Sei sulla pagina 1di 13

Effect of Non-ionic Surfactants and Its Role in K Intercalation

in Electrolytic Manganese Dioxide


AVIJIT BISWAL, B.C. TRIPATHY, T. SUBBAIAH, D. MEYRICK, MIHAIL IONESCU,
and MANICKAM MINAKSHI

The effect of non-ionic surface active agents (surfactants) Triton X-100 (TX-100) and Tween-20
(Tw-20) and their role in potassium intercalation in electrolytic manganese dioxide (EMD)
produced from manganese cake has been investigated. Electrosynthesis of MnO2 in the absence
or presence of surfactant was carried out from acidic MnSO4 solution obtained from manganese
cake under optimized conditions. A range of characterization techniques, including field
emission scanning electron microscopy, transmission electron microscopy (TEM), Rutherford
back scattering (RBS), and BET surface area/porosity studies, was carried out to determine the
structural and chemical characteristics of the EMD. Galvanostatic (discharge) and potentio-
static (cyclic voltammetric) studies were employed to evaluate the suitability of EMD in com-
bination with KOH electrolyte for alkaline battery applications. The presence of surfactant
played an important role in modifying the physicochemical properties of the EMD by increasing
the surface area of the material and hence, enhancing its electrochemical performance. The
TEM and RBS analyses of the discharged EMD (c-MnO2) material showed clear evidence of
potassium intercalation or at least the formation of a film on the MnO2 surface. The extent of
intercalation was greater for EMD deposited in the presence of TX-100. Discharged MnO2
showed products of Mn2+ intermediates such as MnOOH and Mn3O4.

DOI: 10.1007/s40553-014-0022-9
 ASM International (ASM) and The Minerals, Metals & Materials Society (TMS) 2014

I. INTRODUCTION material derived from EMD. This rechargeable battery


has quickly become the technology of choice in the cordless
THE intense interest in manganese oxides for energy power tool industry and is being increasingly applied in the
applications is driven by the low cost and non-toxicity of electric vehicle (EV) and hybrid electric vehicle (HEV)
these materials when compared with nickel and cobalt industries. The EV and HEV battery markets will likely
oxides. The rich chemistry and diverse crystalline expand rapidly over the next few years, and this will
structures of MnO2 offer a versatility that allows its translate to attractive returns for high-end EMD producers.
use in a range of applications.[1] Of the different types of An estimated two billion manganese versions of lithium
MnO2, such as natural manganese dioxide (NMD), primary (non-rechargeable) batteries and half a billion
chemical manganese dioxide (CMD), activated manga- lithium manganese secondary (rechargeable) batteries are
nese dioxide (AMD), and electrolytic manganese diox- produced per annum, with growth rates in excess of 10 and
ide (EMD), EMD is most widely applied in battery 20 pct, respectively.[9] This increasing demand gives rise to a
industries.[2,3] EMD is composed predominantly of the need for high purity EMD from alternative sources.
c-phase and may be produced cost effectively with low Numerous reports are available regarding the synthesis of
environmental impact. EMD has a high redox potential, EMD (c-MnO2) from various manganese ores[10–14] and its
high rate capability, long storage life, and better role as a cathode in alkaline Zn-MnO2 systems,[3,4,8,15] while
performance over a wide temperature range relative to the possibility of producing EMD from low grade and
its nickel and cobalt counterparts.[4–8] secondary manganese resources has not been fully ex-
It is anticipated that over the next decade, the pricing for plored. Production of EMD from a range of sources will be
high purity EMD will increasingly be affected by the necessary to meet future escalating demand.
demand for lithium-ion batteries containing cathodic In this context, we have produced battery grade
EMD, in the presence of surfactants, from manganese
cake, a side product obtained during the processing of
AVIJIT BISWAL, Ph.D. Scholar, and B.C. TRIPATHY and T. manganese nodules.[16] Surfactants play a significant
SUBBAIAH, Senior Scientists, are with the CSIR - Institute of Minerals role in modifying the growth pattern (and hence the
and Materials Technology, Bhubaneswar 751013, India, and also with
the Academy of Scientific and Innovative Research, Anusandhan physicochemical properties) of electrodeposits through
Bhavan, New Delhi 110001. D. MEYRICK, Senior Lecturer, and adsorption on the MnO2 electrode surface.[17] Surface
MANICKAM MINAKSHI, Lecturer, are with the School of Engineer- adsorption of surfactants influences the kinetics of
ing and Information Technology, Murdoch University, Murdoch, WA electron transfer through blocking of active sites and
6150, Australia. Contact e-mail: minakshi@murdoch.edu.au MIHAIL
IONESCU, Senior Research Scientist, is with the Environment, Lucas
also affects electrostatic interactions between electroac-
Heights Laboratories, ANSTO, Menai, NSW 2234, Australia. tive species in the electrolytic bath. Consequently,
Manuscript submitted March 21, 2014. addition of organic surfactants to the electrolytic bath

METALLURGICAL AND MATERIALS TRANSACTIONS E


affects the morphology and mechanical properties of MnSO4 solution was obtained after leaching of manga-
electrochemically deposited material, leading to altered nese cake, followed by two stages of purification.[16,29]
electrochemical behavior of materials. Use of surfac- All EMD samples were prepared with in situ addition of
tants during preparation of materials for battery appli- a calculated (as per critical micelle concentration)
cations has been widely reported in the literature.[17–23] quantity of surfactant to the electrolytic bath containing
Surfactants may be used during chemical co-precipita- 50 g dm3 MnSO4 and 25 g dm3 H2SO4. The critical
tion,[18,19] liquid co-precipitation,[20] as well as electro- micelle concentrations (CMC) of Triton X-100 (TX-100,
chemical deposition methods.[21,22] Ghaemi et al.[22] SLR) and Tween-20 (Tw-20, SLR) are 0.02 and
have reported improved charge/discharge cycle behavior 0.01 wt pct, respectively. EMD samples prepared from
of EMD prepared in the presence of t-octyl phenoxy 0 t, 5 t, 10 t, 15 t, and 30 t cmc values of TX-100 (where
polyethoxy ethanol (Triton X-100). The absence of this t denotes CMC) have been labeled EMD0, EMDTX5,
surfactant was found to produce EMD unsuitable for a EMDTX10, EMDTX15, and EMDTX30, respectively. Sim-
rechargeable battery application. EMD prepared in the ilarly the EMD produced from the electrolytic bath
presence of cetyltrimethylammonium bromide (CTAB) containing 10, 15 t, and 30 t CMC of Tw-20 has been
or sodium n-dodecylbenzenesulfonate (SDBS) has also labeled EMDTw10, EMDTw15, and EMDTw30, respec-
been reported and found suitable for battery applica- tively. The electro-synthesized EMD was scraped off
tions.[23] The positive effect of use of CTAB and SDBS from the anodic substrate after the deposition and
on the performance of Zn-MnO2 alkaline batteries has washed thoroughly with deionized water before drying
been reported earlier.[23] Triton X-100 is reported[21] to in an oven. The dried mass was ground and sieved
offer the best improvement in capacitance behavior of through a 53-lm mesh to obtain EMD powder. The
EMD, possibly due to strong adsorption of the surfac- resultant product in powder form was washed repeat-
tant, resulting in an increase in specific capacitance of edly with de-ionized water, until the sample was sulfate
the energy storage device. free. The EMD powder was finally dried and cooled in a
The current work evaluates the use of Triton X-100 desiccator and subjected to physical and electrochemical
along with Tween-20 on EMD produced from a characterization.
secondary source. The difference in chemical structure
observed (detailed in results and discussion section) for
B. Structural, Thermal, and Microscopic Analyses
the two non-ionic surfactants (TX100 and Tw20)
intrigued to investigate their role in EMD for energy X-ray diffractograms were recorded for the resultant
storage devices. To the best of our knowledge, no work EMD powders using PANalytical diffractometer (PW
has been reported on the suitability for rechargeable 1830, Philips, Japan) with Cu Ka radiation, k = 1.5418 Å.
battery applications of EMD from manganese cake The scans were recorded in 2h range 15 to 80 deg. BET
prepared in the presence of the above mentioned surface area and pore size analysis were carried out using
surfactants. Most of the reported literature on EMD Quantachrome (Autosorb-iQ) Surface area analyzer. Dif-
material in aqueous solutions is based on the well- ferential thermal analysis (DTA) and thermogravimetry
known protonation or lithiation mechanisms.[24–28] In (TG) (Perkin Elmer Diamond) of the EMD samples were
this work, TEM imaging and Rutherford back scatter- carried out under an inert atmosphere over a temperature
ing analyses gave intriguing insights suggesting that the range between 303 K and 1273 K (30 C and 1000 C)
potassium ion may be intercalated in the c-MnO2 with a heating rate of 278 K (5 C) min1. The surface
structure. The extent of intercalation depends on the morphology of the EMD samples was determined using
type of surfactant used. The influence of these two field emission scanning electron microscope (FESEM,
surfactants on the structural, morphological, and elec- ZEISS: SUPRA55). Fourier transform infrared (FT-IR)
trochemical characteristics of the EMD produced from spectra were recorded on a Nicolet 6070 spectrophotom-
purified manganese sulfate solution obtained from eter in the frequency range 400 to 4500 cm1. The
manganese cake was thus investigated. This work not morphology and interplanar spacings of the products
only examines the suitability of these two surfactants in formed before and after discharge were determined by
improving the electrochemical activity of MnO2 material transmission electron microscopy (TEM) coupled with
but also opens a new possibility for utilization of a energy dispersive analysis (EDS) using a JEOL 2010F
secondary source to produce an energy material. TEM model operated at 200 kV. TEM specimens were
prepared by grinding a small fragment of material scraped
from the pressed pellet under methanol and dispersing on a
perforated carbon support film. Specimens were examined
at liquid nitrogen temperature in a cooling stage, to reduce
II. EXPERIMENTAL beam damage and contamination effects. Rutherford back
scattering analysis (RBS) was used in conjunction with
A. Materials and Methods for Producing EMD
proton induced X-ray emission (PIXE) to determine the
Electrolytic manganese dioxide (EMD) was produced concentrations of elements such as O, C, Na, Ni, Co, Mn,
from purified MnSO4 solutions obtained from manga- and P. The RBS and PIXE measurements were made
nese cake at an anodic current density of 200 Am2 in a simultaneously on an ion beam accelerator using a
glass cell, as reported in our previous work.[16] Purified 2.6 MeV proton beam at ANSTO facilities.

METALLURGICAL AND MATERIALS TRANSACTIONS E


C. Electrochemical Characterization species is first formed[24,26,30–32] together with some solid
The pellet preparation, cell arrangement, and all other intermediates such as MnOOH(s) and Mn2O3(s). Being
conditions and parameters were identical to those of our unstable in acidic solution,[2] the Mn3+ ion undergoes a
previous work.[29] For cyclic voltammetric (CV) experi- disproportionation reaction forming Mn4+ and Mn2+.
ments, a standard three-electrode cell was used. For this Mn2þ 3þ 
sol ! Mnads þ e ½3
purpose, EMD was used as the working electrode. The
counter electrode was a zinc foil, which was separated from
the main electrolyte (9 M KOH) by means of a porous frit 2Mn3þ 4þ 2þ
ads ! Mnads þ Mnsol : ½4
to avoid the spallation of zinc inserted into the MnO2.
Mercury-mercury oxide (Hg/HgO) served as the reference
During the electrodeposition process, Mn3+ ions may
electrode. Reported potentials are relative to Hg/HgO
be trapped in the MnO2 lattice, possibly resulting in
with a slow scan rate of 50 lV/s. For the galvanostatic
defects in the crystal structure.[22] Adsorption of surfac-
experiments the applied discharge current was 20 mA
tants at the substrate/electrolytic solution interface may
with a cut-off voltage (COV) of 0.9 V, while the applied
inhibit the rate of Eq. [3] due to the blocking of the
charge current was 40 mA with a cut-off voltage of 1.9 V.
active growth sites, thereby allowing electrodeposition
The discharge capacities were recorded for up to 15 cycles.
preferentially on the crevices. The electron/ion transfer
The galvanostatic measurements were carried out using a
kinetics depends largely on the degree of surface
BITRODE deep cycle battery tester (LCN1-25-24, USA).
coverage of the electrode due to mechanical blocking
that is caused through electrostatic interactions.[33,34]
Coverage of the electrode surface may change the
III. RESULTS AND DISCUSSION electrical double layer characteristics and thus affect
A. Electrolysis the interfacial energy, dielectric constant, potential, and
current distribution at the electrodes, resulting in
The current efficiency (CE) and energy consumption modified crystal growth. Hence, organic surfactants
(EC) of electrodeposited EMD prepared in the presence of may play a crucial role in facilitating the formation of
TX-100 and Tw-20 surfactants are given in Table I. In the compact deposits with greater surface area. However,
absence of surfactant as additive in the bath, 89 pct CE was the effect of surfactant concentration should be well
observed. With the introduction of a small amount of understood. Higher concentrations may lead to irregular
surfactant (TX-100), the CE slightly increased to 89.5 pct morphology owing to higher ohmic potential drop and
with slight decrease in EC from 1.65 to 1.64 kWh kg1. electrode overvoltage.[35]
With an increase in the concentration of TX-100, the
current efficiency decreased, and an increase in energy
consumption was observed, while with the addition of ten B. Structure and Function of Surfactant in the Solution
times the CMC of Tw-20, the CE increased to 98 pct, while
EC decreased from 1.65 to 1.46 kWh kg1 during elec- Surfactants lower the surface tension of a liquid as
trolysis. This suggests strong adsorption of the additives on well as liquid–liquid interfacial tension or liquid–solid
the electrode surface, as suggested by Ghaemi et al.[22] interfacial tension. Surfactants are amphiphilic mole-
In general, the electrodeposition of manganese diox- cules with both hydrophilic and lipophilic moieties. The
ide from an acidic sulfate solution proceeds through the adsorption and desorption of these amphiphilic mole-
following reactions: cules at the solid liquid interface were possible in this
work with the given applied potential range. Critical
At anode: Mn2þ þ 2H2 O ! MnO2 þ 4Hþ þ 2e ½1 micelle concentration (CMC) is the concentration of a
surfactant above which micelle formation occurs. The
existence of micelles does not preclude the existence of
At cathode: 2Hþ þ 2e ! H2 : ½2 individual surfactant molecules in solution. Further
addition of surfactant after reaching the CMC will
At the anode, the formation of MnO2 does not take ultimately increase the no. of micelles. The structure of a
place in a single step; rather, Mn3+ as an intermediate material electrodeposited in the presence of a surfactant

Table I. Effect of the Surfactants TX-100 and Tw-20 on Current Efficiency (CE) and Energy Consumption (EC) during Electrode-
position of MnO2

Amount
Surfactant in CMC CE (pct) EC (kWh kg1) BET Surface Area (m2 g1) Open Circuit Voltage (V)
Nil 89 1.65 101.908 1.550
TX 10 t 89.5 1.64 125.843 1.605
15 t 87 1.74 173.982 1.578
30 t 70 2.07 114.685 1.482
Tw 10 t 98.3 1.46 146.971 1.548
15 t 95.7 1.58 152.525 1.508
30 t 85.8 1.72 133.188 1.480

METALLURGICAL AND MATERIALS TRANSACTIONS E


can be related to the micelle population. With an
increase in the concentration of surfactant (from 10 to e
30 t in the current study), a different morphological
pattern may be created in the diffusion layer of the
electrode.[33] Surfactant adsorbed to charged surfaces d
may arrange in bilayers called admicelles, forming a
molecular layer with head groups down on the surface

Intensity (a.u)
and a second layer with head groups facing the c
solution.[36] The CMC values reported for non-ionic
surfactants Tw-20 and TX-100 are 8.5 9 104 and
9.2 9 104 M, respectively.[37] The surfactant with low- b
er CMC implies higher adsorbed amounts. Hence, to 110
make a comparative study of the effect of surfactants on 021 240 002
121 221
the electrochemical processes, the two non-ionic surfac- 061 a
tants (Tw and TX) with different chemical structures
were chosen.
In the case of TX-100, the hydrophobic aromatic 15 20 25 30 35 40 45 50 55 60 65 70 75 80
hydrocarbon is oriented toward the solution.[34] With an 2θ (degrees)
increase in surfactant concentration (beyond the CMC),
surfactants may assume a range of conformations Fig. 1—X-ray diffraction patterns of (a) EMD0, (b) EMDTX15, (c)
including the interweaving of hydrophobic chains of EMDTX30, (d) EMDTw15, (e) EMDTw30.
adjacent molecules and the formation of admicelles.[38]
The underlying mechanism of adsorption is an attrac- 50
tion between the oxygen atoms in polyoxyethylene MnO2 region
45
groups of TX-100 or Tw molecule and the Mn+2
species, forming Mn+2-TX and Mn+2-Tw pairs, respec- 40
Transmittance / a.u.

EMDTX30
tively. These pairs will be affected by surfactant con- 35
centration and will influence the electrodeposition of
30 EMDTW30
MnO2. Hence, the concentration of surfactant should be
limited to a particular value, above which there will be 25 EMDTW15
undesirable consequences in terms of physical and 20
electrochemical characteristics of deposited EMD. EMDTX15
15
1420
10 EMD0
C. Structural Characterization 1635 1130
5 3375
Figure 1 shows the X-ray diffraction (XRD) patterns 754
of EMD samples prepared at different concentrations of 0
4000 3500 3000 2500 2000 1500 1000 500
surfactants. The diffraction patterns show in situ addi-
-1
tion of either surfactant to the electrolytic bath had no Wavenumber / Cm
significant effect on the crystalline phase of c-MnO2. The
peak intensity for all samples is weak, with peak Fig. 2—FTIR spectra of EMD samples (a) EMD0, (b) EMDTX15, (c)
broadening indicating a typical amorphous quality to EMDTw15, (d) EMDTw30, (e) EMDTX30.
the deposit. All of the diffraction peaks can be indexed
to an orthorhombic phase of c-MnO2 with lattice with the water crystallization,[40] and a broad absorption
constants: a = 8.70 Å, b = 2.90 Å and c = 4.41 Å. band at ~3400 cm1 is due to OH stretching vibra-
This is in good agreement with the standard values tions.[40–43] Additional peaks at 968 and 1025 cm1 are
(JCPDS card no. 65-1298; a = 9.27 Å, b = 2.87 Å, attributed to Mn-O-H structural vibrations in agree-
c = 4.53 Å). The peaks at 2h values of ~22, ~37.2, ~42, ment with the spectral features reported by McBride.[44]
~56, and ~68 deg correspond to the 110, 021, 121, 240/
221, and 002/061 crystal planes of c-MnO2, respectively. D. Thermal Analysis
The FTIR spectra recorded for the EMD samples in
the presence and absence of surfactant (Figure 2) show Thermodynamic stability as well as electrical conduc-
significant IR peaks at different frequency ranges. The tivity of manganese dioxide are strongly governed by the
broad absorption band detected in the finger print presence of water molecules in the crystal lattices; the
region of 400 to 600 cm1 confirms the structural water content causes a variation in the crystal lattice
vibrations occur in MnO6 octahedral in c-MnO2 struc- with a consequent effect on electrical conductivity. The
ture.[39] The absorption band at 1100 cm1 can be water content enhances the transportation of active
attributed to the MnO2 stretching mode and/or O-H ionic species and consequently increases the specific
bending vibrations[39] associated with hydrogen bond- capacitance of manganese dioxide.[45] Hence to deter-
ing, indicating the presence of bound water mole- mine the role of thermal stability, thermal analysis
cules.[40,41] The strong absorption band at ~1630 cm1 (TGA and DTA) was carried out for all EMD samples
can be attributed to O-H bending vibrations associated (Figures 3(a) and (b)). Release of physisorbed water

METALLURGICAL AND MATERIALS TRANSACTIONS E


below 383 K (110 C) is observed for all samples. The decomposition of MnO2 to Mn2O3,[48] and is insignif-
mass loss in the region of 383 K to 573 K (110 C to icant in case of EMDTX15 (at ~768 K (~495 C) and
300 C) is due to removal of chemically bound water EMDTw15 (at ~770 K (~497 C)). Endo peaks observed
associated with c-MnO2. This is followed by mass loss in ~1173 K (~900 C) attributed to the decomposition of
the temperature range 673 K to 873 K (400 C to Mn2O3 to Mn3O4. From the TG and DTA data, it may
600 C) and 873 K to 1173 K (600 C to 900 C), which be concluded that the more significant water loss in the
indicates thermal decomposition resulting in the forma- early stage of heating for EMDTX30, EMDTw30, and
tion of Mn2O3 and Mn3O4, respectively.[46,47] Among EMD may be due to a loose structure with a greater
the doped and undoped sample, the mass loss during quantity of interlayer water. This corresponds to the
this step is substantially higher for EMDTX30, EMD0, instability in the structure responsible for the deteriora-
and EMDTw30. However, EMDTX15 and EMDTw15 tion of these EMD samples during reduction–oxidation
showed quite stable behavior on thermal analysis in battery applications. However, EMDTX15 and
throughout the temperature range. EMDTw15 showed less intense peaks, suggesting that
Differential Thermal Analysis (DTA) shown in bound water molecules responsible for the electrochem-
Figure 3(b) indicates that all EMD samples in this work ical activity of EMD are more strongly bound in these
show a sharp intense peak around 100 C corresponding EMD samples, imparting structural stability to the
to the loss of physisorbed water. A broad endo peak at material.
~543 K (~270 C) for EMDTX30 and at 578 K (305 C)
for EMD0 corresponds to the removal of all the
E. Surface Morphology
structural water. This peak was absent in other EMD
samples. The sharp endo peak observed at (833 K) Surfactants may play an important role in modifying
560 C for EMD sample corresponds to the thermal the structure of EMD by controlling the nucleation and
growth mechanism during electrodeposition[49] pro-
cesses, which in turn impinge on the stability and
34
rechargeability of EMD.[22] The presence of surfactant
32 in the solution strongly affects the particle size of EMD
30 material. In this work, smaller grains have been formed
Weight loss / %

28
in both EMDTX15 and EMDTw15 samples (Figures 4(a)
and (b)) relative to those obtained from EMD produced
26 in the surfactant-free bath. However, with increase in
(e)
24 surfactant concentration (in EMDTX30 and EMDTw30),
(d)
22 the grain size increases. All the EMD samples consist of
(c) particles with an angular shape having a rough surface.
20
On magnifying the surface of the EMD grain, it was
18 (b) observed that EMD0 seems to consist of nanopores,
16 (a) while EMDTX15 and EMDTX30 show fibrous- and spike-
like growth, respectively. EMDTw15 consists of spherical
0 200 400 600 800 1000
nanoparticles in the range of 300 nm with an agglom-
Temperature / °C erated form.
(a)
F. Surface Area and Pore Size
(e)
4 Surface area and porosity play significant roles in
controlling the electrochemical activity of EMD. Sur-
(d)
(c)
face area reached a maximum at 15 t for each of the
2
surfactants (Table I) for the EMD produced in this
Weight loss / %

(b)
work. Pore size distribution of the EMD samples was
0 carried out to evaluate its effect on rate capability of
(a) EMD produced in the presence of surfactant. All EMD
-2 samples showed isotherms of type-IV pattern with
strong hysteresis behavior. The adsorption–desorption
-4 isotherm as well as the shape and position of the
hysteresis loop were affected by the morphology and
specific pore size.
-6
0 200 400 600 800 1000 Figure 5 shows the nitrogen adsorption–desorption
Temperature / °C isotherm of the doped (presence of surfactants) and
undoped (absence of surfactants) EMD samples. The
(b) hysteresis loops obtained for both EMD0 and EMDTX30
are a hybrid of type H2, H3, and H4; however,
Fig. 3—(a) TGA curves of (a) EMDTX15, (b) EMDTw15, (c)
EMDTw30, (d) EMD0, and (e) EMDTX30. (b) DTA curves of (a)
EMDTX15, EMDTw15, and EMDTw30 show the behavior
EMD0, (b) EMDTX15, (c) EMDTX30, (d) EMDTw15, and (e) of H2 type hysteresis loop with a higher surface area
EMDTw30. relative to the EMD0 and EMDTX30. The wide hysteresis

METALLURGICAL AND MATERIALS TRANSACTIONS E


Fig. 4—a FESEM images of EMD0 (a1, a2), EMDTX15 (b1, b2), EMDTX30 (c1, c2). b FESEM images of EMDTw15 (d1, d2), EMDTw30 (e1, e2).

loop indicates that all the EMD samples are predom- surface area of EMD increases markedly (Table I),
inantly composed of mesopores. Mesopores with radius limited by a certain concentration (CMC value) of
of 1.87 to 1.97 and 1.68 to 1.95 nm are obtained for surfactant, after which it decreases. EMDTX15 and
EMDTX15 and EMDTw15, respectively, versus EMD0 EMDTw15 show the highest surface area of 173.9 and
with a pore radius of 1.58 to 1.96 nm. The pore size 152.5 m2 g1 in comparison with EMD0, which shows a
diameters and distribution of all samples were calculated surface area of 101.9 m2 g1. In general, EMD prepared
using the Barrett–Joyner–Halenda (BJH) method. It in the presence of surfactant shows a higher surface area
was also observed that with addition of surfactant, the than that produced in the absence of surfactant. As

METALLURGICAL AND MATERIALS TRANSACTIONS E


Fig. 4—continued.

reported by[50,51] charge transfer processes between the at the starting potential. The CV is characterized by a
electrolyte and the crystal phase can be affected by well-defined reduction peak (C1 = 0.390 V) and an
micro/meso porosity, and a decrease in the pore size oxidation peak (A1 = 0.09 V) which is a fingerprint for
may lead to reduction of accessible electrochemical electron transfer processes corresponding to a redox
active sites. To support this fact, doped EMD samples behavior suitable for energy storage and battery applica-
obtained in this work with high surface area as well as tions. For TX-added EMD, it is observed that the
with greater pore size show improved cyclic behavior cathodic and anodic currents in Figure 6(b) are higher,
(discussed in the Section III–G) relative to ‘‘non-surfac- and the reduction and oxidation peaks (at C1 = 0.450
tant’’ material, due to greater electrochemically accessi- and A1 = 0.030) have become more defined. The effect of
bility of sites available for proton diffusion inside MnO2 adding more TX (TX30) during electrodeposition of
cathode. EMD also had an impact on the voltammogram shown in
Figure 6(c). The redox peaks have shifted to lower
potentials (at C1 = 0.430 and A1 = 0.025) with the
G. Electrochemical Activity
separation between the peaks reduced to 0.040 V, indi-
The electrochemical behavior of the EMD samples was cating excellent reversibility. However, the area under the
characterized by potentiostatic (potential controlled) and peak for TX30 is less than that for TX15. Further
galvanostatic (current controlled) experiments. Cyclic addition of TX did not influence the redox process (not
voltammograms of the doped and undoped EMD mate- shown here). TX has both hydrophilic and lipophilic
rials are shown in Figure 6. The cyclic voltammogram of groups so that when its concentration reaches or sur-
EMD0 (Figure 6(a)) is initiated in the cathodic region and passes CMC, the surfactant tends to form micelles, and
then reversed back to the anodic region before it finishes the adsorption of Mn2+ ions would be comparatively

METALLURGICAL AND MATERIALS TRANSACTIONS E


70

60

Volume (ccg STP)


50

-1
40

30 adsorption
desorption
20

0.0 0.2 0.4 0.6 0.8 1.0


Relative pressure ( P/P 0 )
EMD0

60 70
55
60
50
Volume ( ccg-1 STP )

Volume( ccg-1 STP)


45 50
40
40
35
30 30
25 adsorption adsorption
desorption 20
20 desorption
15 10
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure ( P/ P0) Relative pressure (P/P0)
EMD TX15 EMD TX30
45
55
40
50
35
Volume (ccg STP)

45
Volume (ccg-1 STP)

40 30
-1

35 25
30
20
25
15 adsorption
20 adsorption
desorption
desorption 10
15

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure (P/P 0) Relative Pressure( P/P 0 )
EMD Tw15 EMD Tw30

Fig. 5—Nitrogen adsorption and desorption isotherms.

low. The effect of varying the concentration of TX-100 in oxidation of Mn2+/3+ reactions. In the case of the non-
the EMD is clearly reflected in the voltammograms. The ionic surfactant, Tw20, the voltammogram given in
EMD produced in the presence of surfactant contains Figure 6(d) does not reveal much influence of Tw in
adsorbed surfactant molecules and thus has a higher terms of the peak intensities relative to EMD0. Although
surface area to that of the bare EMD (undoped), resulting the nature of crystallization is not altered due to the
in higher redox peak currents during the reduction and presence of TX-containing EMD (as evidenced by XRD,

METALLURGICAL AND MATERIALS TRANSACTIONS E


0.004 0.004 - 0.025 V

A1
0.002 - 0.097 V 0.002

Current / A
Current / A
A1
0.000 0.000

C1
-0.002 C1
-0.002 - 0.390 V
- 0.430 V
-0.004
-0.004 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3
Potential vs. Hg/HgO / V
Potential vs. Hg/HgO / V
(c) TX30
(a) EMD0

0.030 V
0.004 0.004

A1

0.002 0.002 - 0.085 V

Current / A
Current / A

A1
0.000
0.000

-0.002 C1
-0.002 C1 - 0.450 V

- 0.450 V -0.004
-0.004 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3
Potential vs. Hg/HgO / V
Potential vs. Hg/HgO / V
(d) TW15
(b) TX15

Fig. 6—Cyclic voltammograms of the EMD samples (a) containing no surfactant, (b) Triton (TX15), (c) Triton (TX30), and (d) Tween (Tw15).

Figure 1), the change in morphology and larger surface intensity, corresponding to conductive carbon as an
area (Table I) was reflected in the electrochemistry, additive in the cathode material to improve the conduc-
suggesting a higher rate of nucleation when TX-100 is tivity. In addition to the parent MnO2 diffraction lines,
present in the electrolyte. the evolution of new diffraction peaks corresponding to
The results of cyclic voltammetry for these studies MnOOH and Mn3O4 is seen in Figure 8. The diffraction
were also confirmed by galvanostatic charge/discharge pattern obtained for EMD material prepared in the
studies (Figure 7). As expected, the available discharge presence or absence of surfactants shows the surfactants
capacity for the TX-added EMD showed 273 mAh g1 did not influence the formation of the discharged
(Figure 7(b)), while the surfactant-free EMD showed products. The product obtained is in accordance with
just 170 mAh g1 (Figure 7(a)), additional capacity of the well-known mechanism of c-MnO2 as an insertion
38 pct. The difference in capacities between the various electrode in aqueous KOH electrolyte involves H+
concentrations of TX (Figures 7(b) to (c)) is negligible insertion forming the manganese oxy hydroxides
(only 20 mAh g1). On the other hand, non-ionic (MnOOH) and manganese (III) oxide (Mn3O4).[24–26]
surfactant (Tw) in Figure 7(d) showed only a slight The TEM images of the discharged material, on the
improvement of 181 mAh g1 relative to EMD0. The other hand, show a different but intriguing picture. The
galvanostatic and potentiostatic results of this study TEM images of the cathode (c-MnO2) material prepared
agree well. in the presence of surfactants before and after discharge
To investigate the role of the discharge mechanism in are shown in Figure 9. The TEM image of the cathode
the alkaline KOH electrolyte, the cathode material after material before discharge (Figure 9(a)) showed the
discharge was thoroughly washed with distilled water, parent MnO2 and carbon as additive, but after discharge
and the variation in crystal structure analyzed through (for undoped; Figure 9(b)), the regions A and B
XRD before and after the electrochemical processes. correspond to MnO2 with a small amount of potassium
The XRD results are displayed in Figure 8. The XRD (EDS analysis not shown here). For TX-added EMD,
pattern shows a peak at 2h = 27 deg with higher Figure 9(c), all regions labeled A to E showed MnO2

METALLURGICAL AND MATERIALS TRANSACTIONS E


1.6 EMD materials exhibit a sharp drop in potential at
1.3 V, although the cells have been tested under identical
conditions. These curves in Figures 7(a) and (d) show
1.2 typical anodic limited behavior, suggestive of passiv-
Cell Potential / V

(a) (b) ation of zinc on the surface of the MnO2 cathode. This
(d)
(c) has been confirmed through the simultaneous PIXE and
0.8 RBS analyses of the EMD material before and after
discharge. A typical bar graph containing several steps
corresponding to elements such as K, Mn, Zn, and C is
0.4
given in Figure 10. The variation in the quantity of
manganese and carbon across the electrodes before and
after discharge appears to be fairly constant. However,
it is quite important to note the differences seen in the
0.0
0 100 200 300
concentrations of the potassium and zinc. For the
-1 electrode before discharge, the concentrations of potas-
Discharge Capacity / mAh g
sium and zinc are nil, as is expected given there is no
potassium and zinc in the fresh sample. For the
Fig. 7—First discharge behavior of the EMD samples (a) containing
no surfactant, (b) Triton (TX15), (c) Triton (TX30), and (d) Tween discharged electrode, the concentration of potassium
(Tw30). ion is increased, particularly in the case of TX-100 as the
surfactant. This reflects the TEM microscopy analysis
and further supports the potassium intercalation. A
C Ο MnO 2 Δ Mn 3 O4 further notable point in the RBS analysis, in Figure 10,
C Carbon • MnO OH is the high concentration of zinc for EMD0 and Tw30
electrodes. This shows that the zinc has been precipi-
• Ο• Ο
Ο Δ tated on the cathode as Zn(OH)2, thus inhibiting further
(d)
• • Ο
electron transfer. Hence, the two cells are anode limited
Ο

Ο • with a sharp fall of potential as seen in the discharge
• Ο
Δ Ο curve Figures 7(a) to (d).
(c)
The cumulative discharge capacities of the EMD samples
Ο• Ο in the presence and absence of surfactant TX and Tw are
• ΟΔ • Ο
(b) • shown in Figure 11. The charge and discharge character-
• Ο•
istics enable evaluation of the suitability of the prepared
Ο EMD as a battery material. In Figure 6, it was clearly
Ο
(a)
Δ • • Ο indicated that surfactant works up to an optimum level,
above which an alteration in the cyclic behavior is observed.
Interestingly both surfactants employed in this work
20 30 40 50 60 70 showed their best performance at their 15 t (CMC 9 15)
2 theta / degrees value. EMDTX15 and EMDTw15 showed cumulative dis-
charge capacity of ~2213 and ~2242 mAh g1 at 15th cycle,
Fig. 8—X-ray diffraction patterns of the discharged EMD samples respectively, against a cumulative discharge capacity of
(a) containing no surfactant, (b) Triton (TX15), (c) Triton (TX30), ~1300 mAh g1 for EMD0 at 11th cycles which remain
and (d) Tween (Tw30). constant so far. The discharge capacity of 280 mAh g1 was
obtained from the EMD0 against 170 mAh g1 for the
reported values for EMD in the absence of additives.[52] The
with a larger amount of potassium. For Tw-added initial discharge capacity, above and below which the
EMD, Figure 9(d), except the regions 3 and 4, all discharge capacity decreases, of 270 mAh g1 was observed
regions are identified as K free MnO2 suggesting only for EMDTX15, where as the initial discharge capacity of
a small amount of potassium ion is present. Almost all EMDTw15 was 240 mAh g1, somewhat less than the
the reported literature on MnO2 in alkaline electrolyte EMDTw10 (275 mAh g1). In spite of this, the EMDTX15
describe either protonation[24–27] or lithiation[3,15,28] as showed better cycling stability and higher cumulative
the mechanism involved during discharge but an discharge capacity than EMD0, EMDTw10, and EMDTw30.
unusual observation of potassium ion insertion or as The available cumulative discharge capacity for EMD0 after
surface film formation is evidenced in this study. 10th cycles was quite low of ~1300 mAhg g1, so the cycling
Interestingly, the extent of potassium ion increases with studies were halted well before the cut-off range. The
TX-100 as the surfactant, and this conclusion is further characteristic property of non-ionic surfactant is a tendency
supported by Rutherford backscattering experiments, to accumulate in electrolytic solution that enhances the
discussed in the next section. The observed microscopy adsorption. In the case of TX, nucleation and growth of
results are in parallel with the cyclic voltammograms electrodeposited MnO2 grains in a preferential direction was
and discharge behavior, indicating that the surfactant promoted while favoring a high surface area of the deposit.
facilitates K+ insertion. TX-100 renders the surface more hydrophobic and acts as a
A close inspection of Figures 7(a) and (d) reveals that molecular spacer[21,53] through the steric factor of the phenyl
the discharge curves for surfactant-free and Tw-added ring, while the Tw counterpart is without a phenyl ring. This

METALLURGICAL AND MATERIALS TRANSACTIONS E


Fig. 9—TEM images of EMD (c-MnO2) samples (a) before and (b) after discharge containing no surfactant, (c) discharged Triton (TX30), and
(d) discharged Tween (Tw15).

25 difference in structure observed for non-ionic surfactant led


Relative Composition / at. (%)

K
EMD 0 EMD (TX30 ) EMD (TW30 ) Mn
a role for Mn deposition by facilitating uniform current
20 Zn distribution resulting in Mn-TX complexes. Hence, the
C presence of TX molecules facilitated the superior electro-
15 chemical performance of discharge capacity 275 mAh g1.
The enhancement of cycle life in the presence of the two
surfactants may be attributed to the following: EMDTX15
10
offers high surface area with large pore size (mesopores)
with nanofibrous morphology (as per the FESEM images)
5 which decreases the solid state diffusion path length of
protons and electrons into and out of the bulk of MnO2, as
0 happened in MnO2 films.[54] As explained before, the large
pore size facilitates the proton diffusion inside the MnO2
Before / After Before / After Before / After
discharge discharge discharge electrode, during charge–discharge cycling resulting good
cyclic stability. Second in terms of conductance, EMDTX15
Fig. 10—Simultaneous PIXE and RBS analyses of the possible ele- consists of smaller particles with porous structures, and
ments present in the discharged EMD samples containing no surfac- EMDTw15 composed of spherical nanoparticles (as shown in
tant, Triton (TX30), and (d) Tween (Tw30). FESEM image), which allow better contact with the

METALLURGICAL AND MATERIALS TRANSACTIONS E


2250
300 e
2000 d
c
Discharge capacity(mAh g-1)

discharge capacity(mAhg )
-1
250 1750
b
1500
200

Cumulative
1250
150
a
1000

100 750
e
d 500
50 c
b 250
a
0 0
2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
Cycle number Cycle number
(a) (b)

2250
300 d
Discharge capacity( mAh g -1)

2000

discharge capacity (mAhg )


c

-1
250 1750
1500 b
200
Cumulative
1250
150 a
1000
750
100 d
500
50
c
250
a b 0
0 0 2 4 6 8 10 12 14 16
0 2 4 6 8 10 12 14 16
Cycle number
Cycle number
(c) (d)

Fig. 11—(a) Discharge capacities versus cycling behavior of (a) EMD0, (b) EMDTX30, (c) EMDTX5, (d) EMDTX10, (e) EMDTX15. (b) Cumulative
discharge capacities versus cycling behavior of (a) EMD0, (b) EMDTX30, (c) EMDTX5, (d) EMDTX10, (e) EMDTX15. (c) Discharge capacities ver-
sus cycling behavior of (a) EMD0, (b) EMDTw30, (c) EMDTw10, (d) EMDTw15. (d) Cumulative discharge capacities versus cycling behavior of
(a) EMD0, (b) EMDTw30, (c) EMDTw10, (d) EMDTw15.

graphite powder. Thermal stability of EMDTX15 and 2. FESEM images indicate that presence of surfactants
EMDTw15 is also responsible for the good cyclic behavior in the electrolytic bath plays a key role in control-
of EMDTX15 and EMDTw15 in comparison to EMD0, which ling the growth and nucleation of the deposited
resists the capacity fading during charge–discharge cycling. material and hence modifies the structure accord-
ingly giving porous morphology to the material.
3. The presence of surfactant enhances the electro-
IV. CONCLUSIONS chemical activity of EMD, not only by modifying
the structure but also by increasing the BET surface
EMD samples have been prepared from purified area and porosity of the EMD samples. EMD
manganese sulfate solution obtained from manganese doped with surfactants has good cycle life in spite
cake in the presence or absence of surfactants. The of a decrease in initial discharge capacity, relative
surfactants played a role in the intercalation behavior. A to surfactant-free EMD.
unique observation of potassium ion intercalation in 4. The presence of surfactant affects the thermal sta-
c-MnO2 structure is reported. From the present study, bility of the material and also brings mild changes
the following conclusions can be drawn. in spectroscopic analytics. There optimum concen-
tration of surfactant must be essential for the posi-
1. EMD samples prepared in the presence and absence
tive consequences in the performance and behavior
of surfactants show the characteristic peak of
of the EMD.
c-MnO2 and are electrochemically active. Addition
5. TEM imaging and RBS analysis showed evidence
of surfactant and surfactant type (either TX-100 or
for potassium intercalation with the extent being
Tw-20) has no significant effect on the crystal struc-
greater for TX-100 as the surfactant. The
ture of EMD.

METALLURGICAL AND MATERIALS TRANSACTIONS E


discharged product of the EMD material showed 20. H. Zhang, Y. Wang, C. Liu, and H. Jiang: J. Alloys Compd., 2012,
evidence of protonation. vol. 517, pp. 1–8.
21. S. Devaraj and N. Munichandraiah: J. Electrochem. Soc., 2007,
vol. 154, pp. A901–A909.
22. M. Ghaemi, L. Khosravi-Fard, and J. Neshati: J. Power Sources,
2005, vol. 141, pp. 340–50.
23. R. Khayat Ghavami, Z. Rafiei, and S.M. Tabatabaei: J. Power
ACKNOWLEDGMENTS Sources, 2007, vol. 164, pp. 934–46.
24. K. Kordesch, J. Gsellmann, M. Peri, K. Tomantschger, and R.
The authors are thankful to Prof. B.K. Mishra, Chemelli: Electrochim. Acta, 1981, vol. 26, pp. 1495–1504.
Director, CSIR-IMMT, Bhubaneswar, for permitting 25. A. Kozawa and J.F. Yeager: J. Electrochem. Soc., 1965, vol. 112,
to publish the paper. Authors also thank to the Minis- pp. 959–63.
try of Earth Sciences for partly sponsoring the work. 26. H.S. Wroblowa and N. Gupta: J. Electraonal. Chem., 1987,
The first author, A. Biswal is thankful to Council of vol. 238, pp. 93–102.
27. Y.F. Yao, N. Gupta, and H.S. Wroblowa: J. Electroanal. Chem.,
Scientific and Industrial Research (CSIR), India for 1987, vol. 223, pp. 107–17.
awarding Senior Research Fellowship. The author 28. M. Minakshi, P. Singh, and D.R.G. Mitchell: J. Electrochem. Soc.,
(M.M.) wishes to thank both the Australian Research 2007, vol. 154, pp. A109–13.
Council (ARC) and Australian Nuclear Science and 29. A. Biswal, B.C. Tripathy, T. Subbaiah, D. Meyrick, and M.
Minakshi: J Solid State Electrochem., 2013, vol. 17, pp. 1349–56.
Engineering (AINSE/AINGRA 08048) for providing 30. S. Nijjer, J. Thonstad, and G.M. Haarberg: Electrochim. Acta,
experimental time to enable work on TEM and RBS 2000, vol. 46, pp. 395–99.
facilities. 31. W.H. Kao and V.J. Weibel: J. Appl. Electrochem., 1992, vol. 22,
pp. 21–27.
32. F.-Q. Xue, Y.-L. Wang, W.-H. Wang, and X.-D. Wang: Electro-
REFERENCES chim. Acta, 2008, vol. 53, pp. 6636–42.
33. Z. Kozarac, S. Nikolic, I. Ruzic, and B. Cosovic: J. Electroanal.
1. O. Schilling and J.R. Dahn: J. Appl. Crystallogr., 1998, vol. 31, Chem., 1982, vol. 137, pp. 279–92.
pp. 396–406. 34. I. Felhosi, J. Telegdi, G. Palinkas, and E. Kalman: Electrochim.
2. Y. Chabre and J. Pannetier: Prog. Solid State Chem., 1995, vol. 23, Acta, 2002, vol. 47, pp. 2335–40.
pp. 1–130. 35. D.S. Wen and B.X. Wang: Int. J. Heat Mass Transf., 2002, vol. 45,
3. M. Minakshi, P. Singh, T.B. Issa, S. Thurgate, and R. De Marco: pp. 1739–47.
J. Power Sources, 2004, vol. 138, pp. 319–22. 36. F.J. Srydlowskl, D.L. Dunmire, E.E. Peck, R.L. Eggers, and W.R.
4. D. Balachandran, D. Morgan, G. Ceder, and A. van de Walle: J. Matson: Anal. Chem., 1981, vol. 53, pp. 193–96.
Solid State Chem., 2003, vol. 173, pp. 462–75. 37. A.M. Rodriguez, M.A.C. Vilchez, and R.D. Alvarez: J. Colloid
5. A. Urfer, G.A. Lawrance, and D.A.J. Swinkles: J. Appl. Electro- Interface Sci., 1997, vol. 187, pp. 139–47.
chem., 1997, vol. 27, pp. 667–72. 38. B. Factor, B. Muegge, S. Workman, E. Bolton, J. Bos, and M.M.
6. K. Kordesch and M. Weissenbacher: J. Power Sources, 1994, Richter: Anal Chem., 2001, vol. 73, pp. 4621–24.
vol. 51, pp. 61–78. 39. H. Abbas and S.A. Nasser: J. Power Sources, 1996, vol. 58, pp. 15–
7. W. Jantscher, L. Binder, D.A. Fiedler, R. Andreaus, and K. 21.
Kordesch: J. Power Sources, 1999, vol. 79, pp. 9–18. 40. J.B. Fernandes, B.D. Desai, and V.N.K. Dalal: Electrochim. Acta,
8. S. Chou, F. Cheng, and J. Chen: J. Power Sources, 2006, vol. 162, 1983, vol. 28, pp. 309–15.
pp. 727–34. 41. M.V. Ananth, S. Pethkar, and K. Dakshinamuthi: J. Power
9. K. Kordesch, W. Harer, W. Taucher, and K. Tomantscher: Pro- Sources, 1998, vol. 75, pp. 278–82.
ceedings of 22nd Intersociety Energy Conversion Engineering Con- 42. P. Ruetschi: J. Electrochem. Soc., 1984, vol. 131, pp. 2737–44.
ference, 1987, vol. 2, p. 1102. 43. J. Fitzpatrick, L.A.H. Maclean, D.A.J. Swinkels, and F.L. Tye: J.
10. H. Malankar, S.S. Umare, K. Singh, and M. Sharma: J. Solid Appl. Electrochem., 1997, vol. 27, pp. 243–53.
State Electrochem., 2010, vol. 14, pp. 71–82. 44. M.B. McBride: Soil Sci. Soc. Am., 1987, vol. 51, pp. 1466–72.
11. A.M.A. Hashem, K.S. Abou-El-Sherbini, S. Zein-El-Abedin, and 45. J.P. Zheng, P.J. Cygan, and T.R. Jow: J. Electrochem. Soc., 1995,
H. Abbas: J. Mater. Sci. Technol., 2006, vol. 22, pp. 25–30. vol. 142, pp. 2699–2703.
12. A.M.A. Hashem, K.S. Abou-El-Sherbini, M.H. Askar, and A.M. 46. J.H. Sharp and D.M. Tinsley: J. Therm. Anal., 1971, vol. 3, pp. 43–
Hashim: J. Mater. Sci. Technol., 2001, vol. 17, pp. 351–54. 48.
13. A.G. Kholmogorov, A.M. Zhyzhaev, U.S. Kononov, G.A. 47. J.B. Fernandes, B.D. Desai, and V.N.K. Dalal: J. Power Sources,
Moiseeva, and G.L. Pashkov: Hydrometallurgy, 2000, vol. 56, 1985, vol. 16, pp. 1–43.
pp. 1–11. 48. M.I. Zaki, M.A. Hassan, L. Pasupulety, and K. Kumari: Ther-
14. J.M.M. Paixdo, J.C. Amaral, L.E. Memoria, and L.R. Freitas: mochim. Acta, 1997, vol. 303, pp. 171–80.
Hydrometallurgy, 1995, vol. 39, pp. 215–22. 49. J.O. Besenhard, J. Gurtler, P. Komenda, and A. Paxions: J. Power
15. M. Minakshi, P. Singh, T.B. Issa, S. Thurgate, and R. De Marco: Sources, 1987, vol. 20, pp. 253–58.
J. Power Sources, 2004, vol. 130, pp. 254–59. 50. H. Adelkhani and M. Ghaemi: Solid State Ion., 2008, vol. 179,
16. A. Biswal, K. Sanjay, M.K. Ghosh, T. Subbaiah, and B.K. pp. 2278–83.
Mishra: Hydrometallurgy, 2011, vol. 110, pp. 44–49. 51. S.M. Davis, W.L. Bowden, and T.C. Richards: J. Power Sources,
17. Suhasini and A.C. Hegde: J. Electroanal. Chem., 2012, vol. 676, 2005, vol. 139, pp. 342–50.
pp. 35–39. 52. M.M. Thackeray: Prog. Solid State Chem., 1997, vol. 25, pp. 1–71.
18. Y. Li, H. Xie, J. Wang, and L. Chen: Mater. Lett., 2011, vol. 65, 53. P. Matejka, B. Vickova, J. Vohlidal, P. Pancoska, and V.
pp. 403–05. Baumruk: J. Phys. Chem., 1992, vol. 96, pp. 1361–66.
19. R. Jiang, T. Huang, J. Liu, J.H. Zhuang, and A. Yu: Electrochim. 54. S.C. Pang, M.A. Anderson, and T.W. Chapman: J. Electrochem.
Acta, 2009, vol. 54, pp. 3047–52. Soc., 2000, vol. 147, pp. 444–50.

METALLURGICAL AND MATERIALS TRANSACTIONS E

Potrebbero piacerti anche