Sei sulla pagina 1di 17

Article

0(00) 1–17
! The Author(s) 2018

All-fabric flexible Article reuse guidelines:


sagepub.com/journals-permissions
DOI: 10.1177/1528083718804208
supercapacitor for journals.sagepub.com/home/jit

energy storage

Weihua Luo1,2, Xinxin Li1,3 and Jonathan Y Chen1

Abstract
An all-fabric solid-state flexible supercapacitor has been fabricated using three types of
commercial woven fabrics made of carbon fiber, activated carbon fiber, and polyester
fiber, respectively. The activated carbon fiber fabric is viscose-based twill through car-
bonization and activation, followed by a deposition of CeO2 or ZnO nanoparticles
through a hydrothermal process. The resultant fabric supercapacitor displays a high
specific capacitance of 13.24 mF cm2 at a scan rate of 0.2 mV s1, an excellent capaci-
tance retention of 87.6% at 5000 charge/discharge cycles, and a high energy density of
4.6  107 Wh cm2 at a power density of 3.31  106 W cm2. The supercapacitor is
capable of bending in any angles without losing its performance, revealing an excellent
flexibility. The facile, cost-effective fabrication process and excellent electrochemical
performance allow this all-fabric solid-state flexible supercapacitor to be potentially
used in new generation of wearable and self-powered electronic devices.

Keywords
Specialty fabrics, technical textiles, carbon fabrics, structure properties

Introduction
Flexible solid-state supercapacitors have attracted an increasing interest for their
potential applications in wearable electronic devices, hybrid electric vehicles and
portable electronic devices such as mobile phones, e-readers, tablets, laptops, etc.
[1]. Compared with conventional supercapacitors, flexible electrochemical super-
capacitors possess several important advantages, such as flexibility, lightweight,
and high safety. Being bendable, twistable, rollable, and foldable, they can take

1
School of Human Ecology, The University of Texas at Austin, Austin, TX, USA
2
College of Material Science and Engineering, Central South University of Forestry and Technology, China
3
Engineering Research Center for Knitting Technology, Ministry of Education, Jiangnan University, China

Corresponding author:
Jonathan Y Chen, School of Human Ecology, The University of Texas at Austin, Austin, TX 78712, USA.
Email: jychen2@austin.utexas.edu
2 Journal of Industrial Textiles 0(00)

advantage of the space that may be inaccessible to the conventional supercapaci-


tors [2]. For the purpose of fabricating high-performance flexible supercapacitors,
it is very important to seek new types of flexible capacitive materials with high
capacitance and to develop novel fabrication approaches.

Flexible capacitive materials


Recently, many carbon nanomaterials including carbon particles [3, 4], carbon
nanotubes (CNT) [5, 6], carbon nanofibers [7, 8], and graphene [9, 10] have been
utilized as electrochemical active materials and fabricated into flexible electrodes.
To enhance the capacitance of flexible supercapacitors, pseudo-capacitive materials
such as transition or lanthanide metal oxides such as MnO2 [11], RuO2 [12], Co3O4
[13], CeO2 [14, 15], or ZnO [16, 17], and conducting polymers such as polyaniline
(PANI) [18], polypyrrole (PPy) [19], or poly(3,4-ethylenedioxythiophene) (PEDOT)
[20], have been fabricated into the electrodes, usually as composites with carbon
materials. For example, flexible supercapacitors fabricated by PPy/graphene oxide
(GO)/ZnO nanocomposites exhibit a specific capacitance of 94.6 F g1 at a current
density of 1 A g1. The application of pseudo-capacitive materials PPy and
ZnO renders a high energy density of 10.65 Wh kg1 and power density of
258.26 W kg1 at 1 A g1 [17].

Fabrication methods of flexible supercapacitors


Fabrication methods play an important role in the performance of flexible super-
capacitors. A main approach for fabricating flexible electrodes is preparing a free-
standing film [21–25] or fiber [26–29] from electrochemically active materials, which
can be fabricated into thin flexible supercapacitors. For instance, a free-standing and
flexible MoS2/GO hybrid film prepared via vacuum filtration presents a high specific
capacitance of 380 F cm3 at 10 mV s1 [21]. A symmetric supercapacitor fabricated
by N/O-enriched carbon cloth delivers a high energy density of 9.4 mW h cm3 [22].
Flexible, free-standing, and porous carbon films in the absence of conducting addi-
tives and binders have been developed by activation of reduced graphene-oxide films
with KOH. Flexible supercapacitors fabricated using these films as electrodes reveal
a very high power density of about 500 kW kg1 with a reasonably high energy
density of 26 Wh kg1 [24]. Besides the planar type electrochemical device, fiber-
shaped supercapacitors fabricated by freestanding CNT fibers [26, 27] or graphene
fibers [28, 29] also render excellent capacitive performance.
Another attractive approach for fabricating flexible electrodes is the deposition
of carbon nanomaterials, metal oxides, and conductive polymers onto a
flexible support including plastics, papers, and textile fibers [30–32]. Polyethylene
terephthalate (PET) film is a relatively common use of a substrate due to its flexi-
bility and excellent mechanical properties. The reduced GO and CNT composites
[33], CNT/PANI nanocomposite [34], nanostructured a-MnO2 [35] have been
deposited on PET substrate to obtain flexible electrodes. Cellulose printing
Luo et al. 3

papers have also been used as a substrate for their porous structure, rough surface
properties, and low price [2, 36, 37]. For being able to be knitted into textiles, a
number of fiber-shaped supercapacitors have been fabricated by utilizing carbon
fiber, cotton thread [38], nanocellulose fibers [39, 40], and TiO2 fibers [41] as sub-
strates. However, these fabrication approaches involve either the high cost of
the capacitive materials or the elaborate process, not favorable for mass produc-
tion [42, 43]. Therefore, it is essential for future research with an aim at
increasing electrochemical performance as well as lowering fabrication cost of
the supercapacitors.

State of current work


In this paper, we present a carbon-based all-fabric flexible supercapacitor based on
an activated carbon fiber fabric, a PET fabric, and a carbon fiber fabric that are
used as electrodes, separators, and current collectors, respectively. The commercial
viscose fiber woven fabrics are firstly carbonized and activated into capacitive
fabrics with high porosity. Then CeO2 or ZnO nanoparticles, which are known
as pseudo-capacitive materials and possess lower cost compared to CNT and gra-
phene, are deposited on the activated carbon fiber fabric through a hydrothermal
process. The resultant all-fabric solid-state flexible supercapacitor delivers a high
device capacitance (13.24 mF cm2), excellent cycle stability (capacitance retention
of 87.6% at 5000 cycles), high power and energy density (4.6  107 Wh cm2 at a
power density of 3.31  106 W cm2). Moreover, this facile, cost-effective and
binder-free fabrication process is favorable for commercial production, showing
great potential for use in new generation self-powered electronic devices.

Materials and experimental methods


Raw materials and modification
The activated carbon fiber fabric used in this study was produced from a viscose
twill fabric precursor (denoted as VF) by carbonization and activation and is com-
mercially available. This activated viscose fabric was labeled as ACVF. It had a
fabric areal density of 154 g m2 and fabric thickness of 0.432 mm. The carbon fiber
fabric was a commercial plain woven fabric with fabric areal density of 212 g m2
and thickness of 0.254 mm. The separator layer in the fabric supercapacitor was a
commercial lightweight PET mesh fabric with fabric areal density of 44 g m2 and
thickness of 0.120 mm.
Two different suspensions were prepared by adding 10% CeO2 and ZnO nano-
particles by weight of the ACVF fabric into water. After ultrasonic dispersion of
the suspension for 30 min, the ACVF sample was immersed in the suspension.
The system was further ultrasonically dispersed for 30 min and then vacuum
dried at 60 C for 24 h. Once dried, the CeO2- and ZnO-modified ACVF fabric
samples were obtained and labeled as CeO2-ACVF and ZnO-ACVF, respectively.
4 Journal of Industrial Textiles 0(00)

Fabrication of all-fabric supercapacitors


Five grams of PVA was slowly dissolved in 50 ml distilled water with agitation at
95 C until a clear homogeneous mixture was achieved. Then 5 g H3PO4 was added,
and the resulting solution was stirred continuously for 1 h. The mixture was used as
electrolyte. Two pieces of ACVF were immersed in the electrolyte solution for 24 h.
Then they were taken out and assembled into an all-fabric supercapacitor as elec-
trodes, together with a piece of the PET fabric as a separator and two pieces of the
carbon fiber woven fabric as current collectors.

Characterization
Nitrogen adsorption and desorption isotherms were measured at 77 K using a
Micromeritics 3Flex automatic system. The specific surface area was determined via
the Brunauer-Emmett-Teller (BET) method, and the pore size distribution (PSD) was
estimated according to Horvath-Kawazoe model on the basis of slit-like pore geom-
etry. The BET theory is widely used for the calculation of surface areas of solids by
physical adsorption of gas molecules that do not chemically react with material sur-
faces. Nitrogen is the generally utilized gaseous adsorbate. The morphologies of the
surface and cross-section of the ACVFs were observed by scanning electron micros-
copy (SEM) (Hitachi S5500), and the chemical element analysis of materials was also
performed using this instrument equipped with energy dispersive spectroscopy (EDS).
The electrochemical measurements were performed 3 h after the assembling of the all-
fabric supercapacitors. Cyclic voltammetry (CV) measurements were carried out using
an Autolab instrument (PGSTAT 128 N) from 0 to 1.0 V. Electrochemical impedance
spectra (EIS) were measured using this instrument in a frequency range of 0.01 Hz to
100 kHz. Galvanostatic charge-discharge (GCD) tests were performed using a BT2000
Battery tester (Arbin Instruments) between 0 and 1.0 V.

Results and discussion


Porosity and morphology of materials
The schematic procedure of electrode fabrication is illustrated in Figure 1. Firstly,
the precursor VFs were carbonized and activated into ACVFs to obtain great
porosity in the fibers. The N2 adsorption/desorption isotherm of ACVF is shown
in Figure 2. N2 uptake increases sharply at low relative pressure and approaches a
plateau at high relative pressure. This suggests that the micropores are formed in
ACVF [44]. Moreover, no apparent adsorption/desorption hysteresis loop is
observed. This suggests that the isotherm of ACVF is typical type I and the micro-
pores are dominating in ACVF. The inserted plot presents the PSD graph of
ACVF. The abbreviated term DV/DW along the Y-axis of the graph represents
differential pore volume. The PSD results reveal that ACVF consists mostly of
micropores with a pore size range of 0.4–1.8 nm and two peaks at 0.46 and
0.54 nm respectively. The BET surface area of ACVF reaches 1179 m2 g1.
Luo et al. 5

Figure 1. Schematic procedure of electrode fabrication.

Figure 2. N2 adsorption/desorption isotherm of activated viscose fabric with insertion of pore


size distribution curve.

The morphology of the fiber surface in the ACVF sample is shown in


Figure 3(a). The diameter of the fiber is approximately 7–9 mm. The cross-section
of the fiber under larger magnification in Figure 3(b) shows a very uneven, rough
surface, which can be ascribed to its great porosity. The EDS analysis of the cross-
section of the fiber is shown in the inserted graph. The two main peaks including C
and O appear in the range of binding energies from 0 to 1000 eV, suggesting that C
and O are the major chemical elements that constitute the ACVF sample.
By calculating the area of each peak, C element holds a dominant position.
The carbon enrichment can be attributed to the carbonization and activation pro-
cess at 800 C [44]. Figure 3(c) and (d) present the SEM images of CeO2-ACVF and
ZnO-ACVF samples, respectively. As shown in Figure 3(c) and (d), a layer of CeO2
6 Journal of Industrial Textiles 0(00)

Figure 3. Scanning electron microscopy images of (a) the activated viscose fabric (ACVF) sur-
face; (b) the ACVF cross-section with an insertion of energy dispersive spectroscopy analysis; (c)
the surface of CeO2-ACVF fiber; and (d) the surface of ZnO-ACVF fiber.

and ZnO are deposited on the surface of fibers, respectively through the hydro-
thermal process.

Structure and electrochemical performance of supercapacitors


Activated carbon has been widely used as an electrode material for supercapacitors
because of its high porosity. In most cases, activated carbon is used as powder or
fiber particle form [45–47]. In this case, the commercial fabric VF was carbonized
and activated into ACVF with high porosity, and then was directly used as two
electrodes to form an electrostatic type (Helmholtz Layer only) supercapacitor.
The CeO2-ACVF and ZnO-ACVF fabrics were also used individually as electrodes
to form two electrochemical type (redox reaction) supercapacitors. These three all-
fabric flexible supercapacitors were all constructed using the commercial PET mesh
fabric as a separator and the conductive carbon woven fabric as two current col-
lectors. The construction of the all-fabric supercapacitors is illustrated in Figure 4.
To assess the electrochemical performance of the all-fabric supercapacitors, CV,
EIS and GC measurements were conducted. Figure 5(a) shows the CV curves of the
ACVF supercapacitor that use the ACVF as electrodes at different scan rates. The
CV loops exhibit a quasi-rectangular shape indicating the capacitive behavior of
double-layer electrostatic capacitor. Figure 5(b) and (c) present test results of the
Luo et al. 7

Figure 4. Structure of the all-fabric supercapacitor.

supercapacitors using the CeO2-ACVF and ZnO-ACVF as electrodes. Their CV


curves reveal an increasingly enlarged loop upon an increase of the scan rates. For
the purpose of comparison, the CV curves of the three supercapacitors at the scan
rate of 20 mV s1 are shown in Figure 5(d). Apparently, the curves of CeO2-ACVF
and ZnO-ACVF exhibit a more rectangular-like shape and a larger size than that of
ACVF, indicating that both the CeO2-ACVF and ZnO-ACVF supercapacitors
possess a better capacitive behavior. Because both CeO2 and ZnO are electric
active materials, a pseudofaradic reaction may take place during the charging
and discharging process [48], as described below

CeO2 þ Hþ þe ! CeOOH ð1Þ

ZnO2 þ Hþ þ e ! ZnOOH ð2Þ

From Figures 5(b) and (c), a remarkable redox peak at 0.3 V and a less distinct one
at 0.2 V are exhibited in the CV curves of ZnO-ACVF and CeO2-ACVF, respectively.
The improvement in capacity behavior of the CeO2-ACVF and ZnO-ACVF super-
capacitors should be due to the reaction between CeO2 or ZnO nanoparticles and the
electrolyte. Although both CeO2 and ZnO nanoparticles indicate the improved cap-
acitive performance of the ACVF supercapacitor for their pseudo-capacitive
8 Journal of Industrial Textiles 0(00)

Figure 5. Cyclic voltammetry curves of (a) activated viscose fabric (ACVF) supercapacitor; (b)
CeO2-ACVF supercapacitor; (c) ZnO-ACVF supercapacitor at different scan rates; and (d) all
supercapacitors at the scan rate of 20 mV s1. (e) Calculated specific capacitances. (f) Nyquist
plots of the supercapacitor electrochemical impedance spectra.

characteristics, the ZnO-ACVF supercapacitor shows a better capacitive behavior


than the CeO2-ACVF supercapacitor. A major reason is probably because ZnO is a
typical transition metal oxide having a better performance for redox reaction, while
CeO2 is not. This phenomenon needs a further investigation.
To compare the capacitance performance of different supercapacitors,
R  their spe-
cific capacitances need to be calculated via the equation Csp ¼ IdV =ðSVÞ,
where Csp is defined as the capacitance per unit area of electrode; I is the current; 
is the scan rate; V is the potential range of the CV scan; and S is the surface area
of the individual electrode [49]. In this case, the actual electrode surface area is
chosen to be 4 cm2. The results of the specific capacitance of the three
Luo et al. 9

supercapacitors are shown in Figure 5(e). The specific capacitance of ACVF is


9.62 mF cm2 at a scan rate of 2 mV s1, which was comparable to other solid-
state devices reported previously, for instance, TiO2 nanotubes (5.9 mF cm2 at
scan rate of 2 mV s1) [50] and CNT/paper (11.07 mF cm2 at scan rate of
10 mV s1) [51], and even sandwich structured fiber (11.88 mF cm2 at scan rate
of 10 mV s1) [52]. This indicates that the carbon-based active fabric material has
great potential for capacitor electrode application. Moreover, the CeO2-ACVF and
ZnO-ACVF electrodes display a larger specific capacitance in comparison with that
of ACVF. For instance, the specific capacitance of ZnO-ACVF increases to
13.24 mF cm2 at a scan rate of 2 mV s1. The enhancement may be attributed to
the pseudofaradic reaction between CeO2 or ZnO nanoparticles and the electrolyte.
For all three types of supercapacitors studied here, the specific capacitance
increases with a decrease in scan rate. In contrast, when the scanning rate increases,
the limited diffusion and migration of the ions causes the ACVF surface area
partially inaccessible to store charge, resulting in a lower specific capacitance [53,
54]. Interestingly, as the scanning rate increases, the specific capacitance of the both
CeO2-ACVF and ZnO-ACVF supercapacitors has a slower downward trend than
that of the ACVF supercapacitor. This may be explained that the faradic reaction,
which accounted for part of the capacitance of the CeO2-ACVF and ZnO-ACVF,
can be completed in a very short time [43], hence their capacitances are less affected
by the scanning rate.
To further evaluate the capacitive behavior of the ACVF-based supercapacitors,
EIS measurements were conducted at a frequency range from 0.01 Hz to 100 KHz.
As shown in Figure 5(f), the Nyquist plots exhibit an indistinct semicircle in the
high frequency region, reflecting a very low charge interfacial transfer resistance at
the electrode/electrolyte interface [38], which is benefited from the unique porous
and open structure of the ACVF. At the low frequencies, the sloping lines demon-
strate the capacitance characteristic. However, the Nyquist plots render a linear
line slightly inclining to the Z0 axis, which is corresponding to Warburg resistance,
the result of the frequency dependence of ion diffusion from electrolyte to elec-
trode/electrolyte interface [55, 56]. The equivalent series resistance (ESR) values
can be obtained at the first intercepting points of the EIS plots on Z0 axis. For the
ACVF supercapacitor, ESR value is about 9.80
. It is assumed that this large ESR
is mainly determined by the relatively poor conductivity of the insulated ACVF
electrodes. So it can be proposed that improving ACVF conductivity can reduce
ESR value and enhance overall performance of the ACVF supercapacitors.
Furthermore, a thick separator layer between two ACVF electrodes is also a
cause to increase ESR value. Through the inclusion of CeO2 and ZnO nanoparti-
cles into ACVF, the CeO2-ACVF and ZnO-ACVF supercapacitors exhibit lower
ESR values of 5.06
and 1.01
, respectively, suggesting that the deposition of
oxide on ACVF is capable of enhancing the electronic and ionic diffusion.
The capacitive performance of the all-fabric supercapacitors was also investi-
gated by the GCD measurements. Figure 6(a) shows the GCD curves of the super-
capacitors at a current density of 4  105 A cm2. Compared with the ACVF
10 Journal of Industrial Textiles 0(00)

Figure 6. Performance of charge/discharge, energy density, and retention. (a) GCD curves at a
current density of 4  105 A cm2; (b) CeO2-activated viscose fabric (ACVF) cycling perform-
ance at the current density of 4  105 A cm2 after 5000 charge/discharge cycles; (c) Ragone
plots; and (d) CeO2-ACVF specific capacitances after 3 h and 4 months of fabrication.

supercapacitor, both the CeO2-ACVF and ZnO-ACVF supercapacitors show a


longer discharge time, denoting a higher capacitance. This is consistent with the
CV results. Symmetry of GCD curves is related to capacitive behavior. For the
ACVF supercapacitor, the charge time and the discharge time are around 237 and
102 s, respectively, indicating an asymmetry of the GCD curve and a low charge
efficiency, which may be attributed to a low electrical conductivity of ACVF.
By incorporating CeO2 and ZnO nanoparticles into ACVF, the charge and dis-
charge time become 228 and 211 s for the CeO2-ACVF supercapacitor, and 323 and
256 s for the ZnO-ACVF supercapacitor, their charge/discharge curves become
fairly symmetric, demonstrating that both CeO2 and ZnO nanoparticles can
improve the charge efficiency of the ACVF supercapacitor.
The stability of supercapacitors is an important factor for commercial applica-
tions. The cycling performance of the CeO2-ACVF supercapacitor was studied by
measuring its charge–discharge curves for 5000 cycles. As shown in Figure 6(b), the
specific capacitance slowly decays to 87.6% of the initial value at the 3000th cycle
and remains almost constant for the remaining 2000 cycles. A recent research on a
nonfabric solid-state supercapacitor with a PVA-based electrolyte reported that its
capacitance retention after 3000 cycles of charge-discharge was around 90% [57].
In comparison to this, the present CeO2-ACVF supercapacitor also holds a good
Luo et al. 11

stability. It should be noted that this study did not include the stability test for the
ZnO-ACVF supercapacitor because of the time-consuming measurement.
Figure 6(c) presents the Ragone plots of different supercapacitors. The energy
densities of CeO2-ACVF and ZnO-ACVF are 3.0  107 Wh cm2 and
4.6  107 Wh cm2 while their power densities are 2.19  106 W cm2 and
3.31  106 W cm2, respectively. Generally, the energy densities increase with
decreasing power densities for all types of supercapacitors. The energy densities
of CeO2-ACVF and ZnO-ACVF are higher than that of ACVF especially for
higher power densities. These demonstrated that CeO2 and ZnO nanoparticles
can enhance the energy storing ability of the ACVF supercapacitor.
About 4 months after fabrication, the CV and EIS measurements were per-
formed again on the CeO2-ACVF supercapacitor. The comparison of the specific
capacitances of the supercapacitor at different scan rates before and after 4 months
is illustrated in Figure 6(d). After 4 months, the specific capacitances decrease from
5.6 and 4.24 mF cm2 to 4.14 and 2.47 mF cm2 at the scan rate of 10 and
20 mV s1, respectively. This may be because of moisture evaporation in the elec-
trolyte which can possibly be improved by utilizing other electrolyte like ionic
liquids. As a result, ESR of the supercapacitor increases.
Being a portable and wearable energy storage device, both flexibility and safety
are very important. Figure 7(a) presents the CV curves of the CeO2-ACVF super-
capacitor measured under various bending angles. The CV curves are in rectangu-
lar shape, and the sizes are nearly the same in various bending angles,
demonstrating that the deformation would not affect the capacitive performance
of the supercapacitor. As shown in Figure 7(b), the CeO2-ACVF supercapacitor is
able to be bent in any angles without losing its performance, showing a high flexible
property. The capacitive performance was also tested upon applied bending cycles.
As shown in Figure 7(c) and (d), the specific capacitance of the supercapacitor does
not decline but increases by 13.4% after 200 bending cycles, denoting that the
bending cycles also would not affect the capacitive performance of the supercapa-
citor. These results demonstrate that this carbon-based flexible fabric supercapa-
citor is suitable for being a portable and wearable energy storage device.

Integration of ACVF supercapacitor with solar cell


Solar energy is probably the most clean and renewable energy in the world, but it is
always limited by time, location, and weather. Integrating solar cells with energy
storage devices such as supercapacitors and batteries may offer a solution to build
stand-alone self-powered systems [16, 58, 59]. To demonstrate potential applica-
tions of the carbon-based all-fabric supercapacitors, a supercapacitor/solar cell
integration was formed by using four fabric supercapacitors and four solar cells
units each capable of charging a 1.2 V AA Ni-Cd rechargeable battery with a
capacity of 400 mAh. When sunlight was available, each supercapacitor was con-
nected with a solar cell, so the solar cell harvested solar energy and stored it in the
supercapacitor. After 120 s charging under the illumination, the voltage of each
12 Journal of Industrial Textiles 0(00)

Figure 7. Influence of bending deformation on CeO2-ACVF capacitive performance. (a) and (b)
CV curves and specific capacitance under various bending angles; (c) and (d): CV curves and
specific capacitances after different bending cycles. ACVF: activated viscose fabric.

Figure 8. The voltage profile for (a) the fabric-based supercapacitor when being charged by the
solar cell and (b) the device of four series supercapacitors in powering a red LED.

supercapacitor reached approximately 0.9–1.1 V (Figure 8(a)). Then the solar


source was turned off, and the four charged supercapacitors were connected to
power a commercial red light-emitting diode (LED). The four series supercapaci-
tors could illuminate the red LED for more than 130 min, and could still maintain a
voltage output around 1.61 V after 130 min (Figure 8(b)).
Luo et al. 13

Considering the flexibility of the fabric, it is feasible to integrate flexible solar


cells and the fabric-based supercapacitors into a self-powered system. This self-
powered device can adopt a flexible multilayer sheeting structure. Its area
and shape can be defined according to different end uses. Thus, the fabric super-
capacitor can be directly integrated into apparel, bag, tent, and other consumer
products, making these fabric-based products a mobile and portable solar power
supply. Furthermore, the technology does not need any new yarn spinning and
fabric-forming infrastructures. A simple short supply chain would be expected.

Summary
Overall, the facile, cost-effective carbon-based fabric supercapacitor exhibits high
capacitive performance with a specific capacitance of 13.24 mF cm2 at scan rate of
0.2 mV s1, a capacitance retention of 87.6% at 5000 charge/discharge cycles, and
an energy density of 4.6  107 Wh cm2 at power density of 3.31  106 W cm2.
The supercapacitor is susceptible to cyclic bending in 0–180 angles without losing
its performance. CeO2 and ZnO nanoparticles can enhance the energy storing
ability of the ACVF supercapacitor. Compared with the ACVF supercapacitor,
both the CeO2-ACVF and ZnO-ACVF supercapacitors show a longer discharge
time, denoting a higher capacitance.

Conclusion
The commercial woven viscose fabrics were firstly carbonized and activated into
capacitive fabrics, and then impregnated with CeO2 and ZnO nanoparticles individu-
ally through a hydrothermal process. The as-prepared activated carbon fiber fabrics
were utilized as electrodes to fabricate all-fabric solid-state flexible supercapacitors,
in which the commercial PET fabric and conductive carbon fiber fabrics were used as
separator and current collector, respectively. The as-constructed all-fabric superca-
pacitors possess a high specific capacitance, an excellent cycling retention of capaci-
tance, a high energy density and power density, and could maintain their
performance even under harsh bending conditions. They hold great potential for
many energy storage applications, such as smart clothing and mobile and portable
self-powered systems. The method to fabricate these carbon-based all-fabric flexible
supercapacitors is also facile and cost-effective, favorable for mass production.

Acknowledgements
The authors would like to acknowledge their thanks to China Scholarship Council for sponsoring
Dr Weihua Luo and Ms Xinxin Li as visiting scholars. Thanks are also extended to Micromeritics
Instrument Corporation for helping the test of ACVF surface and micropore properties.

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, author-
ship, and/or publication of this article.
14 Journal of Industrial Textiles 0(00)

Funding
The author(s) received no financial support for the research, authorship, and/or publication
of this article.

ORCID iD
Jonathan Y Chen http://orcid.org/0000-0001-6868-1824

References
[1] Yan X, Tai Z, Chen J, et al. Fabrication of carbon nanofiber–polyaniline composite
flexible paper for supercapacitor. Nanoscale 2011; 3: 212–216.
[2] Zhang YZ, Wang Y, Cheng T, et al. Flexible supercapacitors based on paper substrates:
A new paradigm for low-cost energy storage. Chem Soc Rev 2015; 44: 5181–5199.
[3] Lu X, Yu M, Wang G, et al. Flexible solid-state supercapacitors: Design, fabrication
and applications. Energy Environ Sci 2014; 7: 2160–2181.
[4] Zhang ZS, Zhai T, Lu XH, et al. Conductive membranes of EVA filled with carbon
black and carbon nanotubes for flexible energy-storage devices. J Mater Chem A 2013;
1: 505–509.
[5] Park S, Vosguerichian M and Bao ZA. A review of fabrication and applications of
carbon nanotube film-based flexible electronics. Nanoscale 2013; 5: 1727–1752.
[6] Niu Z, Dong H, Zhu B, et al. Highly stretchable, integrated supercapacitors based on
single-walled carbon nanotube films with continuous reticulate architecture. Adv Mater
2013; 25: 1058–1064.
[7] Wang K, Zhao P, Zhou XM, et al. Flexible supercapacitors based on cloth-supported
electrodes of conducting polymer nanowire array/SWCNT composites. J Mater Chem
2011; 21: 16373–16378.
[8] Davoglio RA, Biaggio SR, Bocchi N, et al. Flexible and high surface area composites of
carbon fiber, polypyrrole, and poly(DMcT) for supercapacitor electrodes. Electrochim
Acta 2013; 93: 93–100.
[9] Cong HP, Ren XC, Wang P, et al. Flexible graphene–polyaniline composite paper for
high-performance supercapacitor. Energy Environ Sci 2013; 6: 1185–1191.
[10] Pumera M. Graphene-based nanomaterials for energy storage. Energy Environ Sci
2011; 4: 668–674.
[11] Chen YC, Hsu YK, Lin YG, et al. Highly flexible supercapacitors with manganese
oxide nanosheet/carbon cloth electrode. Electrochim Acta 2011; 56: 7124–7130.
[12] Choi BG, Chang SJ, Kang HW, et al. High performance of a solid-state flexible asym-
metric supercapacitor based on graphene films. Nanoscale 2012; 4: 4983–4988.
[13] Xu J, Wang Q, Wang X, et al. Flexible asymmetric supercapacitors based upon Co9S8
nanorod//Co3O4@RuO2 nanosheet arrays on carbon cloth. ACS Nano 2013; 7:
5453–5462.
[14] Maheswari N and Muralidharan G. Supercapacitor behavior of cerium oxide nanopar-
ticles in neutral aqueous electrolytes. Energy Fuels 2015; 29: 8246–8253.
[15] Lu X, Yu M, Wang G, et al. H-TiO2@MnO2//H-TiO2@C core–shell nanowires for
high performance and flexible asymmetric supercapacitors. Adv Mater 2013; 25:
267–272.
[16] Yang P, Xiao X, Li Y, et al. Hydrogenated ZnO core–shell nanocables for flexible
supercapacitors and self-powered systems. ACS Nano 2013; 7: 2617–2626.
Luo et al. 15

[17] Chee WK, Lim HN, Harrison I, et al. Performance of flexible and binderless polypyr-
role/graphene oxide/zinc oxide supercapacitor electrode in a symmetrical two-electrode
configuration. Electrochim Acta 2015; 157: 88–94.
[18] Gu D, Ding C, Qin Y, et al. Behavior of electrical charge storage/release in polyaniline
electrodes of symmetric supercapacitor. Electrochim Acta 2017; 245: 146–155.
[19] Zhao C, Wang C, Yue Z, et al. Wallace, intrinsically stretchable supercapacitors com-
posed of polypyrrole electrodes and highly stretchable gel electrolyte. ACS Appl Mater
Interfaces 2013; 5: 9008–9014.
[20] Duay J, Gillette E, Liu R, et al. Highly flexible pseudocapacitor based on freestanding
heterogeneous MnO2/conductive polymer nanowire arrays. Phys Chem Chem Phys
2012; 14: 3329–3337.
[21] Byun S, Sim DM, Yu J, et al. High-power supercapacitive properties of graphene oxide
hybrid films with highly conductive molybdenum disulfide nanosheets.
ChemElectroChem 2015; 2: 1938–1946.
[22] Qin T, Peng S, Hao J, et al. Flexible and wearable all-solid-state supercapacitors with
ultrahigh energy density based on a carbon fiber fabric electrode. Adv Energy Mater
2017; 7: 1700409.
[23] Zhang LL, Zhao X, Stoller MD, et al. Highly conductive and porous activated reduced
graphene oxide films for high-power supercapacitors. Nano Lett 2012; 12: 1806–1812.
[24] Yan J, Wang Q, Wei T, et al. Recent advances in design and fabrication of electro-
chemical supercapacitors with high energy densities. Adv Energy Mater 2014; 4:
1300816–1300840.
[25] Wang K, Wu H, Meng Y, et al. Conducting polymer nanowire arrays for high per-
formance supercapacitors. Small 2014; 10: 14–31.
[26] Zhang D, Miao M, Niu H, et al. Core-spun carbon nanotube yarn supercapacitors for
wearable electronic textiles. ACS Nano 2014; 8: 4571–4579.
[27] Pendashteh A, Senokos E, Palma J, et al. Manganese dioxide decoration of macro-
scopic carbon nanotube fibers: From high-performance liquid-based to all-solid-state
supercapacitors. J Power Sources 2017; 372: 64–73.
[28] Dong Z, Jiang C, Cheng H, et al. Facile fabrication of light, flexible and multifunc-
tional graphene fibers. Adv Mater 2012; 24: 1856–1861.
[29] Kou L, Huang T, Zheng B, et al. Coaxial wet-spun yarn supercapacitors for high-
energy density and safe wearable electronics. Nat Commun 2014; 5: 3754.
[30] Gao Y, Jin H, Lin Q, et al. Highly flexible and transferable supercapacitors with
ordered three-dimensional MnO2/Au/MnO2 nanospike arrays. J Mater Chem A 2015;
3: 10199–10204.
[31] Sahoo G, Ghosh S, Polaki SR, et al. Scalable transfer of vertical graphene nanosheets
for flexible supercapacitor applications. Nanotechnology 2017; 28: 415702–415710.
[32] Tehrani Z, Thomas DJ, Korochkina T, et al. Large-area printed supercapacitor tech-
nology for low-cost domestic green energy storage. Energy 2017; 118: 1313–1321.
[33] Kim BC, Jeong HT, Raj J, et al. Electrochemical performance of flexible poly(ethylene
terephthalate) (PET) supercapacitor based on reduced graphene oxide (rGO)/single-
wall carbon nanotubes (SWNTs). Synthetic Met 2015; 207: 116–121.
[34] Souza VHR, Oliveira MM and Zarbin AJG. Thin and flexible all-solid supercapacitor
prepared from novel single wall carbon nanotubes/polyaniline thin films obtained in
liquid–liquid interfaces. J Power Sources 2014; 260: 34–42.
16 Journal of Industrial Textiles 0(00)

[35] Long X, Zeng Z, Guo E, et al. Facile fabrication of all-solid-state flexible interdigitated
MnO2 supercapacitor via in-situ catalytic solution route. J Power Sources 2016; 325:
264–272.
[36] Feng JX, Ye SH, Wang AL, et al. Flexible cellulose paper-based asymmetrical thin film
supercapacitors with high-performance for electrochemical energy storage. Adv Funct
Mater 2014; 24: 7093–7101.
[37] Yao B, Yuan LY, Xiao X, et al. Paper-based solid-state supercapacitors with pencil-
drawing graphite/polyaniline networks hybrid electrodes. Nano Energy 2013; 2:
1071–1078.
[38] Zhou Q, Jia C, Ye X, et al. A knittable fiber-shaped supercapacitor based on natural
cotton thread for wearable electronics. J Power Sources 2016; 327: 365–373.
[39] Gui Z, Zhu H, Gillette E, et al. Natural cellulose fiber as substrate for supercapacitor.
ACS Nano 2013; 7: 6037–6046.
[40] Wang Z, Carlsson DO, Tammela P, et al. Surface modified nanocellulose fibers yield
conducting polymer-based flexible supercapacitors with enhanced capacitances. ACS
Nano 2015; 9: 7563–7571.
[41] Tong L, Liu J, Boyer SM, et al. Vapor-phase polymerized poly(3,4-ethylenedioxythio-
phene) (PEDOT)/TiO2 composite fibers as electrode materials for supercapacitors.
Electrochim Acta 2017; 224: 133–141.
[42] Wang JG, Yang Y, Huang ZH, et al. Rational synthesis of MnO2/conducting
polypyrrole@carbon nanofiber triaxial nano-cables for highperformance supercapaci-
tors. J Mater Chem 2012; 22: 16943–16949.
[43] Fan X, Wang X, Li G, et al. High-performance flexible electrode based on electrode-
position of polypyrrole/MnO2 on carbon cloth for supercapacitors. J Power Sources
2016; 326: 357–364.
[44] Huang Y, Peng L, Liu Y, et al. Biobased nano porous active carbon fibers for high-
performance supercapacitors. ACS Appl Mater Interfaces 2016; 8: 15205–15215.
[45] Gamby J, Taberna PL, Simon P, et al. Studies and characterisations of various acti-
vated carbons used for carbon/carbon supercapacitors. J Power Sources 2001; 101:
109–116.
[46] Ma W, Chen S, Yang S, et al. Bottom-up fabrication of activated carbon fiber for all-
solid-state supercapacitor with excellent electrochemical performance. ACS Appl Mater
Interfaces 2016; 8: 14622–14627.
[47] Li ZY, Akhtar MS, Kwak DH, et al. Improvement in the surface properties of activated
carbon via steam pretreatment for high performance supercapacitors. Appl Surf Sci
2017; 404: 88–93.
[48] Frackowiak E and Beguin F. Carbon materials for the electrochemical storage of
energy in capacitors. Carbon 2001; 39: 937–950.
[49] Dong L, Xu C, Yang Q, et al. High-performance compressible supercapacitors based
on functionally synergic multiscale carbon composite textiles. J Mater Chem A 2015; 3:
4729–4737.
[50] Anitha VC, Banerjee AN, Dillip GR, et al. Nonstoichiometry-induced enhancement of
electrochemical capacitance in anodic TiO2 nanotubes with controlled pore diameter.
J Phys Chem C 2016; 120: 9569–9580.
[51] Song Y, Cheng X, Chen H, et al. Highly compression-tolerant folded carbon nanotube/
paper as solid-state supercapacitor electrode. Micro Nano Lett 2016; 11: 586–590.
Luo et al. 17

[52] Choi C, Lee JM, Kim SH, et al. Twistable and stretchable sandwich structured fiber for
wearable sensors and supercapacitors. Nano Lett 2016; 16: 7677–7684.
[53] Li HB, Yu MH, Wang FX, et al. Amorphous nickel hydroxide nanospheres with
ultrahigh capacitance and energy density as electrochemical pseudocapacitor materials.
Nat Commun 2013; 4: 1894–1900.
[54] Li WC, Mak CL, Kan CW, et al. Enhancing the capacitive performance of a textile-
based CNT supercapacitor. RSC Adv 2014; 4: 64890–64900.
[55] Burke A. Ultracapacitors: Why, how, and where is the technology. J Power Sources
2000; 91: 37–50.
[56] Wang DW, Li F, Zhao J, et al. Fabrication of graphene/polyaniline composite paper
via in situ anodic electropolymerization for high-performance flexible electrode. ACS
Nano 2009; 3: 1745–1752.
[57] Peng X, Liu H, Yin Q, et al. A zwitterionic gel electrolyte for efficient solid-state
supercapacitors. Nat Commun 2016; 7: 11782.
[58] Wee G, Salim T, Lam YM, et al. Printable photo-supercapacitor using single-walled
carbon nanotubes. Energy Environ Sci 2011; 4: 413–416.
[59] Guo W, Xue X, Wang S, et al. An integrated power pack of dye-sensitized solar cell and
Li battery based on double-sided TiO2 nanotube arrays. Nano Lett 2012; 12: 2520–2523.

Potrebbero piacerti anche