You are on page 1of 9

Electrochimica Acta 52 (2007) 28062814

Capacitive behavior of nanostructured MnO2 prepared by


sonochemistry method
Alireza Zolfaghari a, , Fatemeh Ataherian b , Mehdi Ghaemi b , Ameneh Gholami b
aChemistry and Chemical Engineering Research Center of Iran, Iran
b Department of Chemistry, Science Faculty, Tarbiat Modares University, Tehran, Iran
Received 20 May 2006; received in revised form 30 September 2006; accepted 3 October 2006
Available online 22 November 2006

Dedicated to the memory of Professor B.E. Conway

Abstract
Nanostructured manganese dioxide has been successfully prepared by a sonochemical method from an aqueous solution of potassium bromate
and manganese sulfate. Changing the proportions of reagents leads either to - or layered structures of MnO2 . The capacitive characteristics of
the samples were systematically investigated in aqueous electrolytes through means of cyclic voltammetry and chronopotentiometry methods.
The electrochemical properties of MnO2 were strongly affected by the pH of electrolyte employed and this material exhibited ideally capacitive
behavior in 0.5 M aqueous Na2 SO4 solution. A maximum specific capacitance of 344 F g1 was obtained for the layered structure determined via
cyclic voltammetry at a scan rate of 5 mV s1 in 0.5 M aqueous Na2 SO4 solution at pH 3.3. Excellent electrochemical reversibility of the materials
was also demonstrated.
Layered structure MnO2 showed higher energy density at high power density than the -structure material.
Impedance spectroscopy studies revealed that charge transfer resistance of the -structure oxide has higher value than that of the layered structure.
2006 Elsevier Ltd. All rights reserved.

Keywords: Manganese dioxide; Supercapacitor; Sonochemistry; Nanostructure; -Structure; Layered structure

1. Introduction The majority of pseudocapacitors fall into two main sub-


classes: (i) metal oxides in aqueous electrolyte [3,4] and (ii)
Supercapacitors are charge-storage devices that possess high conducting polymers that work both in aqueous and non-
ideally power density, exhibit excellent reversibility, and have aqueous electrolytes [5,6].
very long cycle life. They are applicable in electronic devices, Most of work on supercapacitors utilizing pseudocapaci-
such as power electronics, as well as in hybrid electric vehicles tance has focused on metal oxide materials. Amorphous hydrous
and space flight technology [1,2]. ruthenium oxide is the most promising material for super-
Electrochemical supercapacitors can be classified based upon capacitors because of its high specific capacitance, excellent
the mechanism of charge storage: (i) electrical double layer reversibility and long cycle-life [3,7].
capacitor (EDLC) with capacitance mainly arising from the Ruthenium oxide is an expensive material; it is toxic and
charge separation at electrode/electrolyte interface and (ii) pseu- requires the use of a strong acidic electrolyte such as sulfu-
docapacitors with capacitance arising from faradaic reactions ric acid. The acidic media can dissolve the metal oxide over
which, for a variety of different reasons, exhibit capacitative extended cycling leading to fade in the specific capacitance with
behaviour [1]. cycle-life [3]. Of course, the requirement of concentrated acid
does not render the RuO2 technology obsolete, as the success
and wide spread use of Pb-acid batteries illustrates, however,
a low cost technology employing non corrosive electrolytes
Corresponding author. Tel.: +98 21 44580720; fax: +98 21 44580762.
E-mail addresses: azhca@yahoo.com, would certainly find numerous applications. As a result, numer-
zolfaghari@ccerci.ac.ir (A. Zolfaghari). ous metal oxides have also been tested as possible candidates

0013-4686/$ see front matter 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2006.10.035
A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814 2807

for electrochemical supercapacitor devices. Candidate systems Table 1


include IrO2 [8] or CoOx , [9] but they suffer from limitations The synthesis conditions of the various samples
similar to RuO2 , that is, they are expensive and require strong Sample [KMnO4 ] (M) [KBrO3 ] (M) Duration time (min)
acidic or alkaline electrolyte. In addition the potential window S1 0.125 0.250 566
over which they operate reversibly is significantly smaller than S2 0.250 0.500 240
that for RuO2 . On the other hand, MoO3 [10], NiO [3], V2 O5
[11] and MnO2 [1222] systems seem promising due primarily
to their lower cost. A significant body of work has already effects of cavitation produce both strong reductive and oxida-
been published studying the suitability of MnO2 compounds tive radicals which are capable of causing secondary oxidation
for battery applications [23,24], but up to now only few papers or reduction reactions [3638].
dealt with its use in electrochemical supercapacitors [1222]. We report here a sonochemical method for synthesis of
It has already been shown that the nanostructure of the nanostructure manganese dioxide by oxidation of Mn2+ and the
electrode affects dischargecharge properties of the pseudoca- characterization of the oxide as an active material for superca-
pacitor, particularly at large current density [25]. Preparation pacitor electrodes.
of the electrode from nanostructured MnO2 results in elec-
trodes having large specific surface area and enlarges the contact 2. Experimental
area between active and conducting material [14,25,26]. Various
nanostructured composite electrodes of MnO2 were synthesized 2.1. Synthesis
via different methods [2022].
Needle-like manganese oxide was prepared by electrochem- Aqueous solutions of KBrO3 (Merck, research grade) and
ical deposition on the surface of a graphite disc electrode in the MnSO4 (Merck, research grade) in various concentrations were
acidic solution of KMnO4 . The coating exhibited a capacitance irradiated by ultrasound for different time intervals. A 24 kHz
of 45 mF cm2 [20]. ultrasound horn transducer system (UP 200 H, Dr. Hielscher
A manganese dioxide nanowire electrode was made by GmbH) with a 3 mm microtip (titanium alloy) and a sonic power
immersing an anodic aluminum oxide template into a solgel density of 390 W/cm2 was employed. The sonic horn was dipped
solution containing Mn(CH3 COO)2 and citric acid. The result- about 2 cm into the solution. Temperature was held constant
ing electrode was found to have the specific capacity of at 45 C during sonication. All the solutions were made from
165 F g1 [21]. doubly distilled water. Results are reported for two different
Electrochemical deposition of nano-sized manganese oxide synthesis conditions, resulting in significantly different MnO2
onto stainless-steel current collectors in an acidic solution of morphologies, S1 and S2. The ratio of KMnO4 : KBrO3 was,
MnSO4 was also reported. The specific capacitance value of the however, maintained constant. The synthesis conditions are pre-
electrode was 410 F g1 [22]. sented in Table 1. After the reaction was complete, the resulting
However, the use of templates and substrates is undesirable as black solid product was filtered, washed with distilled water to
it introduces heterogeneous impurities and increases the produc- remove by-products, and dried at room temperature in air and
tion cost. Thus research on easily controlled and template-free subsequently characterized as described below.
methods to create nanostructured material is of great significance
and is the focus of the present work. 2.2. Characterization methods
The preparation of -MnO2 3D urchin like nanostructures
by heating aqueous solution of KBrO3 and MnSO4 at 130 C 2.2.1. Chemical composition analysis
for 16 h in an autoclave has been reported [27]. The authors The value of n in Kz MnOn was determined using the poten-
suggest that the high specific area of the resulting material make tiometric titration approach of Vetter and Jaeger [39]. The total
it of interest for Li battery applications but do not report any amount of K in the powders was determined by flame photome-
electrochemical characterization. tery (Schrown 410) and was found to be negligible.
We examine the possibility of shortening the reaction time, The surface water content of each sample was determined by
as well as lowering the reaction temperature by application of weighing the sample before and after heating at 110 C for 2 h.
ultrasonic irradiation to the above process. The structural water content was determined by measuring the
Sonochemistry methods have been successfully applied to the weight loss of the same samples after heating at 400 C for 2 h
preparation of various nanostructured materials including met- [40].
als [2830], metal carbides [31], and metal oxides [3235] at
room temperature, ambient pressure, and short reaction time. 2.2.2. Structural and morphological characterization
Sonochemistry derives its success in creating nanostructured The crystallographic structure of the powder was investigated
materials principally from acoustic cavitation; the formation, using a diffractometer (Philips PC-APD with a Cu anode that
growth, and implosive collapse of bubbles in a liquid. Bubble generated Cu K radiation, = 1.5406 A).
collapse induced by cavitation produces intense local heating, Surface morphology of the sample was examined with a scan-
high pressures, and very short lifetimes. These hot spots have ning electron microscope, SEM (Philips XL30) and an atomic
temperature of roughly 5000 C, pressure of about 500 atm, and force microscope, AFM (DME DS 95) in AC mode. The synthe-
heating and cooling rates greater than 109 K s1 . The chemical sized MnO2 powders were dispersed in doubly distilled water
2808 A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814

(0.25%). A few drops of the suspension were spread on glass compared to the charge of electrode with active material
lamella, and dried at 70 C. The AFM images were obtained by (>101 C g1 ).
scanning an area of 1 m2 of the prepared sample with the scan The average specific capacitance was calculated by integrat-
rate of 1.5 m s1 in ambient air at room temperature. ing either the oxidative or the reductive CV half-cycle over the
The specific surface area and pore size distribution were potential window of the voltammogram. All the specific capac-
determined by BrunauerEmmettTeller (BET) [41] analysis of itances are reported per gram of active material.
N2 adsorptiondesorption isotherms measured on sample using
a Quantasorb, Quantachrom apparatus. 3. Results and discussion

2.3. Analysis of point of zero surface charge 3.1. Theory and proposed mechanism

The pHp.z.c. values of the samples were determined by poten- When synthesis of MnO2 was executed at room temperature
tiometric titration using the following bath procedure: different without sonication only a small amount of product was obtained,
volumes of 0.1 M HCl or 0.1 M NaOH were added to a series even after 24 h. Fortunately this reaction was facilitated by ultra-
of NaCl solutions of different concentrations (1.0-103 M) such sonic radiation.
that the total volume of each solution was 50 ml. After about The redox synthesis of MnO2 in an aqueous media containing
1 h the pH values were noted (blank solution) using a pH meter Mn2+ ions has been proposed to occur following the reaction
(744 Metrohm). Fifty milligrams of solid sample were added [43]:
to each solution, immediately stopper, and shaken well. After
72 h of equilibration with intermittent shaking, the pH val- Mn2+ + 2H2 O MnO2 + 4H+ + 2e (1)
ues of each of the supernatant liquids were noted. Then the It is unlikely that reaction (1) occurs in a single step. The exis-
change of pH (pH) is plotted against initial pH (blank solu- tence of a precursor, such as Mn3+ has been suggested [4446],
tion). The pHp.z.c. was estimated from where pH is equal to originating from the following reaction:
zero [42].
Mn2+ Mn3+ + e (2)
2.4. Electrochemical measurement
Mn3+ can disproportionate to Mn2+ and Mn4+ at acidic concen-
Electrodes were prepared by mixing 60 wt% of MnO2 trations, according to reaction (3), or it may hydrolyze to form
powder as active material with 10 wt% of acetylene black an insulating intermediate MnOOH, according to reaction (4):
(Alfa Aesar, >99.9%, S.A. 80 m2 /g), 25 wt% of graphite 2Mn3+ + 2H2 O Mn2+ + MnO2 + 4H+ (3)
(Alfa Aesar, conducting grade, 200 mesh), and 5 wt% of
poly(tetrafluoroethylene) dried powder (Merck). The three first Mn3+ + 2H2 O MnOOH + 3H+ (4)
constituents were first mixed together to obtain a homogeneous
black powder. The poly(tetrafluoroethylene) powder was then Subsequently, the product of reaction (4) is oxidized to MnO2 :
added with a few drops of ethanol. This resulted in a rubber-like MnOOH MnO2 + H+ + e (5)
paste that was rolled into a film (100200 m thick) on a flat
glass surface. A piece of film (0.25 cm2 ) was cut and pressed The final product contains Mn3+ as MnOOK or MnOOH and
onto stainless steel grid current collector. Mn4+ as MnO2 .
An Autolab, Eco Chemie. B.V. potentiostat/galvanostat was
used for all electrochemical measurements. A beaker type 3.2. Material characteristics
electrochemical cell equipped with a MnO2 based working elec-
trode, an Ag/AgCl electrode (Metrohm AG 9101 Herisau, 3 M 3.2.1. Chemical composition
KCl, 0.207 V versus SHE at 25 C) reference electrode and a The presently accepted formula for MnOn based on
1 cm2 stainless steel grid counter electrode was used. A Lug- the cation vacancy model, proposed by Ruetschi [47] is
gin capillary, whose tip was set at a distance of 12 mm from Mn1xy 4+ Mny 3+ On4xy 2 OH4x+y , where x represents the
the surface of the working electrode, was used to minimize amount of cation vacancy. It has been proposed that manganese
errors due to iR drop in the electrolyte. Aqueous solutions of dioxide contains considerable amounts of hydroxide and water
0.5 M Na2 SO4 with different pH values, achieved by addition of existed mainly in different forms. Hydroxide ions compensate
HCl, employed as the electrolyte, were degassed with purified cation vacancies which may have arisen from Mn4+ ions miss-
argon before measurement. The solution temperature was main- ing during manufacturing of manganese dioxide. It could be
tained at 25 C by means of a water thermostat (RE 104 Ecoline, assumed that four ions of six oxygen ions in octahedron sur-
LAUDA). rounding the vacancy are hydroxide ions. Furthermore some
In order to determine the contribution of acetylene black Mn4+ ions may reduce partially to Mn3+ , and hydroxide ions
and graphite to the overall charge capacity of the electrode, a are requisite to compensate the lacking of positive charge due
cyclic voltammetry, CV, experiment was performed on a blank to Mn4+ ions replaced by Mn3+ ions. Finally water molecules
electrode containing no MnO2 . The charge storage capacity could be adsorbed mainly on the surface of the nano-sized pore
of this electrode was found to be negligible (104 C g1 ) throughout the microporous material [48]. The amount of n
A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814 2809

Fig. 1. XRD patterns of prepared samples: (a) S1 and (b) S2. The peaks are
noted with ().

could be evaluated through the potentiometric titration [39] and


in terms of n, there are (2n 3) Mn4+ ions and (4 2n) Mn3+
ions [40].
The water contents and chemical formula of prepared samples
are presented in Table 2. The amount of Mn3+ and water content
of S2 were found to be higher than those of S1.

3.2.2. Bulk and morphological properties


Fig. 1 shows X-ray diffraction (XRD) pattern of the prepared
samples. The diffraction peaks of S1 in Fig. 1a correspond to the
(1 2 0), (1 3 1), (2 3 0), (3 0 0), (0 0 2), (1 6 0), (2 4 2), (4 2 1) and
(0 0 3) planes, from left to right, respectively. All these peaks
can be indexed to pure orthorhombic -MnO2 (JCPDS card
14644, a = 6.36 A, b = 10.15 A, c = 4.09 A). No characteristic
peaks were observed for other impurities such as - or -MnO2 ,
Mn(OH)2 and Mn2 O3 . In Fig. 1b, the pattern is dominated by
two major peaks (e.g. at about 12 and 22 2) corresponding Fig. 2. AFM topography image: (a) S1 and (b) S2.
to (0 0 2) and (2 1 2) planes, respectively which are signature
features of synthetic birnessites material [49], and indicates that
there is layered structure in S2. It seems that the longer reac- than 90% of the population is within the 2050 and 8090 nm
tion time required producing S1-type MnO2 results in a more range for S1 and S2, respectively. These results are in agreement
crystallized structure. with the SEM images of S1 and S2 (Fig. 3).
AFM images show MnO2 powder samples to exhibit spher- The specific surface area and the pore size distribution of
ical nanostructure (Fig. 2). Analysis of the AFM images using S1 and S2 from the BET test are presented in Table 2. The
image processing software (DME, version 1.5.0.4) exhibit a volume of micropores compared to the volume of mesopores is
spherical structure with a narrow size distribution in which more negligible for both samples. The microporous volume of S1 is

Table 2
The water content, chemical formula, specific surface area and pore diameter of prepared samples

Sample Surface Structural Mn1xy 4+ Mny 3+ On4xy 2 OH4xy SBET Dav (A) Microporous Mesoporous
water (%) water (%) (m2 g1 ) volume (cm3 g1 ) volume (cm3 g1 )

S1 17.0 6.3 Mn0.838 4+ Mn0.017 3+ O1.403 2 OH0.597 301 46.0 0.02 0.33
S2 16.4 8.2 Mn0.762 4+ Mn0.066 3+ O1.246 2 OH0.754 161 69.1 <0.01 0.29
2810 A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814

Fig. 4. Variation of pH as a function of initial pH for (a) S1 and (b) S2 in ()


0.05 M, (+) 0.1 M, () 0.2 M, and () 0.5 M aqueous NaCl solution.

3.4. Electrochemical properties


Fig. 3. SEM image of (a) S1 and (b) S2 (the X sign shows a single particle with
diameter of 60 and 90 nm, respectively). 3.4.1. Cyclic voltammetry
Typical cyclic voltammograms, CVs, measured in aqueous
0.5 M Na2 SO4 solution for prepared samples are shown in Fig. 5.
slightly higher than that of S2 (0.04 cm3 g1 ), but the average The potential ranges were 0.00.8, 0.01.0, and 0.21.2 V ver-
pore diameter of S2 is bigger than that of S1. The lower pore sus Ag/AgCl and each measurement was taken at a sweep rate
diameter of S1 leads to the augmentation of its BET surface of 5 mV s1 .
area. These CVs are relatively rectangular in shape and exhibit
near mirror-image current response on voltage reversal thus
indicating ideal capacitive behavior for both samples.
3.3. Characteristics of point of zero surface charge It may be concluded from a comparison of these curves that
the shape of CV is not significantly affected by changing the
The effect of pH on the charge of manganese oxide surface chemical composition, structural phases, specific surface area
may be represented by [50]: and pore diameter of prepared samples.
The specific capacitance (SC) of the electroactive material
can be estimated based on the following equation [51]:
MnOH2(surf) + Mn(OH)(surf) + H+ MnO(surf) + 2Haq +
positive surface p.z.c. negative surface voltammetric charge q
(6) C= = (7)
potential window oxide loading V w
The change in pH following addition of 50 mg of sample was The specific capacitances of the layered structure manganese
monitored as a function of pH for four different NaCl solutions. dioxide, S2, and -MnO2 , S1, determined from calculation of
pH control was achieved by the addition of either HCl or NaOH. the charge from the CVs in Fig. 6 are ca. 344 and 275 F g1
The point of zero charge, p.z.c., is expressed in terms of the in the potential window of 0.01.0 V at sweep rate 5 mV s1 ,
pH at which there is no net H+ or OH adsorption, i.e. where respectively.
pH following the addition of the sample is 0. Based on Fig. 4, The capacity measurements along with the BET results show
the pHp.z.c. s of S1 and S2 were found to be 2.21 and 1.89, that the higher surface area of S1 is not the only factor that
respectively. At pH close to the pHp.z.c. the net surface adsorption determines the capacitance. The other important factor is pore
or desorption decreases for both samples. size of electrode material [1]. The average pore diameter of the
A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814 2811

Fig. 6. Variation of the voltammetric charge density (q) with respect to sweep
rate (): (a) extrapolation of q to = from the q vs. 1/2 plot given the outer
Fig. 5. Cyclic voltammograms of (a) S1 and (b) S2 electrodes over different charge and (b) extrapolation of q to = 0 from the q1 vs. 1/2 plot given the
potential ranges. total charge at () pH 3.3 and () pH 6 for S1, () pH 3.3 and () pH 6 for
S2.
S2 material is 50% larger than that of S1. It is possible that the
larger specific capacitance of S2 is due to the fact that the larger
rate 5 mV s1 in 0.5 M aqueous Na2 SO4 solution with different
pores of this material are more readily accessible. In addition, it
pH is listed in Table 3. As shown, the highest specific capacitance
should be noted that the capacitance of this material is the result
belongs to layered structure MnO2 at pH 3.3.
of a redox process and not limited to electrostatic charging of
These results show that the increment of proton concentration
the double layer at the MnO2 electrode surface.
at constant concentration of Na+ cation results in an enhance-
Since, the electrochemical reversibility is one of the
ment of specific capacitance. This effect may originate from
predominant factors influencing the power capabilities of elec-
the penetration of cations into the sample pores that may be
trochemical supercapacitors the electrochemical reversibility of
attributed to high mobility and small size of proton compare to
the prepared samples was examined systematically by changing
Na+ cation. This conclusion has been presented in other pub-
the potential window of CV at 0.20.8, 0.01.0 and 0.21.2 V
lished work [52]. At pH 2.3, the capacitance suddenly decreases
versus Ag/AgCl that is shown in Fig. 5.
compare to higher pHs. This decrease could be resulted from
Rectangular and symmetric iE responses were observed at
the instability of the MnO2 electrode in the strong acidic solu-
S1 and S2 electrodes between 0.2 and 0.8 V indicating that
tion [52] or it could be arisen from approaching of electrolytes
both samples exhibited high electrochemical reversibility in
pH to pHp.z.c. resulting in reduction of surface adsorp-
this potential range and sweep-rate. As the potential range
tion. Therefore the capacitance measurement was limited to
over which the CV experiment was conducted was increased
pH > 3.
from (0.20.8 V) to (0.01.0 V), an irreversible anodic peak
developed between 0.8 and 1.0 V for both samples. No cor-
responding cathodic response is appeared. As the potential is Table 3
further expanded, a quasi-reversible faradaic process is observed The specific capacitance of the various samples in 0.5 M aqueous Na2 SO4
solution with different pH at scan rate 5 mV s1
just positive of 1.0 V. Charges and discharges of both electrodes
are not electrochemically reversible over the extended potential Sample Specific capacitance (F g1 )
range. pH = 2.3 pH = 3.3 pH = 3.7 pH = 4.5 pH = 6

S1 118 258 236 212 204


3.4.1.1. Effect of electrolytes pH. The specific capacitance of S2 171 344 316 262 229
the various samples in the potential window of 0.01.0 V at scan
2812 A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814

Table 4
The estimated charge density (C g1 ) at very low and very high sweep rate in
the various electrolytes pH for different samples
Sample pH = 3.3 pH = 6

q0 qT q0 qT

S1 122 213 102 179


S2 152 357 84 278

3.4.1.2. Effect of scan rate. Voltammetric charges were deter-


mined as a function of sweep-rate for electrodes of S1 and S2
materials, and the dependence of q on 1/2 for S1 and S2 are
shown in Fig. 6a. Voltammetric charge, q, increases with 1/2 Fig. 7. Constant current charge/discharge curves for (a) S1 and (b) S2 electrodes
(decreases with ). In Fig. 6b, the inverse of the voltammetric measured at 2.5 A g1 in 0.5 M aqueous Na2 SO4 solution at pH 3.3 between 0
charge, q1 , is plotted as a function of the square root of sweep and 1 V.
rate, 1/2 .
At high sweep rate, the diffusion of protons and sodium Ragone plots of different electrodes at various charge/
cations is limited to the outer surface of electrode that is more discharge current densities are shown in Fig. 8. The power den-
accessible. Therefore the extrapolation of q to = from the sity (p.d.) and the energy density (e.d.) of the electrodes can be
q versus 1/2 plot (Fig. 6a) gives the outer charge q0 , which estimated from:
is the charge on the active outer surface. The extrapolation of q Q V E
to = 0 from the q1 versus 1/2 plot (Fig. 6b) gives the total P= = (9)
2tw tw
charge qT , that is, the charge involved to the whole active inner
and outer surfaces [53]. where P, Q, V, t, w and E are indicating of the average
The estimated charges at very low and very high sweep rates power density, total charge delivered, the potential window of
at various pH for the different samples are presented in Table 4. discharge, discharge time, loading prepared sample and total
At pH 6, the higher value of q0 for S1 than that of S2 could energy density, respectively. Although energy density decreases
reveal the higher active outer surface of S1. At pH 3.3, q0 of S2 as power density increases for both samples, it declines more
is more than that of S1, which is due to determining effect of rapidly for S1 type electrodes than for S2 type. The fairly high
the high mobility of proton in the charge/discharge process. The energy density of S2 at high power density provides one of
comparison of qT for both samples at particular pH indicates the basic requirements of a successful electrode material for
that the accessible surface sites of S2 are much more than those supercapacitors applications [2].
of S1. At lower pH, the higher values of qT for both samples
show that the access to surface sites is easier for proton than 3.4.3. Impedance spectroscopy behaviors
sodium cations. Since, this augmentation for S2 is about twice Electrochemical impedance spectroscopy, a powerful tech-
as high as that of S1, and S2 with larger pore diameter reveals nique for the investigation of the capacitive behavior of
more specific capacitance, the charge storage could be a cation electrochemical cells, has been also used to characterize our
diffusion driven process. materials. Fig. 9 presents Nyquist plots obtained at electrodes
made from S1 and S2 type MnO2 at 0 V versus Ag/AgCl over
3.4.2. Charge and discharge properties of electrodes the frequency range 0.01105 Hz. Both of the plots are similar
To investigate the performance of the composite electrode, in form, composed of a semicircle at high and intermediated
both electrodes were charged and discharged at various current
densities. Fig. 7 shows the chronopotentiogram of S1 and S2
electrodes. The cutoff voltages for charging/discharging were
0.0 and 1.0 V at current density of 2.5 A g1 . A good linear varia-
tion of potential versus time was observed for both curves, which
is another typical characteristic of an ideal capacitor. A small iR
drop is observed, indicating a conductive characteristic of the
oxide material [22]. The specific capacitance was obtained from:
I t
C= (8)
V
S2 electrodes show the largest specific capacitance, 295 F g1
when the current density is 1 A g1 . Also, there is a less of a
decrease in specific capacitance with increasing current density
for S2 electrodes than for S1 electrodes. This indicates the
higher accessible surface area in S2. Fig. 8. Ragone plots obtained for () S1 and () S2 electrodes.
A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814 2813

BET measurements have shown that the larger capacitance of


S2 than that of S1 could be explained by an increase in average
pore size of S2 compared to S1 that was thought to facilitate
electrolytes ion accessibility on the electrode surface. The elec-
trochemical impedance spectroscopy also tends to confirm the
above hypothesis.
In addition, the nanostructure samples demonstrate excellent
reversibility and high power density.

Acknowledgements

The financial support of National Committee of Nanotech-


nology in Ministry of Science, Research and Technology of Iran
Fig. 9. Nyquist plots of the electrodes () S1 and () S2 in 0.5 M aqueous is acknowledged. The authors would like to thank Dr. Wendy G.
Na2 SO4 solution at pH 3.3. Pell for her constructive comment.
frequencies, followed by a relative linear response at the low-
frequencies. At very high frequencies, the measured resistance References
is composed of the following terms: the ionic resistance of elec-
[1] B.E. Conway, Electrochemical Supercapacitors, Kluwer-Plenum, New
trolyte, the intrinsic resistance of the active material, and the
York, 1999.
contact resistance at the active material/current collector inter- [2] A. Burke, J. Power Sources 91 (2000) 37.
face [54]. It can be seen that the above resistance measured at [3] J.P. Zheng, P.J. Cygan, T.R. Jow, J. Electrochem. Soc. 142 (1995) 2699.
S1 type electrodes is almost the same that at S2 type electrodes. [4] K.C. Liu, M.A. Anderson, J. Electrochem. Soc. 143 (1996) 124.
At high frequencies, Fig. 9 shows the presence of a semicir- [5] P. Novak, K. Muller, K.S.V. Santanam, O. Hass, Chem. Rev. 97 (1997) 207.
[6] P. Soudan, P. Lucas, H.A. Ho, D. Jobin, L. Breau, D. Belanger, J. Mater.
cle. Similar features at carbon double layer electrodes have been
Chem. 11 (2001) 773.
described as a pseudo-charge transfer resistance associated with [7] Y.S. Yoon, W.I. Cho, J.H. Lim, D.J. Choi, J. Power Sources 101 (2001)
the active carbon-current collector interface [54]. It seems that 126.
the increasing of the contact surface area between the active car- [8] B.E. Conway, V. Birss, J. Wojtowicz, J. Power Sources 66 (1997) 1.
bon and the S2 as active material lead to a decrease in the contact [9] C. Lin, J.A. Ritter, B.N. Popov, J. Electrochem. Soc. 145 (1998) 4097.
[10] W. Sugimoto, T. Ohnuma, Y. Murakami, Y. Takasu, Electrochem. Solid-
resistance and a minimization of the high frequency semicircle.
State Lett. 4 (2001) A145.
In our experiments, the amplitude of the semicircle changes [11] H.Y. Lee, J.B. Goodenough, J. Solid State Chem. 148 (1999) 81.
with the nature of the electrode active material, and it is smaller [12] H.Y. Lee, J.B. Goodenough, J. Solid State Chem. 144 (1999) 220.
for S2 than S1. The BET study showed that the pore size of S2 [13] S.C. Pang, M.A. Anderson, T.W. Chapman, J. Electrochem. Soc. 147 (2000)
is bigger than that of S1 which improves the electrolyte accessi- 444.
[14] H.Y. Lee, S.W. Kim, H.Y. Lee, Electrochem. Solid-State Lett. 4 (2001)
bility to S2. Therefore the pseudo-charge transfer resistance of
A19.
the S1 is more than that of S2. [15] C.C. Hu, T.W. Tsou, Electrochem. Commun. 4 (2002) 105.
At low frequencies, the imaginary part of the impedance, [16] S.F. Chin, S.C. Pang, M.A. Anderson, J. Electrochem. Soc. 149 (2002)
for both samples, increases. As the redox process giving rise to A379.
the pseudo-capacitative behavior is primarily a surface process, [17] M. Toupin, T. Brousse, D. Belanger, Chem. Mater. 14 (2002) 3946.
[19] M. Toupin, T. Brousse, D. Belanger, Chem. Mater. 16 (2004) 3184.
the low frequency behavior results from penetration of the ac
[20] M. Wu, G.A. Snook, G.Z. Chen, D.J. Fray, Electrochem. Commun. 6 (2004)
signal inside the porous matrix increasing with the decrease of 499.
frequency. [21] X. Wang, X. Wang, W. Huang, P.J. Sebastian, S. Gamboa, J. Power Sources
140 (2005) 211.
4. Conclusion [22] T. Shinomiya, V. Gupta, N. Miura, Electrochim. Acta 51 (2006) 4412.
[23] R. Patrice, B. Gerand, J.B. Leriche, L. Seguin, E. Wang, R. Moses, K.
Brandt, J.M. Tarascon, J. Electrochem. Soc. 148 (2001) A448.
Nanostructured, -manganese dioxide, S1, and layered struc- [24] S. Bach, J.P. Pereira-Ramos, N. Baffier, Solid State Ionics 80 (1995) 151.
ture manganese dioxide, S2, were prepared using ultrasonic [25] M. Hibono, H. Kawaoka, H. Zhou, I. Honma, J. Power Sources 124 (2003)
radiation. Structural selectivity of the synthesis of manganese 143.
dioxide could be controlled by dilution of reactants. AFM and [26] A. Yu, R. Frech, J. Electrochem. Soc. 149 (2002) A99.
[27] C. Wu, Y. Xie, D. Wang, J. Yang, T. Li, J. Phys. Chem. B 107 (2003) 13583.
SEM studies demonstrate the nanostructured electrode material [28] K.S. Suslick, S.B. Choe, A.A. Cichowlas, M.W. Grinstaff, Nature 353
was constructed from spherical particles. (1991) 414.
Maximum specific capacitance values as high as 344 and [29] Yu. Koltypin, G. Katabi, R. Prozorov, A. Gedanken, J. Non-Cryst. Solids
257 F g1 were obtained for the layered and -MnO2 , respec- 201 (1996) 159.
tively at scan rate 5 mV s1 in 0.5 M aqueous Na2 SO4 solution [30] K. Okitsu, Y. Mizukoshi, H. Bandow, Y. Maeda, T. Yamamoto, Y. Nagata,
Ultrason. Sonochem. 3 (1996) 249.
at pH 3.3. These capacitance values compare favorable with that [31] T. Hyeon, M. Fang, K.S. Suslick, J. Am. Chem. Soc. 118 (1996) 5492.
reported in the literature for MnO2 electrodes deposited directly [32] X. Cao, Yu. Koltypin, G. Katabi, I. Felner, A. Gedanken, J. Mater. Res. 12
on templates and substrates [2022]. (1997) 405.
2814 A. Zolfaghari et al. / Electrochimica Acta 52 (2007) 28062814

[33] N. Arul Dhas, A. Gedanken, J. Phys. Chem. B 101 (1997) 9495. [44] J.Ph. Petitpierre, Ch. Comninellis, E. Plattner, Electrochim. Acta 35 (1990)
[34] R. Vijaya Kumar, Y. Diamant, A. Gedanken, Chem. Mater. 12 (2000) 281.
2301. [45] S. Rodriguez, N. Munichndraiah, A.K. Shukla, J. Appl. Electrochem. 28
[35] A. Patra, E. Sominska, S. Ramesh, Yu. Koltypin, Z. Zhong, H. Minti, R. (1998) 1235.
Reisfeld, A. Gedanken, J. Phys. Chem. B 103 (1999) 3361. [46] S. Nijjer, J. Thonstad, G.M. Haarberg, Electrochim. Acta 46 (2000) 395.
[36] T.J. Mason, Sonochemistry, Oxford University Press Inc., New York, 1999. [47] P. Ruetschi, J. Electrochem. Soc. 131 (1984) 2737.
[37] K.S. Suslick (Ed.), Ultrasound: Its Chemical, Physical and Biological [48] S.W. Donne, F.H. Feddrix, R. Glockner, S. Marion, T. Norby, Solid State
Effects, VCH, Weinheim, 1988. Ionics 152/153 (2002) 695.
[38] A. Gedanken, Ultrason. Sonochem. 11 (2004) 47. [49] S. Ching, D.J. Petrovay, M.L. Jorgensen, S.L. Suib, Inorg. Chem. 36 (1997)
[39] K.J. Vetter, N. Jaeger, Electrochim. Acta 11 (1966) 401. 883.
[40] D.K. Walanda, G.A. Lawrance, S.W. Donne, J. Power Sources 139 (2005) [50] M.J. Gray, M.A. Malati, M.W. Rophael, J. Electroanal. Chem. 89 (1978)
325. 135.
[41] S. Brunauer, P.H. Emmett, E. Teller, J. Am. Chem. Soc. 60 (1938) [51] C.C. Hu, C.C. Wang, Electrochem. Commun. 4 (2002) 554.
309. [52] S. Wen, J.W. Lee, I. Yeo, J. Park, S. Mho, Electrochim. Acta 50 (2004)
[42] S.S. Tripathy, S.B. Kanungo, S.K. Mishra, J. Colloid Interf. Sci. 241 (2001) 849.
112. [53] S. Ardizzone, G. Fregonara, S. Trasatti, Electrochim. Acta 35 (1990) 263.
[43] S.V. Kordesh, Batteries(Manganese Dioxide), vol. 1, Marcel Dekker, [54] J. Gamby, P.L. Taberna, P. Simon, J.F. Fauvarque, M. Chesneau, J. Power
NewYork, 1974. Sources 101 (2001) 109.

You might also like