Vous êtes sur la page 1sur 9

CARBON

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

available at www.sciencedirect.com

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

Electrical and piezoresistive properties of multi-walled carbon


nanotube/polymer composite films aligned by an electric field
A.I. Oliva-Aviles

a,*
,

F. Aviles b, V. Sosa

Centro de Investigacion y de Estudios Avanzados, Unidad Merida, Departamento de Fsica Aplicada, A.P. 73-Cordemex, 97310 Merida,
Yucatan, Mexico
b
Centro de Investigacion Cientfica de Yucatan, Unidad de Materiales, Calle 43, No. 130 Col. Chuburna de Hidalgo, 97200 Merida, Yucatan,
Mexico

A R T I C L E I N F O

A B S T R A C T

Article history:

Aligned multi-walled carbon nanotube (MWCNT)/polymer composite films are prepared by

Received 11 October 2010

solution casting in the presence of an alternating electric field. Application of 7 kV/m at a

Accepted 6 March 2011

frequency of 60 Hz to the polymer composite melt induces MWCNT alignment in the direc-

Available online 12 March 2011

tion of the applied field, which is maintained after polymer crystallization. The electrical
conductivity and piezoresistive response of electric-field-aligned and randomly oriented
0.10.75 wt% MWCNT/polysulfone films are evaluated. Electrical conductivity is 35 orders
of magnitude higher for composites with electric-field-aligned MWCNTs than for randomly
oriented composites. MWCNT alignment inside the polymer matrix also increases the film
piezoresistive sensitivity, enhancing the strain sensing capabilities of the composite film.
 2011 Elsevier Ltd. All rights reserved.

1.

Introduction

Carbon nanotube (CNT)/polymer composites are among the


most promising advanced materials. The outstanding physical properties of CNTs [1,2] and the versatility of polymers
render multifunctional properties to CNT/polymer composites having both sensing and actuating capabilities [36]. Such
multifunctional properties are currently demanded by novel
mechanical, electronic, biological and energy applications
[7,8]. Baughman et al. [9] are among the pioneering groups
who first reported electromechanical actuation of CNTs. They
found that application of an electrochemical voltage to a
single-walled carbon nanotube (SWCNT) sheet induces sheet
deformation, i.e. a piezoelectric response. The variation of
CNT electrical resistivity with applied strain (piezoresistivity)
has been investigated since the pioneering work of Tombler
et al. [10]; large, reversible changes in electrical conductivity
were observed by this group when they locally deformed a
metallic SWCNT using an atomic force microscope. The

intrinsic coupling between electrical resistivity and mechanical deformation of CNTs renders multifunctional properties
and sensing capabilities to composite materials employing
these nanostructures [1115]. A highly desirable feature of
these composite materials is the possibility of aligning CNTs
inside the host matrix. Alignment of CNTs in CNT/polymer
composites induces anisotropy in the composite, improving
its physical properties in the alignment direction [1619].
Among the CNT alignment techniques commonly attempted,
application of an electric field during composite fabrication
constitutes an attractive option. Manipulating CNTs by an
electric or magnetic field during composite processing has
the advantage of avoiding generation of high shear forces in
the melt that may damage the CNT structure [20]. However,
although electromagnetic alignment of isolated single- and
multi-walled CNTs has been quite successful [2123], the attempts of aligning CNTs inside a polymer matrix [2428] have
been only moderately successful. The current state of the art
of magnetic CNT alignment inside a polymer demands very

* Corresponding author: Fax: +52 9999 812923.


E-mail address: andresivan19@hotmail.com (A.I. Oliva-Aviles).
0008-6223/$ - see front matter  2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.carbon.2011.03.017

2990

CARBON

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

high magnetic fields (>9 T) to overcome the viscous forces of


the polymer, and only limited improvements in the electrical
conductivity of the resulting composite have been observed
[19,24,25].
For example, Steinert and Dean [19] partially oriented
SWCNTs inside a polyethylene terephthalate matrix using a
9.4 T magnetic field to form a composite film. They employed
0.5, 1.0 and 3.0 wt% SWCNT loading concentrations and
found better alignment for the lowest CNT concentration.
However, similar values of electrical conductivity
(5 106 S/m) were found for the films that were fabricated
with and without magnetic field. Martin et al. [26] reported
MWCNT alignment in an epoxy matrix using direct current
(DC) and alternating current (AC) electric fields. Based on optical transmission micrographs and conductivity measurements, they reported that although both DC and AC electric
fields can induce alignment of MWCNTs, AC fields are more
efficient to this aim. This fact was evidenced by the formation
of a more uniform and better aligned CNT network for the AC
field. Park et al. [27] conducted a detailed study of the effect of
AC electric field on SWCNT alignment into a viscous photopolymerizable matrix. These authors found that the electrical
conductivity and dielectric properties of the composite can be
controlled by varying the magnitude, frequency and time of
application of the AC electric field. They suggest that optimum conditions for SWCNT alignment occur when applying
an electric field of 43.5 kV/m at 100 Hz. Recently, Ma et al.
[28] aligned pristine and oxidized MWCNTs in a polymethylmethacrylate (PMMA) matrix, using an AC electric field of
15 kV/m at 500 Hz during in situ polymerization. Alignment
of 0.1 wt% MWCNTs inside the PMMA matrix was evident at
the macroscale, and also reflected on the Raman spectrum
and electrical conductivity of the composite. Besides electrical conductivity, it seems reasonable to suppose that the
anisotropy of an aligned network may also increase the piezoresistive sensitivity of the composite. Recently, the use of
CNTs and its polymer composite films as strain sensors has
been investigated [1215,29]. Dharap et al. fabricated a
10 lm thick film made of entangled, randomly oriented bundles of SWCNTs (buckypaper) finding a good piezoresistive response for small deformation loading cycles [13]. Kang et al.
showed that the use of a polymer matrix in strain sensing
applications prevents CNT slippage and hence improves the
strain transfer across the polymer sensor [14]. Bautista et al.
reported the strain sensing capabilities of piezoresistive
MWCNT/polysulfone films with randomly oriented MWCNTs.
The maximum gage factor found for these polysulfone films
was 0.74, corresponding to films loaded with MWCNTs at
0.5 wt% [15]. Theodosiou and Saravanos [11] modeled the
piezoresistive mechanism of CNT/polymer composites
employing a multiscale modeling strategy. They predict that
CNT alignment in the direction of the applied strain should
increase the piezoresistive sensitivity of the composite. However, there seems to be no experimental studies of the piezoresistive behavior of CNT/polymer composites with aligned
CNTs.

1
2

The aim of this work is to align pristine MWCNTs inside a


thermoplastic polymer matrix by means of an alternating
electric field, and investigate the influence of MWCNT alignment in the electrical and piezoresistive properties of the
composite films.

2.

Experimental

2.1.

Fabrication of MWCNT/polysulfone composites

MWCNTs grown by chemical vapor deposition (Baytubes


C150P) were supplied by Bayer Material Science1. The
MWCNTs have inner and outer diameters of approximately
4 nm and 13 nm, respectively, and length in the range of 1
4 lm, yielding a distribution of CNT aspect ratio ranging from
80 to 300 [30]. The polymer matrix used in this work was
polysulfone (PSF, UDEL P1700) supplied by Solvay Advanced
Polymers2. PSF is an engineering thermoplastic with high
thermal stability and low moisture absorption characteristics.
MWCNT/PSF composite films (nominally 200 lm thick) were
fabricated by solution casting using 0.1, 0.3, 0.5 and 0.75
weight percent (wt%). To fabricate the composite films, 2 g
of PSF pellets were dissolved in 10 ml of chloroform (CHCl3)
forming a viscous PSF solution (Sol 1). In parallel, MWCNTs
were dispersed in CHCl3 (Sol 2) using an ultrasonic probe
(VC750, Sonics & Materials, 150 W, 20 kHz) for 30 min in a cold
bath. This CNT sonication was not applied continuously but
in intervals of five minutes, leaving 30 s between each interval
for cooling. Sol 1 and Sol 2 were then mixed in a common beaker, and the MWCNT/PSF/CHCl3 final solution was cast onto
two types of glass molds, depending upon the application or
not of the electric field. A custom made glass mold with aluminum electrodes was fabricated and employed for application of the AC electric field (EAC), while a common petri dish
was employed for the reference films fabricated without electric field application, see Fig. 1. The value of EAC applied to the
films was 6.0 and 7.3 kV/m for an electrode separation of
20 mm (electrical characterization) and 30 mm (piezoresistive
characterization), respectively. This was achieved through
alternating voltages of 110 and 220 V at 60 Hz. Several initial
tests (not reported here) showed that CNT alignment was promoted by short electrode separation, and hence a 20 mm separation was chosen for samples designed for electrical
characterization. However, conventional tension testing demands the use of longer specimens to promote a uniform
stress state within the specimen, and a 30 mm electrode separation had to be employed for the specimens designed for
piezoresistive characterization (tension).
After casting, CHCl3 was evaporated for approximately
12 h which transformed the viscuous solution into a solid
film. Solid 200 lm thick MWCNT/PSF films were obtained
in this way and subsequently dried in an oven at 100 C for
24 h, to eliminate residual solvent. Five specimens were obtained from each of these films and employed for morphological, electrical and piezoresistive characterization. Although
the aligning process was repeated several times, the results

Bayer Material Science. Leverkusen, Germany. www.baytubes.com.


Solvay Advanced Polymers. Alpharetta, Georgia, USA. www.solvayadvancedpolymers.com.

CARBON

4 9 ( 20 1 1 ) 2 9 8 929 9 7

2991

Fig. 1 MWCNT/PSF/CHCl3 solution casting with (left) and without (right) application of EAC.

reported herein correspond to samples extracted from the


same batch (for identical wt%), since it is known that electrical properties of MWCNT/polymer films are highly dependent
on the final dispersion state of the CNTs [31].

2.2.

Morphologic examination

MWCNT/PSF films fabricated with and without EAC were observed at the macroscale using an optical microscope Motic
Digital Microscope DM143 (NTSC System, 10 magnification).
At the nanoscale, transmission electron microscopy (TEM)
images were obtained using an FEI-TITAN TEM microscope
operated at 300 kV. For TEM analysis, MWCNT/PSF films were
sliced using an ultramicrotome (Leica Ultracut UCT), taking
into account the direction of the applied field. For the case
of aligned films, the microtoming direction corresponds to y
in Fig. 2, which is perpendicular to the direction of EAC (x in
Fig. 2).

2.3.

Electrical characterization

DC electrical resistance of MWCNT/PSF films was measured


at room temperature (25 C) with a high impedance electrometer Keithley 6517B. For the resistance measurement, two silver paint electrodes located at the film edges were used. The
specimen length was 15 mm with 3 mm long electrodes,

Fig. 2 MWCNT/PSF specimen for electrical


characterization. Dimensions in mm.

leaving an effective span (L) of 9 mm between the silver


electrodes, see Fig. 2 (dimensions in mm). The specimen
width was 5 mm and its nominal thickness was 200 lm. In
order to minimize surface effects in the measurements, silver
paint electrodes were painted completely covering the ends of
the specimens.
A DC voltage of 40 V was applied between the electrodes
and the electrical conductivity (re) was calculated using the
measured electrical resistance (R) and the specimen dimensions as,
re

L
AR

where A is the cross-sectional area of the specimen.

2.4.

Piezoresistive characterization

MWCNT/PSF films in dumbbell geometry were tested in axial


tension for the piezoresistive characterization, Fig. 3. The
specimens were 25 mm long, with 3 mm narrowest (central)
width and 200 lm thickness. Specimen geometry and
dimensions were selected by rescaling the specimen type IV
recommended by the ASTM standard D638 [32]. End tabbing
was achieved by wrapping 5 mm of the film ends with conventional masking tape. The dog-bone geometry ensured
fracture at the gage section, sufficiently far from the stress

Fig. 3 Specimen setup for piezoresistive characterization.

2992

CARBON

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

concentration zones. Tensile specimens were tested in a


small universal testing machine specially designed for thin
film materials [33]. Displacement was applied to the film by
means of the machine cross-head motion at a speed of
1 mm/min, while the corresponding force was measured by
the machine load cell and converted to axial stress (r).
Mechanical strain (e) was calculated as the machine crosshead displacement normalized by the gage length of the test
specimen. Two copper cables were cemented on the film surface using silver paint on one side of the specimen, to act as
electrodes (see Fig 3). The cables were centered at the midspan of the specimen, separated 7 mm.
Electrical resistance was measured in situ during tensile
testing employing a portable Fluke 289 multimeter with data
logging capabilities. Prior to mechanical testing, the initial
electrical resistance (Ro) was recorded for 30 s to guarantee
stability of the readings. Only MWCNT/PSF films loaded with
0.75 wt% MWCNTs (aligned and randomly oriented) and
0.5 wt% MWCNTs (aligned) were characterized in terms of
piezoresistivity. The high electric resistance of films with randomly oriented MWCNTs at 0.5 wt% and films with lower
MWCNT concentrations prohibited their piezoresistive characterization. The piezoresistive sensitivity of the MWCNT/
PSF films was quantified by the gage factor (k), defined as,
k

DR 1

Ro e

k was measured in the elastic zone of the film (e < 1.5%),


although some tests were conducted up to film fracture.
MWCNT/PSF films were also tensile tested under cycling loading, employing ten loadingunloading cycles in the linear DR
vs. e regime (e < 1.5%) followed by a final loading up to film
fracture.

3.

Results and discussion

3.1.

Morphology

Transmission optical images (10) of MWCNT/PSF films with


different MWCNT concentrations are shown in Fig. 4. The
direction of EAC in Fig. 4 is vertical, and the scale bar is
1 mm. For 0.1%, 0.3% and 0.5% wt%, alignment of MWCNTs
within the polymer is evidenced by the formation of a
MWCNT network aligned in the direction of the applied electric field. For 0.1 wt%, good dispersion and distribution of
MWCNTs in the polymer is observed for both cases, aligned
and randomly oriented. As the CNT concentration increases,
MWCNTs tend to agglomerate due to increased nanotube-tonanotube interactions, but alignment is still evident up to
0.5 wt%. For the highest concentration employed (0.75 wt%),
it is not possible to appreciate differences between films fabricated with and without application of EAC using this optical
technique. For this high weight loading, the morphology of
the network looks plain at this scale because of the high density of MWCNTs which compose the network. The macroscopic morphology of the aligned films found in this work is
similar to that reported by Ma et al. [28] in their pioneering
work using 0.1 wt% MWCNTs aligned in a PMMA matrix during in situ polymerization.

Fig. 5 shows TEM images of 0.5 wt% MWCNT/PSF films fabricated without (Fig. 5a) and with (Fig. 5b) application of EAC.
To avoid possible confusion between the cutting marks of
the ultramicrotome and the CNT orientation, the microtome
cutting direction was along the y-axis in Fig. 2, while the
direction of EAC was along the x-axis. The direction of EAC
was tracked during the cutting and imaging process and is
indicated in Fig. 5b by an arrow. In this case, the direction of
EAC in Fig. 5b is not vertical given a practical limitation of
our TEM set-up to rotate the sample around the z-axis (out
of the film plane).
When the film is fabricated without assistance of electric
field, individual MWCNTs look curled and randomly oriented,
Fig. 5a. For films fabricated with assistance of EAC, however,
some of the MWCNTs look fairly straight and aligned in the
direction of the electric field, Fig. 5b. A relatively long MWCNT
of about 0.7 lm length (bottom left) is observed in Fig. 5b
aligned in the direction of EAC. The CNT, however, is not totally straight but kinks to form two straight sections. In the
same figure, a moderately shorter CNT (top right) is observed
forming a small angle with the direction of EAC. Shorter CNTs
are shown curled and/or randomly oriented. Analysis of several TEM images consistently suggested that the MWCNTs inside the polymer are only partially aligned at the nanoscale
(as individual nanotubes), and this local alignment is more
effective for longer CNTs than for shorter ones. The currently
proposed mechanism for electrical alignment of CNTs suggests that, in presence of an electric field, CNTs experience
a polarization which leads to a torque acting on the nanotube
[21,22,26]. This torque tends to align the nanotube in the field
direction against the viscous drag forces of the surrounding
polymer. In addition, for an AC electric field, inhomogeneities
in the electric field should appear at the CNT ends and at its
defective sites, e.g. coulombic attraction between oppositely
charged ends. Thus, the non-uniform electric field in the
vicinity of the nanotube tips results in movement of the induced dipoles towards the zone with the highest field
strength, a behavior called dielectrophoresis [22,26,34]. Since
it is expected that longer CNTs have larger dipole moments,
this hypothesis may explain why longer MWCNTs tend to
align easily than shorter ones in our experimental system.

3.2.

Electrical properties

Fig. 6 presents a logarithmic plot of the measured DC electrical conductivity of the examined MWCNT/PSF films. The
average conductivity of the neat PSF (unfilled polymer) was
measured as 1.55 1015 S/m and is indicated in Fig. 6 as
0 wt%. For 0.1 wt%, inclusion of randomly oriented MWCNTs
to the polymeric matrix increased the conductivity in 8 orders
of magnitude with respect to the neat polymer. This significant increment in conductivity confirms that a percolative
network of MWCNTs has been formed at 0.1 wt%. This result
is in agreement with a previously measured percolation
threshold of 0.068 wt%, reported by Bautista et al. for the
same material system with randomly oriented MWCNTs
[15]. The conductivity of the composite films with randomly
oriented MWCNTs at 0.1 wt% is however, still of the order of
107 S/m. At the same concentration (0.1 wt%), the conductivity of the aligned films in the direction of EAC is three orders of

CARBON

4 9 ( 20 1 1 ) 2 9 8 929 9 7

2993

Fig. 4 Optical images of MWCNT/PSF films. Rows correspond to identical MWCNTweight percent and columns to absence or
presence of electric field during film fabrication. Scale bar is 1 mm.

Fig. 5 TEM images of 0.5 wt% MWCNT/PSF films fabricated without and with assistance of electric field. (a) EAC = 0, (b)
EAC = 6 kV/m.

magnitude higher than that of films with randomly oriented


MWCNTs. A similar trend between the conductivity of films
with electric-field-aligned MWCNTs and randomly oriented
ones was observed for films with 0.3% and 0.5 wt% MWCNTs,
being the electrical conductivity five orders of magnitude
higher for the aligned films. It is thus evident that alignment
of MWCNTs inside the polymer greatly improves the conductivity of the composite films in the alignment direction. The

existence of an aligned CNT network inside the polymer matrix enhances the electron flux in a preferential direction
through the composite because the probability of contact between nanotubes increases in the aligning direction [6,11]. For
0.75 wt%, the conductivity of samples fabricated with and
without application of EAC is similar, so the electric field does
not show a significant effect on the electrical conductivity of
these composites. At this high concentration the CNT

2994

CARBON

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

3.3.

Fig. 6 Electrical conductivity of MWCNT/PSF films


fabricated with and without application of EAC.

network is densely packed, see Fig. 4, forming a network


whose conductivity is insensitive to the applied electric field.
This effect is likely caused by a decreased MWCNT mobility at
high concentrations, which promotes agglomerations and
hence difficults CNT alignment, see e.g. [19,31]. Therefore,
correlating the microstructure of the CNT network (Fig. 4)
with the electrical conductivity of the films (Fig. 6), it is clear
than films with relatively low CNT concentrations are more
prone to form well dispersed networks of aligned CNTs when
an AC electric field is applied.

Strain sensing capabilities

In order to investigate the piezoresistive response of the films,


electrical resistance of MWCNT/PSF films was measured
in situ while the films were mechanically strained. Fig. 7
shows the normalized change in electrical resistance (DR/Ro)
vs. mechanical strain (e) within the linear elastic regime of
the polymer film (e < 1.5%) for 0.5 and 0.75 wt%. Experimental
measurements are represented by dots while solid lines are
linear fittings.
Table 1 presents the average gage factors (k, Eq. (2)) and
their standard deviations obtained from the linear fit of the
DR/Ro vs. e curves of five replicates, as shown in Fig. 7. As mentioned earlier, the films with 0.5 wt% fabricated without assistance of EAC presented low conductivity, see Fig. 6, which did
not allow the detection of a piezoresistive signal for those
films.

Table 1 Measured gage factors for MWCNT/PSF films.


EAC (kV/m)

MWNTCs (wt%)

Gage factor (k)

7.3
7.3
0
0

0.5
0.75
0.5
0.75

2.78 0.42
1.49 0.30
a
0.70 0.18

a Not possible to measure.

Fig. 7 Piezoresistive characterization of MWCNT/PSF films in the film elastic zone. (a) 0.5 wt% EAC = 7.3 kV/m, (b) 0.75 wt%
EAC = 7.3 kV/m, (c) 0.75 wt% EAC = 0.

2995

4 9 ( 20 1 1 ) 2 9 8 929 9 7

For specimens fabricated with the aid of EAC (Figs. 7a and


b) a fairly linear piezoresistive behavior is observed for the
whole strain range examined in Fig. 7 (e 6 1.5%). The linear
behavior in this deformation range was consistently observed
in the samples that were aligned by EAC. On the other hand,
for the same range e, a change in slope was typically observed
for films fabricated without assistance of electric field
(Fig. 7c). The piezoresistive behavior of films with 0.75 wt%
randomly oriented MWCNTs may be described by two linear
functions with different slopes, the first one having an initially low slope (k = 0.70) for e < 0.6%, followed by a second linear zone of increased slope for e > 0.6%. The average gage
factor associated to the second slope (k = 1.48, not listed in Table 1) is close to the value obtained for films with the same
CNT weight loading (0.75 wt%) but fabricated with assistance
of electric field, see Table 1. This second slope could correspond to an initial CNT rearrangement driven by the strain
applied to the film. The largest gage factor (k = 2.78) was
achieved by films with 0.5 wt% MWCNTs aligned by EAC. This
gage factor is about four times higher than the one obtained
for films fabricated without assistance of the electric field
(see Table 1). Thus, MWCNT alignment inside the polymer
matrix promotes increased sensitivity and linearity of the
piezoresistive (DR/Ro vs. e) response. Increased piezoresistive
sensitivity for polymer composites with aligned CNTs was recently predicted by Theodosiou and Saravanos using multiscale modeling [11], although it had not been
experimentally reported. Theodosiou and Saravanos predict
a strong correlation between CNT alignment and piezoresistive sensitivity, with increased sensitivity as the CNTs align
in the direction of the applied strain. According to these
authors, for CNTs oriented 30 off the loading axis the decrease in sensitivity with respect to the aligned case is about
30%. Incidentally, the gage factor predicted by these authors
for the system with CNTs aligned in the direction of the applied strain is k = 2.457, in good agreement with the values
found herein for films with aligned CNTs (k = 2.78). A true
quantitative comparison with Theodosiou and Saravanos
work, however, cannot be made since their modeling strategy
does not include the effect of the matrix and the CNT weight
percent employed in the piezoresisitive analysis is not
available.
Table 1 also shows that gage factors for films with 0.75 wt%
are lower than those for films with 0.5 wt%. The decrease of
the piezoresistive sensitivity with increased CNT concentration has also been theoretically predicted [11,35]. The electrical conductivity of the composite film depends, among many
other factors, on the CNT concentration. For high concentrations there is more contact between CNTs, which increases
the density of conductive paths, and hence the film conductivity. When the network is densely packed, no further important increments in electrical conductivity are expected with
increasing CNT content, see Fig. 6. Thus, the relation between
electrical conductivity and CNT weight loading is monotonic,
reaching a plateau for high loadings. For piezoresistivity, on
the other hand, the relation between strain sensitivity and
CNTweight loading seems to be non-monotonic. As discussed
in previous works, the dominant parameters on the film
piezoresistive response seem to be the geometry of the percolative network, the CNT nano-structure and the relative con-

tribution of the tunneling resistance to the total electrical


resistance [15,35,36]. Both, the construction of the CNT network and the capability of increasing the tunneling resistance
are optimized for low CNT concentrations, since a better dispersion is achieved [35]. However, very low CNT weight loadings yield poor film conductivity, which greatly increases
the difficulty of measuring DR. Thus, although the most sensitive CNT concentration might occur slightly above electrical
percolation (see Hu et al. [35]), the existent compromise between piezoresistive sensitivity and electrical conductivity
demand the use of CNT loadings well above percolation, so
changes in electrical resistance can be measured by unsophisticated instruments.
The piezoresistive response of MWCNT/PSF films was also
investigated up to film fracture. Fig. 8 shows a DR/Ro vs. e
curve for a film with 0.5 wt% aligned MWCNTs. A fairly linear
behavior is detected for small deformations (e < 1.5%), but the
complete curve is clearly nonlinear. When the mechanical
deformation exceeds 1.5%, the curve increases in a nonlinear
fashion, gradually increasing its slope until film fracture. At
film fracture, a sudden increase in DR occurs since the continuity of the MWCNT network inside the polymer is disrupted.
Nonlinearity of the piezoresistive response of CNT/polymer
composites under high deformation has been previously observed [15,37]. Park et al. [37] suggested that for deformations
in the elastic range of the polymer matrix, CNTs are still in
contact inside the polymer matrix so that piezoresistive effects can be explained by common percolative theories. However, for higher deformations (close to the elasto-plastic
mechanical transition), CNT contacts decrease and the tunneling resistance may play a dominant role in the nonlinear
behavior. According to Hu et al. [35], increasing the tunneling
resistance improves the composite electromechanical sensitivity, and that may be the reason for the increased sensitivity
(slope) observed in the large deformation zone. In our case, it
is also possible that the large applied strain yields certain CNT
reorientation inside the polymer, since our TEM observations
suggest that our films are only partially oriented.
MWCNT/PSF films were also tested using ten loading
unloading cycles in the linear elastic range, followed by a final
loading excursion up to film fracture, see Fig. 9. For each cycle,
the film stress increases linearly with the applied tensile
20
16

R/Ro(%)

CARBON

0.5 wt.%
EAC=7.3 kV/m

12
8
4
0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

(%)

Fig. 8 Piezoresistive characterization of an aligned 0.5 wt%


MWCNT/PSF film up to film fracture.

2996

CARBON

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

Fig. 9 Loadingunloading tests of a 0.5 wt% aligned MWCNT/PSF film. (a) Stress (r)-strain (emech) curves, (b) comparison of
emech and eCNT as a function of time.

strain up to a maximum, which corresponds to the point


where the displacement of the machine cross-head was
stopped, Fig. 9a. After this maximum (but before strain reversal), a slight decrease in stress occurs while the machine
cross-head movement was fixed, evidencing polymer relaxation. Upon application of strain reversal (unloading) the stress
returns to zero with a slope similar to that observed during
the loading excursion. Fig. 9b presents strain vs. time curves
for MWCNT/PSF films with 0.5 wt% aligned CNTs. The solid
line represents the mechanical strain applied by the crosshead displacement of the universal testing machine and has
been labeled as emech. The dashed line (eCNT) represents the
strain measured by the CNT-composite film and was obtained
using Eq. (2), the measured changes in electrical resistance,
and the gage factor of the tested film (k = 3.05). Good agreement between the applied strain and the strain measured
by the MWCNT/PSF film is observed for all cycles. A slight decay in the eCNT signal over the time when the applied emech remained constant (30 s) is observed in the first three cycles,
which may be related to viscoelastic relaxation of the polymer
matrix, as pointed out in Fig. 9a. The eCNT (DR) signal of the
nanostructured sensor returns to zero after strain reversal
and increases again upon reloading, evidencing the capability
of the polymer composite film to follow the applied strain under cycling loading.

4.

Conclusions

A simple method to fabricate composite films based on a thermoplastic polymer (polysulfone) with MWCNTs aligned in a
preferential direction by means of an electric field was developed. The application of an AC electric field of relatively low
magnitude (7 kV/m) and low frequency (60 Hz) promoted the
formation of an aligned MWCNT network inside the polymer
matrix. The oriented morphology of the network was evident
at the macroscale, although only partial alignment was detected at the nanoscale. The aligned CNT network resulted
in improved electrical and piezoresistive sensing capabilities
of the composite films. For CNT weight loadings of 0.1, 0.3
and 0.5 wt%, MWCNT alignment into the polymer matrix improved the electrical conductivity of the films by 35 orders of

magnitude with respect to films with randomly oriented


MWCNTs. No improvement in electrical conductivity was observed for films with 0.75 wt%. The piezoresistive sensitivity
(gage factor) and linearity of the DR/Ro vs. e signal improved
for films with electric-field-aligned CNTs, and was slightly
higher than that of common metal foil gages (2.1). The gage
factor was higher for films with 0.5 wt% than for films with
0.75 wt%. The aligned films also presented a piezoresistive response capable of reproducing the applied strain upon several
loadingunloading cycles. The piezoresistive properties
exhibited by these MWCNT/polymer films, the capability of
tailoring the CNT network by means of an electric field, and
the strong structureproperty correlation observed, provide
further directions towards the development of strain-sensing
nanostructured materials.

Acknowledgments
This work was supported by CONACYT (Mexico) through Project Number 79609 of Dr. Aviles. Facilities and support for
mechanical testing were kindly provided by Dr. A.I. Oliva at
CINVESTAV-Merida. The authors also wish to thank Emilio
Corona, Fidel Gamboa, Beatriz Heredia (CINVESTAV) and
Alejandro May (CICY) for their valuable technical support.
Assistance of Arturo Ponce and Monica Ceniceros (CIQA) with
TEM analysis is strongly appreciated. Authors wish to express
their gratitude to the detailed review conducted by one of the
anonymous reviewers of this work (reviewer #1).

R E F E R E N C E S

[1] Thostenson ET, Ren ZF, Chou TW. Advances in the science
and technology of carbon nanotubes and their composites: a
review. Compos Sci Technol 2001;61:1899912.
[2] Che JW, Cagin T, Goddard WA. Thermal conductivity of
carbon nanotubes. Nanotechnology 2000;11:659.
[3] Yu X, Rajamani R, Stelson KA, Cui T. Carbon nanotube-based
transparent thin film acoustic actuators and sensors. Sens
Actuators A 2006;132:62631.

CARBON

4 9 ( 20 1 1 ) 2 9 8 929 9 7

[4] Hierlod C, Jungen A, Stampfer C, Helbling T.


Nanoelectromechanical sensors based on carbon nanotubes.
Sens Actuators A 2007;136:5161.
[5] Li C, Thostenson ET, Chou TW. Sensors and actuators based
on carbon nanotubes and their composites: a review. Compos
Sci Technol 2008;68:122749.
[6] Seidel GD, Lagoudas DC. A micromechanics model for the
electrical conductivity of nanotubepolymer
nanocomposites. J Compos Mater 2009;43:91741.
[7] Vo-Dinh T, Cullum BM, Stokes DL. Nanosensors and biochips:
frontiers in molecular diagnostics. Sens Actuators B
2001;74:211.
[8] Deshmukh S, Ounaies Z. Single walled carbon nanotube
(SWNT)-polymide nanocomposites as electrostrictive
materials. Sens Actuators A 2009;155:24652.
[9] Baughman RH, Cui CX, Zakhidov AA, Iqbal Z, Barisci JN,
Spinks GM, et al. Carbon nanotube actuators. Science
1999;284:13404.
[10] Tombler TW, Zhou CW, Alexseyev L, Kong J, Dai H, Liu L, et al.
Reversible electromechanical characteristics of carbon
nanotubes underlocal-probe manipulation. Nature
2000;405:76972.
[11] Theodosiou TC, Saravanos DA. Numerical investigation of
mechanisms affecting the piezoresistive properties of CNTdoped polymers using multi-scale models. Compos Sci
Technol 2010;70(9):131220.
[12] Grow RJ, Wang Q, Cao J, Wang D, Dai H. Piezoresistance of
carbon nanotubes on deformable thin-film membranes. Appl
Phys Lett 2005;86:931047.
[13] Dharap P, Li Z, Nagarajaiah S, Barrera EV. Nanotube film
based on single-wall carbon nanotubes for strain sensing.
Nanotechnology 2004;15:37982.
[14] Kang I, Schulz MJ, Kim JH, Shanov V, Shi D. A carbon
nanotubes strain sensor for structural health monitoring.
Smart Mater Struct 2006;15:73748.
[15] Bautista-Quijano JR, Aviles F, Aguilar JO, Tapia A. Strain
sensing capabilities of a piezoresistive MWCNT-polysulfone
film. Sens Actuators A 2010;159:13540.
[16] Ren ZF, Huang ZP, Xu JU, Wang JH. Synthesis of large arrays of
well-aligned carbon nanotubes on glass. Science
1998;282:11057.
[17] Sennet M, Welsh E, Wright JB. Dispersion and alignment of
carbon nanotubes in polycarbonate. Appl Phys A
2003;76:1113.
[18] Lanticse LJ, Tanabe Y, Matsui K. Shear-induced preferential
alignment of carbon nanotubes resulted in anisotropic
electrical conductivity of polymer composites. Carbon
2006;44:307886.
[19] Steinert B, Dean D. Magnetic field alignment and electrical
properties of solution cast PET-carbon nanotubes composite
films. Polymer 2009;50:898904.
[20] Fu SY, Chen ZK, Hong S, Han CC. The reduction of carbon
nantube (CNT) length during the manufacture of CNT/
polymer composites and a method to simultaneously
determine the resulting CNT and interfacial strengths.
Carbon 2009;47:3192200.

2997

[21] Bubke K, Gnewuch H, Hempstead M, Hammer J, Green MLH.


Optical anisotropy of dispersed carbon nanotubes induced by
an electric field. Appl Phys Lett 1997;71(14):19068.
[22] Yamamoto K, Akita S, Nakayama Y. Orientation and
purification of carbon nanotubes using ac electrophoresis. J
Phys D Appl Phys 1998;31:L346.
[23] Liu X, Spencer JL, Kaiser AB, Arnold WM. Electric-field
oriented carbon nanotubes in different dielectric solvents.
Curr Appl Phys 2004;4:1258.
[24] Camponeschi E, Vance R, Al-Haik M, Garmestani H,
Tannenbaum R. Properties of carbon nanotube-polymer
composites aligned in a magnetic field. Carbon
2007;45:203746.
[25] Garmestani H, Al-Haik M, Dahmen K, Tannenbaum R, Li D,
Sablin SS, et al. Polymer-mediated alignment of carbon
nanotubes under high magnetic fields. Adv Mater
2003;22:191821.
[26] Martin CA, Sandler JKW, Windle AH, Schwarz MK, Bauhofer
W, Schulte K, et al. Electric field-induced aligned multi-wall
carbon nanotubes networks in epoxy composites. Polymer
2005;46:87786.
[27] Park C, Wilkinson J, Banda S, Ounaies Z, Wise K, Sauti G.
Aligned single-wall carbon nanotubes polymer composite
using an electric field. J Polymer Sci B 2006;44:175162.
[28] Ma PC, Zhang W, Zhu Y, Ji L, Zhang R, Koratkar N, et al.
Alignment and dispersion of functionalized carbon
nanotubes in polymer composites induced by an electric
field. Carbon 2008;46:70610.
[29] Wichmann MHG, Buschhorn ST, Gehrmann J, Schulte K.
Piezoresistive response of epoxy composites with carbon
nanoparticles under tensile load. Phys Rev B 2009;80:24543718.
[30] Aviles F, Ponce A, Cauich-Rodrguez JV, Martnez GT. TEM
examination of MWCNT oxidized by mild experimental
conditions. Fullerenes Nanotubes Carbon Nanostruct 2010;
Accepted.
[31] Aguilar JO, Bautista-Quijano JR, Aviles F. Influence of carbon
nanotube clustering on the electrical conductivity of polymer
composite films. Express Polym Lett 2010;4:2929.
[32] ASTM Standard D638. Standard test method for tensile
properties of plastics. ASTM International, West
Conshohocken, PA, USA, 2002.
[33] Huerta E, Corona JE, Oliva AI, Aviles F, Gonzalez J. Universal
testing machine for mechanical properties of thin materials.
Rev Mex Fis 2010;56(4):31722.
[34] Pohl HA. Dielectrophoresis. Cambridge: Cambridge
University Press; 1978.
[35] Hu N, Karube Y, Arai M, Watanabe T, Yan C, Li Y, et al.
Investigation on sensitivity of a polymer/carbon nanotube
composite strain sensor. Carbon 2010;48:6807.
[36] Zhang W, Suhr J, Koratkar N. Carbon nanotube/polycarbonate
composites as multifunctional strain sensors. J Nanosci
Nanotechnol 2006;6:9604.
[37] Park M, Kim H, Youngblood JP. Strain-dependent electric
resistance of multi-walled carbon nanotubes/polymer
composite films. Nanotechnology 2008;19:05570555711.

Vous aimerez peut-être aussi