Sei sulla pagina 1di 14

Applied Nanoscience

https://doi.org/10.1007/s13204-019-01032-2

ORIGINAL ARTICLE

Functional nanocomposites for 3D printing of stretchable


and wearable sensors
Mohammad Abshirini1 · Mohammad Charara1 · Parisa Marashizadeh1 · Mrinal C. Saha2 · M. Cengiz Altan3 ·
Yingtao Liu4 

Received: 12 February 2019 / Accepted: 9 April 2019


© King Abdulaziz City for Science and Technology 2019

Abstract
This paper presents highly flexible strain sensors fabricated by extrusion-based 3D printing of electrically conductive nano-
composites consisting of multi-walled carbon nanotube (MWNT)/polydimethylsiloxane (PDMS). The effects of printing
parameters and nanocomposite formulation on the piezoresistive behavior of the 3D-printed sensors are investigated. Experi-
mental results demonstrate that the 3D printing-induced alignment of MWNTs results in the enhancement of piezoresistive
sensing function of the nanocomposites. Detailed analyses are performed using the optimized sensors to characterize their
sensing performance, including load rate dependency, repeatability under long-term cyclic loads, and relaxation behavior.
The 3D-printed strain sensors demonstrate high flexibility, stretching to 146% strain before fracture, and exhibit a linear
piezoresistive response up to 70% strain with a gauge factor of 12.15. The distribution of nanotubes in the polymer and
the piezoresistive mechanism of the material are explored by in situ micro-mechanical testing under a scanning electron
microscope (SEM). The developed sensors are attached on gloves to monitor the motion of a human hand, demonstrating
their application wearable electronics.

Keywords  Piezoresistivity · Stretchable sensors · 3D printing · Carbon nanotube · Conductive nanocomposites

Introduction et al. 2016; Martinez et al. 2013), and wearable devices for
human health monitoring (Amjadi et al. 2014; Choi et al.
Recent scientific progress in the use of electronic devices 2017; Trung et al. 2017). These applications require flexible
for strain sensing applications has attracted significant strain sensors with high sensitivity and durable performance.
attention in the fields of soft robotics (Fallahi et al. 2017; In general, strain sensing range and sensitivity are two key
Zheng et al. 2015), electronic skins (Ho et al. 2016; Hong parameters used to evaluate the quality of these sensors. Par-
ticularly, the sensitivity is quantified in the form of gauge
factor (GF), and is defined as the ratio of relative electrical
Electronic supplementary material  The online version of this
article (https​://doi.org/10.1007/s1320​4-019-01032​-2) contains resistance change to applied strain. Conventional metallic
supplementary material, which is available to authorized users. strain gauges offer a low GF (around 2) and a limited sens-
ing range (< 5% strain), while semiconductor-based sensors
* Yingtao Liu offer high GFs but a limited detection range, making neither
yingtao@ou.edu
of them suitable for wearable sensing applications (Zhang
1
School of Aerospace and Mechanical Engineering, et al. 2017).
University of Oklahoma, 865 Asp Ave. #301C, Norman, Elastomeric nanocomposites enhanced by conductive
OK 73019, USA nanofillers are one of the most popular advanced materials
2
School of Aerospace and Mechanical Engineering, University for stretchable strain sensing applications. Polydimethylsi-
of Oklahoma, 865 Asp Ave. #208, Norman, OK 73019, USA loxane (PDMS) (Chowdhury et al. 2018; Lu et al. 2012),
3
School of Aerospace and Mechanical Engineering, University polyurethane (Zhang et al. 2013), and natural rubber (Selvan
of Oklahoma, 865 Asp Ave. #205, Norman, OK 73019, USA et al. 2016) have been employed as elastomeric polymers
4
School of Aerospace and Mechanical Engineering, University in these nanocomposites. The addition of conductive nano-
of Oklahoma, 865 Asp Ave. #218A, Norman, OK 73019, fillers to these flexible matrices endows the base material
USA

13
Vol.:(0123456789)
Applied Nanoscience

with electrical properties which can be exploited to meas- jet printing (Agarwala et al. 2018; Goh et al. 2018a), and direct
ure its deformation. Both capacitance-based and resistance- writing (Kim et al. 2012; Vatani et al. 2013). Recently, extru-
based strain sensing mechanisms have been reported in sion-based 3D printing technology has been reported as a rap-
literature, though the electrical resistance-based method idly expanding approach to fabricate customizable structures
has been more commonly reported (Abshirini et al. 2018; and prototyping of elastomers, such as PDMS, and related
Cao et al. 2015; Luo et al. 2018). The change in electrical elastomeric nanocomposites (Charara et al. 2019a; Guo et al.
resistance is believed to be induced by the reorganization of 2017; Hwang et al. 2018; Truby et al. 2018). Due to its thixo-
the electrically conductive network formed by nanofillers tropic flow properties, PDMS can be 3D printed under ambi-
under mechanical stimulus (i.e., tension or compression). ent conditions and used to create relatively thin multi-material
Researchers have explored a variety of nanofillers, including devices, such as a bionic ear (Mannoor et al. 2013). Although
carbon black (Zheng et al. 2017), carbon nanofiber (Charara the printed PDMS is able to maintain its geometric fidelity
et al. 2019b; Chowdhury et al. 2016), graphene (Park et al. after curing, the low elastic modulus of the pre- and post-cured
2015), and carbon nanotubes (CNTs) (Ryu et al. 2015; Liu PDMS and its deformation under gravity prior to curing still
et al. 2012), to increase the conductivity of elastomers. Of restricts the 3D geometries that can be fabricated due to its
these, multi-walled carbon nanotubes (MWNT) have been inability to maintain shape fidelity directly post deposition.
widely used to improve these polymers’ conductivity due Hinton et al. reported a 3D printing method for PDMS in a
to their excellent electrical characteristics at low concentra- hydrophilic support bath, aiming to enable true freeform fab-
tion (Hu et al. 2008; Kang et al. 2006; Lipomi et al. 2011). rication of complex structures (Hinton et al. 2016). Dispersion
Piezoresistive response in nanocomposite materials has been of nanoparticles in PDMS can significantly improve materi-
previously observed (Obitayo and Liu 2012). By measuring als’ elastic modulus and viscosity, leading to 3D printable and
the electrical resistance change resulting from mechanical highly elastic nanocomposites. Carbon-based nanoparticles,
deformation in these materials, the applied strains can be including graphene and CNTs, have been dispersed in elasto-
calculated. mers and 3D printed for electronic and biomedical applications
Numerous studies have been performed in recent years (Jakus et al. 2015; Wang et al. 2017). Most previous work
to produce a sensor with the capability of detecting strains has focused on nanocomposite characteristics, 3D printability,
up to 100%. However, in most of the cases, these sensors and potential applications. However, there is a lack of studies
exhibit a non-linear sensing response over a wide strain exploring the effects of 3D printing and curing parameters
range (Liu et al. 2015; Yamada et al. 2011) and a low GF at on the nanocomposite’s properties and sensing response due
large strains (Cai et al. 2013; Cohen et al. 2012; Liu et al. to the various nano- and micro-structures formed within the
2015; Yamada et al. 2011). Recently, several approaches material during the manufacturing process (Goh et al. 2018b).
have been developed to fabricate sensors with high sensi- In this study, an electrically conductive nanocomposite is
tivity and stretchability, such as micro/nanocracking-assisted 3D printed on a pristine elastomer substrate to fabricate highly
conductive network resulting in sensors with high GFs but stretchable and wearable strain sensors. To obtain optimal
very low sensing range (Chen et al. 2016; Hoang et al. 2016; piezoresistive sensing function, nanocomposites are printed
Kang et al. 2014) and the creation of porous nanocomposite varying processing parameters, including needle diameter,
materials (Li et al. 2017; Liu et al. 2016; Wu et al. 2017). curing temperature, and MWNT concentrations. All the
However, the fabrication process to create such sensors is printed materials are tested under cyclic tensile load to iden-
often highly complex and time consuming. Therefore, it is tify their sensitivity and linear sensing response range. The
important to explore a novel strain sensor fabrication tech- optimized sensors are further tested to validate their reliability
nique which allows for the facile manufacturing of highly and repeatability under various cyclic loads and loading rates.
stretchable, durable sensors with a wide sensing range. The piezoresistive sensing mechanism was demonstrated by
Most reported nanocomposite strain sensors were manufac- comparing the MWNT conductive network using in  situ
tured by molding and casting (Amjadi et al. 2016; Obitayo and micro-mechanical testing in a scanning electron microscope
Liu 2012). Recently, various additive manufacturing methods (SEM). Finally, the fabricated sensors are used as wearable
have been used for fabricating conductive polymers. Fused sensors to detect the motions on a human hand and fingers.
deposition modeling, based on the melting of thermoplastic
filaments to manufacture a desired 3D model, has been widely
used for the printing of conductive thermoplastic polymers Experimental
(Costa et al. 2014; Gnanasekaran et al. 2017; Kumar et al.
2018; Kwok et al. 2017; Leigh et al. 2012; Wei et al. 2015; Materials
Yu et al. 2017). Other printing techniques have been used for
manufacturing conductive polymers, including inkjet printing- Unless otherwise stated, all the materials in this study were
based deposition of ink droplets (Michelis et al. 2015), aerosol used as received. SYLGARD 184 PDMS base elastomer

13
Applied Nanoscience

(part A) and curing agent (part B) were purchased from Dow of various diameters were used to print the nanocomposites
Corning. MWNTs with an average aspect ratio of 100 and on pristine PMDS substrates. The prepared MWNT/PDMS
diameter of 50–90 nm were purchased from Sigma-Aldrich. nanocomposites were printed into an electrically conduc-
Tetrahydrofuran (THF) purchased from Sigma-Aldrich was tive pattern of two connected 30 mm long parallel lines.
used as the solvent. The designed 3D model was converted into G-code using
the Repetier and Slic3r open source software. The extru-
Preparation and 3D printing of nanocomposites sion pressure is adjusted based on the volumetric flow rate
required depending on the filament diameter (in this case the
Pristine PDMS sheets were first manufactured using an alu- syringe diameter) and the nozzle diameter (in this case the
minum mold. The PDMS prepolymer was prepared by mix- needle diameter). For this work, printhead speed of 1 mm/s
ing the base elastomer and curing agent at the 10:1 part A to is used to deposit the nanocomposite. The printed conduc-
part B weight ratio, as recommended by the manufacturer. tive pattern was cured at various temperatures (25 °C, 65 °C,
The mixed prepolymer was degassed in a vacuum desicca- and 130 °C) in an oven to explore the curing temperature
tor at room temperature for 30 min and cured in an oven at effects on the obtained piezoresistive sensing function. A
65 °C for 4 h. The manufactured pristine PDMS sheets were layer of pristine PDMS was coated on the printed nanocom-
used as substrates during the 3D printing process. posites and cured at 130 °C for 20 min before mechanical
The electrically conductive nanocomposites were pre- testing. The schematic of the strain sensor fabrication pro-
pared by dispersing MWNT in the PDMS base elastomer cess is illustrated in Fig. 1b. The modified 3D printer, the
using the solvent-assisted ultrasonication method. MWNT printing process, and the final shape of the fabricated strain
and PDMS base elastomer were added into 30 ml THF and sensor are shown in Fig. 2c. The dimensions of the final sen-
mixed for 5 min using a magnetic stirrer at 350 RPM. After sors are 50 × 30 × 1.8 mm. The PDMS substrate and sealing
mixing, a 750 watt probe sonicator was used to sonicate layer are both 0.9 mm.
the PDMS base elastomer/THF/MWNT solution for 30 min.
The mixture was kept on a hot plate at 65 °C and stirred Piezoresistive characterization
at 100 RPM overnight to evaporate all the THF solvent. of the nanocomposite
Finally, the PDMS curing agent was added to the prepared
nanocomposites and mixed for 3 min before loading into a 3 Three parameters, including MWNT concentration, needle
mL syringe and 3D printing on the pristine PDMS substrate. diameters, and curing temperature were optimized to obtain
The schematic of the material preparation process is shown the highest piezoresistive sensing function of the developed
in Fig. 1a. Nanocomposites with various MWNT concentra- strain sensors. Cyclic tensile tests were performed for each
tions were printed and characterized to identify the optimal set of the fabricated strain sensors using an Instron single
nanocomposite formulation with the highest piezoresistive column uniaxial test machine at maximum strains of 10, 15,
sensing function. 20, 25, and 30%. The electrical resistance of the tested sam-
A Tronxy X5S system was modified to 3D print the con- ples was continuously recorded using an Agilent 34401A
ductive nanocomposites by installing a geared syringe pump multimeter during the mechanical tests for all samples. The
(Abshirini et al. 2018). Plastic syringes capped with needles experimental setup is shown in Fig. 3. The load/unload

Fig. 1  Schematics of the a nanocomposite material synthesis and b strain sensor fabrication processes

13
Applied Nanoscience

Fig. 2  a Modified 3D printer,


b depositing nanocomposite
on the PDMS substrate, c final
shape of the printed strain sen-
sor on pristine PDMS substrate
(ruler for scale shown in mil-
limeter and centimeter)

concentrations: 1.5 wt%, 2 wt%, 2.5 wt%, and 3 wt%.


All nanocomposite sensors were printed using a needle
of 0.41 mm diameter and cured at 130 °C. The optimal
nanocomposite with the highest piezoresistive sensitivity
was identified using the cyclic mechanical tests discussed
above. The nanocomposite with the selected CNT content
was used to print sensors using four needle diameters of
0.26, 0.41, 0.71 and 1.19 mm, cured at 130 °C, and tested
under the same cyclic load conditions as discussed above
to explore the effect of needle diameter on the proper-
ties of the printed nanocomposite sensors. Finally, the
nanocomposites printed with the optimal formulation and
needle size were cured at three different temperatures of
25 °C, 65 °C, and 130 °C to investigate the effect of cur-
Fig. 3  Experimental setup of piezoresistive sensing tests under cyclic ing temperature on their piezoresistive sensing function.
tensile load Nanocomposite sensors printed using all the optimized
parameters and formulations were tested to evaluate their
procedure was repeated for six cycles to ensure the repeat- sensing reliability and repeatability. Uniaxial tensile loads
ability of the sensing response in each test. The normalized at the rate of 1 mm/min were applied to identify the max-
electrical resistance change was calculated to quantify the imum strain that the sample carried before fracture. In
sensing behavior of the 3D-printed sensors using Eq. 1. addition, cyclic tensile loads were applied to find the maxi-
mum strain range in which the sensor showed a linear pie-
R − Ro
Normalized resistance change = 100 × (%) (1) zoresistive response. Moreover, the load rate dependency
Ro
of the printed sensors was investigated under cyclic tensile
where Ro is the initial electrical resistance of the sensor and load at loading rates varying from 5 to 200 mm/min, at a
R is the real-time measured electrical resistance. In addition, constant 10% maximum strain. The piezoresistive sens-
GF was calculated to quantify the piezoresistive sensitivity ing response was recorded throughout all the experiments.
of the 3D-printed sensors using Eq. 2. The long-term sensing and fatigue behavior of the printed
sensors was tested under tensile loads at a loading rate of
3 mm/min and 10% max strain for 400 cycles. Finally, the
( )
R−R0

GF =
R0 (2) stress and electrical resistance relaxation behaviors of the
𝜀 sensors were investigated by holding the sample at 10%
where ε (mm/mm) is the applied strain on the sample. strain load for 8000 s while recording electrical resistance
The effect of the MWNT concentration on the piezore- and stress data.
sistive behavior of the printed sensors was evaluated
by testing nanocomposites with four different MWNT

13
Applied Nanoscience

In situ characterization of the piezoresistive sensing Human body motion recognition


mechanism
The feasibility of using the developed strain sensors as wear-
Understanding the piezoresistive sensing mechanism is able sensors was explored. Five strain sensors were printed
critical for the optimization of 3D printed sensor behav- and mounted on a glove to detect finger motions, and the
ior. In this paper, in situ micro-mechanical tensile tests sixth sensor was attached on the wrist area to detect wrist
were conducted to investigate the reorganization of the bending. Electrical resistance data were recorded throughout
electrically conductive MWNT network in the nano- the test to validate the sensors’ performance as wearable
composites under tensile load. A printed nanocomposite electronic devices.
sensor was manufactured without a pristine PDMS seal-
ing layer, clamped on a micro-mechanical testing stage
(Gatan Microtest 200), and tested in an SEM (TESCAN Results and discussion
VEGA II), while SEM images were captured at different
strain increments, as shown in Fig. 4. The reorganization Distribution of the nanocomposite pattern
of the MWNT network in the nanocomposite was dem- in the sensor
onstrated by comparing the SEM images at 0% and 20%
strain. Figure 5a shows the SEM images of the cross-section of the
strain sensor, detailing the MWNT distribution within the
printed nanocomposite. The half oval shape with a width of
598 μm and a height of 228 μm was the conductive nano-
composite embedded between two layers of pristine PDMS.
Due to surface wetting, the cured cross-section of the 3D
printed nanocomposite sensor was in the shape of half oval
instead of half circle, as outlined in Fig. 5a. Good disper-
sion of the MWNT in the PDMS matrix can be observed in
the high-magnification SEM image in Fig. 5b. No MWNT
agglomerates or fractured nanotubes were observed during
SEM imaging. The width and thickness of the deposited pat-
tern using different needles were measured from the SEM
images captured from cross-sections and tabulated in Table 1.

Optimization of MWNT concentration


in nanocomposites

Fig. 4  In situ SEM analysis of the strain sensor using  a micro- To study the effect of MWNT concentration on the piezore-
mechanical tensile stage sistive sensing function of the conductive nanocomposite,

Fig. 5  SEM images of the nanocomposite: a printed conductive nanocomposite on pristine PDMS substrate, b distribution of MWNT in the
PDMS matrix

13
Applied Nanoscience

Table 1  Width and thickness of the deposited pattern after curing nanocomposites with 3 wt%, 2.5 wt%, and 2 wt% MWNT
Needle size (mm) Width (mm) Thickness (mm) were 1.15, 1.56, and 2.51, respectively. However, the average
GF of nanocomposite with 1.5 wt% increased up to 12.98.
0.26 0.394 0.135 Nanocomposites with 1 wt% MWNT were also printed
0.41 0.598 0.228 and tested following the same testing procedure. However,
0.71 1.122 0.452 their poor electrical conductivity did not offer a measur-
1.19 1.861 0.736 able piezoresistive sensing function. Since the piezoresis-
tive sensing mechanism is due to the reorganization of the
conductive network of MWNTs under external load, an
appropriate amount of MWNTs can form the most effec-
tive micro-scale connections among MWNTs, resulting in
the nanocomposites being most sensitive to external loads.
Therefore, nanocomposites with 1.5 wt% MWNT are con-
sidered the optimal material for sensor development in this
paper.

Optimization of nanocomposite printing


parameters

The effect of the printing needle diameter was investigated


by printing nanocomposites with 1.5 wt% of MWNT using
four different needle diameters. As shown in Fig. 7a, the
increased needle diameters resulted in wider nanocomposite
patterns. The piezoresistive sensing function of these sen-
Fig. 6  The effect of CNT concentration on the strain sensing response sors is expected to be different, since the MWNT distribu-
tions and alignment in the nanocomposites can be affected
by the shear flow generated during the printing process. The
sensors fabricated with four different MWNT formulations measured piezoresistive sensing functions of the four types
were tested via cyclic loading at max strains between 10 and of sensors tested under 10–30% strain are shown in Fig. 7b.
30% at a loading rate of 2 mm/min. The electrical resist- It should be noted that the piezoresistive sensitivity was
ance changes of nanocomposites with various MWNT con- enhanced by reducing the needle diameter from 1.19 mm to
centrations are shown in Fig. 6. By reducing the MWNT 0.41 mm. The average GF is 13.01 for the sensors fabricated
loading, the piezoresistive sensitivity of the printed nano- with needles of 0.41 mm diameter compared to 1.92 for the
composites improved, as indicated by the increased slopes sensors printed using 1.19 mm diameter needles.
at lower MWNT contents in Fig. 6. The average GFs of

Fig. 7  a Printed sensors of various widths due to the effect of needle diameter, b piezoresistive sensing function of sensors printed by various
needle diameter

13
Applied Nanoscience

The effect of needle diameter on the piezoresistive sens-


ing function can be due to the alignment of the MWNT
and formation of the conductive network within the printed
nanocomposites. During the printing process, the flow rate
of nanocomposites near the needle wall was lower than that
near the center of the needle, resulting in the shear stress
within the nanocomposite caused by the extrusion. The shear
stress at a surface element parallel to the needle wall can be
written as:
𝜕u
𝜏(y) = 𝜇
𝜕y (3)

where μ is the dynamic viscosity of the flow, u is the flow


velocity along the needle wall, and y is the distance away
from the needle wall. Shear stress caused by shear flow can
align MWNT nanoparticles along the printing direction, as
shown in Fig. 8. Since the same amount of MWNT nano- Fig. 9  Effect of curing temperature on the piezoresistive sensing
function
composites was printed due to the constant extrusion motor
speed, the generated flow rate and shear stress using smaller
needles were higher than those using larger needles, result- of the printed sensors was reduced when low curing temper-
ing in better alignment of MWNT within nanocomposites. ature was used. The average GFs of nanocomposites cured at
Similar phenomena have been reported in literature when 25 °C, 65 °C, and 130 °C were 1.8, 7.98, and 12.91, respec-
micronozzles were used to deposit nanofillers in nanocom- tively, when tested under 10–30% of tensile strain. It should
posites (Compton and Lewis 2014; Farkash and Brandon be noted that it only took 10 min to fully cure the printed
1994; Shofner et al. 2003). However, if the needle diameter nanocomposite at 130 °C, compared to 96 h when cured at
is too small, the printed nanocomposite sensors may have 25 °C. Hence, it can be estimated that the alignment of the
discontinuous MWNT alignment and MWNT breakage MWNTs was disrupted during the long curing process when
during printing, resulting in reduced piezoresistive sensing a low curing temperature was employed. To take advantage
function. The poor sensing behavior can be more significant of CNT alignment generated during the printing process, the
under high tensile strains. This is evident by the non-linear microstructure of the conductive MWNT network should be
response of the 0.26 mm needle size after 20% strain, which “locked” quickly by curing the nanocomposites as quickly
is likely due to poor and disrupted MWNT network in this as possible. The ideal curing process should be the in situ
sensor under high extension condition. curing of nanocomposites during the printing. Otherwise,
The curing temperature effect on the piezoresistive sens- the initial orientation of the MWNTs in the nanocomposites
ing function was characterized by printing sensors using the can change and reduce the piezoresistive sensing function of
1.5 wt% MWNT nanocomposite and a 0.41 mm needle, and the printed nanocomposites.
immediately curing the printed nanocomposites at three dif- According to the obtained results, the best piezore-
ferent temperatures of 25 °C, 65 °C, and 130 °C. Figure 9 sistive sensing function can be achieved using the 1.5
shows the piezoresistive response of these samples tested wt% MWNT printed on the pristine PDMS substrate
under 10–30% of tensile strain. The piezoresistive sensitivity using 0.41 mm needle and cured at 130 °C. The optimal

Fig. 8  Schematic of nanotube
alignment during the printing
and realignment during the
loading

13
Applied Nanoscience

nanocomposite formulation and manufacturing param- Characterization of piezoresistive sensing function


eters were used to fabricate all the stretchable strain sen-
sors that were characterized in the following sections. To evaluate the maximum strain that the sensors carried
prior to fracture, continuous uniaxial tensile load was
applied at the load rate of 0.2 mm/min. Figure 10a shows
the sensor mounted on the Instron uniaxial test machine
before applying the load and the sensor at the moment before

Fig. 10  a Printed sensor under 0% and 146% tensile strains; b Pie- ing rate of 5  mm/min, 25  mm/min, 50  mm/min; d durability test of
zoresistive sensing response of the flexible strain sensor from 10 to printed sensor for 400 cycles; e relaxation test showing the resistance
70% tensile strain; c effect of loading rate on piezoresistive sensing and stress change under 10% strain for 8000 s; f FE results showing
behavior and details of the relative piezoresistive change under load- the strain distribution in the layers of the sensor at 40% strain

13
Applied Nanoscience

failure. The maximum strain the sensors carried was 146% as the change in the real-time value from the initial values,
tensile strain, which was close to the reported maximum divided by the initial value) throughout the experiment are
strain that PDMS can sustain (elongation, 160%) (Cai et al. depicted in Fig. 10e. Both stress and resistance degraded due
2013). This high fracture strain allows the potential applica- to the viscoelastic behavior of the elastomer polymer in con-
tion of the developed sensor under high strain conditions. stant extension. However, the stress reduction was around 8%
The piezoresistive response of the printed sensors at dif- compared to 3.4% for the resistance, which revealed that the
ferent max strains was characterized to identify the linear relaxation behavior of the stress was higher than that of the
strain sensing range. Cyclic tensile load tests at various max piezoresistance response, likely due to additional relaxation
strains were conducted starting from 10% strain, increment- in the pristine PDMS layers, and not the conductive pattern.
ing by 5% up to 70% strain. The piezoresistive response Finite element analysis (FEA) simulations and experi-
of the sensor was measured continuously and is shown in mental validations were conducted to analyze the differ-
Fig. 10b. The electrical resistance of the nanocomposites ence between the overall applied strain on the sensor and
above 70% strain was beyond the working range of the mul- the localized tensile strain carried by the printed nano-
timeter used in this study, therefore, the sensing performance composite pattern embedded in the sensors. Detailed FEA
above 70% tensile strain is not reported in this paper. The models and experimental validations of the FEA model
sensing results in Fig. 10b shows that the sensor exhibits a using a video extensometer were discussed in the Support-
linear response throughout the tested 10–70% strain range, ing Material. Numerical results showing the longitudinal
with an average GF of 12.15, which is about six times higher strain distribution in the substrate, nanocomposite pattern,
than that of conventional metallic strain gauges (Yamada and the top PDMS layer at the 40% applied tensile strain
et al. 2011). The wide linear sensing range (at least up to are depicted in Fig. 10f. The FEA simulation and experi-
70% tensile strain) with a high GF demonstrated the poten- mental results showed that the true tensile strain carried by
tial strain sensing capability under large deformations. the printed nanocomposite was 6% smaller than the average
To evaluate the load rate dependency of the printed sen- strain applied to the sensor during mechanical testing for
sors, cyclic tensile tests at 10% max strain were conducted nanocomposite sensors with 1.5 wt% MWNT. Although this
under various loading rates. As shown in Fig. 10c, the pie- smaller strain can increase the calculated GF of the strain
zoresistive response initially improved with the increased sensor, the difference in applied strain and actual strain on
loading rate, indicating enhanced piezoresistive sensitivity at the nanocomposite was so small that the correction factor
a higher loading rate, but the sensing behavior stabilized after was not included in the experimental analysis to simplify the
the loading rate reached 50 mm/min. This demonstrated that data processing procedures.
the viscoelastic behavior of the material was affected at load-
ing rates of 50 mm/min or lower, with the effect disappear- In situ characterization of the piezoresistive sensing
ing at loading rates of 50–200 mm/min. The detailed cyclic mechanism
piezoresistive response of the sensor for the tests at loading
rates of 5, 25, and 50 mm/min is also shown in Fig. 10c. In situ micro-mechanical testing under SEM was conducted
An important parameter to verify the robustness of a to illustrate the piezoresistive sensing mechanism by dem-
strain sensor is its long-term sensing response. The piezore- onstrating the reorganization and rearrangement of MWNT
sistive behavior of the sensor under a tensile fatigue test of conductive network within the printed nanocomposite. The
400 cycles loaded up to 10% strain is shown in Fig. 10d. SEM image of a relaxed nanocomposite sample (0% strain) is
The cyclic resistance change degraded gradually in the first shown in Fig. 11. Although only a low percentage of the
100 cycles and then became relatively stable. The reduction MWNTs was on the surface exposed to the SEM imaging,
occurred in both the peaks and valleys in the initial cycles. the realignment of MWNT was visible in several zones. After
The peak to valley resistance response was illustrated sepa- applying a 20% tensile strain on the sample, the rotation and
rately for three different points throughout the test: cycles sliding of MWNT were observed, and is shown in the magni-
20–25 (start), 200–205 (middle), and 380–385 (end). The fied SEM images (P1–P5 zone) in Fig. 11. The zoomed SEM
relative resistance that changed in the cycles of 20–25 was images with solid borders show the initial orientation of the
around 5.1% higher than that of cycles of 200–205 and MWNT (at 0% strain), while the pictures with dashed border
380–384. This result shows the acceptable long-term per- show the reorganization of MWNT network under 20% ten-
formance of the printed sensors. sile strain. The reduced gap between two adjacent nanotubes
To understand the relaxation behavior of the strain sensor, can be seen in zones P1 and P3. Additionally, zones P2–P4
a 10% strain tensile load was applied and held for 8000 s. The show two separated nanotubes connecting after applying
sensor’s stress and resistance were recorded continuously by the load. The cumulative effect of the rearrangement of the
the Instron test machine load cell and multimeter, respec- MWNTs in the conductive network resulted in the reduc-
tively. The normalized stress and resistance change (defined tion of the resistance of the conductive nanocomposite by

13
Applied Nanoscience

Fig. 11  SEM images of the


working area in the in situ
micro-mechanical testing under
SEM: SEM images for five dif-
ferent zones (P1–P5) comparing
the initial unstrained state (show
in the boxes with solid borders)
and final loaded state at 20%
strain (show in boxes with
dashed borders)

increasing the tensile load. This piezoresistive sensing func- as wearable sensors, with potential applications in robotics
tion of the strain sensor can be observed in Fig. 10c, d. and biomedical sensors.

Human body motion recognition Conclusion

To validate the application of the 3D-printed sensor as wear- Piezoresistive strain sensors consisting of PDMS/MWNT
able sensors, a “smart” glove was prepared for human hand nanocomposites were 3D printed on a PDMS substrate and
motion detection. Five sensors were attached to each finger characterized for potential wearable sensor applications. The
on the glove to detect the finger motion, and one sensor was detailed nanocomposite formulation and 3D printing param-
attached to the wrist area to monitor the bending motion of eters were first optimized to obtain the sensors with the high-
the wrist, as shown in Fig. 12a. Two tests were performed to est piezoresistive sensitivity. The strain sensing capability of
validate the sensors. First, the fingers and wrist were folded the sensors was evaluated by testing the printed sensors under
and unfolded repeatedly, while the piezoresistive sensing cyclic tensile loads. The experimental results showed that the
data were recorded simultaneously. Relative electrical resist- nanocomposites with an average GF of 13.01 in the 10–30%
ance changes between 22 and 55% were obtained during the strain range could be manufactured using a PDMS/MWNT
tests, as shown in Fig. 12b. In the second test, there were formulation with 1.5 wt% MWNTs, printed by 0.41 mm
10 s holding period after the bending and relaxation. The diameter needles, and cured at 130 °C. The 3D printing pro-
recorded piezoresistive sensing results shown in Fig. 12c cess was able to enhance the alignment of MWNTs in the
were able to capture the bending and holding of each motion. nanocomposites, leading to improved sensor performance.
In particular, the plateau area in each cycle represented the The developed sensors with optimal properties carried up to
holding time in both the bending and relaxation periods of 146% tensile strain before fracture, demonstrating their high
the motion. The peaks of the plots showed the relaxation and stretchability and flexibility. Experimental results showed
the valleys showed the bending motion. These experiments a linear sensing response, with an average GF of 12.15 in
demonstrated the potential of using the 3D-printed sensors the 10–70% strain range. Additionally, the highly repeatable

13
Applied Nanoscience

Fig. 12  a Smart glove demon-


strating the six possible individ-
ual sensing modes (five for the
fingers and one for the wrist);
b sensing response of repeated
bending of human hand joints;
c sensing response of hold and
bending of hand joints

sensing response was observed in the long-term 400-cycle References


tensile test. The long-term relaxation test showed the sen-
sors only suffered a modest 3.4% resistance degradation after Abshirini M, Charara M, Liu Y, Saha M, Altan MC (2018) 3D print-
8000 s under a tension load, due to the viscoelastic properties ing of highly stretchable strain sensors based on carbon nanotube
nanocomposites. Adv Eng Mater 20:1800425
of the PDMS polymer. The application of the strain sensors Agarwala S, Goh GL, Le Dinh T-S, An J, Peh ZK, Yeong WY, Kim Y-J
in wearable electronics was investigated by attaching six sen- (2018) Wearable bandage-based strain sensor for home healthcare:
sors on a glove to detect bending in the human fingers and the combining 3D aerosol jet printing and laser sintering. ACS Sens
wrist. The results indicated that the fabricated sensors were 4:218–226
Amjadi M, Pichitpajongkit A, Lee S, Ryu S, Park I (2014) Highly
able to monitor the motion of the human body joints and can stretchable and sensitive strain sensor based on silver nanowire–
be used in the wearable sensing applications. elastomer nanocomposite. ACS Nano 8:5154–5163

13
Applied Nanoscience

Amjadi M, Kyung KU, Park I, Sitti M (2016) Stretchable, skin-mount- Ho DH, Sun Q, Kim SY, Han JT, Kim DH, Cho JH (2016) Stretch-
able, and wearable strain sensors and their potential applications: able and multimodal all graphene electronic skin. Adv Mater
a review. Adv Funct Mater 26:1678–1698 28:2601–2608
Cai L, Song L, Luan P, Zhang Q, Zhang N, Gao Q, Zhao D, Zhang Hoang PT, Salazar N, Porkka TN, Joshi K, Liu T, Dickens TJ, Yu
X, Tu M, Yang F (2013) Super-stretchable, transparent carbon Z (2016) Engineering crack formation in carbon nanotube-silver
nanotube-based capacitive strain sensors for human motion detec- nanoparticle composite films for sensitive and durable piezoresis-
tion. Sci Rep 3:3048 tive sensors. Nanoscale Res Lett 11:422
Cao H, Thakar SK, Oseng ML, Nguyen CM, Jebali C, Kouki AB, Hong SY, Lee YH, Park H, Jin SW, Jeong YR, Yun J, You I, Zi G, Ha
Chiao J-C (2015) Development and characterization of a novel JS (2016) Stretchable active matrix temperature sensor array of
interdigitated capacitive strain sensor for structural health moni- polyaniline nanofibers for electronic skin. Adv Mater 28:930–935
toring. IEEE Sens J 15:6542–6548 Hu N, Karube Y, Yan C, Masuda Z, Fukunaga H (2008) Tunneling
Charara M, Abshirini M, Saha MC, Altan MC, Liu Y (2019a) Highly effect in a polymer/carbon nanotube nanocomposite strain sensor.
sensitive compression sensors using three-dimensional printed Acta Mater 56:2929–2936
polydimethylsiloxane/carbon nanotube nanocomposites. J Intell Hwang HH, Zhu W, Victorine G, Lawrence N, Chen S (2018) 3D-print-
Mater Syst Struct 18:1045389X19835953 ing of functional biomedical microdevices via light- and extru-
Charara M, Luo W, Saha MC, Liu Y (2019b) Investigation of light- sion-based approaches. Small Methods 2:1700277
weight and flexible carbon nanofiber/poly dimethylsiloxane nano- Jakus AE, Secor EB, Rutz AL, Jordan SW, Hersam MC, Shah RN
composite sponge for piezoresistive sensor application. Adv Eng (2015) Three-dimensional printing of high-content graphene
Mater 1801068. https​://doi.org/10.1002/adem.20180​1068 scaffolds for electronic and biomedical applications. ACS Nano
Chen S, Wei Y, Wei S, Lin Y, Liu L (2016) Ultrasensitive cracking- 9:4636–4648
assisted strain sensors based on silver nanowires/graphene hybrid Kang I, Schulz MJ, Kim JH, Shanov V, Shi D (2006) A carbon nano-
particles. ACS Appl Mater Interfaces 8:25563–25570 tube strain sensor for structural health monitoring. Smart Mater
Choi DY, Kim MH, Oh YS, Jung S-H, Jung JH, Sung HJ, Lee HW, Lee Struct 15:737
HM (2017) Highly stretchable, hysteresis-free ionic liquid-based Kang D, Pikhitsa PV, Choi YW, Lee C, Shin SS, Piao L, Park B, Suh
strain sensor for precise human motion monitoring. ACS Appl K-Y, T-i Kim, Choi M (2014) Ultrasensitive mechanical crack-
Mater Interfaces 9:1770–1780 based sensor inspired by the spider sensory system. Nature
Chowdhury S, Olima M, Liu Y, Saha M, Bergman J, Robison T (2016) 516:222
Poly dimethylsiloxane/carbon nanofiber nanocomposites: fabrica- Kim JT, Pyo J, Rho J, Ahn J-H, Je JH, Margaritondo G (2012) Three-
tion and characterization of electrical and thermal properties. Int dimensional writing of highly stretchable organic nanowires. ACS
J Smart Nano Mater 7:236–247 Macro Lett 1:375–379
Chowdhury SA, Saha MC, Patterson S, Robison T, Liu Y (2018) Kumar N, Jain PK, Tandon P, Pandey PM (2018) Additive manufac-
Highly conductive polydimethylsiloxane/carbon nanofiber com- turing of flexible electrically conductive polymer composites via
posites for flexible sensor applications. Adv Mater Technol CNC-assisted fused layer modeling process. J Braz Soc Mech
4(1):1800398 Sci Eng 40:175
Cohen DJ, Mitra D, Peterson K, Maharbiz MM (2012) A highly elastic, Kwok SW, Goh KHH, Tan ZD, Tan STM, Tjiu WW, Soh JY, Ng ZJG,
capacitive strain gauge based on percolating nanotube networks. Chan YZ, Hui HK, Goh KEJ (2017) Electrically conductive
Nano letters 12:1821–1825 filament for 3D-printed circuits and sensors. Appl Mater Today
Compton BG, Lewis JA (2014) 3D-printing of lightweight cellular 9:167–175
composites. Adv Mater 26:5930–5935 Leigh SJ, Bradley RJ, Purssell CP, Billson DR, Hutchins DA (2012) A
Costa P, Silvia C, Viana J, Mendez SL (2014) Extruded thermoplas- simple, low-cost conductive composite material for 3D printing
tic elastomers styrene–butadiene–styrene/carbon nanotubes of electronic sensors. PloS One 7:e49365
composites for strain sensor applications. Compos Part B: Eng Li Q, Li J, Tran D, Luo C, Gao Y, Yu C, Xuan F (2017) Engineering
57:242–249 of carbon nanotube/polydimethylsiloxane nanocomposites with
Fallahi A, Bahramzadeh Y, Tabatabaie S, Shahinpoor M (2017) A enhanced sensitivity for wearable motion sensors. J Mater Chem
novel multifunctional soft robotic transducer made with poly C 5:11092–11099
(ethylene-co-methacrylic acid) ionomer metal nanocomposite. Lipomi DJ, Vosgueritchian M, Tee BC, Hellstrom SL, Lee JA, Fox CH,
Int J Intell Robot Appl 1:143–156 Bao Z (2011) Skin-like pressure and strain sensors based on trans-
Farkash M, Brandon D (1994) Whisker alignment by slip extrusion. parent elastic films of carbon nanotubes. Nat Nanotechnol 6:788
Mater Sci Eng, A 177:269–275 Liu Y, Rajadas A, Chattopadhyay A (2012) A biomimetic struc-
Gnanasekaran K, Heijmans T, Van Bennekom S, Woldhuis H, Wijnia S, tural health monitoring approach using carbon nanotubes. Jom
de With G, Friedrich H (2017) 3D printing of CNT-and graphene- 64(7):802–807
based conductive polymer nanocomposites by fused deposition Liu Z, Qi D, Guo P, Liu Y, Zhu B, Yang H, Liu Y, Li B, Zhang C, Yu J
modeling. Appl Mater Today 9:21–28 (2015) Thickness-gradient films for high gauge factor stretchable
Goh GL, Agarwala S, Tan YJ, Yeong WY (2018a) A low cost and strain sensors. Adv Mater 27:6230–6237
flexible carbon nanotube pH sensor fabricated using aerosol jet Liu H, Huang W, Gao J, Dai K, Zheng G, Liu C, Shen C, Yan X,
technology for live cell applications. Sens Actuators B: Chem Guo J, Guo Z (2016) Piezoresistive behavior of porous carbon
260:227–235 nanotube-thermoplastic polyurethane conductive nanocomposites
Goh GL, Agarwala S, Yeong WY (2018b) Directed and on-demand with ultrahigh compressibility. Appl Phys Lett 108:011904
alignment of carbon nanotube: a review toward 3D. Print Electr Lu N, Lu C, Yang S, Rogers J (2012) Highly sensitive skin-mounta-
Adv Mater Interfaces 6:1801318. https ​ : //doi.org/10.1002/ ble strain gauges based entirely on elastomers. Adv Func Mater
admi.20180​1318 22:4044–4050
Guo SZ, Qiu K, Meng F, Park SH, McAlpine MC (2017) 3D printed Luo W, Charara M, Saha MC, Liu Y (2018) Fabrication and char-
stretchable tactile sensors. Adv Mater 29:1701218 acterization of porous CNF/PDMS nanocomposites for sensing
Hinton TJ, Hudson A, Pusch K, Lee A, Feinberg AW (2016) 3D print- applications. Appl Nanosci 1–9
ing PDMS elastomer in a hydrophilic support bath via freeform
reversible embedding. ACS Biomater Sci Eng 2:1781–1786

13
Applied Nanoscience

Mannoor MS, Jiang Z, James T, Kong YL, Malatesta KA, Soboyejo Wang X, Jiang M, Zhou Z, Gou J, Hui D (2017) 3D printing of polymer
WO, Verma N, Gracias DH, McAlpine MC (2013) 3D printed matrix composites: a review and prospective. Compos Part B:
bionic ears. Nano letters 13:2634–2639 Eng 110:442–458
Martinez RV, Branch JL, Fish CR, Jin L, Shepherd RF, Nunes RM, Wei X, Li D, Jiang W, Gu Z, Wang X, Zhang Z, Sun Z (2015) 3D print-
Suo Z, Whitesides GM (2013) Robotic tentacles with three- able graphene composite. Sci Rep 5:11181
dimensional mobility based on flexible elastomers. Adv Mater Wu S, Zhang J, Ladani RB, Ravindran AR, Mouritz AP, Kinloch AJ,
25:205–212 Wang CH (2017) Novel electrically conductive porous PDMS/
Michelis F, Bodelot L, Bonnassieux Y, Lebental B (2015) Highly carbon nanofiber composites for deformable strain sensors and
reproducible, hysteresis-free, flexible strain sensors by inkjet conductors. ACS Appl Mater Interfaces 9:14207–14215
printing of carbon nanotubes. Carbon 95:1020–1026 Yamada T, Hayamizu Y, Yamamoto Y, Yomogida Y, Izadi-Najafabadi
Obitayo W, Liu T (2012) A review: Carbon nanotube-based pie- A, Futaba DN, Hata K (2011) A stretchable carbon nanotube
zoresistive strain sensors. J Sens 2012:652438. https​: //doi. strain sensor for human-motion detection. Nat Nanotechnol 6:296
org/10.1155/2012/65243​8 Yu WW, Zhang J, Wu JR, Wang XZ, Deng YH (2017) Incorporation
Park JJ, Hyun WJ, Mun SC, Park YT, Park OO (2015) Highly stretch- of graphitic nano-filler and poly (lactic acid) in fused deposition
able and wearable graphene strain sensors with controllable sen- modeling. J Appl Polym Sci 134:44703. https​://doi.org/10.1002/
sitivity for human motion monitoring. ACS Appl Mater Interfaces app.44703​
7:6317–6324. https​://doi.org/10.1021/acsam​i.5b006​95 Zhang R, Deng H, Valenca R, Jin J, Fu Q, Bilotti E, Peijs T (2013)
Ryu S, Lee P, Chou JB, Xu R, Zhao R, Hart AJ, Kim S-G (2015) Strain sensing behaviour of elastomeric composite films contain-
Extremely elastic wearable carbon nanotube fiber strain sensor ing carbon nanotubes under cyclic loading. Compos Sci Technol
for monitoring of human motion. ACS Nano 9:5929–5936. https​ 74:1–5
://doi.org/10.1021/acsna​no.5b005​99 Zhang S, Cai L, Li W, Miao J, Wang T, Yeom J, Sepúlveda N, Wang C
Selvan NT, Eshwaran S, Das A, Stöckelhuber K, Wießner S, Pötschke (2017) Fully printed silver-nanoparticle-based strain gauges with
P, Nando G, Chervanyov A, Heinrich G (2016) Piezoresistive nat- record high sensitivity. Adv Electr Mater 3:1700067
ural rubber-multiwall carbon nanotube nanocomposite for sensor Zheng WJ, An N, Yang JH, Zhou J, Chen YM (2015) Tough Al-algi-
applications. Sens Actuators, A 239:102–113 nate/poly (N-isopropylacrylamide) hydrogel with tunable LCST
Shofner M, Lozano K, Rodríguez-Macías F, Barrera E (2003) for soft robotics. ACS Appl Mater Interfaces 7:1758–1764
Nanofiber-reinforced polymers prepared by fused deposition Zheng Y, Li Y, Li Z, Wang Y, Dai K, Zheng G, Liu C, Shen C (2017)
modeling. J Appl Polym Sci 89:3081–3090 The effect of filler dimensionality on the electromechanical per-
Truby RL, Wehner M, Grosskopf AK, Vogt DM, Uzel SG, Wood RJ, formance of polydimethylsiloxane based conductive nanocompos-
Lewis JA (2018) Soft somatosensitive actuators via embedded 3D ites for flexible strain sensors. Compos Sci Technol 139:64–73.
printing. Adv Mater 30:1706383 https​://doi.org/10.1016/j.comps​citec​h.2016.12.014
Trung TQ, Ramasundaram S, Lee N-E (2017) Transparent, stretchable,
and rapid-response humidity sensor for body-attachable wearable Publisher’s Note Springer Nature remains neutral with regard to
electronics. Nano Res 10:2021–2033 jurisdictional claims in published maps and institutional affiliations.
Vatani M, Lu Y, Lee K-S, Kim H-C, Choi J-W (2013) Direct-write
stretchable sensors using single-walled carbon nanotube/polymer
matrix. J Electron Packag 135:011009

13
Reproduced with permission of copyright owner. Further reproduction
prohibited without permission.

Potrebbero piacerti anche