You are on page 1of 38

Polymer Reviews

ISSN: 1558-3724 (Print) 1558-3716 (Online) Journal homepage: https://www.tandfonline.com/loi/lmsc20

Advances in Responsively Conductive Polymer


Composites and Sensing Applications

Jianwen Chen, Yutian Zhu, Jinrui Huang, Jiaoxia Zhang, Duo Pan, Juying Zhou,
Jong E. Ryu, Ahmad Umar & Zhanhu Guo

To cite this article: Jianwen Chen, Yutian Zhu, Jinrui Huang, Jiaoxia Zhang, Duo Pan, Juying
Zhou, Jong E. Ryu, Ahmad Umar & Zhanhu Guo (2021) Advances in Responsively Conductive
Polymer Composites and Sensing Applications, Polymer Reviews, 61:1, 157-193, DOI:
10.1080/15583724.2020.1734818

To link to this article: https://doi.org/10.1080/15583724.2020.1734818

Published online: 10 Mar 2020.

Submit your article to this journal

Article views: 1215

View related articles

View Crossmark data

Citing articles: 25 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lmsc20
POLYMER REVIEWS
2021, VOL. 61, NO. 1, 157–193
https://doi.org/10.1080/15583724.2020.1734818

REVIEW

Advances in Responsively Conductive Polymer Composites


and Sensing Applications
Jianwen Chena,b, Yutian Zhua, Jinrui Huangc, Jiaoxia Zhangd, Duo Pane,f,
Juying Zhoug,j, Jong E. Ryuh, Ahmad Umari, and Zhanhu Guoj
a
College of Materials, Chemistry and Chemical Engineering, Hangzhou Normal University, Yuhang
District, Hangzhou, China; bState Key Laboratory of Polymer Physics and Chemistry, Changchun
Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun, China; cKey Laboratory of
Biomass Energy and Material, Jiangsu Province; Key Laboratory of Chemical Engineering of Forest
Products, National Forestry and Grassland Administration; National Engineering Laboratory for Biomass
Chemical Utilization, Institute of Chemical Industry of Forest Products, Chinese Academy of Forestry,
Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, Nanjing, Jiangsu
Province, China; dSchool of Materials Science and Engineering, Jiangsu University of Science and
Technology, Zhenjiang, China; eCollege of Chemical and Environmental Engineering, Shandong
University of Science and Technology, Qingdao, Shandong, China; fKey Laboratory of Materials
Processing and Mold (Zhengzhou University), Ministry of Education, National Engineering Research
Center for Advanced Polymer Processing Technology, Zhengzhou University, Zhengzhou, China;
g
School of Chemistry and Chemical Engineering, Guangxi University for Nationalities, Nanning, China;
h
Department of Mechanical and Aerospace Engineering, North Carolina State University, Raleigh, NC,
USA; iDepartment of Chemistry, College of Science and Arts, Promising Centre for Sensors and Electronic
Devices (PCSED), Najran University, Najran, Kingdom of Saudi Arabia; jIntegrated Composites Laboratory
(ICL), Department of Chemical & Biomolecular Engineering, University of Tennessee, Knoxville, TN, USA

ABSTRACT ARTICLE HISTORY


Conductive polymer composites (CPCs) have attracted intensive Received 12 September 2019
attention for several decades because they can endow the materials Accepted 20 February 2020
with not only good processability but also various functionalities
KEYWORDS
except the electrical conductivity. It is known that the electrical
Responsively conductive
resistance of the CPCs is dominated by the conductive networks polymer composites;
in the polymer matrix. Therefore, tiny change of the conductive strain/pressure sensors;
networks can lead to remarkable changes in the output electrical temperature sensors; vapor
signal of the CPCs. Utilizing this stimulus-response behavior of con- sensors; liquid sensors
ductive networks to the environment conditions, CPCs can be used
to design sensitive sensors to detect or monitor the environment
conditions, such as the strain/stress, pressure, temperature, solvent
or vapor. This review systematically outlines the preparation, micro-
structures, properties, and the stimulus-response mechanisms to the
environment conditions of the CPCs as well as their applications in
various sensitive sensors, including strain sensors, pressure sensors,
liquid sensors, vapor sensors, and temperature sensors. Finally, the
open question and future challenge of utilizing the stimulus-
response behavior of CPCs to design versatile sensors are discussed.

CONTACT Yutian Zhu ytzhu@hznu.edu.cn College of Materials, Chemistry and Chemical Engineering, Hangzhou
Normal University, No. 2318 Yuhangtang Rd., Cangqian, Yuhang District, Hangzhou 311121, China, Jinrui Huang
jrhuang@caf.ac.cn Key Laboratory of Biomass Energy and Material, Nanjing 210042, Jiangsu Province, China, Zhanhu
Guo zguo10@utk.edu Integrated Composites Laboratory (ICL), Department of Chemical & Biomolecular
Engineering, University of Tennessee, Knoxville, TN 37996, USA.
ß 2020 Taylor & Francis Group, LLC
158 J. CHEN ET AL.

1. Introduction
Incorporation of functional nanoparticles into polymer matrix can effectively combine
the respective advantages of each individual component, thus creating the polymer com-
posites with not only outstanding processability but also versatile functionalities.1–20
Filling the electrically conductive nanoparticles, such as carbonaceous or metallic par-
ticles, into the insulated polymer matrix, the polymer material can achieve an insulator/
conductor transition when a percolated conductive pathway inside the polymer matrix
is formed to generate the conductive polymer composites (CPCs).21–27 According to
their electrical resistivity, the CPCs can be used as the antistatic or anticorrosive materi-
als,28–30 electromagnetic interference (EMI) shielding materials,31–35 and conduc-
tors.22,23,36,37 Interestingly, it was found that the conductive network in specific polymer
matrix had a response to the external environments, such as mechanical stress, tempera-
ture, solvent, vapor, and so on.38,39 Since the charge transportation highly depends on
the conductive network, tiny change in the external environments may result in the
change of the electrically conductive network, thus causing a remarkable change in the
electrical resistance of the CPCs. Therefore, this “stimuli-responsive” behavior of the
CPCs can be utilized to design highly sensitive sensors to detect or monitor the changes
in the external environments.
Although great progress has been made in utilizing the “stimuli-responsive” behavior
of the CPCs to design sensors, there is still no systematical review paper to summarize
the recent progress in this filed. Up to now, most of the review papers about the CPCs
mainly focused on their electrical properties and applications.21–23 In addition, there are
some recent review articles that only reviewed the strain sensing behavior of the flexible
CPCs.40,41 In this article, we give an overview of the category, fabrication, properties of
the “stimuli-responsive” CPCs and their applications in sensors. According to the
“stimuli-responsive” behaviors, the CPC sensors can be divided into strain/pressure sen-
sor, liquid/vapor sensor, and temperature sensor. The mechanisms and properties of the
CPC-based sensors are discussed in this review.

2. Preparation of CPCs
For the insulating polymer-based CPCs, various electrically conductive fillers such as
carbon black (CB), carbon nanotube (CNT), carbon fiber (CF), graphene (GE), graphite,
and metal particles, are dispersed into the polymer matrix to achieve an insulator/con-
ductor transition for the composites.42–47 Based on the categories of the polymers, the
CPCs can be divided into thermoplastic plastic, thermoplastic elastomer (TPE), thermo-
setting plastic, and rubber based-conductive composites. Among these CPCs, thermo-
plastic plastic and TPE based-conductive composites have been the biggest category of
insulating polymer-based CPCs because their production techniques are relatively sim-
ple and capable of large-scale production. Moreover, an attractive advantage of the
thermoplastic plastic and TPE based-conductive composites is that they are reworkable,
which can easily be recycled. Thermosetting plastic based-conductive composites have
high thermal stability at high temperatures, thus can be applied in the high temperature
required fields, for example, encapsulating electronics. However, the recycle of this type
of composites is still a challenge. TPE and rubber based-conductive composites possess
POLYMER REVIEWS 159

the high stretchability, and have potential applications in strain and pressure sensors.
The preparation method of CPCs is critical to the filler dispersion as well as the con-
structed conductive pathways in the polymer matrix, and thus influences the electrical
and mechanical properties of the composites.
Up to now, three types of preparation methods for CPCs are developed, including
melt compounding,48–53 in situ polymerization54–58 and solution mixing.59–61 Melt com-
pounding is a technique to directly disperse the conductive fillers into the polymer melt
via the shearing force, which agrees well with the requirement in current industries. In
situ polymerization is a method that can form the polymer chains and simultaneously
well disperse the fillers at a molecular scale. This usually gives good filler dispersion in
the polymer matrix. Solution mixing is another alternative method, which can achieve a
homogeneous dispersion of polymer chains and conductive fillers in a solvent. Uniform
dispersion of fillers in polymer matrix will be achieved after the removal of the organic
solvent. Among these preparation methods of CPCs, melt compounding technique is
the most commonly used fabrication method for the thermoplastic plastic, TPE, and
rubber based-conductive composites because this method is relatively simple and easy.
Moreover, this method is also widely used in industry because CPCs can be manufac-
tured in a continuous manner by this method. However, the filler breakage and its dis-
persion states need be considered to optimize the properties of the composites. In situ
polymerization can fabricate CPCs with a better filler dispersion and a stronger inter-
facial strength between the filler and the polymer matrix. However, this method is only
suitable for some limited polymer systems. Compared to melt compounding and in situ
polymerization, solution mixing is a time-consuming and high-cost process, which
limits its applications in industry.
In addition, the thermal or UV induced crosslinking is a critical procedure during
the preparation of the thermosetting plastic and rubber-based CPCs.62–81 Epoxy has
been often used as the polymer matrix for the preparation of CPCs via the thermal or
UV induced crosslinking method.62–65,70–75 For example, we recently fabricated a highly
tough, strong, and stiff MWCNTs/epoxy conductive composite with an ultralow perco-
lation threshold by shear-intensive mechanical stirring and then curing the MWCNT/
epoxy pre-mixture at high temperature.70 The remarkable electrical and mechanical
properties of the resulting composites are attributed to the specific three-dimensional
cellular structures inside the composites. Furthermore, these different preparation meth-
ods may be sometimes required to design some complex CPCs. For example, Xiang
et al.82 prepared a high performance strain sensor of CNTs/thermoplastic polyurethane
(TPU) composites by solution mixing and melt extruding into filaments using a single-
screw extruder.

3. Structures and properties of CPCs


The electrical and mechanical properties of the CPCs highly depend on the conductive
network in the polymer matrix. Usually, as the filler content is increased, the electrical
conductivity is increased since the continuity of the conductive network is gradually
improved, thus achieving an insulator-to-conductor transition of the CPCs. However,
the mechanical properties of the CPCs are usually deteriorated as the inorganic filler
160 J. CHEN ET AL.

Figure 1. Schematic illustrations of (a) segregated phase structure and (b) double percola-
tion structure.

content is increased. To quantitatively describe the conductive network, the percolation


theory is developed. Based on the percolation theory, there is a minimum loading of
the conductive filler, i.e., the electrical percolation threshold, to build up an electrically
percolated conductive network in the polymer matrix. According to the percolation the-
ory, the electrical conductivity of the composite can be described by a scaling law of
filler content as the following equation:
raðp  pc Þt for p > pc (1)
where r is the electrical conductivity of the composites, p is the content of filler, pc is
the percolation threshold, and t is a critical exponent that describes the dimensionality
of the conductive network. Since the percolation threshold is highly relative to the
electrical and mechanical properties, intensive attentions have been paid to design the
CPCs with low percolation thresholds. One feasible strategy is to start with the conduct-
ive filler, such as improving the quality of the fillers as well as their dispersion level in
the polymer matrix. For example, it was reported that single-walled carbon nanotubes
(SWCNTs) possess higher length-to-diameter ratios than multi-walled carbon nanotubes
(MWCNTs) and thus can be used to fabricate the CPCs with a lower percolation
threshold.83 Another effective strategy is to utilize the phase morphology of the CPCs
to improve the electrical property, which includes the segregated structure22,23,84–101
or double percolation structure.102–130 In the CPCs containing the segregated phase
structure, all the conductive particles are dispersed between the polymeric matrix
particles, thus resulting in the formation of a percolated conductive pathway among the
polymeric particles, as illustrated in Figure 1a. Since the polymeric matrix particles play
a role of excluded volume, a small amount of conductive fillers can achieve electrical
percolation. For the CPCs containing the double percolation structure (Figure 1b),
the conductive fillers are selectively dispersed in one phase or at the interface between
a co-continuous polymer blend to build up the conductive network, which can
significantly lower the percolation threshold of the composites.
Beside the electrical property, the mechanical property of CPCs is another critical aspect
for practical applications of CPCs. In general, doping a high loading of conductive fillers
into polymer matrix significantly changes the mechanical properties of the composites
because of the interaction between the conductive filler and polymer matrix. Normally,
POLYMER REVIEWS 161

Figure 2. Schematic diagrams of resistive-type (top images) and capacitive-type strain/pressure


sensors, respectively.

the strength and stiffness are usually increased compared to the unreinforced polymers.
However, the elongation at break toughness is normally sacrificed with increasing the
nanoparticle content, which limits the application of the materials. To improve the mech-
anical properties, the chemical modification of fillers is often utilized to improve the inter-
actional interaction between the filler and polymer.60 But paradoxically, chemical
modification of conductive fillers will cause the decline of intrinsic conductivity of filler
simultaneously. Therefore, it needs to find a balance between the electrical property and
mechanical property. Moreover, although the double percolation structure and segregated
structure allow us to build the conductive network at extremely low filler loadings, they
also harm the mechanical property especially in tensile properties because of the weak
interfacial interactions between the two immiscible polymer domains.5,22 To alleviate this
problem, the addition of the compatibilizer into an immiscible polymer blend can enhance
the interfacial interactions between the two polymers, thus improve the mechanical prop-
erties of the composites. However, one should be careful not to destroy the phase morph-
ology of the CPCs with the addition of compatibilizer. In our previous study, we
presented an example to balance the electrical properties and mechanical properties of the
162 J. CHEN ET AL.

Figure 3. (a) Schematic diagram of the strain sensing mechanisms of the composites filled with dif-
ferent dimensional conductive fillers. (a1) CB particles; (a2) CNT particles; (a3) CB-CNT hybrid particles.
(b) The change of resistance (DR/R0) versus time during the stretching-releasing cycles. (b1) 2 phr
CNTs/IR composite; (b2) 10 phr CB/IR composite; (b3) 3 phr CNTs/CB(1/1)/IR composite; (b4) 4phr
CNTs/CB(1/2)/IR composite. Reprinted with permission from Ref. 133. Copyright (2018) Elsevier Ltd.

Figure 4. Schematic of the processing procedure for preparing the PDMS/MWCNT composite
with randomly distributed MWCNTs (a) and the elastomer with a self-segregated structure (b). (c)
Electrical conductivity of the PDMS/MWCNT composites with random distribution of MWCNTs and
self-segregated structure as a function of the MWCNT loadings. (d)DR/R0 as a function of multiple
stretching and releasing cycles with 30% strain for the self-segregated samples with 0.12 vol%
MWCNTs and the conventional samples with 0.6 vol% MWCNTs. Reprinted with permission from Ref.
230. Copyright (2017) The Royal Society of Chemistry.

CPCs containing a double percolation structure.5 In that work, styrene-ethylene/butylene-


styrene (SEBS) triblock copolymer was used as the compatibilizer to improve the mechan-
ical properties of CB/PS/PP composites with the double percolation structure.
POLYMER REVIEWS 163

Figure 5. (a) Schematic illustration of the three layer-structures of the strain sensor of PU-PEDOT:PSS/
SWCNT/PU-PEDOT:PSS on a PDMS substrate; (b) Schematic illustration of strain sensors attached to
the face to sense the emotions and daily activities; (c) Time-dependent responses of DR/R0 for the
sensor attached to different places of the face to sense the emotions. For the measurements of c-1
and c-3, the sensor is attached on the forehead. For the measurement of c-2 and c-4, the sensor
is attached on the skin near the mouth. c-1 and c-2 are the electrical signals when the subject
is laughing, while c-3 and c-4 are the electrical signals when the subject is crying. Reprinted with
permission from Ref. 156. Copyright (2015) American Chemical Society.

4. Applications of stimuli-responsive conductive polymer composites


CPCs have important applications in many fields, such as antistatic materials, electromag-
netic interference (EMI) shielding, and conductors.131 It is known that the electrical
164 J. CHEN ET AL.

Figure 6. (a) Schematic illustration of the fabrication of the stretchable and transparent PDMS/CNTs/
PDMS strain sensor with a sandwich-like structure. (b) and (c) are the SEM and TEM images of the
PDMS/CNTs/PDMS strain sensor, which clearly visualize the ultrathin CNT layer sandwiched between
the PDMS films. (d) Electrical resistance of the sandwich-like PDMS/CNTs/PDMS sensors with different
CNT layer thicknesses. Oval images are the photos of the different sensors on a color pattern. The
inserted photo on top right corner shows that the sensor can light a LED lamp. (e) Response of DR/
R0 with various cyclic strains for the PDMS/CNTs/PDMS strain sensor (CNT layer density: 0.16 mg/cm2)
at a strain rate of 10 mm/min. (f) Dependence of DR/R0 on the cycle number for 100 stretching-releas-
ing cycles at a maximum strain of 100% and a rate of 50 mm/min. Reprinted with permission from
Ref. 232. Copyright (2020) Elsevier.

conductivity of the CPCs highly depends on the conductive networks. Therefore, slight
change in the conductive networks causes a significant change in the conductivity. Utilizing
the “stimuli-responsive” behavior of the conductive networks, the CPCs can be used as
smart sensors to monitor or detect the changes in the external fields, such as strain sen-
sor,132–157 pressure sensor,158–162 vapor sensor,163–184 liquid sensor,185–191 and temperature
sensor.192–210 Compared to traditional sensors made of metal, semiconductor or ceramic
materials, the CPCs-based sensors exhibit various advantages, such as light weight, corrosion
resistance, easy processing, and adjustable sensing performance. Therefore, the CPCs-based
smart sensors become an emerging but fast-growing field in polymer materials, which has
attracted intensive attentions from both academy and industry.

4.1. Strain/pressure sensor


There is a growing market demand for soft tactile sensors (strain/pressure sensor) to
detect the loaded strains or pressures on the materials. This tactile sensor has important
applications in human motion detection, health-monitoring, human-machine interfaces,
soft robotics, structural health monitoring, etc.150,211–221 On the other hand, high per-
formance tactile sensors can be found everywhere in nature. For example, human skin
POLYMER REVIEWS 165

Figure 7. (a) Schematic illustration of the fabrication of stretchable capacitive-type sensor with an
array of AgNWs. (b) The change of capacitance (DC/C0) versus strain during the stretching and releas-
ing test. (c) The change of capacitance (DC/C0) versus strain for the 1st stretching and the 100st
stretching. (d) The change of capacitance (DC/C0) versus normal pressure. (e) Real-time Response of
the sensor to water droplets with weight of 0.06 g each. Reprinted with permission from Ref. 224.
Copyright (2014) The Royal Society of Chemistry.

is the most sensitive tactile sensor that can perceive very slight pressure or shear and
send the monitored data to the brain. The strain or pressure sensors could roughly be
divided into two categories, i.e., resistive-type and capacitive-type sensors, as illustrated
in Figure 2. For the resistive-type sensors, its resistance depends on its inner conductive
networks. Therefore, the change of resistance can be used to detect the applied strain or
pressure on the sensors. Capacitive-type sensors are composed of a dielectric layer and
two stretchable electrodes,222–225 as illustrated in Figure 2. In general, the capacitance
is related to the distance between these two stretchable electrodes, which can also be
utilized to monitor the deformation of CPCs.
166 J. CHEN ET AL.

Figure 8. (a) Schematic illustration of the fabrication of the capacitive-type sensor with uniform
microtower patterns on the bottom electrode. (b) Schematic illustration of the sensing mechanisms of
micropatterned sensor and nonpatterned sensor. (c) Dependence of the capacitance change on the
applied pressure for micropatterned sensor and the nonpatterned sensor, respectively. (d) Real-time
response to the pressure of 0.5, 2, and 10 kPa for the two types of sensors. Reprinted with permission
from Ref. 159. Copyright (2018) WILEY-VCH.

It is reported that the resistance of the CPCs depended on the external strain because
of the piezoresistivity effect. Piezoresistivity is a phenomenon to describe the resistance
variation of the conductive material upon applying a mechanical force.37,226 Therefore,
CPCs can thus be utilized to design the resistive-type strain sensors. To quantitatively
characterize the sensitivity of the resistive-type strain sensors, a parameter of gauge
factor (GF) is defined as GF ¼ (DR/R0)/e, whereDR is the change in the resistance with
strain, R0 is the initial resistance, and e is the applied strain. In general, the high GF
means the high sensitivity of the strain sensor. Based on the inner microstructures of
conductive nanoparticles, there are two types of resistive-type strain sensors. One is the
filling-type strain sensor, in which the conductive fillers are randomly filled into the
polymer matrix. The filling-type strain sensors have many advantages such as good
processability, high mechanical properties, low cost, and ease for the large-scale continu-
ous production. Another one is the sandwich-like strain sensor, in which the conductive
particles form a conductive layer sandwiched between the two polymer films. In this
section, we will discuss these two types of CPCs-based strain sensors in turn.
One common route to obtain filling-type piezoresistive strain sensor is to fill various
conductive nanofillers, including sliver nanowires, CNTs, GE, and CB, into the elasto-
meric matrix. For instance, we recently systematically investigated the strain sensing
behavior of filling-type strain sensors composed of different dimensional conductive
POLYMER REVIEWS 167

nanoparticles and the stretchable isoprene rubber (IR).133 It was found that sensing
mechanisms of filling-type strain sensors highly depended on the conductive particle
dimensionality (Figure 3). For example, 0-dimensional carbon black (CB) particles tend
to build up a conductive network with a large number of weak CB-CB contacts, as illus-
trated in Figure 3a1. As a result, the CB-networks are easily broken under stretch
(Figure 3a1), causing a sharp increase in the resistance of the composite (Figure 3b2).
Moreover, these CB particles also easily return to their original positions to fast recover
the networks under releasing. It is observed that the change of the resistance (DR/R0) of
the CB/IR composites can return to 0 under releasing, indicating good recoverability of
the sensor. On the contrary, 1-dimensional CNTs with a large aspect ratio tend to
entangle with each other to form the relatively firm networks (Figure 3a2), which are
less sensitive to the strain stimulus. Therefore, CNTs/IR composites exhibit a weaker
strain sensitivity than CB/IR composites (Figure 3b1). Furthermore, some irreversible
damage of CNT-network is caused during the stretch, which could never be recovered
under the following releasing. This causes the continuous increase of the maximum and
minimum values of DR/R0 with increasing the stretching–releasing cycle number, indi-
cating the poor recoverability of the CNTs filler composite. More interesting, it was
found that the hybrid CNT-CB network can well compensate the disadvantages of CB/
IR and CNTs/IR composites, which possess a low percolation threshold, high sensitivity
and good recoverability (Figure 3a3, b3, and b4).
Instead of directly filling conductive fillers into stretchable polymer matrix, some
complex hierarchical structures, including double percolation structures154,227–229 and
segregated structures,230,231 were also used to improve the sensing properties of the
filling-type strain sensors. For instance, Wang et al.230 fabricated the PDMS/MWCNT
composites with randomly distributed MWCNTs (conventional samples, Figure 4a) and
the composites with a self-segregated structure (self-segregated samples, Figure 4b),
respectively. It is found that the electrical conductivities of the self-segregated samples
are much higher than the corresponding conventional samples because of the volume
exclusion effect of the cross-linked PDMS particles (Figure 4c). Moreover, it is also
observed that the self-segregated sample exhibit a higher piezoresistive sensitivity
than the conventional samples (Figure 4d), which is ascribed to the quicker damage-
reconstruction of the conductive pathways during the stretching-releasing tests.
Another route to fabricate the resistive-type strain sensor is to sandwich a conductive
layer between two stretchable polymeric layers. Compared to the filling-type strain
sensor, the preparation of this type of strain sensor is time-consuming and hard to
achieve mass production. However, the advantage of the sandwich-like strain sensor is
the high strain sensitivity, which could detect the feeble strain stimuli. For example,
Roh et al.156 fabricated a stretchable, transparent, ultrasensitive, and patchable
strain sensor consisting of the three-layer stacked nanohybrid structure of polyurethane
(PU)-poly(3,4-ethylenedioxythiophene) polystyrenesulfonate (PEDOT:PSS)/SWCNT/
PU-PEDOT:PSS on a PDMS substrate, as illustrated in Figure 5a. The obtained
strain sensor possesses a high stretchability (The applied strain can reach 100%.), optical
transparency (62%), and high sensitivity (GF ¼ 62). This strain sensor can be used to
detect the emotional expressions, such as laughing and crying by attaching it to the face
of the human (Figure 5b). Different emotional expressions can be detected via the time-
168 J. CHEN ET AL.

dependent DR/R0 responses, as shown in Figure 5c. Very recently, we developed a sim-
ple spray deposition and transfer method to sandwich an ultrathin and uniform CNT
layer (ca. tens of nanometers) between two PDMS films to construct the stretchable and
transparent strain sensor, as shown in Figure 6a–c.232 The secret of successfully generat-
ing the ultrathin conductive layer is to uniformly disperse the CNTs in isopropanol to
form a dilute solution (0.04 wt%), which is then uniformly sprayed on the surface of
polytetrafluoroethylene to form the ultrathin CNT film. This sandwich-like strain sensor
exhibits good optical transparency (transmittance: 62.3% at 550 nm) and relatively low
electrical resistance (5.36 MX), which can light a LED lamp (Figure 6d). As the sensor
is stretched from 0 to 20, 40, 60, 80, and 100% strain step, DR/R0 generally increases
with time, indicating good responses of the sensor to the strain (Figure 6e). Moreover,
this transparent sandwich-like strain sensor also shows outstanding repeatability and
stability for repeated 100 stretching-releasing cycles (Figure 6f).
Capacitive-type sensor normally consists of two parallel plate electrodes that are sand-
wiched in a dielectric layer (Figure 2). Basically, the capacitance (c) of the sensor can be
determined according to Equation (2):
C / Ae=d (2)
where A is the area of two electrodes, d is the distance between the two plate electrodes,
and e is the permittivity of dielectric layer. When the external force is applied on the
capacitive-type sensor, the changes in A, d, or e cause the change in capacitance, which
endows the sensor the capability of detecting the strain or pressure. Compared to the
resistive-type sensors, capacitive-type sensors possess the advantages of good linearity,
reversibility and fast response time. For instance, Yao et al.224 fabricated a stretchable
capacitive-type sensor with an array of silver nanowire (AgNW) conductors, as illus-
trated in Figure 7a. This capacitive-type sensor is able to detect both the strain and
pressure, as shown in Figure 7b–e. More interestingly, it is observed that this sensor
exhibits good linearity and reversibility during stretching and releasing (Figure 7b).
Moreover, the sensor also exhibits excellent stability even after 100 stretching-releasing
cycles (Figure 7c). Furthermore, this sensor also has the pressure sensing capability,
which can realize the real-time pressure monitoring. For example, the dropping of small
water droplets (0.06 g/droplet) is on-line monitored, as shown in Figure 7e.
To further increase the sensitivity of the capacitive-type sensor, a strategy of design-
ing the micropatterned structures on the dielectric layer or electrode was developed to
fabricate the high-performance capacitive sensors.159,233–235 For example, Bao and cow-
orkers234 designed the ordered pyramidal arrays on the flexible dielectric layer to build
highly sensitive capacitive-type sensor. Recently, Wan et al.159 created uniform micro-
tower patterns on the PDMS film by utilizing the lotus leaf as the molding template.
This patterned PDMS film was then coated by ultrathin AgNWs to serve as the bottom
electrode to assemble the capacitive-type sensor, as illustrated in Figure 8a. Clearly, the
change in the height of microtowers should be very sensitive to the external pressure
(Figure 8b), and thus resulted in the remarkable change of capacitance according to the
principle described in Equation 2. Figure 8c presents the variation of capacitance change
as a function of applied pressure for the micropatterned sensor and the nonpatterned
sensor, respectively. It is clear that the micropatterned sensor exhibits a much higher
sensitivity to the applied pressure than the nonpatterned sensor. Moreover, it is also
POLYMER REVIEWS 169

observed that the micropatterned sensor can present a stable real-time response with
quick loading/unloading pressures as low as 0.2 kPa, which indicates that this sensor
possesses both high sensitivity and excellent repeatability (Figure 8d).

4.2. Vapor/liquid sensor


In addition, it was found that the host polymer of the CPCs could be swollen in the
presence of organic vapors or liquids. Clearly, the connectivity between the conductive
fillers will be decreased as the polymer matrix swells, thus causes an increase of sample
resistance. On the basis of this feature, the CPCs can also be used to design the vapor
sensors and liquid sensors to detect the change of environmental conditions.
To quantitatively describe the sensitivity of the vapor sensor, the resistance relative
amplitude or relative differential resistance response (AR) is defined by Equation 3:
ðR  R0 Þ
AR ¼ (3)
R0
where R is the resistance of CPCs exposed to analytic vapors, R0 is the initial resistance
of the CPCs exposed to a nitrogen flow. There are many factors that can affect the elec-
trical responsivity of CPC sensors against organic solvents, such as the electrical proper-
ties of the conductive fillers, dispersion degree of conductive fillers in the polymer
matrix, and the physical properties of the polymer matrix. For instance, Dong et al.236
reported that the CB filled polymethacrylate composites exhibited a lower percolation
threshold and a slower response rate when the glass transition temperature (Tg) of the
host polymer was higher. This is because that the polymer with a higher Tg generally
possesses a higher viscosity, which suppresses the diffusion of solvent molecules into
the polymer matrix and increases the response time of the composites against the
organic vapors.
Moreover, the hierarchical structure of the CPCs is another critical factor that domi-
nates the vapor sensing behavior of the CPCs. For example, Castro et al.170 utilized the
layer by layer assembly technique to prepare the vapor sensitive CPCs composed of an
array of CNTs. Taking benefit from the exceptional electrical properties of CNTs and
highly specific 3D surface structure, the CPCs-based sensor shows an excellent sensitiv-
ity and selectivity to nine analytic solvent vapors, including isopropanol, tetrahydro-
furan, dichloromethane, n-heptane, cyclohexane, methanol, ethanol, water and toluene.
It was found that both good selectivity and sensitivity were caused by the intensity dif-
ference between their responses to different vapors for the poly(lactic acid) (PLA)
based- and PMMA based-sensors, as shown in Figure 9a and b. Clearly, different signal
amplitudes are observed when the sensors are exposed to different solvent vapors.
Moreover, the initial slope and general shape for the chemo-resistive responses are also
different, which can well identify the solvent types.
Recently, Kennedy et al.237 prepared the highly-conductive MWCNTs/poly(vinylidene
fluoride) (PVDF) composites, which could be used as filament feedstock for 3D print
sensors with different geometries (Figure 10a). Interestingly, these printed MWCNTs/
PVDF sensors could be used to detect vapors by the changes in the resistance during
exposure (Figure 10b). It was found that the printed sensors could detect volatile
organic compound (VOC) vapors within 20 seconds and remain the sensing properties
170 J. CHEN ET AL.

Figure 9. Chemo-resistive response of PLA–1%CNT (a) and PMMA–1%CNT (b) sensors for cyclic expo-
sures to nine different types of vapors. Reprinted with permission from Ref. 170. Copyright
(2011) Elsevier.

Figure 10. (a) Preparation of MWCNTs/PVDF composites for 3D printing; (b) the change of R/R0 over
time when the printed sensor is exposed to several common volatile organic compounds. Reprinted
with permission from Ref. 237. Copyright (2017) Royal Society of Chemistry.

over at least 25 solvent–vacuum cycles. The high sensitivity of the printed MWCNTs/
PVDF sensors is attributed to the vapors’ ability to swell the PVDF matrix, thus result-
ing in a strong response in resistance. In addition, Qiang et al.174 fabricated the 3D
vapor grown carbon nanofiber (VGCF)/clay/silicon rubber (SR) (VCS) coated polyur-
ethane (PU) foam composites (VCS@PU) by a facile dip-coating method, which exhib-
ited a rapid and stable vapor sensing response (Figure 11). Figure 11a shows that the
VCS@PU foams possess an excellent cyclic performance in n-hexane vapor with a high
responsivity. It is observed from the shape of DR/R0-t curve that the minimum and
maximum DR/R0 values of the composite are almost the same in each cycle of saturated
organic vapor and air (150 s each cycle). Moreover, the responsivity is increased rapidly
as the vapor concentration increases, which can be used to identify the concentration of
organic solvent in the atmosphere. The high sensitivity of the VCS@PU composites to
the organic vapor is attributed to the isolation effect of insulating clay fillers,
as illustrated in Figure 11c. Clearly, the isolation of insulating clays effectively results
POLYMER REVIEWS 171

Figure 11. (a) Electrical response of VCS@PU foams for cyclic exposure to saturated n-hexane vapor;
(b) Electrical response of VCS@PU foam composites upon exposure to n-hexane vapor under different
concentrations (ranged from 1 to 5 ppm in sequence); (c) Schematic diagram of conductive network
evolution of VCS@PU foam during the detection of the organic vapor process. Reprinted with
permission from Ref. 174. Copyright (2019) Elsevier.

in a remarkable damage of VGCF network even for the slight swelling of composites,
thus causes an increase in the electrical resistance of the composites.
When CPCs are immersed into liquid, solvent sorption leads to the swelling of
polymer matrix, thus increasing the distance between the neighboring conductive fillers
and reducing the tunneling current. Therefore, the resistivity of the CPCs is increased
as the swelling occurs. Based on this mechanism, the response of the CPCs after
immersing them into liquid can be utilized to sense liquid. For instance, Rentenberger
et al.190 produced the poly(e-caprolactone) (PCL) þ 4% MWCNTs/PLA ¼ 50/50 wt%
blend multifilament fibers and melt-spin them into a textile sensor for liquids. The
textile sensors were tested for cyclic liquid sensing in ethyl acetate, acetone and
ethanol respectively, as shown in Figure 12a. Clearly, PCL has a solubility parameter
of 19.2 MPa0.5, which is close to that of acetone (19.9 MPa0.5) and ethyl acetate
(18.2 MPa0.5).238 Therefore, these two solvents can be the moderate solvents for PCL/
4 wt%MWCNTs/PLA fibers, which only swell both polymers, without destructing them.
However, ethanol has a solubility parameter of 26.5 MPa0.5,238 which is the poor solvent
for the textile sensors. As a result, polymer swelling in ethanol is negligible. Figure 12b
shows the relative resistance as a function of time for wettings at t ¼ 100 s, 5400s,
10,700 s, 16,000 s and 21,300 s. The observed sharp increase of the relative resistance
change when the textile is wetted by ethyl acetate indicates the high sensitivity
172 J. CHEN ET AL.

Figure 12. (a) Textile sensor fixed between two Cu electrodes and placed on a Teflon substrate
with the solvent line drawn initially; (b) Relative resistance change on textile sample (PCL þ 4 wt%
MWCNT/PLA ¼ 50/50 wt%) upon 2 mL ethyl acetate for 5 times of wettings at t ¼ 100, 5400, 10,700,
16,000 and 21,300 s. Reprinted with permission from Ref. 190. Copyright (2011) Elsevier.

of the liquid sensor. Thereafter, the relative resistance change begins to drop
immediately during the drying process.

4.3. Temperature sensor


When the CPCs are heated, the resistivity increase will generally be observed because of
the expansion of the host polymer, which will destruct the conductive network. This
phenomenon is defined as the positive temperature coefficient (PTC) effect. For
instance, Rybak et al.195 filled the conductive silver nanoparticles (Ag) into the high
density polyethylene (HDPE), polybutylene terephthalate (PBT) and poly(m-xylene
adipamide) (MXD6), which generated the CPCs with a high intensity of PTC effect, i.e.,
a sharp increase in the resistivity. Figure 13 shows the temperature dependence of
resistivity for the composites with Ag filled in different polymers. The HDPE-20Ag
composites exhibit the highest intensity of the PTC effect. It is observed that the
resistivity increases by more than 10 orders of the magnitude, which can be the
promising materials for temperature sensors. However, no obvious PTC effect
is observed for the MXD6–20Ag composites, which is attributed to the low thermal
expansion of the MXD6 matrix. Moreover, Jeon et al.239 found that the PTC effect was
largely depended on the thermal properties of the host polymers. For the conductive
particles filled amorphous polymers, the resulting composites normally exhibited a small
PTC effect (1 to 3 orders magnitude change in resistivity), which was attributed to
the gradual expanding of amorphous polymer with increasing the temperature. On the
contrary, the conductive particles filled semi-crystalline polymers showed a larger PTC
effect because of the transformation of the crystalline phase to amorphous phase, thus
resulting in a large volume expansion near the melting point. For example, they filled
Ni particles into a polymer blends with both semi-crystalline polyethylene (PE) and
polyethylene oxide (PEO) to design the highly sensitive temperature sensor. More
interestingly, this sensor shows a strong PTC effect around 35 to 42  C, which is applic-
able for monitoring the temperature of human body.
POLYMER REVIEWS 173

Figure 13. Dependence of the electrical resistivity on temperature for the composites with AG filled in
HDPE, PBT, or MXD6, respectively. Reprinted with permission from Ref. 195. Copyright (2010) Elsevier.

The PTC effect is a typical temperature/resistivity phenomenon for a solid CPC.


On the other hand, a negative temperature coefficient (NTC) effect, i.e., the resistivity
is decreased upon heating, is always accompanied with the PTC effect during the
increase of the temperature from room temperature to melting point of the polymer
matrix. As the temperature is increased, most of CPCs normally exhibit the PTC
effect before the melting temperature (Tm) of the crystalline polymer matrix or the Tg
of the amorphous polymer matrix, and then show the NTC effect with further
increase of the temperature. For example, Liu et al.240 melting-mixed MWCNTs into
HDPE to obtain the MWCNTs/HDPE composites. It was found that the composites
showed both PTC and NTC phenomena near the melting temperature of HDPE. This
non-monotonic resistance-temperature dependence of CPCs makes the materials hard
to be practically applied as the temperature sensors. As a result, it needs a precise
design in the hierarchical structures of CPCs to achieve the monotonic resistance-tem-
perature dependence (PTC or NTC only). Generally, monotonic NTC composites are
rare and difficult to be obtained because the thermal expansion of host polymer with
increasing the temperature is normally an unavoidable phenomenon. Recently, we fab-
ricated the foamed carbon black/chlorinated poly(propylene carbonate) (CB/CPPC)
composites (Figure 14a,b) with both high sensitivity and outstanding reproducibility
of the NTC effect during the increase of the temperature.210 It was found that this
crosslinked CB/CPPC foam can detect the change of temperature, as shown in Figure
14c–e. The NTC effect of the foam is attributed to the gas expansion-induced thin-
ning of cell walls upon heating, which decreases the distance between the CB particles.
As a result, the electrical resistance of the foam is decreased as the temperature
increases, as illustrated in Figure 14c. As the temperature is decreased, the shrinkage
174 J. CHEN ET AL.

Figure 14. (a) and (b) The morphologies of CB/CPPC foam containing 0.20 vol% CB. (c) Schematic
illustration of microstructure changes in crosslinked CB/CPPC foams during the heating and cooling.
(d) Electrical response of the crosslinked CB/CPPC foam (0.35 vol% CB) in the heating-cooling cycles.
The temperature is ranged from room temperature to 60  C. (e) The demonstration experiment to
utilize the brightness of LED lamp connected with the crosslinked CB/CPPC foam (0.35 vol% CB) to
monitor the change of temperature. Reprinted with permission from Ref. 210. Copyright (2018) Royal
Society of Chemistry.

Figure 15. (a) Schematic diagram of microstructural development for the GE/PA6/UHMWPE compo-
sites during one heating–cooling cycle. (b) Volume resistivity as a function of time for the 0.4 vol%
GE/PA6/UHMWPE composites. Reprinted with permission from Ref. 196. Copyright (2017) Royal
Society of Chemistry.

of gas causes the recovery of the cell walls and the restoration of the conductive net-
works. Figure 14d shows that the crosslinked CB/CPPC foams exhibit both outstand-
ing sensitivity and reproducibility of the resistivity-temperature behavior in the
heating-cooling cycles, which can serve as the temperature sensor as shown in the
demonstration experiment in Figure 14e.
POLYMER REVIEWS 175

Figure 16. (a) Schematic illustration of the preparation of multifunctional temperature and humidity
sensor, as well as its mechanisms of temperature-sensing and humidity-sensing. (b) The change of
resistance of TSP versus temperature. (c) Capacitance of HSP versus humidity. Reprinted with permis-
sion from Ref. 207. Copyright (2017) American Chemical Society.

Negative temperature coefficient (NTC) effect was also observed for the CPCs other than
foam composites. For example, Guo and coworkers196 introduced graphene into a PA6/
ultrahigh molecular weight polyethylene (UHMWPE) binary blend to form a unique segre-
gated and double-percolated segregated structure, i.e., graphene nanosheets were selectively
distributed in PA6 between the isolated UHMWPE particles. Upon heating process, the
expanded PA6 polymer chains yield a tension and the deformation, which can reduce the
gap at contact junctions and increase the contact between the neighbor GE nanosheets, as
illustrated in Figure 15a. As a result, the resistivity of composite is decreased, which exhibits
a classic NTC effect (Figure 15b). In the cooling process, the polymer chains shrink and the
deformation gradually recovers as the temperature decreases, and the conductive network
restores to its original state (Figure 15a). Therefore, the resistivity of the composite will also
restore to its original value, as shown in Figure 15b.

4.4. Multifunctional sensor


For most previous works about the responsive CPCs, they can only monitor one sin-
gle stimulus (such as strain, pressure, temperature, vapor, or liquid) at a time.
However, it is known that human skin is an outstanding multifunctional sensor in
nature, which can perceive various stimuli, such as strain, pressure, shear, tempera-
ture, and humidity simultaneously. For making up the drawback of artificial sensors
based on CPCs, some recent researches try to develop multifunctional sensors that
can detect multiple kinds of stimuli simultaneously or separately.207,208,241 To achieve
different functions in one unit, multiple-sensing parts should be assembled inside the
sensor. For example, Zhao et al.207 stacked the temperature-sensing layer and the
humidity-sensing layer within the PDMS (Figure 16a), to fabricate flexible and
176 J. CHEN ET AL.

Figure 17. (a) DR/R0 of the sensor (0.1 g CNT/0.3 g KH550/PDMS composite) as a function of time to
monitor the finger motion; (b) DR/R0 of the sensor (0.1 g CNT/1.5 g KH550/PDMS composite) as a
function of time to monitor the human breathing; (c) DR/R0 of the sensors (0.1 g CNT/0.3 g KH550/
PDMS composites and 0.1 g CNT/1.5 g KH550/PDMS composites) as a function of temperature; (d) DR/
R0 of the sensor (0.1 g CNT/1.5 g KH550/PDMS composite) as a function of time in response to n-hexane.
Reprinted with permission from Ref. 208. Copyright (2018) American Chemical Society.

bifunctional sensors. These two layers of graphene woven fabrics (GWFs) serve as the
temperature-sensing part (TSP) and the humidity-sensing part (HSP), respectively.
Since PDMS matrix has a high thermal expansion coefficient, the PDMS layer expands
significantly as the temperature increases. As a result, the GWF embedded in the
PDMS follows the deformation of PDMS, thus causing the increase of resistance of
GWF (i.e., PTC effect). It was found that the resistance of GWF is significantly
increased with rising the temperature when GWF is attached to the PDMS matrix
(Figure 16b), which confirms the mechanism of the temperature-sensing behavior of
the GWFs/PDMS composites. On the other hand, HSP contains the humidity-sensitive
material, i.e., cellulose acetate butyrate (CAB), which can absorb moisture to increase
the capacitance (Figure 16c).
Recently, Deng and Fu coworkers208 designed the multifunctional sensors composed
of PDMS substrate and conductive CNT/3-aminopropyltriethoxysilane (KH550) layer.
Because of the formation of well-defined cracks on the conductive layer, the sensor can
severe as the strain sensor to monitor the finger motion and breathing, as shown in
Figure 17a, b. Moreover, it is also observed that the relative resistance change of the
POLYMER REVIEWS 177

sensor (0.1 g CNT/1.5 g KH550/PDMS composite) exhibits a roughly linear response to


temperature (Figure 17c), which thus can be used to monitor the temperature. The tem-
perature sensing mechanism of the sensor is attributed to the thermal expansion of the
PDMS layer, which causes the change of the conductive network simultaneously.
Because PDMS can be swelled in some organic solvents, the increase in the volume of
PDMS matrix causes the break-up of the conductive network. As a result, the electrical
resistance of the CNT/KH550/PDMS composites is increased. Based on this mechanism,
this sensor has potential applications for solvent sensing, as illustrated in Figure 17d.

5. Conclusions and prospect


When the CPCs are stimulated by an external environment, the internal conductive net-
works in the polymer matrix will be changed, which will thus cause the change in the
output electrical signal of the CPCs. This stimuli-responsive behavior has been exploited
for the design of versatile sensors, including strain sensors, pressure sensors, liquid sen-
sors, vapor sensors, and temperature sensors to detect the tiny changes in the environ-
ment conditions. These emerging applications of CPCs attract more and more attention
in the last 10 years.
Although great potential can be seen in using the responsive CPCs to prepare various
sensors, there is still some distance to fully satisfy all the requirements of the commer-
cial sensors. More efforts are required to achieve the merits of high sensitivity, good sta-
bility and reproducibility of the CPCs-based sensors.

1. The principle of the CPCs-based sensors is based on the change of output elec-
trical signal (resistance or capacitance) caused by the external stimuli (such as
strain, pressure, temperature or vapor). However, in most cases, there is a non-
linear relationship between the electrical signal and the stimuli. Therefore, these
sensors cannot quantitatively respond to the external stimuli, which limits their
practical applications.
2. The stability and reproducibility of CPCs-based sensors are still not satisfactory,
especially for the filling-type CPCs. For this type of CPCs, the conductive par-
ticles form a random conductive network inside the polymer matrix. Clearly, the
random conductive network is hard to provide a stable and repeatable response
to the outside stimuli.
3. Fast response to the stimuli is still a challenge for some CPCs-based sensors. For
example, the temperature sensors based on the thermal expansion of host poly-
mer normally need at least several seconds to respond to the change in tempera-
ture. When CPCs-based temperature sensors are used in the e-skins, the slow
response time cannot give us the warning of preventing further injury.

To overcome these above challenges, it needs more precise design of the internal
architecture of the CPCs. Gazing into the future, the CPCs-based sensors are an exciting
and fairly open-ended field, encouraging more researchers to join this field to explore
novel and interesting sensors.
178 J. CHEN ET AL.

Abbreviations
CPCs conductive polymer composites
CB carbon black
CNT carbon nanotube
CF carbon fiber
GE graphene
PP polypropylene
PA6 polyamide 6
SWCNTs single-walled carbon nanotubes
MWCNTs multiwalled carbon nanotubes
SEBS styrene-ethylene/butylene-styrene
PDMS polydimethylsiloxane
GF gauge factor
PU polyurethane
PEDOT PSS poly(3,4-ethylenedioxythiophene) polystyrenesulfonate
OH-SB hydroxyl-poly-(styrene-block-butadiene-block-styrene)
PLA poly(lactic acid)
PCL poly(e-caprolactone)
VGCF vapor grown carbon nanofiber
PVDF poly(vinylidene fluoride)
SR silicon rubber
PTC positive temperature coefficient
Ag silver nanoparticles
NTC negative temperature coefficient
GWFs graphene woven fabrics
Tm melting temperature
Tg glass transition temperature
TSP temperature-sensing part
UHMWPE ultrahigh molecular weight polyethylene
HDPE high density polyethylene
AgNW silver nanowire

Funding
This work was financially supported by the Zhejiang National Science Fund for Distinguished
Young Scholars (LR20E030003), National Natural Science Foundation of China for General
Program (21774126), Youth Science Funds (51803234) and the start-up fund from Hangzhou
Normal University.

References
1. Balazs, A. C.; Emrick, T.; Russell, T. P. Nanoparticle Polymer Composites: Where Two
Small Worlds Meet. Science 2006, 314, 1107–1110. DOI: 10.1126/science.1130557.
2. Sarkar, B.; Alexandridis, P. Block Copolymer-Nanoparticle Composites: Structure,
Functional Properties, and Processing. Prog. Polym. Sci. 2015, 40, 33–62. DOI: 10.1016/
j.progpolymsci.2014.10.009.
3. Yan, N.; Liu, H.; Zhu, Y.; Jiang, W.; Dong, Z. Entropy-Driven Hierarchical
Nanostructures from Cooperative Self-Assembly of Gold Nanoparticles/Block Copolymers
under Three-Dimensional Confinement. Macromolecules 2015, 48, 5980–5987. DOI:
10.1021/acs.macromol.5b01219.
POLYMER REVIEWS 179

4. Zhang, L.; Chen, Y.; Li, Z.; Li, L.; Saint-Cricq, P.; Li, C.; Lin, J.; Wang, C.; Su, Z.; Zink,
J. I. Tailored Synthesis of Octopus-Type Janus Nanoparticles for Synergistic Actively-
Targeted and Chemo-Photothermal Therapy. Angew. Chem. Int. Ed. 2016, 55, 2118–2121.
DOI: 10.1002/anie.201510409.
5. Chen, J.; Cui, X.; Sui, K.; Zhu, Y.; Jiang, W. Balance the Electrical Properties and Mechanical
Properties of Carbon Black Filled Immiscible Polymer Blends with a Double Percolation
Structure. Compos. Sci. Technol. 2017, 140, 99–105. DOI: 10.1016/j.compscitech.2016.12.029.
6. Zhang, Y.; He, Y.; Yan, N.; Zhu, Y.; Hu, Y. Inorganic Nanoparticle Induced Morphological
Transition for Confined Self-Assembly of Block Copolymers within Emulsion Droplets.
J. Phys. Chem. B 2017, 121, 8417–8425. DOI: 10.1021/acs.jpcb.7b06701.
7. Liu, M.; Li, B.; Zhou, H.; Chen, C.; Liu, Y.; Liu, T. Extraordinary Rate Capability
Achieved by a 3D “Skeleton/Skin” Carbon Aerogel–Polyaniline Hybrid with Vertically
Aligned Pores. Chem. Commun. 2017, 53, 2810–2813. DOI: 10.1039/C7CC00121E.
8. Hu, J.; Lin, J.; Zhang, Y.; Lin, Z.; Qiao, Z.; Liu, Z.; Yang, W.; Liu, X.; Dong, M.; Guo, Z.
A New anti-Biofilm Strategy of Enabling Arbitrary Surfaces of Materials and Devices
with Robust Bacterial anti-Adhesion via a Spraying Modified Microsphere Method.
J. Mater. Chem. A 2019, 7, 26039–26052. DOI: 10.1039/C9TA07236E.
9. Lin, J.; Chen, X. Y.; Chen, C. Y.; Hu, J. T.; Zhou, C. L.; Cai, X. F.; Wang, W.; Zheng, C.;
Zhang, P. P.; Cheng, J.; et al. Durably Antibacterial and Bacterially Antiadhesive Cotton
Fabrics Coated by Cationic Fluorinated Polymers. ACS Appl. Mater. Interfaces 2018,
10, 6124–6136. DOI: 10.1021/acsami.7b16235.
10. Zhang, Y.; An, Y.; Wu, L.; Chen, H.; Li, Z.; Dou, H.; Murugadoss, V.; Fan, J.; Zhang, X.;
Mai, X.; Guo, Z. Metal-Free Energy Storage Systems: combining Batteries with Capacitors
Based on a Methylene Blue Functionalized Graphene Cathode. J. Mater. Chem. A 2019, 7,
19668–19675. DOI: 10.1039/C9TA06734E.
11. Qian, Y. X.; Yuan, Y. H.; Wang, H. L.; Liu, H.; Zhang, J. X.; Shi, S.; Guo, Z. H.; Wang, N.
Highly Efficient Uranium Adsorption by Salicylaldoxime/Polydopamine Graphene Oxide
Nanocomposites. J. Mater. Chem. A 2018, 6, 24676–24685. DOI: 10.1039/C8TA09486A.
12. Li, S. W.; Yang, P. P.; Liu, X. H.; Zhang, J. X.; Xie, W.; Wang, C.; Liu, C. T.; Guo, Z. H.
Graphene Oxide Based Dopamine Mussel-like Cross-Linked Polyethylene Imine
Nanocomposite Coating with Enhanced Hexavalent Uranium Adsorption. J. Mater. Chem.
A 2019, 7, 16902–16911. DOI: 10.1039/C9TA04562G.
13. Gu, H.; Xu, X.; Cai, J.; Wei, S.; Wei, H.; Liu, H.; Young, D. P.; Shao, Q.; Wu, S.; Ding, T.;
Guo, Z. Controllable Organic Magnetoresistance in Polyaniline Coated Poly(p-Phenylene-2,
6-Benzobisoxazole) Short Fibers. Chem. Commun. 2019, 55, 10068–10071. DOI: 10.1039/
C9CC04789A.
14. Ma, Y.; Hou, C.; Zhang, H.; Zhang, Q.; Liu, H.; Wu, S.; Guo, Z. Three-Dimensional Core-
Shell Fe3O4/Polyaniline Coaxial Heterogeneous Nanonets: Preparation and High
Performance Supercapacitor Electrodes. Electrochim. Acta 2019, 315, 114–123. DOI: 10.
1016/j.electacta.2019.05.073.
15. Wang, Y.; Jiang, D.; Zhang, L.; Li, B.; Sun, C.; Yan, H.; Wu, Z.; Liu, H.; Zhang, J.; Fan, J.;
et al. Hydrogen Bonding Derived Self-Healing Polymer Composites Reinforced with
Amidation Carbon Fibers. Nanotechnology 2020, 31, 025704. DOI: 10.1088/1361-6528/
ab4743.
16. Guo, Y.; Yang, X.; Ruan, K.; Kong, J.; Dong, M.; Zhang, J.; Gu, J.; Guo, Z. Reduced
Graphene Oxide Heterostructured Silver Nanoparticles Significantly Enhanced Thermal
Conductivities in Hot-Pressed Electrospun Polyimide Nanocomposites. ACS Appl. Mater.
Interfaces 2019, 11, 25465–25473. DOI: 10.1021/acsami.9b10161.
17. Guo, Y. Q.; Ruan, K. P.; Yang, X. T.; Ma, T. B.; Kong, J.; Wu, N. N.; Zhang, J. X.; Gu,
J. W.; Guo, Z. H. Constructing Fully Carbon-Based Fillers with a Hierarchical Structure to
Fabricate Highly Thermally Conductive Polyimide Nanocomposites. J. Mater. Chem. C
2019, 7, 7035–7044. DOI: 10.1039/C9TC01804B.
18. He, Y. X.; Chen, Q. Y.; Liu, H.; Zhang, L.; Wu, D. Y.; Lu, C.; OuYang, W.; Jiang, D. F.;
Wu, M. F.; Zhang, J. X.; et al. Friction and Wear of MoO3/Graphene Oxide Modified
180 J. CHEN ET AL.

Glass Fiber Reinforced Epoxy Nanocomposites. Macromol. Mater. Eng. 2019, 304,
1900166. DOI: 10.1002/mame.201900166.
19. Cui, X.; Zhu, G.; Pan, Y.; Shao, Q.; Zhao, C.; Dong, M.; Zhang, Y.; Guo, Z.
Polydimethylsiloxane-Titania Nanocomposite Coating: Fabrication and Corrosion
Resistance. Polymer 2018, 138, 203–210. DOI: 10.1016/j.polymer.2018.01.063.
20. Si, W.; Sun, J.; He, X.; Huang, Y.; Zhuang, J.; Zhang, J.; Murugadoss, V.; Fan, J.; Wu, D.;
Guo, Z. Enhancing Thermal Conductivity via Conductive Network Conversion from High
to Low Thermal Dissipation in the Polydimethylsiloxane Composites. J. Mater. Chem. C
2020, in press, DOI: 10.1039/C9TC06968B.
21. Deng, H.; Lin, L.; Ji, M.; Zhang, S.; Yang, M.; Fu, Q. Progress on the Morphological
Control of Conductive Network in Conductive Polymer Composites and the Use as
Electroactive Multifunctional Materials. Prog. Polym. Sci. 2014, 39, 627–655. DOI: 10.
1016/j.progpolymsci.2013.07.007.
22. Zheng, Z.; Olayinka, O.; Li, B. 2S-Soy Protein-Based Biopolymer as a Non-Covalent
Surfactant and Its Effects on Electrical Conduction and Dielectric Relaxation of Polymer
Nanocomposites. Eng. Sci. 2018, 4, 87–99. DOI: 10.30919/es8d766.
23. Xie, L.; Zhu, Y. Tune the Phase Morphology to Design Conductive Polymer Composites:
A Review. Polym. Compos. 2018, 39, 2985–2996. DOI: 10.1002/pc.24345.
24. Zheng, Y.; Wang, X.; Wu, G. Chemical Modification of Carbon Fiber with
Diethylenetriaminepentaacetic Acid/Halloysite Nanotube as a Multifunctional Interfacial
Reinforcement for Silicone Resin Composites. Polym. Adv. Technol. 2020, 31, 527–535.
DOI: 10.1002/pat.4793.
25. Zheng, Y.; Chen, L.; Wang, X.; Wu, G. Modification of Renewable Cardanol onto Carbon
Fiber for the Improved Interfacial Properties of Advanced Polymer Composites. Polymers
2019, 12, 45. DOI: 10.3390/polym12010045.
26. Gu, H.; Xu, X.; Dong, M.; Xie, P.; Shao, Q.; Fan, R.; Liu, C.; Wu, S.; Wei, R.; Guo, Z.
Carbon Nanospheres Induced High Negative Permittivity in Nanosilver-Polydopamine
Metacomposites. Carbon 2019, 147, 550–558. DOI: 10.1016/j.carbon.2019.03.028.
27. Jiang, D.; Wang, Y.; Li, B.; Sun, C.; Wu, Z.; Yan, H.; Xing, L.; Qi, S.; Li, Y.; Liu, H.; et al.
Flexible Sandwich Structural Strain Sensor Based on Silver Nanowires Decorated with
Self-Healing Substrate. Macromol. Mater. Eng. 2019, 304, 1900074. DOI: 10.1002/mame.
201900074.
28. Zhu, G.; Cui, X.; Zhang, Y.; Chen, S.; Dong, M.; Liu, H.; Shao, Q.; Ding, T.; Wu, S.; Guo,
Z. Poly (Vinyl Butyral)/Graphene Oxide/Poly (Methylhydrosiloxane) Nanocomposite
Coating for Improved Aluminum Alloy Anticorrosion. Polymer 2019, 172, 415–422. DOI:
10.1016/j.polymer.2019.03.056.
29. Zhang, J.; Zhang, W.; Wei, L.; Pu, L.; Liu, J.; Liu, H.; Li, Y.; Fan, J.; Ding, T.; Guo, Z.
Alternating Multilayer Structural Epoxy Composite Coating for Corrosion Protection of
Steel. Macromol. Mater. Eng. 2019, 304, 1900374. DOI: 10.1002/mame.201900374.
30. Guo, X. K.; Ge, S. S.; Wang, J. X.; Zhang, X. C.; Zhang, T.; Lin, J.; Zhao, C. X. X.; Wang,
B.; Zhu, G. F.; Guo, Z. H. Waterborne Acrylic Resin Modified with Glycidyl Methacrylate
(GMA): Formula Optimization and Property Analysis. Polymer 2018, 143, 155–163. DOI:
10.1016/j.polymer.2018.04.020.
31. Wei, H.; Wang, H.; Li, A.; Cui, D.; Zhao, Z.; Chu, L.; Wei, X.; Wang, L.; Pan, D.; Fan, J.;
et al. Multifunctions of Polymer Nanocomposites: Environmental Remediation,
Electromagnetic Interference Shielding, and Sensing Applications. ChemNanoMat 2020, 6,
174–184. DOI: 10.1002/cnma.201900588.
32. Jiang, D.; Murugadoss, V.; Wang, Y.; Lin, J.; Ding, T.; Wang, Z.; Shao, Q.; Wang, C.; Liu,
H.; Lu, N.; et al. Electromagnetic Interference Shielding Polymers and Nanocomposites -
A Review. Polym. Rev. 2019, 59, 280–337.
33. Wang, C.; Murugadoss, V.; Kong, J.; He, Z.; Mai, X.; Shao, Q.; Chen, Y.; Guo, L.; Liu, C.;
Angaiah, S.; Guo, Z. Overview of Carbon Nanostructures and Nanocomposites for
Electromagnetic Wave Shielding. Carbon 2018, 140, 696–733. DOI: 10.1016/j.carbon.2018.
09.006.
POLYMER REVIEWS 181

34. Wu, N.; Xu, D.; Wang, Z.; Wang, F.; Liu, J.; Liu, W.; Shao, Q.; Liu, H.; Gao, Q.; Guo, Z.
Achieving Superior Electromagnetic Wave Absorbers through the Novel Metal-Organic
Frameworks Derived Magnetic Porous Carbon Nanorods. Carbon 2019, 145, 433–444.
DOI: 10.1016/j.carbon.2019.01.028.
35. Wu, N.; Liu, C.; Xu, D.; Liu, J.; Liu, W.; Liu, H.; Zhang, J.; Xie, W.; Guo, Z. Ultrathin
High-Performance Electromagnetic Wave Absorbers with Facilely Fabricated Hierarchical
Porous Co/C Crabapples. J. Mater. Chem. C 2019, 7, 1659–1669. DOI: 10.1039/
C8TC04984J.
36. Ma, P.-C.; Siddiqui, N. A.; Marom, G.; Kim, J.-K. Dispersion and Functionalization of
Carbon Nanotubes for Polymer-Based Nanocomposites: A Review. Compos. Part A Appl.
Sci. Manuf. 2010, 41, 1345–1367. DOI: 10.1016/j.compositesa.2010.07.003.
37. Wei, H.; Wang, H.; Xia, Y.; Cui, D.; Shi, Y.; Dong, M.; Liu, C.; Ding, T.; Zhang, J.; Ma,
Y.; et al. An Overview of Lead-Free Piezoelectric Materials and Devices. J. Mater. Chem. C
2018, 6, 12446–12467. DOI: 10.1039/C8TC04515A.
38. Liu, H.; Li, Q.; Bu, Y.; Zhang, N.; Wang, C.; Pan, C.; Mi, L.; Guo, Z.; Liu, C.; Shen, C.
Stretchable Conductive Nonwoven Fabrics with Self-Cleaning Capability for Tunable
Wearable Strain Sensor. Nano Energy 2019, 66, 104143. DOI: 10.1016/j.nanoen.2019.
104143.
39. Gu, H.; Zhang, H.; Ma, C.; Sun, H.; Liu, C.; Dai, K.; Zhang, J.; Wei, R.; Ding, T.; Guo, Z.
Smart Strain Sensing Organic–Inorganic Hybrid Hydrogels with Nano Barium Ferrite as
the Cross-Linker. J. Mater. Chem. C 2019, 7, 2353–2360. DOI: 10.1039/C8TC05448G.
40. Liu, H.; Li, Q.; Zhang, S.; Yin, R.; Liu, X.; He, Y.; Dai, K.; Shan, C.; Guo, J.; Liu, C.; et al.
Electrically Conductive Polymer Composites for Smart Flexible Strain Sensors: A Critical
Review. J. Mater. Chem. C 2018, 6, 12121–12141. DOI: 10.1039/C8TC04079F.
41. Chen, J. W.; Yu, Q. L.; Cui, X. H.; Dong, M. Y.; Zhang, J. X.; Wang, C.; Fan, J. C.; Zhu,
Y. T.; Guo, Z. H. An Overview of Stretchable Strain Sensors from Conductive Polymer
Nanocomposites. J. Mater. Chem. C 2019, 7, 11710–11730. DOI: 10.1039/C9TC03655E.
42. Li, Y.; Zhang, T.; Jiang, B.; Zhao, L.; Liu, H.; Zhang, J.; Fan, J.; Guo, Z.; Huang, Y.
Interfacially Reinforced Carbon Fiber Silicone Resin via Constructing Functional Nano-
Structural Silver. Compos. Sci. Technol. 2019, 181, 107689. DOI: 10.1016/j.compscitech.
2019.107689.
43. Ma, R.; Wang, Y.; Qi, H.; Shi, C.; Wei, G.; Xiao, L.; Huang, Z.; Liu, S.; Yu, H.; Teng, C.;
et al. Nanocomposite Sponges of Sodium Alginate/Graphene Oxide/Polyvinyl Alcohol as
Potential Wound Dressing: In Vitro and in Vivo Evaluation. Compos. Part B Eng. 2019,
167, 396–405. DOI: 10.1016/j.compositesb.2019.03.006.
44. Zhang, Z. Z.; Zhang, J. X.; Li, S. Y.; Liu, J. P.; Dong, M. Y.; Li, Y. C.; Lu, N.; Lei, S. Y.;
Tang, J. J.; Fan, J. C.; Guo, Z. H. Effect of Graphene Liquid Crystal on Dielectric
Properties of Polydimethylsiloxane Nanocomposites. Compos. Part B Eng. 2019, 176,
107338. DOI: 10.1016/j.compositesb.2019.107338.
45. Dong, M.; Li, Q.; Liu, H.; Liu, C.; Wujcik, E. K.; Shao, Q.; Ding, T.; Mai, X.; Shen, C.;
Guo, Z. Thermoplastic Polyurethane-Carbon Black Nanocomposite Coating: Fabrication
and Solid Particle Erosion Resistance. Polymer 2018, 158, 381–390. DOI: 10.1016/j.poly-
mer.2018.11.003.
46. Dong, M.; Wang, C.; Liu, H.; Liu, C.; Shen, C.; Zhang, J.; Jia, C.; Ding, T.; Guo, Z.
Enhanced Solid Particle Erosion Properties of Thermoplastic Polyurethane-Carbon
Nanotube Nanocomposites. Macromol. Mater. Eng. 2019, 304, 1900010. DOI: 10.1002/
mame.201900010.
47. Wu, Z.; Cui, H.; Chen, L.; Jiang, D.; Weng, L.; Ma, Y.; Li, X.; Zhang, X.; Liu, H.; Wang,
N.; et al. Interfacially Reinforced Unsaturated Polyester Carbon Fiber Composites with a
Vinyl Ester-Carbon Nanotubes Sizing Agent. Compos. Sci. Technol. 2018, 164, 195–203.
DOI: 10.1016/j.compscitech.2018.05.051.
48. Pan, Y.; Li, L.; Chan, S. H.; Zhao, J. Correlation between Dispersion State and Electrical
Conductivity of MWCNTs/PP Composites Prepared by Melt Blending. Compos. Part A
Appl. Sci. Manuf. 2010, 41, 419–426. DOI: 10.1016/j.compositesa.2009.11.009.
182 J. CHEN ET AL.

49. Ezat, G. S.; Kelly, A. L.; Mitchell, S. C.; Youseffi, M.; Coates, P. D. Effect of Maleic
Anhydride Grafted Polypropylene Compatibilizer on the Morphology and Properties of
Polypropylene/Multiwalled Carbon Nanotube Composite. Polym. Compos. 2012, 33,
1376–1386. DOI: 10.1002/pc.22264.
50. Cheng, H. K. F.; Pan, Y.; Sahoo, N. G.; Chong, K.; Li, L.; Chan, S. H.; Zhao, J.
Improvement in Properties of Multiwalled Carbon Nanotube/Polypropylene
Nanocomposites through Homogeneous Dispersion with the Aid of Surfactants. J. Appl.
Polym. Sci. 2012, 124, 1117–1127. DOI: 10.1002/app.35047.
51. Gorrasi, G.; Bredeau, S.; Di Candia, C.; Patimo, G.; De Pasquale, S.; Dubois, P.
Electroconductive Polyamide 6/MWNT Nanocomposites: Effect of Nanotube Surface-
Coating by in Situ Catalyzed Polymerization. Macromol. Mater. Eng. 2011, 296, 408–413.
DOI: 10.1002/mame.201000336.
52. Huang, Y. Y.; Ahir, S. V.; Terentjev, E. M. Dispersion Rheology of Carbon Nanotubes in
a Polymer Matrix. Phys. Rev. B 2006, 73, 125422. DOI: 10.1103/PhysRevB.73.125422.
53. Villmow, T.; Kretzschmar, B.; P€ otschke, P. Influence of Screw Configuration, Residence
Time, and Specific Mechanical Energy in Twin-Screw Extrusion of Polycaprolactone/
Multi-Walled Carbon Nanotube Composites. Compos. Sci. Technol. 2010, 70, 2045–2055.
DOI: 10.1016/j.compscitech.2010.07.021.
54. Wu, F.; Lu, Y.; Shao, G.; Zeng, F.; Wu, Q. Preparation of Polyacrylonitrile/Graphene
Oxide by in Situ Polymerization. Polym. Int. 2012, 61, 1394–1399. DOI: 10.1002/pi.4221.
55. Ye, Y.-S.; Chen, Y.-N.; Wang, J.-S.; Rick, J.; Huang, Y.-J.; Chang, F.-C.; Hwang, B.-J.
Versatile Grafting Approaches to Functionalizing Individually Dispersed Graphene
Nanosheets Using RAFT Polymerization and Click Chemistry. Chem. Mater. 2012, 24,
2987–2997. DOI: 10.1021/cm301345r.
56. Wang, X. W.; Zhang, C. A.; Wang, P. L.; Zhao, J.; Zhang, W.; Ji, J. H.; Hua, K.; Zhou, J.;
Yang, X. B.; Li, X. P. Enhanced Performance of Biodegradable Poly(Butylene Succinate)/
Graphene Oxide Nanocomposites via in Situ Polymerization. Langmuir 2012, 28,
7091–7095. DOI: 10.1021/la204894h.
57. Nayak, S.; Bhattacharjee, S.; Singh, B. P. Preparation of Transparent and Conducting
Carbon Nanotube/N-Hydroxymethyl Acrylamide Composite Thin Films by in Situ
Polymerization. Carbon 2012, 50, 4269–4276. DOI: 10.1016/j.carbon.2012.05.010.
58. Liu, K.; Chen, L.; Chen, Y.; Wu, J.; Zhang, W.; Chen, F.; Fu, Q. Preparation of Polyester/
Reduced Graphene Oxide Composites via in Situ Melt Polycondensation and
Simultaneous Thermo-Reduction of Graphene Oxide. J. Mater. Chem. 2011, 21,
8612–8617. DOI: 10.1039/c1jm10717h.
59. Sahoo, N. G.; Rana, S.; Cho, J. W.; Li, L.; Chan, S. H. Polymer Nanocomposites Based on
Functionalized Carbon Nanotubes. Prog. Polym. Sci. 2010, 35, 837–867. DOI: 10.1016/j.
progpolymsci.2010.03.002.
60. Spitalsky, Z.; Tasis, D.; Papagelis, K.; Galiotis, C. Carbon Nanotube-Polymer Composites:
Chemistry, Processing, Mechanical and Electrical Properties. Prog. Polym. Sci. 2010, 35,
357–401. DOI: 10.1016/j.progpolymsci.2009.09.003.
61. Li, W.; Tang, X.-Z.; Zhang, H.-B.; Jiang, Z.-G.; Yu, Z.-Z.; Du, X.-S.; Mai, Y.-W.
Simultaneous Surface Functionalization and Reduction of Graphene Oxide with
Octadecylamine for Electrically Conductive Polystyrene Composites. Carbon 2011, 49,
4724–4730. DOI: 10.1016/j.carbon.2011.06.077.
62. El Sawi, I.; Olivier, P. A.; Demont, P.; Bougherara, H. Processing and Electrical
Characterization of a Unidirectional CFRP Composite Filled with Double Walled Carbon
Nanotubes. Compos. Sci. Technol. 2012, 73, 19–26. DOI: 10.1016/j.compscitech.2012.08.
016.
63. Combessis, A.; Bayon, L.; Flandin, L. Effect of Filler Auto-Assembly on Percolation
Transition in Carbon Nanotube/Polymer Composites. Appl. Phys. Lett. 2013, 102, 011907.
DOI: 10.1063/1.4773994.
POLYMER REVIEWS 183

64. Thostenson, E. T.; Chou, T.-W. Real-Time in Situ Sensing of Damage Evolution in
Advanced Fiber Composites Using Carbon Nanotube Networks. Nanotechnology 2008, 19,
215713. DOI: 10.1088/0957-4484/19/21/215713.
65. Thostenson, E. T.; Chou, T.-W. Carbon Nanotube Networks: Sensing of Distributed Strain
and Damage for Life Prediction and Self Healing. Adv. Mater. 2006, 18, 2837–2841. DOI:
10.1002/adma.200600977.
66. Souier, T.; Santos, S.; Al Ghaferi, A.; Stefancich, M.; Chiesa, M. Enhanced Electrical
Properties of Vertically Aligned Carbon Nanotube-Epoxy Nanocomposites with High
Packing Density. Nanoscale. Res. Lett. 2012, 7, 1–8.
67. Dang, Z.-M.; Jiang, M.-J.; Xie, D.; Yao, S.-H.; Zhang, L.-Q.; Bai, J. Supersensitive Linear
Piezoresistive Property in Carbon Nanotubes/Silicone Rubber Nanocomposites. J. Appl.
Phys. 2008, 104, 024114. DOI: 10.1063/1.2956605.
68. Bokobza, L. Some Issues in Rubber Nanocomposites: New Opportunities for Silicone
Materials. Silicon 2009, 1, 141–145. DOI: 10.1007/s12633-009-9010-6.
69. Bloor, D.; Graham, A.; Williams, E. J.; Laughlin, P. J.; Lussey, D. Metal-Polymer
Composite with Nanostructured Filler Particles and Amplified Physical Properties. Appl.
Phys. Lett. 2006, 88, 102103. DOI: 10.1063/1.2183359.
70. Huang, J.; Li, N.; Xiao, L.; Liu, H.; Wang, Y.; Chen, J.; Nie, X.; Zhu, Y. Fabrication of a
Highly Tough, Strong, and Stiff Carbon Nanotube/Epoxy Conductive Composite with an
Ultralow Percolation Threshold via Self-Assembly. J. Mater. Chem. A 2019, 7,
15731–15740. DOI: 10.1039/C9TA04256C.
71. Ma, L.; Zhu, Y.; Feng, P.; Song, G.; Huang, Y.; Liu, H.; Zhang, J.; Fan, J.; Hou, H.; Guo,
Z. Reinforcing Carbon Fiber Epoxy Composites with Triazine Derivatives Functionalized
Graphene Oxide Modified Sizing Agent. Compos. Part B Eng. 2019, 176, 107078. DOI: 10.
1016/j.compositesb.2019.107078.
72. Gong, X.; Liu, Y.; Wang, Y.; Xie, Z.; Dong, Q.; Dong, M.; Liu, H.; Shao, Q.; Lu, N.;
Murugadoss, V.; et al. Amino Graphene Oxide/Dopamine Modified Aramid Fibers:
Preparation, Epoxy Nanocomposites and Property Analysis. Polymer 2019, 168, 131–137.
DOI: 10.1016/j.polymer.2019.02.021.
73. He, Y.; Yang, S.; Liu, H.; Shao, Q.; Chen, Q.; Lu, C.; Jiang, Y.; Liu, C.; Guo, Z. Reinforced
Carbon Fiber Laminates with Oriented Carbon Nanotube Epoxy Nanocomposites:
Magnetic Field Assisted Alignment and Cryogenic Temperature Mechanical Properties. J.
Colloid Interface Sci. 2018, 517, 40–51. DOI: 10.1016/j.jcis.2018.01.087.
74. Wu, Z.; Gao, S.; Chen, L.; Jiang, D.; Shao, Q.; Zhang, B.; Zhai, Z.; Wang, C.; Zhao, M.;
Ma, Y.; et al. Electrically Insulated Epoxy Nanocomposites Reinforced with Synergistic
Core–Shell SiO2@MWCNTs and Montmorillonite Bifillers. Macromol. Chem. Phys. 2017,
218, 1700357. DOI: 10.1002/macp.201700357.
75. He, Y.; Chen, Q.; Yang, S.; Lu, C.; Feng, M.; Jiang, Y.; Cao, G.; Zhang, J.; Liu, C. Micro-
Crack Behavior of Carbon Fiber Reinforced Fe3O4/Graphene Oxide Modified Epoxy
Composites for Cryogenic Application. Compos. Part A Appl. Sci. Manuf. 2018, 108,
12–22. DOI: 10.1016/j.compositesa.2018.02.014.
76. Lee, J.; Lim, M.; Yoon, J.; Kim, M. S.; Choi, B.; Kim, D. M.; Kim, D. H.; Park, I.; Choi,
S. J. Transparent, Flexible Strain Sensor Based on a Solution-Processed Carbon Nanotube
Network. ACS Appl. Mater. Interfaces 2017, 9, 26279–26285. DOI: 10.1021/acsami.
7b03184.
77. Gupta, N.; Rao, K. D. M.; Srivastava, K.; Gupta, R.; Kumar, A.; Marconnet, A.; Fisher,
T. S.; Kulkarni, G. U. Cosmetically Adaptable Transparent Strain Sensor for Sensitively
Delineating Patterns in Small Movements of Vital Human Organs. ACS Appl. Mater.
Interfaces 2018, 10, 44126–44133. DOI: 10.1021/acsami.8b16282.
78. Wang, Z.; Gao, W.; Zhang, Q.; Zheng, K.; Xu, J.; Xu, W.; Shang, E.; Jiang, J.; Zhang, J.;
Liu, Y. 3D-Printed Graphene/Polydimethylsiloxane Composites for Stretchable and Strain-
Insensitive Temperature Sensors. ACS Appl. Mater. Interfaces 2019, 11, 1344–1352. DOI:
10.1021/acsami.8b16139.
184 J. CHEN ET AL.

79. Wu, S.; Zhang, J.; Ladani, R. B.; Ravindran, A. R.; Mouritz, A. P.; Kinloch, A. J.; Wang,
C. H. Novel Electrically Conductive Porous PDMS/Carbon Nanofiber Composites for
Deformable Strain Sensors and Conductors. ACS Appl. Mater. Interfaces 2017, 9,
14207–14215. DOI: 10.1021/acsami.7b00847.
80. Yu, H.; Lian, Y.; Sun, T.; Yang, X.; Wang, Y.; Xie, G.; Du, X.; Gou, J.; Li, W.; Tai, H. Two-
Sided Topological Architecture on a Monolithic Flexible Substrate for Ultrasensitive Strain
Sensors. ACS Appl. Mater. Interfaces 2019, 11, 43543–43552. DOI: 10.1021/acsami.9b14476.
81. Zheng, Y.; Li, Y.; Li, Z.; Wang, Y.; Dai, K.; Zheng, G.; Liu, C.; Shen, C. The Effect of
Filler Dimensionality on the Electromechanical Performance of Polydimethylsiloxane
Based Conductive Nanocomposites for Flexible Strain Sensors. Compos. Sci. Technol.
2017, 139, 64–73. DOI: 10.1016/j.compscitech.2016.12.014.
82. Xiang, D.; Zhang, X.; Li, Y.; Harkin-Jones, E.; Zheng, Y.; Wang, L.; Zhao, C.; Wang, P.
Enhanced Performance of 3D Printed Highly Elastic Strain Sensors of Carbon Nanotube/
Thermoplastic Polyurethane Nanocomposites via Non-Covalent Interactions. Compos. Part
B Eng. 2019, 176, 107250. DOI: 10.1016/j.compositesb.2019.107250.
83. Balberg, I.; Binenbaum, N.; Wagner, N. Percolation Thresholds in the Three-Dimensional
Sticks System. Phys. Rev. Lett. 1984, 52, 1465–1468. DOI: 10.1103/PhysRevLett.52.1465.
84. Yoonessi, M.; Gaier, J. R. Highly Conductive Multifunctional Graphene Polycarbonate
Nanocomposites. ACS Nano 2010, 4, 7211–7220. DOI: 10.1021/nn1019626.
85. Levine, L. E.; Long, G. G.; Ilavsky, J.; Gerhardt, R. A.; Ou, R.; Parker, C. A. Self-Assembly
of Carbon Black into Nanowires That Form a Conductive Three Dimensional
Micronetwork. Appl. Phys. Lett. 2007, 90, 014101. DOI: 10.1063/1.2425011.
86. Waddell, J.; Ou, R.; Capozzi, C. J.; Gupta, S.; Parker, C. A.; Gerhardt, R. A.; Seal, K.;
Kalinin, S. V.; Baddorf, A. P. Detection of Percolating Paths in Polyhedral Segregated
Network Composites Using Electrostatic Force Microscopy and Conductive Atomic Force
Microscopy. Appl. Phys. Lett. 2009, 95, 233122. DOI: 10.1063/1.3265742.
87. Al-Saleh, M. H.; Sundararaj, U. An Innovative Method to Reduce Percolation Threshold
of Carbon Black Filled Immiscible Polymer Blends. Compos. Part A Appl. Sci. Manuf.
2008, 39, 284–293. DOI: 10.1016/j.compositesa.2007.10.010.
88. Dai, K.; Zhao, S.; Zhai, W.; Zheng, G.; Liu, C.; Chen, J.; Shen, C. Tuning of Liquid
Sensing Performance of Conductive Carbon Black (CB)/Polypropylene (PP) Composite
Utilizing a Segregated Structure. Compos. Part A Appl. Sci. Manuf. 2013, 55, 11–18. DOI:
10.1016/j.compositesa.2013.08.001.
89. Balogun, Y. A.; Buchanan, R. C. Enhanced Percolative Properties from Partial Solubility
Dispersion of Filler Phase in Conducting Polymer Composites (CPCs). Compos. Sci.
Technol. 2010, 70, 892–900. DOI: 10.1016/j.compscitech.2010.01.009.
90. Malliaris, A.; Turner, D. T. Influence of Particle Size on the Electrical Resistivity of
Compacted Mixtures of Polymeric and Metallic Powders. J. Appl. Phys 1971, 42, 614–618.
DOI: 10.1063/1.1660071.
91. Grunlan, J. C.; Gerberich, W. W.; Francis, L. F. Lowering the Percolation Threshold of
Conductive Composites Using Particulate Polymer Microstructure. J. Appl. Polym. Sci. 2001,
80, 692–705. DOI: 10.1002/1097-4628(20010425)80:4<692::AID-APP1146>3.0.CO;2-W.
92. Kusy, R. P.; Turner, D. T. Electrical Conductivity of a Polyurethane Elastomer Containing
Segregated Particles of Nickel. J. Appl. Polym. Sci. 1973, 17, 1631–1633. DOI: 10.1002/app.
1973.070170528.
93. Gupta, S.; Ou, R. Q.; Gerhardt, R. A. Effect of the Fabrication Method on the Electrical
Properties of Poly(Acrylonitrile-co-Butadiene-co-Styrene)/Carbon Black Composites. J.
Elec. Mater. 2006, 35, 224–229. DOI: 10.1007/BF02692439.
94. Pang, H.; Bao, Y.; Xu, L.; Yan, D.-X.; Zhang, W.-Q.; Wang, J.-H.; Li, Z.-M. Double-
Segregated Carbon Nanotube-Polymer Conductive Composites as Candidates for Liquid
Sensing Materials. J. Mater. Chem. A 2013, 1, 4177–4181. DOI: 10.1039/c3ta10242d.
95. Bao, Y.; Xu, L.; Pang, H.; Yan, D.-X.; Chen, C.; Zhang, W.-Q.; Tang, J.-H.; Li, Z.-M.
Preparation and Properties of Carbon Black/Polymer Composites with Segregated and
POLYMER REVIEWS 185

Double-Percolated Network Structures. J. Mater. Sci. 2013, 48, 4892–4898. DOI: 10.1007/
s10853-013-7269-x.
96. Jurewicz, I.; Worajittiphon, P.; King, A. A. K.; Sellin, P. J.; Keddie, J. L.; Dalton, A. B.
Locking Carbon Nanotubes in Confined Lattice Geometries - A Route to Low Percolation
in Conducting Composites. J. Phys. Chem. B 2011, 115, 6395–6400. DOI: 10.1021/
jp111998p.
97. Ou, R.; Gupta, S.; Parker, C. A.; Gerhardt, R. A. Fabrication and Electrical Conductivity
of Poly(Methyl Methacrylate) (PMMA)/Carbon Black (CB) Composites: Comparison
between an Ordered Carbon Black Nanowire-like Segregated Structure and a Randomly
Dispersed Carbon Black Nanostructure. J. Phys. Chem. B 2006, 110, 22365–22373. DOI:
10.1021/jp064498o.
98. Pang, H.; Piao, Y.-Y.; Cui, C.-H.; Bao, Y.; Lei, J.; Yuan, G.-P.; Zhang, C.-L. Preparation and
Performance of Segregated Polymer Composites with Hybrid Fillers of Octadecylamine
Functionalized Graphene and Carbon Nanotubes. J. Polym. Res. 2013, 20, 304.
99. Pang, H.; Chen, C.; Bao, Y.; Chen, J.; Ji, X.; Lei, J.; Li, Z.-M. Electrically Conductive Carbon
Nanotube/Ultrahigh Molecular Weight Polyethylene Composites with Segregated and Double
Percolated Structure. Mater. Lett. 2012, 79, 96–99. DOI: 10.1016/j.matlet.2012.03.111.
100. Bharati, A.; Cardinaels, R.; Seo, J. W.; Wubbenhorst, M.; Moldenaers, P. Enhancing the
Conductivity of Carbon Nanotube Filled Blends by Tuning Their Phase Separated Morphology
with a Copolymer. Polymer 2015, 79, 271–282. DOI: 10.1016/j.polymer.2015.09.080.
101. Grunlan, J. C.; Gerberich, W. W.; Francis, L. F. Electrical and Mechanical Behavior of
Carbon Black–Filled Poly(Vinyl Acetate) Latex–Based Composites. Polym. Eng. Sci. 2001,
41, 1947–1962. DOI: 10.1002/pen.10891.
102. Mao, C.; Zhu, Y.; Jiang, W. Design of Electrical Conductive Composites: tuning the
Morphology to Improve the Electrical Properties of Graphene Filled Immiscible Polymer
Blends. ACS Appl. Mater. Interfaces 2012, 4, 5281–5286. DOI: 10.1021/am301230q.
103. Qi, X.-Y.; Yan, D.; Jiang, Z.; Cao, Y.-K.; Yu, Z.-Z.; Yavari, F.; Koratkar, N. Enhanced
Electrical Conductivity in Polystyrene Nanocomposites at Ultra-Low Graphene Content.
ACS Appl. Mater. Interfaces 2011, 3, 3130–3133. DOI: 10.1021/am200628c.
104. Xu, Z.; Zhang, Y.; Wang, Z.; Sun, N.; Li, H. Enhancement of Electrical Conductivity by
Changing Phase Morphology for Composites Consisting of Polylactide and Poly(Epsilon-
Caprolactone) Filled with Acid-Oxidized Multiwalled Carbon Nanotubes. ACS Appl.
Mater. Interfaces 2011, 3, 4858–4864. DOI: 10.1021/am201355j.
105. Huang, J.; Mao, C.; Zhu, Y.; Jiang, W.; Yang, X. Control of Carbon Nanotubes at the
Interface of a co-Continuous Immiscible Polymer Blend to Fabricate Conductive
Composites with Ultralow Percolation Thresholds. Carbon 2014, 73, 267–274. DOI: 10.
1016/j.carbon.2014.02.063.
106. Gubbels, F.; Jerome, R.; Vanlathem, E.; Deltour, R.; Blacher, S.; Brouers, F. Kinetic and
Thermodynamic Control of the Selective Localization of Carbon Black at the Interface of
Immiscible Polymer Blends. Chem. Mater. 1998, 10, 1227–1235. DOI: 10.1021/cm970594d.
107. Gao, C.; Zhang, S.; Lin, Y.; Li, F.; Guan, S.; Jiang, Z. High-Performance Conductive
Materials Based on the Selective Location of Carbon Black in Poly(Ether Ether Ketone)/
Polyimide Matrix. Compos. Part B Eng. 2015, 79, 124–131.
108. Zhang, L.; Wan, C.; Zhang, Y. Morphology and Electrical Properties of Polyamide 6/
Polypropylene/Multi-Walled Carbon Nanotubes Composites. Compos. Sci. Technol. 2009,
69, 2212–2217. DOI: 10.1016/j.compscitech.2009.06.005.
109. Al-Saleh, M. H.; Sundararaj, U. Nanostructured Carbon Black Filled Polypropylene/
Polystyrene Blends Containing Styrene-Butadiene-Styrene Copolymer: Influence of
Morphology on Electrical Resistivity. Eur. Polym. J. 2008, 44, 1931–1939. DOI: 10.1016/j.
eurpolymj.2008.04.013.
110. Cui, L.; Zhang, Y.; Zhang, Y.; Zhang, X.; Zhou, W. Electrical Properties and Conductive
Mechanisms of Immiscible Polypropylene/Novolac Blends Filled with Carbon Black. Eur.
Polym. J. 2007, 43, 5097–5106. DOI: 10.1016/j.eurpolymj.2007.08.023.
186 J. CHEN ET AL.

111. Pan, Y.; Liu, X.; Hao, X.; Stary, Z.; Schubert, D. W. Enhancing the Electrical Conductivity
of Carbon Black-Filled Immiscible Polymer Blends by Tuning the Morphology. Eur.
Polym. J. 2016, 78, 106–115. DOI: 10.1016/j.eurpolymj.2016.03.019.
112. Gao, X.; Zhang, S.; Mai, F.; Lin, L.; Deng, Y.; Deng, H.; Fu, Q. Preparation of High
Performance Conductive Polymer Fibres from Double Percolated Structure. J. Mater.
Chem. 2011, 21, 6401–6408. DOI: 10.1039/c0jm04543h.
113. Ma, L.-F.; Bao, R.-Y.; Huang, S.-L.; Liu, Z.-Y.; Yang, W.; Xie, B.-H.; Yang, M.-B. Electrical
Properties and Morphology of Carbon Black Filled PP/EPDM Blends: effect of Selective
Distribution of Fillers Induced by Dynamic Vulcanization. J. Mater. Sci. 2013, 48,
4942–4951. DOI: 10.1007/s10853-013-7275-z.
114. Yuan, J.-K.; Yao, S.-H.; Sylvestre, A.; Bai, J. Biphasic Polymer Blends Containing Carbon
Nanotubes: Heterogeneous Nanotube Distribution and Its Influence on the Dielectric
Properties. J. Phys. Chem. C 2012, 116, 2051–2058. DOI: 10.1021/jp210872w.
115. Calberg, C.; Blacher, S.; Gubbels, F.; Brouers, F.; Deltour, R.; Jer?Me, R. Electrical and
Dielectric Properties of Carbon Black Filled co-Continuous Two-Phase Polymer Blends. J.
Phys. D: Appl. Phys. 1999, 32, 1517–1525. DOI: 10.1088/0022-3727/32/13/313.
116. Bose, S.; Bhattacharyya, A. R.; Bondre, A. P.; Kulkarni, A. R.; P€ otschke, P. Rheology,
Electrical Conductivity, and the Phase Behavior of Cocontinuous PA6/ABS Blends with
MWNT: Correlating the Aspect Ratio of MWNT with the Percolation Threshold. J.
Polym. Sci. B Polym. Phys. 2008, 46, 1619–1631. DOI: 10.1002/polb.21501.
117. Khare, R. A.; Bhattacharyya, A. R.; Kulkarni, A. R.; Saroop, M.; Biswas, A. Influence of
Multiwall Carbon Nanotubes on Morphology and Electrical Conductivity of PP/ABS
Blends. J. Polym. Sci. B Polym. Phys. 2008, 46, 2286–2295. DOI: 10.1002/polb.21560.
118. Thongruang, W.; Balik, C. M.; Spontak, R. J. Volume-Exclusion Effects in Polyethylene
Blends Filled with Carbon Black, Graphite, or Carbon Fiber. J. Polym. Sci. B Polym. Phys.
2002, 40, 1013–1025. DOI: 10.1002/polb.10157.
119. Sun, Y.; Guo, Z.-X.; Yu, J. Effect of ABS Rubber Content on the Localization of
MWCNTs in PC/ABS Blends and Electrical Resistivity of the Composites. Macromol.
Mater. Eng. 2010, 295, 263–268. DOI: 10.1002/mame.200900242.
120. Gubbels, F.; Blacher, S.; Vanlathem, E.; Jerome, R.; Deltour, R.; Brouers, F.; Teyssie, P.
Design of Electrical Composites: Determining the Role of the Morphology on the
Electrical Properties of Carbon Black Filled Polymer Blends. Macromolecules 1995, 28,
1559–1566. DOI: 10.1021/ma00109a030.
121. Gubbels, F.; Jerome, R.; Teyssie, P.; Vanlathem, E.; Deltour, R.; Calderone, A.; Parente, V.;
Bredas, J. L. Selective Localization of Carbon Black in Immiscible Polymer Blends: A
Useful Tool to Design Electrical Conductive Composites. Macromolecules 1994, 27,
1972–1974. DOI: 10.1021/ma00085a049.
122. Li, Y.; Shimizu, H. Conductive PVDF/PA6/CNTs Nanocomposites Fabricated by Dual
Formation of Cocontinuous and Nanodispersion Structures. Macromolecules 2008, 41,
5339–5344. DOI: 10.1021/ma8006834.
123. Chen, Y.; Yang, Q.; Huang, Y.; Liao, X.; Niu, Y. Influence of Phase Coarsening and Filler
Agglomeration on Electrical and Rheological Properties of MWNTs-Filled PP/PMMA
Composites under Annealing. Polymer 2015, 79, 159–170. DOI: 10.1016/j.polymer.2015.10.027.
124. P?Tschke, P.; Bhattacharyya, A. R.; Janke, A. Morphology and Electrical Resistivity of Melt
Mixed Blends of Polyethylene and Carbon Nanotube Filled Polycarbonate. Polymer 2003,
44, 8061–8069. DOI: 10.1016/j.polymer.2003.10.003.
125. Thongruang, W.; Spontak, R. J.; Balik, C. M. Bridged Double Percolation in Conductive
Polymer Composites: An Electrical Conductivity, Morphology and Mechanical Property
Study. Polymer 2002, 43, 3717–3725. DOI: 10.1016/S0032-3861(02)00180-5.
126. Al-Saleh, M. H. Carbon Nanotube-Filled Polypropylene/Polyethylene Blends: compatibilization
and Electrical Properties. Polym. Bull. 2016, 73, 975–987. DOI: 10.1007/s00289-015-1530-1.
127. Lu, C.; Wang, R.; Hu, X.; N.; Cao, Q-q.; Huang, X-h.; He, Y-x.; Zhang, Y-q. Influence of
Morphology on PTC Effect for Poly (Ethylene-co-Butyl Acrylate)/nylon6 Blends with
POLYMER REVIEWS 187

Multiwall Carbon Nanotubes Dispersed at Interface and in Matrix. Polym. Bull. 2014, 71,
545–561. DOI: 10.1007/s00289-013-1076-z.
128. Sumita, M.; Sakata, K.; Asai, S.; Miyasaka, K.; Nakagawa, H. Dispersion of Fillers and the
Electrical-Conductivity of Polymer Blends Filled with Carbon-Black. Polym. Bull. 1991,
25, 265–271. DOI: 10.1007/BF00310802.
129. Tan, Y.; Fang, L.; Xiao, J.; Song, Y.; Zheng, Q. Grafting of Copolymers onto Graphene by
Miniemulsion Polymerization for Conductive Polymer Composites: improved Electrical
Conductivity and Compatibility Induced by Interfacial Distribution of Graphene. Polym.
Chem. 2013, 4, 2939. DOI: 10.1039/c3py00164d.
130. Tchoudakov, R.; Breuer, O.; Narkis, M.; Siegmann, A. Conductive Polymer Blends with
Low Carbon Black Loading: Polypropylene/Polyamide. Polym. Eng. Sci. 1996, 36,
1336–1346. DOI: 10.1002/pen.10528.
131. Lyu, H.; Liu, J.; Liu, H.; Liu, C.; Lu, Y.; Sun, K.; Fan, R.; Wang, N.; Lu, N.; Guo, Z.;
Wujcik, E. K. An Overview of Electrically Conductive Polymer Nanocomposites toward
Electromagnetic Interference Shielding. Eng. Sci. 2018, 2, 26–42. DOI: 10.30919/es8d615.
132. Chen, M.; Li, K.; Cheng, G.; He, K.; Li, W.; Zhang, D.; Li, W.; Feng, Y.; Wei, L.; Li, W.;
et al. Touchpoint-Tailored Ultra-Sensitive Piezoresistive Pressure Sensors with a Broad
Dynamic Response Range and Low Detection Limit. ACS Appl. Mater. Interfaces 2018, 11,
2551–2558. DOI: 10.1021/acsami.8b20284.
133. Chen, J.; Li, H.; Yu, Q.; Hu, Y.; Cui, X.; Zhu, Y.; Jiang, W. Strain Sensing Behaviors of
Stretchable Conductive Polymer Composites Loaded with Different Dimensional
Conductive Fillers. Compos. Sci. Technol. 2018, 168, 388–396. DOI: 10.1016/j.compscitech.
2018.10.025.
134. Sekitani, T.; Noguchi, Y.; Hata, K.; Fukushima, T.; Aida, T.; Someya, T. A Rubberlike
Stretchable Active Matrix Using Elastic Conductors. Science 2008, 321, 1468–1472. DOI:
10.1126/science.1160309.
135. Alexopoulos, N. D.; Bartholome, C.; Poulin, P.; Marioli-Riga, Z. Structural Health
Monitoring of Glass Fiber Reinforced Composites Using Embedded Carbon Nanotube
(CNT) Fibers. Compos. Sci. Technol. 2010, 70, 260–271. DOI: 10.1016/j.compscitech.2009.
10.017.
136. Bautista-Quijano, J. R.; Aviles, F.; Aguilar, J. O.; Tapia, A. Strain Sensing Capabilities of a
Piezoresistive MWCNT-Polysulfone Film. Sensor Actuat. A Phys. 2010, 159, 135–140.
DOI: 10.1016/j.sna.2010.03.005.
137. Gao, S-l.; Zhuang, R.-C.; Zhang, J.; Liu, J.-W.; M€ader, E. Glass Fibers with Carbon
Nanotube Networks as Multifunctional Sensors. Adv. Funct. Mater. 2010, 20, 1885–1893.
DOI: 10.1002/adfm.201000283.
138. Hu, N.; Karube, Y.; Arai, M.; Watanabe, T.; Yan, C.; Li, Y.; Liu, Y.; Fukunaga, H.
Investigation on Sensitivity of a Polymer/Carbon Nanotube Composite Strain Sensor.
Carbon 2010, 48, 680–687. DOI: 10.1016/j.carbon.2009.10.012.
139. Murugaraj, P.; Mainwaring, D.; Khelil, N. A.; Peng, J. L.; Siegele, R.; Sawant, P. The
Improved Electromechanical Sensitivity of Polymer Thin Films Containing Carbon
Clusters Produced in Situ by Irradiation with Metal Ions. Carbon 2010, 48, 4230–4237.
DOI: 10.1016/j.carbon.2010.07.026.
140. Rogers, J. A.; Someya, T.; Huang, Y. Materials and Mechanics for Stretchable Electronics.
Science 2010, 327, 1603–1607. DOI: 10.1126/science.1182383.
141. Shin, M. K.; Oh, J.; Lima, M.; Kozlov, M. E.; Kim, S. J.; Baughman, R. H. Elastomeric
Conductive Composites Based on Carbon Nanotube Forests. Adv. Mater. 2010, 22,
2663–2667. DOI: 10.1002/adma.200904270.
142. Eswaraiah, V.; Balasubramaniam, K.; Ramaprabhu, S. Functionalized Graphene Reinforced
Thermoplastic Nanocomposites as Strain Sensors in Structural Health Monitoring. J.
Mater. Chem. 2011, 21, 12626–12628. DOI: 10.1039/c1jm12302e.
143. Gao, L.; Chou, T.-W.; Thostenson, E. T.; Zhang, Z.; Coulaud, M. In Situ Sensing of
Impact Damage in Epoxy/Glass Fiber Composites Using Percolating Carbon Nanotube
Networks. Carbon 2011, 49, 3382–3385. DOI: 10.1016/j.carbon.2011.04.003.
188 J. CHEN ET AL.

144. Granero, A. J.; Wagner, P.; Wagner, K.; Razal, J. M.; Wallace, G. G.; Panhuis, M. I. H.
Highly Stretchable Conducting SIBS-P3HT Fibers. Adv. Funct. Mater. 2011, 21, 955–962.
DOI: 10.1002/adfm.201001460.
145. Hwang, J.; Jang, J.; Hong, K.; Kim, K. N.; Han, J. H.; Shin, K.; Park, C. E. Poly(3-
Hexylthiophene) Wrapped Carbon Nanotube/Poly(Dimethylsiloxane) Composites for Use
in Finger-Sensing Piezoresistive Pressure Sensors. Carbon 2011, 49, 106–110. DOI: 10.
1016/j.carbon.2010.08.048.
146. Kollosche, M.; Stoyanov, H.; Laflamme, S.; Kofod, G. Strongly Enhanced Sensitivity in
Elastic Capacitive Strain Sensors. J. Mater. Chem. 2011, 21, 8292–8294. DOI: 10.1039/
c0jm03786a.
147. Shang, S.; Zeng, W.; Tao, X-m. High Stretchable MWNTs/Polyurethane Conductive
Nanocomposites. J. Mater. Chem. 2011, 21, 7274. DOI: 10.1039/c1jm10255a.
148. Wakuda, D.; Suganuma, K. Stretchable Fine Fiber with High Conductivity Fabricated by
Injection Forming. Appl. Phys. Lett. 2011, 98, 073304. DOI: 10.1063/1.3555433.
149. Ahn, J.-H.; Je, J. H. Stretchable Electronics: materials, Architectures and Integrations. J.
Phys. D Appl. Phys. 2012, 45, 103001. DOI: 10.1088/0022-3727/45/10/103001.
150. Lu, N.; Lu, C.; Yang, S.; Rogers, J. Highly Sensitive Skin-Mountable Strain Gauges Based
Entirely on Elastomers. Adv. Funct. Mater. 2012, 22, 4044–4050. DOI: 10.1002/adfm.
201200498.
151. Lin, L.; Deng, H.; Gao, X.; Zhang, S.; Bilotti, E.; Peijs, T.; Fu, Q. Modified Resistivity-
Strain Behavior through the Incorporation of Metallic Particles in Conductive Polymer
Composite Fibers Containing Carbon Nanotubes. Polym. Int. 2013, 62, 134–140. DOI: 10.
1002/pi.4291.
152. Lin, L.; Liu, S.; Fu, S.; Zhang, S.; Deng, H.; Fu, Q. Fabrication of Highly Stretchable
Conductors via Morphological Control of Carbon Nanotube Network. Small 2013, 9,
3620–3629. DOI: 10.1002/smll.201202306.
153. Lin, L.; Liu, S.; Zhang, Q.; Li, X.; Ji, M.; Deng, H.; Fu, Q. Towards Tunable Sensitivity of
Electrical Property to Strain for Conductive Polymer Composites Based on Thermoplastic
Elastomer. ACS Appl. Mater. Interfaces 2013, 5, 5815–5824. DOI: 10.1021/am401402x.
154. Deng, H.; Ji, M.; Yan, D.; Fu, S.; Duan, L.; Zhang, M.; Fu, Q. Towards Tunable
Resistivity-Strain Behavior through Construction of Oriented and Selectively Distributed
Conductive Networks in Conductive Polymer Composites. J. Mater. Chem. A 2014, 2,
10048–10058. DOI: 10.1039/C4TA01073F.
155. Li, M.; Li, H.; Zhong, W.; Zhao, Q.; Wang, D. Stretchable Conductive Polypyrrole/
Polyurethane (PPy/PU) Strain Sensor with Netlike Microcracks for Human Breath
Detection. ACS Appl. Mater. Interfaces 2014, 6, 1313–1319. DOI: 10.1021/am4053305.
156. Roh, E.; Hwang, B. U.; Kim, D.; Kim, B. Y.; Lee, N. E. Stretchable, Transparent,
Ultrasensitive, and Patchable Strain Sensor for Human-Machine Interfaces Comprising a
Nanohybrid of Carbon Nanotubes and Conductive Elastomers. ACS Nano 2015, 9,
6252–6261. DOI: 10.1021/acsnano.5b01613.
157. Li, Q.; Liu, H.; Zhang, S.; Zhang, D.; Liu, X.; He, Y.; Mi, L.; Zhang, J.; Liu, C.; Shen, C.;
Guo, Z. Superhydrophobic Electrically Conductive Paper for Ultrasensitive Strain Sensor
with Excellent Anticorrosion and Self-Cleaning Property. ACS Appl. Mater. Interfaces
2019, 11, 21904–21914. DOI: 10.1021/acsami.9b03421.
158. Chen, X.; Liu, H.; Zheng, Y.; Zhai, Y.; Liu, X.; Liu, C.; Mi, L.; Guo, Z.; Shen, C. Highly
Compressible and Robust Polyimide/Carbon Nanotube Composite Aerogel for High-
Performance Wearable Pressure Sensor. ACS Appl. Mater. Interfaces 2019, 11,
42594–42606. DOI: 10.1021/acsami.9b14688.
159. Wan, Y.; Qiu, Z.; Hong, Y.; Wang, Y.; Zhang, J.; Liu, Q.; Wu, Z.; Guo, C. F. A Highly
Sensitive Flexible Capacitive Tactile Sensor with Sparse and High-Aspect-Ratio
Microstructures. Adv. Electron. Mater. 2018, 4, 1700586. DOI: 10.1002/aelm.201700586.
160. You, X.; He, J.; Nan, N.; Sun, X.; Qi, K.; Zhou, Y.; Shao, W.; Liu, F.; Cui, S. Stretchable
Capacitive Fabric Electronic Skin Woven by Electrospun Nanofiber Coated Yarns for
POLYMER REVIEWS 189

Detecting Tactile and Multimodal Mechanical Stimuli. J. Mater. Chem. C 2018, 6,


12981–12991. DOI: 10.1039/C8TC03631D.
161. Lee, J.; Kwon, H.; Seo, J.; Shin, S.; Koo, J. H.; Pang, C.; Son, S.; Kim, J. H.; Jang, Y. H.;
Kim, D. E.; Lee, T. Conductive Fiber-Based Ultrasensitive Textile Pressure Sensor for
Wearable Electronics. Adv. Mater. 2015, 27, 2433–2439. DOI: 10.1002/adma.201500009.
162. Zhang, Y.; Fang, Y.; Li, J.; Zhou, Q.; Xiao, Y.; Zhang, K.; Luo, B.; Zhou, J.; Hu, B. Dual-
Mode Electronic Skin with Integrated Tactile Sensing and Visualized Injury Warning.
ACS Appl. Mater. Interfaces 2017, 9, 37493–37500. DOI: 10.1021/acsami.7b13016.
163. Lonergan, M. C.; Brett, E. J. S.; Doleman, J.; Beaber, S. A.; Grubbs, R. H.; Lewis, N. S.
Array-Based Vapor Sensing Using Chemically Sensitive, Carbon Black-Polymer Resistors.
Chem. Mater. 1996, 8, 2298–2312. DOI: 10.1021/cm960036j.
164. Li, J. R.; Xu, J. R.; Zhang, M. Q.; Rong, M. Z. Carbon Black/Polystyrene Composites as
Candidates for Gas Sensing Materials. Carbon 2003, 41, 2353–2360. DOI: 10.1016/S0008-
6223(03)00273-2.
165. Shevade, A. V.; Ryan, M. A.; Homer, M. L.; Manfreda, A. M.; Zhou, H.; Manatt, K. S.
Molecular Modeling of Polymer Composite-Analyte Interactions in Electronic Nose
Sensors. Sens. Actuators B Chem. 2003, 93, 84–91. DOI: 10.1016/S0925-4005(03)00245-4.
166. Dong, X. M.; Fu, R. W.; Zhang, M. Q.; Zhang, B.; Rong, M. Z. Electrical Resistance
Response of Carbon Black Filled Amorphous Polymer Composite Sensors to Organic
Vapors at Low Vapor Concentrations. Carbon 2004, 42, 2551–2559. DOI: 10.1016/j.car-
bon.2004.05.034.
167. Lewis, N. S. Comparisons betweenMammalian and ArtificialOlfaction Based on Arrays
ofCarbon Black  PolymerComposite Vapor Detectors. Acc. Chem. Res. 2004, 37, 663–672.
DOI: 10.1021/ar030120m.
168. Feller, J. F.; Guezenoc, H.; Bellegou, H.; Grohens, Y. Smart Poly(Styrene)/Carbon Black
Conductive Polymer Composites Films for Styrene Vapour Sensing. Macromol. Symp.
2005, 222, 273–280. DOI: 10.1002/masy.200550436.
169. Castro, M.; Lu, J.; Bruzaud, S.; Kumar, B.; Feller, J.-F. Carbon Nanotubes/Poly(Epsilon-
Caprolactone) Composite Vapour Sensors. Carbon 2009, 47, 1930–1942. DOI: 10.1016/j.
carbon.2009.03.037.
170. Castro, M.; Kumar, B.; Feller, J. F.; Haddi, Z.; Amari, A.; Bouchikhi, B. Novel e-Nose for
the Discrimination of Volatile Organic Biomarkers with an Array of Carbon Nanotubes
(CNT) Conductive Polymer Nanocomposites (CPC) Sensors. Sens. Actuators B Chem.
2011, 159, 213–219. DOI: 10.1016/j.snb.2011.06.073.
171. Li, Y.; P€ otschke, P.; Pionteck, J.; Voit, B. Electrical and Vapor Sensing Behaviors of
Polycarbonate Composites Containing Hybrid Carbon Fillers. Eur. Polym. J. 2018, 108,
461–471. DOI: 10.1016/j.eurpolymj.2018.09.027.
172. Bora, A.; Mohan, K.; Pegu, D.; Gohain, C. B.; Dolui, S. K. A Room Temperature
Methanol Vapor Sensor Based on Highly Conducting Carboxylated Multi-Walled Carbon
Nanotube/Polyaniline Nanotube Composite. Sens. Actuators B Chem. 2017, 253, 977–986.
DOI: 10.1016/j.snb.2017.07.023.
173. Benlikaya, R.; Slobodian, P.; Proisl, K.; Cvelbar, U.; Morozov, I. Ascertaining the Factors
That Influence the Vapor Sensor Response: The Entire Case of MWCNT Network Sensor.
Sens. Actuators B Chem. 2019, 283, 478–486. DOI: 10.1016/j.snb.2018.11.160.
174. Qiang, F.; Dai, S.-W.; Zhao, L.; Gong, L.-X.; Zhang, G.-D.; Jiang, J.-X.; Tang, L.-C. An
Insulating Second Filler Tuning Porous Conductive Composites for Highly Sensitive and
Fast Responsive Organic Vapor Sensor. Sens. Actuators B Chem. 2019, 285, 254–263. DOI:
10.1016/j.snb.2019.01.043.
175. Wu, W.; Shi, N.; Zhang, J.; Wu, X.; Wang, T.; Yang, L.; Yang, R.; Ou, C.; Xue, W.; Feng,
X.; et al. Electrospun Fluorescent Sensors for the Selective Detection of Nitro Explosive
Vapors and Trace Water. J. Mater. Chem. A 2018, 6, 18543–18550. DOI: 10.1039/
C8TA01861H.
190 J. CHEN ET AL.

176. Singhal, P.; Mazumdar, P.; Rattan, S. One Pot Synthesis of Free Standing Highly
Conductive Polymer Nanocomposite Films: Towards Rapid BTX Vapor Sensor. Polym.
Eng. Sci. 2018, 58, 1074–1081. DOI: 10.1002/pen.24669.
177. Sheng, J.; Zeng, X.; Zhu, Q.; Yang, Z.; Zhang, X. Facile Fabrication of CNT-Based
Chemical Sensor Operating at Room Temperature. Mater. Res. Express. 2017, 4, 125701.
DOI: 10.1088/2053-1591/aa9ac7.
178. Yan, H.; Zhong, M.; Lv, Z.; Wan, P. Stretchable Electronic Sensors of Nanocomposite
Network Films for Ultrasensitive Chemical Vapor Sensing. Small 2017, 13, 1701697. DOI:
10.1002/smll.201701697.
179. Li, Y.; Zheng, Y.; Zhan, P.; Zheng, G.; Dai, K.; Liu, C.; Shen, C. Vapor Sensing
Performance as a Diagnosis Probe to Estimate the Distribution of Multi-Walled Carbon
Nanotubes in Poly(Lactic Acid)/Polypropylene Conductive Composites. Sens. Actuators B
Chem. 2018, 255, 2809–2819. DOI: 10.1016/j.snb.2017.09.098.
180. Patel, S. V.; Cemalovic, S.; Tolley, W. K.; Hobson, S. T.; Anderson, R.; Fruhberger, B.
Implications of Thermal Annealing on the Benzene Vapor Sensing Behavior of PEVA-
Graphene Nanocomposite Threads. ACS Sens. 2018, 3, 640–647. DOI: 10.1021/acssensors.
7b00912.
181. Gao, J.; Wang, H.; Huang, X.; Hu, M.; Xue, H.; Li, R. K. Y. A Super-Hydrophobic and
Electrically Conductive Nanofibrous Membrane for a Chemical Vapor Sensor. J. Mater.
Chem. A 2018, 6, 10036–10047. DOI: 10.1039/C8TA02356E.
182. Marriam, I.; Wang, X.; Tebyetekerwa, M.; Chen, G.; Zabihi, F.; Pionteck, J.; Peng, S.;
Ramakrishna, S.; Yang, S.; Zhu, M. A Bottom-up Approach to Design Wearable and
Stretchable Smart Fibers with Organic Vapor Sensing Behaviors and Energy Storage
Properties. J. Mater. Chem. A 2018, 6, 13633–13643. DOI: 10.1039/C8TA03262A.
183. Chiou, J. C.; Wu, C. C.; Huang, Y. C.; Chang, S. C.; Lin, T. M. Effects of Operating
Temperature on Droplet Casting of Flexible Polymer/Multi-Walled Carbon Nanotube
Composite Gas Sensors. Sensors 2016, 17, 4. DOI: 10.3390/s17010004.
184. Liu, H.; Huang, W.; Yang, X.; Dai, K.; Zheng, G.; Liu, C.; Shen, C.; Yan, X.; Guo, J.; Guo,
Z. Organic Vapor Sensing Behaviors of Conductive Thermoplastic Polyurethane-Graphene
Nanocomposites. J. Mater. Chem. C 2016, 4, 4459–4469. DOI: 10.1039/C6TC00987E.
185. Poetschke, P.; Andres, T.; Villmow, T.; Pegel, S.; Bruenig, H.; Kobashi, K.; Fischer, D.;
Haeussler, L. Liquid Sensing Properties of Fibres Prepared by Melt Spinning from
Poly(Lactic Acid) Containing Multi-Walled Carbon Nanotubes. Compos. Sci. Technol.
2010, 70, 343–349. DOI: 10.1016/j.compscitech.2009.11.005.
186. Pioggia, G.; Francesco, F. D.; Ferro, M.; Sorrentino, F.; Salvo, P.; Ahluwalia, A.
Characterization of a Carbon Nanotube Polymer Composite Sensor for an Impedimetric
Electronic Tongue. Microchim. Acta 2008, 163, 57–62. DOI: 10.1007/s00604-008-0952-y.
187. Poetschke, P.; Kobashi, K.; Villmow, T.; Andres, T.; Paiva, M. C.; Covas, J. A. Liquid
Sensing Properties of Melt Processed Polypropylene/Poly(Epsilon-Caprolactone) Blends
Containing Multiwalled Carbon Nanotubes. Compos. Sci. Technol. 2011, 71, 1451–1460.
DOI: 10.1016/j.compscitech.2011.05.019.
188. Villmow, T.; Pegel, S.; John, A.; Rentenberger, R.; P€ otschke, P. Liquid Sensing: smart
Polymer/CNT Composites. Mater. Today 2011, 14, 340–345. DOI: 10.1016/S1369-
7021(11)70164-X.
189. Villmow, T.; Pegel, S.; P€
otschke, P.; Heinrich, G. Polymer/Carbon Nanotube Composites
for Liquid Sensing: Model for Electrical Response Characteristics. Polymer 2011, 52,
2276–2285. DOI: 10.1016/j.polymer.2011.03.029.
190. Rentenberger, R.; Cayla, A.; Villmow, T.; Jehnichen, D.; Campagne, C.; Rochery, M.;
Devaux, E.; P€ otschke, P. Multifilament Fibres of Poly(Epsilon-Caprolactone)/Poly(Lactic
Acid) Blends with Multiwalled Carbon Nanotubes as Sensor Materials for Ethyl Acetate
and Acetone. Sens. Actuators B Chem. 2011, 160, 22–31.
191. Xu, Z.; Wang, N.; Li, N.; Zheng, G.; Dai, K.; Liu, C.; Shen, C. Liquid Sensing Behaviors of
Conductive Polypropylene Composites Containing Hybrid Fillers of Carbon Fiber and
Carbon Black. Compos. Part B Eng. 2016, 94, 45–51. DOI: 10.1016/j.compositesb.2016.03.047.
POLYMER REVIEWS 191

192. Koratkar, N.; Modi, A.; Lass, E.; Ajayan, P. Temperature Effects on Resistance of Aligned
Multiwalled Carbon Nanotube Films. J. Nanosci. Nanotech. 2004, 4, 744–748. DOI: 10.
1166/jnn.2004.109.
193. Miaudet, P.; Bartholome, C.; Derre, A.; Maugey, M.; Sigaud, G.; Zakri, C.; Poulin, P.
Thermo-Electrical Properties of PVA-Nanotube Composite Fibers. Polymer 2007, 48,
4068–4074. DOI: 10.1016/j.polymer.2007.05.028.
194. Xiang, Z.-D.; Chen, T.; Li, Z.-M.; Bian, X.-C. Negative Temperature Coefficient of
Resistivity in Lightweight Conductive Carbon Nanotube/Polymer Composites. Macromol.
Mater. Eng. 2009, 294, 91–95. DOI: 10.1002/mame.200800273.
195. Rybak, A.; Boiteux, G.; Melis, F.; Seytre, G. Conductive Polymer Composites Based on
Metallic Nanofiller as Smart Materials for Current Limiting Devices. Compos. Sci. Technol.
2010, 70, 410–416. DOI: 10.1016/j.compscitech.2009.11.019.
196. Zhao, S.; Lou, D.; Zhan, P.; Li, G.; Dai, K.; Guo, J.; Zheng, G.; Liu, C.; Shen, C.; Guo, Z.
Heating-Induced Negative Temperature Coefficient Effect in Conductive Graphene/
Polymer Ternary Nanocomposites with a Segregated and Double-Percolated Structure. J.
Mater. Chem. C 2017, 5, 8233–8242. DOI: 10.1039/C7TC02472J.
197. Cui, C.-H.; Pang, H.; Yan, D.-X.; Jia, L.-C.; Jiang, X.; Lei, J.; Li, Z.-M. Percolation and
Resistivity-Temperature Behaviours of Carbon Nanotube-Carbon Black Hybrid Loaded
Ultrahigh Molecular Weight Polyethylene Composites with Segregated Structures. RSC
Adv. 2015, 5, 61318–61323. DOI: 10.1039/C5RA08847J.
198. Sajid, M.; Gul, J. Z.; Kim, S. W.; Kim, H. B.; Na, K. H.; Choi, K. H. Development of 3D-
Printed Embedded Temperature Sensor for Both Terrestrial and Aquatic Environmental
Monitoring Robots”, 3D Print. Addit. Manuf. 2018, 5, 160–169. DOI: 10.1089/3dp.2017.0092.
199. Jasmi, F.; Azeman, N. H.; Bakar, A. A. A.; Zan, M. S. D.; Haji Badri, K.; Su’ait, M. S.
Ionic Conductive Polyurethane-Graphene Nanocomposite for Performance Enhancement
of Optical Fiber Bragg, Grating Temperature Sensor. IEEE Access 2018, 6, 47355–47363.
DOI: 10.1109/ACCESS.2018.2867220.
200. Hou, Y-l.; Zhang, P.; Xie, M-m. Thermally Induced Double-Positive Temperature
Coefficients of Electrical Resistivity in Combined Conductive Filler-Doped Polymer
Composites. J. Appl. Polym. Sci. 2017, 134, 44876. DOI: 10.1002/app.44876.
201. Zhang, P.; Wang, B-b. Positive Temperature Coefficient Effect and Mechanism of
Compatible LLDPE/HDPE Composites Doping Conductive Graphite Powders. J. Appl.
Polym. Sci. 2018, 135, 46453. DOI: 10.1002/app.46453.
202. Lai, F.; Wang, B.-B.; Zhang, P. Enhanced Positive Temperature Coefficient in Amorphous
PS/CSPE-MWCNT Composites with Low Percolation Threshold. J. Appl. Polym. Sci.
2019, 136, 47053. DOI: 10.1002/app.47053.
203. Yurddaskal, M.; Erol, M.; Celik, E. Carbon Black and Graphite Filled Conducting
Nanocomposite Films for Temperature Sensor Applications. J. Mater. Sci. Mater. Electron.
2017, 28, 9514–9518. DOI: 10.1007/s10854-017-6695-y.
204. Zhang, P.; Hou, Y.; Wang, B. VO2-Enhanced Double Positive Temperature Coefficient
Effects of High Density Polyethylene/Graphite Composites. Mater. Res. Express 2018, 6,
035702. DOI: 10.1088/2053-1591/aaf589.
205. Li, M.; Wang, Y.; Zhang, Y.; Zhou, H.; Huang, Z.; Li, D. Highly Flexible and Stretchable
MWCNT/HEPCP Nanocomposites with Integrated near-IR, Temperature and Stress
Sensitivity for Electronic Skin. J. Mater. Chem. C 2018, 6, 5877–5887. DOI: 10.1039/
C8TC01331D.
206. Zhao, S.; Li, G.; Liu, H.; Dai, K.; Zheng, G.; Yan, X.; Liu, C.; Chen, J.; Shen, C.; Guo, Z.
Positive Temperature Coefficient (PTC) Evolution of Segregated Structural Conductive
Polypropylene Nanocomposites with Visually Traceable Carbon Black Conductive
Network. Adv. Mater. Interfaces 2017, 4, 1700265. DOI: 10.1002/admi.201700265.
207. Zhao, X.; Long, Y.; Yang, T.; Li, J.; Zhu, H. Simultaneous High Sensitivity Sensing of
Temperature and Humidity with Graphene Woven Fabrics. ACS Appl. Mater. Interfaces
2017, 9, 30171–30176. DOI: 10.1021/acsami.7b09184.
192 J. CHEN ET AL.

208. Zhou, X.; Zhu, L.; Fan, L.; Deng, H.; Fu, Q. Fabrication of Highly Stretchable, Washable,
Wearable, Water-Repellent Strain Sensors with Multi-Stimuli Sensing Ability. ACS Appl.
Mater. Interfaces 2018, 10, 31655–31663. DOI: 10.1021/acsami.8b11766.
209. Zou, H-z.; Zhang, X.; Zheng, S-d.; Yang, W.; Liu, Z-y.; Yang, M-b.; Feng, J-m. PVDF/CF
Conductive Composites with High Sensitivity and Stable Reproducibility of Positive
Temperature Coefficient Effect. Acta Polym. Sin. 2017, 8, 1215–1219. DOI: 10.1016/j.com-
positesa.2016.12.001.
210. Cui, X.; Chen, J.; Zhu, Y.; Jiang, W. Lightweight and Conductive Carbon Black/
Chlorinated Poly(Propylene Carbonate) Foams with a Remarkable Negative Temperature
Coefficient Effect of Resistance for Temperature Sensor Applications. J. Mater. Chem. C
2018, 6, 9354–9362. DOI: 10.1039/C8TC02123F.
211. Giurgiutiu, V.; Zagrai, A.; Bao, J. J. Piezoelectric Wafer Embedded Active Sensors for
Aging Aircraft Structural Health Monitoring. Struct. Health Monit. 2002, 1, 41–61. DOI:
10.1177/147592170200100104.
212. Kang, I.; Schulz, M. J.; Kim, J. H.; Shanov, V.; Shi, D. A Carbon Nanotube Strain Sensor
for Structural Health Monitoring. Smart Mater. Struct. 2006, 15, 737–748. DOI: 10.1088/
0964-1726/15/3/009.
213. Burton, A. R.; Lynch, J. P.; Kurata, M.; Law, K. H. Fully Integrated Carbon Nanotube
Composite Thin Film Strain Sensors on Flexible Substrates for Structural Health
Monitoring. Smart Mater. Struct. 2017, 26, 095052. DOI: 10.1088/1361-665X/aa8105.
214. Amjadi, M.; Pichitpajongkit, A.; Lee, S.; Ryu, S.; Park, I. Highly Stretchable and Sensitive
Strain Sensor Based on Silver Nanowire-Elastomer Nanocomposite. ACS Nano 2014, 8,
5154–5163. DOI: 10.1021/nn501204t.
215. Helmer, R. J. N.; Farrow, D.; Ball, K.; Phillips, E.; Farouil, A.; Blanchonette, I. A Pilot
Evaluation of an Electronic Textile for Lower Limb Monitoring and Interactive
Biofeedback. In 5th Asia-Pacific Congress on Sports Technology, Subic, A., Fuss, F. K.,
Alam, F. and Clifton, P. Eds., 2011, pp. 513–518DOI: 10.1016/j.proeng.2011.05.123.
216. Giorgino, T.; Tormene, P.; Lorussi, F.; De Rossi, D.; Quaglini, S. Sensor Evaluation for
Wearable Strain Gauges in Neurological Rehabilitation. IEEE Trans. Neural Syst. Rehabil.
Eng. 2009, 17, 409–415. DOI: 10.1109/TNSRE.2009.2019584.
217. Lorussi, F.; Scilingo, E. P.; Tesconi, M.; Tognetti, A.; De Rossi, D. Strain Sensing Fabric
for Hand Posture and Gesture Monitoring. IEEE Trans. Inform. Technol. Biomed. 2005, 9,
372–381. DOI: 10.1109/TITB.2005.854510.
218. Liu, C.-X.; Choi, J.-W. Patterning Conductive PDMS Nanocomposite in an Elastomer
Using Microcontact Printing. J. Micromech. Microeng. 2009, 19, 085019. DOI: 10.1088/
0960-1317/19/8/085019.
219. Xiao, X.; Yuan, L.; Zhong, J.; Ding, T.; Liu, Y.; Cai, Z.; Rong, Y.; Han, H.; Zhou, J.; Wang,
Z. L. High-Strain Sensors Based on ZnO Nanowire/Polystyrene Hybridized Flexible Films.
Adv. Mater. 2011, 23, 5440–5444. DOI: 10.1002/adma.201103406.
220. McEvoy, M. A.; Correll, N. Materials That Couple Sensing, Actuation, Computation, and
Communication. Science 2015, 347, 1261689–1261689. DOI: 10.1126/science.1261689.
221. Majidi, C. Soft Robotics: A Perspective-Current Trends and Prospects for the Future. Soft
Robot 2014, 1, 5–11. DOI: 10.1089/soro.2013.0001.
222. Cai, L.; Song, L.; Luan, P.; Zhang, Q.; Zhang, N.; Gao, Q.; Zhao, D.; Zhang, X.; Tu, M.;
Yang, F.; et al. Super-Stretchable, Transparent Carbon Nanotube-Based Capacitive Strain
Sensors for Human Motion Detection. Sci. Rep. 2013, 3, 3048DOI: 10.1038/srep03048.
223. Cohen, D. J.; Mitra, D.; Peterson, K.; Maharbiz, M. M. A Highly Elastic, Capacitive Strain
Gauge Based on Percolating Nanotube Networks. Nano Lett. 2012, 12, 1821–1825. DOI:
10.1021/nl204052z.
224. Yao, S.; Zhu, Y. Wearable Multifunctional Sensors Using Printed Stretchable Conductors
Made of Silver Nanowires. Nanoscale 2014, 6, 2345–2352. DOI: 10.1039/c3nr05496a.
225. Lipomi, D. J.; Vosgueritchian, M.; Tee, B. C.; Hellstrom, S. L.; Lee, J. A.; Fox, C. H.; Bao,
Z. Skin-like Pressure and Strain Sensors Based on Transparent Elastic Films of Carbon
Nanotubes. Nature Nanotech. 2011, 6, 788–792. DOI: 10.1038/nnano.2011.184.
POLYMER REVIEWS 193

226. Zhang, S.; Liu, H.; Yang, S.; Shi, X.; Zhang, D.; Shan, C.; Mi, L.; Liu, C.; Shen, C.; Guo, Z.
Ultrasensitive and Highly Compressible Piezoresistive Sensor Based on Polyurethane
Sponge Coated with a Cracked Cellulose Nanofibril/Silver Nanowire Layer. ACS Appl.
Mater. Interfaces 2019, 11, 10922–10932. DOI: 10.1021/acsami.9b00900.
227. Duan, L.; D’Hooge, D.; R.; Spoerk, M.; Cornillie, P.; Cardon, L. Facile and Low-Cost
Route for Sensitive Stretchable Sensors by Controlling Kinetic and Thermodynamic
Conductive Network Regulating Strategies. ACS Appl. Mater. Interfaces 2018, 10,
22678–22691. DOI: 10.1021/acsami.8b03967.
228. Ji, M.; Deng, H.; Yan, D.; Li, X.; Duan, L.; Fu, Q. Selective Localization of Multi-Walled
Carbon Nanotubes in Thermoplastic Elastomer Blends: An Effective Method for Tunable
Resistivity–Strain Sensing Behavior. Compos. Sci. Technol. 2014, 92, 16–26. DOI: 10.1016/
j.compscitech.2013.11.018.
229. Lin, Y.; Liu, S.; Chen, S.; Wei, Y.; Dong, X.; Liu, L. A Highly Stretchable and Sensitive Strain
Sensor Based on Graphene–Elastomer Composites with a Novel Double-Interconnected
Network. J. Mater. Chem. C 2016, 4, 6345–6352. DOI: 10.1039/C6TC01925K.
230. Wang, M.; Zhang, K.; Dai, X. X.; Li, Y.; Guo, J.; Liu, H.; Li, G. H.; Tan, Y. J.; Zeng, J. B.;
Guo, Z. Enhanced Electrical Conductivity and Piezoresistive Sensing in Multi-Wall
Carbon Nanotubes/Polydimethylsiloxane Nanocomposites via the Construction of a Self-
Segregated Structure. Nanoscale 2017, 9, 11017–11026. DOI: 10.1039/C7NR02322G.
231. Wang, S.; Zhang, X.; Wu, X.; Lu, C. Tailoring Percolating Conductive Networks of
Natural Rubber Composites for Flexible Strain Sensors via a Cellulose Nanocrystal
Templated Assembly. Soft Matter 2016, 12, 845–852. DOI: 10.1039/C5SM01958C.
232. Chen, J.; Zhu, Y.; Jiang, W. A Stretchable and Transparent Strain Sensor Based on
Sandwich-like PDMS/CNTs/PDMS Composite Containing an Ultrathin Conductive CNT
Layer. Compos. Sci. Technol. 2020, 186, 107938. DOI: 10.1016/j.compscitech.2019.107938.
233. Cho, S. H.; Lee, S. W.; Yu, S.; Kim, H.; Chang, S.; Kang, D.; Hwang, I.; Kang, H. S.;
Jeong, B.; Kim, E. H.; et al. Micropatterned Pyramidal Ionic Gels for Sensing Broad-Range
Pressures with High Sensitivity. ACS Appl. Mater. Interfaces 2017, 9, 10128–10135. DOI:
10.1021/acsami.7b00398.
234. Mannsfeld, S. C. B.; Tee, B. C. K.; Stoltenberg, R. M.; Chen, C. V. H. H.; Barman, S.;
Muir, B. V. O.; Sokolov, A. N.; Reese, C.; Bao, Z. Highly Sensitive Flexible Pressure
Sensors with Microstructured Rubber Dielectric Layers. Nature Mater. 2010, 9, 859–864.
DOI: 10.1038/nmat2834.
235. Wang, X.; Gu, Y.; Xiong, Z.; Cui, Z.; Zhang, T. Silk-Molded Flexible, Ultrasensitive, and
Highly Stable Electronic Skin for Monitoring Human Physiological Signals. Adv. Mater.
2014, 26, 1336–1342. DOI: 10.1002/adma.201304248.
236. Dong, X. M.; Luo, Y.; Xie, L. N.; Fu, R. W.; Zhang, M. Q. Conductive Carbon Black-
Filled Polymethacrylate Composites as Gas Sensing Materials: Effect of Glass Transition
Temperature. Thin Solid Films 2008, 516, 7886–7890. DOI: 10.1016/j.tsf.2008.06.003.
237. Kennedy, Z. C.; Christ, J. F.; Evans, K. A.; Arey, B. W.; Sweet, L. E.; Warner, M. G.;
Erikson, R. L.; Barrett, C. A. 3D-Printed Poly(Vinylidene Fluoride)/Carbon Nanotube
Composites as a Tunable, Low-Cost Chemical Vapour Sensing Platform. Nanoscale 2017,
9, 5458–5466. DOI: 10.1039/C7NR00617A.
238. Hansen, C. M. Hansen Solubility Parameters; CRC Press: New York, 2000.
239. Jeon, J.; Lee, H.-B.-R.; Bao, Z. Flexible Wireless Temperature Sensors Based on Ni
Microparticle-Filled Binary Polymer Composites. Adv. Mater. 2013, 25, 850–855. DOI: 10.
1002/adma.201204082.
240. Liu, F.; Zhang, X.; Li, W.; Cheng, J.; Tao, X.; Li, Y.; Sheng, L. Investigation of the
Electrical Conductivity of HDPE Composites Filled with Bundle-like MWNTs. Compos.
Part A-Appl. S 2009, 40, 1717–1721. DOI: 10.1016/j.compositesa.2009.08.004.
241. Ho, D. H.; Sun, Q.; Kim, S. Y.; Han, J. T.; Kim, D. H.; Cho, J. H. Stretchable and
Multimodal All Graphene Electronic Skin. Adv. Mater. Weinheim. 2016, 28, 2601–2608.
DOI: 10.1002/adma.201505739.

You might also like