You are on page 1of 5

Research Article

www.acsami.org

Ultrasensitive Strain Sensor Produced by Direct Patterning of Liquid


Crystals of Graphene Oxide on a Flexible Substrate
M. Bulut Coskun,†,°,• Abozar Akbari,‡,• Daniel T. H. Lai,*,§ Adrian Neild,† Mainak Majumder,‡
and Tuncay Alan*,†

Laboratory for Micro Systems, Department of Mechanical and Aerospace Engineering, Monash University, Melbourne 3800,
Australia

Nanoscale Science and Engineering Laboratory, Department of Mechanical and Aerospace Engineering, Monash University,
Melbourne 3800, Australia
§
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

College of Engineering and Science, Victoria University, Melbourne 3011, Australia


*
S Supporting Information

ABSTRACT: Ultrasensitive flexible strain sensors were developed


Downloaded via MONASH UNIV on May 4, 2023 at 22:55:40 (UTC).

through the combination of shear alignment of a high concentration


graphene oxide (GO) dispersion with fast and precise patterning of
multiple rectangular features on a flexible substrate. Resistive
changes in the reduced GO films were investigated under various
uniaxial strain cycles ranging from 0.025 to 2%, controlled with a
motorized nanopositioning stage. The devices uniquely combine a
very small detection limit (0.025%) and a high gauge factor with a
rapid fabrication process conducive to batch production.

KEYWORDS: strain gauge, rGO, direct GO patterning, shear alignment, high sensitivity, flexible substrate

1. INTRODUCTION graphite. GO sheets are essentially graphene layers decorated


Wearable, flexible strain sensors have attracted significant with oxygen-bearing functional groups26 which can be removed
attention in a myriad of applications ranging from biomedical to obtain reduced graphene oxide (rGO).
health monitoring1,2 to soft robotics.3,4 Thanks to their superior We developed a new class of highly flexible and ultrasensitive
mechanical and electrical properties and synergy with flexible rGO strain sensors based on our shear-alignment technique27
substrates, graphene-based devices offer improved measure- of liquid crystal nematic phase of GO dispersion which can
ment sensitivities over conventional micromachined sensors,5−7
produce large-area films with controlled structural properties.
and there has been a considerable interest in their use for strain
sensing.8−22 Here, we improved this technique to cast precisely defined rGO
The electrical resistance of pristine graphene changes when layers on flexible polydimethylsiloxane (PDMS) substrates
subjected to strain.11 However, approaches based on structural which are selectively masked by a temporary vinyl layer (Figure
deformation of a single layer of graphene necessitate very high 1). The direct casting on masked substrates takes place at room
strain inputs to open the band gap of the material,10 making temperature, is not sensitive to the substrates being used,27 and
them unsuitable for detection of very low strains. A substantial can generate well-defined patterned structures without
improvement in sensitivity can be obtained using multilayered
configurations9,12,13,23 in which the contacting graphene flakes subsequent transfer processes, resulting in improvements over
form an electrical network. The introduction of strain alters the well-established techniques such as spin coating.28 Similarly, the
overlapping areas between flakes, significantly changing the processing time is significantly shorter than that for vacuum
conductivity of the device. filtration8 methods. Finally, the geometric definition is better
Moreover, the use of pristine or chemical vapor deposited controlled in comparison to that of spray coating.9 This
graphene, while offering a very high quality, requires transfer provides a rapid and scalable technique. However, it is known
techniques or subsequent nanofabrication steps.24 Yet, to make
that, during the drying of sprayed drops, nonuniform deposits
full use of the potential offered by miniaturized flexible devices
in practical applications, the devices should be produced may compromise the uniformity of the films produced.29
through rapid, cheap processes conducive to high-volume
production.25 This challenge can be addressed using a cheaper Received: May 26, 2016
alternative: reduced graphene oxide (rGO), produced in Accepted: August 4, 2016
dispersion17 through the oxidative chemical processing of Published: August 4, 2016

© 2016 American Chemical Society 22501 DOI: 10.1021/acsami.6b06290


ACS Appl. Mater. Interfaces 2016, 8, 22501−22505
ACS Applied Materials & Interfaces Research Article

Figure 1. Masked shear-alignment process. (a) The process starts with a blank substrate. (b) Then, a vinyl mask with the desired features is placed
on the substrate, (b) 0.5 mL of a GO dispersion is poured on the edge of the substrate, and (c) the dispersion is cast by moving a Doctor Blade. (d)
The mask is peeled off, and (e) remaining features on the substrate are reduced at 110 °C for 8 h and diced. (f) Photograph of the rGO film on the
substrate after the casting process. (g) FTIR spectra of the graphene-based sensor before and after reduction by thermal treatment. (h) XRD pattern
of the graphene-based sensor before and after reduction by thermal treatment.

2. EXPERIMENTAL SECTION the oxygen functional groups from the GO sheets to obtain a
conductive rGO film. In this regard, a heat treatment technique was
2.1. Preparation and Concentration of GO Dispersion. The used (at 110 °C for 8 h) to reduce the GO film (Figure 1e).
high concentration nematic liquid crystal phase of GO dispersion used During the drying and heat treatment stages, the thickness of the
in the fabrication is a shear-thinning fluid which allows casting of a GO film decreased substantially. A photograph of the rGO sensor
uniform liquid thin film under high shear stress imposed by the motion resulting from this casting process is shown in Figure 1f.
of a doctor blade, while a highly viscous phase in the absence of shear 2.5. Film Characterization. The GO film was characterized via X-
stress ensures uniformity during the subsequent film drying step. A low ray diffraction (XRD) and Fourier transform infrared (FTIR) methods
concentration GO (∼0.25 mg mL−1) was prepared by Hummers’ before and after reduction. The FTIR spectra were obtained using an
method30 followed by concentrating the dispersion by the dehydration attenuated total reflectance Fourier transform infrared (ATR-FTIR)
technique,27 where a cross-linked polyacrylate copolymer hydrogel was instrument (PerkinElmer, United States) in the range of 600−4000
used to absorb the water content until the desired concentration of 35 cm−1 at an average of 32 scans with a resolution of 4 cm−1. In the case
mg/mL was achieved. Polarized light microscopy using an LC- of GO, there was a strong and broad peak in the range of 3100−3700
PolScope imaging system (40× 0.75/0.85NA objective) was used to cm−1 and a peak at 1650 cm−1 corresponding to the stretching of
confirm the nematic phase of the GO dispersion.31,32 With the LC- hydroxyl groups, possibly from COOH groups in GO and/or the
Polscope technique, it is possible to detect the retardance along with presence of H2O in GO interlayers. Peaks at 1728, 1205, and 1070
their dominant orientation simultaneously at all points of the image cm−1 are attributed to stretching of the CO in carbonyl groups, C−
without having to mechanically rotate the sample (Figure S1). O−C in epoxide groups, and C−O in alkoxy groups, respectively.
2.2. PDMS Substrate Preparation. A mixture of Sylgard 184 Peaks at 1383 and 1400 cm−1 are attributed to the vibration of O−H
curing agent and base with a mass ratio of 1:10 was prepared. The in C−OH groups. The XRD patterns of the GO and rGO films were
mixture was then poured in to a 6″ Petri dish and placed in a recorded using a Phillips 1140 diffractometer with a Cu Kα line
desiccator for half an hour for degassing. Later, degassed PDMS was generated at 40 kV and 25 mA at a scan rate of 1°/min and a step size
thermoset for 6 h at 70 °C on a hot plate. Finally, the cured PDMS of 0.02°.
substrate was plasma treated with 18 W power under 0.3 mbar
pressure for 5 min (Harrick Plasma PDC-32G) to introduce oxygen-
bearing functionalities to improve adhesion with the GO dispersion. 3. RESULTS AND DISCUSSION
2.3. Masking Process. Rectangular patterns (35 × 10 mm) on the The FTIR spectra of the cast rGO film (Figure 1g) showed that
vinyl tape were drawn using CutStudio and programmed to be cut by a the oxygen functional groups experienced a reduction as
vinyl plotter (Roland SV-8). The plotter removes the rectangular
features from the vinyl tape (so that later the openings will be filled evidenced by a significant decrease in the intensity of the
with GO in the casting process). The desired sensor patterns were relevant oxygen functional group peaks and a presence a new
attached onto a plasma-treated 800 μm-thick PDMS substrate, peak at 1550 cm−1 which corresponds to a CC bond,
effectively masking selected regions on the substrate confirming reduction of the GO by the thermal treatment.
2.4. GO Film Casting. A simple lab-scale Doctor Blade (MTI Similarly, the XRD spectrum (Figure 1h) revealed that the
Corporation, United States) was used as a shear-alignment apparatus interlayer distance of the rGO film decreased from ∼0.91 nm
to cast a continuous GO film. A 0.5 mL of nematic phase GO (2θ = 9.68°) for the GO film to ∼0.39 nm (2θ = 22.8°),
dispersion (35 mg mL−1) was deposited at the edge of the mask confirming reduction of the GO sensor in agreement with the
(Figure 1b). To create the thin film, the Doctor Blade was then moved observed FTIR spectra for these two samples. The end result
over the surface of the vinyl. This enabled spreading of the GO
dispersion into the patterned features shown in Figure 1c and crucially after casting was a uniform 6.4 μm-thick continuous film of
provided the shear required for alignment of the graphene flakes. rGO on an 800 μm-thick flexible PDMS substrate, as shown in
Removal of the vinyl mask resulted in the desired GO patterns on the Figures 2a and b. The reduction step19 resulted in a wrinkled
substrate (Figure 1d), which was left overnight at room temperature to surface with a 928 nm root-mean-square (rms) surface
dry. With the casting process complete, the final step was to remove roughness (Figure 2c), as measured with an optical
22502 DOI: 10.1021/acsami.6b06290
ACS Appl. Mater. Interfaces 2016, 8, 22501−22505
ACS Applied Materials & Interfaces Research Article

Figure 2. (a) Scanning electron micrograph of the deposited film. (b)


Optical micrograph showing the film cross section. (c) Surface
topography of the film resulting from the thermal reduction step. (d)
Line profiles in the direction parallel to the motion of the Doctor
Blade obtained with an optical profiler.

profilometer (Bruker Contour GT-I). The line profile in Figure


2d indicates a maximum step height of 3 μm along the direction
of the motion of the Doctor Blade, parallel to the axis in which
the strain measurements were made.
A typical strain gauge consists of a sensory conduction path
that generates a measurable electrical signal when subjected to
strain. For the present case, the dominant sensing mechanism is
the modification of the conduction path due to changes in the
overlapping area of, or the creation of gaps between, the
neighboring rGO sheets. As the thickness of the rGO film
(∼6.4 μm) is much smaller than that of the substrate (∼800
μm), it can be assumed that strain distribution across the
thickness is uniform9 and that the expansion response of the
system is dominated by the stiffness of the substrate. Figure 3. (a) Schematic of the test setup. The rGO-coated PDMS
To assemble the sensors, after making electrical connections sample is clamped on a fixed block from one side and on moveable
XYZ stage actuated by a picomotor. (b) Schematic of the sample
at both ends, the rGO-coated samples were clamped at both showing the direction of the applied strain (εxx) that leads to a slight
ends with one end attached to a fixed block and the other end contraction (εyy) in perpendicular direction driven by PDMS’s Poisson
connected to a linear XYZ actuator stage (Newport M562) as ratio (0.5). (c) Readout circuit based on an op-amp (Texas
shown in Figures 3a and b. The stage was driven by a Instruments, LM324N) used in an inverted configuration to detect
picomotor (Newport 8310CL motor and 8751CL closed loop the strain-induced changes in resistance.
driver) programmed to produce a displacement range of up to
300 μm with a 63.5 nm step resolution. To measure the
resistance changes of the rGO sensor under the applied strains, sensors.13 Experiments were conducted for strain levels
a read-out circuit based on an inverting op-amp configuration increasing from 0.025 to 2%. Figure 4a shows the dynamic
was implemented. The circuit consisted of a fixed input resistor response of the sensor under 4 different maximum strain levels
and the rGO film which acted as the variable feedback resistor (i.e., 0.05, 0.4, 0.8, and 1.2%) and demonstrates a perfect match
(Figure 3c). The change in the rGO film’s resistance was in the initial inclination of the output voltage. Figure 4b shows
measured by means of voltage variations for the corresponding the change in resistance of the PDMS-based sensor for 50 strain
strains as cycles of 0.6% maximum strain. The results indicate that the
response is repeatable and consistent. Figure 4c demonstrates
⎛R ⎞ the percentage change in resistance at increasing deformations.
Vout = −Vin⎜ RGO ⎟ The data were acquired by averaging the results from 3 separate
⎝ R SMD ⎠ (1)
experiments (each using different samples from a single batch)
where Vin is the input voltage and RRGO (base resistance ∼9.2 and show a 5.8% standard deviation. These results indicate a
kΩ) and RSMD (fixed 10 kΩ) are the resistance of the rGO film very high change in resistance, reaching over 400% at 2% strain
and surface mount resistor, respectively. Here, as Vin (−1 V) with two linear regimes below and above 0.8% strain.
and RSMD are fixed, the circuit provides a linear input-output We hypothesize that this is due to a transition in the effective
relation which would not be obtainable from a voltage divider mechanism that modifies the conduction path. For lower strains
or Wheatstone bridge circuit. Hence, through the use of eq 1, (below 0.8%), the resistance change is due to a change in the
changes in RRGO can be determined. During the experiments, overlapping areas of the neighboring rGO flakes. Whereas at
Vout was monitored and recorded using a data acquisition card increasing strains up to 2%, the separation of larger rGO flake
(NI PCI 6251) and Labview with a 100 Hz sampling frequency. groups or islands is responsible for the change in resistance.
The setup was placed under an optical microscope located on a Furthermore, above 2% strain, a permanent inclination in the
vibration isolation table to minimize environmental vibrations. device’s voltage output under rest position was observed. This
The rGO sensor was able to detect strains well below 0.1%, is due to the increase in base resistance of the devices which
which is rarely achieved in the literature for graphene-based indicates formation of irreversible cracks in the rGO film at
22503 DOI: 10.1021/acsami.6b06290
ACS Appl. Mater. Interfaces 2016, 8, 22501−22505
ACS Applied Materials & Interfaces Research Article

4. CONCLUSION
In summary, a new type of flexible and highly sensitive rGO-
based strain sensor was demonstrated. The device combines a
low cost and scalable deposition method and a simple detection
circuitry with a very high sensitivity (gauge factor up to 261).
The sensor can accurately detect strains as small as 0.025%
under cyclic loadings, hence providing significant advantages
for continuous real-time monitoring of minute deformations.
Direct pattern transfer through a masking process primarily
allows multiple copies of identical devices to be produced by a
single fabrication step. It also provides significant potential to
further improve the sensor size and performance through
geometric modifications.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acsami.6b06290.
Formation of liquid crystals of graphene oxide,
experimental results for rGO on the polyamide 66
substrate, and tabulated gauge factors of the previous
works from the literature (PDF)

■ AUTHOR INFORMATION
Corresponding Authors
*E-mail: Tuncay.Alan@monash.edu.au.
*E-mail: Daniel.Lai@vu.edu.au.
Figure 4. (a) Response of rGO-coated PDMS to cyclic strain cycles Present Address
(0.05, 0.4, 0.8, and 1.2%) recorded with a 100 Hz sampling frequency. °
M.B.C.: Department of Mechanical Engineering, University of
(b) Response of the film after 50 cycles of strains under 0.6%. Voltage Texas at Dallas, Richardson, Texas 75080, United States.
output here is converted to the rate of change of resistance via eq 1.
(c) Percentage change in the resistance of rGO. The slope of the linear Author Contributions

fits defines the gauge factor as 137.2 ± 7.9 for strains below 0.8% and M.B.C. and A.A. contributed equally to the work. A.A.
261.2 ± 15.1 for strains above 0.8%. prepared the rGO samples and characterized material proper-
ties. M.B.C. fabricated the flexible sensors, performed the
experiments and analyzed the data. D.L and T.A. oversaw the
research. The manuscript was written through contributions of
higher strain inputs. Hence, the measurement range of the all authors. All authors have given approval to the final version
present devices were set from 0 to 2% strain. of the manuscript.
The common way to relate electrical changes to the Notes
mechanical parameters for the strain sensors is the gauge The authors declare no competing financial interest.


factor (GF). This is effectively a measure of the sensitivity of
the strain sensor and is acquired from the slope of the rate of ACKNOWLEDGMENTS
change of the resistance-strain curve23
Part of this work was performed at the Melbourne Centre for
ΔR Nanofabrication (MCN), the Victorian Node of the Australian
GF =
( R ) National Fabrication Facility (ANFF). The authors acknowl-
ε (2) edge support received from the Monash University Post-
graduate Publications Award and Monash Centre for Atomi-
The GF was calculated by fitting a line to the data shown in cally Thin Materials.
Figure 4c for the two regimes. Using the slope of these portions
of the response curve yields GF values of 137.3 and 261.2 for
strains below and above 0.8%, respectively. As the strain cycles
■ REFERENCES
(1) Patel, S.; Park, H.; Bonato, P.; Chan, L.; Rodgers, M. A Review of
are still reversible above 0.8%, the sensor can easily be tuned to Wearable Sensors and Systems with Application in Rehabilitation. J.
have higher gauge factors at lower strains by prestraining. The Neuroeng. Rehabil. 2012, 9, 1−17.
acquired gauge factors are comparable to or better than (2) Scilingo, E. P.; Gemignani, A.; Paradiso, R.; Taccini, N.;
previously reported values,8,9,12−24 (Table S1) while offering a Ghelarducci, B.; DeRossi, D. Performance Evaluation of Sensing
Fabrics for Monitoring Physiological and Biomechanical Variables.
substantially simpler design and fabrication process. A similar IEEE Trans. Inf. Technol. Biomed. 2005, 9, 345−352.
set of experiments was conducted for rGO casted on a (3) Viry, L.; Levi, A.; Totaro, M.; Mondini, A.; Mattoli, V.; Mazzolai,
polyamide 66 substrate, and the results are demonstrated in B.; Beccai, L. Flexible Three-Axial Force Sensor for Soft and Highly
Figure S2. Sensitive Artificial Touch. Adv. Mater. 2014, 26, 2659−2664.

22504 DOI: 10.1021/acsami.6b06290


ACS Appl. Mater. Interfaces 2016, 8, 22501−22505
ACS Applied Materials & Interfaces Research Article

(4) McEvoy, M.; Correll, N. Materials That Couple Sensing, (24) Lee, Y.; Bae, S.; Jang, H.; Jang, S.; Zhu, S.-E.; Sim, S. H.; Song,
Actuation, Computation, and Communication. Science 2015, 347, Y. I.; Hong, B. H.; Ahn, J.-H. Wafer-Scale Synthesis and Transfer of
1261689. Graphene Films. Nano Lett. 2010, 10, 490−493.
(5) Kang, J. W.; Lee, J. H.; Hwang, H. J.; Kim, K.-S. Developing (25) Yong, K.; Ashraf, A.; Kang, P.; Nam, S. Rapid Stencil Mask
Accelerometer Based on Graphene Nanoribbon Resonators. Phys. Lett. Fabrication Enabled One-Step Polymer-Free Graphene Patterning and
A 2012, 376, 3248−3255. Direct Transfer for Flexible Graphene Devices. Sci. Rep. 2016, 6, 1−8.
(6) Kumar, M.; Bhaskaran, H. Ultrasensitive Room-Temperature (26) Treossi, E.; Melucci, M.; Liscio, A.; Gazzano, M.; Samorì, P.;
Piezoresistive Transduction in Graphene-Based Nanoelectromechan- Palermo, V. High-Contrast Visualization of Graphene Oxide on Dye-
ical Systems. Nano Lett. 2015, 15, 2562−2567. Sensitized Glass, Quartz, and Silicon by Fluorescence Quenching. J.
(7) Tian, H.; Shu, Y.; Wang, X.-F.; Mohammad, M. A.; Bie, Z.; Xie, Am. Chem. Soc. 2009, 131, 15576−15577.
Q.-Y.; Li, C.; Mi, W.-T.; Yang, Y.; Ren, T.-L. A Graphene-Based (27) Akbari, A.; Sheath, P.; Martin, S. T.; Shinde, D. B.; Shaibani, M.;
Resistive Pressure Sensor with Record-High Sensitivity in a Wide Banerjee, P. C.; Tkacz, R.; Bhattacharyya, D.; Majumder, M. Large-
Pressure Range. Sci. Rep. 2015, 5, 8603−8608. Area Graphene-Based Nanofiltration Membranes by Shear Alignment
(8) Fu, X.-W.; Liao, Z.-M.; Zhou, J.-X.; Zhou, Y.-B.; Wu, H.-C.; of Discotic Nematic Liquid Crystals of Graphene Oxide. Nat.
Zhang, R.; Jing, G.; Xu, J.; Wu, X.; Guo, W. Strain Dependent Commun. 2016, 7, 10891−10902.
Resistance in Chemical Vapor Deposition Grown Graphene. Appl. (28) Robinson, J. T.; Zalalutdinov, M.; Baldwin, J. W.; Snow, E. S.;
Phys. Lett. 2011, 99, 213107. Wei, Z.; Sheehan, P.; Houston, B. H. Wafer-scale Reduced Graphene
(9) Hempel, M.; Nezich, D.; Kong, J.; Hofmann, M. A Novel Class of Oxide Films for Nanomechanical Devices. Nano Lett. 2008, 8 (10),
Strain Gauges Based on Layered Percolative Films of 2d Materials. 3441−3445.
(29) Majumder, M.; Rendall, C.; Li, M.; Behabtu, N.; Eukel, J. A.;
Nano Lett. 2012, 12, 5714−5718.
Hauge, R. H.; Schmidt, H. K.; Pasquali, M. Insights into the Physics of
(10) Zhao, J.; Zhang, G.-Y.; Shi, D.-X. Review of Graphene-Based
Spray Coating of Swnt Films. Chem. Eng. Sci. 2010, 65, 2000−2008.
Strain Sensors. Chin. Phys. B 2013, 22, 057701.
(30) Hummers, W. S., Jr; Offeman, R. E. Preparation of Graphitic
(11) Kim, K. S.; Zhao, Y.; Jang, H.; Lee, S. Y.; Kim, J. M.; Kim, K. S.;
Oxide. J. Am. Chem. Soc. 1958, 80, 1339−1339.
Ahn, J.-H.; Kim, P.; Choi, J.-Y.; Hong, B. H. Large-Scale Pattern
(31) Oldenbourg, R.; Mei, G. New Polarized Light Microscope with
Growth of Graphene Films for Stretchable Transparent Electrodes. Precision Universal Compensator. J. Microsc. 1995, 180, 140−147.
Nature 2009, 457, 706−710. (32) Tkacz, R.; Oldenbourg, R.; Mehta, S. B.; Miansari, M.; Verma,
(12) Li, X.; Zhang, R.; Yu, W.; Wang, K.; Wei, J.; Wu, D.; Cao, A.; Li, A.; Majumder, M. Ph Dependent Isotropic to Nematic Phase
Z.; Cheng, Y.; Zheng, Q. Stretchable and Highly Sensitive Graphene- Transitions in Graphene Oxide Dispersions Reveal Droplet Liquid
on-Polymer Strain Sensors. Sci. Rep. 2012, 2, 870−875. Crystalline Phases. Chem. Commun. 2014, 50, 6668−6671.
(13) Trung, T. Q.; Tien, N. T.; Kim, D.; Jang, M.; Yoon, O. J.; Lee,
N. E. A Flexible Reduced Graphene Oxide Field-Effect Transistor for
Ultrasensitive Strain Sensing. Adv. Funct. Mater. 2014, 24, 117−124.
(14) Wang, Y.; Yang, R.; Shi, Z.; Zhang, L.; Shi, D.; Wang, E.; Zhang,
G. Super-Elastic Graphene Ripples for Flexible Strain Sensors. ACS
Nano 2011, 5, 3645−3650.
(15) Zhao, J.; He, C.; Yang, R.; Shi, Z.; Cheng, M.; Yang, W.; Xie, G.;
Wang, D.; Shi, D.; Zhang, G. Ultra-Sensitive Strain Sensors Based on
Piezoresistive Nanographene Films. Appl. Phys. Lett. 2012, 101,
063112.
(16) Yuan, W.; Zhou, Q.; Li, Y.; Shi, G. Small and Light Strain
Sensors Based on Graphene Coated Human Hairs. Nanoscale 2015, 7,
16361−16365.
(17) Liu, Q.; Zhang, M.; Huang, L.; Li, Y.; Chen, J.; Li, C.; Shi, G.
High-Quality Graphene Ribbons Prepared from Graphene Oxide
Hydrogels and Their Application for Strain Sensors. ACS Nano 2015,
9, 12320−12326.
(18) Bae, S.-H.; Lee, Y.; Sharma, B. K.; Lee, H.-J.; Kim, J.-H.; Ahn, J.-
H. Graphene-Based Transparent Strain Sensor. Carbon 2013, 51,
236−242.
(19) Chiacchiarelli, L. M.; Rallini, M.; Monti, M.; Puglia, D.; Kenny,
J. M.; Torre, L. The Role of Irreversible and Reversible Phenomena in
the Piezoresistive Behavior of Graphene Epoxy Nanocomposites
Applied to Structural Health Monitoring. Compos. Sci. Technol. 2013,
80, 73−79.
(20) Lee, C.; Jug, L.; Meng, E. High Strain Biocompatible
Polydimethylsiloxane-Based Conductive Graphene and Multiwalled
Carbon Nanotube Nanocomposite Strain Sensors. Appl. Phys. Lett.
2013, 102, 183511.
(21) Park, J. J.; Hyun, W. J.; Mun, S. C.; Park, Y. T.; Park, O. O.
Highly Stretchable and Wearable Graphene Strain Sensors with
Controllable Sensitivity for Human Motion Monitoring. ACS Appl.
Mater. Interfaces 2015, 7, 6317−6324.
(22) Rinaldi, A.; Proietti, A.; Tamburrano, A.; Ciminello, M.; Sarto,
M. S. Graphene-Based Strain Sensor Array on Carbon Fiber
Composite Laminate. IEEE Sens. J. 2015, 15, 7295−7303.
(23) Kim, Y.-J.; Cha, J. Y.; Ham, H.; Huh, H.; So, D.-S.; Kang, I.
Preparation of Piezoresistive Nano Smart Hybrid Material Based on
Graphene. Curr. Appl. Phys. 2011, 11, S350−S352.

22505 DOI: 10.1021/acsami.6b06290


ACS Appl. Mater. Interfaces 2016, 8, 22501−22505

You might also like