) by
manipulating the LC alignment, which is more efficient
than that of amorphous polymers. Moreover, the
alignment of AZs can be photo-manipulated in a 3D way,
whose ordering can be then transferred to other photoinactive mesogens by MCM.
It is well known that tolane () is one of the
most common core structures for design of highly
birefringent LC molecules because of its longer molecular
conjugation length. Using copolymerization method, LC
copolymers containing both tolane and AZ moieties in
the side chain have been developed, and a large change in
birefringence was obtained in a homogeneously aligned
state. To increase the photosensitivity and induce a larger
birefringence for photonic applications, a concept of
molecular architecture of AZ-containing LC materials
was proposed.

A tolane group is directly attached


onto the 4- or 4- position of an AZ
molecule to prepare an azotolane
mesogen, which shows far longer
molecular conjugation length than one
single tolane group or AZ moiety.

Fig. 11. Molecular architectures of


LCPs containing an azotolane
mesogen for photoinduced a large
change in birefringence.

All the prepared azotolane LCPs showed a wide nematic


LC range and exhibited a phase transition temperature higher
than 200 oC. In general, a NLC phase is quite
advantageous for photonic applications because its
sensitivity to external stimuli is higher than that of other LC
phases. Upon irradiation with actinic light, a huge
birefringence larger than 0.35 at 633 nm was obtained by the
photochemical phase transitions of the azotolane materials with
three benzene rings (P3AT, P3TA, P3AT-NO2).

Furthermore, the large birefringence can be obtained even


at the telecommunication wavelength (1550 nm) by the
photoinduced phase transition from a nematic LC to an
isotropic state.
A more larger birefringence ( >0.65 at 633 nm) was
obtained for the azotolane LCPs with a longer molecular
conjugation length (four benzene rings, P4TTA, P4TAT,
P4ATT) by photoinduced change in alignment of azotolane
mesogens.
The largest birefringence of about 0.76 was obtained for
the LCPs possessing the longest molecular length of an
azotolane mesogen with five benzene ring (P5ATTA). P5ATTA
with two AZ units in one azotolane mesogen also showed
the most efficient change in alignment, indicating that the
high birefringence in the homogeneously aligned state
(neno) could be efficiently converted to an change in
birefringence.

4.4. Holographic Applications


Holography is a unique technique that enables
concomitant recording of both phases and amplitudes of
light waves. The most fascinating feature of holography is
that it can record and display a complete 3D image of an object.
In holography, the phase and amplitude of light waves
are recorded by periodic alternation of physical properties
of materials.

According to the manner of recording of interference


patterns, holograms are mainly classified into two types.
One is an amplitude-type hologram, in which the
interference pattern is recorded as a density variation in
recording media.
The other is a phase-type hologram, in which fringe
patterns are recorded as a change in surface structure or
refractive index.
Theoretically, diffraction of phase-type holograms is always
higher than diffraction of amplitude-type ones.

Fig. 12. Holographic gratings recorded in LCPs. A) Classification of holograms.

AZ-containing LCPs with photoresponsive functions can


be easily modulated into hierarchical patterns by
adjusting the input light with wavelength, intensity,
polarization, phase, interference, and so on.
Therefore, holograms can be recorded in AZ-containing
LC materials by photoinducing an orientation change of
LC molecules in a periodic pattern obtained from
interference of two coherent beams, the object and the
reference beams. Since a large refractive-index
modulation can be obtained in AZ-containing LC materials,
the obtained hologram is phase-type.
Wendorff et al. showed for the first time that holographic
gratings can be inscribed in LCPs composed of AZ
moieties and mesogenic groups.

Although the LMWLC AZs show good mobility of


mesogens, it is often difficult to record holographic
gratings with narrow fringe spacing (i.e., high resolution)
and high stability because of the high mobility of
mesogens. As far as the stability is concerned, AZcontaining LCPs are superior because of their restricted
mobility of mesogens at a temperature below their Tg.
By photochemical phase transition and photoinduced
change in LC alignment, phase-type holograms have been
inscribed in side-chain LCPs containing AZ moieties.
Since the photochemical phase transition of LCPs
containing an AZ moiety can be reversibly and quickly
carried out, formation and highly repeatable recording of
holograms have been achieved in nanosecond timescales
and reversibly. In fact, it has been confirmed that the
holographic gratings are formed in AZ-containing LCPs in
~200 ns with a laser pulse (pulse duration = 20 ps).

Fig. 12C

* Owing to spatial modulation of molecular alignment in the interference pattern, an


alternating arrangement of LC and isotropic phases can be clearly observed with POM.
Such a system has been proved capable of holographic recording of 3D objects with high
resolution.

Fig. 12C. A refractive-index grating with


alternating patterned domains of LC and
isotropic phases is under a surface-relief
grating.

Fig. 12C.
Holographic recording
of 3D objects and
reconstructed images.

5. LCE Actuators
LCEs can be prepared by crosslinking conventional LCPs
into a network, which integrates
mechanical properties of elastomers
regular ordering of LC materials
by the network topologies.
In 3D crosslinked LCEs, the initial ordering of mesogens
can be fixed by the crosslinkers, which might give rise to
quick change in shape due to a fast orderdisorder
transition, induced by slight changes in the orientational
order of mesogens.
Combination of the anisotropic aspects of LC phases and
the rubber elasticity of polymer networks enables LCEs to
show unique feature.

A large deformation can be induced in response to


external stimuli such as temperature, electricity, pH or
humidity.
When a chromophore (e.g., an AZ) as a mesogen is
introduced into LCEs, shape and volume change in
response to light can be produced, which can directly
convert light energy into mechanical power.
Photomechanical and photomobile properties can be
obtained by the
photochemical phase transition
photoinduced contraction and expansion
Photoinduced 3D motions

5.1. Photochemical Phase Transition


LCEs combine characteristics of two systems, rubbery
elastomers with 3D network and self-organized LCPs with
LC ordering, which have been regarded as a novel class
of soft materials.
In 1975, de Gennes et al. first proposed the concept of
LCEs as artificial muscles by taking advantage of their
substantial uniaxial contraction in the direction of the
director axis. Then he theoretically predicted the
possibility of a large deformation of LCEs induced by
phase transition.
Finkelmann et al. pioneered in the preparation of LCEs
with mesogens.

The crosslinking density has a great influence on the


macroscopic properties and the phase structures of
LCEs with crosslinked networks.
1 On one hand, the mobility of mesogenic segments is
reduced with an increase in density of cross-linking
points, and consequently the mobility of mesogens in
the vicinity of a cross-link is suppressed.
2 On the other hand, such confinement from crosslinking
may bring about larger change than a linear LCP in
response to external stimuli. In fact, a linear LCP can be
regarded as a special LCE with a crosslinking density of
zero.

5.2. Photoinduced Contraction and Expansion


Confining mesogens in crosslinked networks, LCEs usually
show thermoelastic properties:
They contract along the alignment direction of mesogens
upon thermally-induced LC-isotropic phase transition, and
expand upon cooling below the phase transition
temperature.
If photoresponsive mesogens such as AZ are incorporated
into LCEs, photo-induced 2D motions such as contraction
and expansion are expected by the above-mentioned
photochemical phase transition (or photochemically
induced decrease in LC ordering).

Fig. 2A
* An AZ molecule exhibits a large change in molecular configuration upon photoisomerization,
and the distance between the 4- and 4- carbons in benzene rings contracts from 9.0 of
trans -AZ to 5.5 of cis-AZ. Such a photoinduced molecular change results in a large
contraction ratio, which arouses intensive interest of scientists.
* Eisenbach first reported that amorphous polymers crosslinked with AZs contracted upon UV
irradiation to induce trans-cis isomerization, while it expanded by irradiation with visible light,
which caused cis-trans back-isomerization. However, the observed contraction was only 0.15
0.25%, which is too small from the viewpoint of practical applications.

Different from amorphous elastomers, photoinduced deformation has


been greatly enhanced in LCEs.
* In 2001, Finkelmann et al. succeeded in inducing a contraction ratio of 20% in LCEs with
polysiloxane main chains and AZ chromophores at the crosslinks upon UV exposure to cause
the trans-cis isomerization of AZs. After irradiation was stopped, the elastomers thermally
returned to the original state due to cis-trans back-isomerization of the AZ mesogens.
* As far as photomechanical effects are concerned, the subtle variation in nematic LC order
upon trans-cis isomerization of the AZ mesogens causes a large uniaxial deformation along
the director axis when the LC molecules are strongly associated by covalent crosslinking to
form a 3D polymer network.
(monomer: main chain)
(LC 1: side chain 1)
(LC 2: side chain 2)
(non-LC: side chain 3)
(cross-linker)
(photoisomerizable cross-linker)

(monomer: main chain)


(LC 2:
side
chain 2)
(cross-linker)

(LC 1: side chain 1)

5.3. Photoinduced 3D Motions


Although LCEs show excellent 2D motional properties of
contraction and expansion, photoinduced 3D motions
with diverse of ways such as bending, twisting and
rotation are expected from the viewpoint of practical
applications.
It is expected that photomechanical and photomobile
performances with a quick response to actinic light can
be obtained in LCEs composed of only AZs as
photoresponsive mesogens.

5.3.1. Photoinduced Bending


Fig. 14AB

Fig. 14 Photoresponsive properties of freestanding


LCE films. A) Scheme of preparation of freestanding
LCE films with an AZ LC monomer and a
crosslinker.
B)
Plausible
mechanism
of
photoinduced bending of monodomain LCE films.

*
Photoresponsive
behavior
of
monodomain-aligned LCE films was first
reported by Ikeda et al.
* The LCE films were prepared by in-situ
photopolymerization of an AZ-containing
LC monomer and a diacrylate with an AZ
moiety in a glass cell, pre-coated with
rubbed PI films as alignment layers.
* The obtained freestanding films showed
bending and unbending behaviors induced
by irradiation with unpolarized UV and then
visible light, respectively.
* It was observed that the monodomainaligned LCE film bent toward the irradiation
direction of the incident UV light along the
rubbing direction, and the bent film reverted
to the initial flat state after exposure to
visible light.

surface
region

Fig. 14B
* Since the molar extinction coefficient of AZ moieties at around 360 nm is large, photons are
absorbed only in the surface of LCE films (a thickness of several tens of m) and trans-cis
photoisomerization takes place only in the surface area of these freestanding LCE films,
which leads to an analogous bilayer structure with mesogenic slabs of two different polymers
that respond differently, and bending can be obtained like a bimetal way (
). As a result, the volume contraction is generated only in the surface layer, thus
causing the LCE films to bend toward the light source.
* Since only a part of the polymer films is involved in deformation in these materials, one can
induce bending much faster than other modes of photoinduced deformations. It has been
demonstrated that both the alignment ordering of AZ mesogens and crosslinking density
strongly influence the bending performances of the AZ-containing LCE films.

Anisotropic bending &


unbending behavior of
azobenzene LC gels by
light exposure (Adv.
Mater. 2003)

Fig. 14C Photographs of precise


control of the bending direction of
polydomain LCE films by linearly
polarized light (left) and the plausible
mechanism (right).

Fig. 14C

* A variety of LCE films with different alignments of AZ mesogens were prepared and examined
to elucidate the influence of the alignment on the photoinduced bending behavior.
* Ikeda et al. succeeded in achieving a photoinduced directioncontrollable bending in single
polydomain LCE films.
* Only by changing the polarization direction of the actinic light, the bending of polydomain LCE
films can be induced repeatedly and precisely along any chosen direction, and they bent toward
the irradiation direction of the incident light with bending occurring parallel to the direction of
LPL.

Fig. 1B Photomechanical effect in LCEs. The


homogenous alignment of LCs produces
bending towards the actinic light, whereas the
homeotropic alignment of LCs causes bending
away from the light source.

Fig. 1B

* Interestingly, homeotropically-aligned LCE films were found to show a completely different


bending behavior; they bent away from the actinic light source upon UV exposure.
* Furthermore, LCE films with hybrid alignment mesogens were prepared, in which a
homogenous alignment was on one surface and a homeotropic alignment was on the other
surface. Upon irradiation with unpolarized UV light on the homogeneous surface, the hybrid
LCE films bent toward the light source along the alignment direction, whereas the film bent
away from the light source when the homeotropic surface was irradiated.
* Furthermore, upon irradiation from both surfaces of the films, the bending speed was greatly
enhanced at the same time comparing with homogeneously or homeotropically aligned LCEs.
* Recently, Broer et al. prepared LCE films with a twisted configuration of AZ moieties,
showing a large amplitude bending motion and a large amplitude coiling motion upon
exposure to UV light, which arises from the 90o twisted LC alignment configuration.

Large amplitude light-induced


motion in high elastic modulus
polymer actuators (J. Mater.
Chem. 2005)
Broer et al. prepared LCE films with a
twisted configuration of AZ moieties,
showing a large amplitude bending
motion and a large amplitude coiling (
) motion upon exposure to UV light,
which arises from the 90o twisted LC
alignment configuration.

To fabricate the actuators, diacrylate dopants


containing azobenzene moieties (A6MA) were
blended with liquid crystalline diacrylate hosts
(C6M) and photopolymerized into crossedlinked
networks in a twisted configuration.

Fig. 7 UV-induced coiling of a film in the twisted configuration.


This film is composed of 8 wt% A6MA in a C6M host. In panel (a), the sample is pictured before UV irradiation,
while in panels (b)(e) the sample coils under UV intensity of roughly 100 mW cm-2. The sample is clamped in a
fixed position in panels (a)(e). In the final panel, (f), the film was moved nearer to the UV source (to roughly 1 cm)
in an attempt to show the maximum deformation. The film dimensions are 24 mm 2 mm 17 m.

5.3.2. Novel LCE Materials


Photomechanical Effects of Ferroelectric Liquid-Crystalline
Elastomers Containing Azobenzene Chromophores (Angew. Chem. Int.
Ed. 2007, 46, p.881)

* Photodeformable smart materials that can undergo a shape or volume


change in response to light are attracting increasing attention. On the one
hand, light is a clean energy and can be controlled rapidly and remotely; on
the other hand, by using the deformation of these materials, one can convert
light energy into mechanical energy directly (photomechanical effects).
* Most photodeformable smart materials contain
photochromic compounds (such as azobenzene). The chromophores change
their molecular structure upon exposure to light; thus, their incorporation into
polymer systems gives rise to conformation changes of the polymer chains
and concomitant changes in the physical and chemical properties of the
polymer solutions and solids through photoisomerization, including
photoinduced contractions/expansions of rubbery networks and swollen gels.
* However, the low elastic modulus and low yield strength of gels provide
important limitations in the performance of actuation, while for the solid
polymer networks, deformations of less than 10% limit their practical
applications.

* Large photoinduced contractions/expansions have been acquired by


Finkelmann et al. and other research groups by incorporating azobenzene
derivatives into liquid-crystalline elastomers (LCEs) as a trigger.
* LCEs with the rubber elasticity of polymer networks exhibit a simultaneous
anisotropic orientational symmetry of liquid-crystalline (LC) phases. The
driving force for their large changes in shape is suggested to arise from a
variation of LC alignment order: upon irradiation with UV light, LC systems
containing azobenzene chromophores experience a reduction in alignment
order and even an LC-I phase transition as a result of the transcis
photoisomerization of the azobenzene moieties, because the rodlike transazobenzene moieties stabilize the LC alignment, whereas the bent cis forms
lower the LC order parameter.
* Furthermore, a 3D deformation photoinduced bending has been
achieved by using nematic LCE films containing azobenzenes swollen in
suitable solvents or heated above their glass transition temperatures in air.
As a consequence of the limitation of absorption of photons, a surface
contraction caused by the photoinduced change in alignment of liquid
crystals contributes to the bending. Several examples of the threedimensional deformations have been reported, such as twisting of the
azobenzene-containing LCEs.

* In comparison with the contraction/expansion that is a two-dimensional


action, the bending could be advantageous for artificial hands and medical
microrobots that are capable of completing particular manipulations.
* However, the following three factors limit their actual applications:
1) a slow response in the order of minutes or seconds
2) difficulty in inducing effective bending at room temperature in air
3) the low force generated upon photoirradiation.
* In this work, they found that the bending speed was different in the nematic
LCE films with various crosslinking densities. The larger the order parameter
and the mobility of polymer segments, the faster the bending. According to
these results, if we use films with a higher order of LCs and a lower glass
transition temperature (Tg), it should be easy to acquire high-speed bending
at room temperature. It is well-known that ferroelectric liquid crystals have
two advantages:
(1) their high degree of order of mesogens,
(2) the fact that their molecular alignment can be controlled quickly by
applying an electric field, because of the presence of spontaneous
polarization.

* To achieve a better photoresponse of LCEs, FLCs were explored to


prepare LCEs because of their inherent nature of quick response and high
degree of mesogenic order. Therefore, ferroelectric LCE films with a high LC
order and a low Tg were prepared by photopolymerization of an LC monomer
and a crosslinker, which were pre-aligned under an electric field.
* UV irradiation induced the LCE films to bend at Troom toward the actinic light
source with a tilt to the rubbing direction of the alignment layer. The bending
process was completed within 0.5 s of irradiation from a laser beam,
demonstrating its quick photoresponse.
* In addition, the mechanical force generated by photoirradiation reached
about 220 kPa, similar to the contraction force of human muscles (around
300 kPa). This fast and strong mechanical response to light may lead to
potential applications of the ferroelectric LCEs in artificial muscles, microoptomechanical systems, and other photodriven mechanical devices.

Fig. 15 Novel LCE materials for photoinduced 3D motions. A) Change of the load on
ferroelectric LCE films when exposed to UV light at 366 nm with different intensities at
50 C. The right is the illustration of setup for measurement using a thermomechanical
analyzer (Simadzu, TMA-60). Both ends of a film were clamped and 20 mN was initially
loaded on the film; then the film was heated to 50oC and exposed to UV light. As the
length of the film was kept unchanged, the increase of the load indicates the generation
of the mechanical force upon photoirradiation.

Three-Dimensional Photomobility of Crosslinked Azobenzene


Liquid-Crystalline Polymer Fibers (Adv. Mater. 2010, 22, p.1361)

* It is well known that humans muscles consist of many bundles of fibers


and their anisotropic contractions are induced by electric stimuli. To construct
artificial muscles, LCE fibres were fabricated due to their highly mechanical
flexibility. LCE fibers with a highly orientational order of mesogens along the
fiber axis were fabricated. Such LCE fibres showed a good photoresponse
upon exposure to UV light. They bent toward the irradiation direction and
reverted to the initial state upon
exposure to visible light. The flexibility
of photoinduced 3D motions is greatly
enhanced since the LCE fibers can be
photo-driven towards any desired
direction.
Chemical structures of the LC
monomers (A6AB6 and A6AB6OH)
and the crosslinker (MDI)

Fig. 15 Preparation and photomechanical


properties of LCE fibres. The illustration below
shows the plausible mechanism of photo induced
3D motions with good flexible nature.

* The CLCP fibers were prepared by


two-step reactions, Firstly, the LC
monomers were polymerized by radical
polymerization. Then the obtained
copolymers were mixed with MDI, and
the mixtures were formed into fibers by
dipping a tip of a toothpick into the
mixture and pulling the mixtures with
the toothpick as quickly as possible. By
DSC measurements, it was found that
the CLCP fibers exhibited a glasstransition temperature (Tg) of around 60
oC.

The CLCP fibers show a high order of mesogens


along the fiber axis.

* When the CLCP fiber was exposed to UV light at 366 nm, the CLCP fiber
bent toward the actinic light source along the fiber axis. The bent fibers
reverted to the initial flat state upon irradiation with visible light at >540 nm.
The photoinduced bending of the CLCP fibers was reversible simply by
changing the wavelength of the actinic light, similar to that of the CLCP
films.
The CLCP fibers have a high degree of alignment of the mesogens along
the fiber axis; therefore, irradiation with UV light leads to a reduction of the
alignment order of the azobenzene mesogens along the fiber axis.
* Moreover, the extinction coefficient
of the azobenzene moieties at around
360nm is large (~2.0104 Lmol-1cm-1)
and almost all the incident photons
are absorbed at the surface region.
This means that the CLCP fibers with
a high concentration of azobenzene
moieties can generate an alignment
change only in the surface region of
the fibers upon exposure to UV light.
As a result, an uneven distribution of
the molecular alignment is produced
and the bending motion can be
induced in the fiber.

Changing the orientational order in the CLCPs leads to internal stresses and
changes of the sample shape.

The authors measured the stress generated in the CLCP fibers upon exposure to UV light by
thermomechanical analysis (TMA). It was difficult to measure the generated stress in a
single fiber due to the measurement limit of TMA. The authors bundled three pieces of the
fiber together side by side and fixed the top and bottom of the CLCP fibers using epoxy glue.
* The bundled CLCP fibers were fixed by clamping both ends of the fibers and heating to 90oC,
which is higher than the Tg value of the CLCP fiber. Initial stress was loaded onto the bundled
fibers to keep their length constant. The stretching direction was parallel to the fiber axis.
* Upon irradiation with UV light, the
generated stress increased and
reached 120 kPa and 210 kPa (a
value that is similar to the stress in
human muscles ~300 kPa). When
the light intensity was 45 and
110mWcm-2, respectively. It was
found that a higher intensity of UV
light could generate larger stress.
* Additionally, the bending time of
the fibers decreased with an
increase in light intensity because
actinic light with a higher intensity
produces a higher concentration of
cis-azo moieties and, thus, a larger
surface contraction.

* A three-dimensional control of a
bending direction in the CLCP
fibers
Since the shape of the CLCP fiber
was approximately cylindrical, the
CLCP fiber could be irradiated with
light under the same conditions from
every side. The freestanding CLCP
fiber was placed on an aluminum
block, heated with a thermocouple
to 100 oC, and irradiated with UV
and visible light. The direction of the
bending upon photoirradiation could
be controlled by changing the
irradiation direction of the actinic
light.

Palffy-Muhoray et al. demonstrated that the mechanical deformation of an LCE sample


doped with AZ dyes becomes very large in response to nonuniform visible-light illumination.
When a laser beam is irradiated from above onto a dye-doped LCE sample floating on water,
the LCE swims away from the laser beam, with an action resembling that of flatfish.

Fast liquid-crystal elastomer swims into the dark (Nature Mater. VOL 3 ,
2004, p. 307)
* Liquid-crystal elastomers are rubbers whose constituent molecules are
orientationally ordered. Their salient feature is strong coupling between the
orientational order and mechanical strain.
Changing the orientational order gives rise to internal stresses, which lead to strains
and change the shape of a sample. * Orientational order can be affected by changes
in externally applied stimuli such as light.
* This work demonstrated that by dissolving rather than covalently bonding azo
dyes into an LCE sample, its mechanical deformation in response to non-uniform
illumination by visible light becomes very large (more than 60o bending) and is
more than two orders of magnitude faster than previously reported.
Rapid light-induced deformations allow LCEs to interact with their environment in
new and unexpected ways. When light from above is shone on a dye-doped LCE
sample floating on water, the LCE swims away from the light, with an action
resembling that of flatfish () such as skates or rays. The authors first
analyse the propulsion () mechanism in terms of momentum transfer.

methylsiloxane
monomer backbone

Ar+ Laser

45

5 mm 5 mm
0.32 mm

mesogenic biphenyl sidechain

n : dashed line
Laser off

tri-functional crosslinker

azo dye Disperse Orange


Chemical structures of the nematic LCE & doped dye

Monodomain LCE samples consist of a


poly(methylsiloxane) backbone with attached
mesogenic biphenyl side-chain units. Using a trifunctional cross-linker, samples under preparation
are first weakly cross-linked, then stretched
unidirectionally to align the mesogens and
establish nematic order, and then finally fully
cross-linked to form a transparent birefringent
monodomain.

Ar+ Laser on
(514.5nm, 600mW)

Optomechanical response (bending and


relaxation dynamics ) of an LCE sample.

* To see how DDLCE materials might


interact with their environment, their
interaction with fluids are studied by
illuminating from above LCE samples
floating on water. They find, astonishingly,
that samples will swim away from the light.
The direction of the observed motion, at
least initially, is perpendicular to the
nematic director.
* This swimming behavior is remarkable,
because no linear momentum is
transferred from the radiation field to the
sample in the horizontal direction of motion.
The system is therefore a motor, where
motion is caused by the transfer of
energy, but not momentum, to the system.
Interaction of a dye-doped LCE sample with a liquid
A 5-mm-diameter disk of DDLCE sample (0.32 mm thick)
is floating nearly motionless on the surface of a water
reservoir (~2 cm in depth) when illuminated from above by
an Ar+ laser with peak intensity of 1.1 W cm2 and beam
waist of 3 mm. The sample will swim away from the light.

* Regarding momentum transfer between the LCE sample and


water during swimming, we conjecture the following
mechanism for the propulsion. If the illuminating beam is
centered on the disk, the disk contracts along the director (and
expands in the other two directions to conserve volume)
forming a saddle shape.

saddle shape

dle m
Sad

oves

m
la c e
disp

en t

saddle curve
(// n )

* If the deformed disk is slightly displaced, say by water or air


convection, capillary waves or other noise, away from the laser
spot, the shape will change again with the elevated shoulders
moving in the direction opposite to the displacement.
* This motion will push the water under the shoulders in the
same direction as the motion of the shoulders, in the direction
opposite to the displacement, and, by Newtons third law,
water will in turn push the LCE sample in the direction of the
original displacement. The force from the water will accelerate
the sample away from the light, and lead to a further
displacement in the same direction, resulting in yet greater
force from the water.

* The original configuration with the illuminated sample under


the beam is thus unstable; any in-plane perturbation in the
sample position normal to the director induce a force in the
same direction, which will push the sample from under the
illuminating beam.

saddle shape

dle m
Sad

oves

m
la c e
disp

en t

saddle curve
(// n )

* DDLCEs undergo large and rapid shape changes on


illumination. Samples floating on water swim away from the
light, because the compact deformed configuration increases
the energy of the system. The system is an unusual motor; the
light source must supply not only energy, but also positional
information, to bring about motion.

5.3.3. Photomobility of LCE-Laminated Films


Generally, the 3D photo-manipulation of LCE films can be greatly
accelerated by increasing the experimental temperature. Below Tg of
LCE materials, the photomechanical movements are usually very slow
and sometimes difficult to be observed.
By molecular design, photoresponsive behavior at Troom was obtained
in LCE films prepared with an AZ-containing monomer and a
crosslinker. Benefit from this, diverse complicated 3D movements like
an inchworm walk and a flexible robotic arm motion as well as
photo-driven plastic motor were explored very recently, which is
completely photomobile performance of LCEs.

Comparing with photoinduced 3D motions of bending and unbending


behavior, photo-driven rotation is undoubtedly more challengeable
from engineering point of view.
Ikeda et al. first developed light-induced rotation based on AZcontaining LCE films at room temperature.
* They prepared continuous rings by connecting both ends of the LCE
films, in which AZ mesogens were homogenously aligned along the
circular direction of the rings. Upon simultaneous irradiation with UV
light (from the downside right) and visible light (from the upside right),
the ring rolled intermittently towards the actinic light source, resulting
almost in a 360o roll at Troom.

Photomobile Polymer Materials: Towards Light-Driven Plastic Motors


(Angew. Chem. Int. Ed. 2008, 47, p.4986)

* Because light is a good energy source that can be controlled remotely, instantly, and precisely,
light-driven soft actuators could play an important role for novel applications in wide-ranging
industrial and medical fields.
* Liquid-crystalline elastomers (LCEs) are unique materials having both properties of LCs
and elastomers, and a large deformation can be generated in LCEs, such as reversible
contraction and expansion, and even bending, by incorporating photochromic molecules,
such as an azobenzene, with the aid of photochemical reactions of these chromophores.
* If materials absorb light and change their shape or volume, they can convert light energy
directly into mechanical work (the photomechanical effect) and could be very efficient as a
single-step energy conversion. Still, these photomobile materials would be widely applicable
because they can be controlled remotely just by manipulating the irradiation conditions.
* It is well known that when azobenzene derivatives are incorporated into LCs, the LC-I phase
transition can be induced isothermally by irradiation with UV light to cause transcis
photoisomerization, and the I-LC reverse-phase transition by irradiation with visible light to
cause cistrans back-isomerization. This photoinduced phase transition (or photoinduced
reduction of LC order) has led successfully to a reversible deformation of LCEs containing
azobenzene chromophores just by changing the wavelength of actinic light.
* Although the photoinduced deformation of LCEs previously reported is large and interesting, it
is limited to contraction/expansion and bending, preventing them from being used for actual
applications. Herein the authors report potentially applicable rotational motions of azobenzenecontaining LCEs and their composite materials, including a first light-driven plastic motor with
laminated films composed of an LCE film and a flexible polyethylene (PE) sheet.

Size of the LCE ring: 18 mm3 mm (diameter: 6 mm).


Thickness of the LCE ring: 20 m.

The LCE films were prepared by photopolymerization of a


mixture of an LC monomer containing an azobenzene
moiety (molecule 1) and an LC diacrylate with an
azobenzene moiety (molecule 2) with a ratio of 20/80
mol/mol, containing 2 mol% of a photoinitiator in a glass
cell coated with rubbed polyimide alignment layers.
The photopolymerization was conducted at a
temperature at which the mixture exhibited a smectic
phase. The glass transition temperature of the LCE films is
at about Troom, allowing the LCE films to work at room
temperature in air, as the films are flexible enough at this
temperature.

* They prepared a continuous ring of the LCE film by


connecting both ends of the film.
* The azobenzene mesogens were aligned along the
circular direction of the ring. Upon exposure to UV light
(366 nm, 200 mWcm-2) from the downside right and visible
light (>500 nm, 120 mWcm-2) from the upside right
simultaneously, the ring rolled intermittently toward the
actinic light source, resulting almost in a 360o roll at Troom.
This is the first example of this kind of photoinduced motion
in a single layer film, although the rolling of the LCE ring
used herein was slow, and stopped when the ring was
broken by irradiation.

The photoisomerization of the azobenzene


moieties, which is a trigger of the photoinduced
deformation of LCE films, occurs only in the
surface region of the films facing the incident light,
and in the bulk of the films the transazobenzene
moieties remain unchanged owing to the limitation
of light absorption by the azobenzene mesogens.

surface
region

We assume that a bilayer structure may be simply produced by


light irradiation in a single LCE film:
The first layer in the surface region of the LCE film contracts by
photoirradiation while the second layer in the bulk of the film shows
no photochemical reactions but changes its shape just by following
the first layer.
On the basis of this assumption, photoactive LCE layers are needed
only in the surface region of the films facing light sources, and the
rest of the films can be replaced by other materials

By preparing a laminated structure


composed of an LCE layer and a flexible
plastic sheet, both photoresponsive and
good mechanical properties can be
provided simultaneously in a simple
laminated film, enabling us to induce
various movements of these films by light
without deterioration () or breakages
() of the materials.
We fabricated the laminated films
composed of an LCE layer and an
unstretched low-density PE film having
good flexibility and mechanical properties at
Troom.

surface
region

LCE layer

Adhesion
layer
unstretched
low-density PE
film

To evaluate the driving force of the LCE laminated films to change their
shapes by photoirradiation, we measured the internal stress generated in the
films upon exposure to UV light with a thermomechanical analyzer.

It should be mentioned that the


LCE laminated films continued to
generate the force during
photoirradiation without breaking,
whereas the LCE single-layer
films cracked after short light
irradiation at high intensities,
owing to insufficient mechanical
strength. The above results
demonstrate the possibility of
inducing 3D motion with the LCE
laminated films by light irradiation
at Troom.

We prepared a plastic belt of the LCE laminated film by connecting both ends
of the film, and then placed the belt on a homemade pulley system.
By irradiating the belt with UV light from top right
and visible light from top left simultaneously, we
induced a rotation of the belt to drive the two
pulleys in a counterclockwise direction at Troom.
* Upon exposure to UV light, a local contraction
force is generated at the irradiated part of the belt
near the right pulley along the alignment direction
of the azobenzene mesogens, which is parallel to
the long axis of the belt. This contraction force
acts on the right pulley, leading it to rotate in the
counterclockwise direction.
* At the same time, the irradiation with visible
light produces a local expansion force at the
irradiated part of the belt near the left pulley,
causing a counterclockwise rotation of the left
pulley.
* These contraction and expansion forces
produced simultaneously at the different parts
along the long axis of the belt give rise to the
rotation of the pulleys and the belt with the same
direction.

* The rotation then brings new parts of the belt to


be exposed to UV and to visible light, which
enables the motor system to rotate continuously.
Reverse rotation of this belt could also be
induced just by changing the irradiation positions
of the UV and visible light.

With the LCE films and their composite materials, the authors have
successfully developed new photomechanical devices, including the first
light-driven plastic motor.
* They can convert light energy directly into mechanical work without the aid
of batteries, electric wires, or gears. The size of the samples is in the range of
millimeters for the demonstration, but is not in principle material-limited, so
numerous applications even on the nanoscale are possible, especially where
efficient power supply to mechanical system is battery-free and noncontact.

In addition to the photodriven rotation, the LCE-laminated films


exhibited other photomobility of sophisticated 3D motions like
inchworm () walk and a flexible robotic arm. The LCElaminated film moves forward upon alternate irradiation with UV and
visible light at room temperature. As light can be handled remotely,
instantly and precisely, these plastic materials can work as main
driving parts of light-driven actuators without the aid of batteries,
electric wires and gears.

inchworm

Photomobile polymer materialsvarious 3D movements (J. Mater. Chem.,


2009, 19, p.60)

Composition of a crosslinked
azobenzene liquid-crystalline
polymer and a flexible
polymer film can provide a
variety of simple devices that
can walk in one direction like
an inchworm and move like
a robotic arm induced by
light.

Photoinduced inchworm walk of the CLCP laminated film by alternate irradiation with UV (366 nm,
240mWcm-2) and visible light (>540 nm, 120mWcm-2) at room temperature.
Size of the film: 11 mm 5 mm
CLCP laminated part: 6 mm 4 mm.
Thickness of the layers of the film: PE, 50 m; CLCP, 18 m

Photoinduced flexible robotic arm motion of the CLCP laminated film by alternate irradiation with UV
(366 nm, 240mWcm-2) and visible light (>540 nm, 120mWcm-2) at room temperature.
Size of the film: 34 mm 4 mm
CLCP laminated part: 8 mm 3 mm & 5mm 3mm
Thickness of the layers of the film: PE, 50 m; CLCP, 16 m

copolymers
A heteropolymer or copolymer is a polymer derived from
two (or more) monomeric species, as opposed to a
homopolymer where only one monomer is used.
Since a copolymer consists of at least two types of
constituent units, copolymers can be classified based on
how these units are arranged along the chain. These
include:
(1) Alternating copolymers with regular alternating A and B
units ( )
(2) Periodic copolymers with A and B units arranged in a
repeating sequence (e.g. (A-B-A-B-B-A-A-A-A-B-B-B)n)

(3) Statistical copolymers are copolymers in which the


sequence of monomer residues follows a statistical rule. If
the probability of finding a given type monomer residue at a
particular point in the chain is equal to the mole fraction of
that monomer residue in the chain, then the polymer may
be referred to as a truly random copolymer ( ).
(4) Block copolymers comprise two or more homopolymer
subunits linked by covalent bonds ( ). The union of the
homopolymer subunits may require an intermediate
subunit,
known
as
a
junction
block.
Block
copolymers with two or
three distinct blocks are
called diblock copolymers
and triblock copolymers,
respectively.

(5) Graft copolymers are a special type of branched


copolymer in which the side chains are structurally distinct
from the main chain. The illustration
depicts a special
case where the main chain and side chains are composed of
distinct homopolymers. However, the individual chains of a
graft copolymer may be homopolymers or copolymers.
Moiety: a part of a molecule that may include either
whole functional groups or parts of functional groups as
substructures

Several photoinduced motion or deformation was found in novel


systems containing AZ chromophores. (H. Yu and T. Ikeda,
Photocontrollable Liquid-Crystalline Actuators, Adv. Mater. 2011, 23, 21492180,
164~166, 167)

Refs.

Kurihara et al. presented the first observation of controlled translational motion of


polystyrene (PS) microparticles on the surface of LC films, which is driven entirely
by light.
Okano et al. proposed photoinduced expansion and condensation behavior in
pure AZ LMWLCs on the surface of water. The photochemical processes of those
AZ LMWLCs with room-temperature LC phases were regarded as the driven force
for the light-induced deformation.
Broer et al. fabricated micro-actuators by using LC inks and inkjet printers, which
acted as artificial cilia, being photo-driven by UV light.

Sec. 6 (LCBC Actuators) Refs. 9, 173, 204, 205


:
Refs. 164~166, 167(a)~(c), 9(a)~(d), 173(a)~(c), 204, 205

7. Outlook
In various systems of LMWLCs, LCPs, LCEs and LCBCs,
photo controllable LC actuators have been achieved by
introduction of photochromic molecules into LC materials.
* Light can conveniently manipulate order-disorder and
order-order changes of LC materials (Fig. 1A), which
induces a large modulation in refractive index by
photochemical
processes
such
as
photocontrolled
alignment,
photochemical
phase
transition,
and
photoinduced
cooperative
motion, leading to their
photonic applications.

* Through crosslinking in LCEs, the LC ordering is fixed in 3D


networks, giving birth to photomechanical and
photomobile effect in macroscopic scales. Even though
the photoinduced change in the isomerization of one
single photochromic molecule is smaller than 1 nm (Fig.
2A), it can be transferred into visible objects of photodriven movements of LCEs by the alteration of mesogenic
ordering, which are competitive and promising for many
applications as soft actuators.

* Combing microphase separation with SMCM in one system


of LCBCs, the LC materials can precisely adjust the
orientation of nanostructures by their photocontrollable
properties, extending their applications in the promising
nanotechnology.

The benefits that arose from recent progress in living


polymerization methods, characterization techniques, and
materials chemistry, photomechanical and photomobile
properties as well as templated nanofabrication methods
are currently the hot topics that dominate the research
fieeld of LC actuators.
* Although LCE actuators directly transfer light into
mechanical energy, at the present time they cannot
compete with optoelectronic materials, and their
efficiency and speed for light energy conversion are still
far from optimal.
* Furthermore, direct utilization of sunlight might be
another key requirement for industrial applications of
such materials.

As one of the most interesting supramolecularly selfassembled materials, LCBC actuators in particular provide
a unique chance for photocontrollable orientation of
regularly ordered nanostructures with parallel or
perpendicular distributions. But other more complicated
patterns such as helical structures have not yet been
obtained by SMCM.
* Moreover, the time-consuming process and thermal
stability of the fabricated regular nanostructures need
improvement. Novel materials for the recycling of such
nanotemplates still remain challenging and require
further intensive investigation. Future investigation of
photocontrollable LC actuators must focus on these
considerations.

You might also like