You are on page 1of 10

Home Search Collections Journals About Contact us My IOPscience

Synthesis, characterization and capacitive performance of hydrous manganese dioxide

nanostructures

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2011 Nanotechnology 22 125703

(http://iopscience.iop.org/0957-4484/22/12/125703)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 128.114.34.22
This content was downloaded on 02/01/2015 at 18:07

Please note that terms and conditions apply.


IOP PUBLISHING NANOTECHNOLOGY
Nanotechnology 22 (2011) 125703 (9pp) doi:10.1088/0957-4484/22/12/125703

Synthesis, characterization and capacitive


performance of hydrous manganese
dioxide nanostructures
Jintao Zhang1 , Wei Chu2 , Jianwen Jiang1 and X S Zhao1,2
1
Department of Chemical and Biomolecular Engineering, National University of Singapore,
4 Engineering Drive 4, Singapore 117576, Singapore
2
College of Chemical Engineering, Sichuan University, Chengdu 610065,
People’s Republic of China

E-mail: chezxs@nus.edu.sg

Received 14 October 2010, in final form 27 December 2010


Published 14 February 2011
Online at stacks.iop.org/Nano/22/125703

Abstract
Hydrous manganese dioxide nanostructures were synthesized via a catalytic oxidation reaction
mechanism at mild temperatures. It was found that the morphology of the manganese dioxide
nanostructures was significantly influenced by the pH of the reaction system. With increasing
pH the morphology of manganese dioxide nanostructures changed from urchin-like structures
to nanobelts. The capacitive performance was investigated by using cycle voltammetry and
galvanostatic charge/discharge techniques. Hydrous manganese dioxide nanostructures obtained
from a basic solution exhibited a capacitance of 262 F g−1 at a current density of 250 mA g−1
and a capacitive retention of 75% after 1200 cycles, suggesting that this is a promising electrode
material for supercapacitors. The high specific capacitance is attributed to the hydrous nature
coupled with a high surface area (181 m2 g−1 ) of the manganese dioxide nanostructure.
S Online supplementary data available from stacks.iop.org/Nano/22/125703/mmedia
(Some figures in this article are in colour only in the electronic version)

1. Introduction used as supercapacitor electrodes [10]. Therefore, much


effort has been made to synthesize manganese oxides of
Supercapacitors, as energy storage devices, possess a higher various nanostructures [9, 11–13]. Li and co-workers [14]
specific power density than batteries and a higher specific described a hydrothermal method for synthesizing manganese
dioxide nanostructures via a redox reaction mechanism with
energy density than conventional capacitors [1, 2]. Based
manganese sulfate and ammonium persulfate as the precursor
on the charge storage mechanism, there are two types of
at 140 ◦ C. It was found that the hydrothermal time and
supercapacitors, namely electrical double layer capacitors
temperature had a great influence on the capacitive properties
(EDLCs) and pseudocapacitors. The pseudocapacitance comes
of the resultant manganese oxide along with the crystallinity,
from the reversible redox reaction of electroactive materials,
structure, and morphology [11, 15]. It was also reported that
such as conducting polymers and transition metal oxides [3]. the redox reaction between manganese sulfate and ammonium
However, the poor cycling stability of conducting polymers persulfate can be carried out at room temperature with
during charge/discharge processes limits their applications [4]. silver ions as a catalyst [16, 17]. Hollow structures of
Among transition metal oxides, hydrous ruthenium dioxide manganese oxide were synthesized under acidic conditions by
(RuO2 ) offers a high pseudocapacitance [5, 6]. The high using the catalytic method [18]. Electrochemical deposition
cost of this noble metal is a big hurdle for practical uses [7]. methods were also employed to prepare manganese oxide
Manganese oxides have attracted growing interest because of nanostructures [12, 19]. The physicochemical properties of
their low cost, environmental friendliness, natural abundance, the manganese oxide nanomaterials are significantly influenced
and relatively high energy density [8, 9]. But bulk manganese by the preparation method because different methods result in
oxides display a poor rate capability and reversibility when different surface states and defects [20, 21].

0957-4484/11/125703+09$33.00 1 © 2011 IOP Publishing Ltd Printed in the UK & the USA
Nanotechnology 22 (2011) 125703 J Zhang et al

In this work, hydrous manganese dioxide nanostructures


with different morphologies were synthesized under different
pH conditions. The sample synthesized under basic conditions
was observed to exhibit a capacitance as high as 262 F g−1 at
a current density of 250 mA g−1 and a capacitive retention of
75% after 1200 cycles.

2. Experimental details

The hydrous manganese dioxide nanostructures were synthe-


sized using a low-temperature catalytic method [17]. In a
typical synthesis, 1.4 g of manganese acetate (Mn(CH3 COO)2 )
and 2.4 g of potassium persulfate (K2 S2 O8 ) were respectively
dissolved in 60 ml of deionized (DI) water. Then 150 μl of
silver nitrate (100 mM) was added into the K2 S2 O8 solution.
The two solutions were mixed under stirring at 60 ◦ C. After
12 h, the precipitates were filtrated off, washed with DI water
and ethanol, and dried at 60 ◦ C for 12 h. Another two Figure 1. XRD patterns of MnO2 –A (a), MnO2 –N (b), and
MnO2 –B (c).
samples were also synthesized similarly except for adding 1 ml
NH3 ·H2 O and 1 ml H2 SO4 , respectively, to regulate the pH.
The solid samples thus obtained under basic (pH ∼ 11), neutral was added to paste the slurry on the GCE. A 0.5 M Na2 SO4
(pH ∼ 8), and acidic (pH ∼ 1) conditions are designated as solution was employed as the electrolyte.
MnO2 –B, MnO2 –N, and MnO2 –A, respectively.
Powder x-ray diffraction (XRD) measurements were 3. Results and discussion
performed on a Shimadzu diffractometer operated using the
reflection mode with Cu Kα radiation at a step size of 3.1. Synthesis and characterization of hydrous manganese
0.06◦ s−1 . Fourier transform infrared (FT-IR) spectra were dioxide nanostructures
recorded on a Perkin-Elmer 580B infrared spectrophotometer In the hydrothermal system, Mn(CH3 COO)2 was oxidized
using the KBr pellet technique. Thermogravimetric analysis by K2 S2 O8 at 60 ◦ C in the presence of AgNO3 as a catalyst
(TGA) data were collected on a thermal analysis instrument according to the following reaction [16, 18]:
(SDT 2960, TA Instruments, New Castle, DE) with a heating
rate of 10 ◦ C min−1 in an air flow of 100 ml min−1 . X- Ag+
Mn2+ + S2 O28− + 2H2 O −→ MnO2 + 2SO24− + 4H+ . (1)
ray photoelectron spectroscopy (XPS) measurements were
conducted on a PHI-5300 ESCA spectrometer (PerkinElmer) Figure 1 shows the XRD patterns of samples MnO2 –
with an energy analyzer working in the pass energy mode at A, MnO2 –N, and MnO2 –B. The XRD pattern of MnO2 –A
35.75 eV and the Al KR line was used as the excitation source. (figure 1(a)) shows a series of diffraction peaks, which can
Nitrogen adsorption–desorption isotherms were measured on a be indexed to the characteristic reflections of α -MnO2 (JCPDS
Micromeritics ASAP 2020 system at −196 ◦ C. The specific card no. 44-0141) [22, 23]. As for the XRD pattern of samples
surface areas were calculated using the Brunauer–Emmett– MnO2 –N (figure 1(b)) and MnO2 –B (figure 1(c)), the main
Teller (BET) method. The pore volumes were estimated peaks can also be indexed to the α -MnO2 phase. The broader
from the nitrogen volume adsorbed at a relative pressure peaks indicate smaller crystalline particles or poor crystallinity
of 0.95. Pore size distribution curves were computed from of MnO2 –N and MnO2 –B than that of MnO2 –A. The peak at
the desorption branches of the isotherms using the Barrett, about 22.2◦ observed on sample MnO2 –N as indicated by a star
Joyner, and Halenda (BJH) method. The morphologies of was attributed to the diffraction of an impurity. The present
samples were characterized using a field-emission scanning results indicate that the acidic condition was favorable for the
electron microscope (FESEM; JEOL, JSM-6700F) and a high- formation of highly crystalline α -MnO2 .
resolution transmission electron microscopy (HRTEM; JEOL- Figure 2 shows the FESEM images of MnO2 –B, MnO2 –
2100, 200 kV), respectively. N, and MnO2 –A. The low-magnification images of MnO2 –
Electrochemical measurements were performed with B (figure 2(a)) and MnO2 –N (figure 2(c)) show the large
an Autolab PGSTAT302N at ambient temperature (about aggregated clusters. The enlarged image shown in figure 2(b)
22 ◦ C). The electrochemical cell (see figure S1 available at reveals that MnO2 –B consists of nanowires, which are
stacks.iop.org/Nano/22/125703/mmedia) had a three-electrode entangled with one another to form big aggregates. In contrast,
configuration with a bright Pt plate as the counter electrode and MnO2 –N consists of nanoplates and nanowires. The sample
an Ag/AgCl electrode as the reference electrode. A polished synthesized under acidic conditions, MnO2 –A, was composed
glassy carbon working electrode (GCE) of diameter 5 mm was of intercrossed nanorods with sharp tips, exhibiting an urchin-
used as a current collector. An electroactive sample was mixed like morphology.
with acetylene black (mass ratio = 80:20) and ethanol to Figure 3 shows the HRTEM images of MnO2 –B, MnO2 –
achieve a slurry. A small amount of Nafion solution (0.5 wt%) N, and MnO2 –A. It can be seen that MnO2 –B is composed

2
Nanotechnology 22 (2011) 125703 J Zhang et al

Figure 2. FESEM images of MnO2 –B (a), (b), MnO2 –N (c), (d), and MnO2 –A (e), (f).

of nanowires. The enlarged image shown in figure 3(b) To investigate the formation mechanism of the manganese
reveals a belt shape of the nanowires. The HRTEM dioxide nanostructures synthesized under different pH condi-
image of the individual nanobelts shows the clear lattice tions, intermediates were collected after different times and
fringes. The lattice spacing between adjacent lattice planes characterized using the FESEM technique. Figure 4 shows the
is ∼0.49 nm (figure 3(c)), corresponding to the distance of FESEM images of the intermediates collected after different
two neighboring (200) crystal planes in α -MnO2 [24]. The reaction times. It is seen from figure 4(a1) that the intermediate
corresponding fast Fourier transform pattern (FFT) of the synthesized in an acidic medium after 1 h reaction displays
individual nanobelts shown in figure 3(d) shows the diffraction a three-dimensional (3D) spherical morphology resembling a
spots, indicating a single crystalline nature of the individual ball of nanosheets. Figure 4(a2) shows the morphology of
nanobelts. Sample MnO2 –N (figure 3(e)) contained both the intermediate collected after 4 h, exhibiting the spherical
nanoplates and nanowires. The lattice-resolved HRTEM structures with lots of sharp nanorods covering the surface.
image (figure 3(f)) reveals a lattice spacing of ∼0.24 nm, Upon further increasing the reaction time to 8 h, the spheres
indicating the presence of α -MnO2 phase. The TEM image transformed to an urchin-like structure consisting of nanorods.
in figure 3(g) shows that MnO2 –A consists of intercrossed An enlarged FESEM image shown in figure S2 (available
nanorods. The enlarged image of typical nanorods shown in at stacks.iop.org/Nano/22/125703/mmedia) reveals that the
figure 3(h) exhibits a pyramidal morphology of the nanorods. nanorods were epitaxially grown from nanosheets on the
The HRTEM image depicted in figure 3(i) reveals the lattice surface of the macrospheres. According to the FESEM
spacing of the nanorods of ∼0.69 nm, the distance between results, it is believed that the urchin-like structure formed
two (110) crystal planes, confirming the formation of α - via a ripening process or dissolution–recrystallization process
MnO2 [15]. On the basis of the FESEM and HRTEM results, because of the great difference between the morphologies of
it can be concluded that the morphology and structure of the the intermediates and the final product. Initially, the fast
manganese oxide materials are strongly dependent on the pH formation of MnO2 resulted in the aggregation of nanosheets
of the synthesis system. to form spheres under acidic solutions. Then, the epitaxial

3
Nanotechnology 22 (2011) 125703 J Zhang et al

Figure 3. TEM and HRTEM images of MnO2 –B (a)–(c), MnO2 –N (e), (f), and MnO2 –A (g)–(i). The corresponding FFT pattern of HRTEM
image (c), (d).

growth of nanorods on the surface of macrosphere occurred nanowires and nanoplates (figure 4(b3)). The evolution of
with the slower reaction rate due to the decrease in the MnO2 –B exhibits a similar process, with the initially formed
concentration of reactants. As the reaction proceeded, the clusters with small nanosheets undergoing a transformation to
whole system was transferred to a thermodynamically stable nanobelts with increasing reaction time (figures 4(c1)–(c3)).
environment due to the exhaustion of reactants [18]. With On the basis of the above observations the formation
increasing reaction time the MnO2 nanosheets dissolved and of MnO2 –A, MnO2 –N and MnO2 –B is believed to undergo
then recrystallized along the previously formed nanorods a similar ripening process, which is characteristic of
because of the one-dimensional (1D) growth habit of the α - a dissolution–recrystallization process. However, great
MnO2 crystal [14, 25, 26], resulting in the formation of the difference between the morphologies of intermediates and the
urchin-like structures [22]. final products were observed. The formation of MnO2 –A
Figures 4(b1)–(b3) show the morphology of the interme- agrees well with what has been observed previously [22, 27].
diates during the formation of MnO2 –N. The FESEM image Our experimental results indicated that 1D structures
of the intermediate collected after 1 h reveals nanorod clusters (e.g. nanowires and nanobelts) preferably formed at high pH,
along with few having a lamellar structure. Further increasing although other factors might also influence the growth of MnO2
the reaction time led to the formation of clusters with larger nanostructures.
nanorods (figure 4(b2)), which may be due to growth of newly Figure 5 shows the FT-IR spectra of the samples. The
formed MnO2 colloids on the early formed ones. As the strong peaks of MnO2 –B, MnO2 –N, and MnO2 –A appearing
reaction proceeded, the intermediate gradually transformed to at the wavenumbers lower than 800 cm−1 are due to

4
Nanotechnology 22 (2011) 125703 J Zhang et al

Figure 4. FESEM images of the intermediate samples during the formation of MnO2 –A (a1)–(a3), MnO2 –N (b1)–(b3), and
MnO2 –B (c1)–(c3). The intermediates were collected after 1 h (a1), (b1), (c1), 4 h (a2), (b2), (c2), and 8 h (a3), (b3), (c3). All scale bars are
400 nm.

the characteristic Mn–O stretching vibrations in manganese


oxides [28]. For MnO2 –B and MnO2 –N, the peaks centered
at about 3400 cm−1 and about 1633 cm−1 are attributed to
the O–H stretching vibration and O–H bending vibration of
adsorbed water molecules, respectively [21]. The FT-IR results
indicate the presence of absorbed water within the structures
of as-prepared samples. As for MnO2 –A only broad humps
related to adsorbed water are seen (figure 5(c)), indicating the
decrease in the water content in MnO2 –A.
Figure 6 shows the TGA and differentiated TGA (DrTGA)
curves of the samples. Three major weight loss events can
be seen from the DrTGA curves. The first weight loss event
below 250 ◦ C originated from the physically adsorbed water
while the subsequent weight loss below 500 ◦ C arose from the
removal of structural water [29], confirming that the samples
were hydrated. The last weight loss event at about 550 ◦ C Figure 5. FT-IR spectra of MnO2 –B (a), MnO2 –N (b), and
was due to the transformation of manganese dioxide (MnO2 ) MnO2 –A (c).
to manganese sesquioxide (Mn2 O3 ), releasing oxygen [30].
In addition, the weight loss below 500 ◦ C for sample MnO2 –
3.2. Capacitive performances of hydrous manganese dioxide
A (∼5 wt%) is lower than that of MnO2 –B and MnO2 –N
nanostructures
(∼8 wt%), indicating a decrease in the content of water in
MnO2 –A. The TGA results are in good agreement with the FT- The capacitive performances of MnO2 –B, MnO2 –N, and
IR data. MnO2 –A were characterized using cyclic voltammetry. Cyclic

5
Nanotechnology 22 (2011) 125703 J Zhang et al

Figure 6. TGA (a) and DrTGA (b) curves of samples MnO2 –B,
MnO2 –N, and MnO2 –A. Figure 7. (a) CVs of MnO2 –B in 0.5 M Na2 SO4 at different scan
rates. (b) Dependence of anodic and cathodic currents on scan rate
for MnO2 –B, MnO2 –N, and MnO2 –A, respectively.
voltammograms (CVs) of the MnO2 –B electrode (figure 7(a))
were measured in 0.5 M Na2 SO4 electrolyte at scan rates
of 100, 50, 25, 10, 5, and 2 mV s−1 , respectively. For MnO2 –B possesses the highest area under the CV curves, it
comparison, the CV curves of MnO2 –N and MnO2 –A are indicates the highest capacitance. The CVs indicated that
charge storage is strongly dependent on the morphology of
shown in figure S3 (available at stacks.iop.org/Nano/22/
manganese oxide nanostructures.
125703/mmedia). The CV curves show a nearly rectangular
The capacitive performance was also investigated using
shape, indicating good electrochemical capacitive behavior.
the charge/discharge technique, and the results are shown in
No redox peaks can be seen on the CV curves at different scan
figure 9. It is seen that the charge/discharge profiles of all
rates, indicating that the charge–discharge process of the active
samples exhibit an isosceles triangle-like behavior at a current
materials occurred at a pseudo-constant rate over the whole
density of 1000 mA g−1 , indicating a good reversibility. The
potential window [15]. The current of MnO2 –B, MnO2 –N,
specific capacitance (C, F g−1 ) can be calculated from the
and MnO2 –A at 0.5 V in response to the scan rate is shown
charge–discharge curves using the following equation [11]:
in figure 7(b). The good linear relationship confirms the good
capacitive behavior, because the capacitive current is linear to i t
C= (2)
the scan rate under cycle voltammetry [31]. mE
The CV curves of MnO2 –B, MnO2 –N, and MnO2 –A in where i in A is the charge–discharge current, t in s is the
0.5 M Na2 SO4 electrolyte at scan rates of 10 and 100 mV s−1 charge or discharge time, m in g is the mass of the active
are shown in figure 8. At a scan rate of 10 mV s−1 , the electrode material, and E in V is the voltage range of charge
CV profiles (figure 8(a)) of MnO2 –B, MnO2 –N, and MnO2 – or discharge. The specific capacitances of MnO2 –B, MnO2 –
A are similar rectangular shapes, showing good capacitive N, and MnO2 –A at the current density of 1000 mA g−1 were
behavior at a low scan rate. At an increased scan rate of 199, 139, and 121 F g−1 , respectively. Table 1 summarizes
100 mV s−1 , the rectangular shape of voltammograms of the specific capacitances at different current densities. It is
all samples is maintained with only slight distortion at the seen that the specific capacitance in 0.5 M Na2 SO4 electrolyte
switching potential, which reveals that samples possess good follows MnO2 -B > MnO2 -N > MnO2 –A at a given current
capacitive behavior at a higher scan rate. It can be noted that density. MnO2 –B shows the highest specific capacitance
the area under the CV curves represents the total amount of (262 F g−1 ) at a current density of 250 mA g−1 . The
stored charge, because the specific capacitance C (F g−1 ) can specific capacitance is larger than that reported in the literature,
be calculated using C = Q/(E × m) (where Q is the charge including flowerlike MnO2 (121.5 F g−1 at a current density of
in C , E is the potential window in V, and m is the mass of 1000 mA g−1 ) [21], α -MnO2 (the maximum capacitance with
active material in g) [8]. As a result, a large area of an electrode a value of 152 F g−1 at a current density of 500 mA g−1 ) [23],
generates a high Q , thus a large capacitance at a given and MnO2 nanostructures (205 F g−1 at a current density of
potential window and the same mass of electrode material. As 200 mA g−1 ) [32].

6
Nanotechnology 22 (2011) 125703 J Zhang et al

Figure 8. CVs of MnO2 –B, MnO2 –N, and MnO2 –A in 0.5 M


Na2 SO4 at scan rates of 10 mV s−1 (a) and 100 mV s−1 (b).

Figure 10. N2 adsorption–desorption isotherms (a) and the


corresponding pore size distributions (b) of MnO2 –B, MnO2 –N, and
MnO2 –A. The inset of (a) shows the enlarged N2
adsorption–desorption isotherm of MnO2 –A.

Table 1. Specific capacitance of MnO2 –B, MnO2 –N, and MnO2 –A


in 0.5 M Na2 SO4 at different current densities.
Specific capacitance (F g−1 )
5000 2000 1000 500 250
Sample mA g−1 mA g−1 mA g−1 mA g−1 mA g−1
MnO2 –B 167 176 199 234 262
MnO2 –N 119 124 139 160 193
MnO2 –A 100 110 121 140 161
Figure 9. Charge–discharge curves of MnO2 –B, MnO2 –N, and
MnO2 –A in 0.5 M Na2 SO4 at a current density of 1000 mA g−1 .
MnO2 –A. Indeed, the BJH pore size distribution curves derived
from the desorption branches (figure 10(b)) reveal the presence
Figure 10 shows the N2 adsorption–desorption isotherms of bimodal mesopores for MnO2 –B. However, the bimodal
and BJH pore size distribution curves. MnO2 –B and MnO2 – distribution is not obvious for MnO2 –N. For sample MnO2 –A,
N (figure 10(a)) exhibit a similar isotherm with a hysteresis only pores with average diameter of about 3.8 nm can be seen.
loop in the relative pressure range of 0.7 < P/P0 < 1.0, Based on the FESEM and TEM observations, the small pores
indicating the existence of mesopores [15, 21]. In contrast, can be attributed to the interstitial space from the interleaving
the N2 adsorption–desorption isotherms of MnO2 –A (inset of nanowires or nanorods.
figure 10(a)) exhibit a hysteresis loop in a broader relative Table 2 reports the specific surface area, total pore volume,
pressure range of 0.45 < P/P0 < 1.0. The abrupt and average particle size of MnO2 –B, MnO2 –N, and MnO2 –
nitrogen uptake occurring at 0.9 < P/P0 suggests the A. The BET surface area (181 m2 g−1 ) and total pore volume
presence of an interstitial space between the nanostructures (0.86 cm3 g−1 ) of MnO2 –B are larger than that of both
(e.g. nanowires, nanorods) of MnO2 –B, MnO2 –N, and MnO2 –N and MnO2 –A. The high surface area and large

7
Nanotechnology 22 (2011) 125703 J Zhang et al

Figure 11. Cycle life of MnO2 –B, MnO2 –N, and MnO2 –A in 0.5 M
Na2 SO4 at a current density of 5000 mA g−1 between 0 and 0.9 V.

Table 2. Specific surface area, total pore volume, and average


particle size of MnO2 –B, MnO2 –N, and MnO2 –A.
BET surface Pore volume Average particle
Sample area (m2 g−1 ) (cm3 g−1 ) size (nm)
Figure 12. Core-level XPS curves of O 1s (a) and Mn 2p (b) for
MnO2 –B 181 0.86 33 MnO2 –B and MnO2 –A before and after 1200 cycles.
MnO2 –N 99 0.52 61
MnO2 –A 31 0.08 192
also showed a decrease in the intensity. These observations
indicate the partial fracture of Mn–O–Mn band, suggesting an
pore volume of MnO2 –B are essential for providing a high
irreversible transformation from MnO2 to MnOOH.
capacitance. Besides, an appropriate content of adsorbed
The changes in the XPS signals of O 1s and Mn 2p can
water is favorable for enhancing the diffusion of ions within
be understood on the basis of the charge storage mechanism
the active material [5, 33]. The present results indicate
of MnO2 . Here, two mechanisms are outlined to help us
that hydrous MnO2 with a high surface area is favorable for
understand charge storage in MnO2 -based electrodes. For
achieving a good capacitive performance.
crystalline MnO2 , the dominant mechanism is redox reactions
To evaluate the long-time cycling stability of MnO2 –
between manganese (III) and manganese (IV) involving
B, MnO2 –N, and MnO2 –A, the specific capacitances were
insertion and extraction of protons or alkali cations in the bulk
measured by charge–discharge experiment at a current density
of the active material according to the following equations [35]:
of 5000 mA g−1 . The relationship of capacitance retention to
the cycle number is shown in figure 11. After 1200 cycles, MnO2 + H+ + e− ⇔ MnOOH (3)
the capacitance retention of MnO2 –B is about 75%, indicating
a good stability. MnO2 –N exhibits a comparable capacitance MnO2 + M+ + e− ⇔ MnOOM. (4)
retention of about 71%. As for MnO2 –A, the capacitance
As for amorphous MnO2 , the mechanism is mainly due
retention decreased to about 56% in the first 300 cycles. Then,
to a surface faradic reaction involving surface adsorp-
the capacitance is almost constant with a retention of about
tion/desorption of electrolyte cations on the MnO2 elec-
55% after 1200 cycles.
trode [29]:
XPS was employed to characterize MnO2 –B and MnO2 –
A before and after a stability test. The core-level XPS signals (MnO2 )surface + M+ + e− ⇔ (MnO− +
2 M )surface (5)
of O 1s and Mn 2p are shown in figure 12. The O 1s
spectra were deconvoluted into oxide (Mn–O–Mn), hydroxide where M+ = H+ , Na+ , K+ , Li+ .
(Mn–O–H), and water (H–O–H) [34]. Table 3 summarizes For MnO2 –A with a good crystallinity it is believed
the deconvoluted data. For MnO2 –B and MnO2 –A, the peak that the charge storage process was mainly governed by the
centered at about 529.6 eV (figure 11(a)) is ascribed to lattice insertion and extraction of Na+ ions from the electrolyte into
oxygen O2− from Mn–O–Mn. The peaks at about 532 eV the manganese oxide nanorods. Based on the XPS results the
are due to hydroxyl groups in Mn–O–H. Figure 12(b) exhibits content ratio of Mn–O–Mn decreased from 59.8% to 13.6%
Mn 2p1/2 and Mn 2p3/2 signals centered at about 642.0 and whereas that of Mn–OH increased from 27.9% to 52.5% after
653.5 eV [11]. After 1200 cycles, a higher content of 1200 cycles of charge–discharge. The observed changes on
hydroxide oxygen (Mn–OH) and water (H–O–H) and a lower the XPS signals of O 1s can be attributed to the irreversible
content of oxide (Mn–O–Mn) were observed on MnO2 –B and transformation from MnO2 to MnOOH, because the lattice
MnO2 –A. The core-level XPS signals of Mn 2p (figure 12(b)) strain caused by the insertion–extraction of Na+ during the

8
Nanotechnology 22 (2011) 125703 J Zhang et al

Table 3. Core-level XPS analysis of samples MnO2 –B and MnO2 –A before and after the stability test.
O 1s (eV)
Sample Mn 2p3/2 Mn–O–Mn Mn–OH H–O–H
MnO2 –B As-prepared BD (eV)a 642.0 529.6 531.5 533.4
Area % 57.8 36.1 6.1
After 1200 cycles BD (eV)a 642.1 529.7 531.9 533.8
Area % 41.4 38.0 20.6
MnO2 –A As-prepared BD (eV)a 641.8 529.6 531.3 533.2
Area % 59.8 27.9 12.3
After 1200 cycles BD (eV)a 641.9 529.8 531.9 533.2
Area % 13.2 52.5 34.3
a
Binding energies of peak position.

charge–discharge process would result in the fracture of the [10] Lee H Y, Kim S W and Lee H Y 2001 Electrochem. Solid-State
Mn–O–Mn band, eventually resulting in the poor cycling Lett. 4 A19–22
stability of MnO2 –A [36]. In contrast, for MnO2 –B the XPS [11] Subramanian V, Zhu H, Vajtai R, Ajayan P M and Wei B 2005
J. Phys. Chem. B 109 20207–14
data exhibited a small increase in the hydroxide (36.1–38.0%) [12] Rajendra Prasad K and Miura N 2004 Electrochem. Commun.
and a small decrease (57.8–41.4%) in oxide, suggesting a weak 6 1004–8
influence from the strain effect. The poor crystallinity and the [13] Zhang J T, Ma J Z, Jiang J W and Zhao X S 2010 J. Mater. Res.
small particle size of MnO2 –B favored release of the lattice 25 1476–84
strain caused by the insertion–extraction of Na+ , leading to the [14] Wang X and Li Y D 2003 Chem.—Eur. J. 9 300–6
[15] Xu M, Kong L, Zhou W and Li H 2007 J. Phys. Chem. C
good capacitance retention. 111 19141–7
[16] House D A 1962 Chem. Rev. 62 185–203
4. Conclusions [17] Dekker A O, Levy H A and Yost D M 1937 J. Am. Chem. Soc.
59 2129–32
[18] Li Z, Ding Y, Xiong Y, Yang Q and Xie Y 2005 Chem.
Hydrous manganese dioxide nanostructures prepared under Commun. 918–20
different pH conditions exhibited different physical and [19] Chou S-L, Wang J-Z, Chew S-Y, Liu H-K and Dou S-X 2008
chemical properties. It was found that the specific capacitance Electrochem. Commun. 10 1724–7
of the sample prepared under basic conditions exhibited the [20] Cheng F Y, Chen J, Gou X L and Shen P W 2005 Adv. Mater.
best capacitive performance including the highest capacitance 17 2753–6
[21] Ni J, Lu W, Zhang L, Yue B, Shang X and Lv Y 2009 J. Phys.
of 262 F g−1 at a current density of 250 mA g−1 and the best Chem. C 113 54–60
stability (a capacitance retention of 75% after 1200 cycles). [22] Li Z, Ding Y, Xiong Y and Xie Y 2005 Cryst. Growth Des.
The high surface area and hydrous nature are attributed to 5 1953–8
the highest specific capacitance of the sample. The long-time [23] Tang N, Tian X, Yang C and Pi Z 2009 Mater. Res. Bull.
cycling stability is dependent on the crystalline nature of MnO2 44 2062–7
[24] Xiao W, Xia H, Fuh J Y H and Lu L 2009 J. Power Sources
and the charge storage process. 193 935–8
[25] Wang X, Li Y, Jiao X and Chen D 2002 J. Am. Chem. Soc.
Acknowledgment 124 2880–1
[26] Huang X, Lv D, Yue H, Atta A and Yang Y 2008
Nanotechnology 19 225606
Financial support from Ministry of Education with a project [27] Yu P, Zhang X, Wang D, Wang L and Ma Y 2009 Cryst.
number of MOE2008-T2-1-004 is greatly appreciated. Growth Des. 9 528–33
[28] Sui N, Duan Y, Jiao X and Chen D 2009 J. Phys. Chem. C
113 8560–5
References [29] Devaraj S and Munichandraiah N 2008 J. Phys.Chem. C
112 4406–17
[1] Long J W, Dunn B, Rolison D R and White H S 2004 Chem. [30] Wang L, Tang F, Ozawa K, Chen Z-G, Mukherj A, Zhu Y,
Rev. 104 4463–92 Zou J, Cheng H-M and Lu G Q M 2009 Angew. Chem. Int.
[2] Zhang L L and Zhao X S 2009 Chem. Soc. Rev. 38 2520–31 Edn 48 7048–51
[3] Qu D and Shi H 1998 J. Power Sources 74 99–107 [31] Liu T C, Pell W G, Conway B E and Roberson S L 1998
[4] Zhang Y, Feng H, Wu X, Wang L, Zhang A, Xia T, Dong H, J. Electrochem. Soc. 145 1882–8
Li X and Zhang L 2009 Int. J. Hydrog. Energy 34 4889–99 [32] Subramanian V, Zhu H W and Wei B Q 2008 Chem. Phys. Lett.
[5] Zheng J P, Cygan P J and Jow T R 1995 J. Electrochem. Soc. 453 242–9
142 2699–703 [33] McKeown D A, Hagans P L, Carette L P L, Russell A E,
[6] Zhang J T, Ma J Z, Zhang L L, Guo P Z, Jiang J W and Swider K E and Rolison D R 1999 J. Phys. Chem. B
Zhao X S 2010 J. Phys. Chem. C 114 13608–13 103 4825–32
[7] Simon P and Gogotsi Y 2008 Nat. Mater. 7 845–54 [34] Wei W, Cui X, Chen W and Ivey D G 2009 J. Power Sources
[8] Toupin M, Brousse T and Belanger D 2004 Chem. Mater. 186 543–50
16 3184–90 [35] Athouël L, Moser F, Dugas R, Crosnier O, Bélanger D and
[9] Toupin M, Brousse T and Belanger D 2002 Chem. Mater. Brousse T 2008 J. Phys. Chem. C 112 7270–7
14 3946–52 [36] Lin C-C and Chen H-W 2009 Electrochim. Acta 54 3073–7

You might also like