Sie sind auf Seite 1von 11

CARBON

4 9 ( 2 0 1 1 ) 1 2 7 9 1 2 8 9

available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/carbon

Electro-active grapheneNafion actuators


Jung-Hwan Jung, Jin-Han Jeon, Vadahanambi Sridhar, Il-Kwon Oh

Structural Dynamics and Smart Systems Laboratory, Division of Ocean Systems Engineering, School of Mechanical,
Aerospace and Systems Engineering, Korea Advanced Institute of Science and Technology, 335 Gwahak-ro, Yuseong-gu,
Daejeon 305-701, Republic of Korea

A R T I C L E I N F O

A B S T R A C T

Article history:

Electro-active actuators based on graphene reinforced Nafion composite electrolytes were

Received 26 September 2010

developed and their electro-chemo-mechanical properties and actuation performances

Accepted 24 November 2010

were investigated. The tensile strength of the grapheneNafion ionic membrane was signif-

Available online 29 November 2010

icantly improved up to 200% within 1.0 wt.% loading, and Youngs modulus was more than
two times with a minute loading of graphene to Nafion electrolyte. The proton conductivity
and the water-uptake ratio were greatly improved, while apparent changes in the ion
exchange capacity were not observed. Morphological tests, chemical techniques, and scattering techniques were used to study the interaction mechanism between graphene and Nafion, resulting in great improvements of the actuation performances. Present results show
that a minute loading of graphene greatly improves the harmonic responses, the blocking
force and the energy efficiency in Nafion-based ionic polymermetal composite actuators.
 2010 Elsevier Ltd. All rights reserved.

1.

Introduction

Smart materials consisting of polymer composites or gels undergo dramatic and reversible shape deformation in response
to application of external stimuli like electric potential [1],
chemical contact [2], temperature [3], pH [4] and so on. Actuators converting electrical energy into mechanical work find
many applications in medical, mechanical, electrical and
aerospace engineering such as multi-link active catheters,
artificial muscles [5], bio-medical devices [6], micro-pumps,
molecular motors, micro-/nano-robots, micro-manipulators
[7], and biomimetic robots like jelly fish [8].
Though many polymers were used [911], Nafion is the
most widely used; however, the degree of actuation in pristine
Nafion is less than desirable. So, in order to improve the actuation and bending functionalities, doped Nafion, Nafion
reinforced with nano-scaled metal-fillers [12] were used
initially. However, metal doping suffers from disadvantages
like high doping levels, as much as 30% which not only
increase the cost of the actuator, but also hinders the

electrolytic mobility of cations, thereby limiting its bending


and actuation.
The recent discovery of nano-scale carbonaceous materials like graphene, carbon nanotubes (CNT), carbon nano-fibers (CNF), and fullerenes and their remarkable mechanical,
electro-chemical, piezo-resistive and other physical properties has opened the possibility for the development of a new
class of smart nano-materials. Graphene, the basic building
block of most carbonaceous nano-materials, is an intriguing
2D flat material of monolayer carbon atoms whose distinct
properties makes it very promising in applications like fieldeffect transistors, lithium ion batteries, hydrogen storage
[13], sensors [14], as well as a reinforcing filler in high performance polymer composites [1517]. By addition of these carbonaceous nano-fillers like CNF [18], CNT [19] and fullerenes
[20] to ionic polymer, one may overcome the severe limitations associated with high metal doping and can enable
new means to generate vivid motion and large bending in
Nafion-based actuators [21]. Though there are some recent reports on the actuation performance of graphite oxideNafion

* Corresponding author: Fax: +82 42 350 1510.


E-mail address: ikoh@kaist.ac.kr (I.-K. Oh).
0008-6223/$ - see front matter  2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.carbon.2010.11.047

1280

CARBON

4 9 ( 2 0 1 1 ) 1 2 7 9 1 2 8 9

actuators [22], the grapheneNafion composite actuator was


not reported until now.
In this study, we developed grapheneNafion composite
actuators and systematically investigated its actuation and
electro-active behavior. The interfacial interactions of polymer chains with the surfaces of dispersed rigid filler particles
are critical for understanding and improving the performance
of electro-active polymer composites [23]. Several techniques
like morphological tests, chemical techniques, scattering
techniques are used to elucidate the structural properties
and interaction mechanism between Nafion and graphene.
To investigate the reinforcing effect of graphenes in the polymer matrix, many electro-mechanical tests have been performed including stressstrain measurements, dynamic
mechanical analysis, ionic exchange capacity, proton conductivity and water-uptake ratio. Also, improvements of the
graphene-based ionic polymer actuators are described in
terms of harmonic responses and blocking forces.

2.

Experimental

2.1.

Fabrication of composite actuators

Graphene was prepared from graphene oxide by using modified-Hummers method [24] with graphite flakes as the starting material and was then synthesized through pre-reduction

of graphene oxide with sodium borohydride at 80 C for 1 h,


followed by post-reduction with hydrazine at 100 C for 24 h.
Representative morphology of graphene sheets is shown in
Fig. 1a.
The ionic polymer Nafion PFSA (perfluoro sulfonic acid)
were purchased as dispersions containing PFSA/PTFE (polytetrafluoroethylene) copolymer in the acid (H+) form in
alcohol and water. Prior to use, Nafion PFSA polymer dispersions were dried in a vacuum oven under a regulated temperature of 80 C for 12 h. Subsequently, the dried Nafion film
was dispersed in Di methyl acetamide (DMAc) solution by stirring for 6 h at a temperature of 80 C. Graphene was dispersed
in ortho-dichlorobenzene at weight percentages of 0.1% and
1.0% and were agitated for 2 h at room temperature to obtain
stable solutions. These two mixtures were subsequently combined and stirred for another 24 h to obtain stable graphene
Nafion dispersions. The stabilized and well-dispersed grapheneNafion solution was dried in a casting mold shown in
Fig. 1b at 120 C for 12 h and annealed at 140 C for 2 h.
The thickness of the membranes was measured with a
micrometer at multiple points on the membrane surface,
and the average value of the membrane thickness was about
0.34 mm. The electrodes on both sides of membrane were
made through electroless plating based on the platinum
amine complex salt [25]. Representative morphology of the
actuators is shown in Fig. 1c wherein uniform coating of plat-

Fig. 1 SEM micrograph of graphene (a); grapheneNafion composites in molds and their membranes (b); morphology of
actuators after platinum electroless plating with plating thickness of 6 lm. Inset in (c) shows a graphene sheet.

CARBON

4 9 ( 20 1 1 ) 1 2 7 912 8 9

inum layer with the thickness of 6 lm is observed. At the end


of these processes, the polymermetal composite membrane
was placed in the 1.5 N Lithium chloride (LiCl) solution for
more than 1 day for the exchange of Li+ cations with H+ cations. Specimens of the final electro-active composite actuators were prepared with a length of 25 mm, a width of 5 mm
and a thickness of 0.38 mm under hydrated condition.

2.2.

Testing

X-ray diffraction (XRD) of the composites was measured using


DMAX-Ultima III X-ray diffractometer in the range of 290.
SEM (scanning electron microscopy) images were recorded
using a cold field emission scanning electron microscope (S4700, Hitachi, Japan). Atomic force microscope (AFM) measurements were made using a commercial XE-100 (Park Scientific Inc.). Tensile testing of fully hydrated and dry
composites was tested using Universal testing machine, PT200N, Minebea, Japan as per American society of testing and
materials, ASTM D882 standards. The ionic exchange capacity
was calculated using the titration method with NaOH. Surface
analysis by High resolution X-ray Photoelectron Spectrometer
(XPS) (ESCA Multilab 2000 system) was used to detect the
presence of surface elements. ATR-FTIR (Attenuated total
reflection Fourier transform infra red spectroscopy) of the
membranes was tested using Shimadzu IRPrestige-21 spectroscope. The experimental setup for the measurement of
actuation is shown in Fig. 2. A charged couple device (CCD)
camera (XC-HR50) and a laser displacement sensor were used
for sensing displacements and motions, and a NI-PXI system
and a current amplifier (UPM1504) were used for signal generation and actuation of the electro-active artificial muscle.

3.

Results and discussion

3.1.

Structure and morphology

The effect of graphene addition on the degree of crystallinity


of the composites was investigated using Wide angle X-ray

1281

analysis as shown in Fig. 3a. Recast Nafion shows broad diffraction maxima at 2h values of 17.48 corresponding to d which can be attributed to the Teflon-like dospacing of 5.5 A
mains of Nafion [26]. In the case of grapheneNafion, there
was a decrease in intensity of both these peaks. The broad
diffraction peaks in the range of 2h = 1220 resulted from a
convolution of amorphous (2h = 15) and crystalline (2h =
17.5) scattering from the perfluorocarbon chains of Nafion.
By comparing these deconvoluted diffractograms as shown
in Fig. 3b and c, a decrease in the ratio between the area of
the crystalline peak and the amorphous scattering with increased graphene content is apparent from these spectra.
Deconvoluted XRD peaks of grapheneNafion shows the
shifting of crystalline peaks to 14.8. This indicates that a
new type of crystallite was created in the amorphous region
of Nafion, possibly at the interface between graphene and
Nafion due to disruption of the crystalline morphology of Nafion [27]. More evidence of changes in crystallinity can be observed from FTIR spectra. There is a correlation between a
decrease in the crystallinity of Nafion composites and the variation in the intensities and peak position [28] in the region
from 550 to 650 cm1 (Fig. 3d). Our results show a continuous
increase in peak intensities with increasing graphene concentrations, which implies progressive loss of crystallinity of the
composites. This is contrary to reported results of graphenereinforced poly vinyl alcohol (PVA) composites wherein a progressive increase in crystallinity of PVA with an increasing
concentration of graphene has been reported [29]. This variation can be attributed to the unique di-block morphology of
Nafion wherein nano-sized additives such as graphene and
Nafion are embedded in the interstitial regions.
The morphologies of recast Nafion and grapheneNafion
composites were studied by SEM and AFM as shown in
Fig. 4a and c, and b and d. Recast Nafion has typical di-block
morphology, whereas a layered structure in the case of grapheneNafion was observed.
The layered structure of graphene layers which are indicated by arrows in the polymer matrix is observed in Fig. 4b
and d, suggesting that during the solvent evaporation of solu-

Fig. 2 Experimental setup for measurement of tip displacement and motion of electro-active grapheneNafion composite
actuators.

1282

CARBON

4 9 ( 2 0 1 1 ) 1 2 7 9 1 2 8 9

Fig. 3 XRD spectra of recast Nafion and graphene-reinforced composites (a); deconvoluted XRD spectra from peak profile
analysis of recast (b) and 1.0 wt.% grapheneNafion composites (c); FTIR (d) spectra of recast Nafion and graphene reinforced
Nafion composites showing variation in crystallinity.

tion-cast films enhanced alignment of graphene layers parallel to the surface of Nafion fibrils. A similar observation in
graphene oxide nano-sheets [30] and nano-clay reinforced
polymer composites has been reported by various researchers
[31,32].

3.2.

Tensile properties

The stressstrain curves were obtained from uniaxial tensile


testing of the recast Nafion and grapheneNafion membranes
with 0.1 and 1.0 wt.% in dry and wet conditions as shown in
Fig. 5a and b, respectively. GrapheneNafion membranes
showed significant improvement in mechanical stiffness
and strength more than 100%, although only minute amounts
of graphene, in the weight percentages of 0.1% and 1.0% were
added.
In dry condition, the recast Nafion membrane failed at a
maximum stress of 11.5 MPa, whereas 1.0 wt.% graphene
Nafion membranes failed at the maximum stress of
31.39 MPa, an increase of over 280%. The fact that this significant increase was achieved at the 1.0 wt.% graphene loading
is very impressive. The magnitude of increase in tensile
strength is much more than that of other graphene-reinforced polymer composites [29,33]. Youngs modulus of the
1.0 wt.% grapheneNafion composite membrane was twice

that of the recast Nafion membrane as listed in Table 1. The


intensity of increase in tensile strength is also much more
than that of other plate-like fillers such as nano-clay-reinforced Nafion composites [34] or variants of nano-clay like
sulfonated nano-clay reinforced Nafion composites [35].
The mechanical properties of grapheneNafion composites were also studied in completely water saturated Nafion/
graphene membranes, since wet strength and stiffness are
of paramount importance for actuator applications. The equilibrium water uptake that infers no further increase in weight
of membranes even after prolonged immersion in water was
determined by a gravimetric method as previously used in our
publication [20]. A minimum of three sets of experiments
were tested and the error was well within the experimental
limits. The water-uptake ratio [38] strongly affects the actuation performance of IPMC which has the actuation mechanism of the ion migration with water molecules in the ionic
exchangeable membrane. The water-uptake ratio of pristine
recast Nafion membrane, 0.1 and 1.0 wt.% graphene-reinforced membranes were 16.7%, 10.42% and 6.8%. Water
adsorption reduced by more than 100% with a mere 1.0 wt.%
addition of graphene. This large decrease in water uptake is
consistent with the other layered-filler (nano-clay) reinforced
polymer composites [36]. The stressstrain curves of wet
composite samples are shown in Fig. 5b. Irrespective of the

CARBON

1283

4 9 ( 20 1 1 ) 1 2 7 912 8 9

Fig. 4 SEM [(a) and (b)] and AFM [(c) and (d)] morphologies of recast Nafion and grapheneNafion composites.

35

35

30

(iii)

(ii)

20

Stress, MPa

Stress, MPa

(iii)

25

25

15
(i)

10

Recast Nafion (i)


0.1 wt% graphene-Nafion (ii)
1 wt% graphene-Nafion (iii)

5
0

30

10

15

20
(ii)

15

(i)

10
Recast Nafion (i)
0.1 wt% graphene-Nafion (ii)
1 wt% graphene-Nafion (iii)

20

10

Strain, %

15

20

Strain, %

Fig. 5 Stressstrain curves of recast and graphene reinforced Nafion in (a) dry and (b) wet conditions.

Table 1 Properties of recast Nafion and two graphene-reinforced membranes and actuators.
Sample type

Recast Nafion
0.1 wt.% GrapheneNafion
1 wt.% GrapheneNafion

Youngs modulus (MPa)


Dry

Wet

11.5
17.9
26.86

9.85
11.60
19.98

Proton conductivity
(S/cm)

graphene content a progressive decrease in the mechanical


properties of wet samples can be observed.
The stressstrain curves also show a substantial increase
in strain with the addition of graphene; this can be explained
on the basis of the deformation behavior of nano-particle
reinforced di-block copolymers. Though many morphological
models of Nafion have been proposed by various researchers,
the general consensus is that Nafion has the morphology of

1.87 102
3.96 102
11.9 102

Water
uptake (%)
16.7
10.42
6.8

IEC (meq/g)

0.75
0.78
0.95

Blocking
force
(gf, @ DC 2 V)
0.199
0.341
0.426

di-block copolymer. The addition of nano-fillers such as


graphenes into these types of polymers can result in spatial
distribution and a template arrangement of graphene, especially at low loadings. The presence of graphene alters both
the morphology and the orientation of di-block copolymer
micro-domains [37] as can be observed in SEM and AFM.
When nano-particle reinforced di-block copolymers are subjected to tensile testing, an increase in strain can be expected.

1284

CARBON

4 9 ( 2 0 1 1 ) 1 2 7 9 1 2 8 9

Second, the addition of graphenes may also alter the static


(or strained) micro-phase morphology of Nafion, resulting in
improved tensile properties accompanied by an increase in
strain. Semi-crystalline polymers like Nafion are composed
of crystalline lamellae and entangled amorphous chains
show a complex behavior under external tensile strain
wherein the crack propagates in its own plane and the primary mechanism for arresting crack growth is the fibrils
and lamellar structure; however these fibrils alone are not
capable of arresting the crack growth. Hence, the crack propagates in its own plane in a stable manner and this is evident
from the smooth curves of load versus displacement as
shown in Fig. 5a and the fracture morphology is accompanied
by the formation of weld lines as shown in our earlier publication [20]. This type of crack propagation is irregular and is
evident from the waviness of the load versus displacement
diagrams. Waviness is more pronounced in the stressstrain
curves of 1.0 wt.% grapheneNafion composite, as shown in
Fig. 5a and b and from the unevenness in the fractured specimen. Another explanation for a large increase in strain is the
self-healing nature of layered-filler reinforced polymer composites [31].

3.3.

Ion exchange capacity (IEC)

The ionic exchange capacity was calculated using the titration method with NaOH as used in our previous publication
[20]. The variation in ion exchange capacity (IEC) in Table 1
shows a marginal increase with increasing concentrations
of graphene. Nafion belongs to a class of ion exchange polymers with a perfluoroethylene backbone and a side group
containing sulfuric acid groups. It is widely accepted that a
typical Nafion membrane consists of crystalline and amorphous regions [38], and that its hydrated clusters are located
in the amorphous regions as shown in schematically Fig. 6.

Higher graphene concentration showed a progressive decrease in water adsorption, thereby leading to a decrease in
IEC values. The variation of membrane conductance has
two possible origins: inter-particles and intra-particles [39].
The first involves a change in the rate constant of the elementary ion transfer reaction due to interactions between mobile
ions and fixed sites that depend on the hydration states of inter-particles. The second originates from micro-structural
changes within the membrane (intra-particle). In this case
of grapheneNafion composites, a decrease in the degree
and extent of ion clustering can be expected with the addition
of graphene. This decrease in ion clustering and water
adsorption decreases both the cluster diameter and the exchange sites for each cluster, thereby causing marginal variations in the IEC values.

3.4.

Chemical characterization

ATR-FTIR and XPS of composite membrane were used to


determine the extent and nature of chemical interactions between graphene and Nafion. ATR-FTIR spectra shown in
Fig. 7a, of recast Nafion and grapheneNafion composite show
substantial changes in COC symmetric stretching bands at
968 and 980 cm1, whereas considerable broadening occurred
in SO3 symmetric stretching vibrations bands at 1057 cm1
[40]. These changes were also reflected in asymmetric
stretching of sulfonate groups (SO3) 1260 cm1 peak indicating good interactions between the sulfonate moieties of Nafion membrane and the dispersed graphene. The doublybroadening observed in bending vibrations of OH deformation and H3O+ peaks respectively observed at 1729 and
1740 cm1 indicate interactions between the dispersed graphene and the hydrate groups of the Nafion polymer matrix. The
insert in Fig. 7b shows the changes in the shape and intensity
of the S2p peak of the XPS spectra, which reflects the interac-

Fig. 6 Schematic representation of proton conductivity in hydrated grapheneNafion composites and interactions between
graphene and sulfonated groups of Nafion.

CARBON

4 9 ( 20 1 1 ) 1 2 7 912 8 9

1285

Fig. 7 FTIR [(a) and (b)] and C1s XPS [(c) and (d)] spectra of recast and grapheneNafion composite membranes. The inset in (b)
shows the S1p spectra.

tions between the sulfonate moieties of Nafion with dispersed graphene sheets. The C1s signal of the XPS spectra
of the recast Nafion is shown in Fig. 7c. Two prominent peaks
corresponding to CC and CF2 were observed. The deconvolution of the spectra resulted in six peaks. CF2 dominates, followed by a satellite peak at 281.8 eV. Minor components also
appeared at 294.4 eV and are attributed to terminal CF3 groups
[41]. In the case of grapheneNafion composites, substantial
changes in C1s spectra can be observed in Fig. 7d. Deconvolution of the C1s peak revealed the presence of CO (286.34)
and C@O (288.73 eV) in addition to the prominent FCO
(292.84) and CF2 (291.41 eV) peaks of Nafion. The effect of
these interactions on the structure and morphology of grapheneNafion composites is shown schematically in Fig. 6. The
variation in FTIR peaks, combined with the decrease in water
uptake and a decrease in crystallinity as previously discussed
in XRD, indicates that substantial changes in the core-channel morphology of Nafion occurred with the addition of
graphene to Nafion.

3.5.

Harmonic actuation performance

A charge-coupled device (CCD) camera (XC-HR50) and a laser


displacement sensor (LK301, KEYENCE) were used to measure
the tip displacements and bending motions of the composite
actuators. A NI-PXI 6252 system and a current amplifier
(UPM1504) were used for signal generation and actuation of
the composite actuators. For simple harmonic responses,

the time histories of the tip displacements for recast Nafion


and graphene-reinforced composite actuators under an electrical sinusoidal input signal of a maximum voltage of 0.5 V
and an excitation frequency of 0.5 Hz are shown in Fig. 8a.
The tip displacement of the 1.0 wt.% graphene-reinforced
actuator is almost two times that of the recast Nafion-based
ionic polymermetal composite (IPMC) actuator.
Fig. 8b and c show the maximum positive values of the
harmonic responses under the sinusoidal electric inputs with
voltage amplitudes of 0.5 and 1.5 V according to the change of
excitation frequencies. The harmonic response of the
graphene-reinforced composite actuators was improved significantly as the input voltage increased with 1.0 wt.% graphene-reinforced composite actuators showing almost three
times higher tip displacement than recast Nafion actuator.
This increase in tip displacement is higher than the reported
values in single-walled carbon nano tube (SWCNT)-reinforced
IPMC actuators, even though the electrolyte in this study was
water, whereas Landi et al. [21] used ionic lithium salts as
electrolytes.
However, irrespective of the graphene concentration, there
is a marked reduction in displacement with increasing frequency, thereby indicating an inverse relationship between
displacement and frequency. The reason for this decrease is
due to the reduction in the response time of the actuator with
increasing frequency, at which the potential bias is alternated
[42]. However the intensity of this decrease is less than that
observed in the SWCNT-reinforced IPMC actuators [21]. This

1286

CARBON

4 9 ( 2 0 1 1 ) 1 2 7 9 1 2 8 9

of an electro-osmotic pressure induced by the migration of


hydrated cations and diffusion of water from these cluster regions through the nano-channels [43]. So, reinforcing filler
that can effectively block or delay this water migration will
be advantageous in improving the overall actuator performance. Graphene due to its platelet morphology will delay
this water diffusion due to its tortuous path. Overall, our results indicate that graphene at extremely low concentrations
are advantageous vis-a`-vis other carbonaceous nano-fillers
such as SWNCTs or MWNCTs and can greatly improve the
harmonic response and actuation behavior of Nafion-based
IPMC actuators.

3.6.

Blocking forces

A low blocking force of the electro-active IPMC actuator is a


critical problem to be improved for the realistic application.
The blocking force is defined as a required force at the tip
of a bending actuator when the tip position is kept fixed,
i.e. identically zero deflection. The blocking force is a measure of the actuation performance and represents the load
carrying capacity of the electro-active polymers. The blocking
force under DC 2 V was measured by using a load cell (LVS5GA, KYOWA). The blocking force of the graphene reinforced
Nafion actuator was about three times as big as that of the recast Nafion-based IPMC actuator as listed in Table 1. Present
results indicate that the graphene reinforcement can increase the blocking force of the IPMC actuator that was a critical problem of the electro-active polymers for real
applications. In our previous results, the blocking forces of
fullereneNafion and MWCNT-Nafion actuators with
0.1 wt.% loading were 0.128 gf/mm2 [20,44] and 0.186 gf/mm2
[45], respectively. As reported in Table 1, the blocking force
of the grapheneNafion actuator was 0.341 gf/mm2, showing
nearly two times higher than other nano-composite actuators. This can be attributed to the unique self-assembled
lamellar structure of graphenes in the Nafion matrix. Our results showed substantial increase in tensile strength and
modulus even with addition of minute concentrations of
graphene in Nafion, indicating considerable increase in stiffness of the membranes. This increase in stiffness of the
membrane results in higher blocking force.

3.7.
Fig. 8 Actuated tip displacements under sinusoidal
electrical inputs of 0.5 * sin(2p * 0.5 * t) (a), tip-displacement
histories in the harmonic responses at the excitation
voltages of 0.5 V (b) and 1.5 V (c) according to applied
frequencies.

improvement can be attributed to the increased proton conductivity, ionic exchangeable capacity and exceptional electron transport behavior of graphenes. Another explanation
for this behavior can be attributed to the platelet morphology of graphene. In a typical actuator, when an electric field
is applied across the actuator, fast bending motion toward
one of the electrodes occurs followed by a slow relaxation towards the other electrode. This bending motion is the result

Currentvoltage tests and efficiency

To consider the effect of graphenes on power consumption


and energy dissipation of the grapheneNafion composite
actuators, the currentvoltage tests were performed. Fig. 9
shows the hysteresis behaviors of a recast Nafion and 0.1
and 1.0 wt.% grapheneNafion composite actuators in the
range of 0.5 to +0.5 V at a constant resistance of 1.3 X and
the excitation frequency of 0.5 Hz.
The maximum current density of the composite actuators
increases up to three times in comparison with a recast Nafion-based actuator. Much higher current density is due to
large electro-mechanical deformations of the grapheneNafion composite actuator reported in Fig. 8. The dissipated electrical input energy per the area of the electrode surface during
a time period with an excitation frequency can be calculated
by using a following equation.

CARBON

1287

4 9 ( 20 1 1 ) 1 2 7 912 8 9

Time and space integral terms in Eq. (4) will be a constant


value, C, for all actuators. The specific electro-mechanical energy efficiency of the grapheneNafion composite transducers
can be defined in the following form with the reference value
of the recast Nafion-based IPMC transducer [46].
gnano-composite

gnano-composite
grecast Nafion
nano-composite !2

Fig. 9 IV curves of recast Nafion and 0.1 and 1.0 wt.%


grapheneNafion actuators.
RT
itvtdt
Einput
0
A
A
Z 1=fex  
it
Vinput sin2pfex tdt

A
0

Einput

where A, i, v, T, fex , and Vinput are the area of electrode, measured current, input voltage, period, excitation frequency
and voltage amplitude of sinusoidal electrical input. And
the unit kinetic energy, TIPMC , stored in the IPMC actuators
for a cycle can be computed in the following form.
R T R L 1 dwx;t2
q dt
dxdt
0
0 2
TIPMC

2
TIPMC
A
A
where q and xx; t are the equivalent density per length and
the transverse displacement, respectively. Here, the separation of variable by assuming the space-dependent deformed
shape can be applied with a mode shape, Ux, of the IPMC
beam and the time-dependent function, f t, as a sinusoidal
oscillation with the excitation frequency [44].
xx; t Wxf t WTipmax Ux sin2pfex t

And then, the electro-mechanical efficiency of the IPMC actuators can be defined as
g

TIPMC
Einput
R
L 1

0 2


R

T
qU2 xdx
2pfex 2 cos2 2pfex tdt W2Tipmax
0



W2Tipmax

AEinput
4

Einput

WTip

max

Nafion
Wrecast
Tipmax

Nafion
Erecast
input

nano-composite

Einput

Basically, much larger tip displacement means the generation of much higher kinetic energy in the oscillating cantilever actuators. Higher kinetic energy requires the external
work from the electrical input. While higher current densities
were observed in the V-I diagram, the power consumption required to generate a unit bending deformation is much smaller in the grapheneNafion composite actuators. The
dissipated electrical input energy densities of the prepared
IPMC actuators that can be calculated by using Eq. (1) will
be the same to the area of the V-I circles. The values of the
electrical input energy per unit area are listed in Table 2.
The input power to activate the 0.1 and 1.0 wt.% composite
actuators increases slightly. But, at this activation, the generated tip displacements of the 1.0 wt.% composite actuators
increase up to three times as listed in Table 2. Therefore,
the specific electro-mechanical efficiency defined in Eq. (5)
shows two times higher values in comparison with the recast
Nafion-based IPMC actuators. It means that the effect of
graphenes as a reinforcing agent in the Nafion-based IPMC
actuator strongly affects the electro-mechanical energy efficiency, resulting in the reduction of the input power to generate the unit tip displacement of the IPMC actuators.

4.

Conclusions

By incorporation of graphene, significant improvement in the


actuation performance of Nafion-based IPMC actuators was
observed in this study. Morphological studies of graphene
Nafion composites by SEM and AFM show that graphene
platelets are dispersed homogeneously in the Nafion polymer
matrix by solvent recasting method. Tremendous improvements in tensile strength and Youngs modulus were observed even with the addition of minute graphenes. The
proton conductivity increased two times with the addition
of 1.0 wt.% graphenes, whereas a considerable decrease in
water uptake was observed. Chemical studies by XPS and
FTIR showed considerable interactions between graphene
platelets and sulfonated moieties of the Nafion polymer ma-

Table 2 Specific electro-mechanical energy efficiency of IPMC actuators under the sinusoidal electric input with amplitude of
0.5 V and excitation frequency of 0.5 Hz.
IPMC actuators

Recast Nafion IPMC


0.1 wt.% GrapheneNafion IPMC
1.0 wt.% GrapheneNafion IPMC

Electrical input
energy density
(mJ/mm2)
Einput
0.046
0.101
0.305

Maximum tip
displacement (mm)

Specific electromechanical efficiency

WTipmax

gnano-composite

0.040
0.084
0.145

1.0
2.01
1.98

1288

CARBON

4 9 ( 2 0 1 1 ) 1 2 7 9 1 2 8 9

trix. Compared to the recast Nafion-based IPMC actuator,


much larger bending deformation and higher blocking forces
were achieved in the graphene-reinforced composite actuator
with minute addition of graphene. Currentvoltage tests and
subsequent efficiency analysis showed that specific electromechanical efficiency of graphene-reinforced actuators is almost twice than that of recast Nafion.

Acknowledgements
This work was supported by National Research Foundation of
Korea Grant funded by the Korean Government (2010-0018423)
and (2010-0000300). And this work was supported by a grant
from the Fundamental R&D Programs for Core Technology
of Materials funded by Ministry of Knowledge Economy,
Republic of Korea [K00060-282].

[16]

[17]

[18]

[19]

[20]

[21]
R E F E R E N C E S

[22]
[1] Shahinpoor M, Kim KJ. Ionic polymermetal composites: I.
Fundamentals. Smart Mater Struct 2001;10(4):81933.
[2] Gao J, Sansinena JM, Wang HL. Tunable polyaniline chemical
actuators. Chem Mater 2003;15(12):24118.
[3] Kuckling D, Adler HP, Ling L, Habicehr WD, Arndt KF.
Temperature sensitive polymers based on 2-(dimethyl
maleinimido)-N-ethyl-acrylamide: copolymers with Nisopropylacrylamide. Polym Bull 2000;44(3):26976.
[4] Schalkhammer T, Lobmaier C, Pittner F, Leitner A, Brunner H,
Aussenegg FR. The use of metal-island-coated pH-sensitive
swelling polymers for biosensor applications. Sens Actuators
B Chem 1995;24(3):16672.
[5] Lu J, Kim SG, Lee S, Oh IK. A biomimetic actuator based on
ionic networking membrane of poly(styrene-alt-maleimide)incorporated poly(vinylidene fluoride). Adv Funct Mater
2008;18(8):12908.
[6] Jeon JH, Cheng TH, Oh IK. Snap-through dynamics of buckled
IPMC actuator. Sens Actuators B Phys 2010;158(2):3005.
[7] Jeon JH, Oh IK. Selective growth of platinum electrodes for
MDOF IPMC actuators. Thin Solid Films 2009;517(17):528892.
[8] Yeom SW, Oh IK. A biomimetic jellyfish robot based on ionic
polymer metal composite actuators. Smart Mater Struct
2009;18(8):085002-1085002-10.
[9] Wang XL, Oh IK, Lu J, Ju JJ, Lee SW. Biomimetic electro-active
polymer based on sulfonated poly(styrene-b-ethylene-cobutylene-b-styrene). Mater Lett 2007;61(29):511720.
[10] Jeon JH, Kang SP, Lee S, Oh IK. Novel biomimetic actuator
based on SPEEK and PVDF. Sens Actuators B Chem
2009;143(1):35764.
[11] Jeon JH, Oh IK, Kee CD, Kim SJ. Bacterial cellulose actuator
with electrically driven bending deformation in hydrated
condition. Sens Actuators B Chem 2010;146(1):30713.
[12] Fujiwara N, Asaka K, Nishimura Y, Oguro K, Torikai E.
Preparation of gold-solid polymer electrolyte composites as
electric stimuli-responsive materials. Chem Mater
2000;12(6):17504.
[13] Liu W, Zhao YH, Li Y, Jiang Q, Lavernia EJ. Electric field
induced reversible switch in hydrogen storage based on
single-layer and bilayer graphenes. Carbon
2009;47(15):345260.
[14] Soldano C, Mahmood A, Dujardin E. Production, properties
and potential of graphene. Carbon 2010;48(8):212750.
[15] Khan U, May P, ONeill A, Coleman JN. Development of stiff,
strong, yet tough composites by the addition of solvent

[23]
[24]
[25]

[26]
[27]

[28]

[29]

[30]

[31]

[32]

[33]

[34]

[35]

[36]
[37]

exfoliated graphene next term to polyurethane. Carbon


2010;48(14):403541.
Cao Y, Feng J, Wu P. Preparation of organically dispersible
graphene nanosheet powders through a lyophilization
method and their poly(lactic acid) composites. Carbon
2010;48(13):38349.
Satti A, Larpent P, Gunko Y. Improvement of mechanical
properties of graphene oxide/poly(allylamine) composites by
chemical crosslinking. Carbon 2010;48(12):337681.
Wang XL, Oh IK, Kim JB. Enhanced electromechanical
performance of carbon nano-fiber reinforced sulfonated
poly(styrene-b-[ethylene/butylene]-b-styrene) actuator.
Comp Sci Technol 2009;69(13):2098101.
Jung JY, Oh IK. Novel nano-composite actuator based on
sulfonated poly(styrene-b-ethylene-co-butylene-b-styrene)
polymer. J Nanosci Nanotech 2007;7(11):37403.
Jung JH, Sridhar V, Oh IK. Electro-active nano-composite
actuators based on fullerene reinforced Nafion. Comp Sci
Technol 2010;70(4):58492.
Landi BJ, Raffaelle RP, Heben MJ, Alleman JL, Vanderveer W,
Gennett T. Single wall carbon nanotubeNafion composite
actuators. Nano Lett 2002;2(11):132932.
Lian Y, Liu Y, Jiang T, Shu J, Lian H, Cao M. Enhanced
electromechanical performance of graphite oxideNafion
nanocomposite actuator. J Phys Chem C 2010;114(21):965963.
Russell TP. The characterization of polymer interfaces. Annu
Rev Mater Sci 1991;21:24968.
Park S, Ruoff RS. Chemical methods for the production of
graphenes. Nat Nanotech 2009;4(4):21724.
Jeon JH, Yeom SW, Oh IK. Fabrication and actuation of ionic
polymer metal composites patterned by combining
electroplating with electroless plating. Compos A
2008;39(4):58896.
Ludvigsson M, Lindgren J, Tegenfeldt J. Crystallinity in Nafion.
J Electrochem Soc 2000;147(4):13035.
Park YS, Yamazaki Y. Low water/methanol permeable Nafion/
CHP organicinorganic composite membrane with high
crystallinity. Eur Polym J 2006;42(2):37587.
Liang Z, Chen W, Liu J, Wang S, Zhou Z, Li W, et al. FT-IR study
of the microstructure of Nafion membrane. J Membr Sci
2004;233(1):3944.
Xu Y, Hong W, Bai H, Li C, Shi G. Strong and ductile poly(vinyl
alcohol)/graphene oxide composite films with a layered
structure. Carbon 2009;47(15):353843.
Ansari S, Kelarkis A, Estevez L, Giannelis EP. Oriented arrays
of graphene in a polymer matrix by in situ reduction of
graphite oxide nanosheets. Small 2010;6(2):2059.
Fong H, Vaia RA, Sanders JH, Lincoln D, Vreugdenhil AJ, Liu
WD, et al. Self-passivation of polymer-layered silicate nanocomposites. Chem Mater 2001;13(11):41239.
Kojima Y, Usuki A, Kawasumi M, Okaua A, Kurachi T,
Kamigaito O, et al. Fine structure of nylon-6-clay hybrid. J
Polym Sci B Polym Phys 1994;32(4):62530.
Rafiee MA, Rafiee J, Srivatsava I, Wang Z, Song H, Yu ZZ, et al.
Fracture and fatigue in graphene nanocomposites. Small
2010;6(2):17983.
Nam JD, Choi HR, Tak YS, Kim KJ. Novel electroactive, silicate
nano-composites prepared to be used as actuators and
artificial muscles. Sens Actuators A Phys 2003;105(1):8390.
Song MK, Park SB, Kim YT, Kim KH, Min SK, Rhee HW.
Characterization of polymer-layered silicate nanocomposites membranes for direct methanol fuel cells.
Electrochim Acta 2004;50(2):63943.
Choudalakis G, Gotsis AD. Permeability of polymer/clay
nano-composites: a review. Eur Polym J 2009;45(4):96784.
Balazs AC, Emrick T, Russell TP. Nanoparticle polymer
composites: where two small worlds meet. Science
2006;314(5802):110710.

CARBON

4 9 ( 20 1 1 ) 1 2 7 912 8 9

[38] Li J, Wilmsmeyer KG, Hou J, Madsen LA. The role of water in


transport of ionic liquids in polymeric artificial muscle
actuators. Soft Matter 2009;5(13):2596602.
[39] Gardner CL, Anantaraman AV. Studies on ion-exchange
membranes. II. Measurement of the anisotropic conductance
of Nafion. J Elect Chem 1998;449(1):20914.
[40] Iwamoto R, Oguro K, Sato M, Iseki Y. Water in perfluorinated
sulfonic acid Nafion membranes. J Phys Chem B
2002;106(28):69739.
[41] Hoffmann EA, Fekete ZA, Korugic-Karasaz LS, Karasz FE,
Wilsuz E. Theoretical and experimental X-ray photoelectron
spectroscopy investigation of ion-implanted Nafion. J Polym
Sci A Polym Chem 2004;42(3):5516.
[42] Landi BJ, Raffaelle RP, Heben MJ, Alleman JL, VanDerveer W,
Gennett T. Development and characterization of single wall

[43]

[44]

[45]

[46]

1289

carbon nanotubeNafion composite actuators. Mater Sci Eng


B 2005;116(3):35962.
Paquette JW, Kim KJ. Ionomeric electroactive polymer
artificial muscle for naval applications. IEEE J Oceanic Eng
2004;29(3):72937.
Oh IK, Jung JH, Jeon JH, Sridhar V. Electro-chemo-mechanical
characteristics of fullerene-reinforced ionic polymermetal
composite transducers. Smart Mater Struct
2010;19(7):075009-1075009-11.
Oh IK, Jung JY. Biomimetic nano-composite actuators based
on carbon nanotubes and ionic polymers. J Int Mater Syst
Struct 2008;19(3):30511.
Meirovitch L. Analytical methods in vibrations. New
Jersey: Prentice Hall; 1967. p. 16166.

Das könnte Ihnen auch gefallen