You are on page 1of 7

Showing the work on an interface nanostructured array guided

high performance electrochemical actuator presented by


Mr. Guan Wu and Prof. Wei Chen, Suzhou Institute of
Nano-Tech and Nano-Bionics, Chinese Academy of Sciences.

As featured in:

Title: An interface nanostructured array guided high


performance electrochemical actuator
A nanostructuredarray interface with large specic surface area
and fast ion-transmission-channels enables us to achieve an
electrochemical actuator with botha fast-responseand
large-deformation in air. We believe the high performance
actuator will have great potential for bionic ying insects or
robots,haptics for portable consumer devices, a shape or
position controller for adaptive optics, dynamic sensors,
and so on.

See W. Chen et al.,


J. Mater. Chem. A, 2014, 2, 16836.

www.rsc.org/MaterialsA
Registered charity number: 207890

Published on 21 August 2014. Downloaded by Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences on 27/09/2014 07:23:01.

Journal of

Materials Chemistry A
COMMUNICATION

Cite this: J. Mater. Chem. A, 2014, 2,


16836

View Article Online


View Journal | View Issue

An interface nanostructured array guided high


performance electrochemical actuator
G. Wu, G. H. Li, T. Lan, Y. Hu, Q. W. Li, T. Zhang and W. Chen*

Received 18th August 2014


Accepted 21st August 2014
DOI: 10.1039/c4ta04268a
www.rsc.org/MaterialsA

Here we report a novel electrochemical actuator using a hierarchically


architectured nanostructure electrode. Vertically aligned NiO nanowall arrays, which act as an interface layer, are in situ grown on a freestanding graphenecarbon nanotube hybrid lm. The large specic
surface area and fast ion transmission channels of this nanostructured
array interface enable us to achieve large deformation in quick
switching response (18.4 mm per 0.05 s), high strain and stress rates
(8.31% s 1, 12.16 MPa s 1) and excellent durability upon 500 000 times
continuous operations in air.

Similar to natural muscles in terms of light, so, and achievable


strain and stress, electroactive polymers have attracted great
attention for a variety of biomimetic applications including
bionic ying insects or robots, haptics for portable consumer
devices, and dynamic sensors.16 In past decades, ionic electroactive polymer actuators based on polymer gel, conducting
polymer, ionic polymer metal composite (IPMC) and carbon
nanotube (CNT) have been intensively studied due to their
impressive large-strain under low-voltage stimulation and airworking capability.79 Among them, electrochemical actuators,
which are composed of a layer of polymer electrolyte sandwiched between electrodes, have emerged as promising candidates owing to their better controllability and operability in air
aer introducing ionic liquid (IL) as the electrolyte.10 As electromechanical strain in actuators is generated by accumulation
or depletion of excess charges (ions) at the electrodes under an
applied voltage, in general, fast ion transport and excess ion
storage near the electrodes are necessary to generate large strain
and fast switching response.10 However, limited to the ion
diusion process, response for large strain is usually slow.11 It is
a great challenge to develop practically viable electrochemical
actuators with both fast response and large deformation.
i-Lab, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences,
Suzhou 215123, P. R. China. E-mail: wchen2006@sinano.ac.cn
Electronic supplementary information (ESI) available: Experimental, the
structural characteristics of electrodes and properties of the electrolyte layer,
and the actuating deformation movies are described. See DOI: 10.1039/c4ta04268a

16836 | J. Mater. Chem. A, 2014, 2, 1683616841

Recently, carbon nanotube and graphene with high specic


surface area (SSA), exceptional electrical and mechanical properties have received substantial attention as exible electrodes,1215 showing signicantly improved performances like
cycle stability,16 response speed and strain (stress) rate17 of
electrochemical actuators. For instance, Asaka's group utilized
Millimeter-Long Single-Walled Carbon Nanotubes (MLSWCNTs) as actuator electrodes.18 Superb mechanical and
electrical properties of the ML-SWCNT lms endued the IPMC
actuators much faster strain and stress rates (2.28% s 1 and
3.26 MPa s 1). Furthermore, two dimensional graphene electrodes, with porous yet densely packed nanostructure and highion-accessible surface area, exhibited promising foreground in
ionic actuators.19 Due to the unique structures, electrochemical
strains are caused not only by ions transfer at the interface
between electrode and electrolyte layer,20 but also by ion insertion directly into the electrode.21 As a result, the actuator
performance could be greatly improved. Therefore, the material
and structure of electrodes have a leading eect on achieving
high performance electrochemical actuator with both large
deformation and fast response.
In this work, we are motivated to develop a novel electrochemical actuator based on a hierarchically architectured
nanostructure electrode where vertically aligned NiO nanowall
arrays are grown on the surface of the graphenecarbon nanotube hybrid lm and act as an interface layer. Our actuator
presents one of the best actuation properties among previously
reported in ionic polymer actuators. The extraordinary actuation behaviours are generated by large specic surface areas of
the nanostructured array electrode interface layer to accommodate more ions and provide continuous path channels for
ion faster transportation, which are vital to achieve large
deformation in high frequency. The main purpose of our study
is to put forward a proposal to clarify the importance of the
nanostructured array interface in the high performance ionic
actuator and we believe our results will guide the development
of biomimetic technologies ranging from robotics and microsensors to articial muscles.

This journal is The Royal Society of Chemistry 2014

Published on 21 August 2014. Downloaded by Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences on 27/09/2014 07:23:01.

View Article Online

Communication

The hierarchically architectured nanocomposite electrode is


composed of a carbon nanostructure conductive network and
metal oxide nanowall arrays, which are introduced at the
interface layer between electrode and electrolyte layer. By a
facile chemical bath deposition technique, vertically aligned
nickel oxide nanowalls (VA-NiONWs) are in situ grown on one
side of a free-standing hybrid lm composed of reduced graphene oxides (RGO) and multiwalled carbon nanotubes
(MWCNTs). Such RGOMWCNT hybrid lms are prepared
through pp stacking self-assembly. As shown in Fig. 1a, the
nanostructured electrode based actuator is then constructed by
laminating a thermoplastic polyurethane/ionic liquid electrolyte
layer with two pieces of VA-NiONW@RGOMWCNT freestanding lms via simple hot-press. It is worthy to note that NiO
nanostructures, which have been widely investigated as active
electrode materials in energy storage and conversion,22,23 are
selected here mainly to modulate the interface microstructure,
and hence optimize the non-Faradaically driven charge injection and ion migration. On the other hand, taking into account
the elasticity and IL uptake capability for the middle polymer
electrolyte layer, thermoplastic polyurethane (TPU) is used as
the supporting polymer, showing not only high storage capacity
for EMIBF4 (>60 wt%) but also good mechanical properties in
elongation at break (>700%) under such high IL uptake (see
Fig. S1). With such architecture, our actuators benet much
from the unique VA-NiONW@RGOMWCNT nanostructured
electrodes. First, neat assembly of two dimensional RGO and
one dimensional MWCNTs (Fig. 1b and c), guarantees the
superior electrical conductivity and mechanical properties for
lightweight and porous electrodes (around 150 S cm 1 and 35

The preparation of actuator and the morphologies of electrodes. (a) Illustration of structure and assembly procedure of the VANiONW@RGOMWCNT electrode based electrochemical actuator.
Surface (b and d) and cross-sectional (c and e) SEM images of the
RGOMWCNTs and VA-NiONW@RGOMWCNT electrode lms,
scale bar 5 mm and 1 mm, respectively.

Fig. 1

This journal is The Royal Society of Chemistry 2014

Journal of Materials Chemistry A

GPa, respectively), and the whole actuator (146.35 MPa).


Second, exceptional interface nanostructure with high SSA and
fast ion transmission channel greatly facilitates the charge
dynamics near the electrodes, which permit the generation of
large strain and fast switching response. Fig. 1d and e show that
NiO nanosheets nearly 1 mm in length and 1 mm in height are
grown vertically and cross-linked with each other on the RGO
MWCNT hybrid lms, and (111), (200), and (220) peaks are
present in the X-ray diraction pattern of NiO shown in the
cubic phase24 in Fig. S2, which further proves that the NiO
nanosheets were successfully grown on the RGOMWCNT
surface.
Similar to previous reports, accumulation or depletion of
excess charges (ions) at the electrodes under an applied electrical eld generates the swelling of one side and shrinkage of
the other side,25,26 and then the electrochemical actuator strip
bends (Fig. 1a). The bending motion of the newly designed
actuator at an applied square wave voltage of 2.5 V with
dierent frequencies was investigated. Remarkably, within 2 s
the VA-NiONW@RGOMWCNT electrode based actuator could
display very large deformation (Fig. 2a). We further examined
the bending motion of the actuator at lower frequencies and
analysed the durable response of the actuator by providing
sucient time for ion migration. Due to the large bending
motions of our actuator in lower frequency, we use the curvature to evaluate the actuator's bending deformation. The longtime bending motion curvature is shown in Fig. 2b and Scheme
S1. The actuator quick responded to large bending deformation (curvature was 0.93 cm 1) within 1 s (0.5 Hz) and then the
movement gradually suspended (curvature was 1.72 cm 1)
around 50 s (0.01 Hz). Notably, no back relaxation was discovered for an extended interval time of up to 600 s, which is known

Fig. 2 Bending performance of low-voltage-driven actuators. (a)


Photographs of the VA-NiONW@RGOMWCNT electrode based
electrochemical actuator without (centre), and with an applied voltage
of 2.5 V (left/right). The bending motion took place within 2 s. (b)
Curvature change of the actuator and inset show the bending motion
of the actuator within 50 s at 2.5 V. Bending response of the actuator at
an applied square wave voltage of 2.5 V with frequencies of (c) 1 Hz
and (d) 10 Hz. (e) Cycling test for the actuator under 2.5 V and 10 Hz
stimulation, D0 represents the initial displacement value for twenty
cycles.

J. Mater. Chem. A, 2014, 2, 1683616841 | 16837

Published on 21 August 2014. Downloaded by Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences on 27/09/2014 07:23:01.

View Article Online

Journal of Materials Chemistry A

as one of the common drawbacks of traditional IPMC actuators.20,26 The inset photograph in Fig. 2b, which was taken
within 50 s (0.01 Hz), deeply illustrates the ultra-large bending
deformation of our actuator in low frequency.
High frequency (>1 Hz) mechanical response was also presented of our actuator. In high frequency, displacement can be
continuously monitored using a laser displacement meter. At
an applied frequency of 1 Hz, peak to peak displacement as
large as 30 mm was observed (Fig. 2c), which was signicantly
larger than the RGOMWCNT electrode actuator. Interestingly,
when the frequency was further increased to 10 Hz, the actuator
could still display a large peak to peak bending deformation
(18.4 mm in 0.05 s, Fig. 2d) without any notable deterioration
over 500 000 times continuous operation in air (Fig. 2e). The
strain rate varied from 0.48.31% s 1 and was recorded in the
frequency ranged from 1 to 20 Hz under 2.5 V. The maximum
strain speed was up to 8.31% s 1 and corresponded to the stress
rate of 12.16 MPa s 1 at the frequency of 10 Hz, which
demonstrated to be one of the best actuation deformation
compared with similar electrochemical actuators16,18,27,28 and
even the actuation performance could be comparable with other
energy type of graphene based actuators.4,29,30 ESI Movies
present the fast and visible motions of the new actuator at the
voltage of 2.5 V with frequencies of 1 Hz, 2.5 Hz, 5 Hz and 10
Hz, respectively. It can be found that when increasing the
electric stimulus frequency the actuator almost shows no
sacrice in displacement, which alters the widely accepted
sense in the ionic electroactive actuator that large deformation
can only occur at low frequency, but small deformation occurs
at high frequency due to the ion transmission.
In order to further evaluate and understand the interface
electrode structure on such considerable fast and large deformation, electrochemical actuators with disordered NiO@RGO
MWCNT electrode was investigated for comparison. Here, a
disordered NiO@RGOMWCNT electrode was fabricated via a
electrochemical deposition and casting method (see Experimental and Fig. S3). Among these, a novel VA-NiONW@RGO
MWCNT electrode based actuator denitely exhibited
outstanding electromechanical properties in the high applied
frequency range of 120 Hz (Fig. 3a), which naturally correlated
with its specic electrode structure (Fig. 1) and electrochemical
properties (Fig. 3b). Results show that the hierarchically nanostructured electrode lm displayed the largest SSA and capacitance (151.17 m2 g 1, 78.75 F g 1), compared with the
disordered NiO@RGOMWCNT electrode (108.91 m2 g 1, 36.66
F g 1) and RGOMWCNT electrode (31.02 m2 g 1, 11.25 F g 1).
Higher SSA could generate larger capacitance and accommodate more regions for ion accumulation, causing larger volume
expansion and mechanical deformation of the actuator. Moreover, pore size analysis of three kinds of electrode lms further
revealed that hierarchically VA-NiONW@RGOMWCNT electrode with wide pore size distribution (0.47117.2 nm from
micro, meso to macro-pore, as shown in Fig. 3c), could provide
open paths for ion diusion, intercalation and de-intercalation
than the disordered NiO@RGOMWCNT (6.386.2 nm) electrode with tortuous paths and RGOMWCNT (2.7133.6 nm)
electrode. Obviously, VA-NiONW@RGOMWCNT has a

16838 | J. Mater. Chem. A, 2014, 2, 1683616841

Communication

Fig. 3 The comparison of electromechanical and electrochemical


properties of three kinds of electrode based actuators. (a) Bending
displacement of actuators with three kinds of electrodes in the applied
frequency range of 120 Hz under 2.5 V. (b) Cyclic voltammetry (CV)
curves for three kinds of dierent electrode based actuators at 10 mV
s 1 sweep rate. (c) Pore size distribution curves for three kinds of
electrode lms based on the BJH method.

considerable amount of micro-pore among the three kinds of


electrode lms. It is those micro and small meso-pores that
have the great contribution to capacitance due to their sizes
closer to the ion sizes,3133 and then are more suitable for large
volume expansion of the actuator.
For the purpose of comprehensive investigation of the
nanostructure of vertically aligned nanosheets aecting the
actuation performance, we controlled the growth time of VANiONWs in inorganic salt solution and modied their heights
and densities. As shown in the SEM images in Fig. 4a, dierent
heights and densities of VA-NiONWs were formed gradually
when the growth time was increased from 0 to 10 h. Apparently,
thin nanosheets and large pores existed in VA-NiONWs of 5 h
and 7 h growth conditions. However, those large pores had no
contribution for electromechanical properties of the actuator.
Aer 10 h growth, neat nanowall array structures were basically
formed. Concomitantly, the height of NiO nanowalls would
almost not be changed even with the further extension of
growth time, remaining at 1 mm (see Fig. 4b). But the nanowalls
would be thicker appropriately (see the Fig. 4a SEM image of 15
h growth conditions). The corresponding charge storage ability
of actuators is also evaluated in Fig. 4b. The capacitance of
actuators would be enhanced with the growth conditions of VANiONWs within 10 h (30.8 F g 1 for 5 h, 50.3 F g 1 for 7 h and
78.75 F g 1 for 10 h), which complied well with the heights and
densities of VA-NiONWs. At 10 h growth conditions, the interfacial porous structure of VA-NiONW@RGOMWCNTs also
presented the optimum value, which resulted in the largest SSA
(151.17 m2 g 1), compared with 78.5 m2 g 1 for 5 h and 115.02
m2 g 1 for 7 h (see Fig. S4). However, the capacitance
decreased at the 15 h growth time (60 F g 1) because the thicker
nanowalls lowered the SSA (129.61 m2 g 1) as well as the ion
diusion process in the electrode. Fig. 4c shows the bending

This journal is The Royal Society of Chemistry 2014

Published on 21 August 2014. Downloaded by Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences on 27/09/2014 07:23:01.

View Article Online

Communication

Journal of Materials Chemistry A

Fig. 4 Properties of dierent heights and densities of VA-

NiONW@RGOMWCNT lms. (a) SEM images of hierarchically architectured VA-NiONW@RGOMWCNT electrodes were prepared in
inorganic salt solution for dierent times. Surface image scale bar
500 nm, sectional image scale bar 1 mm and (b) shows the calculated
heights and capacitances of VA-NiONWs. (c) Bending actuation
displacement in the frequency range 120 Hz under 2.5 V of VANiONW@RGOMWCNT actuators.

displacement of the VA-NiONW@RGOMWCNT electrode


based actuators. When the growth time reached 10 h, the
actuator exhibited an optimal actuation deformation, which
agrees well with the electrochemical property in Fig. 4b. Obviously, the heights and densities of VA-NiONWs had a great
inuence on the electrochemical and electromechanical properties of actuators.
For understanding the electrochemical behavior of
bimorph actuators, the useful electrochemical impedance
spectroscopy (EIS) characterization was introduced.21,34,35 The
impedance responses of all actuators have been characterized
in the way as illustrated in Fig. S5. We have used the same
distance between the electrode contacts. Nyquist plots of three
kinds of electrode based actuators are shown in Fig. 5a, and
are analysed by the same equivalent circuit as described in the
inset of Fig. 5a. The equivalent circuit contains R0 which is the
resistance of the actuator mainly due to the electrolyte layer,
the element Q1/R1 in parallel represents the contact impedance between the actuator's electrode layers and the metal
electrode, Warburg impedance Zw (ref. 21) and the elements
Q2/R2 in parallel represents the ion intercalation impedance.36
The reason for this model is mainly considered based on the
dierent structural characteristics of electrodes in the three
kinds of actuators. For the RGOMWCNT actuator, an
enlarged inset plot in Fig. 5a is presented by a single depressed
semicircle, indicating deviations from ideal capacitive
behavior. To account for this, the real contact/nanocarbon
interface has oen been described in terms of constant phase
element Q1 and contact resistance R1 and owing to the
conductivity and the porous interface structure of the nanocarbon electrode, the contact resistance R1 is a little larger of a
few hundreds of Ohms, which is similar to our previous
report.16 As we can see from the phase vs. frequency of RGO
MWCNTs in Fig. S6, the phase angle got close to 0 at a

This journal is The Royal Society of Chemistry 2014

Fig. 5 EIS analysis and the actuation mechanism of actuators. (a)


Nyquist plots of RGOMWCNT, disordered NiO@RGOMWCNT and
VA-NiONW@RGOMWCNT based actuators. The inset gures are the
depressed semicircle of Nyquist plots and the equivalent circuit model
of the actuator. Symbols denote Experimental data, while the
continuous lines represent the tted data. (b) Schematic illustration of
strain generated of actuator with excess ions in and out of the electrode layers with voltage applied, black represents RGOMWCNTs and
green is NiO nanowalls.

frequency lower than 1 Hz. In other words, the impedance


response is mainly controlled by the resistance at low
frequency and there was also nearly no diusion process and
then main parameter that controls the actuation properties is
the resistance. When active materials composed of VANiONWs and disordered NiO were added at the interface
between the RGOMWCNT electrode and the electrolyte layer,
the impedance data are found to correspond to a resistance in
the high-frequency region, diusion (Warburg impedance) in
the medium-frequency region, and capacitance in the lowfrequency region, which are represented as symbols in Fig. 5a.
The continuous lines are tted data according to the equivalent circuit. According to the tting (see Table S1), the values
of R0 and R1 indicate that the cells have been properly fabricated and tested under the same conditions. The tting
parameter of R2 is much lower for VA-NiONW@RGOMWCNTs
(1320 U) compared to the disordered NiO@RGOMWCNTs
(3526 U), which means that the in situ grown VANiONW@RGOMWCNTs have a better electron and then ion
transfer process than the disordered NiO@RGOMWCNT
composite electrode. When taking the diusion ability (Zw)
into account, one has the following sequence disordered
NiO@RGOMWCNTs > VA-NiONW@RGOMWCNTs, suggesting that the VA-NiONW@RGOMWCNT electrode is more easy
for ion penetration and transportation. In other words, vertically aligned nanowall structures provide more smoothly
channel for fast charge and electromechanical deformation at

J. Mater. Chem. A, 2014, 2, 1683616841 | 16839

Published on 21 August 2014. Downloaded by Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences on 27/09/2014 07:23:01.

View Article Online

Journal of Materials Chemistry A

high frequency. Considering the capacitive behavior expressed


in terms of Q1 for RGOMWCNTs and Q2 for interface charge
storage of hierarchical nanocomposite electrodes, we can see
that the contribution from Q1 (mF) is nearly at the same level
for three kinds of electrodes. But the interfacial charge (ion)
storage Q2, which is at the level of mF, was greatly improved as
NiO was introduced, showing several orders higher than that
of Q1 and relatively make larger contribution to the electrode
capacitance. When we compare the Q2, VA-NiONW@RGO
MWCNTs > disordered NiO@RGOMWCNTs is obtained. The
capacitance values agree well with the electrochemical and
electromechanical properties shown in Fig. 3a and b.
In the light of the results obtained so far, we speculate the
actuation mechanism for the novel electrode based actuator.
As illustrated in Fig. 5b, the signicantly improved actuation
performance is mainly attributed to the unique structure of
the VA-NiONW interface layer. The RGOMWCNT lms act as
a high electronic transmission carrier. Porous VA-NiONW
interface layers with larger specic surface area could provide
more areas for ion ooding and accumulations33 as well as
faster channels for ion transmission. Combined with the
high uptake capability of the polymer electrolyte layer, large
size cations of ionic liquid could easily accumulate not only
at the interface between the hierarchical electrode and the
easily deformable TPU layer but also in the hierarchical
electrode itself, causing the anisotropic large volume
expansion and mechanical elongation of the actuator.37
Additionally, in situ growth of vertically aligned nanostructures oer good electron transfer path and channels for
ion rapid intercalation and de-intercalation,38 compared to
that of disordered and tortuous structures, which is believed
to be vital in realizing large deformation at high frequency.
The actuation performance may be further improved by
other capacitive active metal oxide, which could provide
faradic charging transfer by the chemical reaction with the
electrolyte layer and a detailed study will be in our future
work.

Conclusions
We have demonstrated a high performance electrochemical
actuator based on the VA-NiONW@RGOMWCNT electrode.
The use of a nanostructured VA-NiONW array interface layer
with a large specic surface area could provide more areas for
ion ooding and accumulation as well as vertically aligned
nanostructures oer good electron transfer path and channels
for ion rapid intercalation and de-intercalation, which enables
us to achieve a high performance actuator with both fast
response and large deformation, including large bending peak
to peak deformation (18.4 mm) under high frequency (10 Hz),
high strain and stress rate (8.31% s 1, 12.16 MPa s 1), and
excellent actuation stability (500 000 cycles). Considering these
remarkable achievements together with exibility, lightweight,
and large bending deformation, we believe that the nanostructured array interface based actuators will have great
potential for bionic ying insects or robots, haptics for portable

16840 | J. Mater. Chem. A, 2014, 2, 1683616841

Communication

consumer devices, shape or position controller for adaptive


optics, dynamic sensor, and so on.

Acknowledgements
This work was supported by the Natural Science Foundation for
Distinguished Young Scientists of Jiangsu Province
(BK2012008), the Hong Kong, Macao and Taiwan Science &
Technology Cooperation Program of China (2012DFH50120),
the National Natural Science Foundation of China (21373263,
11204350, 51303204), the National Basic Research Program of
China (2010CB934700), and the External Cooperation Program
of BIC, Chinese Academy of Sciences (121E32KYSB20130009).

Notes and references


1 R. H. Baughman, C. X. Cui, A. A. Zakhidov, Z. Iqbal,
J. N. Barisci, G. M. Spinks, G. G. Wallace, A. Mazzoldi,
D. De Rossi, A. G. Rinzler, O. Jaschinski, S. Roth and
M. Kertesz, Science, 1999, 284, 13401344.
2 M. Shahinpoor, Y. Bar-Cohen, J. O. Simpson and J. Smith,
Smart Mater. Struct., 1998, 7, R15R30.
3 F. Carpi and E. Smela, Biomedical applications of electroactive
polymer actuators, John Wiley & Sons Ltd, United Kingdom,
2009.
4 Y. Zhao, L. Song, Z. P. Zhang and L. T. Qu, Energy Environ.
Sci., 2013, 6, 35203536.
5 D. R. MacFarlane, N. Tachikawa, M. Forsyth, J. M. Pringle,
P. C. Howlett, G. D. Elliott, J. H. Davis, M. Watanabe,
P. Simon and C. A. Angell, Energy Environ. Sci., 2014, 7,
232250.
6 R. Pelrine, R. Kornbluh, Q. B. Pei and J. Joseph, Science, 2000,
287, 836839.
7 Y. Bar-Cohen, Electroactive Polymer (EAP) Actuators as
Articial Muscles: Reality, Potential and Challenges, SPIE
Press, Bellingham, WA, 2004.
8 E. Smela, Adv. Mater., 2003, 15, 481494.
9 O. Kim, T. J. Shin and M. J. Park, Nat. Commun., 2013, 4,
2208.
10 L. Kong and W. Chen, Adv. Mater., 2014, 26, 10251043.
11 J. Lin, Y. Liu and Q. M. Zhang, Polymer, 2011, 52, 540
546.
12 L. Lu and W. Chen, Adv. Mater., 2010, 22, 37453748.
13 T. Fukushima, K. Asaka, A. Kosaka and T. Aida, Angew.
Chem., 2005, 44, 24102413.
14 L. H. Lu, J. H. Liu, Y. Hu, Y. W. Zhang and W. Chen, Adv.
Mater., 2013, 25, 12701274.
15 J. Kim, J. H. Jeon, H. J. Kim, H. Lim and I. K. Oh, ACS Nano,
2014, 8, 29862997.
16 L. H. Lu, J. H. Liu, Y. Hu, Y. W. Zhang, H. Randriamahazaka
and W. Chen, Adv. Mater., 2012, 24, 43174321.
17 J. Li, W. Ma, L. Song, Z. Niu, L. Cai, Q. Zeng, X. Zhang,
H. Dong, D. Zhao, W. Zhou and S. Xie, Nano Lett., 2011,
11, 46364641.
18 K. Mukai, K. Asaka, T. Sugino, K. Kiyohara, I. Takeuchi,
N. Terasawa, D. N. Futaba, K. Hata, T. Fukushima and
T. Aida, Adv. Mater., 2009, 21, 15821585.

This journal is The Royal Society of Chemistry 2014

Published on 21 August 2014. Downloaded by Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences on 27/09/2014 07:23:01.

View Article Online

Communication

19 L. Lu, J. Liu, Y. Hu and W. Chen, Chem. Commun., 2012, 48,


39783980.
20 Y. Liu, C. Lu, S. Twigg, M. Ghaari, J. Lin, N. Winograd and
Q. M. Zhang, Sci. Rep., 2013, 3, 973.
21 I. Takeuchi, K. Asaka, K. Kiyohara, T. Sugino, K. Mukai and
H. Randriamahazaka, J. Phys. Chem. C, 2010, 114, 14627
14634.
22 G. M. Zhou, D. W. Wang, L. C. Yin, N. Li, F. Li and
H. M. Cheng, ACS Nano, 2012, 6, 32143223.
23 H. G. Zhang, X. D. Yu and P. V. Braun, Nat. Nanotechnol.,
2011, 6, 277281.
24 G. H. Li, X. W. Wang, H. Y. Ding and T. Zhang, RSC Adv.,
2012, 2, 1301813023.
25 J. Lin, Y. Liu and Q. M. Zhang, Macromolecules, 2012, 45,
20502056.
26 Y. Liu, S. Liu, J. Lin, D. Wang, V. Jain, R. Montazami,
J. R. Hein, J. Li, L. Madsen and Q. M. Zhang, Appl. Phys.
Lett., 2010, 96, 223503.
27 N. Terasawa, K. Mukai and K. Asaka, J. Mater. Chem., 2012,
22, 15104.
28 J. J. Liang, Y. Huang, J. Y. Oh, M. Kozlov, D. Sui, S. L. Fang,
R. H. Baughman, Y. F. Ma and Y. S. Chen, Adv. Funct. Mater.,
2011, 21, 37783784.

This journal is The Royal Society of Chemistry 2014

Journal of Materials Chemistry A

29 J. J. Liang, L. Huang, N. Li, Y. Huang, Y. P. Wu, S. L. Fang,


J. Y. Oh, M. Kozlov, Y. F. Ma, F. F. Li, R. Baughman and
Y. S. Chen, ACS Nano, 2012, 6, 45084519.
30 Y. Huang, J. J. Liang and Y. S. Chen, J. Mater. Chem., 2012, 22,
36713679.
31 K. Kiyohara, T. Sugino and K. Asaka, J. Chem. Phys., 2010,
132, 144705.
32 P. Wu, J. Huang, V. Meunier, B. G. Sumpter and R. Qiao, ACS
Nano, 2011, 5, 90449051.
33 F. B. Sillars, S. I. Fletcher, M. Mirzaeian and P. J. Hall, Energy
Environ. Sci., 2011, 4, 695706.
34 T. Mirfakhrai, J. Y. Oh, M. Kozlov, S. L. Fang, M. Zhang,
R. H. Baughman and J. D. W. Madden, J. Electrochem. Soc.,
2009, 156, K97K103.
35 H. Randriamahazaka and K. Asaka, J. Phys. Chem. C, 2010,
114, 1798217988.
36 J. Bisquert, Electrochim. Acta, 2002, 47, 24352449.
37 S. Liu, Y. Liu, H. Cebeci, R. G. de Villoria, J. H. Lin,
B. L. Wardle and Q. M. Zhang, Adv. Funct. Mater., 2010, 20,
32663271.
38 B. Varghese, M. V. Reddy, Z. Yanwu, C. S. Lit, T. C. Hoong,
G. V. S. Rao, B. V. R. Chowdari, A. T. S. Wee, C. T. Lim and
C. H. Sow, Chem. Mater., 2008, 20, 33603367.

J. Mater. Chem. A, 2014, 2, 1683616841 | 16841

You might also like