You are on page 1of 10

Journal of Power Sources 520 (2022) 230866

Contents lists available at ScienceDirect

Journal of Power Sources


journal homepage: www.elsevier.com/locate/jpowsour

Nitrogen-doped graphene fiber electrodes with optimal micro-/meso-/


macro-porosity ratios for high-performance flexible supercapacitors
Feng Han a, c, Weixuan Jing a, *, Qian Wu b, Bian Tian a, Qijing Lin a, Chenying Wang a,
Libo Zhao a, Junshan Liu c, Yu Sun d, Zhuangde Jiang a
a
State Key Laboratory for Mechanical Manufacturing Systems Engineering, Xi’an Jiaotong University, Xi’an, 710049, PR China
b
School of Highway, Chang’an University, Xi’an, 710064, PR China
c
Key Laboratory for Micro/Nano Technology and System of Liaoning Province, Dalian University of Technology, Dalian, 116024, PR China
d
Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, Ontario, Canada

H I G H L I G H T S

• Hierarchically structured nitrogen-doped graphene fibers are fabricated.


• The utilization efficiency of fiber inner interface is improved.
• Nitrogen doping improves wettability, nitrogen content and electrical conductivity.
• The all-solid-state flexible fiber-based supercapacitors have been fabricated.
• Fiber-based supercapacitors exhibit outstanding electrochemical performance.

A R T I C L E I N F O A B S T R A C T

Keywords: The introduction of graphene fiber-based supercapacitors as important power storage components in wearable
Fiber-based supercapacitors electronic products has attracted increasing attention because of their lightweight, good flexibility, high power
Nitrogen-doped graphene fiber density, long lifecycle and excellent charge/discharge capacity. However, their capacitance and energy density
Graphene stacking
are greatly confined as a result of the dense graphene stacking, hydrophobicity and poor electrical conductivity
Utilization efficiency
Electrochemical performance
of graphene fiber electrodes. Here we report a nitrogen-doped graphene fiber electrode with optimal micro-/
meso-/macro-porosity ratios to improve the utilization efficiency of fiber inner interface for accumulated charge
storage and boosted ion transport. Urea is employed as both nitrogen doping source and self-removed template
for graphene fibers to tune the hierarchically porous architecture, wettability, nitrogen content, and electrical
conductivity for improved electrochemical performance. The resulting supercapacitors display superior areal
capacitance of 1,217 mF/cm2 (486.3 F/g) and high energy density of 27 μW h/cm2 (10.8 W h/Kg) in polyvinyl
alcohol/H3PO4 gel electrolyte. These metrics represent the highest values to date among existing all-solid-state
fiber-based supercapacitors based on inorganic electrolyte. Moreover, the fiber-based supercapacitors show good
rate capability and excellent cyclic performance. Our work also provides an understanding of the effect of fiber
structure on electrochemical activity and highlights the importance of full utilization of graphene interior
interface.

1. Introduction sources are generally flat and rigid, it is hard to satisfy the demand for
wearability [4–6]. Fiber-based supercapacitors (FSCs), as promising
In recent years, flexible and wearable electronics has aroused great energy storage components, have undergone intensive development for
interest owing to their wide applications in sensors, collapsible displays, their outstanding advantages of light weight, good flexibility, high
and intelligent skins [1–3]. Flexible power supplies are urgently needed power density, long lifecycle and excellent charge/discharge capacity
to offer the power for these electronics. Since conventional power [7–17]. According to the charge storage mechanism, there are two

* Corresponding author.
E-mail address: wxjing@mail.xjtu.edu.cn (W. Jing).

https://doi.org/10.1016/j.jpowsour.2021.230866
Received 26 June 2021; Received in revised form 16 November 2021; Accepted 30 November 2021
Available online 7 December 2021
0378-7753/© 2021 Elsevier B.V. All rights reserved.
F. Han et al. Journal of Power Sources 520 (2022) 230866

categories of supercapacitors, namely pseudo-capacitors and electrical micro-/meso-/macro-porosity ratio, elevated wettability, additional ni­
double-layer capacitors (EDLCs). The materials used in trogen (N)-active sites, and excellent electrical conductivity. By altering
pseudo-capacitors mainly include transitional metal oxides/hydroxide the surface charge density and porous architecture of graphene layers
and conducting polymers, which are vital for the enhancement of through chemical interactions, the nitrogen doping effectively tunes the
pseudo-capacitance through faradaic redox reactions [18–22]. local electronic structure of electrodes and improves the fiber structure
Compared with pseudo-capacitors, EDLCs based on high specific surface utilization, resulting in enhanced ion diffusion and superior charge
carbon materials have distinct advantages in superior cycling stability, storage ability. The supercapacitor assembled from 20%-NGFs shows a
better reversibility, higher electrical conductivity, and lower cost [22, high areal capacitance of 1,217 mF/cm2 at 0.1 mA/cm2 (486.3 F/g at
23]. Carbonous FSCs are primarily fabricated with activated charcoal, 0.4 A/g), corresponding to an ultrahigh energy density of 27 μW h/cm2
reduced graphene oxide (rGO) and carbon nanotubes (CNTs), which are at a power density of 40 mW/cm2 (10.8 W h/Kg at 16.0 W/kg). To our
promising candidates to be extensively applied in generators [24], knowledge, these metrics represent the highest values to date among
sensors [25], actuators [26], and nonvolatile memory [27]. existing all-solid-state FSCs based on inorganic electrolyte to. Addi­
Owing to the remarkable mechanical/electrical properties and ul­ tionality, the 20%-NGFs based supercapacitor shows good rate perfor­
trahigh theoretical specific surface area (SSA) (2,620 m2/g), graphene is mance, excellent cycling stability and flexibility.
considered as a very promising candidate as an attractive electrode
material for FSCs [28–31]. However, the loosely connected graphene 2. Experimental section
nanosheets are significantly aggregated and closely packed within gra­
phene fibers during dry process because of the strong interaction of 2.1. Preparation of nitrogen-doped graphene fibers (NGFs)
capillary force between layers [32–34]. The severe graphene stacking
causes limited electrolyte penetration and poor utilization of the high GO was fabricated according to the Hummer’s method with modi­
SSA of graphene [35]. Therefore, a variety of materials have been uti­ fications [52]. Then NGFs were fabricated as described below. Urea was
lized to mitigate graphene stacking and promote the charge-storage mixed into an aqueous solution of GO (12 mg/mL). Fibers with different
ability [18–22,35–41]. For instance, 286 mF/cm2 in capacitance and urea/GO weight ratios of 0:1, 1:9, 1:4 and 2:3 (named as GFs,
6.4 μW h/cm2 in energy density were achieved through usage of gra­ 10%-NGFs, 20%-NGFs, and 40%-NGFs, respectively) were synthesized
phene fiber electrodes that incorporated holey graphene [35]. To further to determine the optimal composition of the hybrid fibers for FSCs. The
improve electrochemical performance, conducting polymers (polyani­ mixture solution was treated by ultrasonic processing for 30 min and
line and poly(3,4-ethylenedioxythiophene) and metal oxide (Co3O4 and subsequently injected into a polytetrafluoroethylene (PTFE) tube of 2
MnO2) were incorporated on graphene to provide pseudocapacitance mm in inner diameter using a syringe pump at the current velocity of 10
effect [39,40]. Hence, the capacitance and energy density values were mL/h, which was then heated at 180 ◦ C for 2 h in an oven after sealing
enhanced to 304.5 mF/cm2 and 6.8 μW h/cm2, respectively [18]. airtight both ends of the tube. The preformed amino-functionalized
Furthermore, the introduction of nitrogen heteroatom into graphene reduced-GO (rGO(-NH2)n) fibers with a 1 mm diameter in the wet
fiber for large SSA and high electrochemical activity contributed to the state were released from the tube by N2 flow and dried in the air. The
excellent capacitance of 1,132 mF/cm2 and high energy density of 25.1 length of the fibers remained almost unchanged, but the diameter
μW h/cm2 [41]. greatly decreased due to water loss. Finally, the fibers were placed into a
In the light of the charge storage mechanism of EDLCs, electrode furnace with controlled vacuum and gas flows. Before heating a vacuum
materials can achieve the storage and release of energy through the of 0.1 Pa and ultrapure argon flow of 200 sccm were used to take away
electrostatic interaction of ions at the electrode/electrolyte interface O2 and water vapor for 2 h. The samples were then annealed with a rate
[42,43]. Although several schemes have been proposed to boost the of 2.5 C/min under argon protection, maintained constant at 1000 ◦ C

utilization of fiber structures, efforts have been mainly made to apply for 2 h, and left in the furnace to cool naturally. Ambient air was finally
outer surfaces [44–46]. One strategy is to construct pores on graphene allowed to the furnace and the fibers were retrieved. After thermal
fibers to shorten the ion diffusion distance between graphene layers, annealing, the amino-functionalized rGO fibers were reduced to NGFs.
thus facilitating ion transport to the whole surface area [44]. Electrolyte
ions can only penetrate several nanometers away from the fiber-shaped 2.2. Assembly of NGFs-based supercapacitors (NGFSCs)
electrode surface, leaving the interior parts merely as carbon substrates
and contributing little to charge storage and electrochemical perfor­ The preparation of the gel electrolyte was as follows: firstly, 1 g
mance. Therefore, it is desired to construct graphene fiber electrodes polyvinyl alcohol (PVA) powder was added into 10 mL DI water, fol­
with an increased effective interfacial area to further promote the per­ lowed by 30 min of vigorous stir at 90 ◦ C until it became a clear solution.
formance of graphene-based FSCs (GFSCs). An efficient method is to Secondly, 1 g H3PO4 was added into the mixture when its temperature
synthesize a hollow RGO/conducting polymer composite fiber with dropped to the environment. All-solid-state NGFSCs was then assembled
additional inner interface to produce a higher SSA and present a higher with two fibers and the PVA/H3PO4 gel electrolyte. The fiber was
specific capacitance in the fiber supercapacitor [18]. However, the uti­ immersed in the electrolyte for 1 h and removed to dry in the air for 8 h.
lization of fiber electrode inner structure in FSCs remains low. Wu et al. Two fiber electrodes coated with electrolyte were parallelly put on
reported nitrogen-doped graphene fibers (NGFs) with uniform porous flexible PET substrate by recoating some electrolyte to make a solid-state
structure and demonstrated the significant influence of SSA on their supercapacitor.
electrochemical performance [41]. Nevertheless, the electrochemical
performance of GFSCs is also under the influence of ion-accessible pore 2.3. Materials characterization
size besides SSA [47,48]. Through taking pore curvature into account,
the heuristic theoretical model can be built and substituted for the EDLC The observation of surface morphologies was performed using field
model that a hierarchically porous structure is essential for the ideal emission scanning electron microscopy (FESEM, Hitachi SU-8010). The
carbon electrode to promote charge storage with micropores (<2 nm), tensile stress-strain properties were measured with a microforce tester
enhance ion transport with mesopores (2–50 nm) and act as buffer area (MTS Tytron250). X-ray diffraction (XRD, Brooke Advanced D8 A25)
for electrolyte ions with macropores (>50 nm) [49–51]. We hypothe­ was employed to analyze the crystal structures of fibers. The chemical
sized that improved utilization efficiency of fiber inner interface can be structures were detected with a Bruker VERTEX70 Fourier transform
achieved for enhanced charge storage and shorten ion pathway by infrared (FTIR) spectroscopy. Optical contact angle (OCA) measurement
varying the micro-/meso-/macro-porosity ratios in the fiber electrode. was conducted on a Dataphysics OCA40Micro measuring instrument to
In this work, we report hierarchically porous NGFs with an optimal survey the wettability and contact angles. The surface chemical analyses

2
F. Han et al. Journal of Power Sources 520 (2022) 230866

were characterized using a Thermo Fisher Escalab xi + X-ray photo­


td
electron spectroscopy (XPS). Keithley 2400 Source Meter was utilized to η= × 100% (8)
tc
measure the electrical conductivities. Nitrogen adsorption–desorption
measurement was carried out with a static volumetric gas adsorption where tc and td represent the charge and discharge time, respectively.
instrument (Micromeritics ASAP 2020 Plus HD88) and calculated using
Brunauer-Emmett-Teller (BET) technique to characterize surface area 3. Results and discussions
and pore structure.
The formation mechanism of the NGFs is shown in Fig. 1 and Fig. S1.
2.4. Electrochemical measurements Well-mixed GO/urea was injected into a PTFE tube, followed by baking
it at 180 ◦ C for 2 h after sealing up the two ends of the tube. The mixture
An electrochemical workstation (CHI 660D) was employed to mea­ was uniformly diffused and evenly self-assembled into a 3D network
sure the electrochemical performance, including cyclic voltammetry structure within the tube, followed by drying in air to form amino-
(CV), galvanostatic charge/discharge (GCD) and electrochemical functionalized rGO fibers (also known as NGFs before annealing). The
impedance spectra (EIS). The areal capacitance (CA) and mass capaci­ fiber was heated at 1000 ◦ C in an inert environment such that rGO
tance (CM) of the single fiber electrode can be quantified by: experienced a further reduction, and the amino functionalization was
CA = 4I × t × U − 1 A− 1
(1) decomposed and doped nitrogen into the graphene lattice. Accordingly,
the nitrogen-doped graphene fiber with a hierarchically porous network
CM = 4I × t × U − 1 M − 1
(2) and excellent electrical conductivity was achieved. For comparison, GFs
were also produced with the same procedure but without urea. Notably,
where I represents charge/discharge current, t stands for the discharge nitrogen contents in the graphene fibers increased with the increasing
time, U is equal to the potential range, A represents the surface area of urea loadings in the mixture solution, as quantified by XPS analysis.
the overlapping part between two electrodes, and M is the mass of the The porous fibers were first studied by SEM. Fig. 2a is a presentation
two fiber electrodes, respectively. of the cross-sectional images of GF with a diameter of ~126 μm,
Then, the energy density (E) and power density (P) of the electrode demonstrating the rough and wrinkled surface. The zoomed-in view
materials based on the area/mass of the whole FSCs were determined by: (Fig. 2b) shows that the graphene sheets in GF were densely stacked and
/ highly aligned along the fiber’s main axis. In addition, the cross-
E = CU 2 8 (3) sectional images of GF (Fig. 2c and d) reveal compactly entangled gra­
phene sheets with micropores, which is proved by further BET analysis.
P = E × t− 1
(4) SEM images of 10%-NGF, 20%-NGF, and 40%-NGF with different urea
For the NGFSCs, the surface area (A), volume (V) and mass (M) of the loadings are shown in Fig. 2e–h, Fig. 2i-l, and Fig. 2m–p, respectively.
two electrodes are calculated by: For the NGF with a low urea fraction of 10% (Fig. 2f), more inter­
connected mesoporous surface can be seen compared with that of GF.
A = 2πDL (5) The number of mesopores in the fiber (Fig. 2j) increases as the weight
ratio of urea increased to 20%. From the zoomed-in view of the cross
V = 0.5 πD2 L (6) section of 20%-NGF (inset of Fig. 2l), the fiber is well distributed with
hierarchical pores with a size range from a few tenths of a nanometer to
M = ρV (7) over a hundred nanometers. It is beneficial for such a porous structure of
the fiber to provide plenty of ion-accessible path channels for rapid
where D represents the diameter of the fibers, L is the overlapping length
charge transfer. As a result, the 3D pillared-vertically aligned nitrogen-
between two fibers, and ρ is the density of the fiber, respectively.
doped graphene sheets are constrained within a 1D fibrous structure,
The coulombic efficiency is derived as:
resulting in that our fibers with both the rich pores and effective ion-

Fig. 1. Schematic illustration showing the fabrication process of porous NGFs. (a) Doping mechanism of nitrogen in graphene. (b) The nitrogen-doped graphene fiber
was prepared by injecting well-dispersed solutions comprising GO and urea into a PTFE tube, followed by a confined self-assembly step in an oven at 180 ◦ C for 2 h,
release from the tube by a pressurized nitrogen flow, and thermal reduction at 1000 ◦ C for 2 h.

3
F. Han et al. Journal of Power Sources 520 (2022) 230866

Fig. 2. Microstructures of the hybrid fibers. (a–d) The surface and cross-sectional images of GF, respectively. The inset of (d) shows the densely stacked graphene
sheets in GFs at high magnification. (e− h) The surface and cross-sectional images of 10%-NGF, respectively. The inset of (h) shows more interconnected porous
surface in 10%-NGF than that of GF. (i–l) The surface and cross-sectional images of 20%-NGF, respectively. The inset of (l) shows hierarchical pores with sizes in the
range of several nanometers to over a hundred nanometers. (m–p) The surface and cross-sectional images of 40%-NGF, respectively. The inset of (p) indicates a
portion of small pores in the fiber are further enlarged into large pores.

accessible interfacial area are attractive for high-areal-performance fibers, the FTIR patterns of GFs before and after annealing, 10%-NGFs,
FSCs. When the weight ratio of urea is 40%, a portion of small pores 20%-NGFs, and 40%-NGFs were analyzed (Fig. S4). The GF before
in the fiber is further enlarged into macropores (Fig. 2p), and graphene annealing shows the distinct peaks at 2665, 2308, 1787, 1555 and 1061
sheets become sparsely stacked (Fig. 2n). cm− 1, which originate from CHn bonds, CO2, C– – O, C– – C and C–O,
The stress-strain data of GFs and NGFs before and after annealing are respectively, revealing that the surface of rGO(-NH2)n are attached with
shown in Supplementary Fig. 2. Before annealing the tensile strength abundant oxygen-containing groups. The GF after annealing has their
and elongation of 20%-NGF are 152 MPa and 6%, respectively, exhib­ characteristic peaks around 2308 (CO2 bond), 1728 (C– – O), 1546
iting an increase in tensile strength (126 MPa) and a slight decline in (C–– C) and 1068 cm− 1 (C–O), while there is no obvious peak relating to
elongation (6.7%) compared with that of GF. The tensile strengths are CHn bonds but a new peak relating to sp2 C–C bonds is present, sug­
obviously improved (224 MPa for 20%-NGF and 210 MPa for GF) and gesting a large quantity of oxygen groups are partially decomposed and
elongations are found to decrease (2% for 20%-NGFs and 2.4% for GFs) sp2 carbon regions are restored [55]. For all the NGFs, the peaks around
after high temperature treatment. In comparison, the mechanical 2004 (C–N), 1625 (C– – C), 1284 (C–– N) and 730 cm− 1 (sp2 C–C) confirm
integrity of 40%-NGFs after annealing is much poorer because of a that the NGFs are both further reduced and doped with nitrogen atoms.
largely reduced elongation (1.1%) and prone to breakage during The elemental composition of carbon, oxygen and nitrogen in the
handling, and parts of the fiber structure lack continuity. fibers was examined by XPS (Fig. S5). As shown in Fig. S5a, the survey
The fibers were nitrogen doped during the nitridation synthesis with spectra of the GF before and after annealing reveal the existence of C1s
urea, as further investigated by XRD, FTIR, and XPS (Figs. S3–5). Fig. S3 (284 eV) and O1s (532 eV), while the characteristic peak of N1s is
presents the XRD patterns of the hybrid fibers. GF before annealing has observed in 20%-NGF, suggesting that 20%-NGF has nitrogen incorpo­
the (002) diffraction peak located at 2θ = 11.49◦ , corresponding to rated into the carbon matrices. The high resolution C1s peak of GF
calculated interlayer spacing of 7.7 Å. However, the diffraction peak for before annealing (Fig. S5b) can be deconvoluted into one main peak at
GF shifts to 2θ = 25.86◦ with a d-spacing of 3.44 Å after thermal 284.8 eV (sp2 C–C), and three small peaks including C–O (285.4 eV),
annealing at 1000 ◦ C. Such a decreased interlayer spacing is probably C–– O (286.6 eV) and O–C– – O (288.5 eV) at higher binding energies,
because the oxygen heteroatoms within rGO(-NH2)n are largely removed demonstrating the presence of oxygen heteroatoms. After annealing, the
through high temperature treatment [53,54]. As for NGFs with different GF (Fig. S5c) has more dominance of sp2 C–C in C1s peak, further
urea contents, the XRD patterns exhibit the peak centered at 2θ = illustrating partially decomposition of oxygen functional groups and
27.64◦ , which corresponds to the interlayer spacing of 3.22 Å. The lower better recovery of π-conjugated structure. For 20%-NGF (Fig. S5d), peak
d-spacing of NGFs than that of GFs indicates the change of inner struc­ deconvolution of the C1s spectrum presents three sub-peaks of sp2 C–C,
ture within rGO, resulting from the further reduction of rGO and the C–– N & C–O and C–N & C– – O, which locate at 284.8, 285.6 and 287 eV,
replacement of certain carbon atoms by nitrogen atoms in graphene. respectively [56].
To further confirm thermal reduction and nitrogen-doping of the The XPS characterization was used to further confirm the nitrogen

4
F. Han et al. Journal of Power Sources 520 (2022) 230866

content and bonding configurations to investigate the relationship be­ current density of 0.1 mA/cm2. As compared to that of GF (20.5 mV), the
tween evolution of nitrogen functionalities during synthesis and smaller IR drop can be observed in 10%-NGFs (16.1 mV), 20%-NGFs
capacitive performance of fibers (Figs. 4e and S6). As seen the high (8.2 mV), and 40%-NGFs (14.8 mV). The results may be ascribed to the
resolution N1s peak in Fig. S6d, the obtained NGFs have rather complex smaller overall resistance and shorter diffusion distance in NGFs than
structure with four types of nitrogen, which is different from that of urea that in GFs, which are conducive to better capacitance characteristics
with only one nitrogen functionality (N–H). Four components of N1s [59–61]. The discharge time for GF, 10%-NGF, 20%-NGF, and 40%-NGF
spectrum consist of pyridinic N (N-6), pyrrolic N (N-5), quaternary N (N- is 759s, 1419s, 2434s, and 1878s, respectively. Among these fibers,
G) and oxidized N (N–O), locating at 398.6, 399.9, 401 and 402 eV, 20%-NGF possesses the longest discharge time, and thus contributes to
respectively. Among these functionalities, N-6 and N-5 mainly exist on the highest areal capacitance of all the FSCs. Moreover, the GCD curves
the edge of the graphene nanosheets, and may provide additional elec­ of the GFs, 10%-NGFs, 20%-NGFs, and 40%-NGFs present regular isos­
trochemically active sites to enhance pseudo-capacitance. N-G is doped celes triangles in shape, further verifying the good electrochemical
within the graphitic basal plane, which can help improve electrical performance and reversibility [44].
conductivity of carbon materials and enhance fast charge/discharge [57, For fiber supercapacitors, the specific capacitance based on area is
58]. Hence, nitrogen-doping shows great impact on the improvement of more common and useful factor than that based on mass to evaluate the
specific capacitance owing to the pseudo-capacitive and electrical real application for flexible energy-storage devices. When the current
conductive contributions. As presented in Figs. S6a–c, the N-G peak density is in the range of 0.1–1 mA/cm2, the values of areal capacitance
becomes increasingly apparent with higher urea fractions, resulting were calculated based on the GCD curves, the results of which were
from the transformations of the nitrogen configurations and recon­ summarized in Fig. 3c. The GF presents a good capacitance of 379.5 mF/
struction of the NGFs structure. A greater proportion of N-G during cm2 at 0.1 mA/cm2, which can be compared with the values for most
synthesis under enough urea fractions can be ascribed to the higher FSCs reported previously [18,60,62]. The areal capacitance of 10%-NGF
stability than that of N-6 and N-5 [56]. increases to 709.5 mF/cm2, and it reaches the best value of 1217
To demonstrate the superiority of fiber electrodes for FSCs, the mF/cm2 for 20%-NGF, demonstrating 1.87 times and 3.2 times that of
electrochemical test was performed through CV and GCD in a symmetric the GF electrode, respectively. Nevertheless, the capacitance value de­
two-electrode configuration (Fig. 3), where our fiber and 1 M PVA/ creases to 938.75 mF/cm2 for 40%-NGF. Similar trends were also
H3PO4 gel was used as the electrode and electrolyte, respectively. From observed for all the FSCs under different current densities (Fig. 3c). With
the plots in Fig. 3a, the CV curves of the as-synthesized FSCs are ob­ the increasing current density to 1 mA/cm2, satisfactory CA values of
tained at 2 mV/s with the potential in the range 0–0.8 V. Surprisingly, all NGFs electrodes (356 mF/cm2, 659 mF/cm2, 298 mF/cm2 for 10%-NGF,
the CV curves exhibit a near rectangle in shape behaving as the typical 20%-NGF, and 40%-NGF, respectively) are still preserved, whereas the
EDLCs. The GF presents the smallest closed area by comparing CV GF only maintains 168 mF/cm2. In summary, the 20%-NGF electrode
curves, which was smaller than that of 10%-NGFs. The 20%-NGF ex­ exhibits a high CA value and excellent rate capability.
hibits the surprisingly largest CV area, suggesting the best charge- The Ragone plots in Fig. 3d compare the energy storage performance
storage ability as a fiber electrode. With the increasing urea content to of our FSCs with different nitrogen doping in PVA/H3PO4 gel electrolyte.
40%, the CV area dramatically decreases although it is yet larger than The energy density increases to 27 μW h/cm2 for 20%-NGFSCs, which is
those of GFs and 10%-NGFs, leading to deteriorated electrochemical higher than GFSCs (8.43 μW h/cm2), 10%-NGFSCs (15.77 μW h/cm2)
performance. and 40%-NGFSCs (20.86 μW h/cm2). In addition, a downward trend of
Fig. 3b reveals the GCD curves for the graphene-based FSCs at a the energy density is shown in all the FSCs with increase in the power

Fig. 3. Electrochemical characterization of FSCs in H3PO4/PVA electrolyte. (a) CV curves of FSCs at a rate of 2 mV/s (b) GCD curves at a current density of 0.1 mA/
cm2. (c) The areal capacitance calculated from GCD curves under different current densities. (d) The Ragone plots of the NGFSCs. The rate capabilities of the GFSCs,
10%-NGFSCs, 20%-NGFSCs and 40%-NGFSCs are marked on the diagram, respectively.

5
F. Han et al. Journal of Power Sources 520 (2022) 230866

density. It is noteworthy that the E/PA curve of 20%-NGFSCs is urea in the fiber (Fig. S7) can be further removed through high tem­
comparatively smooth with the best rate capability of 54.1%, whereas perature annealing with the release of large amounts of gaseous prod­
the GFSCs (44.3%), 10%-NGFSCs (46.4%) and 40%-NGFSCs (49.7%) ucts such as CO2, water molecules and N-containing gases. These gases
show the poorer rate performance. decomposed from urea quickly enter the narrow space and cause a
To establish a comprehensive understanding of the influence of ni­ “bombing” effect, thereby overcoming the van der Waals interaction
trogen doping on the capacitive performance enhancement of the between adjacent NGFs nanosheets and forming porous structures. This
NGFSCs, various key characterizations consisting of N2 adsorption/ loose porous structure, including mesoporous and macroporous struc­
desorption, contact angle, electrical conductivity measurements, XPS tures, can increase the gap between adjacent NGFs nanosheets, thereby
spectra and EIS were conducted. N2 adsorption/desorption was first preventing their stacking.
used to achieve and assess the BET SSA (Fig. 4a) and pore size distri­ As shown in Fig. 4a, c, the GF has 78.4% micropores, 19.1% meso­
bution (PSD, Fig. 4b). From N2 adsorption/desorption isotherm test, pores, 2.5% macropores and low SSA (228 m2/g). The low SSA can
together with SEM images (Fig. 2), it is worth noting that the SSA and reduce the available ion-accessible area in GF. Meanwhile, micropores
pore structures of NGFs are distinct from that of GFs and they vary with can strengthen the ion storage for a prominent capacitive performance,
the urea content. The GF isotherm (Fig. 4a) shows typical type-I curves whereas the ultrahigh micropore volume of GF can increase the resis­
with strong N2 adsorption observed at relative atmospheric pressure (P/ tance of electrolyte ion transporting from fiber-shaped electrode surface
P0) <0.1, demonstrating the micropore-rich structures, while the NGFs to the interior parts, leading to the low rate performance at high current
present typical type-IV isotherms with abundant mesopores. More density. Notably, the GF shows the hydrophobic property with an initial
detailed information can be provided by further analysis of the PSD contact angle of 103.8◦ (Fig. 4d), suggesting the raised resistance and
(Fig. 4b) and pore volume percentages (Fig. 4c). With the increasing hindered ion transport at electrode/electrolyte interface because of the
urea fraction from 0 to 40%, the microporosity of the as-prepared fibers inferior wettability [65,66]. A large IR drop (Fig. 3b) of 20.5 mV and
reduces, while at the same time the mesoporosity and macroporosity equivalent series resistance (ESR, Fig. 4f) of 257.2 Ω are obtained due to
increase, suggesting a transition from the micropore-rich structure of the serious blocking of ion diffusion in GF caused by the low electrical
GFs to the mesopore-rich structure of NGFs. Such pore size change is conductivity (175 S/cm, Fig. 4e). Considering all these factors, the GFSC
probably because urea can serve as a self-removed pore former as well as presents a low areal capacitance of 379.5 mF/cm2 and rate capability
nitrogen source to realize synergetic pore structure optimization and (44.3%).
nitrogen doping [63,64]. In the fabrication process of NGFs, residual In contrast to GFs, nitrogen doping leads to an enhanced areal

Fig. 4. Characteristics of the hybrid fiber electrode. (a) Nitrogen adsorption/desorption isotherm. (b) pore size distribution and (c) different pore volume percentages
of GF, 10%-NGF, 20%-NGF, and 40%-NGF. (d) Water contact angle measurement for GFs with different urea contents. (e) The relationship between electrical
conductivities and different nitrogen-active sites. (f) Nyquist plots of the FSCs from GF, 10%-NGF, 20%-NGF, and 40%-NGF.

6
F. Han et al. Journal of Power Sources 520 (2022) 230866

capacitance and rate capability. For instance, 10%-NGF shows a higher generated within the fiber inner interface and the ion path channels are
SSA (305 m2/g) with different 41.6% micropores, 50.6% mesopores and obviously shortened, which significantly favors enhanced electrode/
7.8% macropores. It is worth noting that the areal capacitance of 10%- electrolyte contact, faster charge transport and higher rate capability
NGF is as high as 709.5 mF/cm2. The extraordinary capacitive perfor­ (54.1%) of the 20%-NGF. The improved porosity and increased hydro­
mance can be intensely related to the well-defined mesoporosity and philic groups from N-active sites further enhance surface wettability of
macroporosity, improved wettability, enhanced electrical conductivity the electrode with a decreased contact angle of 81.5◦ (Fig. 4d). The
and introduced N-active sites. Thus, investigating the mechanism of the superior wettability provides abundant accessible sites on fiber surface
impact factors mentioned above is of great importance for improving the for the physical adsorption of electrolyte ions on the one hand, and on
electrochemical performance of 10%-NGF. the other hand it improves ionic conductivity of 20%-NGF. Considering
For 10%-NGF, the microporosity is reduced to 41.6%, but still shows the nitrogen configuration modification (the sum of N-6 and N-5 con­
a large amount of sub-nanometer size distribution (<1 nm). Before tents (3.0%)), more N-active sites are provided for the 20%-NGF to
electrolyte ions access to these sub-nanometer pores, the solvated shell enhance electrochemical activities. In addition, the 20%-NGF exhibits
of ions are removed and the diameter of the bare PO43- ions is ~0.46. the highest electrical conductivity (328 S/cm) among all the NGFs
With the size matching of the ultrasmall pores and bare ions, a large electrodes. Similar to the 10%-NGF, the larger (C–C)% content (86.4%),
number of electrolyte ions are absorbed and stored in the electrodes, the increased overall nitrogen contents (4.6%) and nitrogen configura­
leading to an enhanced specific capacitance [48,67–69]. Additionally, tions (N-G contents (1.3%)) further contribute to the higher reduction
the increased mesopores (50.6%) accelerate charge transport with level and thus better electrical conductivity, decreased IR drop (8.2 mV)
shortened ion pathways, whereas macropores (7.8%) are utilized to be and small ESR (109.2 Ω) in the 20%-NGF. Meanwhile, the 20%-NGF
buffering zone for electrolyte ions to achieve the shortest transport paths exhibits a largest areal capacitance of 1217 mF/cm2, which can be
to inner surface. According to the report by Wang et al., a lot of factors demonstrated from the Nyquist plot with a maximum oblique slope at
influence the mechanism of ion transportation in porous graphene-based low frequency (Fig. 4f).
materials. Among them, the charge transfer resistance and transport Although 40%-NGFs exhibits the highest SSA (509 m2/g), the best
paths are the most critical factors [60]. Therefore, the restacking of wettability (contact angle of 77.2◦ ), the maximum nitrogen content
graphene nanosheets within 10%-NGF is mitigated and the rate per­ (6.7%), there are few detectable micropores (7.9%) and excessive
formance (46.4%) is higher than that of GF due to the synergetic effect of macropores (28.4%) in hierarchical structures. The small micropore
the decreased charge transfer resistance in mesopores and the curtailed population of sub-nanometer pores, the size of which is equivalent to the
transport paths in macropores. size of bare ion PO43-, contributes to the reduced specific capacitance.
Moreover, when the urea fraction increases from 0 to 10%, the Moreover, the high macroporosity, especially for pore size >200 nm, is
wettability is enhanced, as evidenced from the decreased contact angle the primary reason for the decrease of electrical conductivity owing to
of the fibers (103.8◦ to 83.5◦ ). The data shown in Fig. 4d suggest a the lack of continuity in some interior parts of the fibers. As a result, the
transition from hydrophobic GF to hydrophilic 10%-NGFs. Although a areal capacitance and rate capability decreased to 938.75 mF/cm2 and
small contact angle can lower the interfacial ohmic resistance for facil­ 49.7%, respectively.
itating ion diffusion at the interface between electrode and electrolyte, Considering the surface double-layer capacitance mechanism of
the poor wettability of GFs had not been carefully considered to date. graphene-based fiber electrodes, electrolyte ions can only be stored and
The well-developed porosity and plenty of hydrophilic groups (intro­ separated in the range of several nanometers near the electrode/elec­
duced polar C–N bonds) were vital for the 10%-NGFs to achieve superior trolyte interface, leaving the interior parts merely as substrates and
wettability, owing to the additional amino-containing functional groups contributing little to the capacitance performance of FSCs [44–46]. In
(Fig. S8). order to comprehensively access the charge storage ability of our FSCs
Importantly, the addition of 10 wt% urea into GO solutions can and highlight the importance of full utilization of fiber structure, the
enhance the electrical conductivity of 10%-NGF to 253 S/cm. The electrochemical performance were compared between our 20%-NGFSCs
reason for this enhancement is the highly conductive porous networks and the previously reported graphene-based FSCs (Fig. 5). The areal
modified by doped nitrogen atoms and the higher reduction degree in capacitance value of the 20%-NGFSC reaches 1217 mF/cm2, repre­
10%-NGF. As observed in Figs. S5d and 4e, the improved (C–C)% con­ senting the highest areal capacitance value among all the reported
tent (83.6%) is considered to increase the reduction degree of 10%-NGF, porous graphene materials (Supplementary Table 1). For example, it is
and the overall nitrogen contents (2.1%) and N-G contents (0.2%) sug­ three times that of the SG-CPF@GF fiber (391.2 mF/cm2) [46], six times
gest a promoted ion transport. These factors together lead to an that of the rGO coaxial fiber (205 mF/cm2) [70], 33 times that of the
improved electrical conductivity for 10%-NGF, thereby diminishing IR plasma-treated GFs (36.25 mF/cm2) [44]. Furthermore, it is of great
drop (16.1 mV, Fig. 3b) and ESR (228.7 Ω, Fig. 4f) in the FSCs. The 10%- significance to maintain the capacitance of the FSCs at high
NGF electrode has a comparably larger slope of d(-Z′′ )/dZ’ at low fre­ charge-discharge current density for practical power sources. It can be
quency than that of GF electrode (Fig. 4f), revealing a better charge observed from Supplementary Table 1 that the 20%-NGF still preserves
storage behavior owing to the nitrogen doping. the capacitance of up to 659 mF/cm2 even under the high current
Finally, the electrochemical performance of the 10%-NGFs in FSC is density of 1 mA/cm2, which is 2.3 times that of N-40% HG@GF fiber
affected both by nitrogen contents and nitrogen configurations. Among (274.5 mF/cm2) [35] and five times that of RGO + CNT@CMC fiber
these functionalities, N-6 and N-5 can alter the electrochemical activities (132.7 mF/cm2) [14].
of fiber electrode due to their existence on the edge of the graphene The 20%-NGFSC demonstrates an exceptional energy storage per­
nanosheets [57,58]. We observed that the sum of N-6 and N-5 contents formance with the highest energy density of 27 μW h/cm2, surpassing all
in the fiber is 1.7% and it can provide additional N-active sites, resulting existing graphene-based hybrid FSCs (Supplementary Table 1). Notably,
in the improvement of specific capacitance (Fig. 4e). The unique porous the electrochemical performance of our supercapacitors are superior to
electrode architecture with higher mesoporosity, improved wettability, that of previously reported NGFSCs [41] owing to the high utilization
enhanced electrical conductivity and introduced N-active sites has efficiency of hierarchically porous structures via adjusting the micro-/­
synergistic effect in strengthening the areal capacitance and improving meso-/macro-porosity ratios to enhance electrolyte permeation and
the rate capability of 10%-NGF. speed up the ion diffusion kinetics.
The 20%-NGF shows an increased SSA (416 m2/g) with 29.2% mi­ Based on the above analysis of the architecture and capacitive
cropores, 58.2% mesopores, and 12.6% macropores. According to the behavior of NGFs electrode materials, we further investigated the po­
higher SSA and improved hierarchically porous structure with the tential mechanism of high-performance FSCs. In order to achieve ideal
increased amount of mesopores and macropores, more shortcuts are electrochemical performance, the hierarchically porous architecture

7
F. Han et al. Journal of Power Sources 520 (2022) 230866

Fig. 5. Comparison of the electrochemical performance of our 20%-NGFSCs with other previously reported FSCs. The supercapacitors assembled from plasma-
treated GFs (36.25 mF/cm2, 0.2 μW h/cm2) [44], rGO coaxial fiber (205 mF/cm2, 17.5 μW h/cm2) [70], SG-CPF@GF fiber (391.2 mF/cm2, 8.7 μW h/cm2) [46]
and some other fibers are listed in the graph.

with macropores, mesopores, and micropores should be constructed in cycles (Fig. 7b). To prove the practicability of the 20%-NGFSCs for
fiber electrodes. Although micropores are typically dominant in GFs, the powering wearable electronics, four NGFs-based supercapacitors are
dense graphene stacking often leads to small ion-accessible SSA, and the arranged in series, which powers up a lab timer (Fig. 7c). Fig. 7d shows
lack of adequate mesopores and macropores hinders the efficient the light up of 23 LEDs into a logo “XJTU” with four NGFs-based
transport of electrolyte ions because of the wormlike winding routes supercapacitors.
between different layers. In comparison, in our work, when an increased
SSA and hierarchically porous structure with optimal micro-/meso-/ 4. Conclusion
macro-porosity ratios are introduced in the GFs by nitrogen doping, the
ion diffusion and storage are significantly facilitated (Fig. 6). This is In summary, a high-performance FSC assembled by flexible and
because more shortcuts are generated within the hierarchical architec­ freestanding NGFs electrodes has been fabricated via a combination of
ture and ion path channels are greatly shortened by the porous structure. the structure and capacitive performance design. Urea serves as self-
In addition to producing hierarchically porous structures, nitrogen removed pore former as well as nitrogen source to realize synergetic
doping also modifies the local electronic structure of fiber electrodes, pore structure optimization and nitrogen doping, tuning the hierar­
increases surface charge density, and enhances ion binding interaction. chically porous architecture, wettability, nitrogen content, and elec­
Benefiting from the hierarchically porous structure with optimal micro-/ trical conductivity of NGFs for improved electrochemical performance.
meso-/macro-porosity ratio, elevated wettability, improved electrical It has been demonstrated that the distinctive layer porous architecture of
conductivity and additional N-active sites of NGFs, the high perfor­ the fiber electrodes facilitates the ion storage and transport. The optimal
mance of the FSCs can be achieved by means of nitrogen doping. supercapacitor displays a superior specific capacitance of 1,217 mF/cm2
The FSCs in this work have high flexibility and excellent electro­ (486.3 F/g) and a high energy density of 27 μW h/cm2 (10.8 W h/Kg),
chemical performance under various deformations. The CV curves which demonstrates the highest value in all-solid-state FSCs based on
(Fig. 7a) remain almost unchanged as the bending angles increase from inorganic electrolyte to date. Moreover, the FSC assembled from 20%-
0◦ to 180◦ , which plays a significant role in wearable electronics. As NGFs maintains ultrahigh capacitance (659 mF/cm2 at 1 mA/cm2 cur­
observed in Fig. 7b, the NGFSC also exhibits the capacitance retention of rent density), and exhibits excellent cyclic stability (96.5% retention
96.5% over 20000 GCD cycles, illustrating its superior cyclic stability. over 20000 cycles). The NGFSC is a very promising candidate for high-
Moreover, the coulombic efficiency is maintained at 99.2% after 20000 performance and flexible power sources in future wearable electronics.

Fig. 6. The mechanism for the high-performance of the FSCs. Densely stacked GFs with a large quantity of micropores have wormlike S-shaped route between layers
to block the ion transport. The NGFs with optimal micro-/meso-/macro-porosity ratios create more ion path channels to facilitate ion diffusion and storage.

8
F. Han et al. Journal of Power Sources 520 (2022) 230866

Fig. 7. The applications of our FSCs for electronics. (a) CV curves of the NGFSCs under bending. (b) Capacitance retention and coulombic efficiency of the NGFSCs.
The inset shows the charge/discharge curves of the last five cycles for 20%-NGFs. Photographs of four NGFSCs connected to power (c) stopwatch, (d) 23 LEDs,
respectively.

CRediT authorship contribution statement [3] D. Pech, et al., Ultrahigh-power micrometre-sized supercapacitors based on onion-
like carbon, Nanotechnology 5 (2010) 651–654.
[4] Y. Shao, et al., Graphene-based materials for flexible supercapacitors, Chem. Soc.
Feng Han: Conceptualization, Investigation, Writing – original draft, Rev. 44 (2015) 3639–3665.
Writing – review & editing. Weixuan Jing: Supervision, Conceptuali­ [5] L. Li, Z. Wu, S. Yuan, X.-B. Zhang, Advances and challenges for flexible energy
zation, Writing – review & editing. Qian Wu: Data preparation, Inves­ storage and conversion devices and systems, Energy Environ. Sci. 7 (2014)
2101–2122.
tigation. Bian Tian: Data preparation, Investigation. Qijing Lin: [6] O. Gutfleisch, et al., Magnetic materials and devices for the 21st century: stronger,
Resources, Writing – review & editing. Chenying Wang: Resources, lighter, and more energy efficient, Adv. Mater. 23 (2011) 821–842.
Writing – review & editing. Libo Zhao: Resources, Writing – review & [7] Y. Shao, et al., 3D Freeze-casting of cellular graphene films for ultrahigh-power-
density supercapacitors, Adv. Mater. 28 (2016) 6719–6726.
editing. Junshan Liu: Resources. Yu Sun: Conceptualization, Writing – [8] Y. Shao, et al., Flexible quasi-solid-state planar micro-supercapacitor based on
review & editing. Zhuangde Jiang: Conceptualization, Writing – review cellular graphene films, Horizons 4 (2017) 1145–1150. https://pubs.rsc.org/lv/c
& editing. ontent/articlehtml/2017/mh/c7mh00441a-fn1.
[9] H. Cheng, C. Hu, Y. Zhao, L. Qu, Graphene fiber: a new material platform for
unique applications, NPG Asia Mater. 6 (2014) e113.
Declaration of competing interest [10] J. Li, et al., Calligraphy-inspired brush written foldable supercapacitors, Nano
Energy 38 (2017) 428–437.
[11] J. Li, et al., Cladding nanostructured AgNWs-MoS2 electrode material for high-rate
The authors declare that they have no known competing financial and long-life transparent in-plane micro-supercapacitor, Energy Storage Mater. 16
interests or personal relationships that could have appeared to influence (2019) 212–219.
the work reported in this paper. [12] D. Yu, et al., Scalable synthesis of hierarchically structured carbon
nanotube–graphene fibres for capacitive energy storage, Nat. Nanotechnol. 9
(2014) 555–562.
Acknowledgments [13] Y. Meng, et al., All-graphene core-sheath microfibers for all-solid-state, stretchable
fibriform supercapacitors and wearable electronic textiles, Adv. Mater. 25 (2013)
2326–2331.
The authors would like to thank the financial supports by National [14] L. Kou, et al., Coaxial wet-spun yarn supercapacitors for high-energy density and
Natural Science Foundation of China (51805426 & 51975466), Young safe wearable electronics, Nat. Commun. 5 (2014) 3754.
Talent Fund of University Association for Science and Technology in [15] Y. Ma, et al., Conductive graphene fibers for wire-shaped supercapacitors
strengthened by unfunctionalized few-walled carbon nanotubes, ACS Nano 9
Shaanxi (No. 20190402) and Open Research Fund of State Key Labo­ (2015) 1352–1359.
ratory of High Performance Complex Manufacturing, Central South [16] J. Bae, et al., Fiber supercapacitors made of nanowire-fiber hybrid structures for
University (Kfkt2020-08). We also appreciate the support from the In­ wearable/flexible energy storage, Angew. Chem. Int. Ed. 50 (2011) 1683.
[17] X. Pu, et al., Wearable self-charging power textile based on flexible yarn
ternational Joint Laboratory for Micro/Nano Manufacturing and Mea­ supercapacitors and fabric nanogenerators, Adv. Mater. 28 (2016) 98.
surement Technologies. [18] G. Qu, et al., A fiber supercapacitor with high energy density based on hollow
graphene/conducting polymer fiber electrode, Adv. Mater. 28 (2016) 3646–3652.
[19] W. Cai, et al., Transition metal sulfides grown on graphene fibers for wearable
Appendix A. Supplementary data asymmetric supercapacitors with high volumetric capacitance and high energy
density, Sci. Rep. 6 (2016) 26890.
Supplementary data to this article can be found online at https://doi. [20] Q. Zhang, et al., Wrapping aligned carbon nanotube composite sheets around
vanadium nitride nanowire arrays for asymmetric coaxial fiber-shaped
org/10.1016/j.jpowsour.2021.230866.
supercapacitors with ultrahigh energy density, Nano Lett. 17 (2017) 2719–2726.
[21] C. Choi, et al., Improvement of system capacitance via weavable superelastic
References biscrolled yarn supercapacitors, Nat. Commun. 7 (2016) 13811.
[22] Y. Zhao, et al., N-P-O co-doped high performance 3D graphene prepared through
red phosphorous-assisted “cutting-thin” technique: a universal synthesis and
[1] M.D. Lima, et al., Biscrolling nanotube sheets and functional guests into yarns,
multifunctional applications, Nano Energy 28 (2016) 346–355.
Science 331 (2011) 51–55.
[2] Z. Zhang, et al., Integrated polymer solar cell and electrochemical supercapacitor
in a flexible and stable fiber format, Adv. Mater. 26 (2014) 466–470.

9
F. Han et al. Journal of Power Sources 520 (2022) 230866

[23] X. Zhang, et al., Fibrous and flexible supercapacitors comprising hierarchical Chem. A 6 (2018) 896–907. https://pubs.rsc.org/az/content/articlehtml/2018/
nanostructures with carbon spheres and graphene oxide nanosheets, J. Mater. ta/c7ta08362a-fn1.
Chem. A 3 (2015) 12761–12768. [47] X. Yang, C. Cheng, Y. Wang, L. Qiu, D. Li, Liquid-mediated dense integration of
[24] K. Jiang, Q. Li, S. Fan, Spinning continuous carbon nanotube yarns, Nature 419 graphene materials for compact capacitive energy storage, Science 341 (2013)
(2002), 801-801. 534–537.
[25] T. Yamada, et al., A stretchable carbon nanotube strain sensor for human-motion [48] J. Chmiola, et al., Anomalous increase in carbon capacitance at pore sizes less than
detection, Nat. Nanotechnol. 6 (2011) 296–301. 1 nanometer, Science 313 (2006) 1760–1763.
[26] G. Wu, et al., Graphitic carbon nitride nanosheet electrode-based high-performance [49] M. Zhi, C. Xiang, J. Li, M. Li, N. Wu, Nanostructured carbon–metal oxide composite
ionic actuator, Nat. Commun. 6 (2015) 7258. electrodes for supercapacitors: a review, Nanoscale 5 (2013) 72–88.
[27] G. Sun, J. Liu, L. Zheng, W. Huang, H. Zhang, Preparation of weavable, all-carbon [50] C. Largeot, et al., Relation between the ion size and pore size for an electric double-
fibers for non-volatile memory devices, Angew. Chem. Int. Ed. 52 (2013) layer capacitor, J. Am. Chem. Soc. 130 (2008) 2730–2731.
13351–13355. https://onlinelibrary.wiley.com/doi/full/10.1002/ange.2013 [51] M. Salanne, et al., Efficient storage mechanisms for building better supercapacitors,
06770-nss. Nat. Energy 1 (2016) 16070.
[28] C. Liu, Z. Yu, D. Neff, A. Zhamu, B.Z. Jang, Graphene-based supercapacitor with an [52] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc.
ultrahigh energy density, Nano Lett. 10 (2010) 4863–4868. 208 (1958) 1334–1339.
[29] J.R. Miller, R.A. Outlaw, B.C. Holloway, Graphene double-layer capacitor with ac [53] D.H. Long, et al., Preparation of nitrogen-doped graphene sheets by a combined
line-filtering performance, Science 329 (2010) 1637–1639. chemical and hydrothermal reduction of graphene oxide, Langmuir 26 (2010)
[30] X. Yang, et al., A high-performance graphene oxide-doped ion gel as gel polymer 16096–16102.
electrolyte for all-solid-state supercapacitor applications, Adv. Funct. Mater. 23 [54] H. Deng, et al., Toward n-doped graphene via solvothermal synthesis, Chem.
(26) (2013) 3353–3360. Mater. 23 (5) (2011) 1188.
[31] L. Zhang, et al., Porous 3D graphene-based bulk materials with exceptional high [55] J.W. Lee, J.M. Kob, J.D. Kima, Hydrothermal preparation of nitrogen-doped
surface area and excellent conductivity for supercapacitors, Sci. Rep. 3 (2013) graphene sheets via hexamethylenetetramine for application as supercapacitor
1408–1416. electrodes, Electrochim. Acta 85 (2012) 459–466.
[32] L. Sheng, L. Jiang, T. Wei, Z. Fan, High volumetric energy density asymmetric [56] Z. Lin, G. Waller, Y. Liu, M. Liu, C.P. Wong, Facile synthesis of nitrogen-doped
supercapacitors based on well-balanced graphene and graphene-MnO2 electrodes graphene via pyrolysis of graphene oxide and urea, and its electrocatalytic activity
with densely stacked architectures, Small 12 (2016) 5217–5227. toward the oxygen-reduction reaction, Adv. Energy Mater. 2 (2012) 884–888.
[33] J.-J. Shao, W. Lv, Q.-H. Yang, Self-assembly of graphene oxide at interfaces, Adv. [57] D. Hulicova-Jurcakova, M. Seredych, G.Q. Lu, T.J. Bandosz, Combined effect of
Mater. 26 (2014) 5586–5612. nitrogen- and oxygen-containing functional groups of microporous activated
[34] R. Raccichini, A. Varzi, S. Passerini, B. Scrosati, The role of graphene for carbon on its electrochemical performance in supercapacitors, Adv. Funct. Mater.
electrochemical energy storage, Nat. Mater. 14 (2014) 271–279. 19 (2009) 438.
[35] C. Lu, et al., Three-dimensional hierarchically porous graphene fiber-shaped [58] J.P. Pels, F. Kapteijn, J.A. Moulijn, Q. Zhu, K.M. Thomas, Evolution of nitrogen
supercapacitors with high specific capacitance and rate capability, ACS Appl. functionalities in carbonaceous materials during pyrolysis, Carbon 33 (1995) 1641.
Mater. Interfaces 11 (2019) 25205–25217. [59] L. Zhang, et al., Highly conductive and porous activated reduced graphene oxide
[36] Z. Cai, et al., Flexible, weavable and efficient microsupercapacitor wires based on films for high-power supercapacitors, Nano Lett. 12 (2012) 1806–1812.
polyaniline composite fibers incorporated with aligned carbon nanotubes, J. Mater. [60] X. Zheng, K. Zhang, L. Yao, Y. Qiu, S. Wang, Hierarchically porous sheath–core
Chem. A 1 (2013) 258–261. graphene-based fiber-shaped supercapacitors with high energy density, J. Mater.
[37] W. Ma, S. Chen, S. Yang, M. Zhu, Hierarchically porous carbon black/graphene Chem. A 6 (2018) 896–907. https://pubs.rsc.org/az/content/articlehtml/2018/
hybrid fibers for high performance flexible supercapacitors, RSC Adv. 6 (2016) ta/c7ta08362a-fn1.
50112–50118. https://pubs.rsc.org/–/content/articlehtml/2016/ra/c6ra0 [61] D.-W. Wang, F. Li, M. Liu, G.Q. Lu, H.-M. Cheng, 3D aperiodic hierarchical porous
8799j-fn1. graphitic carbon material for high-rate electrochemical capacitive energy storage,
[38] W. Jiang, et al., Space-confined assembly of all-carbon hybridfibers for capacitive Angew. Chem. Int. Ed. 48 (2009) 1525. https://onlinelibrary.wiley.com/doi/full/
energy storage: realizing abuilt-to-order concept for micro-supercapacitors, Energy 10.1002/anie.200702721-nss.
Environ. Sci. 9 (2016) 611–622. [62] K. Wang, Q. Meng, Y. Zhang, Z. Wei, M. Miao, High-performance two-ply yarn
[39] J. Yan, et al., Advanced asymmetric supercapacitors based on Ni(OH)2/graphene supercapacitors based on carbon nanotubes and polyaniline nanowire arrays, Adv.
and porous graphene electrodes with high energy density, Adv. Funct. Mater. 22 Mater. 25 (2013) 1494–1498.
(2012) 2632–2641. [63] D. Cheng, et al., Synergetic pore structure optimization and nitrogen doping of 3D
[40] L. Peng, et al., Ultrathin two-dimensional MnO2/graphene hybrid nanostructures porous graphene for high performance lithium sulfur battery, Carbon 143 (2019)
for high-performance, flexible planar supercapacitors, Nano Lett. 13 (2013) 869–877.
2151–2157. [64] F. er al Wang, Laser-induced nitrogen-doped hierarchically porous graphene for
[41] G. Wu, et al., High-performance wearable micro-supercapacitors based on advanced electrochemical energy storage, Carbon 150 (2019) 369–407.
microfluidic-directed nitrogen-doped graphene fiber electrodes, Adv. Funct. Mater. [65] J. Li, et al., 1T-Molybdenum disulfide/reduced graphene oxide hybrid fibers as
27 (2017) 1702493. high strength fibrous electrodes for wearable energy storage, J. Mater. Chem. A 7
[42] E. Frackowiak, F. Beguin, Carbon materials for the electrochemical storage of (2019) 3143–3149.
energy in capacitors, Carbon 39 (2001) 937. [66] H.B. Wang, T. Maiyalagan, X. Wang, Review on recent progress in nitrogen-doped
[43] Q. Cheng, et al., Polyaniline-coated electro-etched carbon fiber cloth electrodes for graphene: synthesis, characterization, and its potential applications, ACS Catal. 2
supercapacitors, J. Phys. Chem. C 115 (2011) 23584. (2012) 781–794.
[44] J. Meng, et al., Enhancing electrochemical performance of graphene fiber-based [67] J. Han, et al., Generation of B-doped graphene nanoplatelets using a solution
supercapacitors by plasma treatment, ACS Appl. Mater. Interfaces 10 (2018) process and their supercapacitor applications, ACS Nano 7 (2013) 19–26.
13652–13659. [68] N. Jäckel, P. Simon, Y. Gogotsi, V. Presser, Increase in capacitance by
[45] Y. Xu, et al., Holey graphene frameworks for highly efficient capacitive energy subnanometer pores in carbon, ACS Energy Lett. 1 (2016) 1262–1265.
storage, Nat. Commun. 5 (2014) 4554. [69] C. Largeot, et al., Relation between the ion size and pore size for an electric double-
[46] X. Zheng, K. Zhang, L. Yao, Y. Qiu, S. Wang, Hierarchically porous sheath–core layer capacitor, J. Am. Chem. Soc. 130 (2008) 2730–2731.
graphene-based fiber-shaped supercapacitors with high energy density, J. Mater. [70] X. Zhao, B. Zheng, T. Huang, C. Gao, Graphene-based single fiber supercapacitor
with a coaxial structure, Nanoscale 7 (2015) 9399–9404.

10

You might also like