Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Materials Research Express

PAPER

Influence of concentration, length and orientation of multiwall carbon


nanotubes on the electromechanical response of polymer
nanocomposites
To cite this article: J Cob et al 2019 Mater. Res. Express 6 115024

View the article online for updates and enhancements.

This content was downloaded from IP address 132.174.255.215 on 24/10/2019 at 15:57


Mater. Res. Express 6 (2019) 115024 https://doi.org/10.1088/2053-1591/ab447b

PAPER

Influence of concentration, length and orientation of multiwall


RECEIVED
13 June 2019
carbon nanotubes on the electromechanical response of polymer
REVISED
22 August 2019
nanocomposites
ACCEPTED FOR PUBLICATION
13 September 2019
PUBLISHED
J Cob1, A I Oliva-Avilés2,4 , F Avilés3 and A I Oliva1
27 September 2019 1
Centro de Investigación y de Estudios Avanzados del IPN, Unidad Mérida, Departamento de Física Aplicada, A.P. 73-Cordemex, 97310,
Mérida, Yucatán, Mexico
2
Universidad Anáhuac Mayab, División de Ingeniería y Ciencias Exactas, Carretera Mérida-Progreso km 15.5 AP 96-Cordemex, CP 97310,
Mérida, Yucatán, Mexico
3
Centro de Investigación Científica de Yucatán A.C., Unidad de Materiales, Calle 43, No. 130×32 y 34, Col. Chuburná de Hidalgo, 97205
Mérida, Yucatán, Mexico
4
Author to whom any correspondence should be addressed.
E-mail: andres.oliva@anahuac.mx

Keywords: carbon nanotubes, high-density ultrasonic energy, piezoresistivity, aspect ratio, polymer composites

Abstract
Multiwall carbon nanotube (MWCNT)/polymethyl methacrylate (PMMA) composites were
fabricated by solution casting with a range of MWCNT concentrations, two groups of MWCNT
lengths and with or without the application of electric field. As-received (CNTs) and fragmented
(FCNTs) MWCNTs were used to fabricate composites with MWCNT concentrations between
0.05 wt% and 0.7 wt%. A well-controlled procedure to systematically fragment the MWCNTs based
on the application of high-density ultrasonic energy was implemented, and MWCNT length
distributions were obtained by a formal statistical analysis. The relationship between the sparseness of
the MWCNT network (macroscale morphology), electrical conductivity and piezoresistive sensitivity
was evaluated. In order to evaluate the effect of the MWCNT alignment on the composite properties,
specific MWCNT concentrations were selected for composites with CNTs and FCNTs under the
application of an electric field. Composites with FCNTs exhibited a sparser network, higher
transparency, lower electrical conductivity, higher percolation threshold and higher piezoresistive
sensitivity than their counterparts with CNTs. On the other hand, composites with electric field-
aligned MWCNT bundles presented higher electrical conductivity and piezoresistive sensitivity than
their non-aligned counterparts, for both fragmented and non-fragmented cases. The evaluation of the
combined effect of carbon nanotube length and alignment on the electromechanical properties of the
composites represents an important step on the understanding of the structure-property relation of
polymer nanocomposites.

1. Introduction

Multifunctionality has become one of the most desired characteristics for the development of novel materials,
and the exploration of the underlying governing mechanisms is of major importance for the scientific
community. Nowadays, it is of great interest that, during operation, materials and structures exhibit features
additional to those that classically motivated their major use in the first place (e.g. mechanical, for a load bearing
structure). In this field, composite materials have demonstrated to adequately fulfill several demands of the
scientific and technologic communities due the combined and enhanced physical properties of matrix and fillers
[1–5]. When the filler dimensions are in the nanoscale, several mechanisms occurring at such low dimensions
can be tracked (e.g. crack propagation, stress, thermal expansion, chemical modifications) and then, a novel type
of sensors emerge, with sensitivities in many cases higher than those of the conventional commercial sensors
[3, 6]. Among the multifunctional materials, polymer nanocomposites reinforced with carbon nanostructures is

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 115024 J Cob et al

a material which has recently gained a lot of interest; several investigations have been carried out to understand
their responses and their use for specific applications, and some good reviews on the topic can be referred
[3, 7, 8]. As a general outcome, the structure-property relationship of nanocomposites plays a role of paramount
importance on the multifunctional response of these materials. Due to the insulating nature of the polymer
matrix, the state and morphology of the electrically conductive network within the polymer mainly define the
material effective response, therefore defining the performance of these novel sensing materials. As such, the
exploration of the influence of each of the parameters defining the conformation of the conducting network
represents a strategic route towards clarifying the understanding of their multifunctional response, being this
desirable for enhanced engineering applications. In particular, polymethyl methacrylate (PMMA) has been
reported as an adequate thermoplastic matrix for different applications of nanocomposites. In a comprehensive
review, Mathur et al [9] highlighted the importance of the fabrication and applications of carbon nanotube
(CNT)/PMMA nanocomposites, quoting the light-weight and high-strength characteristics of this polymer as
major features. For example, PMMA represents the primary component for orthopedic bone cements and the
incorporation of carbon nanoparticles (such as CNTs) largely influences the mechanical, thermal and
rheological properties of the resulting composites [10]. In a very recent study, Nawar and El-Mahalawy reported
the fabrication of a semi-transparent heterostructure Schottky diode based on a multiwall carbon nanotube
(MWCNT)/PMMA composite using a rapid and low-cost method [11]. The authors suggest that such devices
have the capability of working as self-driven UV photodetectors, evidencing the importance of studying such a
material for photovoltaic applications. Regarding the CNT/PMMA composite processing methods, several
techniques have been reported such as melt-mixing, in-situ polymerization, coagulation and solution casting
[9]. Among them, the solution casting technique has the advantage of exhibiting low viscosity, which facilitates a
uniform mixing and dispersion of the fillers. Also, different solvents can be used such as chloroform, toluene and
dimethyl formamide. Reports which support the use of solution casting for nanocomposite preparation can be
found elsewhere [12–17].
Given this background, this work investigates the role of the concentration, length, and orientation of
MWCNTs on the electrical and piezoresistive response of MWCNT/PMMA composites fabricated by solution
casting. Although there are reports exploring the influence of these parameters on the electromechanical
response separately [18–27], here we pretend to shed light on their collaborative (or combined) effect. To
evaluate the role of the MWCNT length, a methodology has been implemented and validated to fragment the
MWCNTs in a controlled manner. For the MWCNT orientation, electric fields are applied to induce MWCNT
alignment and chaining within the thermoplastic polymer matrix (PMMA) during manufacturing. Macroscale
morphology, electrical conductivity and electromechanical sensitivity (gage factor) are evaluated for the
composites and correlated with the configuration of the internal network and main governing mechanisms.

2. Experimental

2.1. Materials and equipment


Commercial MWCNTs with a purity >95% were used (Cheaptubes Inc, Cambridgeport, USA). The mean
external diameter and length are ∼30 nm and ∼2.6 μm, respectively [18]. For the polymer matrix, PMMA (H15-
002-000 Cristal, Plastiglas de México SA de CV) with a density of 1.2 g cm−3 was used. Acetone, chloroform, and
N, N-dimethylformamide (DMF) used were all of reagent grade. For the fragmentation process, an ultrasonic
bath (Fisher Scientific, 110 W at 40 kHz) and an ultrasonic probe (Fisher Scientific FB-505, 500 W at 20 kHz)
were used in different steps. For the microscale observation of the MWCNTs, a field-emission scanning electron
microscope (SEM, JEOL FESEM-7600) was used; the open source software ‘ImageJ’ was used for measuring the
MWCNT length needed for the statistical analysis.

2.2. Carbon nanotube fragmentation procedure


To investigate the effect of the MWCNT length on the properties of the nanocomposites, a procedure to
fragment them in a controlled manner was implemented. The procedure is based on the application of high-
density ultrasonic energy to a MWCNT/liquid solution, which has demonstrated to scissor the MWCNTs as a
result of the concentrated energy given to the system [28–30]. Figure 1 illustrates the MWCNT fragmentation
process. 18 mg of MWCNTs were mechanically dispersed into 307 g (400 ml) of acetone (weight concentration
of 0.0059 wt%) for 30 min. Then, high-density ultrasonic energy was applied by means of an ultrasonic probe
working at 200 W (45% of its capacity) for 2 h effectively (4 h total, 30 s on and 30 s off), for an energy density of
∼92 GJ m−3.
The solution was then heated at 70 °C for 45 min in a hot plate in order to evaporate the acetone. Although
the boiling point of acetone is reported as 56.2 °C [31], a higher temperature was selected as the acetone needs to
be fully evaporated within polymer molecules, since traces may cause polymer plasticization. The fragmented

2
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 1. Methodology for the fragmentation of MWCNTs by using high-density ultrasonic energy.

MWCNTs (named ‘FCNTs’ hereafter) were washed with distilled water in a conventional ultrasonic bath for 1 h
and the water was then evaporated at 100 °C for 30 min. Finally, FCNTs were removed from the container, dried
in a convection oven at 100 °C for 24 h and then stored.
To validate the effectiveness of the fragmentation process, SEM observation of non-fragmented MWCNTs
(named simply ‘CNTs’ hereafter) and FCNTs was carried out and lengths of individual MWCNTs were
measured using a dedicated software (ImageJ [32]). For the sample preparation, 0.9 mg of CNTs and FCNTs
(in separate experiments) were dispersed in 15.35 g of DMF using an ultrasonic bath for 30 min. A droplet was
carefully deposited onto a silicon substrate and the DMF was evaporated at room conditions. A total of ∼1500
MWCNT individual lengths from 110 SEM images were measured and the data were fit to Log-Normal
distributions. The probability density function of the Log-Normal distribution of a variable ‘x’ (length in this
case) is given by [33],
1
f (x ) = e -[ln x - w] / 2b , x > 0
2 2
(1)
xb 2p
where ω and β are the shape and scale parameters, respectively. From these parameters, the mean (μe), median
(μn) and mode (μo) are calculated as [33],
b2
me = e w + 2 (2a)
mn = e w (2b)

mo = e w - b
2
(2c )

2.3. Fabrication of MWCNT/PMMA nanocomposites


MWCNT/PMMA composites (with CNTs and FCNTs) were fabricated by the solution casting technique
[15, 34], with MWCNT concentration varying from 0 (pure polymer) to 0.7 wt%. According to the electrical
percolation plot (shown later), a plateau value for the electrical conductivity is reached around 0.6 wt% for
systems with CNTs and FCNTs; thus, a maximum MWCNT concentration of 0.7 wt% was selected herein.
Figure 2 shows a schematic of the nanocomposite preparation process. 2 g of PMMA were dissolved in 15 ml of
chloroform and mechanically stirred for 1 h. In a parallel process, MWCNTs were dispersed in 10 ml of
chloroform using an ultrasonic bath for 1 h. When both simultaneous processes finished, solutions were mixed
and the MWCNT/PMMA/chloroform solution was first heated at 60 °C for 20 min promoting solvent
evaporation, and then sonicated in a conventional ultrasonic bath (500 W at 20 kHz) for 20 min. This sequence
was repeated 4 times but reducing the heating time to 15, 10, and finally 5 min (sonication remains at 20 min),
in order to reach a value of dynamic viscosity of the solution which is adequate for the solution casting
process (∼2.5 Pa·s). This value of dynamic viscosity was measured at room temperature following the ASTM
D445 standard [35] and was selected as the casting condition based on preliminary experiments where

3
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 2. Methodology used for the fabrication of MWCNT/PMMA composites by solution casting (without and with application of
electric field).

MWCNT/PMMA composites with uniform and homogeneous dispersion were obtained. After this
sequence, the solution was casted in a Petri dish (no electric field) or in a custom-made cell of glass with
aluminum electrodes (electric field application, to promote MWCNT alignment). The applied rms-voltage
was 220 V at 60 Hz and the electrode separation was 30 mm, rendering an electric field of 7.3 kVrms m−1.

2.4. Electrical and piezoresistive characterization


Electrical conductivity and piezoresistive gage factor (i.e. electromechanical sensitivity) of the MWCNT/PMMA
composites were investigated. The electrical conductivity (σ) was calculated from,
L
s= (3)
AR
where L is the separation between electrodes, A the cross-section area, and R the measured electrical resistance.
For low-resistivity composites (R<1 GΩ) rectangular specimens were prepared, electrodes were silver-painted
at their edges (see figure 3(a)), and R was measured with a high-precision digital multimeter (HP 3458 A). For
high-resistance composites (R>1 GΩ), wafers with diameter of 54 cm and nominal thickness of 350 μm were
prepared (see figure 3(b)) and measured with a Keithley 6517B electrometer and a high resistivity test fixture
(Keithley 8009). To avoid errors caused by charge accumulation, polarization and/or tribological charges, an
alternating polarity test method was used for the electrical measurements of high R. In this method, a voltage of
+500 V was first applied for 20 s and the electrical current was measured. Then, the polarity of the voltage is
automatically inverted (−500 V) and then applied for another 20 s, while the electrical current is again registered.
The sequence was repeated for 2 min and the average value was finally recorded. Eight samples for each case were
measured.
Using the electrical conductivity of the composites, the percolation thresholds were calculated using a power
law to describe statistical percolation in the vicinity of the percolation limit [19]. According to percolation
theory, the electrical conductivity (σ) and the particle weight concentration (f) can be correlated by,
s µ (f - fc )t (4)

where fc is the percolation threshold and t is the critical exponent, which has been associated to the
dimensionality of the percolated network within the composite [19].
For the piezoresistive characterization, composites were fabricated rescaling the type IV specimen of the
ASTM D638 standard [36], see figure 4(a), and the mechanical tests were carried out using a dedicated universal
testing machine for thin materials [37]. Copper wires (AWG-36, 0.127 mm-diameter) were attached with silver
paint onto the sample to act as electrodes with a span of 6 mm. The specimens were tested in axial tension using a
cross-head displacement of 0.33 mm min−1 and R was in situ registered during strain tension testing. The strain

4
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 3. Electrical characterization of MWCNT/PMMA composites. (a) Composites with rectangular geometry (low resistivity),
(b) Composites with wafer geometry (high resistivity). Dimensions whose units are not indicated are in mm.

Figure 4. Piezoresistive characterization of MWCNT/PMMA composites. (a) Geometry and dimensions of the specimens,
dimensions in mm; (b) experimental setup.

(ε) was determined from the machine cross-head displacement, normalized over a gauge length of 13 mm. The
gage factor (k) was then calculated as [38],
DR R 0
k= (5)
e
where R0 is the initial resistance of the specimen without deformation (at ε=0), and ΔR=R−R0. Such
electromechanical sensitivity was extracted from the slope of the linear response of ΔR/R0 with the applied strain
ε in the region 0„ε„0.01. A total of nine samples were tested for each case. Figure 4 illustrates the coupon
(figure 4(a)) and experimental setup (figure 4(b)).

3. Results

3.1. Fragmentation of carbon nanotubes


Figure 5(a) shows representative SEM images of CNT and FCNTs, where the reduction of the length for the
FCNTs can be easily appreciated. Several individual CNTs (left) have lengths ranging 1–3 μm while for the case
of FCNTs (right) most of them look closer or below 1 μm. Figure 5(b) shows the statistical distribution of CNTs
and FCNTs, where the Log-Normal distribution fitting is shown for both cases. It can be clearly evidenced that
the distribution of CNTs is wider and with higher lengths than that of the FCNTs; 86% of the measured FCNTs
have a length below 1 μm, while almost 90% of the counts for the CNTs ranged from 0.8 to 5 μm. The mean,
median and mode length values are also presented within figure 5(b), showing a reduction of length of ∼50%
when the fragmentation process is applied and thus confirming the success of the fragmentation process.
The fragmentation of MWCNTs occurred due to the high-density energy applied to the MWCNT/liquid
system by the ultrasonic probe. The ultrasonic energy promotes bubble nucleation and, when a critical size is
reached, cavitation occurs [28, 29]. During cavitation, the implosion of bubbles due to the pressure changes
generates shock waves causing MWCNT scissoring. According to Huang and Terentjev [39], energy densities

5
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 5. Measurement of MWCNT length. (a) Representative SEM images before (CNTs) and after (FCNTs) the fragmentation
process, (b) Log-Normal fitting of the length distribution measurements for CNTs and FCNTs.

Figure 6. Optical images of CNT/PMMA and FCNT/PMMA composites at different MWCNT concentrations.

above 108 J m−3 can break the MWCNTs as they are in the order of the carbon bonding energies. In this work,
the energy density applied to the system was ∼92 GJ m−3, being this value consistent with the values causing
breakage of MWCNTs. In similar reports, some authors have also shown scissoring and shortening of MWCNTs
when ultrasonic energy was applied [30, 40–42].

3.2. Effect of MWCNT fragmentation on the properties of nanocomposites


3.2.1. Macroscale visualization
Because of the MWCNT length reduction, the morphology of the conductive network changed, which was
visible via optical microscopy. Macroscale optical images were taken from the solid nanocomposites at different
concentrations and the results are shown in figure 6. Pure PMMA (0 wt%) is shown on the left side as a reference
to depict the transparent material. As MWCNT concentration increases, less transparent (darker) composites
are observed for both CNTs and FCNTs. However, the transition is more abrupt for the CNT/PMMA
composites (upper row) which reach a similar appearance for concentrations at 0.1 wt% or above, while the
FCNT/PMMA composites (bottom row) exhibited higher transparency at same MWCNT concentration and
more gradual changes of transparency. Thus, the fragmented MWCNTs (FCNTs) yield higher transparency on
the composites, which can be rationalized in terms of the smaller size of the aggregates and, as such, more space
filled by the matrix.

3.2.2. Electrical conductivity


Figure 7 shows the electrical conductivity of the composites prepared with CNTs and FCNTs. The electrical
conductivity of the pure polymer (0 wt%) was measured in the order of 10−13 S m−1 (black square) and was
added only as a reference. As a general outcome, electrical conductivity of composites with FCNTs (shorter
MWCNTs) is lower than the electrical conductivity of CNTs (longer MWCNTs) for all the investigated
concentrations. For the CNT/PMMA composites, samples of 0.15 wt% and below are still not electrically
percolated as can be noticed from the conductivity values in the order of 10−12–10−11 S m−1. However, an
abrupt increase in the electrical conductivity was measured around 0.2 wt% and then small increments were
observed for higher concentrations, reaching a plateau value close to 10−1 S/m. A similar trend was observed for
the FCNT/PMMA composites, but the sudden increase of electrical conductivity was observed at a higher
concentration, around 0.3 wt%; lower concentrations were still non electrically percolated. The electrical

6
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 7. Electrical conductivity as a function of MWCNT weight concentration for composites with as-received (CNT) and
fragmented (FCNTs). Inset shows the fitting for the percolation threshold estimation.

conductivity of FCNT/PMMA continued increasing until reaching a plateau value close to 10−3 S/m. Thus,
shorter MWCNTs render lower electrical conductivity to the composites and the percolation threshold is at a
higher concentration. The decreasing of the electrical conductivity for composites with FCNTs can be associated
mainly to two causes; shorter MWCNTs make the percolation network less efficient in terms of interparticle
contacts and tunneling distances (therefore increasing the resistance), and the damage in the MWCNTs due to
the scissoring process may affect negatively their electronic configuration and properties [30].
Using the collected data and equation (4), the percolation thresholds fc were estimated as 0.18 wt% and
0.24 wt% for the CNT/PMMA and FCNT/PMMA system, respectively, see inset in figure 7. On the other hand,
values of critical exponent t were estimated as t=4.6 and t=3.8 for CNT and FCNTs, respectively. According
to classical percolation theory, small values of t (close to 1) are associated to two-dimensional conductive
networks while value slightly above of 2 are associated with three-dimensional networks [19, 43, 44]. However, it
must be considered that for polymer nanocomposites, particles are not identical, and dispersion is not perfect,
due to the manufacturing process, being these deviations from the assumptions of the classical percolation
theory. Instead, for these kind of materials high values of t (above t=3) have been reported [30, 45], being
suggest that such high values are correlated to the existence of tunneling barriers and/or the presence of high
aspect ratio particles (such as the MWCNTs) [45]. In a similar material system, Avilés et al [30] found that the
percolation threshold for a MWCNT/PMMA system with shortened MWCNTs was 0.185 wt% while for a
system with longer MWCNTs was estimated as 0.098 wt%. This trend is similar with the trend here observed
regarding the effect of the MWCNT length on such a threshold. In other study by Inam et al [41], authors report
that the electrical percolation threshold of a MWCNT/epoxy system with MWCNTs with an average length of
∼1.6 μm is 0.07 vol%. However, when the system was conformed with MWCNTs with an average length of
∼0.7 μm the percolation threshold was 0.44 vol%, evidencing an increased percolation threshold when shorter
MWCNTs are used, as here observed. Authors state that ‘short’ MWCNTs form less efficient entangled
percolating networks compared with ‘long’ MWCNTs [41].

3.2.3. Piezoresistive properties


To investigate the effect of the MWCNT length on the electromechanical response, piezoresistive
characterization of composites with CNTs and FCNTs was carried out. The concentrations evaluated ranged
from 0.4 wt% to 0.7 wt%; composites with lower concentrations, specially those with FCNTs, exhibited overly
noisy electrical signals (due their high resistance) rendering them unreliable for proper data analysis. Figure 8(a)
shows the measured values of the initial resistance (R0), i.e. resistance at ε=0. R0 values tend to decrease with
increased concentration and, as expected, FCNT/PMMA composites exhibited higher R0 (few MΩ) than CNT/
PMMA composites (tens to hundreds kΩ). This is consistent with the electrical behavior shown in figure 7 and
shows good reproducibility. As an example, figure 8(b) shows representative piezoresistive plots for composites
at 0.5 wt%, where the normalized change of resistance is plotted as function of the axial strain. The gage factors k
(slope) are shown next to each plot and is clear that the sensitivity is higher for the FCNT/PMMA composite
(k=12.1) than for the CNT/PMMA composite (k=7.8). Figure 8(c) summarizes the results of the

7
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 8. Piezoresistive characterization of nanocomposites with CNT and FCNTs. (a) Initial (unloaded) resistance of
nanocomposites as a function of the MWCNT concentration, (b) representative piezoresistive plots of MWCNT/PMMA composites
at 0.5 wt%, (c) average gage factors k of the evaluated composites.

piezoresistive characterization showing the average gage factors for the evaluated concentrations. Excepting for
the highest concentration of 0.7 wt%, the electromechanical sensitivity of the FCNT/PMMA composites is
consistently higher than that of the CNT/PMMA composites. Thus, having shorter MWCNTs in the conductive
network renders a higher sensitivity to the material; in other words, the changes in the electrical resistance
caused by induced deformation are more notorious when the percolated network comprises nanoparticles with
lower aspect ratio (length/diameter). Pham et al [21] reported k values ranging from 1.4 to 15.3 for MWCNT/
PMMA composites fabricated by two methods, solution based and dry blended. Authors state that k was higher
when the composites concentrations were closer to the percolation threshold of the system. The gage factors
reported here agree with the range found in [21] for a similar material system. Avilés et al [30] reported k values
ranging from 3 to 10 for a MWCNT/PMMA system, with higher values for composites with shorter MWCNTs
and closer to the percolation threshold, similar to the values observed in this study. For comparison, the gage
factors exhibited for commercial metal gages are in the order of k∼2 [38].
The effect of MWCNT length on the piezoresistive response has been scarcely reported. A few authors who
addressed this phenomenon did not carried out statistical measurements of the MWCNT lengths but used
commercial MWCNTs with different lengths instead [46–48]. Some reports investigated MWCNT/rubber
composites under compression [46, 48], being this a different scenario than the work presented here. On these
reports, higher gage factors were found for longer MWCNTs, in an opposite trend than the one here presented.
Nevertheless, a similar finding is found in the fact that the initial resistance of the composite (initial
configuration of the network) is correlated with the sensitivity of the composite, i.e. higher initial resistance is
associated to higher gage factors. The higher sensitivity for composites with fragmented (shorter) MWCNTs can
be understood in terms of the less CNT-to-CNT contact points and the existence of a loosely packed conductive
network originated by the short MWCNTs. This causes that the mechanical changes on the system, i.e. applied
strains, generate large changes on the effective resistance rendering high gage factors [3, 30].

8
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 9. Optical images of CNT/PMMA and FCNT/PMMA composites fabricated without (left) and with (right) application of
electric field.

3.3. Effect of MWCNT alignment on the properties of nanocomposites


To explore the effect of orienting the MWCNT conductive network on the piezoresistive response of composites
with CNT and FCNTs, an alternating current electric field was applied during composite fabrication. The
application of electric fields has been reported as a convenient strategy to manipulate the carbon nanotubes
within polymers for composite fabrication, due to the ease of implementation and control [49, 50]. Due to the
very different dielectric properties of conductive particles and surrounding medium (liquid state polymer
matrix) under the electric field application, the free charges at the particle/medium interphase are rearranged
and dipole moments are induced on the particles. Such induced dipoles interact with the applied electric field
and with the neighboring particles promoting then the chaining and orientation of the network in the direction
of the applied field [3, 15, 51, 52]. Figure 9 shows macroscale images of nanocomposites at 0.3 wt% and 0.7 wt%
fabricated with CNTs and FCNTs without (left) and with (right) the application of electric field. For the lower
concentration, a homogeneous dispersion with presence of agglomerated MWCNTs within the polymer is
observed when no electric field is applied, being the network sparser (more transparent) for the FCNT/PMMA
composites. On the other hand, the formation of vertical structures is clearly evidenced when the electric field is
applied at such a low concentration (0.3 wt%). For the CNT/PMMA composites the oriented structures are
finer and denser packed, while thicker MWCNT bundles are visualized for FCNT/PMMA composites. For the
composites at 0.7 wt%, a denser packed (much darker) system is evidenced for both type of MWCNTs, but thick
and vertically oriented agglomerates/bundles are still evidenced when the electric field was applied. Due to the
high concentration, no significant differences between samples with CNTs and FCNT are appreciated at this
length scale.
To further examine the property-structure relationship in the composites, figure 10(a) shows the initial
(unloaded) electrical resistances (R0) for CNT/PMMA and FCNT/PMMA composites at 0.3 wt% and 0.7 wt%,
with application of electric field (EF). Regardless of the application of the electric field, R0 for 0.3 wt% is higher
than that for 0.7 wt%, as expected from the given concentrations. The registered R0 values for 0.3 wt% aligned
composites, with CNTs and FCNTs, are of few MΩ, making thus feasible the piezoresistive characterization for
such a low concentration (which was not possible for non-aligned composites due to the very high R0). The

9
Mater. Res. Express 6 (2019) 115024 J Cob et al

Figure 10. Effect of MWCNT length and concentration on the electrical response of aligned composites. (a) Initial (unloaded)
electrical resistance of aligned nanocomposites with CNTs and FCNTs, (b) average gage factors of the aligned composites.

preferential orientation of the MWCNT bundles within the polymer matrix brings improved electrical
properties in such a direction, due to the more efficient electron traveling through the aligned structure
[3, 15, 53]. It is also observed that aligned FCNT/PMMA composites exhibited a higher average resistance
(∼11 MΩ) than the aligned CNT/PMMA composites (∼2 MΩ), evidencing the effect of the MWCNT length on
such a property.
Figure 10(b) presents the average gage factors of the evaluated aligned composites. In the two cases, CNT and
FCNTs, no clear effect of the MWCNT concentration on the gage factor was appreciated when electric field was
applied, as can be noticed from the overlapped error bars. For the case of aligned CNT/PMMA composites (left),
the average gage factor of the sample at 0.3 wt% (k=5.0) was very similar to that of the aligned CNT/PMMA at
0.7 wt% (k=4.7). As previously mentioned, R0 for the non-aligned CNT/PMMA at 0.3 wt% was very high
(tens of MΩ) making the piezoresistive characterization non feasible for such composites and no k value was
registered. It is thus evidenced that by inducing alignment on composites at low concentrations, an
electromechanical response can be registered with acceptable values of sensitivity. The average k value for the
non-aligned CNT/PMMA composites at 0.7 wt% was k=4.6 (see figure 8(c)) a very similar value than their
aligned counterpart (k=4.7, figure 10(b)). Thus, at this high concentration of non-fragmented MWCNTs no
clear effect of alignment was observed, being this explained in terms of the dense and well-packed conductive
network. Furthermore, when aligned FCNTs conform the conductive network (right in figure 10(b)), a
notorious increasing on the gage factor is observed. Average values of k=9.8 and k=10.1 were measured for
FCNT/PMMA composites at 0.3 wt% and 0.7 wt%, respectively. No k value was registered for non-aligned
FCNT/PMMA composites due to their very high R0 (hundreds of MΩ); by aligning the fragmented MWCNTs at
this low concentration a very important enhancement on the electrical response is clearly evidenced. Higher
sensitivity was also observed for the aligned FCNT/PMMA composites at 0.7 wt%, k=10.1, see figure 10(b),
compared with k=4.7 measured for their non-aligned counterparts, see figure 8(c).
Some authors have reported increased electromechanical sensitivity when the fillers are preferentially
oriented in a specific direction [15, 54–56]. This behavior may be attributed to a better strain transfer and
increased contact resistance along the aligned direction. It has been shown by several authors that by aligning
carbon nanotubes within a polymer matrix anisotropy on the effective properties is induced. Regarding the
composites piezoresistivity, composites with aligned structures render enhanced properties than composites
with structures oriented randomly [3, 15, 54, 55, 57]. Some hypothesis state that the change on the CNT-to-CNT
lateral contacts and the rupture of few misaligned paths are more significant phenomena when aligned
structures conform the composite, being these major causes for the improved sensitivity [3]. Also, the MWCNTs
length affects their rotational and chaining motion under electric fields as some studies have reported [58–60].
Here, the influence of the electric field on the piezoresistive sensitivity was more evident on fragmented (shorter)
MWCNTs, as observed from the high values in figure 10(b) (right). Thus, the combined effect of fragmenting
and aligning the MWCNTs renders higher sensitivity to the nanocomposites, being this a major outcome of this
work. More systematical studies are still needed to be addressed to better understand this combined effect,
towards the control and fabrication of tailored materials with controlled properties.

10
Mater. Res. Express 6 (2019) 115024 J Cob et al

4. Conclusions

MWCNT/PMMA composites were fabricated by solution casting and the effect of the MWCNT concentration,
length and alignment on the electrical and piezoresistive properties of the composites were investigated.
MWCNTs were systematically fragmented by high-density ultrasonic energy rendering a length reduction of
∼50%, assessed by a formal statistical analysis of ∼1500 individual measurements. Sparse MWCNT networks
with more transparent solid composites were observed at the macroscale when FCNTs were incorporated to the
PMMA matrix, which greatly impacts the electro-mechanical properties of the composites. Lower electrical
conductivity and higher percolation threshold were measured for composites with fragmented MWCNTs, being
this associated to the less packed conductive network conformed by fragmented (shorter) MWCNTs. On the
other hand, a significantly higher gage factor was observed for composites with fragmented MWCNTs,
evidencing that the sparser network formed by shorter MWCNTs bring more sensitivity to the composites.
Finally, by combining the MWCNT alignment (induced by application of an electric fields) with the
fragmentation of MWCNTs, high gage factors can be obtained for low concentration composites. The combined
effect of MWCNT concentration, length and alignment on the electromechanical properties of the polymer
nanocomposites represents a main outcome of this study and an important step towards a better understanding
of these multifunctional materials.

Acknowledgments

A I Oliva-Avilés acknowledges the support of the ‘Fondo Sectorial de Investigación para la Educación’ through
the SEP-CONACYT grant No. 235905. Authors acknowledge the technical assistance of Emilio Corona with
tensile tests and LANNBIO for the SEM measurements (CINVESTAV-Mérida). The assistance of Alejandro
May (CICY) for the electrical measurements of high-resistance samples is also strongly appreciated.

ORCID iDs

A I Oliva-Avilés https://orcid.org/0000-0002-6491-2995
F Avilés https://orcid.org/0000-0002-0227-1758
A I Oliva https://orcid.org/0000-0003-1854-7074

References
[1] Chen J, Yan L, Song W and Xu D 2018 Thermal and electrical properties of carbon nanotube-based epoxy composite materials Mater.
Res. Express 5 065051
[2] Li Z, Wang L, Li Y, Feng Y and Feng W 2019 Carbon-based functional nanomaterials: preparation, properties and applications Compos.
Sci. Technol. 179 10–40
[3] Avilés F, Oliva-Avilés A I and Cen-Puc M 2018 Piezoresistivity, strain, and damage self-sensing of polymer composites filled with
carbon nanostructures Adv. Eng. Mater. 20 1701159
[4] Li M, Wang J, Wang S, Zuo T, Sun W, Gu Y and Zhang Z 2019 Effect of microstructure on the piezoresistive behavior of carbon
nanotube composite film Mater. Res. Express 6 025034
[5] Chatterjee M J, Banerjee D and Chatterjee K 2016 Composite of single walled carbon nanotube and sulfosalicylic acid doped
polyaniline: a thermoelectric material Mater. Res. Express 3 085009
[6] Fiorillo A S, Critello C D and Pullano S A 2018 Theory, technology and applications of piezoresistive sensors: a review Sens. Actuators A
Phys. 281 156–75
[7] Ponnamma D, Guo Q, Krupa I, Al-Maadeed M, Varughese K T, Thomas S and Sadasivuni K K 2015 Graphene and graphitic derivative
filled polymer composites as potential sensors Phys. Chem. Chem. Phys. 17 3954–81
[8] Amjadi M, Kyung K-U, Park I and Sitti M 2016 Stretchable, skin-mountable, and wearable strain sensors and their potential
applications: a review Adv. Funct. Mater. 26 1678–98
[9] Mathur R B, Pande S and Singh B P 2010 Properties of PMMA/Carbon nanotubes nanocomposites Polymer Nanotube Nanocomposites:
Synthesis, Properties, and Applications ed V Mittal (Salem USA: Scrivener Publishing LLC)
[10] Ormsby R, McNally T, Mitchell C, Halley P, Martin D, Nicholson T and Dunne N 2011 Effect of MWCNT addition on the thermal and
rheological properties of polymethyl methacrylate bone cement Carbon 49 2893–904
[11] Nawar A M and El-Mahalawy A M 2019 Simple processed semi-transparent Schottky diode based on PMMA-MWCNTs
nanocomposite for new generation of optoelectronics Synth. Met. 255 116102
[12] Coleman J N, Khan U, Blau W J and Gun’ko Y K 2006 Small but strong: a review of the mechanical properties of carbon nanotube–
polymer composites Carbon 44 1624–52
[13] Breuer O and Sundararaj U 2004 Big returns from small fibers: a review of polymer/carbon nanotube composites Polym. Composites 25
630–45
[14] Mathur R B, Pande S, Singh B P and Dhami T L 2008 Electrical and mechanical properties of multi‐walled carbon nanotubes reinforced
PMMA and PS composites Polym. Composites 29 717–27
[15] Oliva-Avilés A I, Avilés F and Sosa V 2011 Electrical and piezoresistive properties of multi-walled carbon nanotube/polymer composite
films aligned by an electric field Carbon 49 2989–97

11
Mater. Res. Express 6 (2019) 115024 J Cob et al

[16] Bautista-Quijano J R, Avilés F, Aguilar J O and Tapia A 2010 Strain sensing capabilities of a piezoresistive MWCNT-polysulfone film
Sens. Actuators A 159 135–40
[17] Cen-Puc M, Pool G, Oliva-Avilés A I, May-Pat A and Avilés F 2017 Experimental investigation of the thermoresistive response of
multiwall carbon nanotube/polysulfone composites under heating-cooling cycles Compos. Sci. Technol. 151 34–43
[18] Avilés F, May-Pat A, Canché-Escamilla G, Rodríguez-Uicab O, Ku-Herrera J J, Duarte-Aranda S, Uribe-Calderón J, González-Chi P I,
Arronche L and La Saponara V 2016 Influence of carbon nanotube on the piezoresistive behavior of multiwall carbon nanotube/
polymer composites J. Intell. Mat. Syst. Struct. 27 92–103
[19] Bauhofer W and Kovacs J Z 2009 A review and analysis of electrical percolation in carbon nanotube polymer composites Compos. Sci.
Technol. 69 1486–98
[20] Hu N, Masuda Z, Yan C, Yamamoto G, Fukunaga H and Hashida T 2008 The electrical properties of polymer nanocomposites with
carbon nanotube fillers Nanotechnology 19 215701
[21] Pham G T, Park Y-B, Liang Z, Zhang C and Wang B 2008 Processing and modeling of conductive thermoplastic/carbon nanotube
films for strain sensing Compos Part B Eng. 39 209–16
[22] Zetina-Hernández O, Duarte-Aranda S, May-Pat A, Canché-Escamilla G, Uribe-Calderon J, Gonzalez-Chi P I and Avilés F 2013
Coupled electro-mechanical properties of multiwall carbon nanotube/polypropylene composites for strain sensing applications
J. Mater. Sci. 48 7587–93
[23] Celzard A, McRae E, Deleuze C, Dufort M, Furdin G and Mareché J F 1996 Critical concentration in percolating systems containing a
high-aspect-ratio filler Phys. Rev. B 53 6209
[24] Sun Y, Bao H-D, Guo Z-X and Yu J 2009 Modeling of the electrical percolation of mixed carbon fillers in polymer-based composites
Macromolecules 42 459–63
[25] Meeuw H, Viets C, Liebig W V, Schulte K and Fiedler B 2016 Morphological influence of carbon nanofillers on the piezoresistive
response of carbon nanoparticle/epoxy composites under mechanical load Eur. Polym. J. 85 198–210
[26] Mecklenburg M, Mizushima D, Ohtake N, Bauhofer W, Fiedler B and Schulte V 2015 On the manufacturing and electrical and
mechanical properties of ultra-high wt% fraction aligned MWCNT and randomly oriented CNT epoxy composites Carbon 91 275–90
[27] Nanni F, Mayoral B L, Madau F, Montesperelli G and McNally T 2012 Effect of MWCNT alignment on mechanical and self-
monitoring properties of extruded PET–MWCNT nanocomposites Compos. Sci. Technol. 72 1140–6
[28] Lucas A, Zakri C, Maugey M, Pasquali M, Schoot P V D and Poulin P 2009 Kinetics of nanotube and microfiber scission under
sonication J. Phys. Chem. C 113 20599–605
[29] Hennrich F, Krupke R, Arnold K, Rojas Stütz J A, Lebedkin S, Koch T, Schimmel T and Kappes M M 2007 The mechanism of
cavitation-induced scission of singlewalled carbon nanotubes J. Phys. Chem. B 111 1932–7
[30] Avilés F, Oliva A I, Ventura G, May-Pat A and Oliva-Avilés A I 2019 Effect of carbon nanotube length on the piezoresistive response of
poly(methyl methacrylate) nanocomposites Eur. Polym. J. 110 394–402
[31] Riddick J A, Bunger W B and Sakano T K 1986 Organic Solvents: Physical Properties and Methods of Purification (New York: Wiley-
Interscience)
[32] https://imagej.net/Welcome Accessed 2019
[33] Balakrishnan N and Chen W S 1999 Handbook of Tables for Order Statistics from Lognormal Distribution With Applications (New York:
Springer)
[34] Cen-Puc M, Oliva-Avilés A I and Avilés F 2018 Thermoresistive mechanisms of carbon nanotube/polymer composites Physica E 95
41–50
[35] ASTM D445 2006 Standard Test Method for Kinematic Viscosity of Transparent and Opaque Liquids (and Calculation of Dynamic
Viscosity) (West Conshohocken, PA, USA: ASTM International))
[36] ASTM D638 2004 Standard Test Method for Tensile Properties of Plastics (West Conshohocken, PA, USA: ASTM International)
[37] Huerta E, Corona J E, Oliva A I, Avilés F and González J 2010 Universal testing machine for mechanical properties of thin materials Rev.
Mex. Fis. 56 317–22 (http://www.ejournal.unam.mx/rmf/no564/RMF005600408.pdf)
[38] Kobayashi A S 1993 Handbook on Experimental Mechanics (New York: VCH Publishers)
[39] Huang Y Y and Terentjev E M 2012 Dispersion of carbon nanotubes: mixing, sonication, stabilization, and composite properties
Polymers 4 275–95
[40] Inam F, Vo T, Jones J P and Lee X 2012 Effect of carbon nanotube lengths on the mechanical properties of epoxy resin: an experimental
study J. Compos. Mater. 47 2321–30
[41] Inam F, Reece M J and Peijs T 2011 Shortened carbon nanotubes and their influence on the electrical properties of polymer
nanocomposites J. Compos. Mater. 46 1313–22
[42] Chen S J, Zou B, Collins F, Zhao X L, Majumber M and Duan W H 2014 Predicting the influence of ultrasonication energy on the
reinforcing efficiency of carbon nanotubes Carbon 77 1–10
[43] Kilbride B E, Coleman J N, Fraysse J, Fournet P, Cadek M, Drury A, Hutzler S, Roth S and Blau W J 2002 Experimental observation of
scaling laws for alternating current and direct current conductivity in polymer-carbon nanotube composite thin films J. Appl. Phys. 9
4024–30
[44] Stauffer D and Aharony A 2003 Introduction to percolation theory 2nd Revised Edn (London: Taylor and Francis)
[45] Chen G, Weng W, Wu D and Wu C 2003 PMMA/graphite nanosheets composite and its conducting properties Eur. Polym. J. 39
2329–35
[46] Dang Z-M, Jiang M-J, Xie D, Yao S-H, Zhang L-Q and Bai J 2008 Supersensitive linear piezoresistive property in carbon nanotubes/
silicone rubber nanocomposites J. Appl. Phys. 104 024114
[47] Zha J-W, Shehzad K, Li W-K and Dang Z-M 2013 The effect of aspect ratio on the piezoresistive behavior of the multiwalled carbon
nanotubes/thermoplastic elastomer nanocomposites J. Appl. Phys. 113 014102
[48] Jiang M-J, Dang Z-M, Xu H-P, Yao S-H and Bai J 2007 Effect of aspect ratio of multiwall carbon nanotubes on resistance-pressure
sensitivity of rubber nanocomposites Appl. Phys. Lett. 91 072907
[49] Goh P S, Ismail A F and Ng B C 2014 Directional alignment of carbon nanotubes in polymer matrices: Contemporary approaches and
future advances Composites Part A 56 103–26
[50] Liu L, Ma W and Zhang Z 2011 Macroscopic carbon nanotube assemblies: preparation, properties, and potential applications Small 7
1504–20
[51] Monti M, Natali M, Torre L and Kenny J M 2012 The alignment of single walled carbon nanotubes in an epoxy resin by applying a DC
electric field Carbon 50 2453–64
[52] Zhu Y F, Ma C, Zhang W, Zhang R P, Koratkar N and Liang J 2009 Alignment of multiwalled carbon nanotubes in bulk epoxy
composites via electric field J. Appl. Phys. 105 054319

12
Mater. Res. Express 6 (2019) 115024 J Cob et al

[53] Theodosiou T C and Saravanos D A 2010 Numerical investigation of mechanisms affecting the piezoresistive properties of CNT doped
polymers using multi-scale models Compos. Sci. Technol. 70 1312–20
[54] Obitayo W and Liu T 2015 Effect of orientation on the piezoresistivity of mechanically drawn single walled carbon nanotube (SWCNT)
thin films Carbon 85 372–82
[55] Li S, Park J G, Wang S, Liang R, Zhang C and Wang B 2014 Working mechanisms of strain sensors utilizing aligned carbon nanotube
network and aerosol jet printed electrodes Carbon 73 303–9
[56] Li A, Bogdanovich A E and Bradford P D 2015 Aligned carbon nanotube sheet piezoresistive strain sensors Smart Mater. Struct. 24
095004
[57] Parmar K, Mahmoodi M, Park C and Park S S 2013 Effect of CNT alignment on the strain sensing capability of carbon nanotube
composites Smart Mater. Struct. 22 075006
[58] Oliva-Avilés A I, Avilés F, Sosa V and Seidel G D 2014 Dielectrophoretic modeling of the dynamic carbon nanotube network formation
in viscous media under alternating current electric fields Carbon 69 342–54
[59] Castellano R J, Akin C, Giraldo G, Kim S, Fornasiero F and Shan J W 2015 Electrokinetics of scalable, electric-field-assisted fabrication
of vertically aligned carbon-nanotube/polymer composites J. Appl. Phys. 117 214306
[60] Oliva-Avilés A I, Zozulya V V, Gamboa F and Avilés F 2016 Dynamic evolution of interacting carbon nanotubes suspended in a fluid
using a dielectrophoretic framework Physica. E 83 7–21

13

You might also like