You are on page 1of 7

Electrochimica Acta 52 (2007) 21892195

Evaluation of electrolytes for redox flow battery applications


M.H. Chakrabarti a , R.A.W. Dryfe b , E.P.L. Roberts a,
a

School of Chemical Engineering and Analytical Science, The University of Manchester,


P.O. Box 88, Manchester M60 1QD, UK
b School of Chemistry, The University of Manchester, P.O. Box 88, Manchester M60 1QD, UK
Received 17 May 2006; received in revised form 1 August 2006; accepted 17 August 2006
Available online 26 September 2006

Abstract
A number of redox systems have been investigated in this work with the aim of identifying electrolytes suitable for testing redox flow battery cell
designs. The criteria for the selection of suitable systems were fast electrochemical kinetics and minimal cross-contamination of active electrolytes.
Possible electrolyte systems were initially selected based on cyclic voltammetry data. Selected systems were then compared by charge/discharge
experiments using a simple H-type cell. The all-vanadium electrolyte system has been developed as a commercial system and was used as the starting
point in this study. The performance of the all-vanadium system was significantly better than an all-chromium system which has recently been
reported. Some metalorganic and organic redox systems have been reported as possible systems for redox flow batteries, with cyclic voltammetry
data suggesting that they could offer near reversible kinetics. However, Ru(acac)3 in acetonitrile could only be charged efficiently to 9.5% of
theoretical charge, after which irreversible side reactions occurred and [Fe(bpy)3 ](ClO4 )2 in acetonitrile was found to exhibit poor charge/discharge
performance.
2006 Elsevier Ltd. All rights reserved.
Keywords: Redox flow battery; Vanadium; Chromium; Ru(acac)3 ; [Fe(bpy)3 ](ClO4 )2

1. Introduction
Redox flow batteries are electrochemical energy storage
devices that utilise the oxidation and reduction of two soluble
redox couples for charging and discharging. They differ from
conventional batteries in that the energy-bearing chemicals are
not stored within at the electrode surface, but in separate liquid
reservoirs and pumped to the power converting device for either
charging or discharging [1,2]. Due to the use of two soluble redox
couples, solid-state reactions with their accompanying morphological changes at the electrodes are absent [3]. Thus, there are
no fundamental cycle life limitations associated with these processes such as shedding or shape changes, which usually occur
in conventional storage batteries.
Despite these advantages, the redox flow battery has not been
widely exploited to date. One disadvantage of the systems developed to date is the use of two separate redox species in the halfcells, leading to the potential for cross-contamination of active

Corresponding author. Tel.: +44 161 306 8849; fax: +44 161 306 4399.
E-mail address: edward.roberts@manchester.ac.uk (E.P.L. Roberts).

0013-4686/$ see front matter 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2006.08.052

electrolytes by transport through the membrane. For example,


there has been little recent interest in the development of the
iron/chromium redox flow cell due to this problem [4]. To redress
this issue, an all-chromium redox electrolyte was investigated at
the University of Manchester and the charge/discharge characteristics of a laboratory scale battery were reported [5,6]. Prior
to this, other workers have performed extensive investigations
on the all-vanadium redox system [3,710] and patented the
technology [11]. In addition, an all-neptunium system has been
evaluated [12], although the hazards of working with radioactive electrolytes are likely to limit the practical application of this
system. Several prototype vanadium systems have been investigated successfully [1316] and some systems are well on their
way to commercial success [17].
Despite such achievements, batteries employing aqueous
electrolytes have a low energy content. The energy output from
the battery is proportional to the potential window of operation available from the background electrolyte. The operating
potential window of aqueous electrolytes is limited due to water
electrolysis [5]. Organic electrolytes, which offer a wider potential window, have been investigated in this study. In addition,
species have been selected which minimize the effect of elec-

2190

M.H. Chakrabarti et al. / Electrochimica Acta 52 (2007) 21892195

trolyte cross-contamination. One approach is to use a single


system which offers three oxidation states, so that the discharged
species is the same on each side of the cell. Such a system would
have the advantage that any cross-contamination would only lead
to some self-discharge, and little or no cell balancing or electrolyte processing would be required. An approach whereby the
same cation is used but with different ligands on each side of the
cell has been suggested [18], but this has not been considered in
this study.
Electrolyte systems can be selected on the basis of the following properties, which are generally desirable for redox flow
batteries [18,19]:

give an indication of the reversibility of redox couples, further


experiments are needed to demonstrate that selected systems can
be used for energy storage. For example, a redox couple may be
reversible, but the charged species may be unstable over long
timescales, which would not be detected by cyclic voltammetry.
In this study systems which were found to exhibit fast kinetics
were tested for their charge/discharge performance in a simple
H-type cell. These experiments aimed to determine whether the
selected systems could be used for energy storage and to provide a preliminary indication of the relative performance of each
system.
2. Experimental

fast kinetics at the electrodeelectrolyte interface;


a relatively large open circuit potential;
reasonable cost;
high solubility in the process electrolyte.

In this study, the following series of redox systems in acetonitrile electrolyte were selected which apparently offered fast
electrode kinetics (based on literature data, e.g. [19] and [20])
and the potential to operate with a single electrolyte using a
species with three oxidation states.
(i) Ruthenium organic complexes
A number of ruthenium organic complexes which can
be both oxidized and reduced electrochemically have been
reported in the literature, and some of these have been
suggested as suitable candidates for a redox flow battery
[21]. Tris(2,2 -bipyridine) ruthenium (II) tetrafluoroborate
[Ru(bpy)3 (BF4 )2 ] has exhibited fast kinetics [19]. In addition, this system offers the possibility of cell voltages of up
to 2.6 V, much higher than is possible in aqueous battery
systems [19]. Ruthenium acetylacetonate [Ru(acac)3 ] has
also been reported as offering fast oxidation and reduction
kinetics [22] and a possible cell voltage of around 1.75 V.
(ii) Tris(2,2 -bipyridine) iron(II) perchlorate
This species can be oxidized and reduced [19] and offers
a possible cell potential of 2.4 V. This compound is available commercially and is significantly cheaper than the
ruthenium complexes.
(iii) Rubrene
Rubrene, a neutral organic species, can be oxidized and
reduced electrochemically [23]. The redox potentials of
these reactions offer a possible cell potential of around
2.3 V. Again this compound is available commercially,
although it is significantly more expensive than the other
redox species.
These systems are compared to the all vanadium redox flow
battery system, which has previously been investigated in detail
([3,711]) and has been commercialized in recent years [17].
In this system, vanadium in four different oxidation states is
used: V(II)/V(III) at the negative electrode and V(IV)/V(V) at
the positive electrode.
Each system was first tested by cyclic voltammetry in order
to evaluate the electrode kinetics. While cyclic voltammetry can

2.1. Electrolytes
Vanadium electrolytes were prepared from vanadium (IV)
sulphate (>99.99% purity, Aldrich), with the V(II)/V(III) couple
generated by electro-reduction. Sulphuric acid was used as the
background electrolyte.
Reagent grade tris(2,2 -bipyridine) ruthenium (II) chloride
is available from Aldrich. Since oxidation of the chloride salt
was known to be irreversible [24], the tetrafluouroborate salt
[Ru(bpy)3 (BF4 )2 ] was prepared by addition of NaBF4 in acetonitrile and precipitation of NaCl. Ruthenium acetylacetonate
[Ru(acac)3 , 97% purity, Aldrich], tris(2,2 -bipyridine) iron(II)
perchlorate (reagent grade, GFS) and rubrene (reagent grade,
Aldrich) were used for the preparation of the respective electrolytes (Caution: perchlorate salts are potentially explosive and
should be handled with appropriate care). Tetraethyl-ammonium
tetrafluoroborate and tetraethyl-ammonium perchlorate were
used as the background electrolyte.
To remove dissolved oxygen, electrolytes were sparged for at
least 10 min with oxygen-free dry argon (aqueous electrolytes)
or nitrogen (organic electrolytes). Water was removed from the
organic electrolytes using zeolite 4A (Merck) to a moisture level
of below 0.005 wt%.
2.2. Cyclic voltammetry
A graphite rod (Goodfellow) of surface area 0.06 cm2 was
used as the working electrode for cyclic voltammogram experiments with the vanadium battery electrolytes. A glassy-carbon
electrode (I.J. Cambria Scientific) of surface area 0.07 cm2 was
used for cyclic voltammetry in organic media. The electrode
was polished with alumina washed with de-ionised water and
acetone following the procedure described in literature [21].
The reference electrode used in aqueous solutions was the
saturated calomel electrode along with a salt bridge. Organic
media required the use of a silver wire quasi-reference electrode (AgQRE). A platinum counter electrode was used in each
case.
Cyclic voltammetry was conducted using a standard threeelectrode cell, with a Autolab/PGSTAT30 potentiostat for potential control. All solutions were de-aerated prior to experiments.
The solution headspace was purged with inert gas for the duration of experiments.

M.H. Chakrabarti et al. / Electrochimica Acta 52 (2007) 21892195

Fig. 1. Schematic diagram of a glass cell apparatus for small-scale charge/


discharge tests of redox couples.

2.3. Charge/discharge experiments


A schematic diagram of the H-type test cell is shown in
Fig. 1. The constant current (galvanostatic) charge/discharge
characteristics of the redox couples were used to evaluate their
performance in a prototype redox flow battery.
Each electrolyte compartment contained 40 ml of electrolyte,
except for the initial charging of the all-vanadium system, where
twice as much electrolyte (80 ml) is required in the anodic compartment. After the first charging of the all-vanadium half of the
electrolyte in the anodic compartment was removed [3].
Graphite felt electrodes (Sigratherm GFA 10) were
employed for charge/discharge experiments in the H-type
glass cell. The graphite felt electrode had dimensions of
30 mm 15 mm 10 mm. Graphite rods were used as current
collectors. An UltrexTM (Membranes-International Ltd.) anion
exchange membrane was used for vanadium charge/discharge
tests and a Neosepta AHA membrane (Eurodia Industrie SA)
was used for organic charge/discharge experiments in all cases.
Membranes were pre-conditioned by exposing them to the
required test solution for at least 6 h prior to experiments. The
circular area of the membrane exposed to the electrolyte in the
cell had a diameter of 27.5 mm.
The charge/discharge experiments were carried out under
constant current conditions. The current was selected on the
basis of preliminary experiments. For the vanadium electrolytes,
which had a much higher conductivity than the organic electrolytes, a high charging current of 100 mA was used. Although
this led to a high cell voltage during charging (4 V) with the
likelihood of side reactions, the aim was to attempt to charge the
cell close to its maximum capacity. For the organic electrolytes,
much lower charging currents were used in order to minimize
ohmic losses and to evaluate whether efficient charging could be
achieved. Discharge currents were selected to ensure that sig-

2191

nificant cell voltages were obtained. For the vanadium system, a


discharge current of 2 mA was found to be suitable while with the
lower conductivity organic electrolytes, it was necessary to use
lower discharge currents (0.5 mA or lower). The applied current
was controlled using a galvanostat and the total cell potential and
the potential of each electrode were monitored [relative to saturated calomel electrode (SCE) or silver quasi reference electrode
(AgQRE)] throughout each experiment. Since a silver quasi reference electrode was used in the organic electrolytes the absolute
value of the electrode potentials is not meaningful, and the measurements can only be used to monitor significant changes in
the potential of each electrode. In aqueous experiments mass
transport was provided by means of sparging the solution with
argon gas, while for organic solutions a magnetic stirrer was
used.
3. Results and discussion
3.1. Cyclic voltammetry
The results of cyclic voltammetry experiments for the vanadium system suggested that the kinetics of the V(II)/V(III)
couple were relatively fast, while the V(V)/V(IV) redox couple was found to be irreversible, consistent with results reported
in literature [2527].
All of the organic electrolyte systems studies demonstrated
reasonably fast (in most cases reversible) kinetics for both oxidation and reduction reactions. The results suggested that all
four systems could be oxidized or reduced, confirming their
suitability for a redox flow battery with the same species occurring in the discharged state. Table 1 compares the species
against each other based on the possible open circuit potential of a battery based on the system, their solubility, electrochemical reaction kinetics, and cost. The information on the
kinetics was obtained from cyclic voltammetry experiments
and from literature data. The Ru(acac)3 and Fe(bpy)3 (ClO4 )2
systems were selected for further study on the basis of their
superior solubility and fast kinetics. The relatively low cost of
the Fe(bpy)3 (ClO4 )2 makes this system particularly attractive.
Cyclic voltammograms for the Ru(acac)3 and Fe(bpy)3 (ClO4 )2
systems are shown in Figs. 2 and 3, illustrating the combination
of multiple redox couples with fast kinetics. The cyclic voltammogram for Fe(bpy)3 (ClO4 )2 indicates that the species can be
reduced at least twice, consistent with previous studies [19]. The
reduction reactions are presumed to be single electron reductions
of Fe(bpy)3 2+ to Fe(bpy)3 + and Fe(bpy)3 .

Table 1
The characteristics of the redox species studied in acetonitrile
Chemical

Expected open circuit potential (V)

Solubility in solvent

Reaction kinetics [19,2224]

Approximately cost per mmol

Ru(acac)3
Ru(bpy)3 (BF4 )2
[Fe(bpy)3 ](ClO4 )2
Rubrene

1.77
2.62
2.41
2.33

High
Poor
Moderate
Poor

Reversible
Quasi-reversible
Reversible
Reversible

12 [i]
21 [i]
2 [ii]
37 [i]

Costs were obtained from SigmaAldrich [i] and GFS Chemicals [ii].

2192

M.H. Chakrabarti et al. / Electrochimica Acta 52 (2007) 21892195

Fig. 2. Cyclic voltammograms recorded at 0.1 V s1 at a GC electrode in: (a)


2 mM Ru(acac)3 and 0.05 M TEABF4 in acetonitrile and (b) 0.05 M TEABF4
in acetonitrile.

3.2. Charge/discharge of the vanadium redox system


During the first charging of the VOSO4 electrolyte, the V(IV)
species must be reduced to V(II) at the cathode and oxidized to
V(V) at the anode. Consequently, as same electrolyte concentration was used in each compartment, twice as much electrolyte
(80 ml) was used in the anodic compartment during the first
charging [9]. For subsequent cycles, equal volumes of electrolyte
were used in each compartment [V(III)/V(II) and V(IV)/V(V)].
The charge/discharge reactions for the second and subsequent
charge/discharge cycles of the VOSO4 electrolyte are shown
below:
Positive half-cell [V(IV)/V(V)]:
VO2+ + H2 O  VO2 + + 2H+ + e
Negative half-cell [V(III)/V(II)]:
V3+ + e  V2+

Fig. 4. Charge/discharge potentialtime profile of 0.1 M VOSO4 solution in 2 M


H2 SO4 using graphite felt electrodes and UltrexTM anion exchange membrane.
Constant charging current of 100 mA for 300 min followed by constant discharge
at 2 mA constant current. Electrode potentials were measured relative to a SCE.

The charge/discharge of 0.1 M VOSO4 was performed using


an UltrexTM AEM. The voltage profile during the charge and discharge (second cycle) is shown in Fig. 4. The cell was charged at
100 mA for 300 min, and was discharged at 2 mA until the cell
voltage dropped to zero. A high charging current was used in
order to attempt to fully charge the cell, with around 2.3 times
the theoretical charge passed. This high charging current also led
to a relatively high voltage during charging (4 V). The open
circuit voltage after charging was high at 1.61 V, and the cell
voltage remained above 1 V during most of the discharge process. The overall efficiency was found to be 5.6% (18.4% charge
efficiency and 31% voltage efficiency) and 0.092 Wh of energy
was recovered from the 80 ml of charged electrolyte. The low
overall efficiency obtained is a consequence of the high charging current used. The charge recovered is around 86% of the
theoretical capacity, so that significant side reactions must have
occurred during charging. Furthermore the average voltage during discharge is around 1.2 V, >80% of the theoretical potential
which can be achieved with the vanadium system. In spite of
the poor overall efficiencies obtained, the discharge results indicate that the vanadium system can achieve high efficiencies, as
expected [28].
Furthermore, the results compare favourably to the allchromium system studied by Bae et al. [6] in a similar cell. The
cell potential was lower during charging in this study, but only
14% of this potential was due to ohmic drop (estimated based
on the electrolyte conductivity and cell geometry), compared to
the 50% reported for the all-chromium system [6]. The lower
ohmic loss is associated with the use of the high conductivity
2 M H2 SO4 supporting electrolyte.
3.3. Charge/discharge of Ru(acac)3 and [Fe(bpy)3 ](ClO4 )2
systems

Fig. 3. Cyclic voltammograms recorded at 0.1 V s1 at a GC electrode in: (a)


2 mM Ru(acac)3 and 0.05 M TEABF4 in acetonitrile and (b) 0.05 M TEABF4
in acetonitrile.

Charge/discharge of electrolytes consisting of ruthenium


acetylacetonate [Ru(acac)3 ] and tetraethylammonium tetrafluoroborate (TEABF4 ) were carried out in the H-type glass cell
using graphite felt electrodes and a Neosepta anion exchange
membrane. The reactions occurring at the electrodes are shown
below:

M.H. Chakrabarti et al. / Electrochimica Acta 52 (2007) 21892195

Fig. 5. Potentialtime profile during charging of 0.1 M Ru(acac)3 with 1 M


TEABF4 in acetonitrile using 1 mA constant current to 8.3% SOC in a stirred
H-type glass cell with graphite felt electrodes and Neosepta anion exchange
membrane. Electrode potentials were measured relative to a AgQRE.

2193

Fig. 7. Potentialtime profile during recharging of 0.1 M Ru(acac)3 and 1 M


TEABF4 in acetonitrile at a constant current of 1 mA to 12% SOC in a stirred
H-type glass cell with graphite felt electrodes and Neosepta anion exchange
membrane. Electrode potentials were measured relative to an AgQRE.

Positive electrode:
[Ru(acac)3 ]  [Ru(acac)3 ]+ + e
Negative electrode:
[Ru(acac)3 ] + e  [Ru(acac)3 ]
Fig. 5 shows the charging profile to 8.3% state of charge
(SOC). A rapid rise in potential occurred beyond 7.6% SOC
possibly due to a side reaction at the positive electrode or an
increasing concentration overpotential. Note that the electrode
potentials shown in Fig. 5 (and the subsequent charge/discharge
data in Figs. 610) were measured relative to a silver quasi reference electrode and hence the absolute values of the potentials are
not meaningful. However, the electrode potential data show that
the rapid rise in the cell potential was associated with an increase
in potential at the positive electrode. Fig. 6 shows the discharge
of the charged Ru(acac)3 at a constant current of 0.5 mA. The
variations in the potential were due to the addition of solvent
during the long discharging process (to make-up for solvent
evaporation). With the low currents used, a relatively high energy

Fig. 6. Potentialtime profile during discharge of the charged 0.1 M Ru(acac)3


with 1 M TEABF4 in acetonitrile in a stirred H-type glass cell with graphite felt
electrodes and Neosepta anion exchange membrane at a constant current of
0.5 mA. Electrode potentials were measured relative to an AgQRE.

Fig. 8. Potentialtime profile during discharge of the recharged 0.1 M Ru(acac)3


and 1 M TEABF4 in acetonitrile at 0.5 mA in a stirred H-type glass cell with
graphite felt electrodes and Neosepta anion exchange membrane. Electrode
potentials were measured relative to an AgQRE.

efficiency was obtained, as shown in Table 2. The open circuit


potential was significantly lower than obtained in the vanadium
system, although the state of charge was much lower in this case.
During discharge the cell potential was relatively low, at around
0.7 V.

Fig. 9. Potentialtime profile during charging of 0.05 M [Fe(bpy)3 ](ClO4 )2 and


0.5 M TEAP in acetonitrile at 0.5 mA to 3% SOC in a stirred H-type glass
cell with graphite felt electrodes and Neosepta anion exchange membrane.
Electrode potentials were measured relative to an AgQRE.

2194

M.H. Chakrabarti et al. / Electrochimica Acta 52 (2007) 21892195

Table 2
A comparison of the key results obtained from the charge/discharge experiments conducted on Ru(acac)3 and [Fe(bpy)3 ](ClO4 )2 in a stirred H-type glass cell
Electrolyte

OCP after charge (V)

Energy efficiency (%)

Cell energy output (mWh)

0.1 M Ru(acac)3 in 1 M TEABF4


0.1 M Ru(acac)3 in 1 M TEABF4 (re-charge)
0.05 M [Fe(bpy)3 ](ClO4 )2 in 0.5 M TEAP

0.85
1.30
1.50

74
57
6

11.0
12.0
0.6

The electrolytes were re-charged at 1 mA constant current


to a SOC of 12%, and the observed potentials are shown in
Fig. 7, with the corresponding discharge at 0.5 mA shown in
Fig. 8. Once again, a rapid increase in both cell and positive electrode potential was observed during charging, but this increase
was delayed by around 200 min compared to the first charging. Although the reasons for this delay are unknown, it may
have been due to an increase in solution temperature (resulting
in a decrease in the activation overpotential) or an increase in
the concentration of electroactive species due to solvent evaporation (decreased activation and concentration overpotentials).
The delay could also have been caused by an increase in the rate
of stirring, enhancing mass transport and thereby reducing the
concentration overpotential. After charging to this higher state
of charge, the open circuit potential rose to 1.3 V, however the
cell voltage during discharge was again low.
The reactions that are expected to occur at the positive
and negative half-cells during charging and discharging of the
[Fe(bpy)3 ](ClO4 )2 system are:
Positive electrode:
Fe(bpy)3 2+  Fe(bpy)3 3+ + e
Negative electrode:
Fe(bpy)3 2+ + e  Fe(bpy)3 +

The charge/discharge profiles for this system in the H-type


cell are shown in Figs. 9 and 10. A lower charging current of

0.5 mA was used, considering the lower concentration of the


active species (0.05 M due to solubility limitations). As it was
difficult to determine whether the cell potential was due to a
reversible electrochemical reaction or a side reaction, charging
was only carried out to 3% SOC.
After charging was completed, the electrolytes were discharged at a constant current of 0.1 mA (Fig. 10). It was necessary to use this low discharge current since it was found that
with higher currents the cell voltage fell rapidly to zero. It
should be noted that the lower state of charge used will lead
to a lower concentration of the active species during discharge,
so the performance would be expected to be poorer. The electrolytes in each half-cell were topped up periodically during the
experiments. The energy efficiency obtained was very low when
compared to the performance obtained from the Ru(acac)3 compound under similar conditions (see Table 2). In addition, the cell
potential obtained was less stable, falling below 0.5 V after less
than 200 min and subsequently falling to zero. The results suggest that the charged species were unstable in the electrolyte or
that the current efficiencies were very low.
Although relatively low efficiencies were observed with all
systems, this is largely associated with the H-cell design. As
has been reported for the vanadium system [711], much higher
efficiencies can be achieved with a practical flow cell design.
Quantitative comparison of the three systems is difficult since
differences in each system necessitated the use of a range of operating conditions. However, the results suggest that the Ru(acac)3
system is superior to the [Fe(bpy)3 ](ClO4 )2 system. Higher
efficiencies were obtained with this system and around 2030
times more energy was recovered during discharge. Although the
[Fe(bpy)3 ](ClO4 )2 offers the possibility of higher open circuit
potentials and lower cost, low energy efficiencies were observed,
probably due to low current efficiencies. Because of the potential advantages of the [Fe(bpy)3 ](ClO4 )2 , it is recommended that
the causes and possible remedies for the low current efficiencies
be investigated. In addition, further studies with a flow cell with
improved transport conditions and lower ohmic losses should be
carried out to determine the viability of both systems.
4. Conclusions

Fig. 10. Potentialtime profile during discharge of the charged 0.05 M


[Fe(bpy)3](ClO4)2 and 0.5 M TEAP in acetonitrile solution at 0.1 mA in a stirred
H-type glass cell with graphite felt electrodes and Neosepta anion exchange
membrane. Electrode potentials were measured relative to an AgQRE.

Metalorganic species may offer high efficiency, high cell


potential systems for redox flow battery applications. Results in a
simple H-type cell indicate that high efficiencies can be achieved
with a ruthenium acetylacetonate system, which has high solubility and stability in an acetonitrile electrolyte. Since the two
redox couples revert to the same species on discharge, cross-over
will not reduce the cycle life time and complex electrolyte reprocessing is not required. Charging the system generates species

M.H. Chakrabarti et al. / Electrochimica Acta 52 (2007) 21892195

with opposite charge, so some loss of efficiency will occur by


transport through either cation or anion exchange membrane
materials. However, a redox flow battery utilizing a low cost
microporous membrane can be envisaged. Further evaluation in
a flow cell is recommended.
Acknowledgements
Funding for this research was provided by the Engineering
and Physical Sciences Research Council (EPSRC). The authors
would also like to give special thanks to Dr. N. Stevens and Dr.
C.H. Bae for their valuable input in this work.
References
[1] I.M. Ritchie, O.T. Siira, Eightth Biennial Congress of International Solar
Energy, 1983, p. 1732.
[2] M. Bartolozzi, J. Power Sources 27 (1989) 219.
[3] M. Skyllas-Kazacos, F. Grossmith, J. Electrochem. Soc. 134 (1987) 2950.
[4] A. Paulenova, S.E. Creager, J.D. Navratil, Y. Wei, J. Power Sources 109
(2002) 431.
[5] C.H. Bae, Ph.D. Thesis, University of Manchester Institute of Science and
Technology, UK, 2001.
[6] C.H. Bae, E.P.L. Roberts, R.A.W. Dryfe, Electrochim. Acta 48 (2002) 279.
[7] E. Sum, M. Rychcik, M. Skyllas-Kazacos, J. Power Sources 16 (1985) 85.
[8] E. Sum, M. Skyllas-Kazacos, J. Power Sources 15 (1985) 179.
[9] M. Skyllas-Kazacos, M. Rychcik, R.G. Robins, A.G. Fane, M.A. Green, J.
Electrochem. Soc. 133 (1986) 1057.

2195

[10] M. Skyllas-Kazacos, D. Kasherman, D.R. Hong, M. Kazacos, J. Power


Sources 35 (1991) 399.
[11] M. Skyllas-Kazacos, M. Rychick, R. Robins, US Patent 4,786,567, US,
1988.
[12] T. Yamayura, N. Watanabe, Y. Shiokawa, J. Alloys Compd. 408 (2006)
1260.
[13] D.A.J. Rand, R. Woods, R.M. Dell, Batteries for Electric Vehicles, John
Wiley & Sons Inc., NY, 1994.
[14] I. Tsuda, K. Nozaki, K. Sakuta, K. Kurokawa, Sol. Energy Mater. Sol. Cells
47 (1997) 101.
[15] A. Shibata, K. Sato, Power Eng. J. 13 (1999) 130.
[16] C. Fabjan, J. Garche, B. Harrer, L. Jorissen, C. Kolbeck, F. Philippi, G.
Tomazic, F. Wagner, Electrochim. Acta 47 (2001) 825.
[17] VRB Power Systems Inc., http://www.vrbpower.com/, 2006.
[18] Y.D. Chen, K.S.V. Santhanam, A.J. Bard, J. Electrochem. Soc. 128 (1981)
1460.
[19] M. Morita, Y. Tanaka, K. Tanaka, Y. Matsuda, T. Matsumura-Inoue, Bull.
Chem. Soc. Jpn. 61 (1988) 2711.
[20] A. Pighin, B.E. Conway, J. Electrochem. Soc. 122 (1975) 619.
[21] Y. Matsuda, K. Tanaka, M. Okada, Y. Takasu, M. Morita, T. MatsumuraInoue, J. Appl. Electrochem. 18 (1988) 909.
[22] A. Endo, Y. Hoshino, K. Hirakata, Y. Takeuchi, K. Shimizu, Y. Furushima,
H. Ikeuchi, G.P. Sato, Bull. Chem. Soc. Jpn. 62 (1989) 709.
[23] K. Itoh, K. Honda, M. Sukigara, J. Electroanal. Chem. 110 (1980) 277.
[24] N. Tokel, A.J. Bard, J. Am. Chem. Soc. 94 (1972) 2862.
[25] H. Kaneko, K. Nozaki, Y. Wada, T. Aoki, A. Negishi, M. Kamimoto, Electrochim. Acta 36 (1991) 1191.
[26] S. Zhong, M. Skyllas-Kazacos, J. Power Sources 39 (1992) 1.
[27] M. Gattrell, J. Park, B. MacDougal, J. Apte, S. McCarthy, C.W. Wu, J.
Electrochem. Soc. 151 (2004) A123.
[28] M. Rychcik, M. Skyllas-Kazacos, J. Power Sources 19 (1987) 45.

You might also like