You are on page 1of 29

Materials Today Energy 8 (2018) 80e108

Contents lists available at ScienceDirect

Materials Today Energy


journal homepage: www.journals.elsevier.com/materials-today-energy/

Review of zinc-based hybrid flow batteries: From fundamentals to


applications
A. Khor a, b, 1, P. Leung c, 1, M.R. Mohamed b, C. Flox a, Q. Xu d, L. An e, R.G.A. Wills f,
J.R. Morante a, A.A. Shah g, *
a
IREC, Catalonia Institute for Energy Research, Sant Adria de Beso
s, Spain
b
Faculty of Electrical & Electronics Engineering, Universiti Malaysia Pahang, Pekan, Malaysia
c
Department of Materials, University of Oxford, Oxford, UK
d
Institute for Energy Research, Jiangsu University, Zhenjiang, China
e
Department of Mechanical Engineering, Hong Kong Polytechnic University, Hong Kong
f
Energy Technology Group, University of Southampton, Southampton, UK
g
School of Engineering, University of Warwick, UK

a r t i c l e i n f o a b s t r a c t

Article history: Zinc-based hybrid flow batteries are one of the most promising systems for medium- to large-scale
Received 18 September 2017 energy storage applications, with particular advantages in terms of cost, cell voltage and energy den-
Received in revised form sity. Several of these systems are amongst the few flow battery chemistries that have been scaled up and
21 November 2017
commercialized. The existing zinc-based systems rely on zinc electrodeposition in flowing electrolytes as
Accepted 26 December 2017
the negative electrode reaction, which is coupled with organic or inorganic positive active species in
either solid, liquid or gaseous phases. These reactions are facilitated with specific cell architectures under
certain circumstances. To improve the performance and cycle life of these batteries, this review provides
Keywords:
Applications
fundamental information on zinc electrodeposition and summarizes recent developments in the relevant
Electrodeposition flow battery chemistries, along with recent applications. The future challenges and opportunities for this
Plating technology are discussed.
Redox flow batteries © 2017 Elsevier Ltd. All rights reserved.
Zinc

1. Introduction Despite the fact that the all-vanadium redox flow battery is the
most developed system, due to its high reversibility and relatively
Energy storage is a key component for enabling an increased the large power output, the electrolyte cost of such systems exceeds
share of power from renewables such as photovoltaic cells and USD$ 80/kW h [3,4]. The resulting capital cost can be as high as
wind turbines in electrical grids [1,2]. Among the various electro- USD$ 200e750/kW h, which is well beyond the cost target (USD$
chemical energy storage technologies, redox flow batteries (RFBs) 150/kW h by 2023) set by the USA Department of Energy (DoE) to
are considered to be the most realistic candidates for energy stor- match with existing physical energy storage technologies [5]. In
age in the range of several kW/kW h up to tens of MW/MW h. order to reduce the costs of redox flow battery systems, recent
Conventional redox flow batteries, such as the all-vanadium bat- investigations have highlighted the use of low-cost active materials
teries, store energy in the electrolytes in the form of reduced and sourced from both metallic and non-metallic (i.e. organic) com-
oxidized electroactive species, while at least one of the electrode pounds [6e8] or through other strategies [9,10]. Developments in
reactions of the hybrid flow batteries involve a phase change (solid organic-based redox flow battery are still relatively recent (since
or gaseous). Common examples are those systems based on metal 2009 [11]) and further substantial advances are required to over-
electrodeposition and redox reactions of gaseous hydrogen and come the low energy densities of these systems (<15 Wh dm3);
oxygen. the energy densities are restricted by both the solubilities of the
active compounds and the low cell voltages of the existing systems
[6e8,12], particularly due to the negative electroactive species [13].
Table 1 compares the cost (elements and electrolytes) and typical
* Corresponding author.
E-mail address: Akeel.Shah@warwick.ac.uk (A.A. Shah).
energy density of the aforementioned systems with different zinc-
1
These authors contributed equally to this work. based hybrid flow batteries. It should be noted that the electrolyte

https://doi.org/10.1016/j.mtener.2017.12.012
2468-6069/© 2017 Elsevier Ltd. All rights reserved.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 81

Table 1
Costs and energy densities of different aqueous flow battery systems.

Flow battery systems Cell voltage Negative species cost Positive species cost Electrolyte costc/USD$ Typical energy
(open-circuit)/V (element)a/USD$ Kg1 (element)a/USD$ Kg1 (kW h)1 density/W h L1 [14]

ZneFe 1.5 1.9 0.8 5 e


ZneCe 2.4 1.9 12 42 12e20
ZneI 1.3 1.9 13.5b 41 Up to 101
Zn-Organic 1.6 1.9 5 14 2e5
(Benzoquinone)
Zn-air 1.6 1.9 Nil 4 e
ZneBr 1.8 1.9 2.0 5 Up to 65
VeV 1.4 24 24 87 15e43
All-Organic 1.2 5 (est.) 5 (est.) 92 0.5e5
(Viologen-TEMPO)
a
Element costs are from the latest USGS Materials Databook [15].
b
The cost of iodine could reach USD$ 23 Kg-1, including insurance and freight value as estimated in 2016.
c
Estimated by assuming 80% voltage efficiencies for zinc-based flow batteries; 90% voltage efficiency for all-vanadium redox flow batteries; 50% electrolyte utilization for
electrodeposition reaction; 80% electrolyte utilization for liquid-phase redox reaction.

cost is also highly sensitive to the molecular weight and number lithium and sodium) on a large scale. Together with its low-cost
of electron-transfers. For the case of all-organic viologen-TEMPO nature, zinc remains one of the most common anode materials
system, the molecular weight of these molecules are more for primary and secondary batteries. Primary systems based on a
than 170 g mol1 (vs. 65.38 g mol1 for zinc), while involving zinc anode were in the early stages of development e Leclanche 
just one electron-transfer. On the other hand, other organic wet cells (zinc-manganese dioxide) in 1866 [18]. Zinc-carbon and
molecules, such as 1,4-benzoquinone, have smaller molecular zinc-alkaline dry cells (zinc-manganese dioxide) are still used in
weight (108.1 g mol1) and exhibit two electron-transfers. There- various portable devices [19]. These batteries are often associated
fore, the electrolyte cost of most zinc-based systems (<USD$ 42/kW with one-time use as primary cells, although rechargeable systems
h) are lower than all-vanadium (<USD$ 87/kW h) and the recent are available. The poor reversibility was attributed to the cycling-
all-organic viologen-TEMPO (USD$ 92/kW h) systems. induced changes in the materials in both anode and cathode
The choice of low-cost metals (<USD$ 4 kg1) is still limited to (manganese dioxide) materials [38,39]. In conventional static sys-
zinc, lead, iron, manganese, cadmium and chromium for redox/ tems, dendritic morphologies of zinc anodes tend to develop at
hybrid flow battery applications. Many of these metals are highly limiting current densities, as a consequence of non-uniform con-
abundant in the earth's crust (>10 ppm [16]) and annual production centration gradients [40]. Controlling mass transport by convec-
exceeds 4 million tons (2016) [17]. Their widespread availability tion, as in a flow battery configuration, has been demonstrated to
and accessibility make these elements attractive for large-scale be an effective approach to improve the morphology and extend
energy storage applications. With the exceptions of iron (Fe(II)/ the cycle life by minimizing dendritic growth, shape change and
Fe(III)) and manganese (MnO2/MnOOH), most of these metallic passivation of the zinc electrodes [40,41].
elements involve electrodeposition from dissolved species and Since the 1970s, various zinc-based flow batteries have been
mainly serve as negative electrode species. Among these elements, proposed and developed by coupling with different positive elec-
metallic zinc has the highest energy content due to its large volu- trode reactions [42]. Together with the all-vanadium system, zinc-
metric capacity (5.85 Ah cm3) and negative electrode potential in based systems are one of the few flow battery chemistries to be
aqueous media (acidic: 0.76 V vs. SHE; alkaline: 1.29 V vs. SHE) scaled-up and commercialized, for various applications. The exist-
[18,19]. In general, such negative electrode potentials imply ing zinc-based systems use positive electrode reactions based on
hydrogen evolution from water electrolysis. However, zinc elec- inorganic or organic active materials in either solid, liquid or
trodeposition on inert substrates has been demonstrated as a gaseous phases as demonstrated in Fig. 1. These reactions are
relatively efficient process in aqueous electrolytes (both acidic and facilitated with specific cell architectures under different circum-
alkaline) with current efficiencies of over 90%; this is attributed to stances. Unlike other proposed systems [43,44], most zinc-based
the large hydrogen overpotential and suitable electrolyte compo- flow batteries do not require catalysts for typical operations. In all
sitions. Zinc electrodeposition is well established in the industry systems, the negative electrode reaction relies on zinc electrode-
and has been used extensively in electroplating, corrosion protec- position, which takes place in flowing electrolytes as in other
tion and automotive vehicles [16,17,20e22]. At present, conven- hybrid flow batteries. This highlights the importance of the zinc
tional baths for zinc electrodeposition are still mainly based on electrodeposition process and the importance of obtaining suitable
zincate (alkaline) [23], chloride (acid) [24e26] and sulfate (acid) deposit morphologies for long-term cycling.
[27e35] in the electroplating industry. In many of these systems, separators or membranes are used to
In contrast, electrodeposition of other electronegative metals, prevent direct reaction or crossover of the active species between
particularly lithium, sodium and aluminum, are impossible in the two half-cell compartments. However, the involvement of at
aqueous electrolytes and have to be carried out in non-aqueous least one solid-phase electrode reaction (including zinc electrode-
solvents or room-temperature ionic-liquids [36]. For large-scale position) in zinc-based systems provides the possibility of a
energy storage devices, such as redox flow batteries, these sol- ‘membrane-less’ configuration. In certain chemistries, the positive
vents are less attractive in terms of cost and ionic conductivity electrode reactions do not involve soluble species and undergo
(108e1010 S cm1 without salts). The low power density resulting pure solid-phase transformation, while other chemistries involve
from poor ionic conductivity implies larger electrode size or more slow dissolution of the zinc electrodeposit in the presence of
cells in a system required for a given power output, which leads to a certain active species dissolved in the electrolytes, particularly at
significant increase in overall cost per kW h [37]. low concentrations.
Furthermore, many of these solvents are flammable, raising However, most of these zinc-based systems still face the chal-
concerns over safety when used with reactive metal anodes (e.g., lenges of self-discharge, electrode shape-change and formation of
82 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

Fig. 1. (a) Zinc system with solid phase reaction, (b) liquid state reaction, (c) and gaseous phase reaction.

dendrites. Such effects need to be further minimized for improved ores [57]. The atomic number and weight of this element are 30 and
performance and cycle life. In order to advance the existing tech- 65.38, respectively. Pure zinc is neither particularly ductile nor
nologies, it is essential to improve the understanding of zinc elec- malleable. Its density is 7.14 g cm3, it has an electrical resistivity of
trodeposition. A number of review articles have been published in 6.16 mU cm1 and a melting point of 419.5  C [58]. In general, the
the field of redox flow batteries, regarding general chemistries standard electrode potentials of zinc are 0.76 V and 1.29 V in
[1,2,37,45e49], cell architectures [50], mathematical modelling [51] acidic and alkaline media, respectively. Given the dependence of
and cost analysis [4,52]. Specific systems based on attractive ele- chemical activities on other thermodynamic parameters, the
ments, such as vanadium [53e55], lithium [56] and organic ma- equilibrium electrode potential separates the oxidized/reduced
terials [6e8], have also been reviewed. In contrast to previous phases on a phase diagram.
reviews, the present contribution provides an overview of the zinc In Fig. 2, we show a Pourbaix diagram (a certain class of phase
electrodeposition process and a comprehensive summary of the diagram), which plots the equilibrium potential against pH in
existing zinc-based flow battery technologies and their recent ap- aqueous electrolytes. This Pourbaix diagram is based on relatively
plications. The latest advances, future challenges and opportunities low zinc ion concentration (i.e. 106 M). This kind of diagram is
for further development are discussed. useful in determining which species are thermodynamically stable
at a given potential and pH but does not provide information about
2. Electrochemistry of zinc electrodeposition the kinetics of the reaction processes.
Pourbaix diagrams for zinc have been published by a number of
Zinc is a moderately reactive bluish-grey metal; it can be found authors and vary with temperature and electrolyte composition
in nature as sphalerite, smithsonite, hemimorphite and franklite [59e61]. For most existing zinc-based hybrid flow batteries, charge
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 83

ZnðOHÞ2 4HZnO2  þ Hþ (6)

At pH ≈ >13

HZnO2  4ZnO2
2 þH
þ
(7)
In the presence of different anions or salts (e.g., chloride, bro-
mide), the formations of these complexes could take place at
slightly different pH values. For instance, in the presence of chlo-
ride, the Pourbaix diagram is still very similar to that in pure water.
However, in the presence of bromide salts, the region of Zn2þ
species (ZnBr 3 ) stability is c.a. pH 6 (vs. c.a. pH 8.5 in equation (5)).
Hydrogen reactions (evolution or oxidation) could compete or
pair with zinc electrodeposition/dissolution processes to form a
local net reaction. When zinc or other metals M (particularly those
of the current collectors) are exposed to the electrolytes containing
protons/acid, dissolution of these metals take place as a form of
Fig. 2. Pourbaix diagram for zinc in different pH in water [60]. Note: low zinc ion corrosion alongside hydrogen evolution at open-circuit due to the
concentrations (i.e. 106 M). following mixed electrode reactions on the metallic surface [62]:

or electrical energy is stored mainly as an electrodeposit in the M  ne 4Mnþ (8)


reduction process. As indicated in Fig. 2, metallic zinc is electro-
deposited via four different routes with reactants in the forms of
2Hþ þ 2e 4H2 (9)
Zn2þ, Zn(OH)2, HZnO 2
2 and ZnO2 . Depending on the pH of the
aqueous electrolyte, Zn(OH)2, HZnO 2
2 and ZnO2 are often formed In common with protons, metallic zinc reacts with the active
by the hydrolysis of Zn2þ. It is important to note that the traditional species oxidized by the positive electrodes, resulting in energy loss.
notations of HZnO 2
2 and ZnO2 in most Pourbaix diagrams corre- For this reason, separators are used to prevent crossover of the active
spond to the hydrolysed zinc complexes of Zn(OH)-3 and Zn(OH)2 4 species between the two half-cells. For a few cases [63e65], no
used in most battery equations. membrane is required when these positive electrode reactions do not
The Nernst equation can be written for each of these reactants in involve soluble species and undergo solid-phase transformation
equilibrium with metallic zinc. In water at different pH: (NiOOH/Ni(OH)2 [63,64], PbO2/PbSO4 [65]). Furthermore, metallic
zinc can be protected by mean of cathodic protection in the electro-
Zn2þ þ 2e 4Zn h i deposition process. Therefore, the dissolution of zinc takes place only
(1) at open-circuit or during the discharge process. Based on this concept,
Ee ¼ 0:763 þ 0:0295 log Zn2þ
membrane-less systems [66e69] could be possible and have been
introduced recently by exploiting the slow dissolution of zinc in the
ZnðOHÞ2 þ 2Hþ þ 2e 4Zn þ 2H2 O presence of particular active species at low concentration (<0.5 M).
(2) However, corrosion or dissolution of the metallic components of
Ee ¼ 0:439  0:0591pH
redox flow batteries is detrimental to long-term operation and
performance. The dissolved metal ions could lead to unstable redox
HZnO2 þ 3Hþ þ 2e 4Zn þ 2H2 O h i potentials and distort the original chemistries of the flow battery.
(3)
Ee ¼ 0:54V  0:0886pH þ 0:0295log HZnO2 As a result, and also due to the weight issues, metallic components
are not commonly used in most flow battery systems. For this
reason, carbon and polymer based materials are used as the elec-
þ 
ZnO 2
2 þ 4 H þ 2e 4Zn þ 2H2 OEe
trode and cell components, respectively, since these materials do
h i not undergo dissolution or formation of oxides in the oxidation
¼ 0:441V  0:1182pH þ 0:0295log ZnO2 2 (4) process [70,71]. Although carbon-based materials can be eroded
during the oxidation reactions by means of carbon dioxide evolu-
The Nernst equation for reaction (1) shows that it is indepen- tion [72], these materials are suitable for most negative electrode
dent on pH and corresponds to a horizontal line at 0.763 V vs. SHE reactions even under excessive hydrogen evolution [73]. In contrast
in the Pourbaix diagram. In contrast, the other electrodeposition to the electroplating industry, where the majority of substrates are
routes (reactions (2e4)) vary linearly with pH with negative gra- metals, carbon-based substrates are extensively used in hybrid flow
dients. Based on these equations, the Pourbaix diagram can be batteries. Since pure carbon/graphite is brittle and often difficult to
divided into four regions, corresponding to different reactants be- scale up in stacks, composites with polymer binders and conductive
ing thermodynamically stable in the electrodeposition processes. particles [74e77] are often used instead [74e77].
These vertical lines between the four regions are independent of In typical linear sweep voltammetry, zinc electrodeposition can
electrode potential and can be determined by considering the take place in three regions under the conditions that electrolyte
equilibrium constant [60]: flow is static and hydrogen evolution is not particularly dominant
(Fig. 3). In region I, the initial deposition rate is usually slow and
At pH ≈ 8e10 under charge transfer control. In region II, the current density in-
creases, before decreasing dramatically at more negative electrode
Zn2þ þ 2H2 O4ZnðOHÞ2 þ 2Hþ (5) potentials, resulting in a cathodic peak due to diffusion control.
Subsequently, a sharp increase in current density is obtained in
At pH ≈ 10e12 region III associated with hydrogen evolution together with the
84 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

with the study of Raeissi et al. [85], which concluded that


increasing overpotential results in a decrease in basal planes and an
increase in low angle planes in zinc sulphate electrolytes (pH 2).
Zinc electrodeposits are generally classified into 5 categories:
heavy spongy, dendritic, boulder, layer-like, and filamentous mossy
[86]. Based on the electrodeposit crystallinity and microstructure,
these 5 categories are summarized in Table 2, while microscopic
images of these distinct morphologies are shown in Fig. 5 (elec-
trodeposits obtained from alkaline electrolytes). Based on the
example of zincate solutions, heavy spongy morphologies are large
boulder agglomerates and highly branched dendrites, developing
at high current densities and after prolonged deposition. The
mechanisms for typical hexagonal crystals organizing into den-
dritic structure and the growth of these branches into three-
dimensional threadlike networks have been discussed and illus-
trated in detail by Chamoun et al. [40]. Boulders are typically
hexagonal shaped in discrete assemblies. Layer-like structures are
formed by epitaxial growth in the early stages of the deposition
process, while hexagonal platelets have been reported extensively
in acidic solutions [34,87,88]. Filamentous mossy deposits have the
appearance of tangled whiskers with a typical diameter of
Fig. 3. Linear sweep voltammetry for Zn(Ac)2, ZnSO4 and ZnCl2 [78,79].
50e200 nm and lengths that may exceed 5 mm.
electrodeposition process. In this region, zinc electrodeposition is In zinc bromide solutions, mossy morphologies are often
predominantly under mass transport control rather than charge observed in weakly acidic and basic conditions [89,90], and are
transfer control. When the electrode potential sweeps towards considered undesirable for battery operations. To address the
even more negative values, hydrogen evolution becomes dominant morphological issues, organic and inorganic additives are commonly
and appears as a significant side reaction [78]. used in both electroplating and battery applications. In general,
These cathodic potentials can vary with different electrolyte organic additives increase the electrode overpotentials by
compositions, particularly the anion concentration and pH. Fig. 3 their adsorption or partial covering of the electrode surfaces
shows the cyclic voltammetry of zinc electrodeposition on a [33e35,41,87,88], which prevents the growth of the zinc crystals. On
glassy carbon electrode from different anions (acetate, chloride and the contrary, some inorganic additives, such as indium oxide, may
sulphate) in neutral solutions [79]. In each curve, the forward and exhibit high hydrogen overpotentials [62,91]. Despite the improved
reverse scans form a nucleation loop between the cathodic and or more uniform morphologies, zinc electrodeposition may also be
anodic peaks, which has been ascribed to the nucleation process inhibited by some organic additives (e.g., surfactants), resulting in
requiring an activation energy provided by an overpotential [80]. It low half-cell coulombic efficiencies [41,88]. However, certain elec-
can be seen that the cathodic peak potential of the zinc chloride trodeposits are found to contain a high content of carbon due to the
solution is similar to that with acetate anions, while the peak cur- presence of organic additives and possibly impurities in the elec-
rent of the chloride solution is the largest among the three solu- trodeposition process [92,93].
tions. This demonstrates that the effect of chloride ions (Cl) on the
zinc electrodeposition overpotential is similar to that of acetate 3. Improving zinc electrodeposit morphologies for hybrid
ions on a glassy carbon substrate. In contrast, the cathodic peak flow batteries
potential of the sulphate solution is the most negative (about
50 mV more negative of the peaks in the other two solutions). This Improved zinc electrodeposit morphologies are essential for
suggests a more difficult electrodeposition process in sulphate so- long-term charge-discharge cycling. Dendritic growth is an unde-
lutions, which can be explained by the strong interaction between sirable feature of zinc electrodeposition. It entails the risk of
sulphate and zinc (II) ions, as reflected in its positive stability damage to the separator and possibly an electrical short-circuit or
constant (ln b ¼ 2.3). The deposition overpotential is considered to self-discharge. Dendritic growth usually develops from mossy to
be the driving force for electrocrystallization at the electrode sur- pyramidal/acicular form.
face, which includes adsorption, nucleation and growth processes The overall growth and propagation of dendrites is influenced
[24], and is sensitive to the zinc(II) concentration [24] as well as the by diffusion and the type of zinc complex [94,95]. There are usually
types of additives [32e35]. The mechanisms for these phenomena two main sources of convective diffusion: (1) density differences
have been discussed previously in detail [81e84]. between the zinc ions and other compositions; and (2) local forced
In general, the competition between growth and nucleation convection driven by hydrogen evolution. The local convection
determines the granularity of the electrodeposit. The higher the leads to an enhanced rate of dendrite propagation, while formation
nucleation rate caused by the overpotential during the electrode- of dendrites is suppressed with the addition of electrolyte addi-
position process, the finer the crystal grains of the deposit. On the tives. Suitable operating conditions important crucial for obtaining
other hand, the forms of the crystals at the growth centres deter- favourable electrodeposit morphologies, which will be further
mine the general appearance and structure of the electrodeposits, discussed in latter sections (Section 5.). However, the main strate-
as illustrated in Fig. 4. Since the overpotential is the deviation of the gies of improving zinc electrodeposit morphologies are through (1)
electrode potential from its equilibrium value, poor control of the flowing electrolytes and (2) electrolyte additives.
potential distribution often results in undesirable morphologies
(e.g., dendrites) and growth rates [81,82]. 3.1. The use of flowing electrolytes
The relationships between overpotential, impurities and elec-
trolyte additives (organic), and the crystallographic growth of the Flowing electrolyte is an effective and necessary approach for
electrodeposits are shown schematically in Fig. 4. This is consistent zinc-based batteries to minimize the dendritic growth and extend
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 85

Fig. 4. (a) The growth mechanisms of zinc electrodeposition; (b) the relationships among overpotentials, impurtities and electrolyte additives and the crystallographic growth of
the zinc electrodeposits [81,82].

Table 2
Categories of zinc deposit morphology [86]].

Category of zinc deposit morphology

Property Heavy spongy Dendritic Boulder Layer-like Mossy

Appearance Black powder Metallic crystal Grey metal Reflective metal Black powder
Microstructure (under microscope) Agglomerate large boulder Fern, leaf, hexagon Granular boulder Ridge, layer like Filament, whisker
Adherence Non-adherent Non-adherent Adherent Adherent Non-adherent
Porosity Dispersed Dispersed Compact non-porous Compact non-porous Highly porous
Crystallinitya Anisotropically oriented Isotropically oriented Anisotropically oriented Epitaxial oriented Isotropically oriented
Growth current density Highest Very high Moderate Low Lowest
Nucleation site selectivity Non-selective Non-selective Selective Non-selective Highly selective
a
Anisotropically oriented: deposits grow three-dimensionally; bulk and individual deposits are polycrystalline. Isotopically oriented; deposits grow two-dimensionally;
individual deposits are single crystals.

the cycle life to more than several hundred cycles [96]. This is suppressed by flowing electrolytes, especially high current den-
because the flowing electrolyte can distribute the zinc active spe- sities and random cycling conditions are used [97].
cies more uniformly and decrease the width of Nernst diffusion
layer (dN). Under convective-diffusion mass transport control, the
corresponding limiting current density (jL) for zinc electrodeposi- 3.2. The use of electrolyte additives
tion therefore becomes higher considering (where F is the Faraday
constant (96,485 C mol1), D is the diffusion coefficient of zinc ions In the zinc electroplating industry, electrolyte additives are used
(cm2 s1), c0 is the zinc ion concentration in the bulk solution). It is to obtain finer grain sizes, and hence brighter electrodeposits [98].
important to note that dendritic morphologies tend to develop at These additives tend to promote 2-dimensional growth of the
limiting current densities, as a consequence of non-uniform con- crystallites [99]. Previous theories suggest that a higher deposition
centration gradients [40]. overpotential could lead to a faster nucleation rate and finer crystal
The use of flowing electrolyte has been seen as a breakthrough grain sizes [100]. Electrolyte additives, particularly those with large
for zinc-based rechargeable batteries [97]. Previous investigation organic molecules, i.e. surfactants, are often known as ‘brighteners’
reported that dendritic growth began to take place in static elec- (primary and carrier brighteners) in electroplating industry and
trolytes when the current density is as low as 8 mA cm2. This used to increase the deposition overpotential by partially covering
indicates that electrolyte circulation is necessary for typical current and selectively absorbing on the electrode surface, preventing
densities between 20 and 40 mA cm2. Depending on the operating further grain growth and promoting nucleation [101]. This also
conditions, practical experience showed that a linear flow velocity depends on the structure and the size of the additives, as well as the
of about 2 cm s1 within the stack can be sufficient to prevent the additive-substrate and additive-deposit interactions [102].
formation. However, under prolonged charged, dendritic growth Cationic surfactants have been commonly used in zinc electro-
may still take place eventually and may not be completely winning [103e105], including triethylbenzylammonium chloride
(TEBACl), tetrabutylammonium chloride (TBACl) and cetyltrimethyl
86 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

Fig. 5. Morphologies of zinc electrodepositions (heavy spongy, dendrite, boulder, layer like and filamentous mossy) on electrode surface (electrodeposited through alkaline
electrolytes) [86].

ammonium bromide (CTAB). For larger sizes of these surfactants, 4. Types of zinc-based hybrid flow batteries
the hydrophobic part has a tendency to absorb strongly on the
cathode surface [106]. These additives are often effective in refining 4.1. Positive redox couples involving solid phase active materials
the grain size and suppressing the dendrite and hydrogen evolution
[33,107]. However, some of these primary additives are not very Solid-phase active materials, such as lead-dioxide and nickel
soluble, other additives, are sometimes added in the electrolytes to hydroxide, do not involve soluble active species in the electrolytes
address this issue. and undergo pure solid-phase transformation from one solid to
On the other hand, anionic surfactants often contain polar and another in the charge-discharge processes. Since the charged
solubilizing groups, they are more stable and are able to operate at products, including the zinc electrodeposit, are at the electrode
higher temperature [108]. In addition to anionic surfactants, per- surface, separators and membranes are not needed to prevent
fluorosurfactants tend to have excellent stability [109e111] despite direct reaction between these products. The resulting electrolytes
the higher cost. Some other additives, known as “levellers”, are only contain soluble zinc(II) species and serve as common elec-
typically composed of smaller molecules with highly specific trolytes for both half-cell reactions. Compared to conventional
adsorption at the electrode surface [112]. static systems based on zinc oxide electrodes, flowing electrolytes
According to the early theories [113], diffusion of these ‘level- are still necessary to extend the cycle life to more than several
lers’ to the ‘peaks’ is faster than that of diffusion to the ‘valleys’ in hundred cycles, by addressing issues associated with dendrites,
the microprofile of zinc electrodeposits. This results in higher shape change and passivation at the zinc electrodes [96]. This leads
‘leveller’ concentrations at the ‘peaks’, where inhibition of elec- to flow-assisted [116] or single flow [63] architectures under
trodeposition occurs. On the other hand, metallic atoms migrate different circumstances. The capacities of these systems are still
and electrodeposit in the recessed areas. Therefore, true levelling limited by the mass of zinc electrodeposited at the negative elec-
can be achieved [114]. However, suitable concentration of these trode and the mass of the solid active material used in the positive
‘levellers’ is essential to achieve such an effect [113,115]. For electrode reactions. The operational parameters and performance
instance, when the concentration of these ‘levellers’ is too low, of this type of batteries are summarized in Table 3.
both ‘peaks’ and ‘valleys’ do not pose significant inhibition effects,
therefore levelling effect would not be significant. On the contrary, 4.1.1. Zinc-nickel
when the ‘leveller’ concentration is too high, inhibition effects take The concept of alkaline zinc-nickel rechargeable batteries can be
place in both ‘peaks and valley’. traced back to the 1900s [133]. The development of the original
Table 3
Operational parameters and performance of zinc-based hybrid flow batteries or flow-assisted batteries with positive active species in solid, liquid and gaseous phases.

Zinc-based Cell configuration Electrode materials Negative electrolyte Positive electrolyte Exp. OCV/V % System Eff. No. of cycles Year
chemistries compositions compositions

Positive redox couples involving solid phase active materials


Zinc-nickel Single-flow, Negative: cadmium plated Common electrolyte: 1 M ZnO in 10 M KOH: 1.6 Energy: 86 1000 2007 [63,117]
membraneless nickel foil Coulombic: 96 (20 mA cm2)
Positive: sintered nickel
hydroxide
Flow-assisted, Negative: zinc deposited Common electrolyte: ZnO in w.t. 45% KOH 1.5 Energy: > 80 1500 2011 [116,118]
membraneless copper plate Coulombic: >90 (>10 mA cm2)
Positive: sintered nickel oxide
Zinc-lead-dioxide Single-flow, Negative: carbon polymer Common electrolyte: 1.5 M ZnSO4 þ 0.5 M 2.4 Energy: > 65 90 2012 [65] [119]
membraneless Positive: lead-sulphate paste Na2SO4 þ 0.5 M H2SO4 Coulombic: > 80 (20 mA cm2
(static but flow using static cell)
available [119])
Zinc-polymer Single-flow, Negative: zinc plate Common electrolyte: PANI particles þ 2.0 M ZnCl2 þ 2.0 M 1.1 Coulombic: 97 (20 mA cm2) 32 2013 [120]
separator Positive: carbon plate NH4Cl
Separator: microporous
membrane

A. Khor et al. / Materials Today Energy 8 (2018) 80e108


Positive redox couples involving liquid phase active materials
Zinc-iron Divided, membrane Negative: silver plated iron ZnO at 5 M NaOH K4Fe(CN)6 at 5 M NaOH 1.8 Energy: 73 222 1979 [121]
(alkaline) Positive: porous nickel Coulombic: 98 (mA cm2)
Separator: Nafion® XR 475
Divided, membrane Negative: carbon plate 1.1 M ZnCl2 þ 0.8 M FeCl2 þ 2 M 1.6 M ZnCl2 þ 0.6 M 1.5 Energy: 70 127 2016 [122]
Positive: carbon felt NH4Cl þ 2 g L1 PEG8000 FeCl2 þ 0.2 M FeCl3þ 2 M Coulombic: >60 (25 mA cm2)
Separator: Daramic® 175 NH4Cl þ2 g L1 PEG8000
Divided, double Negative: copper mesh 0.5 M Na2 [Zn(OH)4] 1 M FeCl2 þ 1 M HCl 2.0 Energy: 74 >20 2015 [123]
membranes Positive: carbon felt Coulombic: 99 (40 mA cm2)
Separator: Nafion® 212/211;
Fumatech FAA-3
Zinc-cerium Divided, membrane Negative: carbon polymer plate 1.5 M Zn(CH3SO3)2 þ 1.0 M 0.8 M Ce(CH3SO3)3 þ 4.0 M 2.3 Energy: 46 57 2011 [124]
Positive: platinized titanium CH3SO3H CH3SO3H Coulombic: 84 (50 mA cm2)
mesh stack/carbon felt
Separator: Nafon® 117
Single-flow, Negative: planar carbon Common electrolyte: 1.5 M Zn(CH3SO3)2 þ 0.2 M 2.3 Energy: 76 10 2011 [68,125]
membraneless polymer Ce(CH3SO3)3 þ 0.5 M CH3SO3H Coulombic: 90 (20 mA cm2)
Positive: carbon felt
Zinc-iodine Divided, membrane Negative: carbon felt 5.0 M ZnI2 5.0 M ZnI2 1.2e1.4 Energy: 67 40 2015 [126]
Positive: carbon felt Coulombic: 96 (20 mA cm2)
Separator: Nafon® 115
Zinc-polymeric Divided, Membrane Negative: zinc or carbon plate 1.0 M ZnCl2, 1.0 M NH4 Cl Polymeric TEMPO in 1.0 M 1.7 Energy: 80 500 2016 [127]
TEMPO Positive: carbon felt ZnCl2, 1.0 M NH4 Cl Coulombic: > 90 (20 mA cm2)
Separator: size exclusion
dialysis membrane
Zinc-TEMPO Divided, membrane Negative: zinc foil 1 M ZnAc þ 2 M NaCl þ Na2SO4 1 M hydroxyl- 1.7 Energy: 74.0 50 2016 [128]
Positive: carbon felts TEMPO þ 2.0 M Coulombic: 88 (10 mA cm2)
Separator: Nafon® 212 NaCl þ Na2SO4
Zinc-benzoquinone Single-flow, Negative: planar carbon Common electrolyte: 1.5 Energy: 73 12 2016 [66,67]
membraneless Positive: carbon felt 1 M ZnCl2 þ 50 mM 1,2-BQDS Coulombic: 74 (30 mA cm2)
Positive redox couples involving gaseous phase active materials
Zinc-air Single flow, Negative: zinc plated copper Common electrolyte: >1.6 Energy: 72 150 2009 [129]
membraneless plate 0.7 M K2 [Zn(OH)4] þ 7 M KOH þ 0.7 M LiOH Coulombic: 97 (20 mA cm2)
Positive: oxygen composite
electrode:
Ni(OH)2 þ EMD þ NaBiO3
(continued on next page)

87
88 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

static system is associated with the relatively long cycle life of


cadmium-nickel batteries. In comparison, the zinc-nickel static
1977 [130]

2013 [131]

1979 [132]
battery has a higher cell voltage and exhibits a moderate specific
energy of 55e85 Wh kg1 [134] for relatively high power applica-
Year

tions (e.g., bicycles, scooters and electric vehicles). In conventional


static systems, the nickel (hydroxide) electrode is usually sintered
No. of cycles

or pressed, while the zinc electrode is made of a fine porous matrix


of zinc oxide in a discharged state. In most cases, polymer or cel-
lulose based separators are used to prevent short circuits caused by

Nil
70

dendritic growth.
6

The overall electrode reactions can be described as follows:


Coulombic: 95 (20 mA cm2)

Coulombic: 92 (20 mA cm2)

Coulombic: 84 (22 mA cm2)

At positive electrode:

2NiOOH þ 2H2 O þ 2e 42NiðOHÞ2 þ 2OH ; E0 ¼ 0:49V vs: SHE


% System Eff.

Energy: > 80

(10)
Energy: 82

Energy: 66

Overall reaction:

Zn þ 2H2 O þ 2NiOOH42NiðOHÞ2 þ ZnðOHÞ2 ; Ecell ¼ 1:73V


Exp. OCV/V

(11)
c.a. 2.0
>1.5

>1.7

In the presence of separators, the capacities of most static cells


using zinc oxide electrodes (rather than zinc electrodeposition)
often deteriorate over several hundred cycles, primarily due to
2 M ZnBr2 þ KCl þ NaCl

internal short circuits [96,116,118,135]. The use of a flowing elec-


trolyte is considered to be an effective measure against dendrite
Positive electrolyte

formation, shape change and passivation at the zinc electrode. In a


flowing system, a separator is not necessary when an inter-
compositions

Common electrolyte: 2.0 M ZnCl2 þ 4.0 M KCl

electrode gap of >4 mm is used [116,118,135]. This reduces the


overall cost and simplifies the cell design.
Early investigations included a 100 Ah battery developed by
Common electrolyte: 2.0 M ZnBr2

Bronoel et al. [135] with the functions of pulsed charging and


inversion of electrolyte directions. In 2007, a ‘hybrid flow battery’
concept was introduced by Cheng and co-workers [63,117] through
2 M ZnBr2 þ KCl þ NaCl

fundamental studies [63] and lab-scale testing, in which more than


Negative electrolyte

220 cycles were obtained with energy efficiencies of c.a. 88%. In


these investigations, cadmium was used as the substrate material
compositions

for zinc electrodeposition due to its relatively large hydrogen


overpotential.
Following these developments, Ito et al. [116,118] further eval-
uated the performance of a flow system with a particular focus on
the morphology evolution during charge-discharge cycling. Due to
Positive: carbon felt mixed with

the fact that the zinc electrode reaction has a higher coulombic
Negative: carbon polymer
Separator: Nafon® 125 or

efficiency than its nickel oxide counterpart, residual zinc tends to


Positive: porous graphite
Negative: carbon based

Negative: carbon plate


Separator: Nafon® 115
Positive: carbon based

accumulate on the substrate or deposit surface after each cycle.


active materials and
Electrode materials

Depleted zincate ions in the electrolytes and continuous hydrogen


complexing agent

evolution at the electrodeposit surface further facilitate dendritic


Celgard 3400

growth, which is often initiated at the regions of high local current


electrode

electrode

electrode

electrode

densities at the microprofiles. In subsequent cycles, dendritic


growth tends to take place progressively and may not be sup-
pressed by the flowing electrolyte (Fig. 6). However, dendritic
Divided, membrane

growth may not cause serious internal shorting as long as there is


Cell configuration

no strong contact with the positive electrodes. The dendrites may


membraneless

peel off from the electrodeposit under aggressive flows. To avoid


Single flow,

Single flow,
membrane

these issues, reconditioning of the battery was introduced by


conducting a deep discharge to fully remove or dissolve all the
accumulated zinc electrodeposit every 15 cycles. With this pro-
cedure, 1500 charge-discharge cycles were obtained with energy
Table 3 (continued )

efficiencies of over 80%.


In addition to planar electrodes, several researchers [136e138]
Zinc-bromine

Zinc-chlorine
chemistries
Zinc-based

have proposed the use of three-dimensional nickel foam as the


negative electrode in order to increase the current densities and
cycle lives. Cheng et al. [136] demonstrated that it is possible to
achieve reasonable energy efficiencies (>50%) at ultra-high current
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 89

Fig. 6. Electrodeposition of zinc on a planar carbon substrate under flowing electrolyte (V ¼ 2 cm s1): (a) 1st cycle; (b) 5th cycle; (c) 9th cycle [116,118].

densities of up to 300 mA cm2. Within these porous structures, the Thus, the performance is not competitive with conventional flow
morphologies range from smooth (Fig. 7a), spongy (Fig. 7bee) to battery systems that store energy partly or fully through the elec-
dendrite structures (Fig. 7f) by increasing the current density from trolytes. Rather than using filter-press type cells with specific
40 to 300 mA cm2. Although no further cycling information is manifolds to distribute electrolytes and to avoid shunt (bypass or
provided, other work from the same group [137] concluded that an leakage) currents in a stack, electrode plates were immersed in
extended cycle life was achieved with no accumulation of zinc the electrolyte, which was circulated within a sealed system
observed within the porous structures for more than 400 cycles. [116,118,139]. Battery performance could be further improved with
Based on the established laboratory-scale batteries, a scaled-up the use of additives [142,143], and by carefully selecting electrolytes
prototype was developed by Turney et al. [139]. The capacity of the [144,145], electrodes [136e138] and battery architectures [146], as
system was up to 25 kW h, consisting of thirty 833 W h batteries reported in existing studies of static and flow systems. Due to the
connected in series. The overall voltage ranged from 40 to 60 V with use of positive nickel electrodes, the cost of this system was still
energy efficiencies of more than 80% over 1000 cycles, which is more than USD$400/kW h, significantly higher than the DoE cost
approximately a year of data (three cycles per day). The energy used target (USD$ 150/kW h).
for electrolyte pumping was only c.a. 4% of the total energy
throughput [139]. 4.1.2. Zinc-lead-dioxide
Despite the tremendous improvements in recent studies, the Development of the acidic zinc-lead dioxide battery can be
discharge duration and capacities of the systems above were typi- traced back to the 1970s in the description of a few patents
cally limited to less than 4 h and 50 mA h cm2, respectively. This [147e149]. This chemistry is based on low-cost elements in
was mainly attributed to the limited capacity of the nickel positive aqueous electrolytes and exhibits a high open-circuit voltage (c.a.
electrodes (<50 mA h cm2); the negative electrode reactions have 2.45 V); in fact, one of the highest cell voltages in aqueous elec-
been demonstrated to continue for prolonged periods in the charge trolytes (20% higher than conventional lead-acid batteries). The
direction in other alkaline systems (240e500 mA h cm2) [140,141]. overall electrode reactions can be described as follows:

Fig. 7. Zinc electrodeposit morphologies within porous nickel form structures at different current densities: (a) 40 mA cm2, (b) 80 mA cm2, (c) 120 mA cm2, (d) 160 mA cm2, (e)
200 mA cm2, (f) 300 mA cm2 [136].
90 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

At positive electrode: micropores. During the charge and discharge processes, polyaniline
particles undergo doping and de-doping of chloride anions, respec-
PbO2 þ 2Hþ þH2 SO4 þ2e 4PbSO4 þ2H2 O; E0 ¼þ1:36Vvs:SHE tively, with electrochemical behaviour in terms of potentials and
(12) current densities similar to that of their film counterpart. Charge-
discharge cycling was demonstrated in a laboratory scale cell,
Overall reaction:
which achieved a coulombic efficiency of c.a. 97% at 20 mA cm2 over
30 cycles. The current density was significantly higher than those of
PbO2 þ Zn þ 2Hþ þ H2 SO4 4PbSO4 þ Zn2þ þ 2H2 O; Ecell ¼ 2:12V
previous ‘film’ systems (<5 mA cm2). The theoretical energy density
(13) of this system (66.5 W h L1) can be even higher than that of the
In contrast to most lead-dioxide based batteries, most aqueous commercial all-vanadium redox flow batteries [120].
systems are limited to less than 1.5 V due to gas evolution via water
electrolysis. In common with lead-acid batteries (c.a. 2.05 V), the 4.2. Positive redox couples involving liquid phase active materials
relatively high cell voltages are attributed to the highly positive
electrode potential of the lead-dioxide reaction (þ1.69 V vs. SHE). Liquid-phase positive active materials usually dissolve in elec-
Considering that lead (II) ions are almost insoluble in sulfuric acid trolytes and appear as soluble species in both oxidized and reduced
(<6.5 mg L1), any electrodeposition associated with lead (II) ions is forms. Similar to conventional redox flow batteries, the storage ca-
negligible [65]. Therefore, zinc or other metals can be electro- pacities of systems based on these electrode reactions can be
deposited from soluble species in common electrolytes, providing the increased by increasing the volume and the concentration of the
negative electrode reactions for several proposed static and flow positive electrolytes, but are still limited by the mass of the zinc
batteries (zinc-lead-dioxide [65,119,150]; copper-lead-dioxide [151]; electrodeposits at the negative electrode. Unless a separator or ion-
cadmium-lead-dioxide [152]). Since the charged products are on the exchange membrane is used, the charged species could react with
electrode surface, a membrane or separator is not necessarily required the metallic zinc electrodeposit, which leads to energy loss through a
to prevent direct chemical reactions involving these products. self-discharge process. In a few cases, dissolution of the zinc elec-
The zinc-lead dioxide chemistry has not received much atten- trodeposit is particularly slow in the presence of certain active species
tion due to zinc corrosion and the limited current densities at low concentrations, providing the possibility of a membrane-less
(<10 mA cm2) reported in early systems [150]. In a recently or single flow configuration. The liquid phase of positive active ma-
developed static system, experimental results showed that the terials allows the use of porous, three-dimensional electrode mate-
average coulombic and energy efficiencies were up to 90% and 70%, rials, which effectively facilitates the positive electrode reactions in
respectively, over 90 cycles at 20 mA cm2 [65]. Since a thin carbon terms of reversibility and overpotentials. Although many proposed
polymer electrode (0.6 mm thickness) was used rather than heavy species are based on low-cost elements, the solubilities (e.g., cerium:
metallic electrode materials, the specific energy was around 30% c.a. 0.8 M, benzoquinone: <0.8 M, ferrocyanide <1 M) are still lower
higher than that of a commercial lead-acid battery under the same than those of conventional vanadium electrolytes (c.a. 1.5e2 M), with
operating conditions [150]. Subsequently, Pan et al. [119] developed the exception of zinc iodide (>5 M). This limits the energy density for
a membrane-less zinc-lead dioxide flow battery based on single energy storage applications. The operational parameters and per-
flow configuration. It was found that hydrogen evolution can be formance of this type of battery are summarized in Table 3.
effectively suppressed with polymer resin doping within the
graphite composite. However, during the charging process, protons
are continuously generated as described in equation (12), which 4.2.1. Zinc-iron
implies challenges to zinc electrodeposition, as well as corrosion, at An early zinc-iron hybrid flow battery based on alkaline elec-
the negative electrode. The proposed system also suffered from trolytes, was introduced by scientists at Lockheed Missiles and
limited capacities of the lead-dioxide electrodes (<50 mA h cm2 Space Company (LMSC); and is termed the ‘zinc-ferricyanide’
[65]). To compete with other flow battery chemistries for practical battery [140,157]. This system is currently commercialized by ViZn
applications, further developments in terms of the capacity of the Inc. for grid-scale applications [158]. The active materials (zinc and
positive electrode and corrosion inhibition are necessary. iron) are among the most abundant and lowest cost metals
(<USD$ 4 kg1) in the earth's crust. Ferri- and ferrocyanide are
iron complexes known to be highly reversible in alkaline electro-
4.1.3. Zinc-polymer (suspensions) lytes. Inorganic ferricyanide is less hazardous than halogens in
The zinc-polymer rechargeable battery was introduced by Sur- alkaline but releases toxic hydrogen cyanide gas in acidic solu-
ville et al. [153] in 1968. In conventional systems, these polymers tions. The early system was based on solid-state active materials in
appear in solid form and include polyacetylene, polyaniline, poly- both charged and discharged states. For the zinc negative elec-
thiophenes, poly(p-phenylene), polyazulene and polyfluorene trode reaction, electrodeposited zinc and solid zinc oxides were
[154e156]. Among these polymers, polyaniline is commonly used the charged and discharged products, respectively. A zincate so-
in zinc-based systems due to its high conductivity and stability in lution is generated from the solid state zinc oxide during the initial
aqueous electrolytes [154e156]. However, this type of battery suf- discharge cycle. The active positive materials were ferro- and
fers from a limited specific energy since only the outer layer of the ferricyanide as forms of precipitates obtained through crystallizers
polymer film contributes to the energy storage capability. or when the electrolytes reach saturation. Although solid-state
To address this issue, Zhao et al. [120] proposed the use of poly- active materials were used, the solubility of potassium ferrocya-
aniline particles as suspensions in electrolytes, rather than a polymer nide in 1 M potassium hydroxide is c.a. 0.5 M at 25  C [140,141].
film, in a flow cell configuration (See Fig. 1). These suspensions were In alkaline electrolytes, the half-cell electrode reactions
prepared by dispersing the synthesized polyaniline particles in a (discharge) are as follows:
solution containing zinc and ammonia chloride. The resulting sus-
pension electrolyte was circulated through the flow-cell, which was At negative electrode:
essentially a ‘single-flow’ system. This is because the liquid electro-
lyte containing soluble zinc species could penetrate easily through Zn þ 4 OH 4ZnðOHÞ2  0
4 þ 2 e ; E ¼ 1:22V vs: SHE (14)
the separators to the negative half-cell compartment through the
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 91

At positive electrode: Under conditions that ensured the galvanic displacement of zinc by
iron was low, zinc electrodeposition was used rather than iron

FeðCNÞ3 4 0
6 þ e 4FeðCNÞ6 ; E ¼ þ0:36V vs: SHE (15) electrodeposition as the primary reaction at the negative electrode.
On the other hand, the zinc ions have no significant effect on the
positive iron (II/III) redox reaction at the positive electrode. How-
Overall reaction: ever, the incorporation of iron was only observed in the case of low
zinc concentration (or high iron concentration), or at high current
Znþ2FeðCNÞ3  2 4
6 þ4OH 4ZnðOHÞ4 þ2FeðCNÞ6 ; Ecell ¼1:58V
densities. With the use of a Daramic® microporous separator, the
(16) energy efficiencies of this system were more than 60% over 120
cycles at 25 mA cm2, which demonstrates that the system is
In an experimental system, the open-circuit voltage (c.a. crossover-tolerant. Based on this performance and configuration,
1.5e1.8 V) was even higher than the predicted values based on the estimated cost of a 5.5 h system would be around USD$100/
standard electrode potentials (1.58 V). The negative electrode was kWh [122].
typically a cadmium or silver coated substrate, while a nickel plated In addition to pure alkaline or acidic systems, Gong et al. [123]
graphite felt was used as the positive electrode material. A sepa- has proposed an alkaline (zinc) e acidic (iron) system to yield a
rator or membrane was required to avoid crossover of ferricyanide cell voltage as high as 1.99 V. Since the two supporting electrolytes
into the zincate containing negative electrolytes, which would have very different pH values, it is not feasible to use a conventional
result in undesirable precipitates. For instance, zinc ferrocyanide single-membrane configuration. Instead, a double-membrane, tri-
could electrodeposit as an insoluble film at low hydroxide con- ple-electrolyte configuration was used as illustrated in Fig. 8, in
centrations (i.e. < 1 M). Previous experiments show that zinc which a neutral sodium chloride supporting electrolyte was
electrodeposition is the limiting reaction with a corresponding employed in the middle compartment between the two ion-
half-cell coulombic efficiency of c.a. 85%, compared to c.a. 97% ob- exchange membranes. Nafion® 212/211 (Dupont) and FAA-3
tained at the positive half-cell. This system was reported to have a (Fumatech) were used as the cation and anion exchange mem-
relatively long cycle life of up to 250 cycles [140,141]. However, high branes, respectively. The overall cell resistance was observed to be
membrane costs and challenges due to handling solid zinc oxide smaller at higher electrolyte flow rates or with a smaller electrolyte
precipitates are considered the major hurdles for this system, and compartment thickness. It was reported that energy efficiencies
the capacity of the zinc negative electrode is still limited to less were more than 50% at 80 mA cm2 over 20 cycles. Together with a
than 500 mA h cm2 [140,141]. relatively high cell voltage, a peak power density was achieved at
In addition to alkaline systems, a few acidic systems have 660 mA cm2 (70% state of charge). Since protons and hydroxide
recently been introduced [122,159]. In acidic electrolytes, the half- ions can migrate through the ion-exchange membranes, the pH
cell electrode reactions (discharge) are as follows: values of the electrolytes will change during long-term operation,
and will therefore require periodic adjustment. For this particular
At negative electrode: system, further work should focus on more selective ion-
conductive membranes while ensuring that their costs are signifi-
Zn4Zn2þ þ 2e ; E0 ¼ 0:76V vs: SHE (17) cantly lower than perfluorinated counterparts (e.g. Nafion®) [123].

At positive electrode: 4.2.2. Zinc-cerium


The zinc-cerium hybrid flow battery was first patented by Clarke
Fe3þ þ e 4Fe2þ ; E0 ¼ þ0:77V vs: SHE (18) et al. [160,161] in 2004 and has been developed by Plurion Inc. [71].
Overall reaction: A 2 kW system has been installed in their testing facility in Scot-
land. The proposed system uses a carbon polymer substrate and
Zn þ Fe3þ 4Zn2þ þ Fe2þ ; Ecell ¼ 1:53V (19)
For instance, Xie et al. [159] introduced a mixed acid system
using 1 M zinc sulphate with 1.5 M sodium acetate and 1.5 M acetic
acid as the negative electrolyte, while the positive electrolyte
contained 1 M iron (II) chloride in 1.5 M sulphuric acid. The
resulting cell voltage (1.53 V) was comparable to that of the
aforementioned alkaline systems (c.a. 1.5e1.8 V). It was found that
the role of an acetate buffer solution is crucial to battery perfor-
mance since it constrains the concentration of free protons to
within a preferred range, in particular enabling more proton
transfers from the positive to the negative half-cell. Furthermore, a
higher concentration of acid is favourable for the kinetics of the iron
(II)/iron (III) reactions. With the addition of such a buffer solution,
the coulombic efficiency increased from less than 20% to c.a. 91% at
30 mA cm2. It is also important to note that carbon felt was used as
the negative electrode, upon which it tends to be difficult to deposit
zinc without the buffer solution.
Selverston et al. [122] proposed another acidic system
employing mixed solutions of zinc and iron species for both the
negative and positive electrolytes. The pH values of the electrolytes Fig. 8. Three-compartments zinc-iron hybrid flow batteries, using Zn(OH)2 4 /Zn as
were controlled at pH 1 to avoid the formation of insoluble iron negative redox pair in sodium hydroxide solution, NaCl as middle electrolyte and Fe3þ/

hydroxide precipitates, while minimizing hydrogen evolution. Fe as positive redox pair in hydrochloric acid solution [123].
92 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

platinized carbon/titanium at the negative and positive electrodes,


respectively [160,161]. Nafion® cation exchange membranes were
used to separate the two half-cell compartments. The system made
use of methanesulfonic acid as the supporting electrolyte to yield
reasonable solubilities for both the zinc (>2 M) and cerium species
(c.a. 0.8 M) in both oxidized and reduced forms. In contrast, the
solubilities of cerium species are still too low (<0.3 M) in sulphate
electrolytes to enable energy storage applications [162].

At the positive electrode:

Ce4þ þ e 4Ce3þ ; E0 ¼ þ1:28  1:72V vs: SHE (20)

The overall reaction is:

Zn þ 2Ce4þ 4Zn2þ þ 2Ce3þ ; Ecell ¼ up to 2:48V (21)


As stated by Plurion Inc. [160,161], their batteries operated at
60  C and under constant voltage discharge [71]. The discharge
current densities were up to 100 mA cm2 with coulombic effi-
ciencies of more than 70% over 60 cycles. The electrochemical and
charge-discharge performance has been evaluated by a number of
researchers [62,68,88,91,124,125,163e175]. In the work of Leung
et al. [88,91,124], a number of two- and three-dimensional elec-
trode materials were evaluated for the cerium redox reactions. In
terms of chemical stability and overpotentials, a platinized titanium
mesh stack electrode was the most suitable for long-term perfor-
mance. In contrast, carbon-based materials tended to be oxidized in
electrolytes containing cerium species [124].
A high methanesulfonic acid concentration (4 M) is required to
maintain reasonable solubilities for both the cerium (III) and
cerium (IV) species (Fig. 9). To facilitate an efficient zinc electrode
reaction, a lower acid concentration (1 M) was used in the negative
electrolytes. In typical operation, the coulombic and voltage effi-
ciencies were up to 92% and 68%, respectively, at 50 mA cm2 and
50  C [124]. However, a high concentration of acid has a tendency to
transfer towards the negative side through the cation-exchange
membranes, which makes zinc electrodeposition more difficult in
long-term operation. Regarding the zinc half-cell study, no signif-
icant dendrite formation was observed after 4 h under the same
operating conditions. It was reported that the use of indium addi- Fig. 9. The solubility of (a) Ce(III) and (b) Ce(IV) ions in aqueous methanesulfonic and
tives could suppress hydrogen evolution and lead to a lower sulphuric acid electrolytes [125].
deposition overpotential, which in turn leads to a faster electro-
deposition rate and boulder-like agglomerate morphologies [91].
Some common organic and inorganic additives, including cetyl respectively [125]. In the work of Xie et al. [163], up to 1 M of
trimethylammonium bromide and lead oxide, were shown to cerium species was dissolved in a mixture of 2 M methanesulfonic
inhibit corrosion effectively over a long period of time, but their acid and 0.5 M sulfuric acid.
strong blocking effect leads to relatively low half-cell coulombic The use of a mixed acid also tends to improve the kinetics of the
efficiencies [62,91]. cerium species' reactions at < 40  C [163,164]. In addition to
A membrane-less system has also been introduced by the same platinum based electrodes, a number of precious metal-based
authors, making use of the relatively slow dissolution of zinc at low electrodes were evaluated by Nikiforidis et al. [165], who
acid concentrations and the efficient cerium redox reaction using showed that platinum-iridium based electrodes exhibited the best
carbon felt electrodes [68,125]. The common electrolyte contained electrochemical performance. At the same time, lower cost ma-
1.5 M zinc(II) and 0.2 M Ce(III) in 0.5 M methanesulfonic acid. In the terials, including hierarchical porous carbon [166], graphene oxide
given configuration, the discharge cell voltage and energy effi- [167], tin oxide [168] and indium modified carbon felts [169], have
ciencies were still up to 2.1 V and 75%, respectively, at 20 mA cm2 been observed to improve reaction kinetics compared to conven-
and room temperature. The self-discharge process was slow and tional carbon-based materials. The early development of zinc-
took up to 7 h after charging for 30 min [68,125]. To minimize the cerium flow battery has been reviewed by Walsh et al. [176].
self-discharge and allow for the use of higher cerium concentra- Future work on this system should focus on low-cost, chemically
tions at low acid concentrations, mixed acid electrolytes have been stable electrodes and electrolytes to dissolve more cerium species
proposed by a few researchers [125,163,164]. As shown in the sol- at low acid concentrations. Electrolyte additives for improving
ubility charts (Fig. 9a and b), concentrations higher than 0.5 M of morphologies and inhibiting corrosion of the zinc electrodeposits
cerium species in oxidizied and reduced forms can be made soluble are important for operation at elevated temperatures and/or at
at low concentrations of methanesulfonic and sulfuric acid, high acid concentrations.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 93

4.2.3. Zinc-iodine c.a. 91% to c.a. 76% as the concentration of zinc iodide increased
The zinc-iodine battery was first developed as a primary system from 0.5 to 3.5 M, which was attributed to the increasing electrolyte
by Martin for academic demonstration in 1949 [177], employing resistances. A maximum positive electrolyte energy density of
metallic zinc and potassium iodide solution as the active electrode 167 W h L1 was achieved (compared to 20e35 W h L1 for con-
materials. Around a decade later, Jones and Arranaga [178] pro- ventional vanadium electrolytes). It was also found that the addi-
posed another primary system using solid iodate rather than liquid tion of ethanol could induce ligand formation between oxygen on
iodide solution at the positive electrode. Electrical energy was the hydroxyl groups and zinc ions, further expanding the temper-
released by mechanically forcing sulphuric acid solutions into the ature window to 20 to 50  C [126].
battery using compressed carbon dioxide gas. Despite a relatively Following these developments, Weng et al. [14] proposed the
large current output (up to 100 mA cm2), this system is not use of bromide ions (Br) as a complexing agent to stabilize the free
rechargeable and is based on a highly corrosive electrolyte (c.a. 4 M iodine by forming iodine-bromide ions (I2Br) as a means to
sulphuric acid). liberate iodate ions, hence enabling a higher discharge capacity of
Although iodine (I2) has a low solubility, its solubility increases the positive electrolyte. Compared to chloride ions, bromide ions
significantly in the presence of iodide (I) due to the formation of tend to be more stable and exhibit fewer hydrolysis issues. As
polyiodide ions (I 3 ). Since iodide (or triiodide) has a high solubility shown in Fig. 10, both centrosymmetric I 3 and asymmetric I2Br


in aqueous electrolytes (e.g., lithium iodide: c.a. 8.2 M; zinc iodide: tend to have a linear (or nearly linear) trihalide structure, which is
c.a. 5.6 M), it has been considered as an attractive redox couple for thermodynamically stable. With the use of bromide ions, the ca-
flow battery applications [126], although the imported cost of pacity of the positive electrolyte was up to 202 Wh L1, which is
iodine (to the United States) is often as expensive as vanadium among the highest energy density achieved for aqueous flow bat-
metal (USD$ 23 kg1). Following the introduction of the lithium- teries to date. To improve the reaction kinetics and avoid the use of
iodide system (2013), a zinc-iodide flow battery was developed expensive catalysts, porous coordination polymers were proposed
by Li et al. [126] in 2015. The overall electrode reactions are as by Li et al. [179] as a catalyst for the positive iodide reaction. [6].
follows: Nevertheless, poor chemical stability of this material, especially
under electrolyte soaking, is a major challenge for future applica-
At the positive electrode: tions. Despite the high solubility of the iodide redox species, the
overall capacity of this battery is still limited by the zinc negative
electrode. Since the same electrolyte is used for both half-cells,
I   0
3 þ 2e 43I ; E ¼ 0:54V vs: SHE (22)
further cost reductions can be possible with the use of low-cost
separators.
Overall reaction:
4.2.4. Zinc-polymer (dissolved)
In addition to the polymer suspensions described in Section
I  2þ
3 þ Zn43I þ Zn ; Ecell ¼ 1:30V (23)
4.1.3., the use of organic polymers has become popular in the field
Both the negative and positive electrolytes were based on zinc of aqueous organic redox flow batteries. As proposed by Janoschka
iodide salt (ZnI2) in water. A Nafion® membrane was sandwiched et al. in 2012 [180], these organic polymers typically dissolve in
between the two carbon felt electrodes. It is important to note that electrolytes and consist of two parts, namely a redox active moiety
no acid or alkaline was added to the electrolytes, resulting in a and a unit providing sufficient aqueous solubility to prevent pre-
nearly neutral electrolyte (pH 3e4) at 0% state-of-charge. The flow cipitation [181e183]. The same research group [127] made use of
cell was charge-discharge cycled within a range of current densities this type of polymer and introduced an organic-inorganic system
between 5 and 20 mA cm2. The energy efficiencies decreased from by coupling with a zinc anode. This early proposed system used

Fig. 10. (a) Illustration of the concept of bromide as the complexing agent to stabilize iodine; (b) Structure of I2Br ion; (c) Structure of I
3 ion. *The bonding length is obtained from
the first-principles density functional theory calculations [14].
94 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

aqueous electrolytes based on chloride salts. Zinc chloride not only In 2016, Orita et al. [128] proposed a zinc-organic hybrid flow
served as the active species but also as the supporting electrolyte. battery based on soluble TEMPO derivatives by functionalization
Poly(ethylene glycol) methyl ether methacrylate (PEGMA) and [2- with a hydroxyl group. The charge-discharge reactions of the pro-
(methacryloyloxy)ethyl]trimethylammonium chloride (METAC) posed system are as follows:
were used as the copolymers of the active material, which was
2,2,6,6-tetramethylpiperidinyl-N-oxyl (TEMPO) [127]. At positive electrode:
In flow cell studies, a microporous separator was used to pre-
vent the crossover of the polymer active materials, while porous TEMPOþ þ e 4 TEMPO; E0 ¼ þ0:83V vs: SHE (24)
carbon papers or felts were used as electrode materials for both
half-cell reactions. With these organic polymers, the open-circuit
cell voltages were higher than 1.2 V. Compared to METAC-based Overall reaction:
polymeric TEMPO, PEGMA-based polymers have higher solubil-
ities and do not precipitate in concentrated zinc chloride electro- Zn þ 2 TEMPOþ 4Zn2þ þ 2 TEMPO; Ecell ¼ 1:59V (25)
lytes (1 M). In such electrolytes, capacities of up to c.a. 2.4 were The resulting hydroxyl-TEMPO has the ability to dissolve at
achieved, and decreased linearly with current density. In static concentrations as high as 3.6 M. The system used metallic zinc and
cells, the coulombic and energy efficiencies were over 80% and 50%, carbon felts as the negative and positive electrode materials,
respectively, at 3 mA cm2 [127]. respectively. Both cation (Nafion® 212) and anion exchange mem-
In these aqueous systems, the current densities were still branes (Selemion DSV) were tested in two different supporting
limited to less 20 mA cm2. The same research group proposed electrolytes (1 M sodium perchlorate or sodium chloride). In both
another similar system based on polymeric-TEMPO in non-aqueous cases, the open-circuit cell voltages were more than 1.6 V upon
electrolytes [184]. Rather than using regular, linear polymers, this charging. The energy efficiencies were highest (80.4%) at
work used well-defined TEMPO-methacrylate/styrene block co- 10 mA cm2 when a Nafion® 212 membrane and sodium perchlo-
polymers (PTMA-b-PS) featuring micellar structures. Mixtures of rate electrolytes were used. The capacity of the resulting battery
carbonate electrolytes were used for both the negative and positive tended to decrease after cycling. This is attributable to a decrease in
electrolytes. The concentrations of the organic polymer were up to the hydroxyl-TEMPO concentration caused by its side reaction,
13 mg cm3 in the positive electrolytes. Similar to the aforemen- which converts the active material into other organic molecules.
tioned aqueous systems, the charge-discharge capacities decreased Charge-discharge cycling was also performed at high hydroxyl-
linearly with current density, but within a lower current density TEMPO concentration (1 M). The initial energy efficiencies were
range (<5 mA cm2). The coulombic efficiencies were more than lower (67.5%) at 10 mA cm2. The discharge capacity and energy
90% over a range of current densities (0.5e5 mA cm2). With densities were 8 A h L1 and 9.6 W h L1, respectively. However, the
suitable cell architectures and selections of components (e.g., sep- capacity retention dropped rapidly to near 0% within 5 cycles,
arators), improved current densities of more than 35 mA cm2 have possibly due to chemical degradation of the active materials in the
been demonstrated in other similar work using aqueous and non- presence of the generated hydroxyl ions. For these reasons, stable
aqueous electrolytes [7,8]. control of electrolytes and long-term stability of the active mate-
rials are necessary for improved battery performance [128].
4.2.5. Zinc-organic (non-polymer) In the same year, a membrane-less zinc organic hybrid flow bat-
As discussed in Section 4.2.4, organic polymers have been used tery was introduced by Leung et al. [66,67]. The operating concept
in a number of zinc-based batteries [120,155,156]. This section fo- made use of the slow dissolution of the zinc electrodeposit in the
cuses on the particular use of soluble organic materials for zinc- presence of soluble benzoquinone species at low concentrations (c.a.
based flow batteries. The first few metal-organic hybrid flow bat- 0.1 M) in the common electrolytes. Rather than prototypical 1,4-
teries were introduced by Xu et al. [11,185] in 2009. The initial hydroquinone, 1,2-hydroquinone-disulfonic acid was used due to
systems used cadmium and lead at the negative electrode, both of its higher redox potential (þ0.85 V vs. SHE), resulting in a cell voltage
which have less negative potentials (0.42 V vs. SHE) than zinc of more than 1.4 V. In contrast to the aforementioned TEMPO derived
(0.76 V vs. SHE) in aqueous electrolytes, implying that zinc-based species, both the negative and positive electrode reactions undergo
systems are even more attractive in terms of cell voltages and cost. two-electron-transfers processes. Despite relatively low current
Derivatives of quinone and 2,2,6,6,-tetramethylpiperidinyloxyl densities (30 mA cm2), the capital cost of this system is still lower
(TEMPO) based materials, are the most studied positive redox than the DoE cost target, while the electrolyte cost is USD$ 14/kW h or
couples in the recent development of organic-based flow batteries lower. In the absence of a separator, the charged species are free to
[7,8]. Many of these molecules exhibit relatively high positive redox react with the zinc electrodeposit as a self-discharge process. The
potentials (>0.5 V vs. SHE) and high solubilities in aqueous elec- corrosion current densities were estimated to be less than
trolytes (>0.5 M). They are, furthermore, commercially available at 10 mA cm2, which is still lower than the current densities used
low-cost. Unlike many other organic active materials, the chemical (30 mA cm2) [66,67]. The reactions are as follows.
structures of these molecules lead to long-term stability, making
them suitable for energy storage applications. At positive electrode:

(26)
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 95

Overall reaction:

(27)

With the use of single flow configuration, a carbon substrate and These catalysts are based on low-cost and sometimes non-metal
felt were used as the negative and positive electrodes, respectively. materials [190,191]. In all cases, alkaline electrolytes are used. The
A concentrated zinc species (1.5 M) was used in the common electrode reactions are as follows:
electrolyte to facilitate the zinc electrodeposition process and to
avoid mass transport limitations. The battery was charge-discharge At negative electrode:
cycled with an average energy efficiency of ca. 73% at 30 mA cm2
for more than 12 cycles. The low concentration of the organic Zn þ 4 OH 4ZnðOHÞ2  0
4 þ 2 e ; E ¼ 1:22V vs: SHE (28)
species (50 mM) implies a relatively low specific energy compared
to conventional systems, which offsets the cost advantage. To 
ZnðOHÞ2
4 4ZnO þ H2 O þ 2 OH (29)
further minimize the capital cost per kW h, future work should
focus on the use of higher current densities, facilitated by improved
mass transport and cell architecture. Higher specific energies can At positive electrode:
be achieved at higher organic active species concentrations with
the use of a separator (or membrane), which will add further cost to O2 þ 4 e þ 2 H2 O44 OH ; E0 ¼ þ0:4V vs: SHE (30)
the system but may be necessary for certain applications if the cost
can be kept close to the DoE target [66,67].
Overall reaction:
4.3. Positive redox couples involving gaseous phase active materials
2Zn þ O2 42 ZnO; Ecell ¼ 1:62V (31)
In some zinc-based hybrid flow batteries, gaseous active mate- In practice, the working cell voltage of primary systems is typically
rials, particularly oxygen and halogens (e.g., bromine and chlorine), 1.0e1.4 V (significantly lower than the standard potential (1.62 V))
are often generated in the positive electrode reactions during the under considerable current densities (up to 100 mA cm2) [191].
charging processes. Compared to halogens, oxygen as a charged These batteries has been used for medical and telecommunications
species does not pose a safety risk and is continuously supplied from applications [192]. For rechargeable systems, a high charge voltage of
atmospheric air. For these reasons, there is no specific requirement 2 V or higher has been observed, attributable to the high over-
for gas storage, resulting in a simplified cell design. Despite relatively potentials of the oxygen electrode reactions, restricting the current
large overpotentials, bifunctional catalysts enable both reduction and densities to below 30 mA cm2 [191]. These systems have advantages
oxidation reactions to take place at the same electrode. Since oxygen in terms of a flat discharge voltage, safety, low-cost and low self-
(as the charged species) is of negligible concentration in the elec- discharge rate. However, extended cycle life and long discharge
trolytes, a separator or ion-exchange membrane is not required, duration have been hindered by non-uniform zinc electrodeposition
resulting in a membrane-less or single flow system. and (especially) the lack of durable bifunctional catalysts.
In contrast, bromine and chlorine are toxic, and are often To improve the deposition morphologies and avoid dendritic
sequestered or condensed into liquid or solid phases (e.g., forming growth, flowing electrolytes have been used in zinc-air recharge-
polybromide in an oily phase or forming solid chlorine hydrate with able batteries. The earliest system was introduced by the Lawrence
water at 10  C) and stored carefully in special chambers. Upon Berkeley Laboratory [193,194] in 1988. In contrast to conventional
discharge, these halogen species are often mixed with the zinc zinc-air systems using granular powders or pellets (as a gelled
containing solution in the porous electrode of the positive half-cell. mixture), this system stores energy via the electrodeposition of
In typical cases, these halogen species are corrosive to metallic zinc. Porous reticulated carbon and a bifunctional air catalyst were
electrodeposits. Therefore, a separator or ion-exchange membrane used at the negative and positive electrodes, respectively. A
is often used to prevent the permeation of these species towards membrane or separator was not necessarily required. In the zinc
the negative half-cell, ensuring high coulombic efficiencies. The negative half-cell, no apparent shape change or dendritic growth
operational parameters and performance of these batteries are was observed over 600 cycles, with a discharge capacity of c.a.
summarized in Table 3. 200 mA h cm2. The specific energy was estimated to be 110 Wh
kg1, which is comparable to other zinc-air rechargeable systems at
4.3.1. Zinc-air large scale [195]. Considering that most of the materials are inex-
The zinc-air battery was originally developed as a primary pensive, the cost of this system was estimated at the time (1980's)
battery in the 1930s [186] based on the concept of Leclanche  wet to be as low as USD$ 20/kW h [140], which in today's terms is still
cells (zinc-manganese dioxide, 1866) [18]. Since the 1970s, me- likely to be lower than the cost of other zinc-based systems.
chanically rechargeable systems have been introduced, i.e., without A similar system was introduced by Pan et al. [129], as a single
an electrochemical charging process [187]. The spent zinc metal is flow system. A composite electrode was used as the positive elec-
physically replaced with a fresh anode, or as a form of zinc slurry via trode, with the side facing the aqueous electrolyte containing nano-
pumping [188,189]. Rechargeable systems by means of an electro- structured Ni(OH)2 for the evolution reaction, while the other side
chemical approach have become available with the development of facing atmospheric air contained MnO2 doped with NaBiO3 catalyst
bifunctional catalysts for the oxygen evolution/reduction reactions. for the reduction process. At 20 mA cm2 and 60  C, the charge and
96 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

discharge voltages were 1.78 V and 1.32 V, respectively. By circu- 4.3.2. Zinc-bromine
lating the electrolytes, dendritic growth and shape change of the The zinc-bromine battery was first patented in 1885 [198] and
zinc negative electrode were effectively prevented. It was also was developed as a hybrid flow battery by Exxon, Gould and the
shown that energy efficiencies of more than 70% could be achieved National Aeronautics and Space Administration (NASA) in the 1970s
over 150 cycles. [199]. This kind of battery has a high theoretical energy density
Considering that significant energy is consumed during oxygen (440 W h kg1) and a high cell voltage (c.a. 1.8 V). Over the past few
evolution, Wen et al. [196] proposed a dual functional system to use decades, zinc-bromine hybrid flow batteries have been commer-
this energy to generate organic acid through another positive half- cialized and installed by several companies (See Section 6). Scale-
cell, in which raw materials were continuously supplied through up of this system to multi-kW scales has been facilitated by a
circulation as shown in Fig. 11. Organic acids, such as propanoic modular design (Fig. 12).
acid, glyoxylic acid and cysteic acid, are oxidized from raw mate- The specific energies of these commercial systems are still
rials of propanol, glyoxal and xysteine, respectively. A common limited to 60e85 Wh kg1 [200], which is just 20% of the theo-
negative electrode was shared with the two positive half-cells. retical value. In flow cell configurations, zinc bromine electrolytes
During the charging process, propanol is oxidized to propanoic are circulated in both half-cells through external pumps. A qua-
acid and zinc is electrodeposited at the negative electrode. During ternary bromide salt (QBr), such as N-methyl N-ethyl pyrrolidinium
discharge, oxygen is reduced at the other positive electrode and bromide (MEP) or N-methyl N-ethyl morpholinium bromide
metallic zinc returns to the electrolyte to release energy. (MEM), is added to the positive electrolyte as a sequestering/
The energy efficiency of the battery was up to c.a. 59%, which complexing agent [131,201] to sequester the evolved bromine into
means that a percentage of energy consumed by the organic an alternate phase with low vapour pressure during the charging
electro-synthesis can be recovered. Other work from the same process. The sequestered bromine is then stored in an oily phase
authors [197] reported that spongy morphologies can be sup- that remains separated from the main aqueous phase due to the
pressed effectively with lead and tungstate ions, although co- higher specific gravity of the complexed phase. The concentration
electrodeposition of lead and zinc ions may occur. Similar to most ratio of QBr to ZnBr2 is typically 1:3 [202,203], which means 1 M of
alkaline zinc-based batteries, the challenges of dendritic growth, sequestering agent is used with 3 M zinc bromine solution. The
zinc oxide precipitation and shape change of the electrodes need to electrode reactions can be described as follows:
be addressed further. The overpotential of the positive electrode
reaction needs to be minimized with the development of low-cost At the positive electrode:
and durable catalysts. Improved cell designs or architectures are
important for ensuring efficient oxygen transfer to cells within a Br2 þ 2e 4 2Br ; E0 ¼ þ1:087V vs: SHE (32)
stack system.

Overall reaction:

Br2 þ Zn42Br þ Zn2þ ; Ecell ¼ 1:849V (33)


The actual chemical species are much more complicated since
complexing agents are added to form polybromide ions (Br n ; n ¼ 3,
5, 7), and the reactions tend to be more sluggish due to additional
sequestration/dissociation steps during the charge-discharge pro-
cesses. The overpotentials of both electrode reactions are relatively
low (<200 mV at 100 mA cm2) and usually larger for the zinc
negative electrode reactions due to the use of planar electrodes
[130]. As reported by Lim et al. [130], formations of dendrites are
often observed when the current densities are as low as c.a.
15 mA cm2, although prolonged charge (e.g., 10 h) has been
demonstrated.
For commercial systems, the operating pH range is relatively
Fig. 11. Bifunctional zinc-air hybrid flow batteries by using propanol oxidation as a
narrow (between 1 and 3.5) and needs to be controlled throughout
counter electrode reaction [196]. operation. For instance, undesirable mossy morphologies and
bromate formations tend to form above pH 3. When the pH is

Fig. 12. Scale-up of zinc-bromine from 10 kW h cell stack to 600 kW h systems (demonstrated by Redflow Ltd.).
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 97

to operate at current densities as high as 80 mA cm2, with energy


efficiencies of c.a. 80% [217].
Despite the low-cost active materials (zinc and bromine), the
overall cost of the zinc-bromine system is not necessarily lower than
that of commercial all-vanadium redox flow batteries [200], which is
attributed to the use of expensive sequestering/complexing agents
and the relatively low current densities of the systems (<50 mA cm2,
implying more cells are required for an equivalent power output). To
increase the energy density, a single flow cell system was developed,
in which carbon felt electrodes were soaked with active materials
and sealed within the compartment to avoid bromine emissions, as
shown in Fig. 1 [131]. Other recent systems also exhibit reasonable
performance (energy efficiency of c.a. 74% at 40 mA cm2) in the
absence of such sequestering agents [204]. Improved cell architec-
ture and electrolyte circulations could further improve the mor-
phologies of zinc electrodeposition and avoid poor mixing of the
polybromine and aqueous phases [220,221]. The specific energies of
commercial systems (60e85 Wh kg1) still require significant
improvement. Future research directions should focus on approaches
Fig. 13. Influence of bromine sequestration agents (MEP, MEM) on zinc-bromine to improve the energy density and cell performance at lower costs by
electrolytes (2 M Zn2þ, pH 3, at 20 mV s1 and 25  C. means of cell design and materials development.

higher than 6, solid zinc oxide is formed rather than soluble Zn2þ
species (ZnBr 3 ) as suggested by the relevant Pourbaix diagram 4.3.3. Zinc-chlorine
[200]. At lower pH, corrosion and hydrogen evolution become more The zinc-chlorine battery was described by Charles Renard in
dominant and lead to lower coulombic efficiencies [130]. When the 1884 [222]. Further development of this technology was under-
pH is lower than 1, excessive hydrogen evolution was observed and taken by the Energy Development Association Inc. (EDA) in the
resulted in higher overpotentials due to the trapping of gas bubbles 1970s [223]. The system was developed as a form of flow battery
on the surfaces of the electrodes [130]. and exhibited a relatively high cell voltage of 2.12 V. The overall
The influences of supporting electrolytes on electrode reactions electrode reactions are described as follows:
[204,205], solubilities and ionic conductivities [130,206] have been
evaluated for a broad range of salts (Naþ, NHþ   2
4 , Cl , Br , SO4 , At the positive electrode:
H2PO 4 and NO 
3 ). Chloride based electrolytes (up to 0.5 M) tend to
perform poorer than their sulphate, bromide or phosphate coun- Cl2 þ 2e 4 2Cl ; E0 ¼ þ1:36V vs: SHE (34)
terparts, while nitrate electrolytes are not suitable. However, the
ligands of zinc complexes could affect the electrochemical perfor-
mance of the half-cell electrode reactions [205]. The addition of Overall reaction:
bromine sequestration agents (e.g., MEP, MEM) leads to better
Cl2 þ Zn4ZnCl2 ; Ecell ¼ 2:12V (35)
reversibility in terms of anodic to cathodic current ratios in the
cyclic voltammograms as shown in Fig. 13 (due to the aggregation In the charging process, chlorine gas is evolved at the positive
and effective complexation with bromine [131]). electrode and stored in another chamber by mixing with water to
Similar to other hybrid flow batteries, shunt currents cause form solid chlorine hydrate at c.a. 10  C. When the battery undergoes
uneven electrodeposition in each unit cell and result in lower discharge, chlorine is later reduced back to a soluble species. A 10 kW
discharge capacities [71,207]. This can be minimized by appropriate (50 kW h) system was tested, and achieved an energy efficiency of
cell designs with feed channels to each cell, long and narrow c.a. 60% [223] with cycle lives of up to 500e5000 cycles [224].
enough to increase the electrical resistance. To improve the Among the early studies, Jorne et al. [132] evaluated the half-cell
morphology of zinc electrodeposits, organic [208,209] and inor- overpotentials of a similar hybrid flow battery, in which a graphite
ganic additives [210] have been used. In some cases, ionic liquid plate and porous graphite were used as the negative and positive
additives used for bromine sequestration [208,211] and inorganic electrodes, respectively. The two electrodes had active areas of
salts, e.g., zinc perchlorate [71], were observed to reduce the crystal 68 cm2 and were separated by a gap of 0.2 cm. The electrolyte was
size and prevent dendritic formation. Surfactants are known to 2 M zinc chloride solution and was circulated at 0.06 L min1.
suppress zinc dendrites and promote mixing of the aqueous and Despite the use of porous electrodes, high overpotentials were
polyamide phases, enabling the bromine reaction to take place observed at the chlorine positive electrode. The overpotentials for
evenly on the electrode surface [209]. Poor mixing of the the charge and discharge processes were þ120 mV and 200 mV,
polybromide-complex phase (particularly at the bottom of the cell) respectively. This was mainly caused by slow kinetics and high
is commonly observed in zinc-bromine flow batteries [209]. ohmic losses (through the electrolyte) within the porous structure.
For the bromine positive electrode reactions, three-dimensional The ohmic resistance can be reduced by adding potassium chloride
carbon based materials (mainly carbon felts [131,212] and reticu- salts. On the other hand, the overpotential of the zinc negative
lated vitreous carbon [130,213]) have been used in typical systems electrode reaction was relatively small (<50 mV) during both the
to minimize the overpotentials. Recent investigations have further charge and discharge processes. The voltage and columbic effi-
enhanced the reaction kinetics by using electrode materials with ciencies were 78 and 84%, respectively with an energy efficiency of
high activity, such as carbon nanotubes [212,214], active carbon approximately 66% [132].
[215] and carbon black [216,217]. Other approaches include surface The resulting zinc-chlorine hybrid flow battery had a practical
modifications [215,218] and electrochemical exfoliation, such as energy density of 154 Wh kg1, higher than that of similar zinc-
halogenation [219]. Some of these systems have been demonstrated bromine systems (60e85 Wh kg1). Since the system requires
98 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

additional chambers, the estimated cost was around USD$500e750/ 5.2. Ligand chemistry
kW h, significantly higher than the DoE cost target [224]. Energy loss
was attributed to the inefficient processes of hydrogen evolution and For both the negative and positive electrode reactions, redox
dissolution of chlorine at the negative and positive electrodes, species dissolved in the electrolytes often appear as coordinated
respectively, during the charging process. Other side reactions ion forms, although some of them are of mixed stability. At different
include the slow oxidation of graphite by the corrosive chlorine gas, pH and electrolyte compositions, metal ions often appear in
although the electrode life is often expected to exceed the goal of 10 different forms as metal complexes. As suggested by the Pourbaix
years [223]. Future research direction should focus on safer and more diagrams, the reactions and the corresponding potentials can be
efficient chlorine gas storage methods as well as more durable different at different pH and electrolyte composition. For some
electrode materials for long-term operation. complexes, the reactions may take several steps to achieve equi-
librium. The overall charge transfer rate can change if one of these
5. Research issues associated with the operating conditions reaction steps is hindered.
The stability constants of various metal chelates and anions are
More detailed investigations of both the negative and positive summarized elsewhere [115,230,231]. These values are associated
electrode reactions are essential to obtain improved charge- with dissociation constants, which are intimately related to the
discharge performance for further developments. Attention solubilities. Those complex ions with low stability constants, such
should be paid to technical issues associated with operating pa- as chloride, tend to have higher solubilities than their sulphate
rameters (e.g., current density and temperature) and the half-cell counterparts. Electrode reactions may take place at lower electrode
chemistries (pH, ligand chemistry, hydrogen evolution and corro- overpotentials and reaction kinetics can be improved due to
sion). Overcoming these issues is essential to further improving the weaker ligands. Therefore, electrolytes with mixed anions may
battery performance and enabling full commercial potential. further improve both the reaction kinetics and solubilities of the
active species [232e235].
5.1. pH For zinc electrodeposition, it is generally easier to deposit on the
substrate surface using these complexes, with higher resultant
Electrolyte pH is a crucial factor determining the forms and the deposition rates and current efficiencies. However, as discussed in
reaction mechanisms of the dissolved species as well as their sol- Sections 2 and 3, the low deposition overpotential tends to yield
ubilities, conductivities and electrode potentials. Under certain poor brightening and levelling performance. Therefore, suitable
conditions, the relationships between electrode potentials, pH and additives or complexing agents should be incorporated in the
the forms of dissolved species have been established in Pourbaix electrolytes [236]. In contrast, despite improved morphologies for
diagrams. Side reactions are also often influenced by the acid or those complexes with higher stability constants, the deposition
alkaline contents of the electrolytes. For instance, higher acid process tends to be less efficient in terms of current efficiencies,
concentrations tend to promote and suppress hydrogen and oxygen especially at low ion concentrations.
evolutions, respectively [91,124,225]. Moreover, these gaseous re-
actions consume protons or hydroxide themselves, leading to 5.3. Current density
further changes in the pH values of the solutions.
Many systems employ ion-exchange membranes (cation, anion, In zinc-based hybrid flow batteries, the negative and positive
filtration), to transfer proton or hydroxide ion between the two electrode reactions tend to take place under mixed control due to
half-cells in order to maintain electroneutrality. When a proton- the distribution of current, potential and electrolyte flow velocity,
exchange membrane (e.g., Nafion®) is used in an acidic solution, although it is ideal to avoid mass transport limitations by ensuring
protons move from the positive half-cell to its negative counterpart that the reactions are under charge-transfer control. For soluble
to form acid by complexing with the anions during the zinc elec- redox couples involving liquid phase reactions, this can be achieved
trodeposition process (reduction of Zn2þ to Zn0). Therefore, acidity easily by reducing the local current densities with the use of high-
or alkalinity changes are almost inevitable in the battery processes. surface-area or porous electrodes rather than flat, planar electrodes.
The increase in acid concentration is known to be undesirable Depending on the available active areas of these materials,
for the zinc negative electrode reactions in terms of electrodepo- higher current outputs are often expected, compared to planar (or
sition and corrosion processes. Buffer solutions, such as acetate, two-dimensional) electrode materials. Previous investigations
have been used in some acidic systems to control the concentration demonstrate that current densities with typical porous or three-
of free protons within a preferred range [159]. In the case of zinc- dimensional materials, such as carbon felts and carbon papers,
bromine batteries, a narrow range of pH between 1 and 3.5 needs could exceed 200 mA cm2 for liquid-phase reactions (e.g., vana-
to be maintained throughout operation [226]. For some positive dium and cerium) by minimizing the cell resistances with suitable
active species, such as cerium, the influence of a particular acid on cell architectures and membrane selections [235,237,238].
the solubility of the oxidizied species can be contrary to the influ- However, the current densities used in most zinc-based systems
ence on the reduced species. Therefore, a high acid concentration are usually less than 50 mA cm2, restricted by the area available
(>3 M) is required to dissolve a modest concentration of active for zinc electrodeposition at the negative electrode. A number of
species (e.g., 0.8 M). The potential transfer of this mass of acid high-surface-area electrodes, such as carbon felts and nickel foams,
through the membrane implies difficulties for the zinc electrode- have been used in zinc hybrid flow batteries under acidic and
position process. At high acid concentrations, a low current effi- alkaline conditions [136,159]. It was demonstrated that reasonable
ciency of zinc electrode reaction is expected due to excessive energy efficiencies (>50%) can be achieved at ultra-high current
hydrogen evolution and electrodeposit corrosion during the charge densities of up to 300 mA cm2 [136]. Compared to a planar surface,
[227] and discharge processes [62], respectively. In general, it is often much more difficult to electrodeposit zinc in porous
hydrogen evolution can be observed at low pH (pH < 3) but become structures [130,159]. This is possibly due to the requirements of
excessive when the pH is below 1 [130]. In contrast, a high pH may high overpotentials for uniform electrodeposition, characterized by
lead to the precipitation of a Zn(OH)2 layer, which results in low the ‘throwing power’ e the ability to electrodeposit into low cur-
current efficiency [228]. pH buffers may be used to stabilize the pH rent density areas with the same thickness as in the areas of high
in the electrolytes [229]. current density.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 99

Among common zinc electrolytes used in the electroplating both the negative and positive electrode reactions, which are heavily
industry, the throwing power under a standard condition is of the influenced by the operating parameters (e.g., current density) and
following order: high cyanide > low cyanide > acid > alkaline non- electrolyte compositions (e.g., pH). In both acidic and alkaline solu-
cyanide [239]. At low current densities (<20 mA cm2), the current tions, zinc negative electrode reactions tend to be reversible and
efficiencies are often higher than 80% in all of these baths, as shown exhibit high current efficiencies (>85%) at low current densities (c.a.
in Fig. 14, and acidic baths tend to exhibit higher current efficiencies 20 mA cm2) and ambient temperatures [41,88,207,239]. The influ-
than their alkaline counterparts. However, early work reported that ence of temperature on zinc electrodeposition tends to vary signifi-
dendrite-free zinc can be electrodeposited well into the porous cantly in the literature [88,172,248e251]. In both acidic and alkaline
reticulated vitreous carbon foam at 100 mA cm2 with organic electrolytes, there are reports of high coulombic efficiencies at
additives and flowing electrolytes [240]. elevated temperatures [172,248e250], while some authors report
In the existing literature, zinc electrodeposition has mainly been the opposite result due to the stronger depolarizing effect on
carried out on flat, planar electrodes. In some electrolytes, dendritic hydrogen evolution [88,251] and faster corrosion rates [62]. At
growth is observed when the current density is higher than elevated temperatures, additive adsorptions and polarizations tend
15 mA cm2 [130]. At increased current densities, ‘burned’ coatings to become weaker. In some cases, additives may lose their function
can be observed when the supply rate of the zinc ions to the and become insoluble in the electrolytes [252], leading to poorer
cathode surface is insufficient to maintain the electrodeposition efficiency, deposit adherence and morphology [251].
rate. This phenomenon can be avoided by increasing the zinc(II) ion In practice, the influence of temperature should take into ac-
concentration and the limiting current density through convective count the positive electrode reactions. For instance, the positive
mass transport control [241]. In some extreme cases, ultra-high cerium reaction tends to be a sluggish process and appears to be the
current densities of up to 1300 mA cm2 have been reported limiting reaction of the battery at low temperatures (<50  C). With
with the use of very high electrolyte flow rates (up to 5 m s1) the limited area of a platinised titanium mesh (compared to high-
[242]. surface-area carbon felts), the half-cell coulombic efficiency of the
When the current density increases from 30 to 500 mA cm2, cerium reaction was reported to be below 50% at 25  C [124]. As
the surface morphology changes from hexagonal platelets and suggested by fundamental studies [91], elevated temperatures
ridges to a pyramidal texture in acidic sulphate solutions [243]. (50  C) effectively improve the reversibility by virtue of the faster
However, increased current densities often lead to lower current reaction kinetics and higher electrolyte conductivity, diffusion co-
efficiencies caused by side reactions in both the acidic and alkaline efficients and viscosity. In the case of zinc-bromine batteries, the
electrolytes [244,245]. When the current densities are too small, operating temperatures are between 20 and 50  C and have little
electrodeposition rates may be too slow to form any coating, but influence on overall energy efficiencies as shown in Fig. 15. At low
generally bright electrodeposits are obtained [246,247]. In practical temperature, the electrolyte resistivity increases, resulting in a
systems, zinc electrodeposition in acidic and alkaline electrolytes lower voltage efficiency, which is offset by the higher coulombic
have shown reasonable efficiencies in parallel plate flow cells at efficiency attributed to slower self-discharge process caused by
15e50 mA cm2 for prolonged charge and discharge cycles (charge crossover of bromine. At high temperature, the resistance decreases
duration of up to 20 h) [124,129,130]. but with faster crossover of bromine species, with these two effects
partially cancelling each other out [130,207].

5.4. Temperature 5.5. Hydrogen evolution

Temperature has an influence on the electrode potentials, elec- Hydrogen evolution tends to take place concurrently with zinc
trolyte solubilities, electrolyte resistances, reaction kinetics and electrodeposition as a side reaction at the negative electrode, and is
diffusion coefficients of the active species. For zinc-based flow bat- considered as a major loss of current efficiency at high current
teries, it is often difficult to define the optimum temperatures for

Fig. 14. Comparisons of half-cell coulombic efficiency of zinc electrodepositions in Fig. 15. Influence of temperature on system efficiencies of a zinc-bromine hybrid flow
acidic and alkaline electrolytes [239]. batteries [207].
100 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

densities. Depending on the pH of the solution, the hydrogen particularly surfactants, often inhibit corrosion effectively by
overpotential can be as high as 1.0 V with certain organic additives adsorption, and geometric blocking of the zinc surfaces. On the
and substrate materials [253]. When a metal substrate is used, the other hand, inorganic additives, such as lead and indium, often
rates of hydrogen adsorption and evolution tend to be higher at form less chemically active metallic layers on the zinc surface.
high current density, low temperature and low pH [253,254]. In the However, these inhibition effects are sometimes temporary during
absence of organic additives, zinc electrodeposition with over 90% prolonged operation (e.g., 10 h) since the adsorption of organic
half-cell coulombic efficiency is often obtained with carbon-based additives and the metallic layers may not persist, especially under
substrates at relatively high current densities (at 50 mA cm2) in continued hydrogen evolution [62]. Due to the strong adsorption
several acid electrolytes [41,88]. The mechanisms of hydrogen and blocking properties of certain additives, there is no guarantee
adsorption and evolution have been discussed previously in the of achieving high half-cell efficiencies for zinc-based battery
literature [253,255e258]. applications.
In the field of zinc electroplating, cationic surfactants are often
used to suppress hydrogen evolution [33,109,110,259,260], while
preventing the growths of mossy and dendritic morphologies 6. Applications of zinc-hybrid flow batteries
[259,261e263]. In particular, cetyl trimethylammonium-based
surfactants are commonly used and known to suppress hydrogen The increasing share of power from intermittent renewable
evolution at low cost [33]. Other organic inhibitors for suppressing energy sources has led to a rapid development in redox (or hybrid)
hydrogen evolution include tartaric acid, succinic acid and phos- flow batteries, which are one of the few viable options for flexible
phoric acid, which also provide levelling performance for zinc grid-scale applications. Despite various flow battery chemistries,
electrodepositions [264]. However, some of these organic additives only the all-vanadium, zinc-bromine, zinc-cerium, zinc-nickel and
have large molecular weights or long hydrocarbon chains, and the zinc-iron (zinc-ferricyanide) systems have successfully been
strong absorption and inhibiting effects often slow down the scaled-up or commercialized between kW and MW scales. In
charge-transfer processes [265] and decrease the half-cell contrast to all-vanadium batteries, the other systems are all based
coulombic efficiencies [88]. Inorganic inhibitors include soluble on zinc negative electrodes, and all exhibit higher operating volt-
salts or metal oxides of indium, bismuth, tin, gallium, thallium, ages (1.58 V vs. 1.4 V for all-vanadium) (Fig. 16). The chemistries of
bismuth and lead, used in aqueous electrolytes [160,266]. It should these scaled-up or commercial systems were introduced over a
be noted that some of these elements are highly toxic and/or decade ago, and their specifications and installation details are
hazardous to the environment. summarized in Tables 4 and 5, respectively.
Among these zinc-based systems, zinc-bromine batteries have
5.6. Zinc corrosion been the most studied. They have been used in a number of ap-
plications, including load-levelling (up to MW, grid-connected or
Since zinc is an active material, corrosion or self-dissolution of off-grid), power quality control, coupling with renewable energy
zinc leads to energy losses in the battery systems. According to the sources and electric vehicles (Table 4, Fig. 17aee). Load-levelling is a
Pourbaix diagrams, no surface oxides are stable in acidic solutions; strategy to store excess energy from a power plant during off-peak
instead zinc dissolution and hydrogen evolution tend to take place hours and release it when demand increases. Other zinc-based flow
simultaneously at a mixed potential. In general, higher rates of batteries (e.g., zinc-nickel, zinc-cerium and zinc iron) have been
dissolution and hydrogen evolution are observed at higher proton/ piloted at smaller scales (kW) for load-levelling and energy storage
acid concentrations. The rate of dissolution may slow at certain coupling with renewable energy sources. Zinc-bromine hybrid flow
high acid concentrations (e.g., >6 M) due to the restricted avail- batteries have been shown in a pilot to be capable of prolonged
ability of dissociated protons in the solution [62]. In the pH range charging of up to 10 h [130], which is as competitive as commercial
1e4, this process is via cathodic control and the kinetics of all-vanadium systems.
hydrogen evolution dictate the overall corrosion [267]. In alkaline
solutions, zinc corrosion products are mainly zinc oxide and ZnOHþ
(according to thermodynamics), which are not an effective pro-
tection barrier. In the pH range 4e11, the corrosion rates do not vary
significantly, which is attributed to the change in the cathodic re-
action from hydrogen evolution to oxygen reduction. When the
local pH is between 11 and 12, protective zinc oxides may be
formed at certain negative electrode potentials, although minimal
corrosion may take place at lower pH in alkaline solutions [267].
In general, the corrosion rates are strongly influenced by the
morphology and crystallographic planes, which are highly influ-
enced by the electrolyte compositions [268]. Crystallographic and
close packed planes tend to involve higher binding energies of the
surface atoms, resulting in better corrosion resistance. Other
studies have demonstrated that electrodeposits with finer grain
sizes tend to have higher hardness and corrosion resistance
[29,30,269,270]. The use of organic additives in both acidic and
alkaline solutions often improves the electrodeposit morphologies
and offers high current efficiencies of >90% at < 40 mA cm2. The
resulting electrodeposits also show improved corrosion resistance
in corrosive chloride solutions compared to electrodeposits ob- Fig. 16. Cell voltage vs. time response during charge-discharge cycle for all-vanadium
(30 mA cm2 for 2 h) [74], zinc-bromine (15 mA cm2 for 10 h) [130], zinc-nickel
tained in the absence of additives [271]. (10 mA cm2 for c.a. 2 h [63,117]), zinc-ferricyanide (charging at 20 mA cm2 for c.a.
Apart from improving the zinc morphologies, some electrolyte 1.5 h; discharging at 35 mA cm2) [121] and the undivided zinc-cerium (20 mA cm2
additives also serve as corrosion inhibitors. Organic additives, for 30 min) [66,67] hybrid flow batteries.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 101

Table 4
Specifications and installation information of commercial zinc-based hybrid flow batteries.

Zinc-bromine Zinc-cerium Zinc-nickel Zinc-iron (Zinc-ferricyanide)

Cell open-circuit voltage/V 1.8 2.4 1.73 1.53


Energy density/W h L1 Up to 65 12e20 Up to 50 e
Cycle round-trip DC 65e75 63 >85 65e75
energy efficiency/%
Cycle life (cycles) >2000 NG >500 >5000
Total system cost/USD$ kW1 1044 750 e >300
Cost attribution of storage 80 50 e e
module/%
Typical size range/MW h 0.01e5 NG 0.1 0.04e2
Unit design life time/years 5e10 15 >5 20
Stage of development Demonstration/Commercial Demonstration Demonstration Demonstration/Commercial
units units
Major companies involved ZBB Energy (recently as Ensync Plurion Urban Electric Power ViZn Energy System
Energy System), Premium
Power (recently as Vionx
Energy), Kyushu Electric Power,
Meidensha, Primus Power and
Redflow
Number of installations >20 2 1 >5
Largest installation 1 MW in Kyushu, by Kyushu 2 kWe1 MW testing facility in 100 kW in New York, US by 1000 kW in ComEd Bronzeville,
Electric Power Glenrothes, Scotland by Plurion Urban Electric Power US by ViZn Energy System

Table 5
Existing installations and applications of zinc-based hybrid flow batteries.

System Company Customer Basic specification Application Installation date

Zinc-bromine Ensync Energy Detroit Edison, USA 400 kW h Load levelling Jun 2001
(previously ZBB United Energy, Melbourne, Australia 200 kW h Demonstration for Nov 2001
Energy) network storage applications
Nunawading Electrical Distribution 400 kW h Load levelling 2001
Substation in Box Hill, Australia
Australian Inland Energy, Australia 500 kW h Solar energy Jun 2002
Power Light, USA 2  50 kW h Solar energy Nov 2003
Pacific Gas and Electric Co., USA 2 MW h Peak power capacity Oct 2005
Dundalk Institute of Technology, Ireland 125 kW/500 kW h Wind energy Dec 2008
Illinois Institute of Technology, USA 500 kW h Microgrid Sept 2014
Fort Sill,Oklahoma, USA 500 kW h Microgrid 2013
Kyushu Electric Imajuku substation in Kyushu 1 MW/4 MW h Electric-utility applications 1990
Power & Electric Power, Japan
Meidensha
Redflow University of Queenland, Australia 12  120 kW h Solar energy Apr 2011
Department of Albuquerque, New Mexico, USA 2.8 MW h Solar energy 2011
Energy,US
Vionx Energy Massachusetts, USA 0.5 MW/3 MW h Peak power capacity Nov 2016
(previously
Premium Power)
Zinc-cerium Plurion Glenrothes, Scotland 2 kWe1 MW Testing facility 2007
Zinc-nickel Urban Electric New York, USA 100 kW Peak power capacity Jun 2013
Power
Zinc-iron ViZn Energy System Flathead Electric Cooperative, USA 80 kW/160 kWh Utilities and power Mar 2014
BlueSky Energy, Austria 64 kW Solar energy and Microgrid Nov 2013
Ontario, Canada 2 MW/6 MW h Frequency regulation Aug 2015
Ancillary services
Randolph-Macon College, US 48 kW Solar energy Apr 2015
ComEd Bronzeville, USA 1000 kW Solar energy Mar 2016
Idaho National Laboratory, USA 128 kW Microgrid research Jan 2016

Early grid-scale applications of zinc-bromine batteries from demonstration purposes. Meanwhile, research and development
1 kW to 60 kW modules were enabled by the Moonlight project, on the zinc-bromine batteries continued in Australia at Murdoch
partly sponsored by the Japanese government in the 1980s. These University (Parker and co-workers) and ZBB Technologies (now
modules were later used in a 1 MW/4 MW h system (twenty four Ensync Energy Systems, USA). ZBB technologies expanded its
25 kW modules) in Fukuoka by Kyushu Electric Power Company business in the USA and acquired the assets of Johnson Controls,
(Japan), and this has been the largest zinc-bromine system to date. Inc. to manufacture zinc-bromine batteries [272]. In 2004, ZBB
In the 1980s, Exxon's (USA) patent of the zinc-bromine flow battery Technologies was awarded a contract from the California Energy
was transferred to Johnson Control Inc. (USA) and licensed to Commission to demonstrate a 2 MW/2 MW h system for load-
Studiengesellschaft für Energiespeicher and Antriebssysteme, levelling applications [207].
S.E.A. (later known as Powercell, Austria), Toyota Motor (Japan), In the USA, Premium Power Corp. (now Vionx Energy, USA) used
Meidensha (Japan) and Sherwood Industries (Australia). In the technology developed by Powercell (Austria, formerly S.E.A.,
Australia, several of Exxon's systems at 3e20 kW were installed for Austria) but now manufactures all-vanadium batteries. Since the
102 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

2000s, Redflow Ltd. (Australia) and Primus Inc. (USA) have manu-
factured zinc-bromine flow batteries for household and load-
levelling applications (up to 600 kW h). As claimed by Primus
Inc., a membrane-less system was developed and ready for
commercialization [273]. The zinc-cerium technology has been
developed and scaled-up by Plurion Inc. (Scotland) and Applied
Intellectual Capital (USA) to kW scale until recently. Since the
2010s, ViZn Energy Inc. (a former zinc-air battery company, Zinc Air
Inc., USA) has manufactured zinc-iron (zinc-ferricyanide) flow
batteries for load-levelling applications from kW to MW scales
[274].
Following the oil crisis in 1970s, Energy Development Associates
(USA) investigated the use of a 50 kWh zinc-chlorine flow battery
for electric vehicle applications. Since the Exxon project in the
1980s, S.E.A. (Austria) has tested a number of batteries ranging from
5 to 45 kW h in several electric vehicles, accumulating more than
80,000 km [97,207,275]. The company has installed a 45 kW h,
216 V battery in a Volkswagen bus used by the Austrian Postal
Service in mountainous areas. The battery weight is around 700 kg
and allows a maximum range of up to 220 km at 50 km h1
(Fig. 17c) [207].
In the 1990s, electric vehicles with zinc-bromine batteries were
also tested by the University of California (a 35 kW h system) and
demonstrated by Toyota Motors (Japan) as model EV-3036 (7 kW h,
106 V) [207]. Fiat Automobiles (Italy) has also installed an 18 kW h
zinc-bromine battery (72 V, 250 A h) in a Fiat Panda city car. Hot-
zenblitz GmbH (Germany) designed an electric vehicle with a
15 kW h/114 V zinc-bromine battery. Despite a relatively long
lifetime, the lack of an adequate peak power of the zinc-bromine
flow battery has been the main obstacle for electric vehicle appli-
cations [97].

7. Conclusions and future challenges

Over the past few decades, zinc-based hybrid flow batteries


based on various chemistries have been proposed and developed in
both academic institutions and industry. These batteries share the
advantages of high voltage and good reversibility, while using low-
cost and abundant active materials. In all systems, zinc electrode-
position occurs in the presence of a flowing electrolyte e an
essential strategy to extend the cycle life. In contrast, conventional
static batteries have cycle lives limited to less than several hundred
cycles [96]. This is caused by dendritic growth that tends to develop
under prolonged cycling (or at high current densities) and entails
the risk of an electrical short-circuit.
For these reasons, it is essential to control the zinc electrode-
posit morphologies, which are strongly influenced by the electro-
deposition overpotentials and are sensitive to both electrolyte
compositions and additives. Organic additives with large molecules
are often used to increase the overpotentials and obtain finer grain
sizes by absorbing or partially covering at the electrode surfaces,
while some additives with smaller molecules serve as levellers
based on a similar inhibitive effect to smoothen the electrodeposit
morphologies. However, lower half-cell efficiencies are sometimes
observed as the cost of improving the morphology qualities.
In addition to the negative zinc electrode reaction, many pro-
posed systems use various positive electrode reactions based on
inorganic or organic materials in either solid, liquid or gaseous
phases. These reactions are facilitated by specific cell architectures

Fig. 17. (a) The overall applications of zinc-bromine hybrid flow batteries and their
installations: (b) with solar energy sites at Marine Corps Air Station (MCAS) in Mir-
amar, California by Primus Power Corp.; (c) with solar energy sites at New Zealand by
Redflow Ltd.; (d) a Volkswagen bus used for Austrian Postal Service equipped with a
45 kWh/216 V zinc-bromide redox flow batteries by S.E.A.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 103

under different circumstances. For instance, solid-phase active 50 mA cm2 (vs. up to >100 mA cm2 for all-vanadium) partly as a
materials do not involve soluble active species and undergo pure consequence of dendrite issues and the use of planar electrodes. In
solid-phase transformation; therefore a low-cost membrane-less recent studies, high-surface-area (or porous) electrodes have been
architecture is possible. However, the capacities of the existing demonstrated in acidic and alkaline electrolytes to yield reasonable
positive ‘paste’ electrodes (<50 mA h cm2 [65,139]) are still far too energy efficiencies (>50%) at ultra-high current densities (up to
low to match with the negative zinc electrode reactions (up to 300 mA cm2). Metallic fibres or meshes have also been used for
500 mA h cm2). metal electrodepositions as in other types of hybrid flow batteries
In contrast, liquid-phase positive active materials allow storage [276]. The distribution and morphologies of zinc electrodeposits
capacities to be expanded by increasing the volume and the con- can be varied by adjusting carbon surface properties, which can be
centrations of the positive electrolytes but systems are still limited modelled by density functional theory calculations [277,278]. It is
by the effective amount of zinc electrodeposits (up to also advisable to reduce the overall resistance and improve the
500 mA h cm2) as in most hybrid flow batteries. In certain cases, a mass transport of the active species with suitable cell components
similar membrane-less architecture is still possible by making use and architectures (minimized inter-electrode gaps; microporous,
of the relatively slow dissolution of zinc at low concentrations of selective separators; optimized flow fields) as suggested in recent
acid and particular positive active species [68,125]. However, this review paper [279].
also implies low specific energy or energy density. In recent studies, For proof-of-concept chemistries, a series of scientific and
several attractive positive species have been introduced and show technological issues need to be considered:
reasonable performance. For instance, organic active materials have
relatively positive electrode potentials (c.a. 0.8 V) and some un- (i) Understanding the influence of zinc electrodeposit mor-
dergo two-electron transfers. Iodine (inorganic) has solubilities of phologies on electrochemical cell performance.
up to c.a. 5.6 M, resulting in an energy density (of the positive (ii) Extending the storage capacity by obtaining uniform elec-
electrolyte) (c.a. 150e200 W h L1) several times higher than that of trodeposit morphologies with flowing electrolytes and suit-
the all-vanadium electrolytes (20e35 W h L1). These new systems able electrolyte compositions.
are based on simple cell designs, making them even more prom- (iii) Optimizing the solubilities of both oxidized and reduced
ising than the existing commercial batteries. species with suitable complexes and electrolyte
Some other systems involve gaseous positive active materials. compositions.
For example, oxygen and bromine are generated as the charged (iv) Evaluation of two- and three-dimensional electrode mate-
products of the positive electrode reactions. Oxygen is continuously rials in terms of performance and long-term stability for a
supplied from atmospheric air to the batteries. The storage capac- broad range of current densities.
ities of the positive electrolytes are therefore only limited to the (v) Systematic monitoring and controlling of pH changes in
hydroxide solubilities (up to > 20 M). Compared to oxygen, bromine prolonged charge-discharge cycling.
is harmful and needs to be sequestered in an oily phase, which is (vi) Evaluating and minimizing the influence of the crossover of
then stored in special chambers. In single-flow systems, the posi- the active species across the half-cells with selective but low-
tive electrolyte compartment is well-sealed to avoid any possible cost separators (or membranes).
emission or leakage [131]. Despite the low-cost active materials, the (vii) Corrosion of zinc electrodeposit and self-discharge of batte-
overall cost of the zinc-bromine system (current cost: USD$ 400/ ries should be minimized by using suitable electrolyte
kW h) may not be lower than the all-vanadium counterparts due to compositions and additives, while maintaining reasonable
the use of sequestering agents [200]. half-cell coulombic efficiencies.
Zinc-bromine and zinc-ferricyanide are the earliest systems, (viii) Improved cell designs to minimize the overall resistance and
developed for more than 40 years and successfully commercialized improve the mass transport of active species.
up to MW scales. Together with all-vanadium systems, these zinc- (ix) Developing appropriate mathematical models to simulate
based batteries are the only commercialized systems in use, for a and rationalise battery performance.
range of applications from load-levelling to powering electric ve-
hicles. Despite the nature of hybrid flow batteries, commercial zinc- Acknowledgements
based batteries have been demonstrated to undergo prolonged
discharging (or charging) of up to 10 h, which is comparable to all- Authors thank European Regional Development Funds (ERDF,
vanadium systems, making these systems competitive for various FEDER), Generalitat de Catalunya for financial support through the
applications. However, the storage capacities of zinc negative CERCA Program, MINECO for additional support by coordinated
electrodes at this stage are still limited to less than 1000 mA h cm2 project ENE2016-80788-C5-5-R and Fundacio n Ramon Areces
(mostly < 500 mA h cm2). This value needs to be further improved partial funding through BAT-LIMET project.
by identifying suitable electrolyte compositions and charging
strategies.
References
In the long-term, the capital cost should be lower than USD$
150/kW h (DoE target) for effective market penetration. [1] P. Leung, X. Li, C. Ponce de Leon, L. Berlouis, C.T.J. Low, F.C. Walsh, Progress in
Throughout this review, several chemistries (e.g., zinc-organic and redox flow batteries, remaining challenges and their applications in energy
zinc-air) have been identified to be even more promising than the storage, RSC Adv. 2 (2012) 10125e10156.
[2] C. Ponce de Leon, A. Frias-Ferrer, J. Gonzalez-Garcia, D.A. Szanto, F.C. Walsh,
commercial chemistries. With recent advances in computational Redox flow cells for energy conversion, J. Power Sources 160 (2006)
modelling and synthetic techniques, metal complexes or organic 716e732.
compounds of higher voltages and multi-electron-transfers can [3] R.F. Service, Tank for the batteries, Science 344 (2014) 352e354.
[4] G. Kear, A.A. Shah, F.C. Walsh, Development of the all-vanadium redox flow
reduce the cost per kW h effectively, while simultaneously maxi-
battery for energy storage: a review of technological, financial and policy
mizing the power density. Increased power density implies smaller aspects, Int. J. Energy Res. 36 (2012) 1105e1120.
electrodes and fewer cells for a given power output, which can be [5] Grid Energy Storage, U.S. Department of Energy, 2013', http://energy.gov/oe/
more cost effective than using lower cost materials. downloads/grid-energy-storage-december-2013, (Accessed on)
[6] J. Winsberg, T. Hagemann, T. Janoschka, M.D. Hager, U.S. Schubert, Redox-
Despite the relatively high cell voltages, the current densities of flow batteries: from metals to organic redox-active materials, Angew. Chem.
most zinc-based hybrid flow batteries are still limited to less than Int. Ed. 56 (2016) 686e711.
104 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

[7] M. Park, J. Ryu, W. Wang, J. Cho, Material design and engineering of next- evolution visualized within intact cycling alkaline batteries, J. Mater. Chem. A
generation flow battery technologies, Nat. Rev. 2 (2016) 16080. 2 (2014) 2757e2764.
[8] P. Leung, A.A. Shah, L. Sanz, C. Flox, J.R. Morante, Q. Xu, M.R. Mohamed, [39] R. Patrice, B. Gerand, J.B. Leriche, L. Seguin, E. Wang, R. Moses, K. Brandt,
C.P.D. Leon, F.C. Walsh, Recent developments in organic redox flow batteries: J.M. Tarascon, Understanding the second electron discharge plateau in
a critical review, J. Power Sources 360 (2017) 243e283. MnO2-based alkalinecells, J. Electrochem. Power Sources 148 (2001)
[9] J. Li, L. Yang, B. Yuan, G. Li, J.Y. Lee, Combined mediator and electrochemical A448eA455.
charging and discharging of redox targeting lithium-sulfur flow batteries, [40] M. Chamoun, B.J. Hertzberg, T. Gupta, D. Davies, S. Bhadra, B.V. Tassell,
Mater. Today Energy 5 (2017) 15e21. C. Erdonmez, D.A. Steingart, Hyper-dendritic nanoporous zinc foam anodes,
[10] D.N. Østedgaard-Munck, J. Catalano, M.B. Kristensen, A. Bentien, Membrane- NPG Asia Mater. 7 (2015) (Published online 24 April 2015), https://www.
based electrokinetic energy conversion, Mater. Today Energy 5 (2017) nature.com/am/journal/v7/n4/full/am201532a.html.
118e125. [41] J. Pan, Y. Wen, J. Cheng, J. Pan, Z. Bai, Y. Yang, Zinc deposition and dissolution
[11] Y. Xu, Y. Wen, J. Cheng, G. Gao, Y. Yang, Study on a single flow acid Cd- in sulfuric acid onto a graphite-resin composite electrode as the negative
chloranil battery, Electrochem. Commun. 11 (2009) 1422e1424. electrode reactions in acidic zinc-based redox flow batteries, J. Appl. Elec-
[12] J. Winsberg, T. Hagemann, T. Janoschka, M.D. Hager, U.S. Schubert, Redox- trochem. 43 (2013) 541e551.
flow batteries: from metals to organic redox-active materials, Angew. Chem. [42] X. Li, D. Pletcher, C. Ponce de Le on, F.C. Walsh, R.G.A. Wills, Redox flow
Int. Ed. 56 (2017) 686e711. http://onlinelibrary.wiley.com/doi/10.1002/anie. batteries for energy storage using zinc electrodes, in: C. Menictas, M. Skyllas-
201604925/full. Kazacos, T.M. Lim (Eds.), Advances in Batteries for Large- and Medium-scale
[13] P. Leung, T. Martin, M. Liras, A.M. Berenguer, R. Marcilla, A. Shah, L. An, Energy Storage: Applications in Power Systems and Electric Vehicles,
M.A. Anderson, J. Palma, Cyclohexanedione as the negative electrode reac- Woodhead, 2015, pp. 293e315.
tion for aqueous organic redox flow batteries, Appl. Energy 197 (2017) [43] H. Jiang, W. Shyy, M. Wu, L. Wei, T. Zhao, Highly active, bi-functional and
318e326. metal-free B4C-nanoparticle-modified graphite felt electrodes for vanadium
[14] G.M. Weng, Z. Li, G. Cong, Y. Zhou, Y.C. Lu, Unlocking the capacity of iodide redox flow batteries, J. Power Sources 365 (2017) 34e42.
for high-energy-density zinc/polyiodide and lithium/polyiodide redox flow [44] L. Wei, T. Zhao, G. Zhao, L. An, L. Zeng, A high-performance carbon
batteries, Energy Environ. Sci. 3 (2017) 735e741. nanoparticle-decorated graphite felt electrode for vanadium redox flow
[15] 'Statistics and information on the worldwide supply of, demand for, and flow batteries, Appl. Energy 176 (2016) 74e79.
of minerals and materials essential to the U.S. economy, the national secu- [45] M. Skyllas-Kazacos, M.H. Chakrabarti, S.A. Hajimolana, F.S. Mjalli, M. Saleem,
rity, and protection of the environment', National Minerals Information Progress in flow battery research and development, J. Electrochem. Soc. 158
Center, https://minerals.usgs.gov/minerals/pubs/commodity/, (Accessed 20 (2011) R55eR79.
November 2017). [46] A.Z. Weber, M.M. Mench, J.P. Meyers, P.N. Ross, J.T. Gostick, Q. Liu, Redox
[16] London Metal Exchange, https://www.lme.com/Metals/Non-ferrous/ flow batteries: a review, J. Appl. Electrochem. 41 (2011) 1137e1164.
Zinc#tabIndex¼0, (Accessed 1 Septemper 2010). [47] J. Noack, N. Roznyatovskaya, T. Herr, P. Fischer, The chemistry of redox-flow
[17] M. Fleischmanm, L.R. Hill, G. Sundholm, J. Electroanal. Chem. 157 (1983) batteries, Angew. Chem. Int. Ed. 54 (2015) 9776e9809.
359e368. [48] S.H. Shin, S.H. Yun, S.H. Moon, A review of current developments in non-
[18] D.W. McComsey, Handbook of Batteries, third ed., vol. 8, McGraw Hill, 2002, aqueous redox flow batteries: characterization of their membranes for
pp. 184e228. design perspectives, RSC Adv. 3 (2013) 9095e9116.
[19] K. Kordesch, J. Daniel-Ivad, Handbook of Batteries, vol. 36, McGraw-Hill, New [49] F. Pan, Q. Wang, Redox species of redox flow batteries: a review, Molecules
York, 2002, pp. 1168e1185. 20 (2015) 20499e20517.
[20] X.G. Zhang, Corrosion and Electrochemistry of Zinc, first ed., vol. Chapter 1, [50] Q. Huang, Q. Wang, New-Generation, high-energy-density redox flow bat-
Springer, 1996, pp. 1e16. teries, ChemPlusChem 80 (2015) 312e322.
[21] D. Crotty, Zinc alloy plating for the automotive industry, J. Met. Finish. 94 [51] Q. Xu, T.S. Zhao, Fundamental models for flow batteries, Prog. Energy
(1996) 54e58. Combust. Sci. 49 (2015) 40e58.
[22] E. Budman, R.R. Sizelove, Zinc alloy plating, J. Met. Finish. 99 (2001) [52] R.M. Darling, K.G. Gallagher, J.A. Kowalski, S. Ha, F.R. Brushett, Pathways to
334e339. low-cost electrochemical energy storage: a comparison of aqueous and
[23] H.H. Geduld, Zincate or Alkaline Noncyanide Zinc Plating, ASM International, nonaqueous flow batteries, Energy Environ. Sci. 7 (2014) 3459e3477.
Metals Park, Ohio, 1988, pp. 90e106. [53] C. Choi, S. Kim, R. Kim, Y. Choi, S. Kim, H.Y. Jung, J.H. Yang, H.T. Kim, A review
[24] G. Trejo, Y. Meas, P. Ozil, E. Chainet, Nucleation and growth of zinc of vanadium electrolytes for vanadium redox flow batteries, Renew. Sustain.
from chloride concentrated solutions, J. Electrochem. Soc. 145 (1998) Energy Rev. 69 (2017) 263e274.
4090e4097. [54] A. Parasuraman, T.M. Lim, C. Menictas, M. Skyllas-Kazacos, Review of ma-
[25] P. Díaz-Arista, Y. Meas, R. Ortega, G. Trejo, Electrochemical and AFM study of terial research and development for vanadium redox flow battery applica-
Zn electrodeposition in the presence of benzylideneacetone in a chloride- tions, Electrochim. Acta 101 (2013) 27e40.
based acidic bath, J. Appl. Electrochem. 35 (2005) 217e227. [55] K.J. Kim, M.S. Park, Y.J. Kim, J.H. Kim, S.X. Dou, M. Skyllas-Kazacos,
[26] M. Schlesinger, M. Paunovic, Modern Electroplating, fourth ed., John Wiley & A technology review of electrodes and reaction mechanisms in vanadium
Sons Inc., 2000. redox flow batteries, J. Mater. Chem. 3 (2015) 16913e16933.
[27] F.A. Lowenheim, Electroplating, Technical Reference Publications, 1995. [56] Y. Zhao, Y. Ding, Y. Li, L. Peng, H.R. Byon, J.B. Goodenough, G. Yu, A chemistry
[28] K.M.S. Youssef, C.C. Koch, P.S. Fedkiw, Improved corrosion behavior of and material perspective on lithium redox flow batteries towards high-
nanocrystalline zinc produced by pulse-current electrodeposition, Corrosion density electrical energy storage, Chem. Sov. Rev. 44 (2015) 7968e7996.
Sci. 46 (2004) 51e64. [57] 'Zinc-Periodic Table', Royal Society of Chemistry, http://www.rsc.org/
[29] M.S. Youssef, C.C. Koch, P.S. Fedkiw, Influence of additives and pulse elec- periodic-table/element/30/zinc, (Accessed on).
trodeposition parameters on production of nanocrystalline zinc from zinc [58] J.B. Calvert, 'Zinc and Cadmium, University of Denver', http://mysite.du.edu/
chloride electrolytes, J. Electrochem. Soc. 151 (2004) C103eC111. ~jcalvert/phys/zinc.htm.
[30] A. Gomes, M.I. da Silva Pereira, Pulsed electrodeposition of Zn in the pres- [59] B. Berverskog, I. Puigdomenech, Revised Poubaix diagrams for zinc at
ence of surfactants, Electrochim. Acta 51 (2006) 1342e1350. 25e300 oC, Corrosion Sci. 39 (1997) 107e114.
[31] A.E. Saba, A.E. Elsherief, Continuous electrowinning of zinc, Hydrometallurgy [60] 'The Nernst Equation and Pourbaix diagram, ' TLP Library, DoITPoMS, Uni-
54 (2000) 91e106. versity of Cambridge, https://www.doitpoms.ac.uk/tlplib/pourbaix/index.
[32] J. Yu, Y. Chen, H. Yang, Q. Huang, The influences of organic additives on zinc php, (Accessed on).
electrocrystallization from KCl solutions, J. Electrochem. Soc. 146 (1999) [61] T.K.A. Hoang, T.N.L. Doan, K.E.K. Sun, P. Sun, Corrosion chemistry and pro-
1789e1793. tection of zinc & zinc alloys by polymer-containing materials for potential
[33] A. Gomes, M.I.D.S. Pereira, Zn electrodeposition in the presence of surfac- use in rechargeable aqueous batteries, RSC Adv. 5 (2015) 41677e41691.
tants: Part I. Voltammetric and structural studies, Electrochim. Acta 52 [62] P. Leung, C. Ponce-de-Leo n, F. Recio, P. Herrasti, F. Walsh, Corrosion of the
(2006) 863e871. zinc negative electrode of zinc-cerium hybrid redox flow batteries in
[34] A.M. Alfantazi, D.B. Dreisinger, An investigation on the effects of orthophe- methanesulfonic acid, J. Appl. Electrochem. 44 (2014) 1025e1035.
nylene diamine and sodium lignin sulfonate on zinc electrowinning from [63] L. Zhang, J. Cheng, Y.S. Yang, Y.H. Wen, X.D. Wang, G.P. Cao, Study of zinc
industrial electrolyte, Hydrometallurgy 69 (2003) 99e107. electrodes for single flow zinc/nickel battery application, J. Power Sources
[35] A. Gomes, A.S. Viana, M.I.D.S. Pereira, Potentiostatic and AFM morphological 179 (2008) 381e387.
studies of Zn electrodeposition in the presence of surfactants, J. Electrochem. [64] S. Banerjee, Y. Ito, M. Klein, M.E. Nyce, D. Steingart, R. Plivelich et al., 'Nickel-
Soc. 154 (2007) D452eD461. zinc flow battery', US patent 20130113431 A1, (2013).
[36] D. Mun ~ oz-Torrero, P. Leung, E. García-Quismondo, E. Ventosa, M. Anderson, [65] P.K. Leung, Q. Xu, T.S. Zhao, High-potential zinc-lead dioxide rechargeable
J. Palma, R. Marcilla, Investigation of different anode materials for aluminium cells, Electrochim. Acta 79 (2012) 117e125.
rechargeable batteries, J. Power Sources 374 (2018) 77e83. [66] P.K. Leung, T. Martin, A.A. Shah, M.A. Anderson, J. Palma, Membrane-less
[37] K. Gong, Q.R. Fang, S. Gu, S.F.Y. Li, Y.S. Yan, Nonaqueous redox-flow batteries: organic-inorganic aqueous flow batteries with improved cell potential,
organic solvents, supporting electrolytes, and redox pairs, Energy Environ. Chem. Commun. (2017) article in press.
Sci. 8 (2015) 3515e3530. [67] P.K. Leung, T. Martin, A.A. Shah, M.R. Mohamed, M.A. Anderson, J. Palma,
[38] J.W. Gallaway, C.K. Erdonmez, Z. Zhong, M. Croft, L.A. Sviridov, Membrane-less hybrid flow battery based on low-cost elements, J. Power
T.Z. Sholklapper, D.E. Turney, S. Banerjee, D.A. Steingart, Real-time materials Sources 341 (2017) 36e45.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 105

[68] P.K. Leung, C. Ponce de Leon, F.C. Walsh, An undivided zinc-cerium redox [97] P.T. Moseley, J. Garche, Electrochemical Energy Storage for Renewable
flow battery operating at room temperature (295 K), Electrochem. Commun. Sources and Grid Balancing, Chapter 17: Redox Flow Batteries, first ed.,
13 (2011) 770e773. Elsevier, 2015, pp. 309e336.
[69] A. Hazza, D. Pletcher, R. Wills, A novel flow battery: a lead acid battery based [98] V. Konschuter, J. Trans. Electrochem. Soc. 45 (1924) 229.
on an electrolyte with soluble lead(II) Part I: preliminary studies, Phys. [99] J. Torrent-Burgues, E. Guaus, F. Sanz, Initial stages of tin electrodeposition
Chem. Chem. Phys. 6 (2004) 1773e1778. from sulfate baths in the presence of gluconate, J. Appl. Electrochem. 32
[70] H. Liu, L. Yang, Q. Xu, C. Yan, Corrosion behavior of a bipolar plate of (2002) 225e230.
carbonepolythene composite in a vanadium redox flow battery, RSC Adv. 5 [100] A.M. Volmer, Z. Weber, Nuclei formation in supersaturated states, J. Phys.
(8) (2015) 5928e5932. Chem. 119 (1926) 277e301.
[71] C. Ponce de Leo n, A. Frías-Ferrer, J. Gonz
alez-García, D.A. Szanto, F.C. Walsh, [101] M.C. Li, S.Z. Luo, Y.H. Qian, W.Q. Zhang, L.L. Jiang, J.I. Shen, Effective of ad-
Redox flow cells for energy conversion, J. Power Sources 160 (2006) ditives on electrodeposition of nanocrystalline zinc from acidic sulfate so-
716e732. lution, J. Electrochem. Soc. 154 (2007) D567eD571.
[72] H. Liu, Q. Xu, C. Yan, Y. Qiao, Corrosion behavior of a positive graphite [102] M. Mouagnga, L. Ricq, J. Douglade, P. Berc, J. Appl. Electrochem. 37 (2007)
electrode in vanadium redox flow battery, Electrochim. Acta 56 (2011) 283e289.
8783e8790. [103] A. Hosny, Electrowinning of zinc from electrolytes containing anti-acid mist
[73] L. Wei, T.S. Zhao, Q. Xu, X.L. Zhou, Z.H. Zhang, In-situ investigation of surfactant, Hydrometallurgy 32 (1993) 261e269.
hydrogen evolution behaviour in vanadium redox flow batteries, Appl. En- [104] M. Karavasteva, St. Karivanov, Electrowinning of zinc at high current density
ergy 190 (2017) 1112e1118. in the presence of some surfactants, J. Appl. Electrochem. 23 (1993)
[74] M. Kazacos, M. Skyllas-Kazacos, Performance characteristics of carbon plastic 763e765.
electrodes in the all-vanadium redox cell, J. Electrochem. Soc. 136e9 (1989) [105] M. Karavasteva, The electrodeposition of metal impurities during the zinc
2759e2760. electrowinning at high current density in the presence of some surfactants,
[75] S. Zhong, M. Kazacos, R.P. Burford, M. Skyllas-Kazacos, Fabrication and Hydrometallurgy 35 (1994) 391e396.
activation studies of conducting plastic composite electrodes for redox cells, [106] K.N. Srinivasan, S. Venkatakrishna Iyer, Bull. Electrochem. 6 (1990) 35.
J. Power Sources 36 (1991) 29e43. [107] K.E.K. Sun, T.K.A. Hoang, T.N.L. Doan, Y. Yu, X. Zhu, Y. Tian, P. Chen, Sup-
[76] P. Qian, H.M. Zhang, J. Chen, Y.H. Wen, Q.T. Luo, Z. Liu, D.J. You, B.L. Yi, A novel pression of dendrite formation and corrosion on zinc anode of secondary
electrode-bipolar plate assembly for vanadium redox flow battery applica- aqueous batteries, ACS Appl. Mater. Interfaces 9 (2017) 9681e9687.
tions, J. Power Sources 175 (2008) 613e620. [108] B.C. Tripathy, S.C. Das, G.T. Hefter, P. Singh, Zinc electrowinning from acidic
[77] M.S. Yazici, D. Krassowski, J. Prakash, Flexible graphite as battery anode and sulfate solutions: Part I: effects of sodium lauryl sulfate, J. Appl. Electrochem.
current collector, J. Power Sources 141 (2005) 171e176. 27 (1997) 673e678.
[78] M.S. Pereira, L.L. Barbosa, C.A.C. Souza, A.C.M.D. Moraes, I.A. Carlos, The in- [109] C. Juhel, B. Beden, C. Lamy, J.M. Leger, R. Vignaud, Effect of surfactant Forafac
fluence of sorbitol on zinc film deposition, zinc dissolution process and on hydrogen evolution on a zinc electrode, Electrochim. Acta 35 (1990)
morphology of deposits obtained from alkaline bath, J. Appl. Electrochem. 36 479e481.
(2006) 727e732. [110] M. Maja, N. Penazzi, G. Farnia, G. Sandona, Zinc corrosion in NH4Cl and effect
[79] J.X. Yu, H.X. Yang, X.P. Ai, Y.Y. Chen, Effects of anions on the zinc electro- of some organic inhibitors, Electrochim. Acta 38 (1993) 1453e1459.
deposition onto glassy-carbon electrode, Russ. J. Electrochem. 38 (2002) [111] C. Cachet, B. Saidani, R. Wiart, The behavior of zinc electrode in alkaline
321e325. electrolytes I. A kinetic analysis of cathodic deposition, J. Electrochem. Soc.
[80] J. Heinze, A. Rasche, M. Pagels, B. Geschke, On the origin of the so-called 138 (1991) 678e687.
nucleation loop during electropolymerization of conducting polymers, [112] L.M. Mureşan, S.C. Varvara, Levelling and Brightening Mechanisms in Metal
J. Phys. Chem. B 111 (2007) 989e997. Electrodeposition, and Nunez Magdalena, Metal Electrodeposition, vol. 59,
[81] E. Budevski, G. Staikov, W.J. Lorenz, Electrocrystallization: nucleation and Nova Science Publishers, 1955, pp. 756e766.
growth phenomen, Electrochim. Acta 45 (2000) 2559e2574. [113] D.G. Foulke, O. Kardos, Current distribution in microprofiles, Proc. Am.
[82] F.C. Walsh, M.E. Harron, Electrocrystallization and electrochemical control of Electroplaters Soc. 43 (1956) 172e181.
crystal growth: fundamental considerations and electrodeposition of metals, [114] H.Z. Leidheiser, J. Elekrochem. 21 (1991) 565e574.
J. Phys. D Appl. Phys. 24 (1991) 217e225. [115] J.L. Fang, Theory and Application of Electroplating Additives, 2006,
[83] I.N. Stranski, Stoechiom. Verwandtschaftsl, Z. Phys. Chem. 136 (1928) 297. pp. 365e373.
[84] R. Becker, W. Do €ring, Kinetische Behandlung der Keimbildung in übers€ at- [116] Y. Ito, M. Nyce, R. Plivelich, M. Klein, D. Steingart, S. Banerjee, Zinc
tigten Da €mpfen, Ann. Phys. 24 (1935) 719. morphology in zinc-nickel flow assisted batteries and impact on perfor-
[85] K. Raeissi, A. Saatchi, M.A. Golozar, Effect of nucleation mode on the mance, J. Power Sources 196 (2011) 2340e2345.
morphology and texture of electrodeposited zinc, J. Appl. Electrochem. 33 [117] J. Cheng, L. Zhang, Y.S. Yang, Y.H. Wen, G.P. Cao, X.D. Wang, Preliminary
(2003) 635e642. Study of single flow zinc-nickel battery, Electrochem. Commun. 9 (2007)
[86] R.Y. Wang, D.W. Kirk, G.X. Zhang, Effects of deposition conditions on the 2639e2642.
morphology of zinc deposits from alkaline zincate solutions, J. Electrochem. [118] Y. Ito, X. Wei, D. Desai, D. Steingart, S. Banerjee, An indicator of zinc
Soc. 153 (2006) C357eC364. morphology transition in flowing alkaline electrolyte, J. Power Sources 211
[87] N. Soroura, W. Zhanga, E. Ghalia, G. Houlachi, A review of organic additives in (2012) 119e128.
zinc electrodeposition process (performance and evaluation), Hydrometal- [119] J. Pan, Y. Wen, J. Cheng, J. Pan, S. Bai, Y. Yang, Evaluation of substrates for zinc
lurgy 171 (2017) 320e332. http://www.sciencedirect.com/science/article/ negative electrode in acid PbO2-Zn single flow batteries, Chin. J. Chem. Eng.
pii/S0304386X17300373. 24 (2016) 529e534.
[88] P. Leung, C. Ponce de Leon, C.T.J. Low, F.C. Walsh, Zinc deposition and [120] Y. Zhao, S. Si, C. Liao, A single flow zinc/polyaniline suspension rechargeable
dissolution in methanesulfonic acid onto a carbon composit electrode as the battery, J. Power Sources 241 (2013) 449e453.
negative electrode reactions in a hybrid redox flow battery, Electrochim. [121] G.B. Adams, Electrically rechargeable battery, US Pat. 4180623, 25/12/1979.
Acta 18 (2011) 6536e6546. [122] S. Selverston, R. Savinell, J. Wainright, Zinc-iron flow batteries with common
[89] R.A. Putt, Assessment of Technical and Economic Feasibility of Zinc/bromine electrolyte, J. Electrochem. Soc. 164 (2017) A1069eA1075.
Batteries for Utility Load Leveling. Final Report, Researchgate, 1979. [123] K. Gong, X. Ma, K.M. Conforti, K.J. Kuttler, J.B. Grunewald, K.L. Yeager,
[90] N. Clark, P. Eidler, P. Lex, Development of Zinc/Bromine Batteries for Load- M.Z. Bazant, S. Gu, Y. Yan, A zinc-iron redox-flow battery under $100 per kW
levelling Applications: Phase 2 Final Report, SAND99e2691, Sandia, 1999. h of system capital cost, Energy Environ. Sci. 8 (2015) 2941e2945.
[91] P. Leung, C. Ponce de Leon, C.T.J. Low, F.C. Walsh, Ce(III)/Ce(IV) in meth- [124] P. Leung, C. Ponce de Leon, C.T.J. Low, A.A. Shah, F.C. Walsh, Characterization
anesulfonic acid as the positive half cell of a redox flow battery, Electrochim. of a zinc-cerium flow battery, J. Power Sources 196 (2011) 5174e5185.
Acta 56 (2011) 2145e2153. [125] P.K. Leung, C. Ponce de Leo n, F.C. Walsh, The influence of operational pa-
[92] S.C. Yang, An approximate model for estimating the faradaic efficiency loss in rameters on the performance of an undivided zinc-cerium flow battery,
zinc/bromine batteries caused by cell self-discharge, J. Power Sources 50 Electrochim. Acta 80 (2012) 7e14.
(1994) 343e360. [126] B. Li, Z. Nie, M. Vijayakumar, G. Li, J. Liu, V. Sprenkle, W. Wang, Ambipolar
[93] P.A. Adcock, A. Quillinan, B. Clark, O.M.G. Newman, S.B. Adeloju, Measure- zinc-polyiodide electrolyte for a high-energy density aqueous redox flow
ment of polarization parameters impacting on electrodeposit morphology. II: battery, Nat. Commun. 6 (2015), 6303. https://www.nature.com/articles/
conventional zinc electrowinning solutions, J. Appl. Electrochem. 34 (2004) ncomms7303.
771e780. [127] J. Winsberg, T. Janoschka, S. Morgenstern, T. Hagemann, S. Muench,
[94] J.E. Oxley, The Improvement of Zinc Electrodes for Electrochemical Cells, G. Hauffman, J. Gohy, M.D. Hager, U.S. Schubert, Poly(TEMPO)/Zinc hybrid-
NASA, ASA Contractor Report, 1966. flow battery: a novel, “Green,” high voltage, and safe energy storage sys-
[95] K.I. Popov, M.G. Pavlovi c, M.D. Spasojevi c, V.M. Naki c, The critical over- tem, Adv. Mater. 28 (2016) 2238e2243.
potential for zinc dendrite formation, J. Appl. Electrochem. 9 (1979) [128] A. Orita, M.G. Verde, M. Sakai, Y.S. Meng, The impact of pH on side reactions
533e536. for aqueous redox flow batteries based on nitroxyl radical compounds,
[96] D.E. Turney, J.W. Gallaway, G.G. Yadav, R. Ramirez, M. Nyce, S. Banerjee, J. Power Sources 321 (2016) 126e134.
Y.K. Chen-Wiegart, J. Wang, M.J. D'Ambrose, S. Kolhekar, J. Huang, X. Wei, [129] J. Pan, L. Ji, Y. Sun, P. Wan, J. Cheng, Y. Yang, M. Fan, Preliminary study of
Rechargeable zinc alkaline anodes for long-cycle energy storage, Chem. alkaline single flowing Zn-O2 battery, Electrochem. Commun. 11 (2009)
Mater. 29 (2017) 4819e4832. 2191e2194.
106 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

[130] H.S. Lim, A.M. Lackner, R.C. Knechtli, Zinc-bromine secondary battery, [164] G. Nikiforidis, W.A. Daoud, Effect of mixed acid media on the positive side of
J. Electrochem. Soc. 124 (1977) 1154e1157. the hybrid zinc-cerium redox flow battery, Electrochim. Acta 141 (2014)
[131] Q. Lai, H. Zhang, X. Li, L. Zhang, Y. Cheng, A novel single flow zinc-bromine 255e262.
battery with improved energy density, J. Power Sources 235 (2013) 1e4. [165] G. Nikiforidis, L. Berlouis, D. Hall, D. Hodgson, Charge/discharge cycles on Pt
, J.T. Kim, D. Kralik, The zinc-chlorine battery: half-cell overpotential
[132] J. Jorne and Pt-Ir based electrodes for the positive side of the Zinc-Cerium hybrid
measurements, J. Appl. Electrochem. 9 (1979) 573e579. redox flow battery, Electrochim. Acta 125 (2014) 176e182.
[133] T. Michaelowski, Russian Patent, Group XI, No. 5100, (1901). [166] Z. Xie, B. Yang, D. Cai, L. Yang, Hierarchical porous carbon toward effective
[134] D. Coates, A. Charkey, Handbook of Batteries, Chapter 31 Nickel-zinc Batte- cathode in advanced zinc-cerium redox flow battery, J. Rare Earths 32 (2014)
ries, third ed., McGraw Hill, 2002. 973e978.
[135] G. Bronoel, Development of Ni-Zn cells, J. Power Sources 34 (1991) 243e255. [167] Z. Xie, B. Yang, L. Yang, X. Xu, D. Cai, J. Chen, Y. Chen, Y. He, Y. Li, X. Zhou,
[136] Y. Cheng, X. Xi, D. Li, X. Li, Q. Lai, H. Zhang, Performance and potential Addition of graphene oxide into graphite toward effective positive electrode
problems of high power density zinc-nickel single flow batteries, RSC Adv. 5 for advanced zinc-cerium redox flow battery, J. Solid State Electrochem. 19
(2015) 1772e1776. (2015) 3339e3345.
[137] Y. Cheng, Q. Lai, X. Li, X. Xi, Q. Zheng, C. Ding, H. Zhang, Zinc-nickel single [168] Z. Na, X. Wang, D. Yin, L. Wang, Tin dioxide as a high-performance catalyst
flow batteries with improved cycling stability by eliminating zinc accumu- towards Ce(VI)/Ce(III) redox reactions for redox flow battery applications,
lation on the negative electrode, Electrochim. Acta 145 (2014) 109e115. J. Mater. Chem. A 5 (2017) 5036e5043.
[138] Y. Cheng, H. Zhang, Q. Lai, X. Li, D. Shi, L. Zhang, A high power density single [169] G. Nikiforidis, W.A. Daoud, Indium modified graphite electrodes on highly
flow zinc-nickel battery with three-dimensional porous negative electrode, zinc containing methanesulfonate electrolyte for zinc-cerium redox flow
J. Power Sources 241 (2013) 196e202. battery, Electrochim. Acta 168 (2015) 394e402.
[139] D.E. Turney, M. Shmukler, K. Galloway, M. Klein, Banerjee, Development and [170] G. Nikiforidis, L. Berlouis, D. Hall, D. Hodgson, Evaluation of carbon com-
testing of an economic grid-scale flow-assisted zinc/nickel-hydroxide alka- posite materials for the negative electrode in the zinc-cerium redox flow cell,
line battery, J. Power Sources 264 (2014) 49e58. J. Power Sources 206 (2012) 497e503.
[140] N.J. Magnani, R.P. Clark, J.W. Braithwaite, D.M. Bush, P.C. Butler, J.M. Freese, [171] G. Nikiforidis, L. Berlouis, D. Hall, D. Hodgson, Impact of electrolyte
K.R. Grothaus, K.D. Murphy, P.E. Shoemaker, Exploratory Battery Technology composition on the performance of the zinc-cerium redox flow battery
Development and Testing Report for 1985, Report No. SAND86e1266 UC- system, J. Power Sources 243 (2013) 693e698.
94cb, Sandia National Laboratories, 1987. [172] G. Nikiforidis, L. Berlouis, D. Hall, D. Hodgson, A study of different carbon
[141] M. Yang, H. Wu, J.R. Selman, A cycling performance model for the zinc/ composite materials for the negative half-cell reaction of the zinc cerium
ferricyanide battery, J. Chem. Ind. & Eng. (China) 4 (1989) 93e114. hybrid redox flow cell, Electrochim. Acta 113 (2013) 412e423.
[142] Y. Wen, T. Wang, J. Cheng, J. Pan, G. Cao, Y. Yang, Lead ion and tetrabuty- [173] G. Nikiforidis, R. Cartwright, L. Berlouis, D. Hall, D. Hodgson, Factors affecting
lammonium bromide as inhibitors of the growth of spongy zinc in single the performance of the Zn-Ce redox flow battery, Electrochim. Acta 140
flow zinc/nickel batteries, Electrochim. Acta 59 (2012) 64e68. (2014) 139e144.
[143] T.K. Hoang, M. Acton, H.T. Chen, Y. Huang, T.N.L. Doan, P. Chen, Sustainable [174] Z. Xie, Q. Liu, Z. Chang, X. Zhang, The developments and challenges of cerium
gel electrolyte containing Pb 2þ as corrosion inhibitor and dendrite sup- half-cell in zinc-cerium redox flow battery for energy storage, Electrochim.
pressor for the zinc anode in the rechargeable hybrid aqueous battery, Mat. Acta 90 (2013) 695e704.
Today Energy 4 (2017) 34e40. [175] F. Xiong, D. Zhou, Z. Xie, Y. Chen, A study of the Ceþ þ
3 /Ce4 redox couple in
[144] C. Iwakura, H. Murakami, S. Nohara, N. Furukawa, H. Inoue, Char- sulfamic acid for redox battery application, Appl. Energy 99 (2012) 291e296.
geedischarge characteristics of nickel/zinc battery with polymer hydrogel [176] F.C. Walsh, C. Ponce de Le on, L. Berlouis, G. Nikiforidis, L.F. Arenas-Martínez,
electrolyte, J. Power Sources 152 (2005) 291e294. D. Hodgson, D. Hall, The development of Zn-Ce hybrid redox flow batteries
[145] S. Satoshi, O. Tsukasa, I. Yutaka, Y. Takao, 'Nickel-zinc alkaline storage bat- for energy storage and their continuing challenges, ChemPlusChem. 80
tery', US Pat. 3976502, (1976). (2015) 288e311.
[146] S. Higashi, S.W. Lee, J.S. Lee, K. Takechi, Y. Cui, Avoiding short circuits from [177] D.R. Martin, Lecture demonstrations of electrochemical reactions, J. Chem.
zinc metal dendrites in anode by backside-plating configuration, Nat. Com- Educ. 25 (1948) 495e497.
mun. 7 (2016) 11801. [178] J.L. Jones, A.B. Arranaga, A new zinc-iodate primary battery, J. Electrochem.
[147] S. Achiwa, Shiga-Kenritsu Tanki Darguku Grakujutsu Zasshi (Japan) 16 Soc. 105 (1958) 435e439.
(1975) 7. [179] B. Li, J. Liu, Z. Nie, W. Wang, D. Reed, J. Liu, P. McGrail, V. Sprenkle, Metal-
[148] E. Villarreal-Dominguez, 'Lead dioxide-zinc rechargeable type cell and bat- organic frameworks as highly active electrocatalysts for high-energy density,
tery and electrolyte therefore', US Pat. 3964927, (1976). aqueous zinc-polyiodide redox flow batteries, Nano Lett. 16 (2016)
[149] J.E. Stauffer, Lead zinc Battery, EP 1555710, 20/07/2005, 2005. 4335e4340.
[150] M.S.E. Abdo, T.Z. Fahidy, Charge-discharge characteristics of a lead oxide- [180] T. Janoschka, N. Martin, U. Martin, C. Friebe, S. Morgenstern, H. Hiller,
zinc secondary battery system, J. Appl. Electrochem. 12 (1982) 225e230. M.D. Hager, U.S. Schubert, An aqueous, polymer-based redox-flow battery
[151] J.Q. Pan, Y.Z. Sun, J. Cheng, Y.H. Wen, Y.S. Yang, P.Y. Wan, Study on a new using non-corrosive, safe, and low-cost materials, Nature 527 (2015) 78e81.
single flow acid Cu-PbO2 battery, Electrochem. Commun. 10 (2008) [181] T. Janoschka, M.D. Hager, U.S. Schubert, Powering up the future: radical
1226e1229. polymers for battery applications, Adv. Mater. 24 (2012) 6397e6409.
[152] J. Pan, M. Yang, X. Jia, Y. Sun, The principle and electrochemical performance [182] S.I. Imabayashi, N. Kitamura, S. Tazuke, K. Tokuda, Substituent effects on
of a single flow Cd-PbO2 battery, J. Electrochem. Soc. 100 (2013) electrochemical reduction of viologen dimer and trimer with ethylene
A1146eA1152. spacer, J. Electroanal. Chem. 239 (1988) 397e403.
[153] R.D. Surville, M. Josefowicz, L.T. Yu, J. Perichon, R. Buvet, Electrochemical [183] S.I. Imabayashi, N. Kitamura, S. Tazuke, K. Tokuda, The role of intramolecular
chains using protolytic organic semiconductors, Electrochim. Acta 13 (1968) association in the electrochemical reduction of viologen dimers and trimers,
1451e1458. J. Electroanal. Chem. 243 (1988) 143e160.
[154] P. Novak, K. Muller, K.S.V. Santhanam, O. Haas, Electrochemically active [184] J. Winsberg, S. Muench, T. Hagemann, S. Morgenstern, T. Janoschka,
polymers for rechargeable batteries, Chem. Rev. 97 (1997) 207e281. M. Billing, F.H. Schacher, G. Hauffman, J.F. Gohy, S. Hoeppener, M.D. Hager,
[155] S.Q. Li, G.L. Zhang, G.L. Jing, J.Q. Kan, Aqueous zincepolyaniline secondary U.S. Schubert, Polymer/zinc hybrid-flow battery using block copolymer mi-
battery, J. Synthetic Metals 158 (2008) 242e245. celles featuring a TEMPO corona as catholyte, Polym. Chem. 7 (2016)
[156] M. Sima, T. Visan, M. Buda, A comparative study of zincdpolyaniline elec- 1711e1718.
trochemical cells having sulfate and chloride electrolytes, J. Power Sources [185] Y. Xu, Y.H. Wen, J. Cheng, G.P. Cao, Y.S. Yang, A study of tiron in aqueous
56 (1995) 133e136. solutions for redox flow battery application, Electrochim. Acta 55 (2010)
[157] L.W. Hruska, R.F. Savinell, Investigation of factors affecting performance of 715e720.
the iron-redox battery, J. Electrochem. Soc. 128 (1981) 18e25. [186] G.W. Heise, Air-depolarized primary battery, US Patent 1899615 (1933).
[158] R. Amirante, E. Cassone, E. Distaso, P. Tamburrano, Overview on recent de- [187] S. Smedley, X.G. Zhang, Zinc-air: Hydraulic Recharge, Encyclopedia of Elec-
velopments in energy storage: mechanical, electrochemical and hydrogen trochemical Power Sources, Elsevier, 2009, pp. 393e403.
technologies, Energy Convers. Manag. 132 (2017) 372e387. [188] S.I. Smedley, X.G. Zhang, A regenerative zinc-air fuel cell, J. Power Sources
[159] Z. Xie, Q. Su, A. Shi, B. Yang, B. Liu, J. Chen, X. Zhou, D. Cai, L. Yang, High 165 (2007) 897e904.
performance of zinc-ferrum redox flow battery with Ac-/HAc buffer solution, [189] G. Savaskan, T. Huh, J.W. Evans, Further studies of a zinc-air cell employing a
J. Energy Chem. 25 (2016) 495e499. packed bed anode part I: discharge, J. Appl. Electrochem. 22 (1992) 909e915.
[160] R.L. Clarke, B.J. Dougherty, S. Harrison, P.J. Millington, S. Mohanta, Cerium [190] E.A. Schumacher, 'Zinc-Oxygen Cells with Alkaline Electrolyte, Primary Bat-
Batteries, US 2004/0202925 A1, 2004. tery, vol. 1, John Wiley & Sons Inc., 1994, p. 265.
[161] R.L. Clarke, B.J. Dougherty, S. Harrison, J.P. Millington, S. Mohanta, Battery [191] Y. Li, H. Dai, Recent advances in zinc-air batteries, Chem. Soc. Rev. 43 (2014)
with Bifunctional Electrolyte, US 2006/0063065 A1, 2005. 5257e5275.
[162] B. Fang, S. Iwasa, Y. Wei, T. Arai, M. Kumagai, A study of the Ce (III)/Ce (IV) [192] R.P. Hamlen, T.B. Atwater, Handbook of Batteries, Chapter 38, 2002,
redox couple for redox flow battery, Electrochim. Acta 47 (2002) pp. 1210e1262.
3971e3976. [193] P.N. Ross, Feasibility study of a new zinc-air battery concept using flowing
[163] Z.P. Xie, F.J. Xiong, D. Zhou, Study of the Ce3þ/Ce4þ redox couple in mixed- alkaline electrolyte, in: IECEC '86; Proceedings of the Twenty-first Interso-
acid media (CH3SO3H and H2SO4) for redox flow battery application, En- ciety Energy Conversion Engineering Conference, vol. 2, American Chemical
ergy Fuels 25 (2011) 2399e2404. Society, San Diego, CA, 1986, pp. 1066e1072.
A. Khor et al. / Materials Today Energy 8 (2018) 80e108 107

[194] P.N. Ross, Zinc-air Design Concept for the DOE-EHP IDSEP Van, Report no. [222] Z.G. Yang, J.L. Zhang, M.C.W. Kintner-Meyer, X.C. Lu, D.W. Choi, J.P. Lemmon,
24639, Lawrence Berkeley Laboratory, 1988. J. Liu, Electrochemical energy storage for Green grid, Chem. Rev., ACS 111
[195] S. Muller, F. Holzer, O. Haas, Progress towards 20Ah/12V electrically (2011) 3577e3613.
rechargeable zinc/air battery', 192nd meeting of the electrochemical society, [223] D.L. Douglas, J.R. Birk, Secondary battery for electrical energy storage, Annu.
Paries, France, Electrochem. Soc. Proc. 97 (1977) 859e868. Rev. Energy 5 (1980) 61e88.
[196] Y.H. Wen, J. Cheng, S.Q. Ning, Y.S. Yang, Preliminary study on zinc-air battery [224] T.R. Crompton, Battery Reference Book, Chapter 14, third ed., Elsevier Sci-
using zinc regeneration electrolysis with propanol oxidation as a counter ence & Technology Books, Boston : Newnes, Oxford, England, 2000.
electrode reaction, J. Power Sources 188 (2009) 301e307. [225] R.L. Doyle, M.E.G. Lyons, Photoelectrochemical Solar Fuel Production: from
[197] Y.H. Wen, J. Cheng, L. Zhang, X. Yan, Y.S. Yang, The inhibition of the spongy Basic Principles to Advanced Devices. Chapter 2 the Oxygen Evolution Re-
electrocrystallization of zinc from doped flowing alkaline zincate solutions, action: Mechanistic Concepts and Catalyst Design, Springer, 2016.
J. Power Sources 193 (2009) 809e894. [226] G.P. Rajarathnam, A.M. Vassallo, The Zn-br Flow Battery-materials Chal-
[198] C.S. BEADLEY, 'Secondary battery', US Pat. 312802, (1885). lenges and Practical Solutions for Technology Advancement, Springer, 2015.
[199] P.C. Butler, D.W. Miller, A.E. Verardo, Flowing-electrolyte-battery testing and [227] C. Cachet, R. Wiart, Zinc deposition and passivated hydrogen evolution in
evolution, in: 17th Intersoc. Energy Conversion Eng. Conf., Los Angeles, 1982. highly acidic sulphate electrolytes: depassivation by nickel impurities,
[200] G.P. Rajarthnam, The Zinc/bromine Flow Battery: Fundamentals and Novel J. Appl. Electrochem. 20 (1990) 1009e1014.
Materials for Technology Advancement, University of Sydney, Faculty of [228] M. Paunovic, M. Schlesinger, 'Fundamentals of Electrochemical Deposition',
Engineering & Information Technologies, School of Chemical & Biomolecular See R. Winand Chapter 'Electrodeposition of Zinc and Zinc Alloys', fifth ed.,
Engineering, 2016. John Wiley & Sons, 2010, p. 285.
[201] J.D. Jeon, H.S. Yang, J. Shim, H.S. Kim, J.H. Yang, Dual function of quaternary [229] Jr. Diaddario, L. Leonard, 'Method for Improving the Macro Throwing Power
ammonium in Zn/Br redox flow battery: capturing the bromine and for Chloride Zinc Electroplating Baths', US Pat. 6143160, (2000).
lowering the charge transfer resistance, Electrochim. Acta 127 (2014) [230] A.E. Martell, NIST Critically Selected Stability Constants of Metal Complexes:
397e402. Version 8.0, National Institute of Science and Technology (US), 2014.
[202] K.J. Cathro, D. National Energy Research, D. Program, A.D.o. Resources, En- [231] R.K. Cannan, A. Kibrick, Complex formation between carboxylic acids and
ergy, Zinc-bromine Batteries for Energy Storage Applications, Department of divalent metal cations, Journal of the American Chemical Society, J. Am.
Resources and Energy, 1986. Chem. Soc. 60 (1938) 2314e2320.
[203] K.J. Cathro, K. Cedzynska, D.C. Constable, P.M. Hoobin, Selection of quater- [232] L. Li, S. Kim, W. Wang, M. Vijayakumar, Z. Nie, B. Chen, J. Zhang, G. Xia, J. Hu,
nary ammonium bromides for use in zinc/bromine cells, J. Power Sources 18 G. Graff, J. Liu, Z. Yang, A stable vanadium redox-flow battery with high
(1986) 349e370. energy density for large-scale energy storage, Adv. Energy Mater. 1 (2011)
[204] M. Wu, T. Zhao, H. Jiang, Y. Zeng, Y. Ren, High-performance zinc bromine 394e400.
flow battery via improved design of electrolyte and electrode, J. Power [233] S. Kim, M. Vijayakumar, W. Wang, J. Zhang, B. Chen, Z. Nie, F. Chen, J. Hu,
Sources 355 (2017) 62e68. Z. Yang, Chloride supporting electrolytes for all-vanadium redox flow bat-
[205] G.P. Rajarathnam, M. Schneider, X. Sun, A.M. Vassallo, The influence of teries, Phys. Chem. Chem. Phys. 13 (2011) 18186e18193.
supporting electrolytes on zinc half-cell performance in zinc/bromine flow [234] C. Tang, D. Zhou, Methanesulfonic acid solution as supporting electrolyte for
batteries, J. Electrochem. Soc. 163 (2016) A5112eA5117. zinc-vanadium redox battery, Electrochim. Acta 65 (2012) 179e184.
[206] R. Zito, 'Zinc-bromine battery with long term stability', US Pat. 4482614, [235] P. Leung, M.R. Mohamed, A.A. Shah, Q. Xu, M.B. Conde-Duran, A mixed acid
(1984). based vanadium-cerium redox flow battery with a zero-gap serpentine ar-
[207] P.C. Butler, P.A. Eidler, P.G. Grimes, S.E. Klassen, R.C. Miles, Handbook of chitecture, J. Power Sources 274 (2015) 651e658.
Batteries, by D. Linden and T. B. Reddy, Chapter 39-Zinc/Bromine Batteries, [236] S. Srinivasan, M. Pushpavanam, Role of additives in bright zinc deposition
third ed., McGraw-Hill, 2002, pp. 1e20. from cyanide free alkaline baths, J. Appl. Electrochem. 36 (2006) 315e322.
[208] G.P. Rajarathnam, M.E. Easton, M. Schneider, A.F. Masters, T. Maschmeyer, [237] D.S. Aaron, Q. Liu, Z. Tang, G.M. Grim, A.B. Papandrew, A. Turhan,
A.M. Vassallo, The influence of ionic liquid additives on zinc half-cell elec- T.A. Zawodzinski, M.M. Mench, Dramatic performance gains in vanadium
trochemical performance in zinc/bromine flow batteries, RSC Adv. 6 (2016) redox flow batteries through modified cell architecture, J. Power Sources 206
27788e27797. (2012) 450e453.
[209] J.H. Yang, H.S. Yang, H.W. Ra, J. Shim, J.D. Jeon, Effect of a surface active agent [238] Q.H. Liu, G.M. Grim, A.B. Papandrew, A. Turhan, T.A. Zawodzinski,
on performance of zinc/bromine redox flow batteries: improvement in M.M. Mench, High performance vanadium redox flow batteries with opti-
current efficiency and system stability, J. Power Sources 275 (2015) mized electrode configuration and membrane selection, J. Electrochem. Soc.
294e297. 159 (2012) A1246eA1252.
[210] D. Kim, J. Jeon, A Zn(ClO4)2 supporting material for highly reversible zinc- [239] C.A. Loto, Electrodeposition of zinc from acid based solutions: a review and
bromine electrolytes, Bull. Kor. Chem. Soc. 37 (2016) 299e304. experimental study, Asian J. Appl. Sci. 5 (2012) 314e326.
[211] M.E. Easton, P. Turner, A.F. Masters, T. Maschmeyer, Zinc bromide in aqueous [240] C.D. Iacovangelo, F.G. Will, Parametric study of zinc deposition on porous
solutions of ionic liquid bromide salts: the interplay between complexation carbon in a flowing electrolyte cell, J. Electrochem. Soc. 132 (1985)
and electrochemistry, RSC Adv. 5 (2015) 83674e83681. 851e857.
[212] Y. Munaiah, S. Dheenadayalan, P. Ragupathy, V.K. Pillai, High performance [241] K. Sato, K. Yamato, K. Iozumi, Manufacturing of one-side electrogalvanized
carbon nanotube based electrodes for zinc bromine redox flow batteries, ECS steel strip with heavy coating, Transactions ISIJ 23 (1983) 946e953.
J. Solid State Sci. Technol. 2 (2013) M3182eM3186. [242] J.F. Silva Filho, V.F.C. Lins, Crystallographic texture and morphology of an
[213] M. Mastragostino, C. Gramellini, Kinetic study of the electrochemical pro- electrodeposited zinc layer, J. Surf. Coat. Technol. 200 (2006) 2892e2899.
cesses of the bromine/bromine aqueous system on vitreous carbon elec- [243] H. Park, J.A. Szpunar, The role of texture and morphology in optimizing the
trodes, Electrochim. Acta 30 (1985) 373e380. corrosion resistance of zinc-based electrogalvanized coatings, Corrosion Sci.
[214] Y. Munaiah, S. Suresh, S. Dheenadayalan, V.K. Pillai, P. Ragupathy, Compar- 40 (1998) 525e545.
ative Electrocatalytic performance of single-walled and multiwalled carbon [244] Y.A. Naik, T.V. Venkatesha, A new condensation product for zinc plating from
nanotubes for zinc bromine redox flow batteries, J. Phys. Chem. C 118 (2014) non-cyanide alkaline bath, Bull. Mater. Sci. 28 (2005) 495e501.
14795e14804. [245] Y.A. Naik, T.V. Venkatesha, P.V. Nayak, Electrodeposition of zinc from chlo-
[215] L. Zhang, H. Zhang, Q. Lai, X. Li, Y. Cheng, Development of carbon coated ride solution, Turk. J. Chem. 26 (2002) 725e733.
membrane for zinc/bromine flow battery with high power density, J. Power [246] G. Achary, H.P. Sachin, Y. Arthoba Naik, T.V. Venkatesha, Effect of a new
Sources 227 (2013) 41e47. condensation product on electrodeposition of zinc from a non-cyanide bath,
[216] C. Wang, X. Li, X. Xi, P. Xu, Q. Lai, H. Zhang, Relationship between activity and J. Bull. Mater. Sci. 30 (2007) 219e224.
structure of carbon materials for Br2/Br in zinc bromine flow batteries, RSC [247] Y. Arthoba Naik, T.V. Venkatesha, P. Vasudeva, Electrodeposition of zinc from
Adv. 6 (2016) 40169e40174. chloride solution, Turk. J. Chem. 26 (2002) 725e733.
[217] C. Wang, X. Li, X. Xi, W. Zhou, Q. Lai, H. Zhang, Bimodal highly ordered [248] T.J. Tuaweri, E.M. Adigio, P.P. Jombo, A study of process parameters for zinc
mesostructure carbon with high activity for Br 2/Br- redox couple in electrodeposition from a sulphate bath, Int. J. Engin. Sci. Invention 2 (2013)
bromine based batteries, Nanomater. Energy 21 (2016) 217e227. 17e24.
[218] D. Desai, X. Wei, D.A. Steingart, S. Banerjee, Electrodeposition of preferen- [249] S. Afifi, A. Ebaid, M. Hegazy, K. Donya, On the electrowinning of zinc from
tially oriented zinc for flow-assisted alkaline batteries, J. Power Sources 256 alkaline zincate solutions, J. Electrochem. Soc. 138 (1991) 1929e1933.
(2014) 145e152. [250] A.L. Marshall, The electrodeposition of zinc from sulphate solutions, Trans.
[219] Y. Munaiah, P. Ragupathy, V.K. Pillai, Single-step synthesis of halogenated Faraday Soc. 21 (1925) 297e314.
graphene through electrochemical exfoliation and its utilization as elec- [251] F. Galvani, I.A. Carlos, The effect of the additive glycerol on zinc electrode-
trodes for zinc bromine redox flow battery, J. Electrochem. Soc. 163 (2016) position on steel, J. Met. Finish. 95 (1997) 70e72.
A2899eA2910. [252] S. Martin, 'Zinc Electroplating and Baths therefore Containing Carrier
[220] H.S. Yang, J.H. Park, H.W. Ra, C.S. Jin, J.H. Yang, Critical rate of electrolyte Brighteners', US Pat. 4541906, (1985).
circulation for preventing zinc dendrite formation in a zinc-bromine redox [253] D.R. Gabe, The role of hydrogen in metal electrodeposition processes, J. Appl.
flow battery, J. Power Sources 325 (2016) 446e452. Electrochem. 27 (1997) 908e915.
[221] M. Wu, T.S. Zhao, R. Zhang, H. Jiang, L. Wei, A zinc bromine flow battery with [254] L. Mirkova, G. Maurin, I. Krastev, C. Tsvetkova, Hydrogen evolution and
improved design of cell structure and electrodes, Energy Technol. 6 (2) permeation into steel during zinc electroplating; effect of organic additives,
(2017) 333e339. J. Appl. Electrochem. 31 (2001) 647e654.
108 A. Khor et al. / Materials Today Energy 8 (2018) 80e108

[255] M.H. Abd Elhamid, B.G. Ateya, K.G. Weil, H.W. Pickering, Calculation of the [268] A.M. Alfantazi, D.B. Dreisinger, The role of zinc and sulfuric acid concentra-
hydrogen surface coverage and rate constants of the hydrogen evolution tions on zinc electrowinning from industrial sulfate based electrolyte, J. Appl.
reaction from polarization, J. Electrochem. Soc. 147 (2000) 2148e2150. Electrochem. 31 (2001) 641e646.
[256] T. Zakroczymski, V. Kleshnya, J. Flis, Evolution and entry of hydrogen [269] M.S. Youssef, C.C. Koch, P.S. Fedkiw, J. Corros. Sci. 46 (2004) 51.
into iron during cathodic charging in alkaline solution with ethyl- [270] Kh. Saber, C.C. Koch, P.S. Fedkiw, Pulse current electrodeposition of nano-
enediaminetetraacetic acid, J. Electrochem. Soc. 145 (1998) 1142e1148. crystalline zinc, J. Mater. Sci. Eng. A341 (2003) 174e181.
[257] D.M. Drazic, B.E. Conway, J.O.M. Bockris, R.E. White, Modern Aspects of [271] B. Kavitha, P. Santhosh, M. Renukadevi, A. Kalpana, P. Shakkthivel,
Electrochemistry, vol. 19, Plenum Press, New York, 1989. T. Vasudevan, Role of organic additives for zinc plating, J. Surf. Coat. Technol.
[258] H. Vehoff, H. Wipf, Tropics in Applied Physics, vol. 73, Springer Verlag, Berlin, 201 (2006) 3438e3442.
1997. [272] M. Pownall, 'Back to the future with battery development', Business News
[259] J.L. Zhu, Y.H. Zhou, C.Q. Gao, Influence of surfactants on electrochemical Western Australia, https://www.businessnews.com.au/article/Back-to-the-
behavior of zinc electrodes in alkaline solution, J. Power Sources 72 (1998) future-with-battery-development, (Accessed on)
231e235. [273] 'EnergyPod® 2-Long Duration Energy Storage', Primus Power, http://
[260] C. Cachet, M. Keddam, V. Mariotte, R. Wiart, Influence of perfluorinated and primuspower.com/en/product/, (Accessed on)
hydrogenated surfactants upon hydrogen evolution on gold electrodes, [274] P. Alotto, M. Guarnieri, F. Moro, Redox flow batteries for the storage of
Electrochim. Acta 39 (1994) 2743e2750. renewable energy: a review, Renew. Sustain. Energy Rev. 29 (2014)
[261] C. Cachet, Z. Chami, R. Wiart, Electrode kinetics connected to deposit growth 325e335.
for zinc electrodeposition: influence of surfactants, Electrochim. Acta 32 [275] D. Vissers, W. DeLuca, G. Henriksen, Advanced Batteries for Electric Vehicle
(1987) 465e474. Applications', Symposium proceedings of the ACS Meeting, Illinois, Chicago,
[262] C.D. Iacovangelico, F.G. Will, J. Electrochem. Soc. 132 (1985) 851. 1993.
[263] C. Cachet, R. Wiart, Influence of a perfluorinated surfactant on the mecha- [276] P. Leung, J. Palma, E. Garcia-Quismondo, L. Sanz, M.R. Mohamed,
nism of zinc deposition in acidic electrolytes, Electrochim. Acta 44 (1999) M.A. Anderson, Evaluation of electrode materials for all-copper hybrid flow
4743e4751. batteries, J. Power Sources 310 (2016) 1e11.
[264] C.W. Lee, K. Sathiyanarayanan, S.W. Eom, H.S. Kim, M.S. Yun, Novel elec- [277] H.R. Jiang, W. Shyy, M. Liu, Y.X. Ren, T.S. Zhao, Borophene and defective
trochemical behavior of zinc anodes in zinc/air batteries in the presence of borophene as potential anchoring materials for lithium-sulfur batteries: a
additives, J. Power Sources 159 (2006) 1474e1477. first-principles study, J. Mater. Chem. A 6 (2018) 2107e2114.
[265] F. Ganne, C. Cachet, G. Maurin, R. Wiart, E. Chauveau, J. Petitjean, Impedance [278] H.R. Jiang, M.C. Wu, Y.X. Ren, W. Shyy, T.S. Zhao, Towards a uniform distri-
spectroscopy and modelling of zinc deposition in chloride electrolyte con- bution of zinc in the negative electrode for zinc bromine flow batteries, Appl.
taining a commercial additive, J. Appl. Electrochem. 30 (2000) 665e673. Energy 213 (2018) 366e374.
[266] R. Ichino, C. Cachet, R. Wiart, Mechanism of zinc electrodeposition in [279] Q. Xu, Y.N. Ji, L.Y. Qin, P.K. Leung, F. Qiao, Y.S. Li, H.N. Su, Evaluation of redox
acidic sulfate electrolytes containing Pb2þ ions, Electrochim. Acta 41 flow batteries goes beyond round-trip efficiency: a technical review,
(1996) 1031e1039. J. Energy Storage 16 (2018) 108e115.
[267] S. Thomson, N. Birbilis, M.S. Venkatraman, I.S. Cole, Corrosion of zinc as a
function of pH, Corrosion Science Edition 68 (1) (2012) 1e9.

You might also like