Journal Pre-proof Synthesis of a novel hybrid anode nanoarchitecture of Bi2O3/porous-RGO nanosheets for high-performance asymmetric supercapacitor Lakshmanan Gurusamy, Sambandam Anandan, Na Liu, Jerry J. Wu PII:
S1572-6657(19)30757-X
DOI:
https://doi.org/10.1016/j.jelechem.2019.113489
Reference:
JEAC 113489
To appear in:
Journal of Electroanalytical Chemistry
Received Date: 25 April 2019 Revised Date:
12 September 2019
Accepted Date: 12 September 2019
Please cite this article as: L. Gurusamy, S. Anandan, N. Liu, J.J. Wu, Synthesis of a novel hybrid anode nanoarchitecture of Bi2O3/porous-RGO nanosheets for high-performance asymmetric supercapacitor, Journal of Electroanalytical Chemistry (2019), doi: https://doi.org/10.1016/j.jelechem.2019.113489. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Elsevier B.V. All rights reserved.
Graphical Abstract
1
Synthesis of a Novel Hybrid Anode Nanoarchitecture of Bi2O3/Porous-RGO
2
Nanosheets for High-Performance Asymmetric Supercapacitor
3 Lakshmanan Gurusamya, Sambandam Anandanb, Na Liuc, and Jerry J. Wua,*
4 5 6
a
Department of Environmental Engineering and Science, Feng Chia University, Taichung,
7
Taiwan b
8 9
c
Department of Chemistry, National Institute of Technology, Trichy, India
College of New Energy and Environment, Jilin University, Changchun 130021, China
10
*Correspondence, E-mail:
[email protected], Fax: +886-4-24517686; Tel: +886-4-
11
24517250 Ext. 5206
12 13 14
Abstract
15
Novel bismuth oxide (Bi2O3) nanoparticles dispersed on porous reduced graphene oxide
16
nanosheets is prepared using the facile hydrothermal reaction followed by a calcination process
17
in the air atmosphere. At electrochemical study, the electrode materials of Bi2O3/porous-RGO
18
display the capacitance retention to be 81.1% at a current density of 0.5 Ag-1 and α-MnO2-NRs
19
exhibit about 80.7 % at a scan rate of 10 mVs-1 for 3000 cycles in 6 M KOH electrolytes of
20
three-electrode configuration. Moreover, the outstanding capacitance retention of anode and
21
cathode materials mainly due to the porosity (porous-RGO), thermal stability with maximal
22
weight loss rate temperature T(mwlr) reach of 623 oC, a smaller size of Bi2O3 (~7.5± 0.5 nm), and
23
aspect ratio of α-MnO2 nanorods for 5.1 ± 0.9 nm. The assembled asymmetric supercapacitor
24
(ASC) achieves the specific capacitance of 84 Fg-1 at a scan rate of 5 mVs-1 and capacitance 1
25
retention of 91.4% at a current density of 1 Ag-1 by the Bi2O3/porous-RGO//α-MnO2-NRs in
26
PVA/KOH gel electrolyte of two-electrode configuration. Notably, the ASC delivers an energy
27
density of 86 Wh kg-1 (787 mF cm-2) at a power density of 9000 W kg-1. As a result,
28
Bi2O3/porous-RGO and α-MnO2-NRs is considered as a promising candidate for future
29
anode/cathode material in ASC energy storage device.
30 31
Keyword: Nanosheets, Hydrothermal, Porous-RGO, Carbon gasification, Anode materials
32 33
1. Introduction
34
Globally, it needs more efforts in the development, demonstration, and utilization of
35
electrochemical energy storage technology. The worldwide energy crises would occur because
36
the problems have been arisen through the insufficient renewable energy storage and conversion
37
technologies [1]. Among all energy storage technologies, the supercapacitor is the state-of-the-
38
art device, which fulfills next-generation energy demands for developing portable electronic
39
devices, such as camera, mobile phones, memory backup systems, and electrical and hybrid
40
vehicles [2,3]. Furthermore, it bridges the gap between high-energy chemical batteries and high
41
power traditional dielectric capacitor. Generally, supercapacitor electrode materials can be
42
classified into two types, namely electrical double layer electrode materials (EDLEM) and
43
Faradaic electron-transfer electrode materials (pseudocapacitors) [4, 5]. Development of EDLEM
44
mechanisms is based on the Helmholtz electrical double layers (HEDL) upon polarization. For
45
example, carbon nanofibers (CNFs), activated carbons, graphene, and porous-reduced graphene
46
oxide (porous-RGO) are often selected as electrode materials for supercapacitors. The
47
pseudocapacitive electrode stores the electrical charges through the electron transfer between the
2
48
electrodes by the faradic reversible redox reaction, such as transition-metal oxides/hydroxides
49
and conductive polymers [6-8]. However, the pseudocapacitive electrode materials have some
50
disadvantages, such as lower rate performance and electrical conductivity, and capacity loss
51
upon cycles [9-10]. To overcome these drawbacks, we need to prepare a hybrid nanoarchitecture
52
of metal oxide nanoparticles decorated on the carbon framework [11-14].
53
Supercapacitors were more dominating in the commercial market because of the porous
54
carbon materials employed as both anode and cathode electrode for electrical double layer
55
capacitors (EDLCs). Usually, EDLCs can provide fast charging/discharging rate, excellent
56
electrochemical kinetics, and good electrical conductivity. However, they still suffer from the
57
low specific capacitance with decreased energy density to hinder asymmetric supercapacitor
58
(ASC) technology [15, 16]. To overcome this obstacle, people need more efforts by exploring
59
new pseudocapacitive electrode materials, such as Mn3O4, Fe2O3, MoO3, Bi2O3, and so forth.
60
Among those, nanoscale bismuth (Bi)-based materials possess a low environmental impact,
61
cheap cost, and rich in natural resources. Additionally, the implementation of Bi2O3
62
nanoparticles has a suitable voltage window with high theoretical specific capacitance (690 mAh
63
g-1) for the negative side, which makes them promising candidates for anode materials in ASC
64
technology. Moreover, it has been widely used for different kinds of applications, such as
65
catalysts, optical materials, gas sensors, and supercapacitors, etc. Meanwhile, the recent reports
66
on Bi2O3 nanoparticles displayed the pseudocapacitance properties, but exhibited the poor
67
capacitance, lower accessible area, and difficulty for the ion and electron transport from the
68
electrolyte to the electrode surface. To overcome these shortcomings, several studies have
69
rationally prepared and fabricated nano-architecture of Bi2O3 nanoparticles decorated carbon
3
70
nanocomposites to succeed the high surface area for boosting the electrochemical performance
71
[17, 18].
72
Among all the carbon materials, graphene as a novel member of carbon morphology has a
73
freestanding atom-thick monolayer of sp2 hybridized carbon atoms with covalently bonded 2D
74
basal planes. In addition, graphene has been suggested as versatile electrode materials due to its
75
excellent electrical conductivity (5×10-3 S/cm) for 2.6% weight of RGO mixed with 97.4% of
76
H2O, thermal conductivity (~ 5000 Wm-1k-1), and strong young’s modulus (~1 TPa) [19, 20].
77
The restacking layers of graphene nanosheets are remarkably agglomerated due to the impaired
78
van der Waals force of attraction or (π-π) conjugation. The ion diffusion path is perpendicular to
79
the basal plane graphene owing to the ion-accessible surface areas, which have been
80
disintegrated between the adjacent layers of graphene. It is essential to improve the diffusion
81
kinetics of electrolyte ions into the electrode surface. Therefore, the penetration of these
82
electrolyte ions mostly belongs to the porous architectures of interconnected pores of RGO [21,
83
22].
84
The porous-RGO electrodes have been synthesized by several methods, such as self-
85
assembly [23], template approach [24] and activation method [25]. Among these, the hard
86
template technique is one of the most common, inarguable, and effective established methods of
87
preparing porous graphene. Nevertheless, it still possesses many disadvantages, such as
88
expensive cost, multiple steps, lack of structural robustness, massive use of template agents, and
89
inherent difficulty. To solve this problem, the porous-RGO nanosheets were prepared by the
90
effective technique of catalytic carbon gasification (CO, CO2) method without using any
91
template assistance [26, 27]. Hence, the synthetic strategy of the porous-RGO electrode was
92
adopted through the restacked layer behavior of lamellar microstructure of graphene. The
4
93
porous-RGO is more contributing to pseudocapacitive performances (capacity and cyclic
94
stability) of metal oxide nanoparticles due to possessing multiple pore channels with mesoporous
95
architecture. This evidence is the most favorable for interfacial charge transfer, thus facilitating
96
onto superior contact occurring in between the electrolyte and porous graphene materials.
97
Therefore, the porous-RGO architectures are well suitable for supporting redox species of metal
98
oxide nanoparticles [28-30].
99
In this work, the development of high-performance ASC has been conducted with much
100
effort. In the negative side, the Bi2O3/porous-RGO was prepared using catalytic carbon
101
gasification method in the air atmosphere. The anode materials exhibited a highest specific
102
capacitance and capacitance retention due to the nanosized Bi2O3 nanoparticles which were
103
strongly decorated on the porous-RGO nanosheets. Mesoporous morphology with the high
104
specific surface area (SSA) has promoted the full utilization of the transport both ion and
105
electron. In the positive side, the α-MnO2-NRs (nanorods) were prepared using the hydrothermal
106
process. It exhibited the higher electrochemical performance due to nanosized α-MnO2-NRs with
107
the higher aspect ratio of 5.1 ± 0.9 nm and enhanced interfacial charge transfer reaction between
108
electrode/electrolyte surfaces. Inspired by these features, the Bi2O3/porous-RGO and α-MnO2-
109
NRs electrodes could play a decisive role in the overall electrochemical performance of ASC.
110 111
2. Materials and Methods
112
2.1. Materials
113
Graphite powder (200 mesh, ≤ 74 µm), sodium nitrate (NaNO3), sulfuric acid (H2SO4,
114
95%), potassium permanganate (KMnO4, ≥99.0 %), hydrogen peroxide (H2O2), bismuth nitrate
115
Bi (NO3)3), aqueous ammonia (NH3.H2O), manganese sulfate dihydrate (MnSO4.2H2O), ultra-
5
116
pure water (18.2 mΩ), ethanol (C2H5OH), nickel foam (1.6 mm thickness), carbon black, and
117
polytetrafluoroethylene (PTFE), were purchased from Sigma Aldrich. The entire analytical grade
118
chemicals were used without further purification.
119
2.2. Preparation of graphene oxide from graphite powder
120
According to Hummer’s method, we have prepared graphene oxide (GO) from purified
121
natural graphite powders [31, 32]. In a typical synthesis, 1 g of graphite powder was mixed with
122
0.5 g of sodium nitrate (NaNO3) and dissolved in the minimum amount of water. After that, 23
123
ml of concentrated sulfuric acid (H2SO4) was added under a uniform stirring for 2 hours.
124
Furthermore, 5 g of KMnO4 was gradually added to the mixture of above solution then mixed for
125
overnight. The temperature is kept at below 5 oC by adding ice cubes to prevent overheating and
126
explosion. Then, the mixture was continuously stirred for 12 hours and the resulting solution was
127
diluted by adding 500 ml water under vigorous stirring. After that, the suspension was further
128
treated with 30% H2O2 solution (5 ml), realizing a color change of the suspension to light brown
129
to yellow, to confirm the completion of the reaction. The resultant products are washed with HCl
130
and water for several times until to remove the excess of salt impurities. The dry graphene oxide
131
obtained by drying the final product at 60 oC for 24 hrs.
132
2.3. Catalytic carbon gasification synthesis of Bi2O3 nanoparticles (NPs) dispersed on
133
Porous-RGO nanosheets
134
The Bi2O3/porous-RGO electrodes were synthesized via the following process. Concisely,
135
250 mg of graphene oxide (GO) was dispersed in 100 ml of water under sonic-probe irradiation
136
for 10 min. The resulting suspension was subjected to centrifugation at 5000 rpm for 10 min, and
137
thus obtained GO suspension. Then, 200 mg of Bi(NO3)3 was added into GO suspension
138
followed by intense agitation for 10 min. Then, 6 ml of NH3.H2O (30%) dropwise was added to
6
139
the above suspension for 30 min; finally, the resultant mixture light yellow viscous suspension
140
was transferred into a stainless steel vessel and subjected to the hydrothermal reduction at 180 oC
141
for 12 hrs, then cool to room temperature naturally. Afterward, the obtained bulk-Bi2O3/RGO
142
(Intermediate product) was washed with ultra-pure water (18.2 mΩ) and ethanol until to remove
143
the unnecessary impurities. This intermediate product was thermally treated in the furnace at 400
144
o
145
final product of Bi2O3/porous-RGO electrode. The same procedure was followed to prepare the
146
Bi2O3 and RGO under the similar condition but without calcination.
147
2.4. Preparation of α-MnO2 nanorods (NRs)
C for 4 hrs in air. Subsequently, the furnace has to reach the room temperature and collect the
148
The α-MnO2 nanorods were prepared by adding 15 ml of 0.002 M MnSO4.H2O dissolved
149
in 50 ml distilled water with stirring for 10 minutes. Then, 15 ml of 0.005 M KMnO4 was added
150
into the above mixture solution with stirring about 20 minutes. After 30 minutes, the entire
151
solution became clear. Subsequently, the transparent solution was transferred into Teflon-lined
152
stainless steel autoclave (100 ml) of 80 % capacity of the total volume. The autoclave was locked
153
and placed into the muffle furnace and maintained at 140 °C for 12 h. When the reaction was
154
completed, the autoclave temperature was automatically cooled down to room temperature. The
155
obtained precipitate of α-MnO2 nanorods were washed several times by water and ethanol and
156
dried at 70 °C for 6 hours.
157
2.5. Sample characterization
158
The morphology and structural characterization of the entire samples were observed by the
159
field emission scanning electron microscopy (FE-SEM; JEOL, JSM-7610F) and high-resolution
160
transmission electron microscopy (HRTEM; JEOL, JEM2010) at an acceleration voltage of 200
161
kV. The presence of carbon, bismuth, and oxygen element in the nanocomposite of
7
162
Bi2O3/Porous-RGO were evaluated by energy dispersive X-ray spectrum (EDS) and elemental
163
mapping analysis equipped within the HRTEM; JEOL, JEM2010. The XRD pattern was
164
recorded on a Philips XPertPro X-ray diffractometer with Cu Kα radiation (λ=1.5418) of
165
operating on 40 kv, 60 mA at a scanning speed of 5 oC/min. Furthermore, the FT-IR spectra were
166
used to determine the stretching frequency of all as-prepared electrode materials through the KBr
167
pellet technique (Nicolet FT-IR 380 Spectrometer). Nitrogen adsorption/desorption isotherm was
168
measured by the micromeritics ASAP 2020 instrument at temperature 77K. The specific surface
169
area and pore size distribution curves (PSD) of all electrodes was calculated by the BET and BJH
170
method, and the weight of the samples is 0.110 g. Thermogravimetric analysis was carried out
171
(AutoTGA2950-V54A) by flowing the air gas condition for applying the temperature starting
172
from room temperature to 1000 oC under a heating rate of 10 oC min-1.
173
2.6. Electrochemical Measurements
174
2.6.1. Three-electrode fabrication process The working electrodes were prepared as follows: about (80:10:10) % of Bi2O3/porous-
175 176
RGO
active
materials
(8
mg),
conductive
177
polytetrafluoroethylene (PTFE) (1 mg) were mixed together in a test tube containing 1 ml
178
ethanol and vigorously stirred for an overnight. After that, the resulting slurry was coated
179
uniformly onto the nickel foams (about 10 mg of active material over a geometric surface area of
180
1 cm2). Finally, the electrodes were dried for 10 hrs at 80 oC in the open-air atmosphere
181
condition. All the electrochemical measurements were carried out by an electrochemical
182
workstation (Auto lab PGSTAT128N) in a three-electrode system in 6 M KOH aqueous
183
electrolyte solution at room temperature. A Pt-wire and saturated Hg/HgO acted as the counter
184
and reference electrodes, respectively. Cyclic Voltammetry (CV), Galvanostatic Charge and
8
carbon
black
(1
mg),
and
binder
185
Discharge (GCD), and cyclic durability test were performed under an ambient condition within
186
the range of potential -1.2 to 0.2 V (vs Hg/HgO). Electrochemical impedance spectroscopy (EIS)
187
measurements were carried out in the frequency range from 0.01 Hz to 100 kHz. Moreover, the
188
specific capacitance is calculated from the CV curve using the following equation as given
189
below: =
1
190
where V and V0 are the initial and final potential of the CV curve, I is the current density
191
(Ag-1), m is the mass of the electrode materials (gm) on the nickel foam, and ‘s’ is representing
192
the scan rate (mVs-1). Subsequently, the specific capacitance (Csp, Fg-1) of the single electrodes
193
was calculated from the non- linear galvanostatic charge/discharge curve (GC/GD) at different
194
current density based on the formula as follows: &
/! = " #
&'(
$ % 2
195
The discharging energy can be calculated from integrated area of the GD curves, charging
196
energy /! from GC. Where /! is the average discharge capacitance or claimed
197
capacitance when C is the function of U, ‘I’ is the specific discharge current (Ag-1), and ‘t’ is the
198
discharge time (s) [33].
199
2.6.2 Asymmetric supercapacitor fabrication process
200
A solid-state asymmetric supercapacitor fabrication process is presented in supporting
201
information Fig. S1, where α-MnO2 NR and Bi2O3/porous-RGO NS are functioned as positive
202
and negative electrodes, respectively. First, anode or cathode material was crushed for 10
203
minutes then became a fine powder. Then, the slurry was prepared by the weight of anode or
204
cathode material (80:10:10) %, PTFE and carbon black mixed with ethanol (10 mg material into
205
1 ml ethanol). The obtained slurry was added on cutting surface 12 cm-2 of the nickel foam (8 mg 9
206
cm-2) and dried at 60 oC for 8 hours. Next, the polymeric electrolyte was prepared by adding 6 g
207
KOH with 6 g PVA dissolved into the 60 mL distilled water under stirring for 1 hour at 85 oC.
208
The polymeric electrolytes were obtained by the more transparent solution changed in the gel
209
form at increased time and constant temperature. After that, gel electrolytes were added into
210
already coated nickel foams surfaces and dried at 80 oC for overnight. Finally, the cellulose
211
papers were used as a separator and inserted between the anode (+) and cathode (-). The two
212
ends were wrapped with copper tap and connected with the electrochemical workstation (Auto
213
lab PGSTAT128N) in a two-electrode system. Areal capacitances of the electroactive materials
214
were calculated from their GCD curves according to the following equation:
215
Areal capacitances 56 = " × ∆ /5 × ∆V (3)
216
where Ac is the areal capacitance (F/cm2), I is the discharging current density (A/g), ∆ is
217
the discharging time, and ∆V (V) is the potential window. The electrode materials were coated on
218
nickel foam with surface area A (cm2) of 12 cm2 Bi2O3/porous-RGO NS and 12 cm2 α-MnO2
219
NR.
220
The energy density and power density was calculated from the following equations: (4) and (5).
221
Energy density > = 0.5 × × ∆B C 4
222 223
where ED is the energy density (Wh kg-1), Cs is the specific capacitance (Fg-1), and ∆E is the change in the potential window (V), respectively. Power density I> = > × 3600/∆ 5
224 225
where PD is the power density (W kg-1) and ∆ is the discharge time from the GCD curve, respectively.
226 227
3. Results and discussion 10
228
The synthetic procedure of Bi2O3/porous-RGO anode material is displayed in Scheme. 1.
229
First, GO powders were dispersed in water under the ultrasonic assistance, then the Bi(NO3)3
230
solution was added to the GO dispersion during the stirring process. The Bi(NO3)3 was
231
uniformly dispersed on the surface of GO suspension with the help of the sp2 carbon atoms
232
(vacant ‘p’ orbitals), greater surface energy also fewer functional groups [11, 34]. During the
233
hydrothermal reaction, the Bi(NO3)3 was dissociated into the Bi3+ cation and NO3- anionic
234
species at the temperature above 130 oC. The cationic species of Bi3+ is converted into Bi(OH)3
235
by the addition of aqueous ammonium hydroxide. Then, the anionic species of NO3- was further
236
decomposed into NO2- and O2. Furthermore, upon increasing the hydrothermal temperature
237
gradually, reduced size of the Bi2O3 NPs was obtained by the dehydration of H2O molecules.
238
Reduced size of Bi2O3 NPs was uniformly located on the RGO nanosheets because the RGO has
239
excellent physical properties, such as chemisorption and strong electrostatic force of interaction
240
between metal to carbon bond. However, the Bi2O3/porous-RGO anode was prepared by
241
applying calcination temperature at 400 oC for precise time of 4 hours in air atmosphere on bulk-
242
Bi2O3/RGO electrodes. The Bi2O3/porous-RGO Ns were prepared from through catalytic carbon
243
gasification using Bi2O3-NPs (nanocatalysts) obtained in situ to generate mesoporous structures
244
as revealed in Scheme. 1. However, the porous morphology created by the O2 (air oxidation)
245
atoms can react with carbon (RGO-Ns), which was selectively decomposed, in the presence of
246
adjacent Bi2O3 NPs at the temperature lower than carbon combustion temperature. The remaining
247
Bi2O3 NPs and unreacted carbon atoms retained in these composites. The CO2 can also react with
248
H2 (intercalation reaction) according to the reverse water gas shift reaction: CO2 + H2 ↔ H2O +
249
CO (Scheme. 1). The reaction process is expressed as follows in equation (6). Therefore, the
250
formation of porous morphology depends on the number and average size distribution of the
11
251
Bi2O3 NPs. The creation of mesopores on RGO nanosheets can be confirmed by FE-SEM (Fig.
252
S2 (d)) and HR-TEM images in Fig. 1d. 6'TU'V
LMN + P QR + SC WXXXXXXXY Porous − RGO + CO/ SC + P QR 6 253 254
3.1. Morphology and structural analysis of anode materials
255
The FE-SEM image of RGO, Bi2O3, bulk-Bi2O3/RGO (before calcination), and
256
Bi2O3/porous-RGO (after calcination) of as-synthesized materials are as shown in Fig. S2 (a-d).
257
The formation of well-defined Bi2O3 nanoparticles (NPs) can uniformly agglomerate on the
258
surface of porous reduced graphene oxides framework as displayed in Fig. S2 (d). More
259
importantly, the hybrid Bi2O3/porous-RGO architectures formed by the conductive matrix of
260
porous-RGO nanosheets act as a nucleation site for growing a Bi2O3 NPs due to numerous
261
interconnected pores. Besides, the well-aligned Bi2O3 NPs are able to provide plenty of redox
262
active sites on the surface of porous-RGO. Owing to the benefit to facilitation, the quick
263
migration of electrolyte ion into electroactive sites was done at higher charge/discharge current
264
density. Finally, it efficiently improves the electrochemical reaction kinetics during buffer
265
volume change of electrochemical measurements and thus leads to supercapacitor performance.
266
On the other hand, Fig. S2 (c) shows that Bi2O3 NPs were irregularly dispersing on the RGO
267
nanosheets to yield bulk-Bi2O3/RGO (before calcination), whose morphology is not suitable for
268
the penetration of electrolyte ions for capacitive performance. Besides, for comparison, the
269
porous morphology of Bi2O3/porous-RGO electrode materials was prepared at various time
270
intervals (2, 3, and 4 hours) with a constant temperature of 400 oC and the results are presented
271
in Fig. S3.
12
272
The detailed morphological structure of as-synthesized electrode materials, such as the
273
bulk-Bi2O3/RGO and Bi2O3/porous-RGO, was further confirmed by TEM and HR-TEM
274
analysis. The bulk-Bi2O3/RGO nano-hybrid material was successfully prepared by the one-pot
275
hydrothermal method, which shows that the Bi2O3 nanoparticles are anchoring thoroughly on the
276
RGO sheets as displayed in Fig. 1a. Moreover, the bulk-Bi2O3/RGO was converted into the
277
Bi2O3/porous-RGO through the calcination temperature and it is shown in Fig. 1c. Before
278
calcination, the average size of the Bi2O3 nanoparticles (10.6 ± 0.4 nm, Fig. S4 (a)) was
279
calculated from the bulk-Bi2O3/RGO. After calcination, the average size of the Bi2O3
280
nanoparticles (7.5 ± 0.5 nm, Fig. 1b and Fig. S4 (b)) was also calculated from the Bi2O3/porous-
281
RGO (Fig. 1c). As shown in Fig. 1d, the composite of Bi2O3/porous-RGO had the well-defined
282
porous channel, where the porous structure not only increases the surface active site, but also the
283
electrolyte ions easily penetrate both intercalation/de-intercalation processed at higher current
284
density. As can be seen in Fig. 1e, the Bi2O3 nano-particles are well enclosed by the porous
285
structure of graphene. The HR-TEM measurement (Fig. 1f) provided detailed information about
286
crystal lattice of Bi2O3 nanoparticles chemically bonded on the RGO layer in Bi2O3/porous-RGO
287
composite. The different diameter of selected Bi2O3 nanoparticles was calculated in
288
Bi2O3/porous-RGO composite as shown in Fig. S5a. The Bi2O3 nanoparticles encircled (Fig. S5
289
(b-c)) about 1.5 nm thickness on RGO layer (measured by Image-J software) and this is because
290
the size of the Bi2O3 crystallinity increased due to the movement from hydrothermal temperature
291
180 oC to annealing temperature at 400 oC as presented in Fig. 1f. Moreover, the clear lattice
292
fringes with an interplanar distance of 0.32 nm correspond to the Bi2O3 crystal plane (Fig. S5d)
293
of the nanoparticles facet (021), which values were measured from the FFT and IFFT spectrum
294
(Gatan Microscopy suite software) as depicted in Fig. 1g-i.
13
295
The elemental mapping analysis (Fig. S6 (a-c)) of Bi2O3/porous-RGO obviously confirms
296
the homogeneous distribution and coexistence of Bi, O, and C element. Such observations
297
unambiguously confirm that the Bi2O3 nanoparticles have been successfully anchored on RGO
298
nanosheets to form the Bi2O3/porous-RGO material. These results are in good concordance with
299
the XPS spectrum in Fig. S10b. In addition, the EDX spectrum of the corresponding
300
Bi2O3/porous-RGO material (Fig. S6 (d)) was further confirmed in the presence of C, O, and Bi
301
element with increasing weight % and decreasing the atomic %.
302
The crystallographic structure of the Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and
303
Bi2O3 electrode materials were further determined by the X-ray diffraction (XRD) measurements
304
as shown in Fig. 2a. The crystalline structure of the Bi2O3 electrode was observed in the cubic
305
crystal systems of hydrothermally formed bulk-Bi2O3/RGO and tetragonal crystalline structures
306
of calcinated Bi2O3/porous-RGO as displayed in Fig. 2a. This phase transformation may occur
307
because Bi2O3 NPs was able to create the voids morphology of supporting materials as
308
mentioned in the previous publication [35]. Remarkably, two sharp hump RGO diffractions
309
peaks were clearly detected at the position of 2θ=26.48 and 54.73, belonging to the typical (120)
310
and (023) reflection planes. As can be seen that the many typical sharp diffraction peaks intensity
311
of Bi2O3 in Bi2O3/porous-RGO and bulk- Bi2O3/RGO increased gradually due to the presence of
312
graphene, also fewer oxygen-containing functional groups (Fig.S10c) with strong electrostatic
313
force of attraction between Bi2O3 and porous-RGO [36]. Among the diffraction peaks, the
314
leading sharp diffraction peak appears at the position of 2θ=26.5, 27.3, 30.5, and 32.78, which
315
can be indexed to the (120), (021), (222), and (132) crystal planes of enlarging pattern as shown
316
in Fig. S7. It is confirmed that the calculated interlayer spacing and crystallite size of the sharp
317
diffraction peak of Bi2O3 in Bi2O3/porous-RGO calculated at the position of d (021) is 0.32 nm and
14
318
7.3 nm, respectively, according to the Braggs and Scherer equation as presented in supporting
319
information (Table-S1 and Table-S2). These values were further compared in the HR-TEM
320
image, where almost all are consistent with Bi2O3 NPs size and lattice spacing values of FFT
321
spectrum.
322
The porous structure of as-prepared electrode materials was further explored by the
323
nitrogen adsorption−desorption techniques as depicted in Fig. 2b. The corresponding pore-size-
324
distribution (PSD) curved of all four-electrode materials displays that pore distribution ranging
325
from 2 to 7 nm and resultant pore volumes of 0.071, 0.050, 0.008, and 0.0037 cm3/g,
326
respectively. As a result, it is obviously shown the wide PSD curve of Bi2O3/porous-RGO
327
electrode material, i.e., it exhibited a high capacitance due to the high specific surface area and
328
plentiful mesoporous (2-50 nm), which may be favorable for electrical double layer capacitance
329
(EDLCs) of porous graphene. Further, it can also improve the electrical conductivity upon Bi2O3
330
NPs, combining with the porous RGO to developing pseudocapacitive performance.
331
Furthermore, the mesoporous RGO morphology was created by the catalytic effect of Bi2O3 NPs,
332
which were done by simultaneous pyrolysis and the etching of RGO NPs. This kind of pore
333
nanostructure was very important to be employed for improving the rate capability at higher
334
discharge/charge current density. Moreover, the application of mesoporous RGO structures was
335
further utilized for electrolyte ion to rapidly diffuse into the accessible observed porous surface
336
area and their satisfactory amount of pseudocapacitive performance was improved on Bi2O3 NPs
337
involved in the redox reaction mechanisms. Additionally, the N2 adsorption isotherms data was
338
presented in Fig. S8. All of them belong to type IV isotherms with a distinct type of H3
339
hysteresis loop at a high relative pressure (P/P0) between 0.4 to 1.0, which suggests the presence
340
of mesoporous solids characteristics. The calculated specific surface area (SBET) is 40.4, 23.8,
15
341
10.8, and 7.9 m2/g for Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and bare Bi2O3, respectively.
342
As a result, the high specific surface area of Bi2O3/porous-RGO is a potential electrode material
343
for supercapacitor applications.
344
The stretching frequency of functional groups moieties in all the as-prepared materials was
345
determined from FT-IR spectra as seen in Fig. 2c. GO exhibited series of characterized
346
absorption bands, which are obviously related to the different types of oxygen-rich functional
347
groups of both surfaces and edges of the GO nanosheets, such as located at 3427 cm-1 (broad
348
absorbance band of O–H stretching vibration in hydroxyl or carboxyl groups), 3736 cm-1 (sharp
349
deformation peak of O-H distinctive vibration), and 1631 cm-1 (C=C stretching vibrations and
350
breathing vibration of epoxy groups), respectively. The weak peak arising from 1319 cm-1
351
corresponds to a strong C-O stretching frequency of carbonyl groups, while the 1065 cm-1
352
characteristic frequency can be assigned to a robust C-O bond between alcoholic functional
353
group presented on an RGO nanosheet. The weak bands centered around 1371 cm-1 could be
354
ascribed to N-O bond between nitro groups of the RGO nanosheets (‘N’ comes from the
355
precursor of aqueous ammonia or Bi(NO3)3) [37]. These results are good agreements with XPS
356
spectrum of Fig. S10d. Furthermore, the strong narrow peak centered at 831 cm-1 is assigned the
357
predictable vibration of the Bi-O-Bi bonds of Bi2O3 materials. The very strong intense peak at
358
534 cm-1 is absorption band due to the vibration of Bi-O bonds (Fig. S10e) between [BiO6] units.
359
These are the typical vibrations present in Bi2O3/porous-RGO and Bi2O3 electrode materials that
360
confirmed the incorporation of Bi2O3 NPs on the RGO nanosheets [38, 39]. Meanwhile, the bulk-
361
Bi2O3/RGO did not appear metal-oxygen bond [Bi-O] stretching vibration at the position of 534
362
cm-1, which means that hydrothermal temperature is not enough for the formation of (Bi) metal
363
to oxygen bond between these electrode materials and thus hinders the pseudocapacitive
16
364
performance. Therefore, it is necessitated to follow the calcination temperature for formation of
365
metal to oxygen bond and simultaneously to generate the pores on RGO nanosheets. These
366
results are good concordance with the XPS survey spectrum of Fig. S9a and S10a.
367
Generally, the weight loss, thermal stability, and pyrolysis behavior of all as prepared
368
electrode materials have been investigated by the thermogravimetric analysis (TGA) under the
369
flowing of air gas atmospheric condition with a heating rate of 10 oC min-1 as presented in Fig.
370
2d. From the TGA curve, the first weight loss was carried out between 30 oC to 200 oC, which
371
were attributed to the evaporation of physically adsorbed/intercalated water molecules or may be
372
due to moisture in these samples [40, 41]. The second weight loss was located in the temperature
373
range of 200 oC to 350 oC due to the formation of Bi to Bi2O3. More importantly, the third major
374
significant weight loss of the Bi2O3, Bulk-Bi2O3/RGO, RGO and Bi2O3/porous-RGO electrode
375
materials were observed (Table S3) for temperatures belonging to (i) 355 oC to 545 oC, (ii) 415
376
o
377
determine the maximal weight loss rate temperature T(mwlr) around 355 oC and followed by
378
weight loss occurred for 6.2%. Hence, we decided to prepare Bi2O3/porous-RGO anode material
379
for suitable calcination temperature at 400 oC, which may ascribe to the phase transformation of
380
Bi2O3 from bulk-Bi2O3/RGO to Bi2O3/porous-RGO [42]. Furthermore, the content of the case (ii)
381
was observed with weight loss at 33% of RGO (retain) and 67 % of Bi2O3 for bulk-Bi2O3/RGO.
382
Case (iii) can be designated for 91.6% weight loss occurring for RGO nanosheets, which is due
383
to the bulk pyrolysis of the carbon skeleton (RGO sheets). We can also evaluate the mass content
384
of the case (IV) with 27% of RGO (retain) and 73% of Bi2O3, respectively, which are almost in
385
good agreement with an EDX spectrum. Finally, the thermal stability of the Bi2O3/porous-RGO
386
is greater upon compared with other electrodes because the unexceptional T
C to 847 oC, (iii) 522 oC to 734 oC, and (iv) 623 oC to 901 oC. In case (i), it represented to
17
(mwlr)
were attained
387
at the temperature of 623 oC and it was completely decomposed from C to CO or CO2. These
388
explanations designate that the Bi2O3 nanoparticles can catalyze the oxidation of RGO
389
nanosheets that were in contact with them. For 6 hrs at 400 oC, the thermal stability of the
390
Bi2O3/porous-RGO electrode materials exhibits (Fig. S11) T (mwlr) to be 638 oC.
391
3.2. Morphology and structural analysis of cathode material
392
The TEM images of α-MnO2-NRs are shown in Fig. S12 (a). As can be seen in Fig. S12,
393
the pure and uniform α-MnO2-NRs were observed with the length of 165 ± 5 nm (Fig. S13 (a))
394
and its width of 32.7 ± 0.3 nm (Fig. S13 (b)) (calculated in image J software). The SAED pattern
395
(inset) exhibited a line pattern of single crystalline nature from the single nanorods as displayed
396
in Fig. S12 (b). The shape of the α-MnO2-NRs clearly shows a blind stick with an aspect ratio
397
(length/width) of 5.1 ± 0.9 nm as seen in Fig. S12 (c-d). The HR-TEM images of well-defined
398
lattice line are displayed in Fig. S12 (e-f). The FFT and IFFT spectrum and d-spacing values
399
were analyzed using digital microscopy software as shown in Fig. S12 (g-i). The interlayer
400
spacing of α-MnO2-NRs is about d = 0.2845 nm, which corresponds to the (211) crystal planes
401
of the tetragonal structure. The EDX spectrum was used to confirm the elements composed in
402
electrode materials. Therefore, the EDX peaks have demonstrated Mn and O elements present in
403
α-MnO2-NRs and resultant atomic/weight % as shown in Fig. S13 (c). No other impurities were
404
detected, which indicated the high quality of α-MnO2-NRs was obtained from the preparation
405
process.
406
As depicted in Fig.3a, the high yield of α-MnO2-NRs was observed by FE-SEM images.
407
The blind stick shape of the α-MnO2-NRs was obtained and the maximum size of the NRs length
408
and width are 180 nm and 23 nm, respectively. Subsequently, the crystallographic structure and
409
phase purity of α-MnO2-NRs has been measured after hydrothermal time of 12 hours by the
18
410
XRD pattern. As shown in Fig. 3b, the diffraction peaks angles, appeared at 2θ = 12.7°, 15.4°,
411
24.8o, 28.8°, 37.4°, 42.2°, 49.9°, 57.1°, 60.1°, 65.5o, 69.7o, and 73.1o, are assigned to crystal
412
plane (110), (200), (220), (310), (211), (301), (411), (600), (521), (022), (541), and (312)
413
reflection, respectively [43, 44]. The intense and sharp peaks were observed in α-MnO2-NRs,
414
where those exhibited a highly crystalline form of tetragonal structure (JCPDS card PDF file no.
415
44-0141) and no other any impurity peaks detected. According to the Braggs equation, the lattice
416
spacing of d = 0.2855 nm value was calculated at the angle 2θ = 37.4. The pure form of α-MnO2
417
NRs was used to calculate the average crystal size of 14.6 nm along the growth of (211) direction
418
(inset), enlarged pattern in Fig. 3b, according to Scherrer’s equations.
419
Fourier transforms infrared (FTIR) spectroscopy of α-MnO2-NRs was shown in Fig. 3c.
420
The broad peaks appeared at 3480.6 cm-1, which corresponds to stretching vibration of absorbed
421
(H-O-H) water molecules. The peak position of 1619.3 cm-1 is assigned to the bending mode of
422
hydroxyl groups of Mn-OH. The band located at about 522 cm-1 is the characteristic vibration of
423
metal-oxygen (Mn-O) bond in [MnO6] octahedra. The peaks appear at around 711.9 and 1382.7
424
cm-1, which are given to the stretching vibration of an O-Mn-O bond between α-MnO2-NRs [45,
425
46].
426
As presented in Fig. 3d, the shape of the plots shows α-MnO2-NRs have a typical type
427
II/IV isotherm with a discrete hysteresis loop obtained in the pressure range (P/PO) = 0.5-1.0 of
428
the region at 77 K. BET surface areas reveal the α-MnO2 NRs have gravimetric surface areas of
429
44.2 m2/g [47]. The BJH pores size distribution curve (PSD) is depicted (inset) in Fig. 3d. The
430
PSD curves to reach maximum range is 45.6 nm within the average pore size of 3.7 nm and the
431
corresponding pore volume is 0.13 and 0.05 cm3/g. The available surface area of mesopore and
19
432
micropore exists on α-MnO2 NRs and it provides electrolyte ion easily to reach the electrode
433
surfaces [48].
434
3.3. Electrochemical characterization of anode materials
435
The electrochemical properties of Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and Bi2O3
436
anode materials were evaluated by the cyclic voltammetry (CV) test in three-electrode
437
configuration at room temperature as discussed in Fig. 4a-d. Furthermore, the typical CV profile
438
concerning all of the corresponding anode materials is involved in the same potential window of
439
-1.2 to 0.2 V at various scan rate (100 to 5 mVs-1) of the electrolyte solution containing 6 M
440
KOH. The electrochemical performances mainly depended on the concentration and hydrous
441
ionic radius of the electrolyte. Therefore, smaller ionic radius K+ (3.31Ao) is more suitable to
442
enhance ionic mobility on the electrode surface. Based on the CV data, the quasi-reversible
443
faradic redox peaks were observed from Bismuth-metal ion nanocomposites anode materials in
444
an alkaline electrolyte solution. The excellent pseudocapacitive properties were ensured as
445
presented in those nanocomposites (as shown in Fig. 4a-b, d). Without Bismuth-metal ion on
446
RGO nanosheet, it displays the electrical double layer peak, instead of redox peak (as shown in
447
Fig. 4c). More details of Fig. 4a, the low scan rate at 5 mVs-1 shows the negative direction of
448
reduction peaks, which could be seen at the potential range of -0.7 to -1.2 V, representing a
449
reduction of Bismuth-metal ion (+3 oxidation) to the metallic bismuth (0 oxidation). In addition,
450
the positive direction of oxidation peak appears at the potential range of +0.07 to -0.1085 V,
451
which indicates the oxidation of bismuth metallic state (0 oxidation) to the bismuth-metal ion
452
(+3 oxidation). Moreover, the unique Bi2O3/porous-RGO nanocomposites displayed a larger
453
redox peak current density with a higher area of CV curve, which may be because the electrode
454
materials possess the rapid electronic and ionic transport as well to enhance the kinetics of a
20
455
faradic reversible redox reaction. Besides, the more negative reduction current density was
456
observed due to the hydrogen intercalation in the oxide and formation of BiO2– (2 Bi2O3 + 3 OH-
457
+ e- → 4BiO2- + H2O + Had). From the above consequence, the unique Bi2O3/porous-RGO
458
nanocomposites is considered as a better anode material for electrochemical energy storage
459
devices. To the best of our knowledge, the possible conversion of redox mechanism of Bi-based
460
electrode materials had been proposed and described in the following equation given as below:
461
PC SQ /porous − RGO + 3`C S + 6 ↔ 2P + 6S` [49-53]
462
The galvanostatic charge and discharge (GCD) behaviors were carried out under the designated
463
voltage window of -1.2 to 0.2 V for the electrode materials, such as Bi2O3/porous-RGO, bulk-
464
Bi2O3/RGO, RGO, and Bi2O3 in the aqueous 6 M KOH electrolyte solution at a different current
465
density of 1~20 Ag-1, respectively. The GCD profile of the Bismuth-metal ion containing
466
nanocomposites is totally differing from that without Bismuth-metal ion of RGO nanosheets
467
GCD curve, which demonstrated that enough valance state has been transferred between
468
electrode surfaces under the same potential as shown Fig. S14 (a-b), S14 (d). This credit goes to
469
the fast kinetic faradaic redox reaction involved in the pseudocapacitive behavior of Bi2O3
470
nanocomposites. Meanwhile, the RGO shows that the GCD profiles appeared regular triangular
471
shape of the asymmetric GCD curve, which is designated that typical triangular shape of the
472
peak exhibited by electrical double layer mechanisms occurred on the RGO electrode surface as
473
shown in Fig. S14 (c). As can be seen from these experimental results, the lengthy discharging
474
tail-time of the Bismuth-metal ion present anode nanomaterials could be divided into the two-
475
discharge region: (i) horizontal discharge and (ii) vertical discharge region. A horizontal
476
discharge named gradual voltage drops, which clearly shows that discharging time is
477
substantially greater than charging time due to the faradaic reaction occurring to continue over
21
478
the long period of discharging time for the presence of discharging plateau. This can be
479
attributed to the improved pseudocapacitance with EDLC properties containing pore volume of
480
mesoporous RGO nanosheets. Besides, the vertical discharge called for steep voltage drop upon
481
the first few seconds of discharge, which has represented for IR drop (non-optimized electrolyte
482
interface) as presented in the corresponding non-symmetrical peaks [54]. For instances, the same
483
type of the GCD curves has been already reported as a different kind of bismuth present in
484
nanocomposites electrodes, such as BiVO4/SWCNT [55], Bi2O3 [56], and BiVO4: Ag [57].
485
However, the extended discharging tail of the Bi2O3/porous-RGO showed the larger specific
486
capacitance at low current density.
487
According to the equation (1), the specific capacitances of Bi2O3/porous-RGO composites
488
are calculated as 226, 158, 120, 49, and 32 Fg-1 at a scan rate of 5, 10, 20, 50, and 100 mVs-1,
489
respectively. The higher value of capacitance was produced at lower scan rate due to the fast
490
diffusion of electrolyte ion, which had enough time to interact with mesoporous nanocomposite
491
electrode surface, while at high scan rate the ion would not have enough time to approach the
492
electrode surface as depicted in Fig. 5a. According to the equation (2), the specific capacitances
493
of Bi2O3/porous-RGO electrodes materials can reach up to 285.1 Fg-1 at the current density of 1
494
Ag-1. Meanwhile, the composites can still deliver 243, 238, 234, and 123 Fg-1 at the different
495
current density of 5, 10, 15, and 20 Ag-1, respectively. At low current density, it exhibits the
496
higher specific capacitance, which could attribute to the mesoporous morphology with a high
497
electrical conductivity of graphene network. Therefore, the electrolyte ion can rapidly diffuse
498
into the accessible surface area of electroactive materials. For comparative studies, the calculated
499
specific capacitance value of Bi2O3/porous-RGO is much better than that of bulk-Bi2O3/RGO,
500
RGO, and Bi2O3 electrodes at the different current density. Additionally, the rate capability of all
22
501
the prepared electrode materials were compared, such as 285.1, 217.4, 145, and 116 Fg-1 at 1 Ag-
502
1
503
Bi2O as showed in Fig. 5b. At a current density of 20 Ag-1, the rate performance exists in lower
504
status (specific capacitance reduces) due to the greater IR drop and unsatisfactory redox process
505
in the active material. Furthermore, the diffusion-controlled of redox reactions occur in the bulk
506
of pseudocapacitive materials (bulk-Bi2O3/RGO), resulting in low coulombic efficiency and
507
sluggish of reaction kinetics. Moreover, the bulk-Bi2O3/RGO, RGO, and Bi2O3 displayed the low
508
rate performance at the higher current density of 20 Ag-1, which can be attributed to the larger
509
resistance of the electrode material, limited ion-diffusion rate in the electrolyte solution, low
510
electrical conductivity, and large interfacial resistance between electroactive materials and the
511
conducting substrate. The cyclic durability of supercapacitor electrode material was evaluated by
512
the GCD curve under the potential range of -1.2 to 0.2 V at a current density of 0.5 Ag-1, the
513
Bi2O3/porous-RGO anode revealed a higher capacitance, excellent rate capability, and
514
outstanding capacitance retention. Furthermore, the calculated specific capacitances of first
515
cycles are 522.5, 396.2, 287.9, and 216 Fg-1 and final cycles are 425, 274, 193, and 113 Fg-1 for
516
Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and Bi2O3 under the potential range of -1.2 to 0.2 V
517
at a current density of 0.5 Ag-1 as displayed in Fig. 5c. On the other hand, the order of
518
capacitance retention of Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and Bi2O3 anode materials
519
(81.1% > 73.2% > 68.5% > 55.9%) after 3000 cycles presented in Fig. 5d. Additionally, the
520
cyclic durability of Bi2O3/porous-RGO also confirmed to the higher number of cycles, which
521
exhibits (Fig. S15) 77% capacitance retention after 10000 cycles. According to the capacitance
522
retention, the Bi2O3/porous-RGO has exhibited the most outstanding cyclic performance. This
and 123, 98.5, 81, and 39 Fg-1 at 20 Ag-1 for Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and
23
523
result mainly depends on the strong electrical conductivity of porous-RGO nanosheets, which
524
enhances the electrolyte ion penetration on the surface of a working electrode.
525
The electrochemical impedance spectroscopy (EIS) of all the prepared electrode materials
526
were further studied for the solution resistance (Rs), charge to transfer resistance (Rct) (or)
527
polarization resistance, and Warburg impedance (Zw). The impedance data of the Nyquist plots
528
were obtained between the real (Z’) and an imaginary axis (-Z’’) of the impedance spectrum at
529
the electrolyte solution containing 6 M KOH with the frequency range of 10 KHz to 0.1 Hz. Fig.
530
S16 (a) shows the typical Nyquist plots of Bi2O3/porous-RGO, bulk-Bi2O3/RGO, RGO, and
531
Bi2O3 electroactive materials, where the EIS curve could be divided into three frequency range.
532
At higher frequency range, the Bi2O3/porous-RGO shows the intersection point of the real axis
533
solution resistance (Rs) value at 0.54 Ω, which indicates that the superior conductivity is higher
534
than that of the bulk-Bi2O3/RGO, RGO, and Bi2O3. However, the medium frequency ranges of
535
Bi2O3/porous-RGO displays a semicircle arc of the diameter, which is able to measure the value
536
of Rct at 0.10 Ω and it demonstrates the faster electrochemical reaction in between the
537
electrode/electrolyte surface due to exhibiting the higher current density of with
538
pseudocapacitance [58]. Owing to the present advantage of the porous-RGO network, those can
539
offer direct current pathways for good electrical transport, thus effectively reducing the
540
resistance of the electrode materials. Nevertheless, the low-frequency ranges of Bi2O3 /porous-
541
RGO electrode exhibited Warburg impedance (Zw) of a diagonal line of vertical slope spike,
542
which precisely started from 45o as depicted in Fig. S16 (b). It indicates that the ion
543
diffusion/transportation processes took place from an electrolyte solution to electrode surfaces
544
with the porous-RGO network to ensure the resultant capacity of the rate capability behavior
545
predominantly. Meanwhile, the ion/electron penetration process of bulk-Bi2O3/RGO is slower
24
546
than that of the Bi2O3/porous-RGO, which may occur due to the poor electrostatic interaction
547
between the metal to carbon bond and the inaccessible area of RGO nanosheets. Above this
548
electrochemical characterization, it can also confirm the lower value of solution resistance (Rs),
549
charge to transfer resistance (Rct), and higher ion diffusion process (Zw) of the Bi2O3/porous-
550
RGO anode material can possess an excellent supercapactive performance.
551
3.4. Electrochemical characterization of cathode material
552
The α-MnO2-NRs cathode material was investigated in 6 M KOH electrolyte in the three-
553
electrode configuration by cyclic voltammograms (CVs) recorded at a different scan rate of 100-
554
1 mVs-1 within the potential window of 0-0.9 V (vs. Hg/HgO) as depicted in Fig. 6a. As can be
555
seen from the data, α-MnO2-NRs exhibit a pseudo-capacitive behavior via redox reaction,
556
suggesting a good reversibility and greatest electron transfer reaction take place on electrode
557
materials at all potential scan rates. Furthermore, the charge storage mechanisms of α-MnO2-
558
NRs can be understood from the intercalation/deintercalation ions. The K+ ions enter into the
559
electrode surface (α-MnO2-NRs) based on the intercalation/deintercalation phenomenon, which
560
occurs through the reversible reaction between Mn-ionic states during redox phenomenon as the
561
following reaction (MnO2 + K+ + e- ↔ MnO.OK). On the other hand, the electrode surface
562
involves the redox reaction in terms of adsorption/absorption of electrolyte cations (K+) as the
563
following reaction (MnO2 surface + K+ + e- ↔ MnO2- K+ surface) [59, 60]. The α-MnO2-NRs show
564
the specific capacitance of 142, 177, 195, 223, 249, and 369 Fg-1 at a scan rate 100, 50, 20, 10, 5,
565
and 1 mVs-1. At very low scan rate (1 mVs-1), α-MnO2-NRs exhibited the highest specific
566
capacitance due to a good reversibility through anodic peak current density (-ipa) divided by
567
cathodic peak current density (+ipc) equal to 1, (i.e.)
568
rate (100 mVs-1) exhibited for the lowest specific capacitance due to its quasi-reversibility or
–cd .e
Rcf g.h
25
= 1. Nevertheless, the faster scan
cd ij
≠ 1 n61, 62p. Furthermore, the capacitance retention for the α-MnO2-
569
irreversibility
570
NRs was found to be about 80.7%. In addition, initial (1st) and final (3000 th) cycles exhibited the
571
223 and 180 Fg-1, respectively, at a scan rate of 10 mVs-1 as presented in Fig. 6b. The GCD
572
measurements were used to calculate the specific capacitance of α-MnO2 NRs electrode within
573
the potential window of 0 - 0.9 V and depicted in Fig. 6c. A comparison of GCD profiles was
574
made at various current densities from 1 to 10 Ag-1, where the specific capacitance was found
575
755 Fg-1 at current density of 1 Ag-1. However, while the current density increased from 1 to 10
576
Ag-1, the specific capacitance values decreased from 755 to 366 Fg-1 due to quasi-reversible
577
redox reaction increasing from lower to higher current density, indicating a decreased electron
578
transfer rate from 1 to 10 Ag-1. Besides, the resistance behavior of α-MnO2-NRs can be
579
determined by EIS spectra and displayed in Fig. 6d. The charge transfer resistance (Rct = 4.5 Ω)
580
and solution resistance (Rs = 0.4 Ω) were found at higher frequency region, designating rapid
581
charge transfer kinetics and higher electrical conductivity between the electrode material and
582
electrolyte [63]. The Warburg resistance (WR = 5.6 Ω) was originated at lower frequency region,
583
which indicates the kinetics for fast diffusion of electrolyte ion into electrode materials. The Rs,
584
Rct, and WR values were obtained based on the equivalent circuit fit inset in Fig. 6d.
585
3.5. Asymmetric supercapacitor assembled in two-electrode configuration
Rcf kl.Q
586
The solid-state ASC device was assembled by α-MnO2 NRs as a positive electrode (8 mg
587
cm-2) and Bi2O3/porous-RGO NS as a negative electrode (8 mg cm-2) using gel electrolyte of
588
PVA/KOH. As depicted in Fig. 7a, the corresponding cathode (+) and anode (-) materials
589
displayed within different potential for (0 to 0.9 V) and (0.2 to -1.2 V) ranges of three-electrode
590
configuration at a constant scan rate of 50 mVs-1. The results confirm the stable cell voltage up to
591
1.8 V using such an assembled ASC device. The CV data onto the solid-state ASC device was 26
592
measured at various scan rate (100 - 5 mVs-1) under 1.8 V, suggesting operating voltage 1.8 V
593
can readily achieve without decomposition of gel-electrolyte as revealed in Fig. 7b. The ASC
594
device achieved the specific capacitance values equation (1) of 84, 75, 67, 61, and 50 Fg-1,
595
respectively, at a scan rate of 5 to 100 mVs-1. The superior performance of ASC device was
596
further confirmed from GCD curve at a various loading current density (10-1 Ag-1) as shown in
597
Fig. 7c. The ASC device achieved the specific capacitance values of 6, 7, 22, 42, 47, and 53 Fg-1,
598
respectively, at a current density of 10-1 Ag-1. The highest specific capacitance was obtained at 1
599
Ag-1 due to a good reversibility (redox reaction) at 1 Ag-1, which agrees well with CV curves.
600
The GCD measurements of ASC device were further performed within different potential
601
windows within 1.0-1.8 V at a current density of 1 Ag-1 and shown in Fig. 7d, indicating the
602
stable electrochemical behavior of assembled ASC within potential voltage from 0-1.8 V.
603
After 3000 cycles, the ASC devices still exhibited an excellent capacitance retention of
604
91.4 % and loss of capacitance 8.6 % at a constant current density of 1 Ag-1. According to
605
equation (3), the first and final cycles achieved to the areal capacitance of 784 and 717 mF cm-2
606
and depicted in Fig. 8a. The capacitance of ASC devices was calculated as a function of current
607
density as presented in Fig. 8b. A high areal capacitance of 787 mF cm-2 (54 Fg-1, based on the
608
mass of electrode material coated on nickel foam surface at an approximate loading of 8 mg cm-
609
2
610
Ragone plot of energy and power density of ASC devices was calculated according to equation
611
(4) and (5). Furthermore, the ASC has exhibited a volumetric capacitance of 74 Fcm-3 (54 Fg-1)
612
at a current density of 1 Ag-1 as seen in Fig. S17. The solid-state ASC displays the maximum
613
energy density of 86 Wh kg-1 at a power density of 9000 W kg-1 and this remains delivered 10
614
Wh kg-1 at a power density of 11000 W kg-1 calculated at a current density of 1 and 10 Ag-1,
) can be achieved at a current density of 1 Ag-1 and shown in Fig. 8b. As depicted in Fig. 8c, the
27
615
respectively. These values are compared and higher than that of recent reports of Bi-based ASC
616
in reported literature as shown in Table S4. The inset in Fig. 8c has shown the red light emitting
617
diode illuminated by the fabricated Bi2O3/porous-RGO Ns)//α-MnO2-NRs devices, indicating
618
promising electrode materials in the practical application of energy storage.
619 620
4. Conclusions
621
In summary, the unique structure of the Bi2O3/porous-RGO (catalytic carbon gasification)
622
and α-MnO2-NRs (hydrothermal) materials was successfully prepared. In the three-electrode
623
system, the unique morphology of Bi2O3/porous-RGO Ns and α-MnO2-NRs achieved the
624
maximum specific capacitance of 285 and 755 Fg-1 at a current density of 1 Ag-1. Besides,
625
significant rate capability and excellent cyclic durability were observed at a potential window of
626
the anode (-1.2 to 0.2 V) and cathode (0 to 0.9 V) in 6 M KOH electrolyte. This excellent
627
electrochemical performance can be attributed to the synergetic effect of the highly conducting
628
porous-RGO nanosheets possessing mesoporous morphology, nanosized of α-MnO2-NRs, and
629
ultra-small Bi2O3 nanoparticle. In the two-electrode system, the ASC delivered an outstanding
630
areal capacitance of 787 mF cm-2 and excellent capacitance retention of 91.4 % obtained after
631
3000 cycles within the working voltage reaching 1.8 V at a constant current density of 1 Ag-1.
632
The ASC achieved a remarkable performance of the maximum energy density of 86 Wh kg-1 at a
633
power density of 9000 W kg-1 and maintained 10 Wh kg-1 at a power density of 11000 W kg-1 in
634
PVA/KOH gels electrolyte. This work has demonstrated the excellent electrochemical
635
performance of Bi2O3/porous-RGO and α-MnO2-NRs composites, which has laid the foundation
636
for a new class of electrode materials for high-performance aqueous asymmetric supercapacitor.
637
28
638 639
ACKOWLEDGEMENT
640
The authors wish to thank for the financial support by the Ministry of Science and
641
Technology (MOST) in Taiwan under the contract number of 105-2221-E-035-002-MY2. In
642
addition, the support of Taiwan-India joint project (NSC-102-2923-E-035-001-MY3 and DST-
643
GITA/DST/TWN/P-50/2013) is also acknowledged. The support in providing the fabrication and
644
measurement facilities from the Precision Instrument Support Center of Feng Chia University is
645
also acknowledged.
646 647 648
REFERENCES (1)
S.P.S. Badwal, S.S. Giddey, C. Munnings, A.I. Bhatt, A.F. Hollenkamp, Emerging electrochemical energy conversion and storage technologies, Front. Chem. 2 (2014) 79.
(2)
S. Ghosh, O. Inganäs, Conducting polymer hydrogels as 3D electrodes: applications for supercapacitors, Adv. Mater. 11 (1999) 1214–1218.
(3)
M. Winter, R.J. Brodd, What Are Batteries, Fuel Cells, and Supercapacitors?, Chem. Rev. 104 (2004) 4245–4270.
(4)
L. Su, L. Gao, Q. Du, L. Hou, X. Yin, M. Feng, W. Yang, Z. Ma, G. Shao, Formation of micron-sized nickel cobalt sulfide solid spheres with high tap density for enhancing pseudocapacitive properties, ACS Sustain. Chem. Eng. 5 (2017) 9945−9954.
(5)
I. Oh, M. Kim, J. Kim, Controlling hydrazine reduction to deposit iron oxides on oxidized activated carbon for supercapacitor application, Energy. 86 (2015) 292– 299.
29
(6)
G. Wang, L. Zhang, J. Zhang, A review of electrode materials for electrochemical supercapacitors, Chem. Soc. Rev. 41 (2012) 797–828.
(7)
C.H. Kim, B.-H. Kim, Electrochemical behavior of zinc oxide-based porous carbon composite nanofibers as an electrode for electrochemical capacitors, J. Electroanal. Chem. 730 (2014) 1–9.
(8)
Y.Q. Fang, W.H. Wei, W. Chao, J.C. De, S.Q. Feng, One Step construction of nitrogen-carbon derived from bradyrhizobium japonicum for supercapacitor applications with a soybean leaf as a separator, ACS Sustain. Chem. Eng. 6 (2018) 4695−4704.
(9)
Y. Luan, Y. Huang, L. Wang, M. Li, R. Wang, B. Jiang, Porous carbon@MnO2 and nitrogen-doped porous carbon from carbonized loofah sponge for asymmetric supercapacitor with high energy and power density, J. Electroanal. Chem. 763 (2016) 90–96.
(10) C. Guan, J. Liu, Y. Wang, L. Mao, Z. Fan, Z. Shen, H. Zhang, J. Wang, Iron oxidedecorated carbon for supercapacitor anodes with ultrahigh energy density and outstanding cycling stability, ACS Nano. 9 (2015) 5198–5207. (11) H. Fan, R. Niu, J. Duan, W. Liu, W. Shen, Fe3O4@Carbon nanosheets for all-solidstate supercapacitor electrodes, ACS Appl. Mater. Interfaces. 8 (2016) 19475– 19483. (12) Z.-S. Wu, G. Zhou, L.-C. Yin, W. Ren, F. Li, H.-M. Cheng, Graphene/metal oxide composite electrode materials for energy storage, Nano Energy. 1 (2012) 107–131. (13) X. Yao, Y. Zhao, Three-dimensional porous graphene networks and hybrids for lithium-ion batteries and supercapacitors, Chem. 2 (2017) 171–200.
30
(14) H. Wang, X. Sun, Z. Liu, Z. Lei, Creation of nanopores on graphene planes with MgO
template
for
preparing
high-performance
supercapacitor
electrodes,
Nanoscale. 6 (2014) 6577–6584. (15) E. Frackowiak, Carbon materials for supercapacitor application, Phys. Chem. Chem. Phys. 9 (2007) 1774–1785. (16) C.-C. Hu, K.-H. Chang, M.-C. Lin, Y.-T. Wu, Design and tailoring of the nanotubular
arrayed
architecture
of
hydrous
RuO2
for
next
generation
supercapacitors, Nano Lett. 6 (2006) 2690–2695. (17) D. Yuan, J. Zeng, N. Kristian, Y. Wang, X. Wang, Bi2O3 deposited on highly ordered mesoporous carbon for supercapacitors, Electrochem. Commun. 11 (2009) 313–317. (18) W. Fang, N. Zhang, L. Fan, K. Sun, Preparation of polypyrrole-coated Bi2O3@CMK-3 nanocomposite for electrochemical lithium storage, Electrochim. Acta. 238 (2017) 202–209. (19) C. Pei, H. Sun, Z. Zhu, W. Liang, J. An, Q. Zhang, A. Li, Synthesis of novel porous graphene nanocomposite and its use as electrode and absorbent, RSC Adv. 4 (2014) 14042–14047. (20) C. Liu, Z. Yu, D. Neff, A. Zhamu, B.Z. Jang, Graphene-based supercapacitor with an ultrahigh energy density, Nano Lett. 10 (2010) 4863–4868. (21) Q. Ke, J. Wang, Graphene-based materials for supercapacitor electrodes – A review, J. Mater. 2 (2016) 37–54. (22) J.H. Lee, N. Park, B.G. Kim, D.S. Jung, K. Im, J. Hur, J.W. Choi, Restackinginhibited 3D reduced graphene oxide for high performance supercapacitor
31
electrodes, ACS Nano. 7 (2013) 9366–9374. (23) Y. Xu, K. Sheng, C. Li, G. Shi, Self-assembled graphene hydrogel via a one-step hydrothermal process, ACS Nano. 4 (2010) 4324–4330. (24) B.G. Choi, M. Yang, W.H. Hong, J.W. Choi, Y.S. Huh, 3D macroporous graphene frameworks for supercapacitors with high energy and power densities, ACS Nano. 6 (2012) 4020–4028. (25) L.L. Zhang, X. Zhao, M.D. Stoller, Y. Zhu, H. Ji, S. Murali, Y. Wu, S. Perales, B. Clevenger, R.S. Ruoff, Highly conductive and porous activated reduced graphene oxide films for high-power supercapacitors, Nano Lett. 12 (2012) 1806–1812. (26) L. Sun, C. Tian, M. Li, X. Meng, L. Wang, R. Wang, J. Yin, H. Fu, From coconut shell to porous graphene-like nanosheets for high-power supercapacitors, J. Mater. Chem. A. 1 (2013) 6462–6470. (27) Y. Gong, Z. Wei, J. Wang, P. Zhang, H. Li, Y. Wang, Design and fabrication of hierarchically porous carbon with a template-free method, Sci. Rep. 4 (2014) 6349. (28) L. Wan, S. Sun, T. Zhai, S.V. Savilov, V.V. Lunin, H. Xia, Multiscale porous graphene oxide network with high packing density for asymmetric supercapacitors, J. Mater. Res. (2017) 1–12. (29) J. Yang, J. Hu, M. Zhu, Y. Zhao, H. Chen, F. Pan, Ultrahigh surface area meso/microporous carbon formed with self-template for high-voltage aqueous supercapacitors, J. Power Sources. 365 (2017), 362-371. (30) A. Jana, E. Scheer, S. Polarz, Synthesis of graphene–transition metal oxide hybrid nanoparticles and their application in various fields, Beilstein J. Nanotechnol. 8 (2017) 688–714.
32
(31) W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80 (1958) 1339. (32) A. Lerf, H. He, M. Forster, J. Klinowski, Structure of graphite oxide revisited, J. Phys. Chem. B. 102 (1998) 4477–4482. (33) A. Laheäär, P. Przygocki, Q. Abbas, F. Béguin, Appropriate methods for evaluating the efficiency and capacitive behavior of different types of supercapacitors, Electrochemistry Communications 60 (2015) 21–25. (34) P. Xiong, J. Zhu, L. Zhang, X. Wang, Recent advances in graphene-based hybrid nanostructures for electrochemical energy storage, Nanoscale Horiz. 1 (2016) 340– 374. (35) K.-Y. Niu, J. Park, H. Zheng, A.P. Alivisatos, Revealing bismuth oxide hollow nanoparticle formation by the kirkendall effect, Nano Lett. 13 (2013) 5715–5719. (36) P. Madhusudan, J. Yu, W. Wang, B. Cheng, G. Liu, Facile synthesis of novel hierarchical
graphene-Bi2O2CO3
composites
with
enhanced
photocatalytic
performance under visible light, Dalt. Trans. 41 (2012) 14345–14353. (37) Z. Ma, X. Huang, S. Dou, J. Wu, S. Wang, One-pot synthesis of Fe2O3 nanoparticles on nitrogen-doped graphene as advanced supercapacitor electrode materials, J. Phys. Chem. C. 118 (2014) 17231–17239. (38) L. Liu, J. Jiang, S. Jin, Z. Xia, M. Tang, Hydrothermal synthesis of [β]-bismuth oxide nanowires from particles, CrystEngComm. 13 (2011) 2529–2532. (39) W. Fang, N. Zhang, L. Fan, K. Sun, Bi2O3 nanoparticles encapsulated by threedimensional porous nitrogen-doped graphene for high-rate lithium ion batteries, J. Power Sources. 333 (2016) 30–36.
33
(40) S.R. Ede, S. Kundu, Microwave synthesis of SnWO4 nanoassemblies on DNA scaffold: A novel material for high performance supercapacitor and as catalyst for butanol oxidation, ACS Sustain. Chem. Eng. 3 (2015) 2321–2336. (41) G. Zhu, C. Xi, M. Shen, C. Bao, J. Zhu, Nanosheet-based hierarchical Ni2(CO3)(OH)2 microspheres with weak crystallinity for high-performance supercapacitor, ACS Appl. Mater. Interfaces. 6 (2014) 17208–17214. (42) L. Zhang, Y. Hashimoto, T. Taishi, I. Nakamura, Q.-Q. Ni, Fabrication of flowershaped Bi2O3 superstructure by a facile template-free process, Appl. Surf. Sci. 257 (2011) 6577–6582. (43) K.Ahmad, A.Mohammad, S.M.Mobin, Hydrothermally grown α-MnO2 nanorods as highly efficient low cost counter-electrode material for dye-sensitized solar cells and electrochemical sensing applications, Electrochim. Acta. 252 (2017) 549–557. (44) C.Liu,
S.T.Navale,
Z.B.Yang,
M.Galluzzi,
V.B.Patil,
P.J.Cao,
R.S.Mane,
F.J.Stadler, Ethanol gas sensing properties of hydrothermally grown α-MnO2 nanorods, J. Alloys Compd. 727 (2017) 362–369. (45) H.W. and Z.L. and D.Q. and Y.L. and W.Zhang, Single-crystal α-MnO2 nanorods: synthesis and electrochemical properties, Nanotechnology. 18 (2007) 115616. (46) N.Tang, X.Tian, C.Yang, Z.Pi, Q.Han, Facile synthesis of α-MnO2 nanorods for high-performance alkaline batteries, J. Phys. Chem. Solids. 71 (2010) 258–262. (47) P.Umek, R.C.Korošec, A.Gloter, U.Pirnat, The control of the diameter and length of α-MnO2 nanorods by regulation of reaction parameters and their thermogravimetric properties, Mater. Res. Bull. 46 (2011) 278–284. (48) D.P.Dubal, J.G.Kim, Y.Kim, R.Holze, W.B.Kim, Demonstrating the Highest
34
Supercapacitive Performance of Branched MnO2 Nanorods Grown Directly on Flexible Substrates using Controlled Chemistry at Ambient Temperature, Energy Technol. 1 (2013) 125–130. (49) R. Liu, L. Ma, G. Niu, X. Li, E. Li, Y. Bai, G. Yuan, Oxygen-deficient bismuth oxide/graphene of ultrahigh capacitance as advanced flexible anode for asymmetric supercapacitors, Adv. Funct. Mater. 27 (2017) Article 1701635. (50) L. Li, X. Zhang, Z. Zhang, M. Zhang, L. Cong, Y. Pan, S. Lin, A bismuth oxide nanosheet-coated electrospun carbon nanofiber film: a free-standing negative electrode for flexible asymmetric supercapacitors, J. Mater. Chem. A. 4 (2016) 16635–16644. (51) K.C. Chitrada, K.S. Raja, Nanoporous anodic bismuth oxide for electrochemical energy storage, ECS Transactions. 61 (2014) 55-67. (52) W. Zuo, W. Zhu, D. Zhao, Y. Sun, Y. Li, J. Liu, X.W. (David) Lou, Bismuth oxide: a versatile high-capacity electrode material for rechargeable aqueous metal-ion batteries, Energy Environ. Sci. 9 (2016) 2881–2891. (53) B. Wang, J. Wang, Y. Zhang, Y. Mei, P. Lian, Electrochemical performance of Bi2O2CO3 nanosheets as negative electrode material for supercapacitors, Ceram. Int. 43 (2017) 9310–9316. (54) Y. Arora, C.Seth, D. Khushalani, Crafting Inorganic Materials for Use in Energy Capture and Storage. Langmuir. 2018, xxxx. (55) Z. Khan, S. Bhattu, S. Haram, D. Khushalani, SWCNT/BiVO4 composites as anode materials for supercapacitor application, RSC Adv. 4 (2014) 17378–17381. (56) H. Su, S. Cao, N. Xia, X. Huang, J. Yan, Q. Liang, D. Yuan, Controllable growth of
35
Bi2O3 with rod-like structures via the surfactants and its electrochemical properties, J. Appl. Electrochem. 44 (2014) 735–740. (57) Z.J. Zhang, Q.C. Zheng, L. Sun, Synthesis of 2-D nanostructured BiVO4:Ag hybrid as an efficient electrode material for supercapacitors, Ceram. Int. 43 (2017) 16217– 16224. (58) R. Ramya, R. Sivasubramanian, M.V. Sangaranarayanan, Conducting polymersbased electrochemical supercapacitors - Progress and prospects, Electrochim. Acta. 101 (2013) 109–129. (59) A.
Phakkhawan,
P.
Klangtakai1,
A.
Chompoosor,
S.
Pimanpang,
V.
Amornkitbamrung, A comparative study of MnO2 and composite MnO2-Ag nanostructures
prepared
by a hydrothermal
technique
on supercapacitor
applications, Journal of Materials Science: Materials in Electronics. 29 (2018) 9406–9417. (60) S. Maiti, A. Pramanik, S. Mahanty, Interconnected Network of MnO2 Nanowires with
a
“Cocoonlike”
Morphology:
Redox
Couple-Mediated
Performance
Enhancement in Symmetric Aqueous Supercapacitor, ACS Appl. Mater. Interfaces. 6 (2014) 10754–10762. (61) L.M. Johny, S. R, N.S.N. Jothi, P. Sagayaraj, Single crystalline SnO2 nanorods functionalized with TiO2 nanospheres and their electrochemical properties, Ceram. Int. 44 (2018) 14471–14479. (62) G.A. Mabbott, An introduction to cyclic voltammetry, J. Chem. Educ. 60 (1983) 697. (63) A. Kumar, A. Sanger, A. Kumar, Y. Kumar, R. Chandra, An efficient α-MnO2
36
nanorods forests electrode for electrochemical capacitors with neutral aqueous electrolytes, Electrochim. Acta. 220 (2016) 712–720. 649
37
650
Figure Caption
651 652
Scheme 1. Schematic illustration of the synthesis routes towards the Bi2O3/porous-RGO
653
supercapacitor electrode
654
Fig. 1. Morphological characterization of supercapactive electrode materials for (a) bulk-
655
Bi2O3/RGO (inset part shows the SAED pattern), (b) the average size of the Bi2O3 nanoparticles
656
calculated from the Bi2O3/porous-RGO (c) The TEM image of obvious porous nanostructures of
657
Bi2O3/porous-RGO electrode at different magnification shows in (c, d). The HR-TEM image of
658
Bi2O3 nanoparticles encircled by the porous-RGO nanosheets is depicted as (e, f). The lattice
659
distance was calculated from the FFT and IFFT spectrum of Bi2O3/Porous-RGO electrode is
660
clearly illustrations of (g-i)
661
Fig. 2. (a) XRD, (b) Pore size distribution curve (PSD), (c) FTIR, and (d) TGA (inset: maximum
662
weight loss rate temperature T (mwlr) calculated from Bi2O3) for the Bi2O3/porous-RGO, bulk-
663
Bi2O3/RGO, RGO, and Bi2O3 electrode materials
664
Fig. 3. (a) FE-SEM images, (b) XRD and enlarged pattern (inset) in (211) planes, (c) FTIR, and
665
(d) BET (inset pore size distribution curve) of α-MnO2- NRs
666
Fig. 4. Typical CV results in (a) Bi2O3/porous-RGO, (b) bulk-Bi2O3/RGO, (c) RGO and (d)
667
Bi2O3 within the potential range of -1.2 to 0.2 V at a scan rate of 5 mVs-1 to 100 mVs-1
668
Fig. 5. (a) and (b) the Specific capacitance versus scan rate and a current density of all electrode
669
materials, (c) and (d) the calculated specific capacitance and capacitance retention for 3000 cycle
670
number of prepared electrode materials
671
38
672
Fig. 6. (a) CV curves at a different scan rate (100-1mVs-1), (b) cyclic stability at 10 mVs-1, (c)
673
GCD curves at a different current density (10-1Ag-1), and (d) EIS spectra (inset) equivalent
674
circuit fit of α-MnO2-NRs
675
Fig. 7. (a) CV curves compared to anode/cathode materials within their respective stable
676
potential window at a scan rate of 50 mVs-1, (b) CV curves of the ASC at a different scan rate
677
(100-1 mVs-1), (c) GCD curves of ASC at a different loading current density (10-1 Ag-1), and (d)
678
GCD curve of Bi2O3/porous-RGO Ns//α-MnO2-NRs at a different applied potential for 1.8 to 1.0
679
V
680
Fig. 8. (a) Areal capacitance and capacitance retention vs. cycle number of (inset) first and final
681
5 cycles of corresponding ASC devices at a 1 Ag-1, (b) Areal and specific capacitance calculated
682
at a different current density (1-10 Ag-1), and (c) The energy density vs. power density of ASC
683
device and (inset) connected with LED bulb illustration lighting
684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 39
699 700
Scheme. 1
701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 40
723 724
Fig. 1
725 726 727 728 729 730 731 732
41
733
Fig. 2
734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755
42
756
Fig. 3
757 758 759
(a)
760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 43
785
Fig. 4
786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805
44
806
Fig. 5
807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828
45
829
Fig. 6
830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851
46
852 853
Fig. 7
854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874
47
875
Fig. 8
876 877 878 879 880
48
Highlights
Bismuth oxide nanoparticles were successfully dispersed on porous reduced graphene oxide nanosheets by catalytic carbon gasification method
The capacitance retention of Bi2O3/porous-RGO and α-MnO2-NRs has been achieved up to 81.1 % at 0.5 Ag-1 and 80.7 % 10 mVs-1
The ASC delivered an energy density of 86 Wh kg-1 (787 mF cm-2) at a power density of 9000 W kg-1