Accepted Manuscript Polyaniline/β-MnO2 nanocomposites as cathode electrocatalyst for oxygen reduction reaction in microbial fuel cells Xinxing Zhou, Yunzhi Xu, Xiaojie Mei, Ningjie Du, Rongmao Jv, Zhaoxia Hu, Shouwen Chen PII:
S0045-6535(18)30066-3
DOI:
10.1016/j.chemosphere.2018.01.058
Reference:
CHEM 20639
To appear in:
ECSN
Received Date: 19 September 2017 Revised Date:
12 January 2018
Accepted Date: 13 January 2018
Please cite this article as: Zhou, X., Xu, Y., Mei, X., Du, N., Jv, R., Hu, Z., Chen, S., Polyaniline/βMnO2 nanocomposites as cathode electrocatalyst for oxygen reduction reaction in microbial fuel cells, Chemosphere (2018), doi: 10.1016/j.chemosphere.2018.01.058. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT Polyaniline/β-MnO2 nanocomposites as cathode electrocatalyst for oxygen
2
reduction reaction in microbial fuel cells
3
Xinxing Zhou, Yunzhi Xu, Xiaojie Mei, Ningjie Du, Rongmao Jv, Zhaoxia Hu*,
4
Shouwen Chen*
5
Jiangsu Key Laboratory of Chemical Pollution Control and Resources Reuse, School
6
of environmental and biological engineering, Nanjing University of Science and
7
Technology, Nanjing 210018, China
SC
RI PT
1
M AN U
8 9 10 11
14 15
EP
13
TE D
12
*Corresponding author information:
17
Tel.: +86-25-84315532;
18
Fax: +86-25-84315518.
19
E-mail address:
[email protected];
[email protected]
AC C
16
20 21 22 1
ACCEPTED MANUSCRIPT 23
Abstract An efficient and inexpensive catalyst for oxygen reduction reaction (ORR),
25
polyaniline (PANI) and β-MnO2 nanocomposites (PANI/β-MnO2), was developed for
26
air-cathode microbial fuel cells (MFCs). The PANI/β-MnO2, β-MnO2, PANI and
27
β-MnO2 mixture modified graphite felt electrodes were fabricated as air-cathodes in
28
double-chambered MFCs and their cell performances were compared. At a dosage of
29
6 mg cm-2, the maximum power densities of MFCs with PANI/β-MnO2, β-MnO2,
30
PANI and β-MnO2 mixture cathodes reached 248, 183 and 204 mW m-2, respectively,
31
while the cathode resistances were 38.4, 45.5 and 42.3 Ω, respectively, according to
32
impedance analysis. Weak interaction existed between the rod-like β-MnO2 and
33
surficial growth granular PANI, this together with the larger specific surface area and
34
PANI electric conducting nature enhanced the electrochemical activity for ORR and
35
improved the power generation. The PANI/β-MnO2 nanocomposites are a promising
36
cathode catalyst for practical application of MFCs.
37
Keywords: Cathode catalyst; PANI/β-MnO2 nanocomposites; Oxygen reduction
38
reaction; Microbial fuel cell
39
1. Introduction
SC
M AN U
TE D
EP
AC C
40
RI PT
24
Microbial fuel cell (MFC) is a green energy technology that converts chemical
41
energy harvested in organic compounds whether pollutant or not into electrical energy
42
utilizing the power of microorganisms. It has the characteristics of wide range of raw
43
materials and good biocompatibility (Cheng et al., 2006; Anu Prathap et al., 2013).
44
Oxygen has superiority as electron acceptor for MFCs due to its high oxidation 2
ACCEPTED MANUSCRIPT potential, the harmless reduction product (water) (Dong et al., 2012a) and
46
sustainability (Dong et al., 2012b). However, the sluggish kinetics of the oxygen
47
reduction reaction (ORR) in a near neutral medium seriously limits the power density
48
production of MFCs (Wen et al., 2012). To solve this problem, platinum (Pt) and
49
Pt-based materials are widely used as catalyst of ORR in MFC cathodes. Nevertheless,
50
noble metal platinum is expensive, unrenewable, and sensitive to catalyst poisoning
51
which attenuates the performance of MFCs and limits its large-scale application.
52
Because of this, it is necessary to pour a great deal of efforts on developing alternative
53
catalysts which are easy to produce, low consumption, and have the similar or better
54
electrocatalytic activity than Pt for ORR.
M AN U
SC
RI PT
45
Manganese dioxide (MnO2) materials with different crystal structure (α-MnO2,
56
β-MnO2, and γ-MnO2) have been investigated extensively as the cathode catalysts for
57
ORR in MFCs due to its rich resources, low cost, environment friendliness, and good
58
catalytic activity. Zhang et al. (Zhang et al., 2009) compared the three crystal types (α,
59
β and γ) of MnO2 as cathode catalyst in MFCs, they found that β-MnO2 showed the
60
best effective ORR catalytic activity owing to its highest BET surface area and
61
average oxidation state of manganese (AOS: 3.59), which was higher than those of
62
α-MnO2 (3.55) and γ-MnO2 (3.48), since AOS was an indicator of the manganese
63
state in the structures of manganese dioxides and the higher AOS stands for the higher
64
oxidative activity (Zhang et al., 2009). Therefore, β-MnO2 is chosen as the object of
65
this study. However, the poor electronic conductivity of MnO2 remains a major
66
challenge and limits the rate capability for MFC power performance. To address this
AC C
EP
TE D
55
3
ACCEPTED MANUSCRIPT drawback, MnO2-based catalysts are usually supported on highly conductive materials
68
such as polypyrrole (PPy) (He et al., 2007; Tran et al., 2007), polythiophene (PTh) (Li
69
et al., 2009) or polyaniline (PANI) (Li et al., 2008a; Li et al., 2008b) materials. These
70
polymers have attracted more and more attention because of their low cost,
71
convenient synthesis, strong energy-storage capacity and high electrical conductivity.
72
Among the various electroconductive polymers studied to date, PANI has been
73
considered as the most attractive polymer due to its excellent conductivity and
74
environmental stability (Wang et al., 2017). More importantly, PANI also has
75
electrocatalytic activity for ORR (Khomenko et al., 2005). In previous reports,
76
polyaniline/inorganic composites have been proved to have better electrical
77
conductivity (Yuan et al., 2008; Lai et al., 2011). The PANI-MnO2 composites have
78
been reported in the application of supercapacitors (Jiang et al., 2012; Chen et al.,
79
2013). However, the reports of these kinds of PANI-MnO2 especially β-MnO2
80
composites function as air-cathode catalyst in application of MFCs are only few.
TE D
M AN U
SC
RI PT
67
In this paper, we successfully loaded PANI on the surface of β-MnO2 nanorods
82
by in situ chemical oxidative polymerization method and investigated the possibility
83
of PANI/β-MnO2 nanocomposites as cathode catalyst for ORR in an MFC. The
84
performances of MFCs were compared with three kinds of materials (β-MnO2,
85
PANI+β-MnO2 mixture and PANI/β-MnO2 nanocomposites) used as cathode catalysts
86
in MFC. Furthermore, the cathode with different loading amount of PANI/β-MnO2
87
nanocomposites as catalyst were also investigated.
88
2. Experimental
AC C
EP
81
4
ACCEPTED MANUSCRIPT 89
2.1 Materials Aniline was purchased from Shanghai Bai Lingwei Chemical Technology Co.,
91
Ltd. Potassium permanganate (KMnO4) was obtained from Sinopharm Chemical
92
Reagent Co., Ltd. Aniline sulfate was purchased from Aladdin Industrial Corporation.
93
The graphite felt with a thickness of 5 mm was bought from Zibo Jinpeng Carbon Co.,
94
Ltd. Cation exchange membrane (CEM) was purchased from Hangzhou El
95
Environmental Protection Technology Co., Ltd. (IEC = 2.4 mmol g-1). All other
96
reagents used in this work were of analytical grade and were used without further
97
purification. Deionized water was produced by an EPED pure water system and used
98
in all solution.
99
2.2 Synthesis of PANI
M AN U
SC
RI PT
90
To a three-necked flask, 100 mL of hydrochloric acid solution (1 M) and aniline
101
(0.05 mol, 4.65 g) were added sequentially. The solution was stirred and cooled to 0-5
102
o
103
solution dropwise, and the transparent solution gradually turned into mazarine. After
104
the continuous stirring for 3 h, the solution was kept still overnight. The precipitate
105
was collected by filtration, washed with deionized water and dried under vacuum
106
oven at 65 oC and ground into powder (Ma et al., 2007).
107
2.3 Synthesis of β-MnO2 nanorods
TE D
100
AC C
EP
C in an ice bath. Then ammonium peroxydisulfate (1 M, 50 mL) was added to the
108
β-MnO2 was prepared by a previously reported method (Zhang et al., 2009). In a
109
typical procedure, 1.2 g of KMnO4, 2.7 mL of ethanol and 51.3 mL of deionized water
110
were added into a 100 mL beaker. After the dissolution of KMnO4, the solution was 5
ACCEPTED MANUSCRIPT transferred into a Teflon-lined stainless steel autoclave (300 mL) and heated at 125 oC
112
for 24 h. The obtained precipitate was collected by filtration, washed with deionized
113
water and ethanol several times, and dried at 100 oC in a vacuum oven. Finally, the
114
precipitates were annealed at 300 oC in air for 5 h at a ramping rate of 3.5 oC min-1.
115
2.4 Synthesis of PANI/β-MnO2 nanocomposites
RI PT
111
In a 500 mL beaker, aniline sulfate (0.143 g, 0.5 mmol) was dissolved in 100 mL
117
of sulfuric acid, and the solution was cooled to 0-5 oC in an ice bath. Under vigorous
118
stirring, 0.279 g of as-prepared β-MnO2 was added and kept for 2 h. The obtained
119
precipitate was collected by filtration, washed with deionized water and ethanol
120
several times, respectively, then dried in a vacuum oven at 60 oC.
121
2.5 Pretreatment of graphite felt and cation exchange membrane
M AN U
SC
116
First, the graphite felt was cut into small piece with the size of 2 × 2 × 0.5 cm
123
and put into a 250 mL beaker. Next, ethanol (150 mL) was added to immerse the
124
graphite felt completely and ultrasonic for 10 min. Then the graphite felt was rinsed
125
by deionized water and the operation was repeated until the deionized water was
126
transparent. Finally, the graphite felt was dried in a vacuum oven at 60 oC for 24 h.
127
The purified graphite felt was acidic oxidized in a sealed 300 mL Teflon-lined
128
stainless steel autoclave containing 120 mL mixed acid solution (VH SO :VHNO3 = 3:1) at
129
80 oC for 8 h (Zhang et al., 2013). The acidic oxidized graphite felt was washed with
130
deionized water, and sonicated for 30 min in the deionized water until the pH of the
131
rinsed water became neutral. Then they were dried in a vacuum oven at 100 oC for 5 h
132
and saved in a sealed bag.
AC C
EP
TE D
122
2
6
4
ACCEPTED MANUSCRIPT CEM was soaked in 1 M HCl solution at least 24 h for proton exchange, rinsed
134
thoroughly with deionized water and then immersed into deionized water at least 24 h
135
before use.
136
2.6 Electrode fabrication
RI PT
133
Graphite felt was used as the supporting carbon/electrode material, and several
138
air-cathodes were prepared as parallel control. The air-electrode was fabricated by
139
ultrasonic-load method. The as-prepared β-MnO2, PANI+β-MnO2 mixture or
140
PANI/β-MnO2, the binder of polyvinylidene fluorine were mixed by a weight ratio of
141
95:5 in 20 mL of N-methyl-2-pyrrolidone in a ultrasonic bath for 15 min. A piece of
142
acid treated graphite felt (2 × 2 × 0.5 cm) was immersed into the solution and placed
143
in the ultrasonic bath for another 30 min. Then the graphite felt was taken out and
144
dried in a vacuum oven at 60 oC for 10 h to remove the solvent. The above step was
145
repeated several times until gained the settled catalyst loading. The prepared
146
air-cathodes were named as GF, GF-PANI+β-MnO2 and GF-PANI/β-MnO2. We also
147
studied the effect of PANI/β-MnO2 loading amount on the performance of the MFC.
148
The prepared air-electrodes were named as GF-PANI/β-MnO2-x, where x refers to the
149
catalyst loading amount (mg cm-2).
150
2.7 MFC configuration and operation
M AN U
TE D
EP
AC C
151
SC
137
The double-chamber air-cathode MFC was constructed with two inner
152
cylindrical Plexiglas chambers (d × H = 8 × 5.5 cm, 280 mL), each chamber consists
153
of two ports for feed inlet/outlet, gas inlet/outlet and electrode connection parts. They
154
were separated by the CEM. The graphite felt-based electrodes were placed parallel to 7
ACCEPTED MANUSCRIPT the CEM facing each other with a distance of 3 cm each. The anode and cathode were
156
connected by titanium wire as current collector with conductive paste. The anodic
157
chamber was filled with medium containing Na2HPO4·12H2O (11.1 g L-1),
158
NaH2PO4·2H2O (2.96 g L-1), NaAc·3H2O (2.72 g L-1), NH4Cl (0.31 g L-1),
159
MgCl2·6H2O (0.21 g L-1), NaCl (1.0 g L-1), CaCl2 (0.02 g L-1), KCl (0.13 g L-1) and
160
trace element (1 mL L-1), whereas the medium was sterilized at 121 oC for 15 min
161
before use. The culture was inoculated from anaerobic digester sludge collected from
162
Nanjing East sewage treatment plant. The catholyte was comprised of a phosphate
163
buffer solution (PBS, pH=7), which consisted of Na2HPO4·12H2O (11.1 g L-1),
164
NaH2PO4·2H2O (2.96 g L-1) and NaCl (1.0 g L-1). Air was continuously fed into the
165
cathode chamber using an air pump at a rate of 90 mL min-1 to maintain a sustained
166
concentration of oxygen during the MFC operation. All the reactors were controlled at
167
30 oC in a water bath incubator. At the start-up period, an external resistor (1000 Ω)
168
was connected in the circuit. The schematic figure of the MFC is shown in Fig. SM-1.
169
2.8 Characterizations and measurements
EP
TE D
M AN U
SC
RI PT
155
X-ray diffraction (XRD) was conducted on D8 Advance (Bruker) using Cu Kα
171
radiation at 40 kV and with 2θ of 10°-70°. Scanning electron microscope (SEM) was
172
performed on a Quant 250FEG (FEI) machine. Transmission electron microscopy
173
(TEM) analysis was conducted on a TECNAI G2 20 LaB6 (FEI) electron microscope
174
operated at 200 kV. Thermal gravimetric analysis (TGA) was carried out on a
175
SDTA851E instrument (TA-instruments-waters LLC) at a heating rate of 10 oC min-1
176
under flowing air from 50 to 800 oC. Fourier transform infrared (FT-IR) spectra was
AC C
170
8
ACCEPTED MANUSCRIPT recorded on Nicolet is10 (Thermo Fisher Scientific) with the KBr pellet method. The
178
specific surface area and pore size distribution of the catalyst were analyzed using the
179
Brunauer–Emmet–Teller (BET, ASAP-2020, Micromeritics, America) method by the
180
adsorption and desorption of N2. X-ray photoelectron spectroscopy (XPS) analysis
181
was conducted on a PHI Quantera II ESCA System with Al Ka radiation at 1486.8 V.
182
2.9 Electrochemical analysis
RI PT
177
All electrochemical analysis of the MFC was conducted on an electrochemical
184
work station (CHI 604E, China). Polarization curves and power density curves were
185
obtained by varying the external resistance from 5000 to 82 Ω and the voltage was
186
recorded until the value was stable for 15 min by an exquisite multimeter. Current
187
density and power density were calculated by normalizing current and power to the
188
anode surface area (8 cm2) according to the following equation (Karra et al., 2013),
190
M AN U
TE D
189
SC
183
Current density =
Power density =
V RA V2 RA
(1) (2)
where V is the voltage across the external resistor, R is the external resistance, and A
192
is the anode surface area.
AC C
193
EP
191
Cyclic voltammetry (CV) measurement was performed using a three-electrode
194
system, where the prepared air-cathode, an Ag/AgCl electrode and a platinum wire
195
were used as the working, reference and counter electrode, respectively. The CV
196
measurement was recorded from -0.8 to 0.4 V at a scan rate of 50 mV s-1 in PBS (pH
197
= 7) saturated by air. Tafel plots and the exchange current density (i0) were recorded
198
by sweeping the overpotential from 0 to 100 mV at 10 mV s-1, calculated by using 9
ACCEPTED MANUSCRIPT 199
the following empirical equation (Huang et al., 2017a). lgi = lgi0 -
200
βnFη
(3)
2.303RT
Electrochemical impedance spectroscopy (EIS) measurement was carried out at
202
the open circuit potential (OCV) with an AC perturbation of 5 mV over a frequency
203
range from 0.1 Hz to 10 kHz. The anode was used as the working electrode, while the
204
cathode as the counter and reference electrode. The resistance data were fitted with
205
Zsimpwin software (Echem), where the ohmic resistance (Rs) was obtained from the
206
Nyquist impedance plots at the point where -Z” was equal to zero at high frequency,
207
the charge-transfer resistance (Rct) was calculated from a semicircular fit of the
208
charge-transfer impedance in the Nyquist plots (Yuan et al., 2015).
209
3. Results and discussion
210
3.1 Synthesis and characterization of the PANI/β-MnO2 composites
TE D
M AN U
SC
RI PT
201
Fig. 1 shows the synthetic diagram of the PANI/β-MnO2 nanocomposites. The
212
nanocomposites were prepared by in-situ PANI polymerization in the presence of
213
β-MnO2 nanorods.
EP
211
Fig. 2 shows XRD pattern of the PANI, β-MnO2 and PANI/β-MnO2
215
nanocomposites. The XRD pattern of PANI showed two broad peaks at 2θ of 20.5°
216
and 25.3°, which represented the periodicities parallel (100) and perpendicular (110)
217
to the PANI chain, respectively (Fig. 2a) (Ryu et al., 2007; Zhu et al., 2007). As
218
shown in Fig. 2b, all of the diffraction peaks for the as-prepared sample were indexed
219
with the standard XRD pattern of β-MnO2 (JCPDS No. 24-0735, tetragonal symmetry
220
with P42/mnm space group and lattice constants of a = 4.399 nm and c = 2.874 nm)
AC C
214
10
ACCEPTED MANUSCRIPT (Zang et al., 2011). There was no significant difference in the diffraction peaks of the
222
PANI/β-MnO2 nanocomposites (Fig. 2c) and pristine β-MnO2 except the diffraction
223
peaks at 2θ = 20.5° and 25.3° which resulted from the PANI on the β-MnO2 surface,
224
confirming the presence of β-MnO2 in the nanocomposites after in-situ polymerization.
225
No characteristic impurity peak was observed, indicating that high purity
226
PANI/β-MnO2 nanocomposites were produced by the simple polymerization.
RI PT
221
Fig. 3 shows the SEM and TEM images of β-MnO2 nanorods and PANI/β-MnO2
228
nanocomposites. As clearly observed in Fig. 3a and b, the as-prepared β-MnO2 shown
229
1-D nanostructured crystals with smooth surface. TEM images in Fig. 3c and d further
230
confirmed the formation of β-MnO2 nanorods with average length of 10-20 µm and
231
average diameter of 200-300 nm. The length and diameter of the as-prepared β-MnO2
232
in this paper were larger than those reported in the literature (Liu et al., 2017), but
233
more orderly (Liu et al., 2005) and scattered (Zang et al., 2011). Compared with the
234
β-MnO2 nanorods, the synthesized PANI/β-MnO2 nanocomposites in Fig. 3e and f
235
exhibited shorter average length about 8-12 µm and had some granules and cracks
236
(red dashed frame) on the surface. The reason of this phenomenon was that β-MnO2
237
worked as oxidant and template during the aniline monomer polymerization to form
238
PANI on the nanorods, and partial β-MnO2 was consumed (Yao et al., 2013). This
239
would lead to an intimate interface, facilitating donor–acceptor interactions between
240
PANI and β-MnO2 (Zhou et al., 2017). TEM images also proved that some small
241
black granules grown on the nanorods (Fig. 3g and h). We composed that these
242
granules were PANI nanogranules which were confirmed by FT-IR and XPS analysis.
AC C
EP
TE D
M AN U
SC
227
11
ACCEPTED MANUSCRIPT From the results of BET analysis, the specific surface area of PANI/β-MnO2
244
nanocomposites was two times more than that of β-MnO2 nanorods. Introducing the
245
PANI granules and cracks increased the specific surface area and will bring about
246
more active sites for ORR of the PANI/β-MnO2 nanocomposites.
RI PT
243
Fig. SM-2 shows the TGA curves of PANI, β-MnO2 and PANI/β-MnO2. PANI
248
showed continuous weight loss till 660 oC, where the weight loss below 300 oC was
249
due to the removal of physic-adsorbed water, interstitial water or dopant anions (Ni et
250
al., 2010; Sen et al., 2013), and the sharp weight loss to 660 oC was ascribed to the
251
large-scale thermal degradation of the PANI chains (Nirmalesh Naveen et al., 2015),
252
while the small weight reserved at 800 oC was due to the final carbonization of
253
intermediate chemicals (Bhadra et al., 2007). Only a small weight loss (3%) was
254
observed for the β-MnO2 nanorods up to 800 oC. However, a small weight loss
255
between 500 and 650 oC, which was probably owing to the reduction of manganese
256
ions going from tetravalent to trivalent state accompanied with oxygen evolution
257
(Zhang et al., 2007b). As observed in Fig. SM-2, weight loss of the PANI/β-MnO2
258
nanocomposites majorly occurred up to 500 oC, which was attributed to the loss of
259
adsorbed water and the decomposition of the PANI (Ni et al., 2010). Obviously,
260
comparing with pure PANI, the thermal decomposition temperature of PANI in the
261
PANI/β-MnO2 nanocomposites was lower. This can be explained by the fact that a
262
coordination bond between manganese and nitrogen atoms in PANI weakened the
263
interactive forces of the PANI inter-chains (Pham et al., 2009; Anu Prathap et al.,
264
2013). The weight was stable at about 80% when the temperature was above 600 oC
AC C
EP
TE D
M AN U
SC
247
12
ACCEPTED MANUSCRIPT 265
and the residues were mainly composed of β-MnO2. It is considered that the weight
266
ratio of PANI and β-MnO2 in the prepared PANI/β-MnO2 was about 1:4. The structural information and chemical component of PANI/β-MnO2
268
nanocomposites were also identified by the FT-IR spectroscopy, as shown in Fig.
269
SM-3. Typical absorption bands at about 1110, 721, and 474 cm−1 attributed to the
270
Mn-O vibrations of MnO6 octahedra in MnO2 (Wang et al., 2008) (Fig. SM-3a) and
271
characteristic absorption peaks at 807 (C-C and C-H in the benzenoid ring), 1122 (the
272
=N+–H -stretching), 1220 (the C–N+ stretching), 1294 (the C–N stretching of the
273
secondary aromatic amine), 1485 (the aromatic C=C stretching of the benzenoid ring),
274
and 1558 cm−1 (the aromatic C=C stretching of the quinonoid ring) (Fig. SM-3b)
275
(Zhang et al., 2013; Yang et al., 2017) for PANI were clearly observed in the spectra
276
of β-MnO2 and PANI, respectively. As for the nanocomposites, Fig. SM-3c shows all
277
the characteristic absorption bands of PANI and β-MnO2. However, the characteristic
278
peaks of Mn-O vibrations at 721 and 474 cm-1 shifted to 712 and 484 cm-1,
279
respectively, reflecting a mutual interaction between PANI and β-MnO2 that, most
280
likely, was a hydrogen bond formed between oxygen atom of Mn-O and N-H in PANI
281
(Yao et al., 2013). In addition, our observation of both the difference in peak position
282
and intensity between the PANI/β-MnO2 and the pure PANI (Fig. SM-3b), further
283
indicating that there was a mutual interaction between PANI and β-MnO2, which
284
verified PANI has been grown on the surface of β-MnO2 nanorods (Sun et al., 2015).
AC C
EP
TE D
M AN U
SC
RI PT
267
285
Fig. SM-4 shows the nitrogen adsorption-desorption isotherm curves of the
286
β-MnO2 nanorods and PANI/β-MnO2 nanocomposites, and the inset graph exhibits 13
ACCEPTED MANUSCRIPT 287
the pore size distribution for the two materials. Both materials showed type IV
288
adsorption-desorption isotherm curves with a hysteresis loop, indicating a
289
well-ordered mesoporous structure.
290
volumes of β-MnO2 nanorods were 19.9 m2 g-1 and 0.023 cm3 g-1. The BET surface
291
areas of β-MnO2 nanorods was larger than those of the commercial β-MnO2 powder
292
(0.11 m2 g-1) (Zhang et al., 2006), pure phase of tetragonal MnO2 (12.4 m2 g-1) (Xu et
293
al., 2016) and MnO2 nanorods ( 16.17 m2 g-1) (Zhang et al., 2007a). However, the
294
as-prepared PANI/β-MnO2 nanocomposites possessed a larger BET surface areas
295
(39.7 m2 g-1) and total pore volumes (0.042 cm3 g-1) than those of pure β-MnO2
296
nanorods (19.9 m2 g-1) in this paper, PPy/β-MnO2 (10.3 m2 g-1) (Chen et al., 2017),
297
np-RuO2/nr-MnO2 (31.9 m2 g-1) (Xu et al., 2016) and Ag/MnO2 rods (19.5 m2 g-1),
298
tubes (25.6 m2 g-1), wires (34.7 m2 g-1) (Li et al., 2016). That was mainly ascribed to
299
introducing PANI granules onto the β-MnO2 nanorods surface. The inset in the figure
300
shows the pore size distribution curves calculated using the Barrett–Joyner–Halenda
301
equation from the desorption branch of the isotherms (Chen et al., 2013). The BJH
302
analysis exhibited that PANI/β-MnO2 nanocomposites had an intensive pore size
303
distribution of ~3.87 nm, which was classified as a mesopore distribution, with an
304
adsorption average pore diameter of 6.63 nm, that was smaller than that of β-MnO2
305
nanorods (6.90 nm). The high surface area and high mesoporous nature of
306
PANI/β-MnO2 nanocomposites may result in more exposure of catalyst active sites
307
diffusing O2 and electrolyte, leading to enhancement during the ORR process.
308
AC C
EP
TE D
M AN U
SC
RI PT
The measured BET surface areas and total pore
Fig. 4 displays the XPS spectra of PANI/β-MnO2 nanocomposites. Four elements 14
ACCEPTED MANUSCRIPT including C, N, O, and Mn were detected in Fig. 4a, demonstrating the presence of
310
PANI and β-MnO2 in the nanocomposites. Furthermore, to examine the oxidation
311
state of Mn in the PANI/β-MnO2 nanocomposites (Pan et al., 2016), XPS spectrum of
312
Mn 2p is plotted in Fig. 4b. Two peaks located at 654.4 and 642.7 eV corresponded to
313
Mn 2p1/2 and Mn 2p3/2, respectively, which was in agreement with the energy
314
separation (11.7 eV) between Mn 2p1/2 and Mn 2p3/2 reported previously (Zhou et al.,
315
2017). As reported previously, the Mn oxidation state in manganese oxides can be
316
determined from the separation of peak energies (∆Eb) between the two peaks of the
317
Mn 3s components (Luan et al., 2016). The as-prepared PANI/β-MnO2 showed a
318
separation energy of 4.83 eV for the Mn 3s doublet (Fig. 4c), which indicated that the
319
dominant Mn4+ ions existed in the nanocomposites and the chemical composition was
320
β-MnO2. The existence of β-MnO2 can be further varied by O1s peaks, which was
321
further deconvoluted into three major bond components in Fig. 4d (Asif et al., 2016).
322
The three peaks centered at binding energy 530.21, 531.41, and 532.32 eV were
323
corresponding to Mn-O-Mn bond for the tetravalent oxide, Mn-O-H bond for
324
hydroxide and H-O-H bond for residual structural water, respectively (Li et al., 2011).
325
3.2 Electrochemical activity of PANI/β-MnO2 for ORR
SC
M AN U
TE D
EP
AC C
326
RI PT
309
The activity of various cathodes for ORR was investigated with both CV and
327
Tafel curves in PBS (pH = 7). Fig. 5a shows the CV curves of four prepared
328
air-cathodes, GF, GF-β-MnO2-6.0, GF-PANI+β-MnO2-6.0 (β-MnO2:PANI = 4:1
329
according to the results of TGA results of PANI/β-MnO2) and GF-PANI/β-MnO2-6.0
330
in air-saturated PBS. Except for GF electrode, all the electrodes exhibited ORR peaks 15
ACCEPTED MANUSCRIPT between
-0.1
and
-0.3
V.
Among
them,
obvious
ORR
peak
of
the
332
GF-PANI/β-MnO2-6.0 electrode was clearly observed at -0.17 V, which was more
333
positive than GF-PANI+β-MnO2-6.0 electrode (-0.21 V) and GF-β-MnO2-6.0
334
electrode (-0.24 V). Remarkably, the ORR peak current of GF-PANI/β-MnO2-6.0
335
electrode was bigger than other three electrodes. Compared with β-MnO2 and
336
PANI+β-MnO2 mixture catalysts, PANI/β-MnO2 nanocomposites demonstrated higher
337
catalytic activity towards ORR (Zhang et al., 2012), that is because (1) PANI itself
338
had the ability to catalyze ORR (Khomenko et al., 2005); (2) introducing PANI
339
granules onto the β-MnO2 nanorods surface increased the surface areas and total pore
340
volumes of catalyst, which might facilitate the diffusion, adsorption, and transport of
341
O2 (Cheng et al., 2010); (3) the specific interaction between β-MnO2 and PANI, which
342
was confirmed by FT-IR analysis, could enhance the electron delocalization and
343
increase the electrical conductivity, resulting in a higher catalytic activity (Ding, 2009;
344
Zhou et al., 2017).
TE D
M AN U
SC
RI PT
331
Exchange current density (i0) is an important factor to evaluate the ORR
346
performance which is also the main factor to estimate the electrogenic activity and
347
charge transfer coefficient of an electrode (Huang et al., 2017b). The Tafel plots and
348
fitting results are shown in Fig. 5b and i0 calculated for all the cathodes were listed in
349
Table SM-1. The valid linear Tafel regression (R2 > 0.99) was determined by the
350
overpotential ranging from 60 to 80 mV. The i0 value of GF-PANI/β-MnO2-6.0
351
cathode (0.66 × 10-4 A cm-2) was higher than those of GF-β-MnO2-6.0 cathode (0.44 ×
352
10-4 A cm-2) and GF-PANI+β-MnO2-6.0 cathode (0.57 × 10-4 A cm-2), which was also
AC C
EP
345
16
ACCEPTED MANUSCRIPT 2.7 times higher than the GF electrode, implying a faster reaction rate and lower
354
activation energy barrier of forward ORR on the GF-PANI/β-MnO2-6.0 electrode.
355
These results indicated that the catalytic activity for ORR was evidently enhanced
356
after generating PANI granules onto the external surface of β-MnO2 nanorods (Zhao
357
et al., 2009). This trend also consisted with the following power density, polarization
358
curves and EIS results, which further illustrated that PANI/β-MnO2 nanocomposites
359
accelerated the kinetics of catalytic activity, consequently, enhancing the power output
360
of MFCs.
361
3.3 MFC performance
M AN U
SC
RI PT
353
To evaluate the performance of as-prepared ORR catalysts, double-chamber
363
MFC were constructed and operated in batch mode. The constructed MFC shared one
364
anode chamber, but used different cathodes after the MFC entered stable period.
TE D
362
Fig. 6 shows the power density curves, polarization curves, anode and cathode
366
potentials, and Nyquist curves of the MFC fabricated with the prepared electrodes.
367
The data were also listed in Table 1. Open circuit voltage (OCV) is a useful indicator
368
of the electrode performance, usually closely related to the electrode activity. As
369
shown in Fig. 6a, the OCVs of the MFC with different air-electrode were in the order
370
of: GF-PANI/β-MnO2-6.0 (770 mV) > GF-PANI+β-MnO2-6.0 (739 mV) >
371
GF-β-MnO2-6.0 (723 mV) > GF (680 mV). In addition the MFC cathode modified by
372
PANI/β-MnO2 had maximum power density (Pmax) of 248 mW m-2, which was 1.22,
373
1.36 and 1.85 times higher than that of PANI+β-MnO2 mixture (204 mW m-2),
374
β-MnO2 (183 mW m-2) and bare (134 mW m-2), higher than CNT-β-MnO2 (98.7 mW
AC C
EP
365
17
ACCEPTED MANUSCRIPT m-2) (Lu et al., 2011), MnOx/C (161 mW m-2) (Roche et al., 2009), 0.4 wt% Pt/MnO2
376
(165 mW m-2) (Khan et al., 2015), MnO2/CNTs (210 mW m-2) (Zhang et al., 2011)
377
and graphite/β-MnO2 (172 mW m-2) (Zhang et al., 2009). The result suggested that the
378
PANI/β-MnO2 nanocomposites catalyst performed higher catalytic activity and
379
improved the power output. Power density and polarization curves were also
380
measured to determine the performances of the MFC with different loading amount of
381
PANI/β-MnO2 nanocomposites (0, 2.2, 4.1 and 6.0 mg cm-2) as cathode catalyst in Fig.
382
6b. Remarkably, the MFC equipped with the GF-PANI/β-MnO2-6.0 electrode
383
generated the largest Pmax, which was 1.29 and 1.63 times higher than that of
384
GF-PANI/β-MnO2-4.1 electrode (191 mW m-2) and GF-PANI/β-MnO2-2.2 electrode
385
(152 mW m-2). Both OCV and power density increased with the increase of loading
386
amount of PANI/β-MnO2 catalyst, which was owing to increasing active sites and
387
accelerating kinetics towards ORR with the increase of catalyst amount.
TE D
M AN U
SC
RI PT
375
The curves of potential versus current density for the individual cathodes and
389
anodes are shown in Fig. 6c. It was obvious that there was an insignificant difference
390
in the anode potentials, but the cathode potentials were considerably different. This
391
phenomenon also occurred in Fig. 6d, which exhibited the potentials of individual
392
cathodes and anodes of different loading amount of PANI/β-MnO2 nanocomposites as
393
cathode catalyst. These observations implied that the differences in power output
394
resulted from cathodes rather than anodes (Huang et al., 2017a) and the differences in
395
the MFC performances mainly were attribute to variations of the cathode catalysts and
396
the loading amount of catalysts in the catalytic activities.
AC C
EP
388
18
ACCEPTED MANUSCRIPT For ensuring a relatively smooth transfer of electrons, the cathode electrocatalyst
398
must have high conductivity (Hao et al., 2017). All the Nyquist plots of the electrodes
399
with different catalysts and different loading amount of PANI/β-MnO2 catalyst were
400
tested using AC impedance measurements. The Nyquist plots and the equivalent
401
circuit model are shown in Fig. 6e and f. The results indicated that there was a
402
remarkable decrease in the resistance of MFC system after the addition of cathode
403
catalyst. Both the ohmic resistance (Rs) and charge transfer resistance (Rct) of the GF
404
electrode
405
GF-PANI/β-MnO2-6.0 electrode in Fig. 6e. The values of Rs for GF, GF-β-MnO2-6.0,
406
GF-PANI+β-MnO2-6.0 and GF-PANI/β-MnO2-6.0 electrode were estimated to be
407
50.4, 45.5, 42.3 and 38.4 Ω, respectively. Besides, there was an obvious reduction in
408
Rct for catalyst coated cathodes. The GF-PANI/β-MnO2-6.0 electrode exhibited Rct of
409
9.2 Ω, which was lower than GF (93.3 Ω), GF-β-MnO2-6.0 (19.8 Ω), and
410
GF-PANI+β-MnO2-6.0 (14.4 Ω) electrode. The lowest Rs and Rct obtained for
411
GF-PANI/β-MnO2-6.0 electrode might be explained by pronounced synergistic effect
412
between PANI and β-MnO2 of the PANI/β-MnO2 catalyst, which increased the
413
conductivity of the electrode surface and accelerated electron transfer throughout the
414
electrode (Wang et al., 2012). Fig. 6f shows the Nyquist plots of the cathodes with
415
different loading amount of PANI/β-MnO2 catalyst. The Rs of GF-PANI/β-MnO2-6.0
416
electrode was ~1.15, 1.23 and 1.31 times lower than GF-PANI/β-MnO2-4.1,
417
GF-PANI/β-MnO2-2.2 and bare electrode. In addition, there was also a reduction
418
trend in Rct. The Rct of GF-PANI/β-MnO2-6.0 electrode was lower than bare (93.3 Ω),
higher
than
GF-β-MnO2-6.0,
GF-PANI+β-MnO2-6.0
and
AC C
EP
TE D
M AN U
were
SC
RI PT
397
19
ACCEPTED MANUSCRIPT GF-PANI/β-MnO2-2.2 (23.9 Ω) and GF-PANI/β-MnO2-4.1 (15.6 Ω) electrode. This
420
reduction in Rs and Rct could result from the more PANI/β-MnO2 catalyst on the
421
cathode surface, which were deemed to provide excess active sites for oxygen
422
molecules and improve more electrocatalytic activity of the cathode for ORR. The
423
total internal resistance of the MFC was estimated by polarization slope method,
424
which was denoted as Rint and listed in Table 1. It was also in good agreement with
425
power density results, which further indicated that the loading of PANI/β-MnO2
426
catalyst onto graphite felt surface decreased the resistance and accelerated the kinetics
427
activity. Therefore, it will be an efficient ORR catalyst for MFC applications to
428
increase power output.
429
4. Conclusion
M AN U
SC
RI PT
419
Novel PANI/β-MnO2 nanocomposites prepared by hydrothermal and in-situ
431
chemical oxidative polymerization were used as cathode catalyst in MFCs. The
432
PANI/β-MnO2 exhibited higher catalytic activity for ORR than β-MnO2 and
433
PANI+β-MnO2 mixture, owing to the interaction between β-MnO2 and PANI,
434
increased specific surface area and electric conductivity, which facilitated the electron
435
transfer process. Several tests results indicated that the PANI/β-MnO2 cathode did
436
promote the power generation of the MFC, the maximum power density was 1.2-1.4
437
times as high as those by controlled cathodes, while cathode resistance was 1.1-1.2
438
times lower, indicating the promising potential as ORR catalyst in MFC applications.
439
Acknowledgements
440
AC C
EP
TE D
430
We gratefully thank the National Natural Science Foundation of China 20
ACCEPTED MANUSCRIPT 441
(21276128), the Basic Research Program of Jiangsu Province of China (BK20141398)
442
for the financial support.
443
References
444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481
Anu Prathap, M.U., Satpati, B., Srivastava, R., 2013. Facile preparation of polyaniline/MnO2 nanofibers
RI PT
and its electrochemical application in the simultaneous determination of catechol, hydroquinone, and resorcinol. Sens. Actuators, B. 186, 67-77.
Asif, M., Tan, Y., Pan, L., Li, J., Rashad, M., Fu, X., Cui, R., Usman, M., 2016. Improved performance of a MnO2@PANI nanocomposite synthesized on 3D graphene as a binder free electrode for supercapacitors. RSC Adv. 6, 46100-46107.
Bhadra, S., Singha, N.K., Chattopadhyay, S., Khastgir, D., 2007. Effect of different reaction parameters
SC
on the conductivity and dielectric properties of polyaniline synthesized electrochemically and modeling of conductivity against reaction parameters through regression analysis. J. Polym. Sci., Part B: Polym. Phys. 45, 2046-2059.
M AN U
Chen, D., Shen, J. Y., Jiang, X.B., Mu, Y., Ma, D.H., Han, W.Q., Sun, X.Y., Li, J.S., Wang, L.J., 2017. Fabrication of polypyrrole/β-MnO2 modified graphite felt anode for enhancing recalcitrant phenol degradation in a bioelectrochemical system. Electrochim. Acta. 244, 119-128. Chen, L., Song, Z., Liu, G., Qiu, J., Yu, C., Qin, J., Ma, L., Tian, F., Liu, W., 2013. Synthesis and electrochemical performance of polyaniline–MnO2 nanowire composites for supercapacitors. J. Phys. Chem.Solids. 74, 360-365.
Cheng, F.Y., Su, Y., Liang, J., Tao, Z.L., Chen, J., 2010. MnO2-Based Nanostructures as Catalysts for
TE D
Electrochemical Oxygen Reduction in Alkaline Media. Chem. Mater. 22, 898-905. Cheng, S., Liu, H., Logan, B.E., 2006. Power Densities Using Different Cathode Catalysts (Pt and CoTMPP) and Polymer Binders (Nafion and PTFE) in Single Chamber Microbial Fuel Cells. Environ. Sci. Technol. 40, 364-369.
Ding, K.Q., 2009. Cyclic Voltammetrically Prepared MnO2-Polyaniline Composite and Its
EP
Electrocatalysis for Oxygen Reduction Reaction (ORR). J. Chin. Chem. Soc. 56, 891-897. Dong, H., Yu, H., Wang, X., 2012a. Catalysis Kinetics and Porous Analysis of Rolling Activated Carbon-PTFE Air-Cathode in Microbial Fuel Cells. Environ. Sci. Technol. 46, 13009-13015.
AC C
Dong, H., Yu, H., Wang, X., Zhou, Q., Feng, J., 2012b. A novel structure of scalable air-cathode without Nafion and Pt by rolling activated carbon and PTFE as catalyst layer in microbial fuel cells. Water Res. 46, 5777-5787.
Hao, L., Yu, J., Xu, X., Yang, L., Xing, Z., Dai, Y., Sun, Y., Zou, J., 2017. Nitrogen-doped MoS2/carbon as highly oxygen-permeable and stable catalysts for oxygen reduction reaction in microbial fuel cells. J. Power Sources. 339, 68-79.
He, X.M., Li, C., Chen, F.E., Shi, G.Q., 2007. Polypyrrole Microtubule Actuators for Seizing and Transferring Microparticles. Adv. Funct. Mater. 17, 2911-2917. Huang, Q., Zhou, P., Yang, H., Zhu, L., Wu, H., 2017a. CoO nanosheets in situ grown on nitrogen-doped activated carbon as an effective cathodic electrocatalyst for oxygen reduction reaction in microbial fuel cells. Electrochim. Acta. 232, 339-347. Huang, Q., Zhou, P., Yang, H., Zhu, L., Wu, H., 2017b. In situ generation of inverse spinel CoFe2O4 nanoparticles onto nitrogen-doped activated carbon for an effective cathode electrocatalyst of 21
ACCEPTED MANUSCRIPT microbial fuel cells. Chem. Eng. J. 325, 466-473. Jiang, H., Ma, J., Li, C., 2012. Polyaniline–MnO2 coaxial nanofiber with hierarchical structure for high-performance supercapacitors. J. Mater. Chem. 22, 16939. Ma, J.J., Liu, L., Zhu, X.F., Zhang., T.L., 2007. Preparation and properties of polyaniline doped with hydrochloric acid. Chem. Res. Appl. 19, 1233-1237. Karra, U., Manickam, S.S., McCutcheon, J.R., Patel, N., Li, B., 2013. Power generation and organics J. Hydrogen Energ. 38, 1588-1597.
RI PT
removal from wastewater using activated carbon nanofiber (ACNF) microbial fuel cells (MFCs). Int. Khan, M.R., Chen, K.M., Ong, H.R., Cheng, C.K., Rahman, W., 2015. Nanostructured Pt/MnO2 Catalysts and Their Performance for Oxygen Reduction Reaction in Air Cathode Microbial Fuel Cell. Int. J. Electr., Comput., Electron. Commun. Eng. 9, 247-253.
Khomenko, V.G., Barsukov, V.Z., Katashinskii, A.S., 2005. The catalytic activity of conducting polymers toward oxygen reduction. Electrochim. Acta. 50, 1675-1683.
SC
Lai, B., Tang, X., Li, H., Du, Z., Liu, X., Zhang, Q., 2011. Power production enhancement with a polyaniline modified anode in microbial fuel cells. Biosens. Bioelectron. 28, 373-377. Li, J.M., Qu, Z.P., Qin, Y., Wang, H., 2016. Effect of MnO2 morphology on the catalytic oxidation of
M AN U
toluene over Ag/MnO2 catalysts. Appl. Surf. Sci. 385, 234-240.
Li, X.G., Li, A., Huang, M.R., 2008a. Facile high-yield synthesis of polyaniline nanosticks with intrinsic stability and electrical conductivity. Chem. Eur. J. 14, 10309-10317.
Li, X.G., Li, J., Meng, Q.K., Huang, M.R., 2009. Interfacial Synthesis and Widely Controllable Conductivity of Polythiophene Microparticles. J. Phys. Chem. B. 113, 9718–9727. Li, X.G., Lu, Q.F., Huang, M.R., 2008b. Self-stabilized nanoparticles of intrinsically conducting copolymers from 5-sulfonic-2-anisidine. Small. 4, 1201-1209.
TE D
Li, Z., Wang, J., Liu, X., Liu, S., Ou, J., Yang, S., 2011. Electrostatic layer-by-layer self-assembly multilayer films based on graphene and manganese dioxide sheets as novel electrode materials for supercapacitors. J. Mater. Chem. 21, 3397.
Liu, C.S.T., Navale, S.T., Yang, Z.B., Galluzzi, M., Patil, V.B., Cao, P.J., Mane, R.S., Stadler, F.J., 2017. Ethanol gas sensing properties of hydrothermally grown α-MnO2 nanorods. J. Alloys Compd. 727,
EP
362-369.
Liu, X.M., Fu, S.Y., Huang, C.J., 2005. Synthesis, characterization and magnetic properties of β-MnO2 nanorods. Powder. Technol. 154, 120-124. Lu, M., Kharkwal, S., Ng, H.Y., Li, S.F., 2011. Carbon nanotube supported MnO2 catalysts for oxygen
AC C
482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525
reduction reaction and their applications in microbial fuel cells. Biosens. Bioelectron. 26, 4728-4732.
Luan, Y.T., Huang, Y.Q., Wang, L.J., Li, M.X., Wang, R.H., Jiang, B.J., 2016. 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, 90-96.
Ni, W., Wang, D., Huang, Z., Zhao, J., Cui, G., 2010. Fabrication of nanocomposite electrode with MnO2 nanoparticles distributed in polyaniline for electrochemical capacitors. Mater. Chem. Phys. 124, 1151-1154. Nirmalesh Naveen, A., Selladurai, S., 2015. Fabrication and performance evaluation of symmetrical supercapacitor based on manganese oxide nanorods–PANI composite. Mat. Sci. Semicon.Proc. 40, 468-478. Pan, T., Liu, H.Y., Ren, G.Y., Li, Y.N., Lu, X.Y., Zhu, Y., 2016. Metal-free porous nitrogen-doped carbon 22
ACCEPTED MANUSCRIPT nanotubes for enhanced oxygen reduction and evolution reactions. Sci. Bull. 61, 889-896. Pham, Q., Pham, D., Kim, J., Kim, E., Kim, S., 2009. Preparation of polyaniline–titanium dioxide hybrid materials in supercritical CO2. Synth. Met. 159, 2141-2146. Roche, I., Katuri, K., Scott, K., 2009. A microbial fuel cell using manganese oxide oxygen reduction catalysts. J. Appl. Electrochem. 40, 13-21. Ryu, K.S., Jeong, S.K., Joo, J., Kim, K.M., 2007. Polyaniline Doped with Dimethyl Sulfate as a Battery and a Redox Supercapacitor. J. Phys. Chem. B 111, 731-739.
RI PT
Nucleophilic Dopant and Its Electrochemical Properties as an Electrode in a Lithium Secondary Sen, P., De, A., Chowdhury, A.D., Bandyopadhyay, S.K., Agnihotri, N., Mukherjee, M., 2013. Conducting polymer based manganese dioxide nanocomposite as supercapacitor. Electrochim. Acta. 108, 265-273.
Sun, X.W., Gan, M.Y., Ma, L., Wang, H.H., Zhou, T., Wang, S.Y., Dai, W.Q., Wang, H.N., 2015. Fabrication for energy storage. Electrochim. Acta. 180, 977-982.
SC
of PANI-coated honeycomb-like MnO2 nanospheres with enhanced electrochemical performance Tran, H.D., Shin, K., Hong, W.G., D'Arcy, J.M., Kojima, R.W., Weiller, B.H., Kaner, R.B., 2007. A Template-Free Route to Polypyrrole Nanofibers. Macromol. Rapid. Comm. 28, 2289-2293.
M AN U
Wang, J.G., Yang, Y., Huang, Z.H., Kang, F., 2012. Rational synthesis of MnO2/conducting polypyrrole@carbon nanofiber triaxial nano-cables for high-performance supercapacitors. J. Mater. Chem. 22, 16943.
Wang, Y.G., Wu, W., Cheng, L., He, P., Wang, C.X., Xia, Y.Y., 2008. A Polyaniline-Intercalated Layered Manganese Oxide Nanocomposite Prepared by an Inorganic/Organic Interface Reaction and Its High Electrochemical Performance for Li Storage. Adv. Mater. 20, 2166-2170. Wang, Y., Wen, Q., Chen, Y., Qi, L., 2017. A novel polyaniline interlayer manganese dioxide composite
TE D
anode for high-performance microbial fuel cell. J. Taiwan Institute. Chem. Eng. 75, 112-118. Wen, Q., Wang, S., Yan, J., Cong, L., Pan, Z., Ren, Y., Fan, Z., 2012. MnO2–graphene hybrid as an alternative cathodic catalyst to platinum in microbial fuel cells. J. Power Sources. 216, 187-191. Xu, Y.F., Chen, Y., Xu, G.L., Zhang, X.R., Chen, Z.H., Li, J.T., Huang, L., Amine, K., Sun, S.G., 2016. RuO2 nanoparticles supported on MnO2 nanorods as high efficient bifunctional electrocatalyst of
EP
lithium-oxygen battery. Nano Energy. 28, 63-70. Yang, C.X., Dong, W.P., Cui, G. W., Zhao, Y.Q., Shi, X. F., Xia, X.Y., Tang, B., Wang, W.L., 2017. Enhanced photocatalytic activity of PANI/TiO2 due to their photosensitization-synergetic effect. Electrochim. Acta. 247, 486-495.
AC C
526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569
Yao, W., Zhou, H., Lu, Y., 2013. Synthesis and property of novel MnO2@polypyrrole coaxial nanotubes as electrode material for supercapacitors. J. Power Sources. 241, 359-366.
Yuan, C., Su, L., Gao, B., Zhang, X., 2008. Enhanced electrochemical stability and charge storage of MnO2/carbon nanotubes composite modified by polyaniline coating layer in acidic electrolytes. Electrochim. Acta. 53, 7039-7047.
Yuan, H., Deng, L., Tang, J., Zhou, S., Chen, Y., Yuan, Y., 2015. Facile Synthesis of MnO2/Polypyrrole/MnO2 Multiwalled Nanotubes as Advanced Electrocatalysts for the Oxygen Reduction Reaction. ChemElectroChem. 2, 1152-1158. Zang, J., Li, X., 2011. In situ synthesis of ultrafine β-MnO2/polypyrrole nanorod composites for high-performance supercapacitors. J. Mater. Chem. 21, 10965. Zhang, L., Liu, C., Zhuang, L., Li, W., Zhou, S., Zhang, J., 2009. Manganese dioxide as an alternative cathodic catalyst to platinum in microbial fuel cells. Biosens. Bioelectron. 24, 2825-2829. 23
ACCEPTED MANUSCRIPT
595 596 597 598 599
Zhang, W.X., Wang, H., Yang, Z.H., Wang, F., 2007a. Promotion of H2O2 decomposition activity over β-MnO2 nanorod catalysts. Colloids Surf., A. 304, 60-66. Zhang, W.X., Yang, Z.H., Wang, X., Zhang, Y.C., Wen, X.G., Yang, S.H., 2006. Large-scale synthesis of β-MnO2 nanorods and their rapid and efficient catalytic oxidation of methylene blue dye. Catal. Commun. 7, 408-412.
RI PT
Zhang, X., Ji, L., Zhang, S., Yang, W., 2007b. Synthesis of a novel polyaniline-intercalated layered manganese oxide nanocomposite as electrode material for electrochemical capacitor. J. Power Sources. 173, 1017-1023.
Zhang, Y., Mo, G., Li, X., Ye, J., 2012. Iron tetrasulfophthalocyanine functionalized graphene as a platinum-free cathodic catalyst for efficient oxygen reduction in microbial fuel cells. J. Power Sources. 197, 93-96.
SC
Zhang, Y.P., Hu, Y.Y., Li, S.Z., Sun, J., Hou, B., 2011. Manganese dioxide-coated carbon nanotubes as an improved cathodic catalyst for oxygen reduction in a microbial fuel cell. J. Power Sources. 196, 9284-9289.
M AN U
Zhao, F., Slade, R.C., Varcoe, J.R., 2009. Techniques for the study and development of microbial fuel cells: an electrochemical perspective. Chem. Soc. Rev. 38, 1926-1939. Zhou, D., Che, B., Lu, X., 2017. Rapid one-pot electrodeposition of polyaniline/manganese dioxide hybrids: a facile approach to stable high-performance anodic electrochromic materials. J. Mater. Chem. C. 5, 1758-1766.
Zhu, C.L., Chou, S.W., He, S.F., Liao, W.N., Chen, C.C., 2007. Synthesis of core/shell metal oxide/polyaniline nanocomposites and hollow polyaniline capsules. Nanotechnology. 18, 275604.
TE D
594
electrode with surface modification for H2/Br2 fuel cell. J. Power Sources. 242, 15-22.
EP
593
Zhang, L., Shao, Z. G., Wang, X., Yu, H., Liu, S., Yi, B., 2013. The characterization of graphite felt
AC C
570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592
600 601 602 603 24
ACCEPTED MANUSCRIPT 604 605 606
RI PT
607 608
AC C
EP
TE D
M AN U
SC
609
25
ACCEPTED MANUSCRIPT Figure captions Fig. 1. Schematic of the preparation of the PANI/β-MnO2 nanocomposites. Fig. 2. XRD patterns of (a) PANI powders, (b) β-MnO2 nanorods and (c)
RI PT
PANI/β-MnO2 nanocomposites. Fig. 3. SEM images of β-MnO2 (a and b) and PANI/β-MnO2 (e and f), and TEM images of β-MnO2 (c and d) and PANI/β-MnO2 (g and h).
SC
Fig. 4. X-ray photoelectron spectroscopy (XPS) analysis of the PANI/β-MnO2
M AN U
nanocomposites: (a) survey spectra, (b) Mn 2p spectra, (c) Mn 3s spectra, and d O 1s spectra.
Fig. 5. (a) CV curves for ORR in air-saturated PBS (pH=7) at the scan rate of 50 mV s-1 with different catalyst cathodes and (b) Tafel plots of all cathodes at 1 mV s-1 and
TE D
the linear fit for the Tafel plots (inset).
Fig. 6. Power density and polarization curves as function of current density of MFC with cathode loading different types of catalyst (a), and with cathode loading different
EP
amount of PANI/β-MnO2 (b). Individual potential (vs. Ag/AgCl) as function of current
AC C
density of MFC with cathode loading different types of catalyst (c), and with cathode loading different amount of PANI/β-MnO2 (d). Nyquist curves of the EIS data of MFC with cathode loading different types of catalyst (e), and with cathode loading different amount of PANI/β-MnO2 (f).
ACCEPTED MANUSCRIPT Table 1. Performance of the MFC using GF-PANI/β-MnO2 air-cathode. Cathode resistance OCV
OCP
Pmax
Rint (Ω)
Cathode (mV)
(cathode/mV)
-2
(mW m )
(Ω)
RI PT
Rs
Rct
680
244
134
243.7
50.4
93.3
GF-β-MnO2-6.0
723
276
183
175.3
45.5
19.8
GF-PANI+β-MnO2-6.0
739
281
204
164.7
42.3
14.4
GF-PANI/β-MnO2-6.0
770
307
248
155.6
38.4
9.2
GF-PANI/β-MnO2-4.1
730
GF-PANI/β-MnO2-2.2
705
M AN U
TE D EP AC C
SC
GF
280
191
170.2
44.2
15.6
258
152
182.1
47.2
23.9
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT Highlights The PANI/β-MnO2 nanocomposites are synthesized successfully.
The nanocomposites serve as electrocatalysts for the air cathode of MFCs.
The performance of PANI/β-MnO2 nanocomposites modified MFC is enhanced.
The nanocomposites are potential electrocatalysts for air-cathode MFCs.
AC C
EP
TE D
M AN U
SC
RI PT