Você está na página 1de 21

Ruijie Ye

Transfercenter Sustainable Electrochemistry,


Saarland University,
Saarbr€ucken 66125, Germany;
Bio Sensor and Materials Group,
KIST Europe,
Campus E7 1,
Saarbr€ucken 66123, Germany

Dirk Henkensmeier
Fuel Cell Research Center,
Korea Institute of Science and Technology,
Seoul 02792, South Korea;
ET-GT,
University of Science and Technology,
Seoul 02792, South Korea;
Green School,
Korea University,
Seoul 136-713, South Korea

Sang Jun Yoon


Transfercenter Sustainable Electrochemistry,
Saarland University,
Saarbr€ucken 66125, Germany;
Bio Sensor and Materials Group,
KIST Europe,
Campus E7 1,
Saarbr€ucken 66123, Germany;
Center for Membranes,
Advanced Materials Division,
Korea Research Institute of Chemical Technology,
Daejeon 34114, South Korea
Redox Flow Batteries for Energy
Zhifeng Huang
Transfercenter Sustainable Electrochemistry,
Saarland University,
Storage: A Technology Review
Saarbr€ucken 66125, Germany;
Bio Sensor and Materials Group, The utilization of intermittent renewable energy sources needs low-cost, reliable energy
KIST Europe, storage systems in the future. Among various electrochemical energy storage systems,
Campus E7 1, redox flow batteries (RFBs) are promising with merits of independent energy storage and
Saarbr€ucken 66123, Germany
power generation capability, localization flexibility, high efficiency, low scaling-up cost,
Dong Kyu Kim and excellent long charge/discharge cycle life. RFBs typically use metal ions as reacting
Transfercenter Sustainable Electrochemistry,
Saarland University,
species. The most exploited types are all-vanadium RFBs (VRFBs). Here, we discuss the
Saarbr€ucken 66125, Germany; core components for the VRFBs, including the development and application of different
Bio Sensor and Materials Group, types of membranes, electrode materials, and stack system. In addition, we introduce the
KIST Europe, recent progress in the discovery of novel electrolytes, such as redox-active organic
Campus E7 1,
Saarbr€ucken 66123, Germany;
compounds, polymers, and organic/inorganic suspensions. Versatile structures, tunable
Department of Mechanical and properties, and abundant resources of organic-based electrolytes make them suitable for
Aerospace Engineering, cost-effective stationary applications. With the active species in solid form, suspension
Seoul National University,
electrolytes are expected to provide enhanced volumetric energy densities.
Seoul 08826, South Korea
[DOI: 10.1115/1.4037248]
Zhenjun Chang
Transfercenter Sustainable Electrochemistry,
Saarland University,
Saarbr€ucken 66125, Germany;
Bio Sensor and Materials Group,
KIST Europe,
Campus E7 1,
Saarbr€ucken 66123, Germany;
College of Materials Science and Engineering,
Jiangsu University of Science and Technology,
Zhenjiang 212003, China

Sangwon Kim
Transfercenter Sustainable Electrochemistry,
Saarland University,
Saarbr€ucken 66125, Germany;
Bio Sensor and Materials Group,
KIST Europe,
Campus E7 1,
Saarbr€ucken 66123, Germany
1
Ruiyong Chen
Transfercenter Sustainable Electrochemistry,
Saarland University,
Saarbr€ucken 66125, Germany;
Bio Sensor and Materials Group,
KIST Europe,
Campus E7 1,
Saarbr€ucken 66123, Germany
e-mail: r.chen@kist-europe.de

1
Corresponding author.
Manuscript received May 15, 2017; final manuscript received July 5, 2017;
published online September 19, 2017. Assoc. Editor: Kevin Huang.

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-1
C 2018 by ASME
Copyright V

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


1 Introduction as: porous membranes and dense ion-conducting membranes.
The latter can be further separated into three categories: cation
Electrochemical energy storage is currently attracting signifi-
exchange membranes (CEMs), anion exchange membranes
cant interest, considering the expanding market for portable elec-
(AEMs), and membranes which conduct both cations and anions.
tronics, electric vehicles, smart grid, and off-grid energy storage.
Historical development of membranes for flow batteries has been
Among various energy storage systems, redox flow batteries
reviewed elsewhere [10,11]. In this section, we focus on the very
(RFBs) are promising for stationary large-scale applications in
recent development, highlighting sulfuric-acid-doped polybenzi-
terms of cost, reliability, and safety [1]. RFBs consist of cathode
midazole (PBI) membranes, which seem to be promising for
and anode chambers, membranes, and flowable electrolytes.
VRFBs [12,13]. Another example is the use of spongelike porous
Energy is converted in electrochemical reactions and stored by
membranes, in which the pores are filled with highly conductive
keeping the active species of the electrolytes in external tanks.
sulfuric acid and the thin polymeric pore walls can block transport
The energy density of RFBs is determined by the volume of the
of vanadium ions [13–17].
electrolytes, concentration of active species, the cell voltage, and
the number of stacks. The power generation capability is related
to the kinetic behavior of redox-active species and the size of the 2.1 Porous Membranes. Earlier development of flow bat-
teries was restricted by the availability of suitable membranes.
electrodes. Such characters make the RFBs flexible energy storage
systems. For example, a diaphragm was prepared by drilling holes in a
Various metal-based redox couples have been investigated over Plexiglass disk (60 holes of 0.6 mm diameter per 1 cm2) [18].
Such diaphragm cannot prevent mechanical mixing of two elec-
the past few decades [2]. Among them, all-vanadium RFBs
(VFRBs) are the most investigated. Sulfuric acid is typically used trolytes, resulting in low CE. Nowadays, well-defined porous
membranes are commercially available, such as microfiltration
as supporting electrolyte. The maximal energy density of VRFBs
is only about 30 Wh L1. This is limited by the solubility and sta- membranes (block particles above 100 nm), ultrafiltration mem-
bility of the four involved vanadium ions (VO2þ, VO2þ, V3þ, and branes (block particles down to 2 nm), and nanofiltration mem-
branes (a cut-off below 2 nm) [19]. Ultrafiltration membranes can
V2þ) and can be slightly enhanced by using mixed
sulfate–chloride electrolytes [3]. Membranes with low vanadium be used for flow batteries with polymeric redox couples. For
example, a commercial Celgard 2400 microporous membrane
permeability and thus high Coulombic efficiency (CE) and high
energy efficiency (EE) are required. Crossover of vanadium ions (25 lm thick, 39% porosity, and 28 nm average pore diameter)
through the membranes will reduce the cell efficiency and require rejected 86.3% of a redox-active polymer when the molecular
weight of the polymer was 318 kDa [20]. Schubert and coworkers
remixing of electrolytes. Low-cost and high-performance mem-
branes are of great importance. Carbon-based electrodes are gen- developed an aqueous solution of polymers [21], allowing the use
of a cellulose-based dialysis membrane.
erally used to catalyze the VO2þ/VO2þ and V3þ/V2þ redox
For nanofiltration membranes, the size exclusion effect can
reactions. Rational processing and modification of the carbon
electrode can increase the surface area and the number of active even be used to separate protons and vanadium ions [22,23]. For
this purpose, a poly(acrylonitrile) membrane was prepared by a
functional groups on the surface [4,5]. Catalyst particles on the
carbon electrode surface can further promote the kinetics and phase inversion process [22]. Based on the scanning electron
microscope (SEM) images, the preparation process probably
enhance the voltage efficiency (VE) [6]. To enhance the energy
involved the casting of the membrane on a substrate (e.g., glass
storage capability of RFBs, it is necessary to find alternative
energy-dense electrolytes (high concentration of active species plate), followed by immersion in a nonsolvent, e.g., water. Mem-
branes prepared in this way usually show an anisotropic morphol-
and with high cell voltage [7]). Recently, efforts have been
devoted to energy storage systems that use low-cost, environmen- ogy, consisting of large pores and a dense, thin selective layer.
tally safe organic and polymer electrolytes [8] and suspension Critical process parameters include the choice of solvents, proc-
essing temperature, and polymer concentration in the casted mem-
electrolytes with solid active species [9]. Optimization of the
VRFBs and the development of new electrolyte and cell systems brane. In order to improve the wetting behavior of the membrane
surface, the nitrile groups of the porous membranes were partially
for a practical application need to overcome many challenges,
including the fundamental understanding of the electrolyte elec- hydrolyzed by immersion in 10 wt. % NaOH solution. A
trochemistry, development of improved electrode materials, and polyethersulfone-based nanofiltration membrane was further
developed [23]. Different pore structures were obtained by adding
innovation in the critical aspects of cell and stack design.
In this review, we address the critical issues and challenges that a water soluble polymer to the casting solution. CE in the range of
90–95% has been achieved with an EE of 75% at 80 mA cm2.
are related to the performance of RFBs. In Sec. 2, we discuss the
types and properties of membranes and their impact on the VE, Recently, ultrafiltration membranes were also considered for
CE, and EE for the VRFBs. In Sec. 3, we introduce the modifica- the use in VRFBs. Wessling and coworkers [24] coated a porous
polyacrylonitrile membrane with either a 0.75- or a 2.8-lm thick
tion of carbon-based electrode materials, the application of cata-
lysts, and their effect on cell performance. The main focus of Sec. PIM-1 (a polymer of intrinsic microporosity) layer. PIM-1 was
designed to have a high free volume with a rigid, kinked polymer
4 is on organic redox-active species, which possess great potential
backbone (Fig. 1) [25]. The nanosized space between the polymer
to meet the low-cost and high-power requirements. Section 5 dis-
cusses novel electrolytes using solid redox-active compounds in chains is rather hydrophobic, limiting its water absorption and
blocking large vanadium ions. Protons can enter the polymer and
semi-solid RFB systems, which can allow for large improvements
in the energy density. Section 6 presents system configurations for interact with the water molecules in the nanopores of PIM-1. The
RFBs. In the end, we present concluding remarks and prospects CE of these membranes is hardly affected by the current density,
which is around 100% at open circuit voltage and a little lower at
for the future development of RFBs.
current densities up to 40 mA cm2. The VE is similar to Nafion
112. In order to reduce the porosity and enhance the conductivity,
some strategies could be helpful such as hydrolysis of the nitrile
2 Membranes group [26] or copolymerization with less rigid monomers. Note
In flow batteries, membranes separate the two electrode cham- that PIM-1 has poor acid resistance. The structure of the spiro
bers, preventing mixing of anolyte and catholyte. They are perme- compound is complicated, but it can be easily obtained by acid-
able for supporting ions, such as protons or sulfates. However, catalyzed condensation of acetone and catechol. The back reaction
membranes should be impermeable for the reactive species. Mem- of the monomer synthesis opens a pathway for degradation of the
branes should be chemically stable against oxidation and acid- polymer backbone under acidic conditions [27].
catalyzed reactions like hydrolysis, which excludes polymers with In conclusion, porous membranes can effectively block permea-
ester bonds, for example. In general, membranes can be classified tion of the redox compounds through the size exclusion effect.

010801-2 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Further optimization of these membranes could be obtained by performance. Membranes casted from commercial dispersions are
modifying the pore structure, adding inorganic fillers, or blending brittle, unless the solvents are evaporated at a minimum tempera-
with more hydrophilic materials (to enhance proton conductivity). ture of about 60  C [35]. The brittleness can also be reduced by an
annealing step, at which the polymer is heated above its glass
transition temperature [36]. For the use for RFBs, the requirement
2.2 Cation Exchange Membranes. CEMs consist of poly-
of mechanical stability is not as critical as for fuel cells, consider-
mers substituted with acidic groups. For weak ion exchange mate-
ing that membranes (ideally) are always in contact with the liquid
rials, which are less relevant to flow batteries, these groups can be
electrolyte. Future work therefore may focus on membranes cast-
carboxylic acid groups. For strong ion-exchange materials, the
ing under different conditions or solvents [37], in order to reba-
functional groups are mainly sulfonic or phosphonic acids, which
lance the membrane properties (conductivity, permeability, and
easily dissociate into a mobile cation and the respective immobi-
mechanical stability).
lized counter ion. Due to different hydrophilicities of the back-
Many hydrocarbon-based CEMs with a large structural variety,
bone (hydrophobic) and the functional groups (hydrophilic),
developed previously for fuel cells, can be considered for their
CEMs show phase separation into hydrophobic domains (which
use in flow batteries. Structurally, most hydrocarbon-based mem-
provide mechanical integrity) and connected hydrophilic domains
branes are derived from condensation polymers like poly(ether
(which provide conductivity). This phase separation is especially
ether sulfone) or poly(ether ether ketone). Design parameters are
pronounced in perfluorinated materials like Nafion (Fig. 2) [28].
the number, position, and distribution of sulfonic acid groups. Sul-
In a strong contrast to typical porous membranes discussed in Sec.
fonation of polymers leads to a random distribution over the
2.1, the shape and size of the hydrophilic domains for CEMs may
chains and connecting points are activated positions (e.g., ortho to
vary, depending on the water contents, and also show dynamic
ether groups). The use of sulfonated monomers allows controlling
reorientations due to mobility of the polymer chains. CEMs have
the degree of sulfonation and the position of the sulfonic acid
several advantages such as high conductivity, commercial avail-
groups. The distribution of the functional groups can be controlled
ability (mainly developed for chlor-alkali electrolysis, fuel cells,
by changing from random copolymers to block copolymers, which
electrolyzers, and water treatment), high chemical stability (e.g.,
has a large influence on the microstructure and morphology and
Nafion), and structural variety. The main drawback of CEMs is
therefore on the water uptake and conductivity [38]. Another way
that they are susceptible to metal crossover, e.g., vanadium ions.
to enhance the phase separation is to use spacer groups between
During operation, especially at high current densities, this effect is
the aromatic backbone and the sulfonic acid groups. Kim and
less pronounced because protons have a higher diffusivity than
coworkers [39] prepared a polymer in which some of the hexa-
metal ions. At open circuit voltage, cells with a CEM will rapidly
flourinated bisphenol A units bear two bis(propylsulfonic acid)-
self-discharge.
amine substituents. As expected, both conductivity and vanadium
If CEMs are in the protonated form, they will easily show much
permeability increased with the degree of sulfonation. However, a
higher conductivity than a membrane which selectively conducts
membrane with a thickness close to Nafion 117 showed 50%
anions like Cl or SO2 4 , considering protons have the highest lower permeability, but 10% higher conductivity than Nafion 117.
molar conductivity among all known ions [29]. Protons are trans-
Stability test in 1 M VO2þ suggested a good stability with 1.5%
ported not only via the vehicle mechanism (a moving charge) but
weight loss. In another work, Fujiwara and coworkers [40] com-
also via structural diffusion (Grotthus mechanism). In the latter
pared the effect of three different ways to connect the acid groups:
case, charge is not transported by individual protons between the
(1) directly attached to an aromatic ring, (2) with an allylic spacer,
electrodes, but transported by a reorientation of bonds between
and (3) attached as styrenesulfonic acid. Most tested membranes
hydronium ions (H3Oþ) and a neighboring water molecule. For
showed an EE of up to 86%, higher than that of Nafion 212. Espe-
some flow batteries, also the ability of CEMs to conduct Naþ can
cially, the membranes based on styrenyl sulfonic acid showed
be useful.
slightly higher proton conductivities and performed well in the
The most prominent CEM membrane is Nafion, which is avail-
VRFB, indicating that the longest side group has the most benefi-
able in different thicknesses (25, 50, 125, and 175 lm for Nafion
cial impact on the properties. Kwon and coworkers [41] analyzed
211, 212, 115, and 117, respectively). The dependence of VE, CE,
the effect of the membrane thickness (30–150 lm, sulfonated pol-
and EE on the thickness in flow batteries has been well investi-
y(ether ether ketone) (sPEEK) with a degree of sulfonation of
gated [30,31]. In general, thin membranes increase the VE (low
55%) on the VRFB performance. As expected, the VE decreases
resistance), and thick membranes increase the CE (reduced vana-
and the CE increases with the thickness. However, all the tested
dium ion crossover). Between 120 and 240 mA cm2, Nafion 115
membranes showed higher EE than Nafion 212, due to the higher
showed the highest EE, balancing CE and VE [31]. Further opti-
CE (around 95%) of sPEEK membranes than that of Nafion 212
mization is possible by controlling the hygrothermal history of the
(below 80%).
membranes, which influences the degree of crystallinity and chan-
nel structure [32]. Therefore, when handling the membranes, care
needs to be taken to follow an exact protocol. For example, the
water content of a membrane during assembly can have an effect
on the swelling direction (in-plane versus thickness direction) of
the membrane in the assembled cell [33,34]. Also, the preparation
conditions are expected to have a major effect on the membrane

Fig. 2 Structure of Nafion and the cluster-network model:


hydrophilic clusters connected by short narrow channels, short
Fig. 1 Structure of PIM-1, a polymer of intrinsic microporosity curves: Nafion side chains, and dots: sulfonic acid groups [28]

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-3

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Similar to Nafion, the properties of hydrocarbon-based mem- respective solutions [53]. In this approach, the PEI layer bears a
branes can be strongly influenced by the casting solvents. By positive charge on the polymer backbone, which effectively
varying the degree of sulfonation and casting sPEEK membranes blocks vanadium ions, improving the CE from 81% to 93% and
from N,N-dimethylformamide (DMF), N,N-dimethylacetamide the EE from 41.5% to 45.2%. Apparently, the alternating structure
(DMAc), N-methyl-2-pyrrolidone (NMP), and dimethylsulfoxide of layers with positively and negatively charged groups prevents
(DMSO) solutions, Xi et al. [42] found that the membrane with a delamination. A structure without delamination was presented by
degree of sulfonation of 67%, casted from DMF, provided Zhai and coworkers [54]. The authors immersed a Nafion 117
the best efficiencies and a better cycle-life performance at membrane in an acetone solution containing N,N-dimethylami-
80 mA cm2. noethylmethacrylate and subjected it to c-ray irradiation from a
60
It was shown that sulfonated hydrocarbon-based membranes Co source. The radiation degrades Nafion, leading to formation
usually lead to a higher CE than Nafion. As discussed, ultrafiltra- of macromolecular radicals, which act as starting point for free
tion membranes can have high proton/vanadium selectivity, based radical polymerization of the acrylic monomer. If the reaction
on the molecular sieving effect. For comparison, if we consider stops before the reaction proceeds throughout the membrane, the
CEMs as ultrafiltration membranes with the pores as hydrophilic grafted chains are mainly located on the membrane surface.
channels, Kreuer’s comparison of the morphologies between Unreacted monomers and polymer chains, which are not cova-
Nafion and sPEEK membranes (Nafion has wider channels than lently attached to the membrane, need to be carefully extracted
sPEEK) may provide explanation for the enhanced CE [43]. It is with acetone. By applying such vanadium blocking layer, an opti-
plausible that such a sieving effect should be further enhanced by mized membrane showed half the conductivity of Nafion 117, but
the more rigid structure of hydrocarbon-based membranes (lower also just 17% of the vanadium permeability, resulting in a two
chain mobility). This finding is suggestive to optimize Nafion times increase in selectivity. Note that the acrylic monomer con-
membranes. Wang and coworkers [44] analyzed Nafion mem- tained ester groups, which are hydrolytically instable and will
branes with an equivalent weight of 1000, 1200, and 1500 g likely degrade over time in the highly acidic environment of the
mol1 Hþ by small-angle X-ray scattering (SAXS) and calculated VRFB.
that the average pore size decreases from ca. 3.3 nm to 3.15 and All the membranes discussed earlier were prepared by solution
3.05 nm, respectively. In the VRFB, a 47-lm thick EW 1500 casting, which implies that all the components are freely miscible
membrane showed the highest capacity retention, but a lower VE. and soluble in the casting solvent. Radiation grafting is a way to
When the membrane was produced with just 31 lm thickness, it circumvent this limitation. For example, partially fluorinated ethyl-
showed an area resistance comparable to Nafion 115 (EW 1100), ene tetrafluoroethylene (ETFE) membranes are commercially
but a 40% lower VO2þ ion flux. available at a relatively low price. They are nearly insoluble in
A very effective way to improve CEMs is to blend them with common solvents. By electron-beam irradiation, which is a well-
small amounts of a basic polymer. As an example, Kerres and established industrial scale process, radicals are formed throughout
coworkers [45] blended a partially fluorinated, sulfonated polymer the whole membrane. When such an activated membrane is
(based on bisphenol A and biphenyl) with a partially fluorinated immersed in a monomer-containing solution, the monomers start to
PBI. The strong ionic interaction between the acidic and basic diffuse into the membrane and then to polymerize. Usually, the
groups leads to ionic crosslinking, which reduces water swelling crystalline areas of the ETFE substrate absorb much less (or no)
and therefore the average pore size. Accordingly, such membrane monomers, leading to a unique morphology of crystalline,
showed a higher CE (close to 100%) than Nafion 117 and a good unreacted ETFE domains (enhancing mechanical stability) and
stability over 200 cycles. amorphous, grafted domains [55]. By choosing the right combina-
Composite membranes with added components, which are not tion of monomers, copolymers can be also grafted. Gubler and
miscible with polymers, have been developed for use in fuel cells coworkers [56] grafted ETFE films with a styrene/acrylonitrile
[46]. Such strategies have been applied to RFBs. For instance, copolymer. The aromatic groups were sulfonated, and some of the
2.5–10 wt. % of Al2O3, SiO2, or TiO2 nanoparticles were mixed nitrile groups were transferred into amidoxime groups (Fig. 3),
into a sPEEK membrane [47]. Since the particles do not provide which are known to complex vanadium ions. However, the com-
acidic groups, the ion-exchange capacity (IEC) decreases with plexes are not stable under acidic conditions [57]. Therefore, the
increasing the amount of nanoparticles. This leads to decrease of observed good membrane properties (similar conductivity to
water uptake, swelling ratio, and proton conductivity. At a load of Nafion 117, but five times lower VO2þ uptake, resulting in higher
5 wt. %, the addition of nanoparticles significantly decreased the VE and CE than Nafion 212 and 117) are probably a result of elec-
VO2þ permeability of the sPEEK membrane. When the thickness trostatic repulsion of vanadium ions by the protonated amidoximes.
was adjusted to 60–70 lm, the composite membranes showed not
only higher CE but also higher VE than Nafion 117. The stability
in 0.1 M VO2þ/3 M sulfuric acid was good (no visible degradation 2.3 Anion Exchange Membranes. For AEMs, the positively
over 21 days). The concentration of V5þ was very low in this test charged functional groups electrostatically repel cations. This will
(others use 1.7 M [48]), but even 0.1 M is enough to strongly inhibit the crossover of vanadium ions. But in spite of the Donnan
degrade sulfonated Radel [49]. The discharge capacity retention exclusion effect, AEMs were found to absorb some additional sul-
of the membrane with 5 wt. % silica was 53.3%, which was more furic acid when immersed in the electrolytes, which leads to
than twice the value observed for Nafion 117. In another work, higher conductivity than expected for a pure membrane [58].
Nafion was filled with silica nanoparticles with a sulfonated sur- Ramani and coworkers [59] tested the suitability of AEMs for the
face [50]. The additional sulfonic acid groups increased the IEC V/Ce RFBs. The results showed that the employed trimethylam-
from 0.99 to 1.24 mmol g1. Both Nafion 117 and a composite monium functionalized poly(etherketone)-based AEM is chemi-
membrane (similar thickness) showed the same VE of 87%, while cally and mechanically stable and can prevent mixing of the
the composite membrane showed a CE of 94%, compared to 82% active species. After 20 cycles, the capacity retention of a system
for Nafion 117. Xi and coworkers [51] mixed graphene and gra- with Nafion 212 was 55%, while the capacity loss was <5% for
phene oxide into Nafion. Graphene is impermeable for most mole- the AEM-based system.
cules, but permeable for protons, which could boost the The main challenge on AEMs is the low alkaline stability of the
membrane selectivity [52]. In comparison to recast Nafion, the cationic groups, which are typically quaternary ammonium
composite membranes improved self-discharge and EE at 80 mA groups. Marino and Kreuer [60] reported that a spiro-ammonium
cm2 (from 80% to 85%). model compound has a 1000 times longer half-life in 6 M NaOH
Another approach is to coat membranes with a thin blocking at 160  C than the 1,2,3-trimethylimidazolium ion. Another func-
layer. This can be a single layer or alternating layers of Nafion tional group, which is considered unstable, is the 1-methyl pyri-
117 and polyethyleneimine (PEI) by dipping it repeatedly in dinium group [61]. However, under acidic conditions, it becomes

010801-4 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 3 Membrane preparation in a pre-irradiation grafting process. Bifunctionalized ETFE-graft-
poly(styrene-co-acrylonitrile) membranes were obtained via activation by electron-beam
irradiation, grafting of styrene and acrylonitrile, amidoximation, and sulfonation. Under the
acidic conditions, the amidoxime groups are probably in protonated, positively charged form.

attractive again to investigate AEMs with these functional poly(phenylene)-based AEM in VRFBs (Fig. 4). It is known that
groups. Zhang et al. [62] modified chloromethylated Radel with CEMs based on a similar structure degraded in the VRFBs,
pyridine. Zhang and coworkers [63] prepared crosslinked despite the fully aromatic structure. The similar materials, func-
membranes by adding bipyridine to a casting solution containing tionalized with quaternary ammonium groups, showed an
chloromethylated Udel. Zhao and coworkers [64] followed a very improved stability in VO2þ solution. AEMs with an intermediate
similar protocol, but crosslinked bromomethylated PPO (poly IEC value (0.8 mmol g1) showed the lowest vanadium perme-
(2,6-dimethyl-1,4-phenylene oxide) with 1,2-bis(4-pyridyl)ethane, ability (1.4  106 cm2 min1 versus 2.1  106 cm2 min1 for
which is a more flexible linker than bipyridine. All three groups Nafion 212) and a low area resistance of 0.21 X cm2 (compared to
reported excellent chemical stability, high CE, and good VE. Lu 0.32 X cm2 for Nafion 212). In general, the IEC value was shown
et al. [65] attached methylated imidazolium groups to a bromome- to be proportional to degradation rate and VE and negatively cor-
thylated poly(ether ether sulfone) backbone. Membranes with an related to the CE. A direct comparison between a CEM and an
IEC of 1.5 and 2.0 showed a very low permeability. A membrane AEM, both based on Diels–Alder polymers, showed unexpected
with a high IEC of 3.5 showed a permeability close to Nafion 117. findings [70]. Some sulfonated polymers showed a similar CE as
Another big challenge for development of hydroxide- the corresponding AEM, similar to that with Nafion 117. It is con-
conducting membranes is the alkaline degradation of the polymer sidered that the transport processes are strongly affected by water
backbone. It is considered that polymers with only C–C bonds uptake. The bulky, rigid structure of the Diels–Alder polymer
would be more stable than polymers with heteroatoms in the main chains may lead to a high free volume, limiting the effect of the
chain [66]. Similarly, for an Udel membrane functionalized and ionic groups. Finally, a Diels–Alder polymer, in which the ammo-
crosslinked by bipyridine, it was postulated that VO2þ attacks the nium groups are tethered by C6-spacers, was tested for application
ether groups, leading to chain scission [67]. For a sulfonated in a nonaqueous ionic liquid-based flow battery [71]. Membranes
Radel, Hickner and coworker proposed a mechanism which with an IEC of 1.5 were too brittle to be useful, while membranes
involves attack of a V-radical on the electron-rich phenyl rings of with an ICE of 2.5 showed excessive movement of solvent across
the biphenyl unit (which is activated by the ether groups in posi- the membrane. Therefore, an IEC of 2.0 was mostly chosen. No
tion 4 and 40 ) [48]. This suggests that for more stable polymer chemical degradation was found. An increase in membrane resist-
backbones, it is necessary to reduce the electron density of the ance was observed, probably due to mechanical and thermal
phenyl rings, to enhance phase separation between the ionic aging.
groups and the hydrophobic backbones, and to prevent access of The strategy to embed metal oxide nanoparticles into mem-
the reactive species to the backbone. Recently, Zhang et al. [68] branes was also applied to AEMs. Recently, Ramani and
prepared poly(ether ether ketone)s with an adamantane moiety in coworkers [72] added a quaternized silesquioxane additive to a
the backbone. They hypothesized that the adamantane segments quaternized cardo-poly(ether ketone) membrane. The additive
restrict water uptake and membrane swelling, leading to a high was formed in the casting solution, by adding N-(trimethoxysilyl-
stability in the VRFB. As expected, all the AEMs showed a factor propyl)-N,N,N-trimethylammonium chloride to the polymer solu-
about 20 times lower vanadium crossover than Nafion 117. In an tion. Positively charged groups were added to increase the
ex situ stability test in a 0.15 M VO2þ solution, a higher IEC conductivity, while maintaining chemical, mechanical, and ther-
value leads to faster degradation. In terms of efficiencies, the mal stability. Most importantly, by introducing additives and nar-
AEMs with an IEC of 1.8 showed a slightly improved CE and EE. row channels (which form around the nanoparticles) and
Polymers without heteroatoms in the main chain have also been increasing the number of positively charged groups in the chan-
studied. Fujimoto and corkers [69] evaluated a Diels–Alder nels, it is expected that the vanadium crossover can be further

Fig. 4 A Diels–Alder poly(phenylene) with quaternary ammonium groups (left) and sulfonic
acid groups (right)

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-5

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


reduced. By adding up to 40 wt. % of the additives, the IEC was 2.4 Acid-Doped Membranes. Although pure PBI is an insu-
increased from 1.11 up to 1.28 meq Cl g1. This increased the lating material, the basic nitrogen groups of PBI strongly interact
water uptake and the sulfate conductivity, but decreased the VO2þ with strong acids (such as phosphoric acid or sulfuric acid), ren-
permeability. This is in strong contrast to most other studies, dering the acid-doped materials conductive (Fig. 6) [73,74]. It is
which reported that diffusion increases with the IEC. Because the expected that protons and anions (the conjugated base of the dop-
additive increased conductivity and reduced vanadium crossover, ing acid) contribute to the conductivity [75]. Through interaction
a membrane with 20% filler reached an EE of 93% at 20 mA with the first strongly bound acid molecules, additional acid mole-
cm2. However, in an ex situ accelerated test, the conductivity cules can be absorbed by hydrogen bond interactions. Zhao and
decreased within a few days, when immersed in 1.5 M VO2þ/3 M coworkers [76] argued that PBI used in the VRFBs can be consid-
H2SO4. Further analysis by nuclear magnetic resonance (NMR) ered as a porous membrane with pore diameters of 0.5–2 nm,
spectroscopy revealed that the ammonium groups were detached which is much smaller than the size of the ionic clusters in Nafion
from the propyl linkers of the additive. Addition of charged addi- (connected by small channels of 1 nm diameter, see Fig. 2). Note
tives is very promising, but different chemical structures need to that PBI will be fully protonated when immersed in electrolyte.
be tested. Thus, the polymer backbone has a high positive charge density,
As will be discussed later for acid-doped membranes, mem- which can hinder transport of vanadium ions through electrostatic
branes with closed pores (spongelike structures) can be beneficial repulsive forces. In addition, acid absorption will increase the
for use in the VRFB, if the pores are filled with sulfuric acid. pore size, as water uptake does for Nafion membranes [44]. PBI
Then, the pores allow for an overall low resistance, and the poly- membranes showed excellent properties in the VRFB. After
meric pore walls block transport of vanadium ions. Zhang and immersion for 3 months in 1 M V5þ solution, a PBI membrane
coworkers [14] prepared a porous membrane from a chloromethy- showed a weight loss of 2.9%, similar to Nafion (2.1%). The con-
lated polysulfone (CMPSF) solution by water-vapor-induced ductivity of Nafion decreased 5.6%, while PBI decreased only
phase separation, followed by immersion of the membranes in an 4.4% in contrast. No permeation of vanadium ions was measura-
imidazole containing solution for 1–4 days. Due to the heteroge- ble in ex situ tests. In the VRFBs, the CE of a 30-lm thick PBI
neous character of this reaction, reaction with imidazole should membrane was close to 100% at very low current densities. The
proceed from outside to inside. The VE improves for membranes bottleneck is the low conductivity of sulfuric-acid-doped PBI
treated for 1, 2, or 3 days, after which the VE reaches a constant (15.8 mS cm1 versus 50.7 mS cm1 for Nafion 211). As a conse-
value. The membrane immersed in the imidazole solution for 3 quence, the EE of PBI exceeds that of Nafion 211 only at current
days showed a CE of 99% and an EE of 86% at 80 mA cm2. Fur- densities below 40 mA cm2.
thermore, when the membrane was tested over 6000 cycles at By changing the backbone structure of PBI (BIpPBI, see
120 mA cm2, no efficiency loss was observed (Fig. 5). Very sim- Fig. 6), Hong and coworkers [77] obtained a membrane that can
ilar results were obtained, when the imidazole was substituted by absorb about 25% more acid than standard meta-PBI, when
1,4-butanediamine [15]. immersed in the VRFB electrolyte. It has been reported that the

Fig. 5 (a) and (b) Cross-sectional SEM images of a CMPSF membrane immersed for 3 days in an imidazole solu-
tion, (c) dependence of the efficiencies on the current densities, and (d) cycling test at 120 mA cm22. (Reproduced
with permission from Zhao et al. [14]. Copyright 2016 by Wiley.)

010801-6 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 6 Structures of mPBI, pPBI, O-PBI, BIpPBI, and abPBI and interaction of mPBI with sulfuric acid

conductivity increases exponentially with the phosphoric acid molecules. He and coworkers [82] preswelled membranes in phos-
uptake [78]. The area resistance of BIpPBI was about half of that phoric acid, followed by immersion in a sulfuric acid solution.
of PBI. An ex situ test showed a slightly increased VO2þ perme- Apparently, the ADL increased, leading to reduced resistance. In
ability (but far below that for Nafion 115), as expected for a mem- order to further minimize the membrane resistance, a porous
brane with higher acid uptake. After 200 cycles at 50 mA cm2, O-PBI membrane was fabricated recently with a 2–5 lm thin
both mPBI and BIpPBI showed a capacity retention of >80%, defect-free dense blocking layer [12]. When the porogen content
while Nafion 115 already lost >50% of the capacity after 100 (dibutylphthalate) reached 250%, the highly porous membranes
cycles. showed a similar resistance to Nafion 211, but no measurable
As discussed, the positively charged PBI backbones hinder VO2þ permeability. The EE reached 92% and 82% at 20 and
vanadium ions in the electrolyte solution from entering the poly- 100 mA cm2, respectively.
mer. Li and coworkers [13] hypothesized that establishing several
liquid/polymer interfaces, as in a membrane with a closed pores
structure, would block vanadium ions even more efficiently. A 3 Electrodes
cell with a 68-lm thick porous O-PBI membrane can withstand Electrodes used in an RFB are required to have a high surface
over 13,000 cycles between 80 and 120 mA cm2, with good effi- area, suitable porosity, good wettability for electrolytes, low elec-
ciencies (CE close to 100%, VE and EE around 80% after 13,500 tronic resistance, high activity toward the redox reactions (while
cycles, see Fig. 7). Furthermore, the material was used in a 1 kW suppressing water electrolysis), and good (electro)chemical stabil-
stack, which showed stable efficiencies over 100 cycles at 120 mA ity. Due to the commonly corrosive environment in a RFB, there
cm2. Such porous PBI membranes are promising for VRFBs in are limited choices of materials to fabricate electrodes. Carbon-
terms of performance and cost [13]. based materials have been used as electrodes in various RFBs
Moon and coworkers [16] reported the use of porous PBI copol- such as VRFBs, polysulfide/Br, Zn/Br, and V/Ce flow batteries
ymer membranes, which contain 50% para-phenyl and 50% C7 [83–101]. Because of inadequate electrochemical activity, cata-
alkyl chain links. The pores were rather large. Because of the lysts have been employed for facilitating the redox reactions
additional interfacial resistance and poor transport of vanadium [102,103]. So far, electrodes for VRFBs have been mostly investi-
ions through PBI membranes, the CE of a porous PBI membrane gated, as will be discussed in this section.
was higher than that of a dense PBI membrane. An improved VE Since the VRFB was first proposed by Skyllas-Kazacos et al.
was obtained. Wessling and coworkers [17] compared abPBI- [104], various approaches have been reported to modify the
based membranes. Again, a porous membrane filled with sulfuric carbon-based electrode materials, including thermal treatment,
acid showed a higher conductivity than the dense membrane. chemical treatment, electrochemical oxidation, and metal doping.
The chemical structure of PBI can be varied. Not only can the Sun and Skyllas-Kazacos [97] adopted heat treatment of graphite
backbone structure be freely modified (Fig. 6) but it is also possi- felts (GFs) (at 400  C for 30 h) to improve its EE for the VRFBs.
ble to crosslink PBI or to sulfonate activated phenyl rings. Zhang The improved performance was attributed to the enhanced surface
and coworkers [79] prepared a series of crosslinked PBI mem- hydrophilicity and the formation of the oxygen functional groups
branes, in which the ratio of an aniline moiety and a sulfonated (C–O–H and C¼O groups) on the graphite felt surface, which act
phenylether was varied. These membranes showed stable CE as active sites to catalyze the electrochemical reactions, as illus-
close to 100% and EE above 80% for over 300 cycles at 60 mA trated in Fig. 8 [101]. Another method to modify the surface of
cm2. Note that the crosslinker is bisphenol A based, a structure graphite felts is chemical treatment with nitric acid or sulfuric
which is known to be acid labile and therefore will degrade over acid [96]. A significant enhancement in VRFB performance was
time [80,81]. achieved (EE of about 88% at 25 mA cm2) after the graphite felt
Since the membrane resistance decreases with the acid uptake, was treated with concentrated sulfuric acid solution for 5 h. A
efforts were made to increase the acid uptake. It is known that combination of thermal and chemical treatments has also been
PBI membranes absorb more phosphoric acid molecules per employed [105]. The graphite felt was first treated in 98 wt. % sul-
repeat unit (the acid doping level (ADL)) than sulfuric acid furic acid solution for 5 h and then kept at 450  C for 2 h. The

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-7

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


synergetic effect of thermal and chemical treatment increased the Wang [110] reported an Ir-modified carbon felt as positive elec-
number of the active sites on the graphite electrode surface and trode of VRFBs. After Ir-modification, the conductivity of elec-
enlarged the specific surface area of the graphite felt electrode trode materials increased and the cell internal resistance
(from 0.31 to 0.45 m2 g1). Li et al. [106] studied chemical treat- decreased. Jeong et al. [111] proposed Pt-based catalyst synthe-
ment of graphite felts using mixed acids (nitric acid/phosphoric sized by polyol process. To promote the VO2þ/VO2þ redox reac-
acid ¼ 3:1, at 80  C for 8 h). The improved electrochemical prop- tion, Pt/C catalyst was synthesized and coated on the surface of
erties (EE of the VRFB reached about 78% at 20 mA cm2) of the graphite felt using air spray for VRFB tests.
treated graphite felt electrode was ascribed to the increase of the Apart from using noble metals, Gonzalez et al. [112] proposed
interaction between vanadium ions and OH groups that were graphite felt modified with nanodispersed Bi as positive electrode
formed by mixed acid treatment. in VRFBs. The graphite felt was modified by immersion in a
Electrochemical oxidation is another method to modify the Bi2O3 solution in 0.01 M HCl under vacuum for 3 h and then
graphite felts [107]. After electrochemical oxidation treatment, treated in air at 450  C for 3 h. Eventually, Bi-modified graphite
the specific surface area of the graphite felt electrode increased felt contains 1.09 at % Bi on the surface of the fibers with a high
from 0.33 to 0.49 m2 g1. The improvement of electrochemical concentration of defects or holes. The well-dispersed Bi nanopar-
activity for the electrode was ascribed to the increase of the num- ticles on the carbon surface served as stable active sites. Further-
ber of COOH group and the special surface area. Tan et al. [108] more, they reported an additional achievement by incorporating
investigated the activation mechanism of electrochemically Bi nanoparticles into the graphite felt electrode in the negative
treated graphite felts by electrochemical impedance spectroscopy. half-cell of VRFBs [113]. Unlike previous study, Bi was electro-
The decrease in the value of impedance can be explained by the deposited onto thermally oxidized graphite (at 450  C in air for
increase of COOH group on the graphite felt surface. Zhang et al. 3 h) by immersing the graphite felt in 5 mM Bi(NO3)3/1 M HNO3
[109] studied the electrochemical activation of graphite felts at solution and applying 0.2 V for 1 min. Then, the graphite felt
different oxidation degrees. The graphite felt was modified was treated in air at 450  C for 30 min to stabilize the electrode-
through electrochemical oxidation at 100 mA cm2 in 1 M sulfu- posited Bi. The reversibility of the V3þ/V2þ redox reactions and
ric acid solution. Active sites on graphite felt surface for VO2þ/ the long-term cyclability of the electrode were improved by pre-
VO2þ redox reactions were created. As a result, the VE and EE venting H2 evolution through incorporating Bi nanoparticles into
for VRFB with oxidized graphite felts were 4–5% higher than that the graphite felt.
with pristine graphite felts. Metal oxides with relatively low price and comparable catalytic
Sun and Skyllas-Kazacos [98] deposited metal on the surface of properties with noble metals have been investigated for VRFBs
graphite fiber electrodes by impregnation or ion-exchange with [101]. Kim et al. [114] investigated catalytic effects of Mn3O4
solutions containing Pt4þ, Pd2þ, Au4þ, In3þ, Mn2þ, Te4þ, or Ir3þ. nanoparticles formed on carbon felts using a hydrothermal treat-
Among them, the electrode modified by Ir3þ showed the best elec- ment. A carbon felt was soaked in 1 M manganese acetate solution
trochemical properties for the vanadium redox species. Wang and and heated to 200  C for 12 h for hydrothermal reactions.

Fig. 7 (a) Cross-sectional SEM image of a 68 -lm thick porous O-PBI membrane, (b) 12 kW
stack made with that membrane, cycling performance for (c) a single cell, and (d) a 1 kW
stack. (Reproduced with permission from Yuan et al. [13]. Copyright 2016 by Royal Society of
Chemistry.)

010801-8 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Afterward, the modified carbon felt was treated at 500  C for 5 h Carbon nanotubes (CNTs) have been widely used as catalyst
under an Ar atmosphere to obtain the final product. The well- support because of their high-specific surface area, chemical sta-
dispersed Mn3O4 nanoparticles can catalyze the VO2þ/VO2þ bility, and excellent electrical conductivity [119–122]. Zhu et al.
redox reaction. The VRFBs using Mn3O4-modified carbon felts [123] first proposed graphite–carbon nanotube composite electro-
exhibited improved CE, VE, and EE compared to the cell using des for VRFBs. Yan and coworkers [124] also employed CNTs as
pristine carbon felts. Li et al. [115] proposed Nb2O5 nanorods as catalysts for VRFBs. Multiwalled carbon nanotubes (MWCNTs)
efficient catalysts toward both VO2þ/VO2þ and V3þ/V2þ redox and hydroxyl and carboxyl functional MWCNTs were coated on
couples. They investigated and optimized the amount, size, and the surface of the carbon electrodes. Modified carbon electrodes
distribution of Nb2O5 nanorods on graphite felt surface and added showed enhanced electrochemical activities and kinetic behavior
salt containing W in the precursor solutions during synthesis for in the order: carboxyl MWCNTs > hydroxyl MWCNTs > pristine
more uniform distribution and minimum agglomeration. VRFBs MWCNTs, suggesting that the oxygen functional groups could
using Nb2O5 nanorods on graphite felt exhibited about 10.7% significantly facilitate the VO2þ/VO2þ redox reaction.
higher EE at 150 mA cm2, compared to the cell without cata- Nitrogen-doped carbon electrode materials have been explored
lysts. Tseng et al. [116] investigated the effect of adding TiO2 par- with improved electroactivity [125–131]. Also, N-doped
ticles to the negative electrode in VRFBs. Hydrophilic TiO2/C polyacrylonitrile-based graphite felt by a hydrothermal ammoni-
composite electrodes were prepared by a sol-gel method. The spe- ated treatment was used for VRFBs [132]. Wang et al. [133]
cific capacitance of electrode and CE of VRFBs were improved employed N-doped carbon nanotubes on graphite felt (N-CNT/
by adding an adequate amount of TiO2 particles to the carbon GF) by a chemical vapor deposition (CVD) process. The cell
electrode. Wu et al. [117] electrodeposited PbO2 onto graphite using N-CNT/GF electrodes showed increased EE in VRFBs,
felts to improve the activity and kinetic properties. Zhou et al. compared to that with pristine graphite felts. The significantly
[118] decorated graphite felts with CeO2 as a high-performance improved performance was attributed to the unique porous struc-
electrode for VRFBs. Compared to the pristine material, CeO2 ture and increased surface area. Jin et al. [134] developed N-
nanoparticle decorated graphite felts, prepared by a facile precipi- doped graphene sheet (NGS). NGS was synthesized by annealing
tation method, showed higher activity and reversibility toward the graphite oxide with urea at 700–1050  C and used as positive elec-
VO2þ/VO2þ redox reaction. The highest EE was observed at a trodes in VRFBs. The catalytic activity was related to the types of
CeO2 content of 0.2 wt. %. nitrogen species (pyridinic, pyrrolic, quaternary, and oxidic) in
the graphene sheets, not to the N-doping level. Among them, the
quaternary nitrogen was verified as catalytic active center for the
VO2þ/VO2þ reaction.
Bio-derived carbon materials were also investigated as electro-
des for the VRFB. Park et al. [135] employed carbon-based cata-
lysts by corn protein self-assembly. Carbon black nanoparticles
were coated with N-doped graphite layers with oxygen-rich func-
tionalities. This treatment increased catalytic activity toward V3þ/
V2þ and VO2þ/VO2þ redox reactions. As a result, a significant
improvement in the EE of VRFB was observed. The abundant
oxygen active sites and nitrogen defects in the corn protein-
derived N-doped carbon black enhanced the electron transfer rate
and vanadium ion transfer kinetics. Ulaganathan et al. [136] inves-
tigated the multicouple reactions in VRFBs by using coconut
shell-derived mesoporous carbon with high surface area as
electrodes.
Recently, Park et al. [137] proposed a new fabrication method
for highly porous graphite felt electrodes. They conducted etching
treatment on graphite felts by repeating the NiO/Ni redox reac-
tions to produce a high surface area. The etched graphite felts
have stepped-edges incorporating oxygen defects, which showed
increased catalytic activity and wettability, leading to enhanced
electrochemical performance. Gonzalez et al. [138] employed
graphene-modified graphite felts as electrodes for VRFBs by elec-
trophoretic deposition. Blasi et al. [139,140] employed an electro-
spinning method to fabricate carbon nanofibers with oxide
materials. Apart from these studies focusing on electrode materi-
als, Bhattarai et al. [141] found that porous electrodes with flow
channels improved the overall EE by reducing pumping power
and enhancing flow distribution of electrolyte.

4 Organic Redox-Active Species


Compared to conventional metal ion-based RFBs, organic
materials are relatively inexpensive and structurally diverse.
Recent researches on RFBs using organic redox-active materials
have been comprehensively reviewed elsewhere [142]. In this sec-
tion, we introduce some representative organic redox species
developed for RFBs.
Fig. 8 Schematic illustration of the redox reaction mechanism
for (a) VO21/VO21 redox couple in the catholyte and (b) V31/V21 4.1 Benzoquinone. Benzoquinones (BQs) are known as
redox couple in the anolyte on the surface of the carbon felt reversible redox compounds. BQs undergo a reversible two-
electrode in VRFB. (Reproduced with permission from Kim electron process with protons in aqueous media to form hydroqui-
et al. [101]. Copyright 2015 by Royal Society of Chemistry.) none (BQH2) [143]

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-9

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


From the second cycle, the CE was over 90% and maintained
around 93% over the next 10 cycles. Although the hydroxylation
(1) of BQDS leads to a decrease in the cell voltage, BQDS is promis-
ing as active species for catholytes in organic-based aqueous
RFBs due to its high solubility (about 1.7 M at 25  C in water)
and high redox potential (0.85 V) [151]. To inhibit the hydroxyla-
The reduction of BQ in aprotic media proceeds in two steps (first
tion reaction, Hoober-Burkhardt et al. [149] have introduced 3,6-
to anion radical and then to dianion) [144,145]:
dihydroxy-2,4-dimethylbenzenesulfonic acid and achieved a high
CE of 100% over 25 cycles and good capacity retention (0.05%
decay per cycle).
(2) 4.2 Anthraquinone. The redox mechanism of anthraqui-
nones (AQs), taking 9,10-AQ as an example, is similar to that of
BQs. In solid state or aprotic media, the reduction of AQ involves
A BQ-based compound, chloranil (tetrachloro-p-benzoquinone, two steps [152,153]:
BQCl4), exhibiting good electrochemical reversibility with a high
redox potential of E0 ¼ 0.71 V (potentials are referred to the stand-
ard hydrogen electrode (SHE) in this section, if not specified) in
strongly acidic solutions, was used in a single flow acidic Cd/
BQCl4 RFB [146]:
Positive electrode (6)
In aqueous solutions [154], especially under acidic conditions
[151], AQs tend to undergo a rapid reversible two-electron
reduction.
Aspuru-Guzik and coworkers [155] systematically investigated
the structure–property relationships of quinones using computa-
(3) tional approaches. Taking redox potential and solubility into
account, the 9,10-AQs compounds were considered to be suitable
for the negative side of aqueous RFBs, whereas the 1,2-BQs,
2,3-naphthoquinones, and 2,3-AQs could be appropriate for the
Negative electrode
positive side. The redox potentials of quinones can be tuned by
utilizing electron-donating groups (such as –OH and –NH2, which
decrease the electron affinity and redox potential) or electron-
(4) withdrawing groups (such as –SO3H and –NO2, which increase
the redox potential). Substitutions of hydrophilic groups (e.g.,
–OH, –NH2, –COOH, –SO3H, and –PO3H2), especially on the
A smooth copper sheet was employed as a substrate, on which the
positions far from the ketone groups, can increase the solubility.
Cd2þ/Cd redox reaction took place. A mixture of H2SO4–(NH4)2-
Wang et al. [156] reported a nonaqueous hybrid Li/AQ RFB,
SO4–CdSO4 was continuously circulated to pass through the cell.
using graphite felt as positive electrode. In a static cell, an overall
No ionic membrane was needed in this battery. An average CE of
EE of 82% and a specific discharge energy density of 25 Wh L1
99% and EE of 82% were achieved at 10 mA cm2 over 100
were achieved. A metal-free organic/inorganic hybrid aqueous
cycles. However, the energy density of the battery is limited by
RFB was demonstrated by Aziz and coworkers [157], using 1 M
the size of the electrode and the amount of chloranil on it.
9,10-anthraquinone-2,7-disulfonic acid (AQ27DS) in 1 M H2SO4
Tiron (4,5-dihydroxy-1,3-benzenedisulfate, BQDS) is a weakly
(anolyte), 0.5 M Br2 in 3 M HBr (catholyte), and a Nafion 212
acidic aromatic organic compound. Two sulfonic groups on the
membrane:
benzene ring can improve the solubility of BQDS in water. In a
hybrid BQDS/Pb RFB [147], 0.25 M BQDS in 3 M H2SO4
medium was employed as catholyte and 3 M H2SO4 solution as
anolyte. A graphite felt positive electrode and a Pb negative elec-
trode were used, separated by a Nafion 115 membrane. An irre-
versible process occurred for the first cycle (a possible
hydroxylation reaction, resulting in new substances containing p-
benzoquinone structure [148,149]) with a low CE of only 38%
[150]:

(5) Fig. 9 Structures of selected organic compounds

010801-10 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Positive side

(7)
Negative side

(8)

Although this flow battery yielded a peak power density of 0.6 W cm2 at 1.3 A cm2 and a high CE of around 95%, crossover of bro-
mine occurred. Bromine can react with the hydroquinone and oxidize it back to quinone, resulting in capacity loss. Yang et al. [151]
introduced an aqueous all-organic RFB, using 0.2 M BQDS in 1 M H2SO4 as catholyte and 0.2 M 9,10-anthraquinone-2-sulfonic acid
(AQS) as anolyte, equipped with carbon paper electrodes and a Nafion 117 membrane:
Positive side

(9)

Negative side

(10)

Argon gas was purged to the solutions to avoid reoxidation of the reduced form to quinone. A power density of only 0.025 W cm2 was
achieved. Later, the researchers demonstrated another aqueous organic RFB with BQDS (catholyte) and AQ26DS (anolyte), which
showed an EE of 70% at 100 mA cm2 and a high CE (about 100% over 100 cycles) [154]:
Positive side: the same as Eq. (9).
Negative side

(11)

AQDS exhibited fast reaction kinetics (a rapid two-electron redox reaction), high stability, low membrane crossover, and high aqueous
solubility (1 M at pH 0).
Li and coworkers [158] reported 3,4-dihydroxy-9,10-anthraquinone-2-sulfonic acid (ARS) with a redox potential of 0.082 V. For the
RFB test, 0.05 M ARS in 1 M H2SO4 and 0.05 M BQDS in 1 M H2SO4 were used as anolyte and catholyte, respectively:
Positive side: the same as Eq. (9), E0 ¼ 0.908 V (note that this is different from a previously reported value of 0.85 V [151,154])
Negative side

(12)

The electrolytes were sealed to prevent oxidation and pumped into a cell with carbon paper electrodes and a Nafion 212 membrane. A
CE of 99% was achieved. A maximal power density of 10.6 mW cm2 at 80% state-of-charge has been obtained.
All the previously mentioned AQs were tested under acidic conditions. An alkaline quinone RFB was first introduced by Aziz and
coworkers [159]. A commercially available 2,6-dihydroxy-9,10-anthraquinone (2,6-DHAQ), which has a solubility of >0.6 M in 1 M
KOH, was used as negative material. Ferrocyanide was used as positive material:
Positive side

(13)
Negative side

(14)

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-11

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


A 0.5 M 2,6-DHAQ dipotassium salt and 0.4 M K4Fe(CN)6, both The nitroxide radical can be oxidized to oxoammonium cation
dissolved in 1 M KOH, were pumped through a flow cell with car- (reversible process) and reduced to aminoxyl anion (less reversi-
bon paper electrodes and a Nafion 212 membrane. A CE exceed- ble process). The high stability of the nitroxide radical can be
ing 99% with a stable EE of 84% was achieved at 0.1 A cm2 ascribed to (1) the delocalization of the unpaired electron by con-
over 100 cycles. A 2,6-DHAQ exhibited good stability when jugation over the N–O bond and (2) steric restriction from the
heated in 5 M KOH at 100  C for 30 days. Moreover, all the elec- four methyl groups on adjacent carbons.
troactive species remain negatively charged in all the states of Wei et al. [168] reported a nonaqueous RFB using Li metal
charge/discharge, leading to a dramatic decrease in crossover anode and 2 M TEMPO in 2.3 M LiPF6/EC-PC-EMC (EC: ethyl-
through the CEM membrane. ene carbonate, PC: propylene carbonate, and EMC: ethyl methyl
carbonate) electrolyte in a static cell:
4.3 Viologen. Viologens (1,10 -disubstituted 4,40 -bipyridinium Positive electrode reaction
ions) have been investigated as redox indicators in biological stud-
ies [160]. The electrochemical behavior of viologens was further
studied by Bird and Kuhn [161]. The viologens exist in three oxi-
dation states:

(18)
Negative electrode reaction

(15)

(19)
The first reduction step is highly reversible. The second reduction This flow battery with a high concentration and a high voltage
is less reversible, because the fully reduced state is an insoluble delivered an energy density of 126 Wh L1. Another flow cell
uncharged species. Methyl viologen (1,10 -dimethyl-4,40 -bipyridi- with a lithium–graphite hybrid anode, using a diluted catholyte of
nium, MV) is chemically stable, but will dealkylate in alkaline 0.1 M TEMPO in 1.0 M LiPF6, was able to operate at 10 mA
solutions [162]: cm2 with an EE above 75% and a CE above 99%. An enhanced
EE (86%) was achieved when operating at 5 mA cm2. Over 100
cycles, the average capacity retention was 99.8% per cycle.
A 4-methoxy-2,2,6,6-tetramethylpiperidine-1-oxyl, a methoxy-
(16) substituted TEMPO (MeO-TEMPO) was investigated by Takechi
The potential of the first reduction for MV dichloride (MVCl2) is et al. [169]. MeO-TEMPO was mixed with lithium bis(trifluoro-
0.446 V and the second one is 0.88 V [163]. The solubility of methanesulfonyl)imide (LiTFSI) at a molar ratio of 1:1 to form a
MV in water can reach 2.6 M [164]. Aqueous RFBs using violo- self-melting liquid, which could be recognized as a super-cooled
gens were demonstrated by Liu et al. [165] and Schubert and liquid. This liquid, with a concentration of over 2 M consisting of
coworkers [21,166]. Viologens were combined with TEMPO- 17 wt. % of water, was used as catholyte. Li-metal and 1 M
based species in these RFBs, which will be discussed later. LiTFSI in propylene carbonate were used as the anode and elec-
trolyte, respectively. A piece of Liþ-conducting glass ceramics
was employed as separator. MeO-TEMPO exhibited a higher
4.4 TEMPO. TEMPO (2,2,6,6-tetramethylpiperidine-1-oxyl)
redox potential than TEMPO. The charge and discharge capacities
is a heterocyclic nitroxide radical compound. TEMPO and its
at 0.1 mA cm2 were 93% and 92% of the theoretical capacity,
derivatives were investigated and used as organic electrode mate-
respectively. The CE was 99% and the capacity retention was
rials [167]. The nitroxide radical has two redox couples:
84% after 20 cycles. The energy density and specific energy were
200 Wh L1 and 155 Wh kg1, respectively.
By introducing the hydrophilic –OH group at the para-position
for the nitroxide radical compound, Liu et al. [165] obtained a
water solubility up to ca. 2.1 M and demonstrated an all-organic
aqueous RFB using 4-hydroxy-TEMPO as catholyte and MV as
(17) anolyte (0.5 M for both redox-active materials in 1.5 M NaCl):
Positive side

(20)

Negative side

(21)

010801-12 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


electrolyte was necessary. The flow cell with 2 M TEMPTMA
and MV was assembled with two graphite electrodes and an
AEM. At 75 mA cm2, the EE was over 70% and capacity stayed
unchanged over 100 cycles.
Recently, a so-called VIOTEMP was created by combining
viologen unit and TEMPO radical in one single molecule [170],
which can serve as both anode and cathode. The idea is to mimic
vanadium RFB, in which the same electrolytes are employed on
both sides. In the anolyte, during the first charge, the TEMPO rad-
ical was first reduced to its hydroxylamine form, followed by the
reduction of the viologen unit. In the catholyte, two equivalents of
TEMPO were oxidized to its oxoammonium form. Half of the
catholyte was used as sacrificial agent. In addition, an RFB utiliz-
ing VIOTEMP also suffered from fast capacity loss (nearly 50%
after 30 cycles).
Some other organic redox-active species are shown in Fig. 9.
The redox mechanisms of these materials are most related to for-
mation of radicals. With respect to their low water solubility,
these species are often investigated in nonaqueous RFBs or in
form of suspensions. Figure 10 summarizes the solubility of some
organic redox-active species discussed earlier and their redox
Fig. 10 Solubility and redox potential of some organic redox- potentials.
active species

4.5 Redox Polymers. As discussed, redox-active polymers


At 60 mA cm2, the cell delivered 71.5% theoretical capacity and allow the use of low-cost porous membranes. The crossover of
the CE remained above 99% over cycling. VE and EE were active species can be avoided by the size exclusion effect. Schu-
62.1% and 62.5%, respectively. After 100 cycles, the capacity bert and coworkers [21] demonstrated an all-organic aqueous
retention was 89%. RFB based on TEMPO and MV. Two polymers (Fig. 11), P1 con-
Schubert and coworkers [166] developed a so-called taining the TEMPO unit and P2 containing the MV unit, were
TEMPTMA by choosing the ionic trimethylammonium group synthesized and investigated as redox-active materials. The redox
(–TMA) as a substituent and accordingly obtained a water solubil- potential of P1 and P2 were determined as 0.7 V and 0.425 V
ity of up to 3.2 M in 0.3 M NaCl. The TEMPTMA has a redox versus Ag/AgCl, respectively, suggesting a cell voltage of ca.
potential of 0.79 V versus Ag/AgCl, surpassing the 4-hydroxy- 1.12 V. Using a cellulose-based dialysis membrane, a CE of ca.
TEMPO by 0.15 V. Combining TEMPTMA with MV leads to a 98% and an EE of 75% were achieved at 40 mA cm2.
theoretical cell voltage of 1.42 V. Owing to the high mobility of Schubert and coworkers [171] further investigated poly(1-
Cl counter ions in both electrolytes, no additional supporting methyl-10 -(4-vinylbenzyl)-[4,40 -bipyridine]-1,10 -diium dichloride)

Fig. 11 (a) A schematic representation of a polymer-based RFB. The anolyte and catho-
lyte are separated by a size exclusion membrane and (b) electrode reactions. (Reproduced
with permission from Janoschka et al. [21]. Copyright 2015 by Nature Publishing Group.)

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-13

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


and poly(4-methacryloyloxy-2,2,6,6-tetramethylpiperidine-1-oxyl-
co-2-(methacryloyloxy)-N,N,N-trimethylethane ammonium chlo-
ride) for RFBs. The rheological, thermal, and electrochemical
properties were studied. Winsberg et al. [172] reported a hybrid
RFB using TEMPO-methacrylate/styrene block copolymers
(PTMA-b-PS) as cathode active material and Zn2þ/Zn couple for
anode reaction (Fig. 12). The copolymer was composited with a
polar PTMA and an unpolar PS to enable the formation of
core–corona micelles. Such hybrid RFBs showed an excellent
cycling stability over 1000 cycles with 95% capacity retention and
Fig. 12 (a) Redox reactions of TEMPO1/TEMPO and Zn21/Zn0 a stable voltage.
couples and (b) a schematic representation of the copolymer Boron dipyrromethenes (BODIPYs) have chemically reversible
structure of PTMA-b-PS. (Reproduced with permission from oxidation and reduction reactions, allowing their utilization as
Winsberg et al. [172]. Copyright 2016 by Royal Society of
Chemistry.)
bipolar redox-active material for organic flow batteries. Winsberg
et al. [173] prepared two different BODIPY copolymers (P3 and
P4) (Fig. 13). They found that the redox polymers have good elec-
trochemical stability. A new battery was prepared with P3 as ano-
lyte and P4 as catholyte, which showed a mean discharging
voltage of 1.28 V and stable cycling over 90 cycles.

5 Suspension Electrolytes
Different from the conventional solid-state lithium-ion bat-
teries, RFBs store energy in redox-active electrolytes, contained
in external reservoirs. The energy density of RFBs is limited by
the concentration of the active species (typically below 2 M) and
the narrow electrochemical potential window of aqueous electro-
lytes (1.23 V). The development of semi-solid flow batteries,
with a high concentration of active species (in solid form) up to
10–40 M, can lead to an enhanced energy storage capability [9].
As illustrated in Fig. 14, in the semi-solid flow batteries, the solid
active compounds and the conductive additives are well mixed in
the electrolytes [174–178]. With sufficient conductive additives,
an electron transfer percolation network can form, which ensures
that the active materials can effectively accept or dispatch
charges. In this section, we discuss several reported semi-solid
suspensions, including inorganic intercalation-based materials,
redox-active organic compounds, and polysulfides.

5.1 Intercalation-Based Materials. In order to improve the


energy density of flow battery systems, a Liþ-based semi-solid
Fig. 13 (a) Synthesis and (b) redox mechanism of polymers P3
and P4. (Reproduced with permission from Winsberg et al.
RFB was proposed by Chiang and coworkers [9]. With appropriate
[173]. Copyright 2016 by American Chemical Society.) content of active materials (from 0 to 40 vol %), the semi-solid
system using lithium intercalation compounds (LiCoO2, LiFePO4,
and Li4Ti5O12) delivered a high-energy density of about
300–500 Wh L1. When the operating voltage was below 1 V ver-
sus Li/Liþ, an insulating solid electrolyte interphase (SEI) film can

Fig. 14 (a) Schematic illustration of a semi-solid RFB and (b) a percolation network
between the suspension

010801-14 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 15 The cyclic voltammetry properties of (a) 10 mM sulfur, (b)10 mM LiI, and (c) a mixture of 10 mM sulfur and 10 mM LiI
at 1 mV s21. (Reproduced with permission from Chen et al. [194]. Copyright 2016 by Wiley.)

form at the graphite anode due to the decomposition of alkyl car- in the liquid phase and fast redox reaction kinetics of ferrocene/
bonate solvents [9], which hindered the electron transfer between ferrocenium, the full-cell showed small polarization and high
the fluid electrode and the current collector. Using anode materials capacity retention (94% of the theoretical capacity of ferrocene at
with a suitable lithiation potential (1.55 V versus Liþ/Li for 60  C). The power density and energy density based on the vol-
Li4Ti5O12, for instance) can minimize the formation of SEI [179]. ume of catholyte reached 1400 W L1 and 40 Wh L1, respec-
Biendico et al. [180] studied the role of carbon black on tively. Excellent cycling stability (a capacity retention of 80%)
the electrochemical and rheological properties based on was observed over 500 cycles.
LiNi1/3Co1/3Mn1/3O2 suspensions. With an increase in the volume Tomai et al. [191] reported a promising approach to impregnate
percentage of Ketjenblack (KB), the electrolyte resistance and nanoporous carbon beads with redox-active tetrachlorohydroqui-
charge transfer resistance decreased. Meanwhile, the viscosity of none and 1, 5-dichloroanthraquinone to form a flowable, low vis-
the suspensions increased. During static and dynamic measure- cosity slurry electrolyte. Through shuttling protons between the
ments, large capacities can be delivered as the volume percentage slurry electrolytes, the energy density of this flow cell was
of KB increased. Moreover, Lewis and coworkers [181] intro- 19.4 Wh kg1 at 0.26 A g1, based on the total weight of solid
duced a nonionic dispersant of polyvinylpyrrolidone (PVP) into compounds.
the LiFePO4/KB suspensions. Such electrode suspensions allowed
for high content of active materials, without deteriorating the flow
behavior and electronic conductivity. Recently, carbon-free sus- 5.3 Li–S Flow Battery. A concept of semi-solid Li/polysul-
pensions with a low viscosity have been reported, functioning fide batteries was introduced by Cui and coworkers [192]. Carbon
through the collision of active particles with the current collector fiber paper was used as current collector for the polysulfide catho-
[182,183]. However, only a low concentration and low operating lyte and assembled against a passivated Li film. The catholyte was
currents (from 0.001 to 0.02 mA) were applied. The utilization of cycled only between sulfur and Li2S4, to avoid the formation of
the active particles was too low from their preliminary work. insoluble Li2S2 and Li2S. When the concentration of Li2S8
An aqueous RFB based on LiFePO4–LiTi2(PO4)3 was first reached up to 5 M, an energy density of 108 Wh L1 was
reported by Chiang and coworkers [184]. Aqueous suspensions achieved, which is about three times that of the VRFBs. The addi-
with high ionic conductivity are of interest for safe and cost- tion of LiNO3 in the ether solvent contributed to the formation of
effective applications [7]. In static cell configuration, the suspen- an SEI film, which prevented parasitic reactions between the Li
sion containing KB in 1 M LiNO3 aqueous electrolyte showed a anode and the polysulfide.
capacity retention of about 85% after 100 cycles at 10 mA cm2. Lu and coworkers [193] investigated a concentrated sulfur-
Side chemical reactions, such as (1) oxidation reactions due to the impregnated carbon suspension in order to increase the volumetric
presence of dissolved oxygen in the aqueous electrolytes, (2) capacity of a Li/polysulfide RFB. The catholyte with 20 vol % sul-
hydrolysis reaction, and (3) precipitation reaction, caused capacity fur and 26 vol % KB showed a higher volumetric capacity of up
fade over time. to 155 Ah L1 at 4 mA cm2. Furthermore, Chen and Lu [194]
Naþ-based batteries have attracted increasingly attention due to demonstrated a catholyte suspension with mixed liquid lithium
abundance in resources. Recently, Ventosa et al. [185] applied iodide and solid sulfur/carbon active components. As observed
Naþ intercalation materials as suspensions for nonaqueous RFBs, from the cyclic voltammetry analysis for the individual S/S2
using a P2-type NaxNi0.22Co0.11Mn0.66O4 suspension cathode and redox couple (Fig. 15(a)), I3/I redox couple (Fig. 15(b)), and
a NASICON-type NaTi2(PO4)3 suspension anode. NaTi2(PO4)3 their mixture (Fig. 15(c)), there were no electrochemical interfer-
has a flat potential plateau at about 2.1 V versus Naþ/Na and is ences, confirming that LiI is compatible with sulfur. A high volu-
suitable as SEI-free anode [186]. The full-cell delivered a dis- metric capacity of 550 Ah L1 catholyte has been achieved using such
charge capacity of 80 mAh g1 (based on cathode material) at multiple redox reactions of LiI and sulfur. For further application,
0.17 mA cm2 with an average discharge voltage of about 0.8 V. it is necessary to replace the lithium metal anode with semi-solid
Large overpotential was observed from the positive electrode. negative flow electrode, such as silicon anolyte [195]. Table 1
summarizes the test conditions and performance for several
5.2 Organic Materials. Metallocenes have demonstrated selected compounds.
good electrochemical performance in nonaqueous electrolytes Considering the high-energy consumption for pumping suspen-
with high-energy density and CE [187,188]. Ferrocene (Fc) den- sions with a typically high viscosity, Carter et al. [196] modeled
drites were first introduced as a redox mediator in RFBs by Wang the hydrodynamic (residence time of a fluid volume within a cell)
and coworkers [189]. Recently, a novel membrane-free semi-solid and electrochemical efficiency properties (time required to
RFB with a high cell voltage (3.0 V) using a ferrocene-based charge/discharge a single volume of suspensions within a cell) for
catholyte and a passivated lithium anode has been investigated the semi-solid RFBs. High-flow rates increase the pumping-
by Yu and coworkers [190]. Owing to the fast mass diffusion energy dissipation, whereas low-flow rates will give rise to shunt

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-15

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 Selected examples of semi-solid RFBs

Capacity/ Average
Testing Current energy voltage CE
Cathode Anode Electrolytes model density density (V) (%) Ref.

11.87 vol % LiNi1/3Co1/3 Li LiPF6 þ ethylene carbonate Static 2.5 mA cm2 48.8 Ah L1 About 3.6 90 180
Mn1/3O2 þ 13.97 vol % (EC)/dimethyl carbonate
KB (DMC) (1:1 w/w)
20 vol % LiFePO4 Li Propylene Intermittent 1.67 mA cm2 93 Wh L1 — 72.3 181
þ 1.25 vol % KB carbonate þ LiTFSI
þ 0.3 wt % PVP
NaxNi0.22Co0.11 NaTi2(PO4)3 NaPF6 in EC/DMC Static 0.17 mA cm2 9 Wh L1 About 0.9 88 185
Mn0.66O4 þ 1.3 wt % KB þ 1.3 wt % KB
Tetrachlorohydroquinone 1,5- dichloroanthraquinone A mixture of 0.5 M H2SO4 Static 0.26 A g1 19.4 Wh kg1 0.55 — 191
þ 9.8 wt % nanoporous þ 9.8 wt % and 0.05 M HCl aqueous
carbon nanoporous carbon solutions
20 vol % sulfur Li 0.2 M LiClO4/0.1 M Static 4 mA cm2 221 Ah L1 2.1 80 193
þ 26 vol % KB LiNO3 þ 1,3-dioxolane
(DOL)/1,2-dimethoxy-
ethane (DME) (1:1 v/v)
5 M LiI-20 vol % S Li LiClO4/LiNO3 þ DOL/ Static 2.5 mA cm2 550 Ah L1 About 1.9 95 194
þ 26 vol % KB DME (1:1 v/v) (based on
catholyte)

currents. In addition, at low-flow rates, the presence of gradients Electric Industries (SEI) have developed VRFB systems for a wide
in the state-of-charge, current distribution, and overpotential along range of applications [206]. They demonstrated that overall EE of
the length of the flow channels needs to be considered. Rational the previously mentioned systems could be maintained about 80%
operational mode of semi-solid flow cells is thus important. In over 270 thousand cycles [207]. To store wind energy, in 2005 SEI
addition, a suitable battery module design may improve the rheo- installed 4 MW/6 MWh VRFB plant at Subaru Wind Farm in Japan
logical properties of the suspension and reduce the energy con- [208]. The power of each stack is 45 kW, which consisted of 108
sumption from the pump. cells. Recently, MW-class energy storage system has emerged to
meet the market needs. Although SEI demonstrated already several
6 Systems MWh scale VRFB systems, they seem to be impractical due to the
expensive cost. For large-scale systems, the modularization and
6.1 Historical Development. The earlier Fe/Cr systems optimization are important. It is necessary to understand the sys-
(1 kW prototype) were developed in 1980 by NASA Glenn tem configuration from the engineering perspective.
Research Center (Cleveland, OH) [197]. Since then, several
commercial attempts were followed [198–200]. Although many
pilot-scale systems were demonstrated, they were not sufficiently 6.2 System Configuration. A VRFB system is mainly cate-
satisfactory for commercial systems [201]. Commercial attempts gorized by the cell stack, electrolyte storage, and the balance of
using the polysulfide-Br flow batteries were conducted by Regen- plant (BOP). Electrolytes are circulated between the tanks and the
esys [202]. Moreover, Exxon, Johnson Controls (Milwaukee, WI), stack during operation. VRFB system needs a power conditioning
and Meidensha developed large-scale Zn/Br flow batteries. NASA system and a system controller. Through the proper design of
Glenn Research Center and the U.S. Electric Power Research direct current (DC)/alternating current (AC) converter, the system
Institute (Palo Alto, CA) installed a large-scale electric storage can obtain high round-trip efficiency. In the case of system con-
system using a Fe/Cr RFB [203]. Japanese industries featured troller, there is a battery management system (BMS) to operate
prominently for the early development of RFBs. The Japanese the system safely and to maximize the stack performance and sys-
Ministry of International Trade and Industry initiated a tem lifetime by reducing undesired side reactions and local deple-
“Moonlight Project.” Kansai Electric Power Co., Inc. (KEPCO, tion of the redox reactants. A sufficient quantity of electrolyte
Amagasaki, Japan) tested 60 kW Fe/Cr redox battery prototypes should be circulated between storage tank and stack for preventing
[203]. Meidensha Electric Company developed a Zn/Br battery, side reactions. The upper and lower limits of the state-of-charge
and Furukawa Electric Company developed a Zn/Cl battery [203]. are approximately 80–85% and 15–20%, respectively, to avoid
Although previous studies can demonstrate the technical possibil- undesired side reactions. Furthermore, a VRFB system typically
ities, commercialization of large-scale system was not successful, uses centrifugal pumps to prevent leaking or corrosion [209].
due to inherent cross-contamination. Low energy density and high VRFB systems have several advantages from the engineering
primary capital cost also restricted the commercialization [103]. aspects. First, such systems have dominant feature for safety and
In contrast, VRFB systems were successfully commercialized thermal management. Flowing electrolytes can remove heat gen-
since 1990s. The first system appeared in 1993. Thai Gypsum erated in the cell stack. Second, the unique system configuration
Products, Ltd., installed a vanadium battery system for Solar can simplify the manufacturing process. Through the modulariza-
House in Thailand in cooperation with UNSW Center for Photo- tion of stacks, electrolyte containers, plumbing, and electrical sys-
voltaic Devices and Systems [204]. This integrated system tems, VRFB can reduce expensive and complicated production
included solar cells of 2.2 kW and the VRFB system with 12 kWh. steps. Finally, the rebalancing process in the VRFB system is sim-
They developed a 47 V, 36 cell stack for long-term field trials. In ple. For VRFB systems, it is necessary to manage the active spe-
1997, Mitsubishi Chemicals (Kashima-Kita Electric Power, cies crossover and water transport through the membrane.
Tokyo, Japan) installed a 200 kW/800 kWh grid-connected vana- Each single cell of a VRFB stack includes several components,
dium battery for load-leveling. This system could run over 150 such as the ion-exchange membrane, electrodes, and bipolar
charge–discharge cycles and showed a high EE of 80% at plates. In addition, nonconductive plastic frames and sealing
80–100 mA cm2 [205]. Since the installation of a VRFB system materials are also required. A complex manifold design is often
in 1996 at Tatsumi Sub-Station for peak shaving, Sumitomo necessary to reduce undesirable shunt currents in the ionically

010801-16 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


7 Conclusions
RFBs are promising and competitive systems for stationary
energy storage applications for the future. To promote the market
penetration of VRFBs, the performance of membranes and elec-
trodes needs to be further improved. Rational cell/stack design
and system management can lead to a reduction in the construc-
tion and operation costs. RFBs with high energy and power den-
sities can be obtained by developing new active redox couples
with high concentration, high cell voltage, and fast reaction
kinetics. Recent development in active organic molecules opens
new opportunities for sustainable energy storage strategy. Funda-
mental understanding for the physiochemical and electrochemical
properties of the electrolytes is needed. In addition, finding com-
Fig. 16 Schematic of cross section of different flow fields: (a) patible membranes, electrodes, and supporting salts is critical for
flow-by bipolar plate and (b) flow-through bipolar plate. (Repro-
successful applications. The use of energy-dense solid suspension
duced with Permission by Chalamala et al. [209]. Copyright
2014 by IEEE.) electrolytes for RFBs may largely enhance the energy density of
the system and make it possible for their use in mobile
conductive solutions. For the development of stacks, the design of applications.
channel is important for supplying reactants uniformly. There are
two methods for feeding the electrolyte along the stack: flow-by
and flow-through (Fig. 16) [209,210]. In the flow-through geome- Acknowledgment
try, there is no channel to flow the electrolyte in the bipolar plate.
We gratefully acknowledge the support from the basic research
Forced-convection occurs in the porous electrode. Thus, reactants
funding of KIST Europe. D.H. thanks the support from the
and products are easily transferred. On the other hand, in order to
German-Korean joint SME R&D projects program of ZIM-AIF
reduce pressure drop, relatively thick electrodes are required,
and MOTIE/KIAT. Z.H. gratefully acknowledges the scholarship
which leads to large ohmic loss. The overall system efficiency is
from China Scholarship Council (CSC). D.K. thanks the Basic
low for such configuration. In contrast, the pressure drop can be
Science Research Program through the National Research Foun-
reduced when flow-by channels are placed adjacent to the porous
dation of Korea (NRF) funded by the Ministry of Education
electrodes. However, there is no forced-convection flow in the
(NRF-2016R1A6A3A03007749). R.C. thanks Professor R. Hem-
electrode. To tackle the disadvantages of flow-by channels, sev-
pelmann for his continuing support.
eral types of channels have been introduced recently. A serpentine
channel can reduce the thickness of the electrode without severe
pressure drop and cause forced-convection [211]. Interdigitated References
channels can be one of the solutions for optimal channel design [1] Soloveichik, G. L., 2015, “Flow Batteries: Current Status and Trends,” Chem.
[212]. This configuration allows electrolyte to flow in the elec- Rev., 115(20), pp. 11533–11558.
[2] Noack, J., Roznyatovskaya, N., Herr, T., and Fischer, P., 2015, “The Chemis-
trode, perpendicular to the channels, with forced-convection. The try of Redox-Flow Batteries,” Angew. Chem. Int. Ed., 54(34), pp. 9776–9809.
pressure drop at interdigitated channels is higher than that for [3] Li, L., Kim, S., Wang, W., Vijayakumar, M., Nie, Z., Chen, B., Zhang, J., Xia,
flow-by type channels, but lower than that for flow-through ones. G., Hu, J., Graff, G., Liu, J., and Yang, Z., 2011, “A Stable Vanadium Redox-
Flow Battery With High Energy Density for Large-Scale Energy Storage,”
Adv. Energy Mater., 1(3), pp. 394–400.
6.3 Engineering Factors. VRFB system can be optimized by [4] Chakrabarti, M. H., Brandon, N. P., Hajimolana, S. A., Tariq, F., Yufit, V.,
focusing on the engineering factors: system scale-up, structural Hashim, M. A., Hussain, M. A., Low, C. T. J., and Aravind, P. V., 2014,
and operational parameters, and supervisor systems. Efforts were “Application of Carbon Materials in Redox Flow Batteries,” J. Power Sources,
253, pp. 150–166.
made to understand the accurate internal behavior by developing [5] Chen, R., Kim, S., and Chang, Z., 2017, “Redox Flow Batteries: Fundamentals
multiscale, multidimensional, multiphysical, steady-state and and Applications,” Redox: Principles and Advance Applications, M. A. A.
dynamic models. Some models were used to show the transient Khalid, ed., InTech, Rijeka, Croatia.
behavior of VRFB system [213]. Some sophisticated models were [6] Park, M., Ryu, J., and Cho, J., 2015, “Nanostructured Electrocatalysts for All-
Vanadium Redox Flow Batteries,” Chem. Asian J., 10(10), pp. 2096–2110.
also developed to examine the changes of flow field for different [7] Chen, R., and Hempelmann, R., 2016, “Ionic Liquid-Mediated Aqueous
channel designs, aiming at supplying the reactant uniformly over Redox Flow Batteries for High Voltage Applications,” Electrochem. Com-
the electrode areas with a low-pressure drop and reducing shunt mun., 70, pp. 56–59.
currents by advanced bipolar plates [214]. Further studies for [8] Li, B., and Liu, J., 2017, “Progress and Directions in Low-Cost Redox Flow
Batteries for Large-Scale Energy Storage,” Natl. Sci. Rev., 4(1), pp. 91–105.
channel design are needed to reduce pumping power. Further- [9] Duduta, M., Ho, B., Wood, V. C., Limthongkul, P., Brunini, V. E., Carter, W.
more, computational studies are required to solve the problems C., and Chiang, Y. M., 2011, “Semi-Solid Lithium Rechargeable Flow
related to cell elements, nonlinear material behaviors, and optimal Battery,” Adv. Energy Mater., 1(4), pp. 511–516.
geometries. Modeling can also help to optimize the operation [10] Schwenzer, B., Zhang, J., Kim, S., Li, L., Liu, J., and Yang, Z.,
2011, “Membrane Development for Vanadium Redox Flow Batteries,”
parameters of the system [215]. ChemSusChem, 4(10), pp. 1388–1406.
The flow rate is also a significant factor affecting the perform- [11] Doan, T. N. L., Hoang, T. K. A., and Chen, P., 2015, “Recent Development of
ance of the VRFB system. By increasing the flow rate stepwise Polymer Membranes as Separators for All-Vanadium Redox Flow Batteries,”
during operation, the EE of stack can be increased and the power RSC Adv., 5(89), pp. 72805–72815.
[12] Peng, S., Yan, X., Wu, X., Zhang, D., Luo, Y., Su, L., and He, G., 2017, “Thin
consumption of pump can be decreased [216]. Capacity loss can Skinned Asymmetric Polybenzimidazole Membranes With Readily Tunable
be decreased by controlling operating parameters. Capacity loss Morphologies for High-Performance Vanadium Flow Batteries,” RSC Adv.,
may be recovered by periodic mixing of the electrolyte. When the 7(4), pp. 1852–1862.
battery is off, stopping the pump is a simple solution for reducing [13] Yuan, Z., Duan, Y., Zhang, H., Li, X., Zhang, H., and Vankelecom, I., 2016,
“Advanced Porous Membranes With Ultra-High Selectivity and Stability for
the capacity loss. The species crossover can be decreased by con- Vanadium Flow Batteries,” Energy Environ. Sci., 9(2), pp. 441–447.
trolling the hydraulic pressure [216]. Increasing the current den- [14] Zhao, Y., Li, M., Yuan, Z., Li, X., Zhang, H., and Vankelecom, I. F. J., 2016,
sity may be a solution to reduce the crossover, because it can “Advanced Charged Sponge-Like Membrane With Ultrahigh Stability and
shorten the charge and discharge time. Increasing current and Selectivity for Vanadium Flow Batteries,” Adv. Funct. Mater., 26(2),
pp. 210–218.
power density may reduce stack size and overall system costs [15] Zhao, Y., Lu, W., Yuan, Z., Qiao, L., Li, X., and Zhang, H., 2017, “Advanced
[217,218]. To accelerate commercialization of VRFB systems, Charged Porous Membranes With Flexible Internal Crosslinking Structures
more engineering efforts are necessary. for Vanadium Flow Batteries,” J. Mater. Chem. A, 5(13), pp. 6193–6199.

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-17

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


[16] Maurya, S., Shin, S. H., Lee, J. Y., Kim, Y., and Moon, S. H., 2016, [44] Vijayakumar, M., Luo, Q., Lloyd, R., Nie, Z., Wei, X., Li, B., Sprenkle, V.,
“Amphoteric Nanoporous Polybenzimidazole Membrane With Extremely Londono, J.-D., Unlu, M., and Wang, W., 2016, “Tuning the Perfluorosulfonic
Low Crossover for a Vanadium Redox Flow Battery,” RSC Adv., 6(7), Acid Membrane Morphology for Vanadium Redox-Flow Batteries,” ACS
pp. 5198–5204. Appl. Mater. Interfaces, 8(50), pp. 34327–34334.
[17] Luo, T., David, O., Gendel, Y., and Wessling, M., 2016, “Porous Poly(Benzi- [45] Chromik, A., dosSantos, A. R., Turek, T., Kunz, U., H€aring, T., and Kerres, J.,
midazole) Membrane for All Vanadium Redox Flow Battery,” J. Power 2015, “Stability of Acid-Excess Acid–Base Blend Membranes in All-
Sources, 312, pp. 45–54. Vanadium Redox-Flow Batteries,” J. Membr. Sci., 476, pp. 148–155.
[18] Kangro, W., and Pieper, H., 1962, “Zur Frage der Speicherung von Elek- [46] Nambi, K. N., Henkensmeier, D., Jang, J. H., and Kim, H.-J., 2014,
trischer Energie in Fl€
ussigkeiten,” Electrochim. Acta, 7(4), pp. 435–448. “Nanocomposite Membranes for Polymer Electrolyte Fuel Cells,” Macromol.
[19] Koros, W. J., Ma, Y. H., and Shimidzu, T., 1996, “Terminology for Mem- Mater. Eng., 299(9), pp. 1031–1041.
branes and Membrane Processes,” J. Membr. Sci., 120(2), pp. 149–159. [47] Yin, B., Yu, L., Jiang, B., Wang, L., and Xi, J., 2016, “Nano Oxides Incorpo-
[20] Nagarjuna, G., Hui, J., Cheng, K. J., Lichtenstein, T., Shen, M., Moore, J. S., rated Sulfonated Poly(Ether Ether Ketone) Membranes With Improved Selec-
and Rodrıguez-Lopez, J., 2014, “Impact of Redox-Active Polymer Molecular tivity and Stability for Vanadium Redox Flow Battery,” J. Solid State
Weight on the Electrochemical Properties and Transport Across Porous Electrochem., 20(5), pp. 1271–1283.
Separators in Nonaqueous Solvents,” J. Am. Chem. Soc., 136(46), pp. [48] Chen, D., and Hickner, M. A., 2013, “V5þ Degradation of Sulfonated Radel
16309–16316. Membranes for Vanadium Redox Flow Batteries,” Phys. Chem. Chem. Phys.,
[21] Janoschka, T., Martin, N., Martin, U., Friebe, C., Morgenstern, S., Hiller, H., 15(27), pp. 11299–11305.
Hager, M. D., and Schubert, U. S., 2015, “An Aqueous, Polymer-Based [49] Kim, S., Tighe, T. B., Schwenzer, B., Yan, J., Zhang, J., Liu, J., Yang, Z., and
Redox-Flow Battery Using Non-Corrosive, Safe, and Low-Cost Materials,” Hickner, M. A., 2011, “Chemical and Mechanical Degradation of Sulfonated
Nature, 527(7576), pp. 78–81. Poly(Sulfone) Membranes in Vanadium Redox Flow Batteries,” J. Appl. Elec-
[22] Zhang, H., Zhang, H., Li, X., Mai, Z., and Zhang, J., 2011, “Nanofiltration trochem., 41(10), pp. 1201–1213.
(NF) Membranes: The Next Generation Separators for All Vanadium Redox [50] Huang, S. L., Yu, H. F., and Lin, Y. S., 2017, “Modification of Nafion Mem-
Flow Batteries (VRBs)?,” Energy Environ. Sci., 4(5), pp. 1676–1679. brane Via a Sol-Gel Route for Vanadium Redox Flow Energy Storage Battery
[23] Zhang, H., Zhang, H., and, Li, X., 2013, “Nanofiltration Membranes for Vana- Applications,” J. Chem., 2017, p. 4590952.
dium Flow Battery Application,” ECS Trans., 53(7), pp. 65–68. [51] Yu, L., Lin, F., Xu, L., and Xi, J., 2016, “A Recast Nafion/Graphene Oxide
[24] Chae, I. S., Luo, T., Moon, G. H., Ogieglo, W., Kang, Y. S., and Wessling, Composite Membrane for Advanced Vanadium Redox Flow Batteries,” RSC
M., 2016, “Ultra-High Proton/Vanadium Selectivity for Hydrophobic Polymer Adv., 6(5), pp. 3756–3763.
Membranes With Intrinsic Nanopores for Redox Flow Battery,” Adv. Energy [52] Hu, S., Lozada-Hidalgo, M., Wang, F. C., Mishchenko, A., Schedin, F., Nair,
Mater., 6(16), p. 1600517. R. R., Hill, E. W., Boukhvalov, D. W., Katsnelson, M. I., Dryfe, R. A. W.,
[25] Budd, P. M., Elabas, E. S., Ghanem, B. S., Maakhseed, S., McKeown, N. B., Wu, H. A., and Geim, A. K., 2014, “Proton Transport Through One-Atom-
Msayib, K. J., Tattershall, C. E., and Wang, D., 2004, “Solution-Processed, Thick Crystals,” Nature, 516(7530), pp. 227–230.
Organophilic Membrane Derived From a Polymer of Intrinsic Microporosity,” [53] Grosse Austing, J., Kirchner, C. N., Komsiyska, L., and Wittstock, G., 2016,
Adv. Mater., 16(5), pp. 456–459. “Layer-by-Layer Modification of Nafion Membranes for Increased Lifetime
[26] Du, N., Robertson, G. P., Song, J., Pinnau, I., and Guiver, M. D., 2009, “High- and Efficiency of Vanadium/Air Redox Flow Batteries,” J. Membr. Sci., 510,
Performance Carboxylated Polymers of Intrinsic Microporosity (PIMs) With pp. 259–269.
Tunable Gas Transport Properties,” Macromolecules, 42(16), pp. 6038–6043. [54] Ma, J., Wang, S., Peng, J., Yuan, J., Yu, C., Li, J., Ju, X., and Zhai, M., 2013,
[27] Kim, B. G., Henkensmeier, D., Kim, H. J., Jang, H. J., Nam, S. W., and Lim, “Covalently Incorporating a Cationic Charged Layer Onto Nafion Membrane
T. H., 2014, “Sulfonation of PIM-1: Towards Highly Oxygen Permeable Bind- by Radiation-Induced Graft Copolymerization to Reduce Vanadium Ion
ers for Fuel Cell Application,” Macromol. Res., 22(1), pp. 92–98. Crossover,” Eur. Polym. J., 49(7), pp. 1832–1840.
[28] Hsu, W. Y., and Gierke, T. D., 1983, “Ion Transport and Clustering in Nafion [55] Mortensen, K., Gasser, U., Alkan, G. S., and Scherer, G. G., 2008, “Structural
Perfluorinated Membranes,” J. Membr. Sci., 13(3), pp. 307–326. Characterization of Radiation-Grafted Block Copolymer Films, Using SANS
[29] Coury, L., 1999, “Conductance Measurements—Part 1: Theory,” Curr. Sep., Technique,” J. Polym. Sci., 46(16), pp. 1660–1668.
18, pp. 91–96. [56] Nibel, O., Schmidt, T. J., and Gubler, L., 2016, “Bifunctional Ion-Conducting
[30] Jeong, S., Kim, L., Kwon, Y., and Kim, S., 2014, “Effect of Nafion Membrane Polymer Electrolyte for the Vanadium Redox Flow Battery With High
Thickness on Performance of Vanadium Redox Flow Battery,” Korean J. Selectivity,” J. Electrochem. Soc., 163(13), pp. A2563–A2570.
Chem. Eng., 31(11), pp. 2081–2087. [57] Nibel, O., Bon, M., Agiorousis, M. L., Laino, T., and Gubler, L., 2017,
[31] Jiang, B., Wu, L., Yu, L., Qiu, X., and Xi, J., 2016, “A Comparative Study of “Unraveling the Interaction Mechanism Between Amidoxime Groups and
Nafion Series Membranes for Vanadium Redox Flow Batteries,” J. Membr. Vanadium Ions at Various pH Conditions,” J. Phys. Chem. C, 121(12), pp.
Sci., 510, pp. 18–26. 6436–6445.
[32] Jiang, B., Yu, L., Wu, L., Mu, D., Liu, L., Xi, J., and Qiu, X., 2016, “Insights Into [58] Chen, D., Hickner, M. A., Agar, E., and Kumbur, E. C., 2013, “Selective
the Impact of the Nafion Membrane Pretreatment Process on Vanadium Flow Anion Exchange Membranes for High Coulombic Efficiency Vanadium
Battery Performance,” ACS Appl. Mater. Interfaces, 8(19), pp. 12228–12238. Redox Flow Batteries,” Electrochem. Commun., 26, pp. 37–40.
[33] Henkensmeier, D., and Gubler, L., 2014, “Shape Memory Effect in Radiation [59] Yun, S., Parrondo, J., and Ramani, V., 2015, “A Vanadium–Cerium Redox
Grafted Ion Exchange Membranes,” J. Mater. Chem. A, 2(25), pp. 9482–9485. Flow Battery With an Anion-Exchange Membrane Separator,” ChemPlu-
[34] Hink, S., Henkensmeier, D., Jang, J. H., Kim, H. J., Han, J., and Nam, S. W., sChem, 80(2), pp. 412–421.
2015, “Reduced In-Plane Swelling of Nafion by a Biaxial Modification Proc- [60] Marino, M. G., and Kreuer, K. D., 2015, “Alkaline Stability of Quaternary
ess,” Macromol. Chem. Phys., 216(11), pp. 1235–1243. Ammonium Cations for Alkaline Fuel Cell Membranes and Ionic Liquids,”
[35] Moore, R. B., and Martin, C. R., 1988, “Chemical and Morphological Proper- ChemSusChem, 8(3), pp. 513–523.
ties of Solution-Cast Perfluorosulfonate Ionomers,” Macromolecules, 21(5), [61] Merle, G., Wessling, M., and Nijmeijer, K., 2011, “Anion Exchange Mem-
pp. 1334–1339. branes for Alkaline Fuel Cells: A Review,” J. Membr. Sci., 377(1–2), pp. 1–35.
[36] Li, J., Yang, X., Tang, H., and Pan, M., 2010, “Durable and High Performance [62] Zhang, B., Zhang, E., Wang, G., Yu, P., Zhao, Q., and Yao, F., 2015,
Nafion Membrane Prepared Through High-Temperature Annealing Method- “Poly(Phenyl Sulfone) Anion Exchange Membranes With Pyridinium Groups
ology,” J. Membr. Sci., 361(1–2), pp. 38–42. for Vanadium Redox Flow Battery Applications,” J. Power Sources, 282,
[37] Kim, Y. S., Welch, C. F., Hjelm, R. P., Mack, N. H., Labouriau, A., and Orler, pp. 328–334.
E. B., 2015, “Origin of Toughness in Dispersion-Cast Nafion Membranes,” [63] Xu, W., Zhao, Y., Yuan, Z., Li, X., Zhang, H., and Vankelecom, I. F. J., 2015,
Macromolecules, 48(7), pp. 2161–2172. “Highly Stable Anion Exchange Membranes With Internal Cross-Linking
[38] Roy, A., Lee, H. S., and McGrath, J. E., 2008, “Hydrophilic–Hydrophobic Networks,” Adv. Funct. Mater., 25(17), pp. 2583–2589.
Multiblock Copolymers Based on Poly(Arylene Ether Sulfone)s as Novel Pro- [64] Zeng, L., Zhao, T. S., Wei, L., Zeng, Y. K., and Zhang, Z. H., 2016, “Highly
ton Exchange Membranes—Part B,” Polymer, 49(23), pp. 5037–5044. Stable Pyridinium-Functionalized Cross-Linked Anion Exchange Membranes
[39] Yang, S., Ahn, Y., and Kim, D., 2017, “Poly(Arylene Ether Ketone) Proton for All Vanadium Redox Flow Batteries,” J. Power Sources, 331, pp. 452–461.
Exchange Membranes Grafted With Long Aliphatic Pendant Sulfonated [65] Lu, D., Wen, L., Nie, F., and Xue, L., 2016, “Synthesis and Investigation of
Groups for Vanadium Redox Flow Batteries,” J. Mater. Chem. A, 5(5), Imidazolium Functionalized Poly(Arylene Ether Sulfone)s as Anion Exchange
pp. 2261–2270. Membranes for All-Vanadium Redox Flow Batteries,” RSC Adv., 6(8),
[40] Gindt, B. P., Tang, Z., Watkins, D. L., Abebe, D. G., Seo, S., Tuli, S., pp. 6029–6037.
Ghassemi, H., Zawodzinski, T. A., and Fujiwara, T., 2017, “Effects of Sulfo- [66] Arges, C. G., and Ramani, V., 2013, “Two-Dimensional NMR Spectroscopy
nated Side Chains Used in Polysulfone Based PEMs for VRFB Separator,” Reveals Cation-Triggered Backbone Degradation in Polysulfone-Based Anion
J. Membr. Sci., 532, pp. 58–67. Exchange Membranes,” Proc. Natl. Acad. Sci. U.S.A., 110(7), pp. 2490–2495.
[41] Jung, H. Y., Jeong, S., and Kwon, Y., 2016, “The Effects of Different Thick [67] Yuan, Z., Li, X., Zhao, Y., and Zhang, H., 2015, “Mechanism of Polysulfone-
Sulfonated Poly (Ether Ether Ketone) Membranes on Performance of Vana- Based Anion Exchange Membranes Degradation in Vanadium Flow Battery,”
dium Redox Flow Battery,” J. Electrochem. Soc., 163(1), pp. A5090–A5096. ACS Appl. Mater. Interfaces, 7(34), pp. 19446–19454.
[42] Xi, J., Li, Z., Yu, L., Yin, B., Wang, L., Liu, L., Qiu, X., and Chen, L., [68] Zhang, B., Zhang, S., Weng, Z., Wang, G., Zhang, E., Yu, P., Chen, X., and
2015, “Effect of Degree of Sulfonation and Casting Solvent on Sulfonated Wang, X., 2016, “Quaternized Adamantane-Containing Poly(Aryl Ether
Poly(Ether Ether Ketone) Membrane for Vanadium Redox Flow Battery,” Ketone) Anion Exchange Membranes for Vanadium Redox Flow Battery
J. Power Sources, 285, pp. 195–204. Applications,” J. Power Sources, 325, pp. 801–807.
[43] Kreuer, K. D., 2001, “On the Development of Proton Conducting Polymer [69] Sun, C. N., Tang, Z., Belcher, C., Zawodzinski, T. A., and Fujimoto, C., 2014,
Membranes for Hydrogen and Methanol Fuel Cells,” J. Membr. Sci., 185(1), “Evaluation of Diels–Alder Poly(Phenylene) Anion Exchange Membranes in
pp. 29–39. All-Vanadium Redox Flow Batteries,” Electrochem. Commun., 43, pp. 63–66.

010801-18 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


[70] Pezeshki, A. M., Tang, Z. J., Fujimoto, C., Sun, C. N., Mench, M. M., and [95] Shao, Y., Wang, X., Engelhard, M., Wang, C., Dai, S., Liu, J., Yang, Z.,
Zawodzinski, T. A., 2016, “Full Cell Study of Diels Alder Poly(Phenylene) and Lin, Y., 2010, “Nitrogen-Doped Mesoporous Carbon for Energy Stor-
Anion and Cation Exchange Membranes in Vanadium Redox Flow Batteries,” age in Vanadium Redox Flow Batteries,” J. Power Sources, 195(13), pp.
J. Electrochem. Soc., 163(1), pp. A5154–A5162. 4375–4379.
[71] Small, L. J., Pratt, H. D., III, Fujimoto, C. H., and Anderson, T. M., 2016, [96] Sun, B., and Skyllas-Kazacos, M., 1992, “Chemical Modification of Graphite
“Diels Alder Polyphenylene Anion Exchange Membrane for Nonaqueous Electrode Materials for Vanadium Redox Flow Battery Application—Part II:
Redox Flow Batteries,” J. Electrochem. Soc., 163(1), pp. A5106–A5111. Acid Treatments,” Electrochim. Acta, 37(13), pp. 2459–2465.
[72] Yun, S., Parrondo, J., and Ramani, V., 2016, “Composite Anion Exchange [97] Sun, B., and Skyllas-Kazacos, M., 1992, “Modification of Graphite Electrode
Membranes Based on Quaternized Cardo-Poly(Etherketone) and Quaternized Materials for Vanadium Redox Flow Battery Application—I: Thermal
Inorganic Fillers for Vanadium Redox Flow Battery Applications,” Int. J. Treatment,” Electrochim. Acta, 37(7), pp. 1253–1260.
Hydrogen Energy, 41(25), pp. 10766–10775. [98] Sun, B., and Skyllas-Kazacos, M., 1991, “Chemical Modification and
[73] Li, Q., Jensen, J. O., Savinell, R. F., and Bjerrum, N. J., 2009, “High Tempera- Electrochemical Behaviour of Graphite Fibre in Acidic Vanadium Solution,”
ture Proton Exchange Membranes Based on Polybenzimidazoles for Fuel Electrochim. Acta, 36(3–4), pp. 513–517.
Cells,” Prog. Polym. Sci., 34(5), pp. 449–477. [99] Rychcik, M., and Skyllas-Kazacos, M., 1987, “Evaluation of Electrode
[74] Asensio, J. A., Sanchez, E. M., and G omez-Romero, P., 2010, “Proton- Materials for Vanadium Redox Cell,” J. Power Sources, 19(1), pp.
Conducting Membranes Based on Benzimidazole Polymers for High- 45–54.
Temperature PEM Fuel Cells. A Chemical Quest,” Chem. Soc. Rev., 39(8), [100] Zhong, S., Padeste, C., Kazacos, M., and Skyllas-Kazacos, M., 1993,
pp. 3210–3239. “Comparison of the Physical, Chemical and Electrochemical Properties of
[75] Eberhardt, S. H., Toulec, M., Marone, F., Stampanoni, M., B€uchi, F. N., and Rayon- and Polyacrylonitrile-Based Graphite Felt Electrodes,” J. Power Sour-
Schmidt, T. J., 2015, “Dynamic Operation of HT-PEFC: In-Operando Imaging ces, 45(1), pp. 29–41.
of Phosphoric Acid Profiles and (re)Distribution,” J. Electrochem. Soc., [101] Kim, K. J., Park, M. S., Kim, Y. J., Kim, J. H., Dou, S. X., and Skyllas-Kaza-
162(3), pp. F310–F316. cos, M., 2015, “A Technology Review of Electrodes and Reaction Mecha-
[76] Zhou, X. L., Zhao, T. S., An, L., Wei, L., and Zhang, C., 2015, “The Use of nisms in Vanadium Redox Flow Batteries,” J. Mater. Chem. A, 3(33),
Polybenzimidazole Membranes in Vanadium Redox Flow Batteries Leading pp. 16913–16933.
to Increased Coulombic Efficiency and Cycling Performance,” Electrochim. [102] de Leon, C. P., Frias-Ferrer, A., Gonzalez-Garcia, J., Szanto, D., and Walsh,
Acta, 153, pp. 492–498. F., 2006, “Redox Flow Cells for Energy Conversion,” J. Power Sources,
[77] Jang, J. K., Kim, T. H., Yoon, S. J., Lee, J. Y., Lee, J. C., and Hong, Y. T., 160(1), pp. 716–732.
2016, “Highly Proton Conductive, Dense Polybenzimidazole Membranes [103] Skyllas-Kazacos, M., Chakrabarti, M. H., Hajimolana, S. A., Mjalli, F. S., and
With Low Permeability to Vanadium and Enhanced H2SO4 Absorption Capa- Saleem, M., 2011, “Progress in Flow Battery Research and Development,”
bility for Use in Vanadium Redox Flow Batteries,” J. Mater. Chem. A, 4(37), J. Electrochem. Soc., 158(8), pp. 55–79.
pp. 14342–14355. [104] Skyllas-Kazacos, M., Rychcik, M., Robins, R. G., Fane, A. G., and Green, M.
[78] Singh, B., Duong, N. M. H., Henkensmeier, D., Jang, J. H., Kim, H. J., Han, A., 1986, “New All-Vanadium Redox Flow Cell,” J. Electrochem. Soc.,
J., and Nam, S. W., 2017, “Influence of Different Side-Groups and Cross- 133(5), pp. 1057–1058.
Links on Phosphoric Acid Doped Radel-Based Polysulfone Membranes for [105] Yang, Z., Zhang, J., Kintner-Meyer, M. C. W., Lu, X., Choi, D., Lemmon, J.
High Temperature Polymer Electrolyte Fuel Cells,” Electrochim. Acta, 224, P., and Liu, J., 2011, “Electrochemical Energy Storage for Green Grid,”
pp. 306–313. Chem. Rev., 111(5), pp. 3577–3613.
[79] Xia, Z., Ying, L., Fang, J., Du, Y. Y., Zhang, W. M., Guo, X., and Yin, J., [106] Li, W. W., Chu, Y. Q., and Ma, C. A., 2014, “Highly Hydroxylated Graphite
2017, “Preparation of Covalently Cross-Linked Sulfonated Polybenzimidazole Felts Used as Electrodes for a Vanadium Redox Flow Battery,” Adv. Mater.
Membranes for Vanadium Redox Flow Battery Applications,” J. Membr Sci., Res., 936, pp. 471–475.
525, pp. 229–239. [107] Li, X. G., Huang, K. L., Liu, S. Q., Tan, N., and Chen, L. Q., 2007,
[80] Chen, W. F., Lin, H. Y., and Dai, S. A., 2004, “Generation and Synthetic Uses “Characteristics of Graphite Felt Electrode Electrochemically Oxidized for
of Stable 4-[2-Isopropylidene]-Phenol Carbocation From Bisphenol A,” Org. Vanadium Redox Battery Application,” Trans. Nonferrous Met. Soc. China,
Lett., 6(14), pp. 2341–2343. 17(1), pp. 195–199.
[81] Johnson, B. C., Yilg€ or, I., Tran, C., Iqbal, M., Wightman, J. P., Lloyd, D. R., [108] Tan, N., Huang, K. L., Liu, S. Q., Li, X. G., and Chang, Z. F., 2006,
and McGrath, J. E., 1984, “Synthesis and Characterization of Sulfonated Poly “Activation Mechanism Study of Electrochemical Treated Graphite Felt for
(Arylene Ether Sulfones),” J. Polym. Sci., 22(3), pp. 721–737. Vanadium Redox Cell by Electrochemical Impedance Spectrum,” Acta Chim.
[82] Peng, S., Yan, X., Zhang, D., Wu, X., Luo, Y., and He, G., 2016, “A H3PO4 Sin., 64(6), pp. 584–588.
Preswelling Strategy to Enhance the Proton Conductivity of a H2SO4-Doped [109] Zhang, W., Xi, J., Li, Z., Zhou, H., Liu, L., Wu, Z., and Qiu, X., 2013,
Polybenzimidazole Membrane for Vanadium Flow Batteries,” RSC Adv., “Electrochemical Activation of Graphite Felt Electrode for VO2þ/VO2þ
6(28), pp. 23479–23488. Redox Couple Application,” Electrochim. Acta, 89, pp. 429–435.
[83] Bartolozzi, M., 1989, “Development of Redox Flow Batteries. A Historical [110] Wang, W. H., and Wang, W. D., 2007, “Investigation of Ir-Modified Carbon
Bibliography,” J. Power Sources, 27(3), pp. 219–234. Felt as the Positive Electrode of an All-Vanadium Redox Flow Battery,” Elec-
[84] Sum, E., and Skyllas-Kazacos, M., 1985, “A Study of the V(II)/V(III) Redox trochim. Acta, 52(24), pp. 6755–6762.
Couple for Redox Flow Cell Applications,” J. Power Sources, 15(2–3), [111] Jeong, S., Kim, S., and Kwon, Y., 2013, “Performance Enhancement in Vana-
pp. 179–190. dium Redox Flow Battery Using Platinum-Based Electrocatalyst Synthesized
[85] Zhong, S., and Skyllas-Kazacos, M., 1992, “Electrochemical Behavior of by Polyol Process,” Electrochim. Acta, 114, pp. 439–447.
Vanadium(V)/Vanadium(IV) Redox Couple at Graphite Electrodes,” J. Power [112] Gonzalez, Z., Sanchez, A., Blanco, C., Granda, M., Menendez, R., and
Sources, 39(1), pp. 1–9. Santanmarıa, R., 2011, “Enhanced Performance of a Bi-Modified Graphite
[86] Zhong, S., Kazacos, M., Burford, R. P., and Skyllas-Kazacos, M., 1991, Felt as the Positive Electrode of a Vanadium Redox Flow Battery,” Electro-
“Fabrication and Activation Studies of Conducting Plastic Composite Electro- chem. Commun., 13(12), pp. 1379–1382.
des for Redox Cells,” J. Power Sources, 36(1), pp. 29–43. [113] Suarez, D. J., Gonzalez, Z., Blanco, C., Granda, M., Menendez, R., and
[87] Kazacos, M., and Skyllas-Kazacos, M., 1989, “Performance Characteristics of Santamarıa, R., 2014, “Graphite Felt Modified With Bismuth Nanoparticles as
Carbon Plastic Electrodes in the All-Vanadium Redox Cell,” J. Electrochem. Negative Electrode in a Vanadium Redox Flow Battery,” ChemSusChem,
Soc., 136(9), pp. 2759–2760. 7(3), pp. 914–918.
[88] Haddadi-Asl, V., Kazacos, M., and Skyllas-Kazacos, M., 1995, “Conductive [114] Kim, K. J., Park, M. S., Kim, J. H., Hwang, U., Lee, N. J., Jeong, G., and Kim,
Carbon-Polypropylene Composite Electrodes for Vanadium Redox Battery,” Y. J., 2012, “Novel Catalytic Effects of Mn3O4 for All Vanadium Redox Flow
J. Appl. Electrochem., 25(1), pp. 29–33. Batteries,” Chem. Commun., 48(44), pp. 5455–5457.
[89] Haddadi-Asl, V., Kazacos, M., and Skyllas-Kazacos, M., 1995, [115] Li, B., Gu, M., Nie, Z., Wei, X. L., Wang, C. M., Sprenkle, V., and Wang, W.,
“Carbon–Polymer Composite Electrodes for Redox Cells,” J. Appl. Polym. 2014, “Nanorod Niobium Oxide as Powerful Catalysts for an All Vanadium
Sci., 57(12), pp. 1455–1463. Redox Flow Battery,” Nano Lett., 14(1), pp. 158–165.
[90] Radford, G. J. W., Cox, J., Wills, R. G. A., and Walsh, F. C., 2008, [116] Tseng, T. M., Huang, R. H., Huang, G. Y., Hsueh, K. L., and Shieu, F. S.,
“Electrochemical Characterization of Activated Carbon Particles Used in 2013, “Improvement of Titanium Dioxide Addition on Carbon Black Compos-
Redox Flow Battery Electrodes,” J. Power Sources, 185(2), pp. ite for Negative Electrode in Vanadium Redox Flow Battery,” J. Electrochem.
1499–1504. Soc., 160(8), pp. A1269–A1275.
[91] Inoue, M., Tsuzuki, Y., Iizuka, Y., and Shimada, M., 1987, “Carbon Fiber [117] Wu, X. X., Xu, H. F., Lu, L., Zhao, H., Fu, J., Shen, Y., Xu, P. C., and Dong,
Electrode for Redox Flow Battery,” J. Electrochem. Soc., 134(3), Y. M., 2014, “PbO2-Modified Graphite Felt as the Positive Electrode for an
pp. 756–757. All-Vanadium Redox Flow Battery,” J. Power Sources, 250, pp. 274–278.
[92] Kaneko, H., Nozaki, K., Wada, Y., Aoki, T., Negishi, A., and Kamimoto, M., [118] Zhou, H., Xi, J. G., Li, Z. H., Zhang, Z. G., Yu, L. H., Liu, L., Qiu, X. P., and
1991, “Vanadium Redox Reactions and Carbon Electrodes for Vanadium Chen, L. Q., 2014, “CeO2 Decorated Graphite Felt as a High-Performance
Redox Flow Battery,” Electrochim. Acta, 36(7), pp. 1191–1196. Electrode for Vanadium Redox Flow Batteries,” RSC Adv., 4(106), pp.
[93] Xi, J., Zenghua, W., Qiu, X., and Chen, L., 2007, “Nafion/SiO2 Hybride Mem- 61912–61918.
brane for Vanadium Redox Flow Battery,” J. Power Sources, 166(2), [119] Wang, X., Li, W. Z., Chen, Z. W., Waje, M. H., and Yan, Y. S., 2006,
pp. 531–536. “Durability Investigation of Carbon Nanotube as Catalyst Support for Proton
[94] Mohammadi, F., Timbrell, P., Zhong, S., Padeste, C., and Skyllas-Kazacos, Exchange Membrane Fuel Cell,” J. Power Sources, 158(1), pp. 154–159.
M., 1994, “Overcharge in the Vanadium Redox Battery and Changes in [120] Jha, N., Leela, A., Reddy, M., Shaijumoon, M. M., Rajalakshmi, N., and Ram-
Electrical Resistivity and Surface Functionality of Graphite-Felt Electrodes,” aprabhu, S., 2008, “Pt–Ru/Multi-Walled Carbon Nanotubes as Electrocatalysts
J. Power Sources, 52(1), pp. 61–68. for Direct Methanol Fuel Cell,” Int. J. Hydrogen Energy, 33(1), pp. 427–433.

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-19

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


[121] Mohanareddy, A., Rajalakshmi, N., and Ramaprabhu, S., 2008, “Cobalt-Poly- [147] Xu, Y., Wen, Y., Cheng, J., Cao, G., and Yang, Y., 2010, “A Study of Tiron in
pyrrole-Multiwalled Carbon Nanotube Catalysts for Hydrogen and Alcohol Aqueous Solutions for Redox Flow Battery Application,” Electrochim. Acta,
Fuel Cells,” Carbon, 46(1), pp. 2–11. 55(3), pp. 715–720.
[122] Saha, M. S., Li, R., and Sun, X., 2008, “High Loading and Monodispersed Pt [148] Papouchado, L., Petrie, G., and Adams, R. N., 1972, “Anodic Oxidation Path-
Nanoparticles on Multiwalled Carbon Nanotubes for High Performance Proton ways of Phenolic Compounds—Part I: Anodic Hydroxylation Reactions,” J.
Exchange Membrane Fuel Cells,” J. Power Sources, 177(2), pp. 314–322. Electroanal. Chem., 38(2), pp. 389–395.
[123] Zhu, H. Q., Zhang, Y. M., Yue, L., Li, W. S., Li, G. L., Shu, D., and Chen, H. [149] Hoober-Burkhardt, L., Krishnamoorthy, S., Yang, B., Murali, A., Nirmalchan-
Y., 2008, “Graphite–Carbon Nanotube Composite Electrodes for All Vana- dar, A., Surya Prakash, G. K., and Narayanan, S. R., 2017, “A New Michael-
dium Redox Flow Battery,” J. Power Sources, 184(2), pp. 637–640. Reaction-Resistant Benzoquinone for Aqueous Organic Redox Flow
[124] Li, W. Y., Liu, J. G., and Yan, C. W., 2011, “Multi-Walled Carbon Nanotubes Batteries,” J. Electrochem. Soc., 164(4), pp. A600–A607.
Used as an Electrode Reaction Catalyst for VO2þ/VO2þ for a Vanadium [150] Yang, B., Hoober-Burkhardt, L., Krishnamoorthy, S., Murali, A., Surya Pra-
Redox Flow Battery,” Carbon, 49(11), pp. 3463–3470. kash, G. K., and Narayanan, S. R., 2016, “High-Performance Aqueous Organic
[125] Maldonado, S., and Stevenson, K. J., 2005, “Influence of Nitrogen Doping on Flow Battery With Quinone-Based Redox Couples at Both Electrodes,” J.
Oxygen Reduction Electrocatalysis at Carbon Nanofiber Electrodes,” J. Phys. Electrochem. Soc., 163(7), pp. A1442–A1449.
Chem. B, 109(10), pp. 4707–4716. [151] Yang, B., Hoober-Burkhardt, L., Wang, F., Surya Prakash, G. K., and Nar-
[126] Shao, Y., Sui, J., Yin, G., and Gao, Y., 2008, “Nitrogen-Doped Carbon Nano- ayanan, S. R., 2014, “An Inexpensive Aqueous Flow Battery for Large-Scale
structures and Their Composites as Catalytic Materials for Proton Exchange Electrical Energy Storage Based on Water-Soluble Organic Redox Couples,”
Membrane Fuel Cell,” Appl. Catal., B, 79(1), pp. 89–99. J. Electrochem. Soc., 161(9), pp. A1371–A1380.
[127] Gong, K., Du, F., Xia, Z., Durstock, M., and Dai, L., 2009, “Nitrogen-Doped [152] Song, Z., Zhan, H., and Zhou, Y., 2009, “Anthraquinone Based Polymer as
Carbon Nanotube Arrays With High Electrocatalytic Activity for Oxygen High Performance Cathode Material for Rechargeable Lithium Batteries,”
Reduction,” Science, 323(5915), pp. 760–764. Chem. Commun., 2009(4), pp. 448–450.
[128] Sidik, R. A., Anderson, A. B., Subramanian, N. P., Kumaraguru, S. P., and [153] Zhao, L., Wang, W., Wang, A., Yu, Z., Chen, S., and Yang, Y., 2011, “A MC/
Popov, B. N., 2006, “O2 Reduction on Graphite and Nitrogen-Doped Graphite: AQ Parasitic Composite as Cathode Material for Lithium Battery,” J. Electro-
Experiment and Theory,” J. Phys. Chem. B, 110(4), pp. 1787–1793. chem. Soc., 158(9), pp. A991–A996.
[129] Wu, G., Li, D., Dai, C., Wang, D., and Li, N., 2008, “Well-Dispersed High- [154] Guin, P. S., Das, S., and Mandal, P. C., 2011, “Electrochemical Reduction of
Loading Pt Nanoparticles Supported by ShellCore Nanostructured Carbon Quinones in Different Media: A Review,” Int. J. Electrochem., 2011, p.
for Methanol Electrooxidation,” Langmuir, 24(7), pp. 3566–3575. 816202.
[130] Saha, M. S., Li, R., Sun, X., and Ye, S., 2009, “3-D Composite Electrodes for [155] Er, S., Suh, C., Marshak, M. P., and Aspuru-Guzik, A., 2015, “Computational
High Performance PEM Fuel Cells Composed of Pt Supported on Nitrogen- Design of Molecules for All-Quinone Redox Flow Battery,” Chem. Sci., 6(2),
Doped Carbon Nanotubes Grown on Carbon Paper,” Electrochem. Commun., pp. 885–893.
11(2), pp. 438–441. [156] Wang, W., Xu, W., Cosimbescu, L., Choi, D., Li, L., and Yang, Z., 2012,
[131] Shi, L., Gao, Q., and Wu, Y., 2009, “High Performance Oxide Functionalized “Anthraquinone With Tailored Structure for a Nonaqueous Metal-Organic
Nitrogen-Doped Mesocellular Carbon Foam for Biosensor Construction,” Redox Flow Battery,” Chem. Commun., 48(53), pp. 6669–6671.
Electroanalysis, 21(6), pp. 715–722. [157] Huskinson, B., Marshak, M. P., Suh, C., Er, S., Gerhardt, M. R., Galvin, C. J.,
[132] Wo, T., Huang, K. L., Liu, S. Q., Zhung, S. X., Fang, D., Li, S., Lu, D., and Chen, X., Aspuru-Guzik, A., Gordon, R. G., and Aziz, M. J., 2014, “A Metal-
Su, A. Q., 2012, “Hydrothermal Ammoniated Treatment of PAN-Graphite Free Organic-Inorganic Aqueous Flow Battery,” Nature, 505(7482),
Felt for Vanadium Redox Flow Battery,” J. Solid State Electrochem., 16(2), pp. 195–198.
pp. 579–585. [158] Zhang, S., Li, X., and Chu, D., 2016, “An Organic Electroactive Material for
[133] Wang, S. G., Zhao, X. S., Cochell, T., and Manthiram, A., 2012, “Nitrogen- Flow Batteries,” Electrochim. Acta, 190, pp. 737–743.
Doped Carbon Nanotube/Graphite Felts as Advanced Electrode Materials for [159] Lin, K., Chen, Q., Gerhardt, M. R., Tong, L., Kim, S. B., Eisenach, L., Valle,
Vanadium Redox Flow Batteries,” J. Phys. Chem. Lett., 3(16), pp. 2164–2167. A. W., Hardee, D., Gordon, R. G., Aziz, M. J., and Marshak, M. P., 2015,
[134] Jin, J. T., Fu, X. G., Liu, Q., Liu, Y. R., Wei, Z. Y., Niu, K. X., and Zhang, J. “Alkaline Quinone Flow Battery,” Science, 349(6255), pp. 1529–1532.
Y., 2013, “Identifying the Active Site in Nitrogen-Doped Graphene for the [160] Michaelis, L., and Hill, E. S., 1933, “The Viologen Indicators,” J. Gen. Phys-
VO2þ/VO2þ Redox Reaction,” ACS Nano, 7(6), pp. 4764–4773. iol., 16(6), pp. 859–873.
[135] Park, M. J., Ryu, J. C., Kim, Y. S., and Cho, J., 2014, “Corn Protein-Derived [161] Bird, C. L., and Kuhn, A. T., 1981, “Electrochemistry of the Viologens,”
Nitrogen-Doped Carbon Materials With Oxygen-Rich Functional Groups: A Chem. Soc. Rev., 10(1), pp. 49–82.
Highly Efficient Electrocatalyst for All-Vanadium Redox Flow Batteries,” [162] Farrington, J. A., Ledwith, A., and Stam, M. F., 1969, “Cation-Radicals: Oxi-
Energy Environ. Sci., 7(11), pp. 3727–3735. dation of Methoxide ion With 1,10 -Dimethyl-4,40 -Bipyriylium Dichloride
[136] Ulaganathan, M., Jain, A., Aravindan, V., Jayaraman, S., Ling, W. C., Lim, T. (Paraquat Dichloride),” J. Chem. Soc. D, 6, p. 259.
M., Srinivasan, M. P., Yan, Q., and Madhavi, S., 2015, “Bio-Mass Derived [163] Ito, M., and Kuwana, T., 1971, “Spectroelectrochemical Study of Indirect
Mesoporous Carbon as Superior Electrode in All Vanadium Redox Flow Bat- Reduction of Triphosphopyridine Nucleotide—I: Methyl Viologen,
tery With Multicouple Reactions,” J. Power Sources, 274, pp. 846–850. Ferredoxin-TPN-Reductase and TPN,” J. Electroanal. Chem., 32(3),
[137] Park, J. J., Park, J. H., Park, O. O., and Yang, J. H., 2016, “Highly Porous Gra- pp. 415–425.
phenated Graphite Felt Electrodes With Catalytic Defects for High- [164] Haley, T. J., 1979, “Review of the Toxicology of Paraquat (1,10 -Dimethyl-
Performance Vanadium Redox Flow Batteries Produced Via NiO/Ni Redox 4,40 -Bipyridinium Chloride),” Clin. Toxicol., 14(1), pp. 1–46.
Reactions,” Carbon, 110, pp. 17–26. [165] Liu, T., Wie, X., Nie, Z., Sprenkle, V., and Wang, W., 2016, “A Total Organic
[138] Gonzalez, Z., Flox, C., Blanco, C., Granda, M., Morante, J. R., Menendez, R., Aqueous Redox Flow Battery Employing a Low Cost and Sustainable Methyl
and Santanmarıa, R., 2017, “Outstanding Electrochemical Performance of a Viologen Anolyte and 4-HO-TEMPO Catholyte,” Adv. Energy Mater., 6(3),
Graphene-Modified Graphite Felt for Vanadium Redox Flow Battery p. 1501449.
Application,” J. Power Sources, 338, pp. 155–162. [166] Janoschka, T., Martin, N., Hager, M. D., and Schubert, U. S., 2016, “An Aque-
[139] Blasi, A. D., Busaccaa, C., Blasia, O. D., Briguglioa, N., Squadritoa, G., and ous Redox-Flow Battery With High Capacity and Power: The TEMPTMA/
Antonuccia, V., 2017, “Synthesis of Flexible Electrodes Based on Electrospun MV System,” Angew. Chem. Int. Ed., 55(46), pp. 14427–14430.
Carbon Nanofibers With Mn3O4 Nanoparticles for Vanadium Redox Flow [167] Janoschka, T., Hager, M. D., and Schubert, U. S., 2012, “Powering Up the
Battery Application,” Appl. Energy, 190, pp. 165–171. Future: Radical Polymers for Battery Applications,” Adv. Mater., 24(48),
[140] Busacca, C., Blasi, O. D., Briguglio, N., Ferraro, M., Antonucci, V., and Blasi, pp. 6397–6409.
A. D., 2017, “Electrochemical Performance Investigation of Electrospun [168] Wei, X., Xu, W., Vijayakumar, M., Cosimbescu, L., Liu, T., Sprenkle,
Urchin-Like V2O3-CNF Composite Nanostructure for Vanadium Redox Flow V., and Wang, W., 2014, “TEMPO-Based Catholyte for High-Energy
Battery,” Electrochim. Acta, 230, pp. 174–180. Density Nonaqueous Redox Flow Batteries,” Adv. Mater., 26(45), pp.
[141] Bhattarai, A., Wai, N., Schweiss, R., Whitehead, A., Lim, T. M., and Hug, H. 7649–7653.
H., 2017, “Advanced Porous Electrodes With Flow Channels for Vanadium [169] Takechi, K., Kato, Y., and Hase, Y., 2015, “A Highly Concentrated Catholyte
Redox Flow Battery,” J. Power Sources, 341, pp. 83–90. Based on a Solvate Ionic Liquid for Rechargeable Flow Batteries,” Adv.
[142] Winsberg, J., Hagemann, T., Janoschka, T., Hager, M., and Schubert, U. S., Mater., 27(15), pp. 2501–2506.
2017, “Redox-Flow Batteries: From Metals to Organic Redox-Active Materi- [170] Lin, K., G omez-Bombarelli, R., Beh, E. S., Tong, L., Chen, Q., Valle, A.,
als,” Angew. Chem. Int. Ed., 56(3), pp. 686–711. Aspuru-Guzik, A., Aziz, M. J., and Gordon, R. G., 2016, “A Redox-Flow
[143] Jeftic, L., and Manning, G., 1970, “A Survey on the Electrochemical Reduc- Battery With an Alloxazinebased Organic Electrolyte,” Nature Energy, 1(9),
tion of Quinones,” J. Electroanal. Chem., 26(2–3), pp. 195–200. p. 16102.
[144] Ji, X., Banks, C. E., Silvester, D. S., Wain, A. J., and Compton, R. G., 2007, [171] Janoschka, T., Morgenstern, S., Hiller, H., Friebe, C., Wolkersd€ orfer, K.,
“Electrode Kinetic Studies of the Hydroquinone-Benzoquinone System and H€aupler, B., Hager, M. D., and Schubert, U. S., 2015, “Synthesis and Charac-
the Reaction Between Hydroquinone and Ammonia in Propylene Carbonate: terization of TEMPO- and Viologen-Polymers for Water-Based Redox-Flow
Application to the Indirect Electroanalytical Sensing of Ammonia,” J. Phys. Batteries,” Polym. Chem., 6(45), pp. 7801–7811.
Chem. C, 111(3), pp. 1496–1504. [172] Winsberg, J., Muench, S., Hagemann, T., Morgenstern, S., Janoschka, T., Bill-
[145] Rene, A., and Evans, D. H., 2012, “Electrochemical Reduction of Some o- ing, M., Schacher, F. H., Hauffman, G., Gohy, J. F., Hoeppener, S., Hager, M.
Quinone Anion Radicals: Why Is the Current Intensity So Small?,” J. Phys. D., and Schubert, U. S., 2016, “Polymer/Zinc Hybrid-Flow Battery Using
Chem. C, 116(27), pp. 14454–14460. Block Copolymer Micelles Featuring a TEMPO Corona as Catholyte,” Polym.
[146] Xu, Y., Wen, Y., Cheng, J., Cao, G., and Yang, Y., 2009, “Study on a Single Chem., 7(9), pp. 1711–1718.
Flow Acid Cd-Chloranil Battery,” Electrochem. Commun., 11(7), [173] Winsberg, J., Hagemann, T., Muench, S., Friebe, C., H€aupler, B., Janoschka,
pp. 1422–1424. T., Morgenstern, S., Hager, M. D., and Schubert, U. S., 2016, “Poly(Boron-

010801-20 / Vol. 15, FEBRUARY 2018 Transactions of the ASME

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Dipyrromethene): A Redox-Active Polymer Class for Polymer Redox-Flow [196] Brunini, V. E., Chiang, Y.-M., and Carter, W. C., 2012, “Modeling the Hydro-
Batteries,” Chem. Mater., 28(10), pp. 3401–3405. dynamic and Electrochemical Efficiency of Semi-Solid Flow Batteries,” Elec-
[174] Cabana, J., Monconduit, L., Larcher, D., and Palacin, M. R., 2010, “Beyond trochim. Acta, 69, pp. 301–307.
Intercalation-Based Li-Ion Batteries: The State of the Art and Challenges of [197] Johnson, D., and Reid, M., 1985, “Chemical and Electrochemical Behavior of
Electrode Materials Reacting Through Conversion Reactions,” Adv. Mater., the Cr(III)/Cr(II) Half-Cell in the Iron-Chromium Redox Energy Storage Sys-
22(35), pp. E170–E192. tem,” J. Electrochem. Soc., 132(5), pp. 1058–1062.
[175] Dunn, B., Kamath, H., and Tarascon, J.-M., 2011, “Electrical Energy Storage [198] Giner, J., Swette, L., and Cahill, K., 1976, “Screening of Redox Couples and
for the Grid: A Battery of Choices,” Science, 334(6058), pp. 928–935. Electrode Materials,” NASA Lewis Research Center, Cleveland, OH, Report
[176] Zhang, H., Mao, C., Li, J., and Chen, R., 2017, “Advances in Electrode Materi- No. NASA-CR-134705.
als for Li-Based Rechargeable Batteries,” RSC Adv., 7(54), pp. 33789–33811. [199] Ciprios, G., Erskine, W., and Grimes, P. G., 1977, “Redox Bulk Energy Stor-
[177] Yamada, A., Chung, S.-C., and Hinokuma, K., 2001, “Optimized LiFePO4 for age System Study,” Exxon Research and Engineering Co., Government
Lithium Battery Cathodes,” J. Electrochem. Soc., 148(3), pp. A224–A229. Research Labs, Linden, NJ, Technical Report No. NASA-CR-135206.
[178] Yoo, E., Kim, J., Hosono, E., Zhou, H.-S., Kudo, T., and Honma, I., 2008, [200] Michaels, K., and Hall, G., 1980, “Cost Projections for Redox Energy Storage
“Large Reversible Li Storage of Graphene Nanosheet Families for Use in Systems,” United Technologies Corp., Power Systems Div., South Windsor,
Rechargeable Lithium Ion Batteries,” Nano Lett., 8(8), pp. 2277–2282. CT, Technical Report No. NASA-CR-165260.
[179] Ventosa, E., Zampardi, G., Flox, C., La Mantia, F., Schuhmann, W., and [201] Warshay, M., and Wright, L. O., 1977, “Cost and Size Estimates for a Redox
Morante, J., 2015, “Solid Electrolyte Interphase in Semi-Solid Flow Bat- Bulk Energy Storage Concept,” J. Electrochem. Soc., 124(2), pp. 173–177.
teries: A Wolf in Sheep’s Clothing,” Chem. Commun., 51(81), pp. [202] Price, A., Bartley, S., Male, S., and Cooley, G., 1999, “A Novel Approach to
14973–14976. Utility Scale Energy Storage [Regenerative Fuel Cells],” Power Eng. J., 13(3),
[180] Biendicho, J. J., Flox, C., Sanz, L., and Morante, J. R., 2016, “Static and pp. 122–129.
Dynamic Studies on LiNi1/3Co1/3Mn1/3O2-Based Suspensions for Semi–Solid [203] Shigematsu, T., 2011, “Redox Flow Battery for Energy Storage,” SEI. Tech.
Flow Batteries,” ChemSusChem, 9(15), pp. 1938–1944. Rev., 73, pp. 4–13.
[181] Wei, T. S., Fan, F. Y., Helal, A., Smith, K. C., McKinley, G. H., Chiang, Y. [204] Shi, F., 2014, “Design of Flow Battery,” Reactor and Process Design in
M., and Lewis, J. A., 2015, “Biphasic Electrode Suspensions for Li-Ion Semi- Sustainable Energy Technology, Elsevier, Amsterdam, The Netherlands, pp.
Solid Flow Cells With High Energy Density, Fast Charge Transport, and Low- 72–75.
Dissipation Flow,” Adv. Energy Mater., 5(15), p. 1500535. [205] Skyllas-Kazacos, M., 2010, “Energy Storage for Stand-Alone/Hybrid Systems:
[182] Qi, Z., Liu, A. L., and Koenig, G. M., 2017, “Carbon-Free Solid Dispersion Electro-Chemical Energy Storage Technologies,” Stand-Alone and Hybrid
LiCoO2 Redox Couple Characterization and Electrochemical Evaluation Wind Systems: Technology, Energy Storage and Applications, Woodhead Pub-
for All Solid Dispersion Redox Flow Batteries,” Electrochim. Acta, 228, lishing, Cambridge, UK.
pp. 91–99. [206] Chan, K.-Y., and Vanessa Li, C.-Y., 2014, Electrochemically Enabled Sustain-
[183] Qi, Z., and Koenig, G. M., 2016, “A Carbon-Free Lithium-Ion Solid Disper- ability: Devices, Materials and Mechanisms For Energy Conversion, CRC
sion Redox Couple With Low Viscosity for Redox Flow Batteries,” J. Power Press, Boca Raton, FL, pp. 380–384.
Sources, 323, pp. 97–106. [207] Tokuda, N., Kumamoto, T., Shigematsu, T., Deguchi, H., Ito, T., Yoshikawa,
[184] Li, Z., Smith, K. C., Dong, Y., Baram, N., Fan, F. Y., Xie, J., Limthongkul, P., N., and Hara, T., 1998, “Development of a Redox Flow Battery System,” SEI.
Carter, W. C., and Chiang, Y.-M., 2013, “Aqueous Semi-Solid Flow Cell: Dem- Tech. Rev., 45, pp. 88–94.
onstration and Analysis,” Phys. Chem. Chem. Phys., 15(38), pp. 15833–15839. [208] Alotto, P., Guarnieri, M., and Moro, F., 2014, “Redox Flow Batteries for the
[185] Ventosa, E., Buchholz, D., Klink, S., Flox, C., Chagas, L. G., Vaalma, C., Storage of Renewable Energy: A Review,” Renewable Sustainable Energy
Schuhmann, W., Passerini, S., and Morante, J. R., 2015, “Non-Aqueous Rev., 29, pp. 325–335.
Semi-Solid Flow Battery Based on Na-Ion Chemistry. P2-Type [209] Chalamala, B. R., Soundappan, T., Fisher, G. R., Anstey, M. R., Viswanathan,
NaxNi0.22Co0.11Mn0.66O2–NaTi2(PO4)3,” Chem. Commun., 51(34), pp. V. V., and Perry, M. L., 2014, “Redox Flow Batteries: An Engineering
7298–7301. Perspective,” Proc. IEEE, 102(6), pp. 976–999.
[186] Fang, Y., Zhang, J., Xiao, L., Ai, X., Cao, Y., and Yang, H., 2017, “Phosphate [210] Hoberecht, M. A., and Thaller, L. H., 1982, “Design Flexibility of Redox
Framework Electrode Materials for Sodium Ion Batteries,” Adv. Sci., 4(5), Flow Systems,” NASA Lewis Research Center, Cleveland, OH, Technical
p. 1600392. Report No. NASA-TM-82854.
[187] Ding, Y., Zhao, Y., Li, Y., Goodenough, J. B., and Yu, G., 2017, “A High- [211] Aaron, D. S., Liu, Q., Tang, Z., Grim, G. M., Papandrew, A. B., Turhan, A.,
Performance All-Metallocene-Based, Non-Aqueous Redox Flow Battery,” Zawodzinski, T. A., and Mench, M. M., 2012, “Dramatic Performance Gains
Energy Environ. Sci., 10(2), pp. 491–497. in Vanadium Redox Flow Batteries Through Modified Cell Architecture,” J.
[188] Pan, F., Huang, Q., Huang, H., and Wang, Q., 2016, “High-Energy Density Power Sources, 206, pp. 450–453.
Redox Flow Lithium Battery With Unprecedented Voltage Efficiency,” Chem. [212] Perrry, M. L., 2009, “Flow Battery With Interdigitated Flow Field,” United
Mater., 28(7), pp. 2052–2057. Technologies Corporation, Farmington, CT, U.S. Patent No. US9166243 B2.
[189] Huang, Q., Li, H., Gr€atzel, M., and Wang, Q., 2013, “Reversible Chemical [213] Di Noto, V., Guarnieri, M., and Moro, F., 2010, “A Dynamic Circuit Model of
Delithiation/Lithiation of LiFePO4: Towards a Redox Flow Lithium-Ion a Small Direct Methanol Fuel Cell for Portable Electronic Devices,” IEEE
Battery,” Phys. Chem. Chem. Phys., 15(6), pp. 1793–1797. Trans. Ind. Electron., 57(6), pp. 1865–1873.
[190] Ding, Y., Zhao, Y., and Yu, G., 2015, “A Membrane-Free Ferrocene-Based [214] Ma, X., Zhang, H., and Xing, F., 2011, “A Three-Dimensional Model for Neg-
High-Rate Semiliquid Battery,” Nano Lett., 15(6), pp. 4108–4113. ative Half-Cell of the Vanadium Redox Flow Battery,” Electrochim. Acta, 58,
[191] Tomai, T., Saito, H., and Honma, I., 2017, “High-Energy-Density Electro- pp. 238–246.
chemical Flow Capacitors Containing Quinone Derivatives Impregnated in [215] Viswanathan, V., Wang, W., Li, B., Coffey, G., Thomsen, E., Graff, G., Bal-
Nanoporous Carbon Beads,” J. Mater. Chem. A, 5(5), pp. 2188–2194. ducci, P., Kintnereyer, M., and Sprenkle, V., 2014, “Cost and Performance
[192] Yang, Y., Zheng, G., and Cui, Y., 2013, “A Membrane-Free Lithium/Polysul- Model for Redox Flow Batteries,” J. Power Sources, 247, pp. 1040–1051.
fide Semi-Liquid Battery for Large-Scale Energy Storage,” Energy Environ. [216] Li, B., Luo, Q., Wei, X., Nie, Z., Thomsen, E., Chen, B., Sprenkle, V., and
Sci., 6(5), pp. 1552–1558. Wang, W., 2014, “Capacity Decay Mechanism of Microporous Separator-
[193] Chen, H., Zou, Q., Liang, Z., Liu, H., Li, Q., and Lu, Y.-C., 2015, “Sulphur- Based All-Vanadium Redox Flow Batteries and Its Recovery,” ChemSu-
Impregnated Flow Cathode to Enable High-Energy-Density Lithium Flow sChem, 7(2), pp. 577–584.
Batteries,” Nat. Commun., 6, p. 5877. [217] Cho, K. T., Albertus, P., Battaglia, V., Kojic, A., Srinivasan, V., and Weber,
[194] Chen, H., and Lu, Y. C., 2016, “A High-Energy-Density Multiple Redox A. Z., 2013, “Optimization and Analysis of High-Power Hydrogen/Bromine-
Semi-Solid-Liquid Flow Battery,” Adv. Energy Mater., 6(8), p. 1502183. Flow Batteries for Grid-Scale Energy Storage,” Energy Technol., 1(10),
[195] Hamelet, S., Larcher, D., Dupont, L., and Tarascon, J.-M., 2013, “Silicon- pp. 596–608.
Based Non Aqueous Anolyte for Li Redox-Flow Batteries,” J. Electrochem. [218] Perry, M. L., Darling, R. M., and Zaffou, R., 2013, “High Power Density
Soc., 160(3), pp. A516–A520. Redox Flow Battery Cells,” ECS Trans., 53(7), pp. 7–16.

Journal of Electrochemical Energy Conversion and Storage FEBRUARY 2018, Vol. 15 / 010801-21

Downloaded From: http://electrochemical.asmedigitalcollection.asme.org/ on 10/10/2017 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Você também pode gostar