Cao 2018
Cao 2018
Review
a r t i c l e i n f o a b s t r a c t
Article history: As one of the most important components of the vanadium redox flow battery (VRFB), the electrolyte can
Received 18 December 2017 impose a significant impact on cell properties, performance and capital cost. In particular, the electrolyte
Revised 10 April 2018
composition will influence energy density, operating temperature range and the practical applications of
Accepted 11 April 2018
the VRFB. Various approaches to increase the energy density and operating temperature range have been
Available online xxx
proposed. The presence of electrolyte impurities, or the addition of a small amount of other chemical
Keywords: species into the vanadium solution can alter the stability of the electrolyte and influence cell perfor-
Vanadium redox flow battery mance, operating temperature range, energy density, electrochemical kinetics and cost effectiveness. This
Electrolyte additive review provides a detailed overview of research on electrolyte additives including stabilizing agents, im-
Precipitation inhibitor mobilizing agents, kinetic enhancers, as well as electrolyte impurities and chemical reductants that can
Stabilizing agent be used for different purposes in the VRFBs.
Kinetic enhancer
© 2018 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published
Impurity
Immobilizing agents
by Elsevier B.V. and Science Press. All rights reserved.
Reducing agent
Liuyue Cao received her Bachelor’s degree (2011) and Chris Menictas has been actively involved in the energy
Master’s degree (2014) in Materials Science and Engineer- efficiency sector for over 20 years. His research interests
ing from Hunan University, China. She is currently a Ph.D. include: energy storage; flow battery and fuel cell sys-
candidate at the University of New South Wales, Sydney tems; energy harvesting; temperature shifting devices for
where she is working on vanadium redox flow batteries, bio-medical applications, refrigeration and air condition-
with a focus on membrane and electrode materials. ing efficiency optimisation; and thermal morphing. Dr.
Menictas is Head of the Energy Storage and Refrigeration
Research Group in the School of Mechanical and Manu-
facturing Engineering, UNSW Sydney Australia.
∗
Corresponding author at: School of Chemical Engineering, UNSW Sydney, 2052
NSW, Australia.
E-mail address: [email protected] (M. Skyllas-Kazacos).
https://doi.org/10.1016/j.jechem.2018.04.007
2095-4956/© 2018 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by Elsevier B.V. and Science Press. All rights reserved.
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
1. Introduction cost, thus becoming critical to the economics of the VRFB. Vana-
dium pentoxide, V2 O5 , a comparatively cheaper vanadium com-
The vanadium redox flow battery (VRFBs) pioneered at the Uni- pound, is a typical raw material used to prepare the vanadium
versity of New South Wales, Sydney (UNSW) in 1980s [1,2] is electrolyte. It is commercially available in a range of grades and
presently attracting increasing attention and commercial interest purity levels. Although no standard for electrolyte specifications
in both on- and off-grid energy storage applications including wind has been set in the market, an electrolyte of high purity is always
and solar energy storage, load-levelling, peak shaving, back-up favoured by customers to avoid possible negative effects on perfor-
power supply and power arbitrage. By utilization of the V(V)/V(IV) mance, and this may result in a higher cost for the electrolyte so-
and V(III)/V(II) redox couples reactions in the catholyte and anolyte lution. In addition, facile and economic reduction methods are re-
respectively, the VRFB can successfully avoid cross contamination quired to reduce the V(V) solution to lower valence states in both
issues observed in other RFB chemistries that employ different el- preparation and rebalancing processes in order to realize the in-
ements in each half-cell. definite cyclic life of the electrolyte. The choice of chemical reduc-
Fig. 1 depicts the typical structure of a VRFB. Two pumps are tants used could potentially introduce impurities that could impact
used to circulate the vanadium electrolytes from the tanks to the on the long-term properties of the electrolyte.
surface of the electrodes in the cell stack where the electron trans- As discussed above, the current state-of-the-art electrolyte
fer reactions take place. During the charging process, VO2+ ions in composition for the VRFB has many benefits, but has a relatively
the positive half-cell are oxidized to VO2 + while V3+ ions in the low energy density and operating temperature range compared
negative half-cell are reduced to V2+ ions. During the discharge with other battery technologies. The use of high purity vanadium
process, reactions occur in the reverse direction. Differing from raw materials in electrolyte production can also increase the cost
solid-state batteries, the electrolytes are not only ionic conductors, of the VRFB. One economically efficient approach to address these
but also the cell reactants and products. Therefore, the concentra- problems is to employ lower purity vanadium as raw material for
tion of vanadium in the electrolyte determines the energy density electrolyte production and to incorporate additives in the elec-
of the system. Though a high energy density is not the main pri- trolyte to modify the characteristics while keeping the bulk prop-
ority for stationary energy storage systems, it is desirable in terms erties the same. This topic has been receiving considerable atten-
of reducing footprint and tank manufacturing costs. tion in recent years, however no review article has been previ-
Early UNSW studies showed that V2+ concentrations up to 4 M ously published on the specific subject of vanadium electrolyte ad-
could be obtained in HCl [3,4], however, the risk of HCl vapour ditives. This review starts with a brief introduction to the vana-
emissions during operation or at elevated temperatures hindered dium electrolyte chemistry and vanadium salt solubilities, and then
its further development at UNSW. Sulphuric acid, instead, can pro- categorizes and discusses the various electrolyte additives studied
vide reasonable reactant concentration (2 M) and open circuit volt- for the VRFB, on the basis of their function including precipitation
age (1.4 V) without hazardous gas emissions. However, the solubil- inhibitors, immobilizing agents, kinetic enhancers, electrolyte im-
ity of V(V) ions in sulphuric acid solution increases with total sul- purities and chemical reductants. A brief description of electrolyte
phate concentration and decreases with temperature while that of additives used in other vanadium-related RFB systems is also in-
V(IV), V(III) and V(II) show exactly the opposite trends. Therefore, a cluded. Finally, we probe into the future directions and perspec-
composition of 1.5–2 M vanadium in 3–5 M sulphuric acid solution tives of vanadium electrolyte research.
is normally recommended to achieve adequate energy density and
thermal stability for vanadium ions of all valance states. Several 2. Understanding vanadium chemistry and stability
companies including Sumitomo, Rongke Power, Imergy (previously
named Deeya), Glex (previously named Gildemeister), RedT etc, are In the early research studies of the VRFB by Skyllas-Kazacos and
using the vanadium sulphuric electrolyte with such compositions. co-workers at UNSW Sydney, the vanadium electrolytes were re-
One critical factor for the VRFB is that efficient cell perfor- ported to have excellent stability in the temperature range of −5
mance is dependent on the reaction kinetics of the V(V)/V(IV) and to 20 °C for up to a year. Moreover, neither decomposition nor pre-
V(III)/V(II) couples. Several reviews have summarized the methods cipitation was observed during continual charge–discharge testing
used to improve the reaction kinetics with most focussing on elec- even at temperatures up to 60 °C, although the solution composi-
trode modifications, however, only limited enhancement has been tions and state-of-charge range used at each temperature were not
achieved so far [5–9]. disclosed. Further studies did however show that the fully charged
Another key point is the capital cost associated with the elec- positive electrolyte of 2 M V(V) in 2–3 M H2 SO4 was inclined to
trolyte, especially when the VRFB system needs to meet high en- gradually precipitate when stored at high temperatures without
ergy capacity requirements. Typically, for energy to power ratios charge–discharge cycling [10]. Precipitation problems and possible
greater than 4 (or energy storage capacities greater than 4 h), the consequences of flow restrictions or blockages in the operation of
electrolyte contributes a high percentage of the per kWh capital the cell thus became the focus of further detailed investigation. Al-
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Fig. 2. Solubilities of VSO4 , V2 (SO4 )3 and VOSO4 salts as a function of temperature and initial sulphuric acid concentration (based on data from Min Chen MSc thesis, UNSW
1991).
though there have been numerous studies on the chemistries and As illustrated in Fig. 2, the solubilities of VSO4 , V2 (SO4 )3 and
solubilities of different valent vanadium ions in various aqueous VOSO4 increase with increasing temperature, but decrease with
solutions since the 1930s [11–25], results relating to the composi- increasing sulphuric acid concentration as a result of the com-
tion of vanadium species in solution and the equilibrium solubili- mon ion effect associated with the sulphate ions. In contrast how-
ties are rather controversial. These discrepancies are not only due ever, V(V) does not precipitate as a sulphate salt, but under-
to variations in the solution compositions used but also to the lim- goes thermal precipitation according to the following endothermic
ited duration of the precipitation experiments. reaction:
The UNSW group thus conducted a comprehensive study of the
solubilities of the vanadium sulphate salts of each of the V(II), 2VO2 + + H2 O ↔ V2 O5 + 2H+ (1)
V(III) and V(IV) oxidation states in sulphuric acid at various initial
acid concentrations and temperatures and the results are presented This equilibrium shifts to the left with increasing proton con-
in Fig. 2. centration, but shifts to the right with increasing temperature,
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
leading to thermal precipitation of V(V). For example, V(V) con- method [32,33]:
centrations up to 5 M were reported in 7–8 M H2 SO4 [26,27].
−dCt /dt = k(Ct − Ce )n (2)
For the vanadium electrolyte in sulphuric acid therefore, it is
generally found that the dissolution of V(V) ions can be improved The concentration of the total vanadium ions can be deter-
by increasing the sulphuric acid or decreasing the temperature mined by ICP (Inductively coupled plasma mass spectrometry)
while V(IV), V(III) and V(II) ions can be stabilized by elevating the analysis while the concentration of the vanadium ions at differ-
temperature and reducing the sulphuric acid [10,26–33]. These op- ent valent states can be measured by potentiometric titration. The
posing trends in temperature and sulphuric acid concentration for equilibrium vanadium concentration refers to the final stabilized
V(V) relative to those of V(II), V(III) and V(IV) species, mean that vanadium concentration at infinite time, however this is often
practical vanadium battery electrolyte compositions are generally measured after 80 0–10 0 0 h in most precipitation tests. In many
maintained between 1.6 and 2 M vanadium in 4–5 M total sulphate. cases true equilibrium may not be established during these time-
Generally, a 2 M vanadium electrolyte can be employed within the frames and this can lead to errors in calculated precipitation rate
temperature range of 10 to 40 °C where the battery undergoes constants.
regular charge–discharge cycling. For hot climates, the stability of
the V(V) is the dominant factor, so a higher acid concentration 2.2. Precipitation of vanadium ions
would be recommended; while in cold climates, V(V) stability is
not the main concern, so lower acid concentrations would enhance 2.2.1. Pentavalent vanadium ions
the solubilities of the V(II), V(III) and V(IV) species at low tem- The Raman spectroscopy results of V(V) in sulphuric acid by
peratures. On the other hand for a broader operating temperature Kausar and co-workers [34] showed that V(V) exists as VO2 SO4 − ,
range, the vanadium concentration is usually reduced to approxi- VO2 (SO4 )2 3− , VO2 (HSO4 )2 − , VO3 − , dimers with V2 O4 2+ and
mately 1.6 M. While these compositions provide reasonable energy V2 O3 4+ units in highly acidic solutions and V2 O3 4+ , VO2 (SO4 )2 3-
densities for most stationary energy storage applications, higher as well as their copolymer species at higher SO4 2− /HSO4 − concen-
energy densities are desirable in urban areas where a low footprint trations. However, using methods including ESR, static solid state
and volume is required due to limited space or high cost of real 51 V NMR, XRD and TEM, Lu [35] found that monomer species of
estate. VO2 + ions were the main species in 2 M V (V) and 5 M H2 SO4 solu-
When selecting an electrolyte composition for the VRFB, the tion and those monomers can polymerise into V2 O5 gradually with
vanadium and sulphate ion concentrations are thus chosen to en- the increasing temperature and time.
sure that each of the vanadium ions remain within their respective Considering the dissociation reactions of sulphuric acid,
solubility levels for the required operating temperature range. Pre-
cipitation will however occur when the temperature drops below H2 SO4 → H+ + HSO4 − (2)
the pre-set value in the case of the V(II), V(III) and V(IV) solutions
or exceeds the upper limit in the case of V(V). In these instances, HSO4 − ↔H+ + SO4 2− (3)
the solutions become supersaturated and precipitation may occur,
the onset and rate of which will depend upon on a number of fac- and combining the findings mentioned above, it was believed that
tors. the possible interactions occurring in the solution are [36]:
The understanding of the vanadium precipitation mechanisms
is therefore of critical importance for optimization of the vanadium VO3 − + 2H+ ↔ VO2 + + H2 O (4)
electrolyte composition and improvement of the VRFB electrolyte
stability. 2VO2 + + H2 O ↔ V2 O5 + 2H+ (5)
Although there are different theories of the precipitation mech- VO2 + + SO4 2− ↔ VO2 SO4 − (7)
anism, it is widely accepted that precipitation occurs via a two-
step mechanism which starts with the nucleation process followed VO2 + + VO2 SO4 − ↔ (VO2 )2 SO4 (8)
by crystal growth. Nucleation involves the formation of the first
stable centres of the new (solid) phase that are necessary for pre- The interaction between V(V) ions and sulphuric acid was also
cipitation to proceed. These nuclei could be clusters of the solute confirmed by Sepehr and Paddison [37] who found that VO2 + does
molecules (homogeneous nucleation) or particles of an undissolved not interact with HSO4 − but complexes with SO4 2− which is prone
impurity (heterogeneous nucleation) present in solution. The rate to deprotonate and form hydroxyl and then precipitate. This can
of nucleation in supersaturated solutions is controlled by the rate also explain the stability of V(V) ions in H2 SO4 solution which in-
at which a cluster reaches the critical size and may be correlated creases with H2 SO4 concentration and decreases with temperature
to the degree of supersaturation. [10,26,27]. Sulphuric acid is a diprotic acid and the first dissocia-
The crystal growth process is characterised by a significant tion reaction (2) is usually complete while the second reaction (3)
change in the concentration of the solute accompanied by a fast is much weaker. Therefore, it is reasonable that H2 SO4 of higher
increase in crystal mass and volume. The first step in this mecha- concentration can dissociate more H+ ions thus restraining the
nism is the diffusion of the solute particles to the crystal-solution precipitation reaction by shifting reaction (5) to the opposite direc-
interface which is followed by adsorption of these particles at the tion. Among all the three complexing reactions between V(V) and
crystal surface. The third step is the integration of the adsorbed the dissociated sulphuric acid species, reaction (7) has the highest
solute into the crystal lattice and the fourth step is the opposite equilibrium constant, 9.32 ± 0.43 [36] and thus plays a predomi-
dissolution process. Accordingly, the rate of the crystal growth is nant role. Therefore, higher H2 SO4 dissociates more HSO4 − , which
proposed to be either controlled by the rate of diffusion or by the can shift the reaction (6) to produce more VO2 SO4 − rather than
rate of the surface reaction. forming the V2 O5 crystals.
The precipitation reaction order n and rate constant k can However, the precipitation mechanism of V(V) ions proposed
be determined with the vanadium concentration (Ct ) at time t by Vijayakumar and co-workers [29,30] from Pacific Northwest
and equilibrium vanadium concentration (Ce ) based on differential National Laboratory (PNNL) is slightly different. Through 17 O and
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
51V NMR and density functional theory (DFT)-based computational mechanisms were discovered [32,33,43] for the precipitation of
methods, they found that V(V) ions are presented as [VO2 (H2 O)3 ]+ V(III) sulphate. Stagnant solutions with low supersaturation fol-
and change into H3 VO4 before eventually turning to V2 O5 : low a first-order precipitation rate law and the activation energy
was found to be relatively low (e.g. 10.7 ± 1.1 kJ mol−1 for 2 M
[VO2 (H2 O)3 ]+ ↔ VO(OH)3 + H3 O+ (9) V(III) in 5.5 M total sulphate/bisulphate solution at 20 °C), indicat-
ing a diffusion-controlled mechanism. However, the stagnant so-
2H3 VO4 ↔ V2 O5 + 3H2 O (10) lution with relatively high S/V ratio was found to experience two
stages during the whole precipitation process: the first 30% fol-
Rahman [27] and Kausar [38] further investigated the kinetics lowed the second order kinetics while the rest obeyed the first-
of V(V) precipitation in stagnant and stirring states, respectively. In order law. As the first and second order rate law indicated the
the static mode, the precipitation rates in low supersaturated so- diffusion- and reaction-controlled mechanism respectively, it was
lutions follow a first order law and the activation energy is around believed the chemical composition altered during the precipitation
9.0 kJ mol−1 indicating diffusion control. The highly supersaturated process. A remarkable reduction of the V(III) equilibrium concen-
vanadium solution however, obeys a second order precipitation tration further suggested this. Depending on the sulphuric acid ra-
rate law and has an activation energy of 77.2 kJ mol−1 . It was found tio, V(III) precipitated from H2 SO4 in the form of simply V2 (SO4 )3
that agitation can accelerate the mobility of V(V) species and re- or V2 (SO4 )3 •xH2 SO4 •yH2 O. The results also showed that transport
duce the distance between the ions, so the precipitation reaction of the precipitating species to the nuclei is the determining step.
is no longer diffusion controlled. Especially at higher supersatura- The rate of precipitation in stirred V(III) sulphate solution was also
tion, the stirred solution obeys the second order law and is limited found to be seven times faster than in the stagnant solution, ex-
by surface reaction control. hibiting a greater activation energy of around 41 kJ mol−1 . The sur-
face reaction turns to be the determining step as a faster feeding
2.2.2. Tetravalent vanadium ions of the V(III) ions or sulphate to the crystal nuclei site due to the
In contrast to the V(V), the solubility of V(IV) sulphate stirring.
[27,28,31] has been found to deteriorate with increasing sulphuric
acid concentration and reduced temperature. In addition, the rate 2.2.4. Bivalent vanadium ions
of decline in solubility is larger when the concentration of H2 SO4 The trends of V(II) sulphate solubility are analogous to those
is in the range of 0–7 M but smaller at 8 or 9 M H2 SO4 solution of V(IV) and V(III) ions. Mousa and Skyllas-Kazacos also investi-
[28]. The influence of temperature is also more significant in the gated the precipitation kinetics of V(II) in sulphuric acid solutions
lower sulphuric acid concentration range (0–7 M) and less in the [32,33]. All V(II) sulphate solutions precipitated according to the
highly concentrated sulphuric acid (8 or 9 M) [28]. Skyllas-Kazacos first order rate law and reached high equilibrium concentration
and co-workers assumed that the interactions between V(IV) ions values with relatively lower activation energies compared to V(III)
and sulphuric acid proposed by Ivakin and Voronova [39] were the sulphate solutions. The precipitation of V(II) species in sulphuric
possible explanation: acid was a diffusion-controlled process under both stagnant and
stirring conditions, which was attributed to the possibility that
VOSO4 •xH2 O ↔ VO2 + + SO4 2 − + xH2 O (11)
fewer ions are required to form the VSO4 precipitate compared to
V2 (SO4 )3 .
VO2 + + SO4 2− ↔ VOSO4 (12)
VO2 + + 2SO4 2− ↔ VO(SO4 )2 2− (13) 2.2.5. Mixed valent electrolyte in practical VRFBs
The precipitation dynamics in an operating VRFB system might
VO2 + + HSO4 − ↔ VOHSO4 + (14) be different from the results in beaker tests of vanadium ions at
each valance state due to continuous state-of-charge (SOC) varia-
When the concentration of H2 SO4 is relatively low (0–7 M), tions, potential temperature fluctuations, the effects of precipita-
the second dissociation of H2 SO4 occurs more readily so that tion seeding by impurities and additional electrolyte flowing con-
the higher H2 SO4 can lead to more SO4 2− ions, shifting reaction ditions.
(11) towards the precipitation direction. However, in the 8 or 9 M It is undeniably challenging to analyse the vanadium electrolyte
H2 SO4 solution, the second dissociation of H2 SO4 becomes more in-situ with or without additives in the flowing system. However,
difficult [40,41] thus resulting in a slower rate of decrease in sol- Mousa and Skyllas-Kazacos [33] designed a series of experiments
ubility. The varied effect of temperature can also be explained by with mixed V(II)/V(III) electrolytes to simulate the negative half-
the dependence of the sulphuric acid dissociation constant on tem- cell electrolyte at different SOCs. Both stagnant and stirring condi-
perature [41]. tions were applied to investigate the effect of flowing electrolyte.
Through 1 H and 17 O NMR methods and theoretical calculation By using the differential methods, the precipitation mechanism
based on DFT, researchers at PNNL [30,42] held the view that V(IV) of the negative half-cell electrolyte was investigated. The results
ions exist in the form of [VO(H2 O)5 ]2+ and are weakly bonded by showed that the precipitation of the negative half-cell electrolyte
SO4 2− or HSO4 − at the second coordination sphere. The hydrated in the temperature range of 1–20 °C [33] followed the first or-
vanadyl ions can remain stable up to 3 M at a temperature range der reaction law regardless of the movement of electrolyte. Fur-
of –33 to 66 °C. But with increasing sulphuric acid concentration, thermore, both the stagnant and stirred solution co-precipitation
more secondary solvation shells will be occupied by the sulphate tests under different temperatures showed similar results to the
or bisulphate anions, thus hindering the reorientation of V(IV) ions. reported single V(II) or V(III) in sulphuric acid solutions (as dis-
A decrease in temperature also leads to reduced mobility of V(IV), cussed previously), which suggests the independence of the crystal
contributing to precipitation. growth of the precipitates of V(II) or V(III) ions.
However, in the study of the positive half-cell electrolyte by
2.2.3. Trivalent vanadium ions Kazacos et al. [10], the presence of V(IV) was shown to inter-
The solubility of V(III) sulphate follows similar trends to V(IV), act with the precipitation of V(V) species. It is speculated that
decreasing with reducing temperature and increasing sulphuric the complex V(IV) species containing both sulphate and bisul-
acid concentration because of the common ion effect of the sul- phate ions which can lead to the several equilibria in addition to
phate ions. Depending on the degree of supersaturation, two the pure V(V) precipitation reactions (Eqs. (4)–(7)), cause the V(V)
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
dissolution-precipitation processes to become more complicated in by adsorbing on the crystal surface to form a monomolecular layer
mixed V(IV)/V(V) electrolytes. that changes the crystal habit and growth rate. Since the amount
Despite the different co-precipitation behaviours between pos- in solution is sometimes not sufficient to form the monolayer how-
itive and negative half-cell solutions, they still follow the general ever, it is assumed that the additive molecules are adsorbed on
trend that V(V) tends to precipitate at elevated temperatures and certain edges or corners that play a special part in the crystal
low total sulphate concentrations while V(IV), V(III) and V(II) ions growth. Consequently, a small of amount of the additive could be
possess better stability with increasing temperature and lower to- enough to block most of these “active sites” causing a significant
tal sulphate concentrations. Depending on the vanadium species, decrease in crystal growth rate.
the flowing electrolyte in practical VRFB can impose a slight effect
on the precipitation mechanism such as in the case of V(II) or com- 3. Electrolyte additives and impurities in the VRFB
pletely alter the precipitation from diffusion-controlled mechanism
to surface reaction-controlled mechanism. 3.1. Stabilizing agents
In practical operating VRFB systems, unequal permeation rates
of vanadium ions across the membrane can lead to the build-up of Given the opposing trends in precipitation for each of the four
vanadium ions in one half-cell while the heat generated from the vanadium ions in the VRFB electrolyte, a practical composition of
consequent self-discharge reactions can elevate the temperature 2 M vanadium in 5 M total sulphate/bisulphate was originally rec-
within the cell especially during standby periods when the pumps ommended by Skyllas-Kazacos and co-workers for the VRFBs for
are off. Solvent transfer across the membrane during the charge– operation within the temperature range of 15–40 °C V(V) could
discharge reactions can also alter the concentration of the vana- however precipitate at temperatures higher than 40 °C, while V(II)
dium ions in each half-cell, while the transfer of sulphate and/or and V(III) could crystalize if the electrolyte remains below 15 °C
bisulphate ions in the case of anion exchange membranes can also for extended periods. A wider operating temperature range is nor-
lead to decreased solubilities of the V(II) and V(III) sulphates in the mally achieved by either decreasing the vanadium concentration or
negative half-cell electrolyte due to the common ion effect. There- narrowing the operating state of charge (SOC) range of the VRFB.
fore, special attention should be paid to the diffusion behaviour of However, both methods result in a reduced practical energy den-
both the vanadium and bisulphate ions as well as the solvent es- sity. Although energy density is not a critical criterion for VRFBs in
pecially when employing electrolyte with high vanadium concen- most large-scale stationary energy storage applications, higher en-
trations to avoid precipitations during operation. ergy density can always help to reduce the size of tanks and offer
more marketability in urban areas. Approaches to improve the sta-
2.3. Precipitation inhibitors bility of vanadium electrolyte over a wider temperature range and
even at higher concentrations are thus desirable.
Although the exact mechanism through which additives en- Skyllas-Kazacos and co-workers first proposed the term “sta-
hance the stability of supersaturated solutions is not fully under- bilising agent” in VRFB research in the mid-1990s. This refers to
stood, additives can be classified into three categories: dispersion, compounds that can be introduced to the vanadium electrolyte
complexing and threshold agents [33]. to inhibit precipitation when the electrolyte becomes supersatu-
rated at high or low temperatures and allow an increased prac-
2.3.1. Dispersion agents tical vanadium electrolyte concentration and/or greater operating
Dispersion agents are chemicals that decrease the strength of temperature range. The stabilising additives do not increase sol-
the forces of attraction between particles, thereby reducing the ubility, but work by increasing induction time for precipitation.
chance of agglomeration. Dispersion agents are believed to create They can therefore reduce or inhibit the precipitation of vanadium
similar charges on the surface of the particles so that their ten- ions and may also allow the preparation of stable highly super-
dency to form larger particles is reduced. Dispersion agents are saturated vanadium solutions. Inspired by various chemical tech-
usually low molecular weight polymers like lignin and its deriva- nologies including the Bayer process (in which organic impurities
tives are known to inhibit precipitation of alumina trihydrate) and scale
control techniques, they achieved great success in stabilizing vana-
2.3.2. Complexing agents dium electrolytes under highly supersaturated conditions. A wide
Complexing agents are chemicals that are able to form com- variety of additives were screened and several patents on the sta-
plexes with one of the ions involved in the precipitation, thus re- bilizing agents [44,45] and high energy density VRFBs [46,47] were
ducing the activity of the ion and enhancing the stability of the filed in the mid-1990s. An extensive screening study conducted at
solution. Several chemical compounds such as EDTA have been re- UNSW Sydney in the mid- to late-1990s, gave rise to the identi-
ported to reduce scale formation via a complex formation mech- fication of the following additive groups as potential precipitation
anism. Different chelating organic acids, polyelectrolytes, and long inhibitors for one or more of the vanadium ions in the VRFB elec-
chain agents have been used as efficient crystal growth inhibitors trolyte:
for silver, copper, lead, and aluminium salts at different degrees of
supersaturation. Several types of complexing agents have also been (a) surfactants selected from the group consisting of anionic,
used for stabilising calcium sulphate solutions. cationic, amphoteric and non-ionic surfactants;
(b) straight chain, branched chain or cyclic compounds with
2.3.3. Threshold agents polyfunctional groups selected from the group consisting of
While complexing agents are used on a stoichiometric basis polyhydric alcohols, polyamines, polymercapto compounds,
to produce a more soluble species than the precipitating salt, the and organic compounds containing at least two moieties,
main feature of threshold agents is their ability to inhibit precip- each selected from the group consisting of –OH, –NH2 and
itation of certain compounds on a non-stoichiometric basis. Sev- –SH;
eral chemicals have been shown to prevent scale formation when (c) phosphoric acids selected from the group consisting of in-
added to the solution in very small quantities. For example, addi- organic phosphoric acids, organic phosphoric acids, organic
tion of 2 ppm sodium hexametaphosphate is sufficient to prevent polyphosphoric acids and inorganic polyphosphoric acids;
scale formation in calcium bicarbonate solution with a hardness (d) ammonium compounds;
of 200 ppm. Threshold agents decrease the rate of crystal growth (e) phosphonates and phosphoric acid compounds;
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
(f) carbohydrates selected from the group consisting of saccha- ing –PO3 H2 or –PO3 2− were found to be effective in the inhibition
rides, polysaccharides and starches; of concentrated V(V) ions even at elevated temperature. They be-
(g) carboxylic acids selected from the group consisting of lieved that the high level of anionic functionality of such groups
polyaminocarboxylic acids and polyhydroxy-polycarboxylic may be an important factor in sequestrating the cations and in-
acids and hibiting the hydrolysis. The additives can function in three ways:
(h) oxo-acids (i) chelate directly with the lattice ions to form such as (VO2 )3 PO4
in the electrolyte, (ii) adsorb onto the precipitating ion thus pre-
The UNSW researchers proposed that these compounds contain venting the crystal nucleation or (iii) attach to the growing crystals
functional groups that could adsorb onto the surface of the nu- leading to the distortion of further crystal formation. Considering
cleation sites, thereby inhibiting crystal growth and increasing in- V(V) ions will be reduced to V(IV) ions during the VRFB operation,
duction time for precipitation in highly supersaturated vanadium an evaluation of precipitation inhibitor additives for concentrated
solutions. Some organic additives such as compounds with –COOH, V(IV) ions at 20 °C was also carried out. The result showed that
–COO–, –COH, –CO–, –OH, –SH, –NH2 , functional groups were ini- K3 PO4 and (NaPO3 )6 (sodium hexametaphosphate, SHMP) were
tially proposed as stabilizing agents for the negative vanadium good stabilizers for both V(V) and V(IV) ions. A further investiga-
half-cell electrolyte, however many of these were found to slowly tion of these two additives was undertaken not only to evaluate
reduce V(V) to the more soluble V(IV) ions in the positive half- the stabilizing performance of different formulations, but also to
cell. As illustrated in Table 1, compounds such as glycerol, inos- study the electrochemical response, while taking the negative elec-
itol and ammonium oxalate were found to undergo oxidation by trolyte stability into consideration. The mixture of K3 PO4 (1 wt%)
V(V), forming CO2 and leading to gassing and restricted electrolyte and SHMP (1 wt%) was identified to be the best additive for both
flow or electrolyte crossover through the membrane leading to ad- concentrated positive and negative vanadium electrolytes. In addi-
ditional capacity loss [48]. Organic species with reducing function- tion, a solution with 3.5 M V in 5.7 M total SO4 2− /HSO4 − with this
ality are therefore unsuitable options for long-term stabilization additive was recommended for a high energy density VRFB in the
of VRFB electrolytes. A wide range of inorganic compounds was temperature range of 5–40 °C.
therefore screened by the UNSW group and several groups of com- Inorganic additives such as sodium or potassium phosphate
pounds were found to provide excellent stabilisation of the V(V) and sulphate salts, phosphoric acid and SHMP [50] were also fur-
electrolyte at elevated temperatures without oxidation of the addi- ther investigated for the positive half-cell electrolyte of the VRFB
tive and reduction of V(V) to V(IV), as summarised in Table 1. at 40 °C. Amongst these, the combination of 1% H3 PO4 and 2%
For any additive to be practical for use in the VRFB however, (NH4 )2 SO4 offered the longest induction time and lowest precipi-
it must also be effective, or at least not have any detrimental ef- tation rate. The increase of H+ and SO4 2− ions in the V(V) solution
fects in solutions of all of the vanadium oxidation states. This is with these additives was attributed to the shift in the precipita-
especially true since the two half-cell electrolytes are frequently tion and ion pairing reactions respectively. Other phosphates with
remixed to restore capacity during long-term operation of the tetrahedral or ring-shaped structures and (NH4 )2 SO4 all demon-
VRFB. Additive screening was therefore also conducted in solu- strated some improvement in increasing the induction time for
tions of V(IV), V(III) and V(II) and the results are summarised in precipitation. Through a detailed comparison, Skyllas-Kazacos and
Tables 2–4, respectively. The molecular structures of some organic co-workers discovered a series of inorganic additives combined by
additives are given in Fig. 3. tetrahedral or ring-shaped phosphates with NH4 + , K+ , Na+ or H+
Amongst the compounds screened by the UNSW group, phos- and ammonium sulphate for stabilising a 1.8–2 M VRFB electrolyte
phoric acid and ammonium phosphate were selected as the best at high temperatures. They proposed that polyphosphate functions
additives for the VRFB, with the phosphate ions enhancing the sta- by either complexing with VO2 + to obtain (VO2 )3 PO4 or by dis-
bility of the V(V) species at elevated temperatures, and ammo- turbing the growth of V2 O5 nuclei; Furthermore, they suggested
nium ions acting in the negative half-cell solution at low temper- that NH4 + can work better than K+ because of the similar sp3 -
atures. This work led to the addition of phosphoric acid as sta- hybrid molecular structure of VO3 − while H+ can shift the thermal
bilising agent to the standard sulphuric acid electrolyte by several precipitation reaction of V(V) and SO4 2− can stabilise V(V) by ion
VRFB producers in the late 1990s, but otherwise this research was pairing.
largely ignored until very recently when several papers on the ad- By recording the induction time with continuous visual ob-
ditives were published by Skyllas-Kazacos and co-workers. servation of the precipitation status, Mousa and Skyllas-Kazacos
The first journal article on VRFB stabilizing agents was pub- [62] conducted a comprehensive screening of stabilizing additives
lished by Skyllas-Kazacos and co-workers for supersaturated for anolyte of VRFBs at low temperature. Among the added in-
vanadyl electrolytes [31]. The compounds (NaPO3 )6 (sodium hex- organic acids and salts, (NH4 )3 PO4 , (NH4 )2 SO4 , WO5 (H2 O) and
ametaphosphate, SHMP), urea, K2 SO4 and Li2 SO4 were investigated Na2 WO4 prolonged the 2 M V(II) precipitation time from 16 h to
as precipitation inhibitors for 4 M VOSO4 in 3 M H2 SO4 at 4 °C. more than 34 h in 5.0 M total sulphate solution. However, the lat-
They evaluated the precipitation rate by periodically measuring the ter two additives appeared to change the colour of the solution,
concentration of vanadium in the V(IV) solution with and without which indicated the oxidation of V(II) ions. A subsequent titration
additives using atomic absorption spectroscopy (AAS) and visually measurement also confirmed this. (NH4 )3 PO4 and (NH4 )2 SO4 also
monitoring the appearance of any precipitation and colour change. exhibited excellent inhibiting effect in the V(III) solution. In addi-
The addition of Li2 SO4 did not show obvious enhancement but a tion, sodium salts surprisingly presented good stabilizing ability for
negative effect was observed instead. In contrast, the remaining V(III) ions which may be explained by the complexing with oxy-
three additives were found to extend the induction time for precip- gen atoms in the V(III)-sulphate ion pairs. In the low molecular
itation from less than 10 days to more than 90 days and the per- weight organic compound category, compounds with –OH groups
formance improved with increasing content of SHMP, urea, K2 SO4 failed to show any effect in stabilizing V(II) ions except for glyc-
up to 3 wt%, 5 wt% and 2 wt%, respectively. erol. Similar results were observed in V(III) solution but only with
Inspired by threshold agents which are used to prevent scale glycerol and citric acid as effective inhibitors. In the group of poly-
formation in boilers, Rahman and Skyllas-Kazacos [27,49] screened meric organic compounds, only Flucon-100 was found to enhance
a variety of scale inhibitors as electrolyte additives to restrain the the stability of both V(II) and V(III) ions. Dispersing agents, includ-
precipitation of supersaturated V(V) solutions at high temperature ing maleic anhydride, did not extend but rather reduced the in-
such as 50 °C. In their initial screening, threshold agents contain- duction time for V(II) precipitation, but the acrylic acid derivatives
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
(NH4 )2 C2 O4 2 – Room temp. Improved 5–10 wt% V(V) was not completely [44,45]
reduced by the (NH4 )2 C2 O4
additive
(NH4 )2 C2 O4 3 – – SOC dependent 2–3 wt% Solution forms orange oxalate [44,45]
complexes immediately at
100% SOC. But 80%–90% SOC
V(V) solution can remain 2.8 M
after 20 days at 5 °C and the
cell efficiency was improved
Thiourea 2 – Room temp. Improved 5 vol% of 1 M solution All V(V) was reduced [44,45]
Glycerol 2 – Room temp. Improved 1–20 vol% of 98% V(V) has been reduced to V(IV) [44,45]
glycerine and the reduction quantity
depends on the addition
amount
Fructose 6 6 Room temp. Improved 1% Add 3 M V2 O5 to 6 M H2 SO4 [44,45]
with 1% fructose solution and
electrolyse with H2 SO4 as the
other side. Reverse the cell
polarity to make stabilised 6 M
V(V) solutions
Sorbitol 5 6 Room temp. Improved 1% Add 2.5 M V2 O5 to 6 M H2 SO4 [44,45]
with 1% fructose solution and
electrolyse with H2 SO4 as the
other side. Reverse the cell
polarity to make stabilised 5 M
V(V) solutions
Diphenyl-thiocarbazone 2 – Room temp. Deteriorated 0.01–1 wt/vol Form solid products [44,45]
Tri-n–butyl phosphate 2 – Room temp. Deteriorated 1–5 vol% Form solid products [44,45]
EDTA disodium salt 2 – Room temp. Deteriorated 0.5–5 vol% Slightly reducing and form [44,45]
solid products
Tri-sodium citrate 2 – Room temp. Improved 1–5 wt/vol% Slightly reducing [44,45]
Ascorbic acid 2 – Room temp. Improved 1–5 wt/vol% Slightly reducing [44,45]
Malic acid 2 – Room temp. – 1–5 wt/vol% Reducing and gassing [44,45]
Succinic acid 2 – Room temp. – 1–5 wt/vol% Unreacted [44,45]
Salicylic acid 2 – Room temp. – 1–5 wt/vol% Form solid product [44,45]
Glycerine 2.96 – Room temp. Improved 1% [44,45]
Starch 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Glycerol 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Sorbitol 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Myo-inositol 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Mannitol 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Lactose 2 – 45 Significantly 0.03 g/25 mL Reduced [46,47]
improved
D-fructose 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
D-glucose 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Methyl-α -d-glucopyranoside 2 – 45 Deteriorated 0.03 g/25 mL Reduced [46,47]
D gluconic acid (Na) 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Succinic acid 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
Malic acid (Fizz) 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
L-ascorbic acid 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Citric acid 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Trisodium citrate 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Tartaric acid 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Poly(tetrahydrofuran) 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Furfural 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Sodium dodecyl sulphate (Fizz) 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Sodium lauryl sulphate 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Poly 4-styrene 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
Urea 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
Thiourea 2 – 45 Improved 0.03 g/25 mL Reduced [46,47]
Ethanolamine 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Glycine 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Nitrilotriacetic acid 2 – 45 Slightly improved 0.03 g/25 mL Reduced [46,47]
Ethylenediaminetetraacetic acid 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
Di(2-ethylhexyl) phosphate 2 – 45 Deteriorated 0.03 g/25 mL Reduced [46,47]
Boric oxide 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
H3 PO4 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
(NH4 )2 C2 O4 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
(NH4 )2 SO4 2 – 45 Slightly improved 0.03 g/25 mL Slightly reduced [46,47]
K2 S2 O8 2 – 45 Slightly improved 0.03 g/25 mL Unreacted [46,47]
(continued on next page)
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Table 1. (continued)
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Table 1. (continued)
(NH4 )2 SO4 2 5 50 Improved 2–3 wt% The higher dosage, the less [50]
improvement
(NH4 )3 PO4 2 5 50 Significantly 1–2 wt% Good cycling stability at [50]
improved 40mAcm−2 under 45 °C for
more than 150 cycles
K3 PO4 2 5 50 Improved 1–2 wt% [50]
Na7 P5 O16 (SPPP) 2 5 50 Significantly 2 wt% [50]
improved
H3 PO4 + (NH4 )2 SO4 2 5 50 Significantly 1 wt% + 2 wt% [50]
improved
K3 PO4 + (NH4 )2 SO4 2 5 50 Significantly 1–2 wt% + 2 wt% [50]
improved
SPPP 3 5 30 Improved 1 wt% (0.041 M) [48]
K3 PO4 3 5 30 Improved 3 wt% (0.219 M) [48]
H3 PO4 3 5/5.5 30 Significantly 1 wt% (0.16 M) [48]
improved
(NH4 )2 SO4 3 5/5.5 30 Negligible 1 wt% (0.12 M) [48]
(NH4 )2 SO4 + SPPP 3 5/5.5 30 Negligible 2 wt% + 1 wt% [48]
(0.23 M + 0.041 M)
(NH4 )2 SO4 + H3 PO4 3 5/5.5 30 Significantly 2 wt% + 1 wt% Good cycling stability at [48]
improved (0.23 M + 0.16 M) 50/80 mA cm−2 under 30 °C for
around 90 cycles
Glycerol + (NH4 )2 C2 O4 – – Improved 1 wt% + 2 wt% Severe gassing during charging [45,48]
Al2 (SO4 )3 2 5 40 Improved 3 wt% [51]
Na3 PO4 2 5 −5 to 40 Significantly 3 wt% Excellent improvement in [51]
improved ex-situ static test but serve
precipitation (VOPO4 •2H2 O) in
the positive graphite felt after
20 cycles under 50 mA cm−2
K2 SO4 2 5 −5 to 40 Significantly 3 wt% KVSO6 •3H2 O precipitation [51]
deteriorated
Na2 SO4 2 5 40 Slightly 3 wt% [51]
deteriorated
Iso-propanol 1.8 5 40 Improved 10 wt% Colour change. Reduced by [51]
V(V)
Glucose 1.8 5 40 Improved 1 wt% Colour change. Reduced by [51]
V(V)
®
Darven C 2 5 40 Deteriorated 0.3 wt% [51]
Polyethylene glycol 2 5 40 Improved 0.3 wt% Colour change. Reduced by [51]
V(V)
CH3 (CH2 )11 SO3 Na 2 5 40 Deteriorated 0.3 wt% [51]
Polyacrylic acid 1.8 5.1 40 Improved 0.1–1.0 wt% [51]
Polyacrylic acid 2 6 40 Improved 0.1–1.0 wt% [51]
Dimethyl sulfoxide 2 5 40 Improved 10 wt% Color change. Reduced by V(V) [51]
Tween 20 2 5 40 Improved 0.01 mL Colour change. Reduced by [51]
V(V)
Saccharin 2 5 40 Deteriorated 0.1–2 wt% [51]
CH3 SO3 H 2 5 40 Improved 2.1–3 wt% [51]
CH3 SO3 H + Polyacrylic acid 1.8 5 40 Significantly 7 wt% + 0.4 wt% [51]
improved
Boric acid
2 5 40 Deteriorated 1 wt% [51]
Sulphamic acid
2 5 40 Deteriorated 0.3–1.0 wt% [51]
(NH4 )2 HPO4 + MgCl2 2 5.5 −5 to 50 Significantly 0.15 M + 0.0 5 M [52]
improved
(NH4 )2 HPO4 + MgCl2 2 5.5 −5 to 50 Significantly 0.2 M + 0.0 5 M [52]
improved
(NH4 )2 HPO4 2 5.5 40–50 Deteriorated 0.15 M [52]
(NH4 )2 HPO4 + MgCl2 2 5.5 −5 to 50 Improved 0.1 M + 0.1 M [52]
MgCl2 2 5.5 −5 to 50 Improved 0.05 M [52]
CaCl2 2 5.5 −5 to 50 Deteriorated – CaSO4 precipitation [52]
Hexadecyl trimethyl ammonium 1.5 4.5 45 Improved 0.0 0106–0.0 053 M 0.00265 M improved the best [53]
bromide(CTAB) for 95% SOC
D-sorbitol 1.8 4.8 20–60 Improved 1 wt% Enhanced electrochemical [54]
performance
Glucose 1.8 4.8 20–60 Negligible 1 wt% [54]
Fructose 1.8 4.8 20–60 Negligible 1 wt% [54]
Mannitol 1.8 4.8 20–60 Slightly 1 wt% [54]
deteriorated
Phytic acid 1.8 3 25–60 Improved N/A Enhanced electrochemical [55]
performance
(continued on next page)
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Table 1. (continued)
showed good stabilizing performance. It was also noted that poly- with 2 M V in 5 M total sulphate/bisulphate solutions to further
meric additives tend to form precipitates themselves after react- evaluate the influence of the addition of the pre-mentioned agents
ing with the sulphuric acid in the vanadium solutions. Due to the and (NH4 )3 PO4 was proved to be the best precipitation inhibitor at
great inhibiting ability of polyphosphate in V(IV) and V(V) solu- low temperatures.
tions, Mousa and Skyllas-Kazacos further studied the behaviour of The results from the additive screening studies for the positive
polyphosphate-based additives in the negative vanadium solutions. and negative half-cell electrolytes were used by Skyllas-Kazacos
For V(III) solution, sodium tri-, pent– and hexapolyphosphate were and co-workers [48] to select the most promising additives for
all seen to possess great stabilizing ability. However, only sodium use in a high energy density cell employing a 3 M vanadium elec-
pentapolyphosphate (SPPP, Na7 P5 O16 ) showed excellent improve- trolyte. Experiments were carried out in 3 M V(V) in 5 M total sul-
ment of the induction time in V(II) solution. It was assumed that phate/bisulphate solutions at 30 °C and 50 °C and the addition
the linear structure of SPPP may play a role in prevention of V(II) of 1 wt% Na7 P5 O16 , 1 wt% K3 PO4 and 2 wt% (NH4 )2 SO4 + 1 wt%
precipitation. Cyclic tests were carried out in vanadium flow cells H3 PO4 were found to be able to keep the vanadium concentration
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
above 2.6 M after 22 days compared to only 2.4 M in the blank. trolyte flow was observed during the experiment. Thus the addi-
A larger difference in final concentration was observed when the tive comprising 2 wt% (NH4 )2 SO4 + 1 wt% H3 PO4 appeared to be a
temperature was elevated to 50 °C. Thus a high energy density very effective stabilising agent for RFBs with high vanadium con-
VRFB using 3 M V in 5 M total sulphate/bisulphate electrolyte with centration.
2 wt% (NH4 )2 SO4 + 1 wt% H3 PO4 additive was assembled and eval- Other research groups started the investigation of additives to
uated over 90 constant current charge–discharge cycles. Only a enhance the stability of supersaturated vanadium solutions after
slight capacity decrease was observed. The coulombic efficiency 2010 and these results are also summarised in Tables 1–4. For
remained around 100% and the voltage and energy efficiencies example, researchers at PNNL began an investigation of additives
remained relatively constant. No precipitation or restricted elec- in the vanadium electrolyte in the 2010 s. Similar to the results
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
(NH4 )2 C2 O4 3 – 5–45 Depends on 2–3 wt% Stable at high temperatures but [44,45]
temperatures easy to form crystals at low
temperature
Glycerine 2.82 – Room Improved 1% The concentration of V(II) [44,45]
temperature decreased with time but
decreased less compared to the
solution without additives
Starch 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Glycerol 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Sorbitol 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
Myo inositol 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Mannitol 2 3 Room Significantly 0.003 g/25 mL [46,47]
temperature deteriorated
Lactose 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
D-fructose 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
D-glucose 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
Methyl-α -d- 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
glucopyranoside temperature
D-gluconic acid 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Succinic acid 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
Malic acid 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
Ascorbic acid 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Citric acid 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Trisodium 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Tartaric acid 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
Polytetrahydrofuran 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Furfural 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
Sodium dodecyl sulphate 2 3 Room Deteriorated 0.003 g/25 mL Gassing [46,47]
temperature
Sodium lauryl sulphate 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
Poly 4-sytrene 2 3 Room Negligible 0.003 g/25 mL [46,47]
sulphonate temperature
Urea 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Thiourea 2 3 Room Improved 0.003 g/25 mL The electrolyte became [46,47]
temperature dark/purple
Ethanolamine 2 3 Room Deteriorated 0.003 g/25 mL [46,47]
temperature
Glycine 2 3 Room Deteriorated 0.003 g/25 mL Gassing [46,47]
temperature
Nitrilotriacetic acid 2 3 Room Negligible 0.003 g/25 mL Gassing [46,47]
temperature
EDTA 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
DL-(2-ethylhexyl) PO4 2 3 Room Negligible 0.003 g/25 mL Gassing [46,47]
temperature
Boric acid 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
Phosphoric acid 2 3 Room Negligible 0.003 g/25 mL [46,47]
temperature
(NH4 )2 C2 O4 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
(NH4 )2 SO4 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
K2 S2 O8 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
Glycylglycine 2 3 Room Improved 0.003 g/25 mL [46,47]
temperature
(continued on next page)
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Table 4. (continued)
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Table 4. (continued)
of UNSW [31,49], K2 SO4 was found to inhibit the precipitation abandoned as an alternative supporting electrolyte for the VRFB.
of V(IV) ions but was observed to promote the crystallization of The PNNL team found however that in the mixed sulphuric and
V(V) ions in the form of KVSO6 •3H2 O. Due to the low solubility of hydrochloride (2.5 M SO4 2− and 6 M Cl− ) solution, the vanadium
KVSO6 •3H2 O, additives containing K+ were thus not recommended ions of all four valence states can reach up to 2.5 M in the tem-
for stabilizing the VRFB electrolyte [51]. However, according to the perature range from −5 to 40 °C. Furthermore, the 90% SOC elec-
UNSW report, K3 PO4 was found to be able to stabilize both the trolyte of a 3 M VRFB system also showed great stability over 10
V(V) and V(IV) ions, indicating that PO4 3− has a higher interac- days at 50 °C, which further extends the energy density and tem-
tion with V(V) ions to prevent precipitation. Later the PNNL group perature range. Their analysis showed that the presence of Cl− an-
also confirmed a great stabilizing effect of Na3 PO4 for all vana- ions can interact with some of the V(V) ions ([VO2 (H2 O)3 ]+ ), form-
dium ions (2 M V/5 M SO4 2− ) by observing that no precipitation ing a soluble neutral VO2 Cl(H2 O)2 intermediate that slows down
occurred for 30 days at temperatures of −5 to 40 °C. Unexpect- the V2 O5 thermal precipitation reaction. This mechanism has how-
edly, in a flow cell test with 3 wt% Na3 PO4 , the VRFB showed pre- ever been disputed by others who point out that the higher V(V)
cipitation on the positive side after 20 charging–discharging cycles stability is simply due to the higher proton concentration in the
at 50 mA cm−2 and XRD results showed that VOPO4 •2H2 O crystals mixed acid electrolyte formulation that shifts the V2 O5 thermal
were formed during the cycling tests. Similar observations were re- precipitation equilibrium reaction. By adding HCl to the sulphuric
ported by Kausar [38] and Mousa and Skyllas-Kazacos [32,33] for acid electrolyte, a higher proton concentration can be achieved to
precipitation in stirred solutions. The PNNL group thus pointed out stabilise the V(V) ions without exceeding the solubility product of
the necessity of in-situ flow cell stability tests in any additive stud- the V(IV), V(III) and V(II) sulphates in solution. The dissolution of
ies [51], although Skyllas-Kazacos and co-workers did in fact find V(IV) ions in the form of [VO(H2 O)5 ]2+ , still depends on the solu-
that the amount of additive and the total sulphate concentration bility of VOSO4 while V(III) ions ([V(H2 O)6 ]3+ ) are determined by
also play a critical role in determining the effectiveness of particu- the solubility products of V2 (SO4 )3 and VOCl. The mixed acid elec-
lar stabilising additives. trolyte thus improves the solubility of the V(IV), V(III) and V(II)
In a work to study the influence of phosphate as an additive, ions by decreasing the concentration of SO4 2− relative to the stan-
Roznyatovskaya et al. used NMR and DLS on sulphate-based V(V) dard vanadium/sulphuric acid electrolyte.
electrolytes [106]. With DLS it was possible to trace the agglomer- Subsequent DFT and 51 V and 35 Cl NMR studies by the PNNL
ation to larger particles and determine induction times. The NMR group were employed to further elucidate the structure of vana-
measurements showed that at practical concentrations of around dium (V) cations [65]. These showed that V(V) is present as
1.6 M, V(V) was a dimer in the form [V2 O3 (H2 O)8 ]4+ and in diluted [V2 O3 Cl2 •6H2 O]2+ at concentrations greater than 1.75 M and forms
solutions a monomer [VO2 (H2 O)3 ]+ . Phosphate concentrations up chlorine-bonded [V2 O3 Cl2 •6H2 O]2+ with increasing temperature
to 0.15 M resulted in an increase of the induction time of precipi- which hinders the deprotonation reaction, thus producing less
tation of V2 O5 by the formation of a phosphate complex. At higher V2 O5 precipitates. The PNNL group had initially investigated the
phosphate concentrations, precipitation took place as VOPO4 •xH2 O. use of single HCl supporting electrolyte for the VRFBs [66]. No
The phosphate complex is in an equilibrium with the concentration precipitation was observed in 3 M vanadium solutions of each ox-
of vanadium and sulphate, which is why the ideal phosphate addi- idation state in 10 M Cl− solution for longer than 15 days over
tive concentration must be matched to the electrolyte composition. 0–50 °C. The formation of [V2 O3 •4H2 O]4+ and [V2 O3 Cl•3H2 O]3+
In their screening of organic additives for the VRFB, the PNNL was proposed as the reason for the modified stability of V(V) in
group also found that chemicals with C=C, –OH, –CHO and C=O the HCl solution, however similar stabilities were reported by the
were easily oxidized by V(V) ions. Polyacrylic acid was ultimately UNSW group for > 4 M V(V) solutions in 7–8 M H2 SO4 suggesting
selected as a promising additive to stabilize the vanadium elec- that the high proton concentration rather that the Cl ions are in
trolyte over the −5 to 40 °C temperature range [51]. Nevertheless, fact responsible for the increased stability of V(V) [27]. The vana-
in-situ flow cell tests with such additive were not presented. dium hydrochloride acid solution was however seen to be less vis-
The PNNL group also focussed their attention to the investiga- cous than the sulphate solution, which was expected to reduce the
tion of the mixed acid electrolyte system [63]. Skyllas-Kazacos and pump energy consumption, although the issue of HCl loss due to
co-workers had previously shown that 4 M V(III) and V(II) solutions vaporisation was not discussed.
could be prepared in HCl supporting electrolytes [64]. The loss of The mixed acid system was therefore found to be the most
HCl due to vaporisation even at room temperature, however, led appropriate supporting electrolyte by PNNL and by operating at
to the precipitation of V(V) in the positive half-cell, so HCl was higher temperatures, the reaction kinetics also showed some en-
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Fig.3. Continued
hancement compared to the pure sulphate system. A 1 kW/1 kWh eral advantages of the mixed acid system over the conventional
VRFB system with mixed HCl and H2 SO4 supporting electrolyte sulphuric acid system were demonstrated, including higher stack
was developed to further validate this technology [67]. No pre- energy efficiency, lower viscosity, less necessity of cooling system
cipitation was observed during the operation in the SOC range of and higher system efficiency. A patent on RFBs with supporting
15–85% at 80 mA cm−2 at temperatures higher than 45 °C. Sev- electrolyte containing Cl− was granted in 2014 [68]. However, as
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
with any HCl containing supporting electrolyte for the VRFB, the itation in the positive graphite felt. The cell with inositol additive
risk of HCl evaporation and Cl2 evolution during charging cannot also showed slight capacity loss compared to the cell without ad-
be avoided. Although it is claimed that less than 10 ppm Cl2 gas is ditives.
produced at temperatures lower than 40 °C and SOCs lower than The addition of methanesulphonic acid (CH3 SO3 H, MSA) and
80%, Cl2 evolution is still likely to occur in multi-cell stacks where aminomethylsulphonic acid (NH2 CH2 SO3 H, AMSA) in the positive
individual cells can readily go into overcharge as a result of un- vanadium sulphate electrolyte was also investigated [63]. Both
even electrolyte flow distribution, leading to corrosion and poten- MSA and AMSA was able to reduce the degree of precipitation
tial safety issues. Vaporisation of HCl and loss of protons needed of 2 M V(V) in 3 M H2 SO4 at 50 °C and AMSA which contains
to stabilise the V(V) species at elevated temperature is also a crit- the –NH2 and –SO3 H group, showed the better improvement. The
ical concern that requires careful electrolyte maintenance to avoid cell cycling test results also showed that the cell with AMSA in
thermal precipitation during operation. To reduce the potential for the positive electrolyte had higher energy efficiency and larger
HCl vaporisation or Cl2 emissions, low chloride ion concentrations capacity retention (82.7%) after 40 cycles. It was speculated that
(0.1–0.2 M) were thus recommended. these functional groups can disturb the crystal growth by adsorb-
With these risks in mind, the PNNL group thus began inves- ing onto the surface of V(V) ions [56]. MSA was later studied as
tigation of stabilising agents to produce high vanadium ion con- the supporting electrolyte for V(IV)/V(V) redox couple [69,70] as
centrations and great operating temperature ranges for the VRFB its conductivity was comparable to that of HCl or H2 SO4 acid at
without the need for high HCl concentrations. In 2016, the PNNL concentration of 1 M and less corrosive. The experimental results
group filed another patent on vanadium sulphate electrolyte with show that the concentration of V(IV) ions in 2–6 M MSA can reach
binary additives containing chloride ions and phosphate ions [52]. 3 M and can remain stable for several months in the tempera-
MgCl2 was initially reported to be quite effective in stabilizing ture range of 20–60 °C. In fact, up to 4 M of V(IV) ions can be
the vanadium species, probably due to the additional effect of dissolved in MSA solution, but the viscosity was very high even
the Mg2+ ions. However, the single inorganic chloride salt failed at room temperature. Besides, the conductivity of electrolyte de-
to stabilize the electrolyte at low temperatures. Therefore, phos- creased with increasing vanadium concentrations. However, the
phate ions (0.05–0.5 M) were added to modify the low temper- 3 M V(IV) in total 10 M methanesulphonic anion solution had com-
ature stability. The stabilising effect of ammonium phosphate is parable viscosity (12.37 mPa s) and conductivity (0.1 S cm−1 ) to
consistent with the original UNSW patent claims, but contra- that of 3 M V(IV) in 7 M total sulphate solutions. A vanadium/zinc
dicts PNNLs initial study that found that the addition of chem- flow battery was demonstrated with 70% energy efficiency
icals containing PO4 3− , HPO4 2− , H2 PO4 − and polyphosphate is achieved.
ineffectual [51]. The synergetic effect of Mg2+ , Cl− , NH4 + and More chemicals with functional groups such as –NH2 or –OH
HPO4 2− was thus claimed to contribute to the improved stabil- were explored and an investigation was carried out on the ef-
ity of the vanadium sulphate electrolyte over a broad temperature fect of trishydroxymethyl aminomethane additive (1–4 wt%) on the
range, although other additive combinations were not examined by thermal stability of V(V) ions at 40 °C [57]. The additive demon-
PNNL. strated a better thermal stabilizing effect and the use of 3 wt% gave
In the work by Liu and co-workers from Central South Univer- the best electrochemical performance in terms of cyclic voltam-
sity [53], the addition of hexadecyl trimethyl ammonium bromide metry results. The group further studied the VRFB with 3 wt%
(CTAB) into the electrolyte was shown to prevent polymerization trishydroxymethyl aminomethane additive in the positive elec-
of V(V) ions and supress the precipitation of V2 O5 via the interac- trolyte [71]. The modified electrolyte showed much better cyclic
tion between the quaternary ammonium head groups of CTAB mi- behaviour and reduced capacity loss than the pure vanadium elec-
celles and V(V) ions. Moreover, CTAB also demonstrated some cat- trolyte. Moreover, UV–visible spectroscopy confirmed that the elec-
alytic effect towards the redox reaction of V(IV)/V(V) and thus im- trolyte with additive remained the same before and after 40 cy-
proved the electrochemical performance as well. The same group cles. The morphology and structural characteristics of the graphite
later investigated a series of organic additives including fructose, felt electrode also showed no etching or oxidation produced by the
mannitol, glucose, d-sorbitol for the positive electrolyte of VRFBs. additive.
d-sorbitol appeared to be the best additive to reduce the precipita- Amino acids with –NH2 and –COOH such as l-glutamate and l-
tion of V(V) at high temperatures and also showed some improve- arginine were further studied by the same group [58]. The addition
ment in the electrochemical performance [54]. These compounds of 1.0 wt% l-glutamate extended the induction time for precipi-
had been previously screened by Skyllas-Kazacos and co-workers tation of V(V) ions by 5 h compared with the pristine electrolyte
however, and were found to be unsuitable as VRFB electrolyte ad- while the addition of l-arginine decreased the solubility of V(V)
ditives due to oxidation by V(V), giving rise to CO2 formation and ions. Based on the structural differences between these two com-
gassing. The addition of chemicals such as d-sorbitol of simple pounds, they believe it was the –OH and C=O functional groups of
chain structure with –OH groups thus has the risk of CO2 evolu- the l-glutamate that contributed to the improved stability, while
tion accompanied by reduction of V(V) to the more soluble V(IV) the –N=C structure in the l-arginine would decompose and form
species. This can readily be observed by a colour change from the poly-vanadium ammonium of low solubility. The positive effect can
yellow V(V) to the blue/green colour of a fully or partially reduced also be found in l-glumtamic acid which has similar structure to
V(IV) solution that is often not reported, but gives misleading re- l-glutamate [59].
sults in the V(V) thermal precipitation experiments. Chang and co-workers also began studying stabilizing additives
To avoid oxidation by V(V), ring organic compounds such as in- for the VRFB in 2012 [60]. They found that additives such as glyc-
ositol and phytic acid were further studied [55]. Inositol and other erol and hexanehexol, thiols, CTAB and other chain or ring struc-
organic ring compounds had been identified as potential precipita- tural compounds can stabilize 2 M V (V) solution at 60 °C for
tion inhibitors in the early UNSW patents, but these were discon- very long periods but these actually reduced V(V) to V(IV) gradu-
tinued in favour of the more stable inorganic additives. Wu et al. ally thus giving erroneous results. If used in the VRFB electrolytes,
did however find that both inositol and phytic acid exhibited some these additives would lead to self-discharge or gassing in the posi-
improvement in the V(V) stability at temperatures such as 40, 50, tive half-cell during battery operation. The researchers thus shifted
and 60 °C and phytic acid can further extend the induction time. their focus to the investigation of dispersants [60]. Among the sur-
However, the cell operated with phytic acid additive showed a re- factants studied with different carbon atoms all the dispersants
markable capacity loss during the cycling test and showed precip- except for anionic surfactant were found to extend the induc-
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
tion time for vanadium electrolyte precipitation in the temperature age control could be highly significant and deserves further inves-
range of 45–60 °C, with the cationic type being the most effec- tigation.
tive. A further study of dosage amount of the Coulter dispersant
with cationic surfactant showed that the higher additive content 3.3. Kinetic enhancers
increased viscosity and reduced conductivity. Foams were gener-
ated when the addition exceeded 0.3 w/w. Dosage levels of 0.05– The kinetics of V(V)/V(IV) and V(III)/V(II) redox reactions de-
0.1 w/w were thus considered as the most appropriate and no vari- termine activation overpotential and thus influence the voltage
ation in valence or concentration of vanadium was observed from efficiency and current/power densities of the VRFBs. A common
the UV–vis spectroscopy for these dosages. The cell with the 0.05– method to improve the kinetics is to modify the electrodes by the
0.1 w/w Coulter cationic surfactant also showed a higher energy introduction of oxygen functional groups [72,73], by deposition of
efficiency and the same coulombic efficiency as the cell without metal or metal oxides [74–76], by nitrogen or other heterogeneous
additives. atom doping [77–79], etc. Although these electrode modification
Tables 1–4 summarize all of the investigated stabilising agents methods have effectively enhanced the reversibility of the vana-
for vanadium solutions of different valances reported to date. Dis- dium redox reactions during electrochemical measurements and
crepancies can be found in results from different groups, such short-term cell cycling studies, any decay in their catalytic effect
as the effects of (NH4 )2 SO4 [48,50] in V(V) solution, K2 SO4 , during long-term cycling or cell operation is not understood. Oxy-
[31,51,61] Na2 SO4 [51,61], and (NH4 )2 SO4 [51,61] in V(IV) solu- gen functional groups will undergo reduction at different poten-
tion. These controversies are not only derived from the differ- tials during cell cycling, while deposited metals and oxides can
ences between the concentrations of V and S in the electrolyte gradually dissolve in the electrolyte, limiting their effectiveness. On
but are also caused by the variations in the amount of additives the other hand, a simpler approach that was proposed by Skyllas-
used in each study. When the concentration of the additive is Kazacos in the 1980s is to add the electrocatalyst to the electrolyte
close to or exceeds its own solubility in the vanadium solutions, and carry out in-situ electro-activation by incorporating the ad-
it will form seed crystal and thus induce precipitation in the vana- ditive onto the electrode surface during charge–discharge cycling.
dium solutions when they become supersaturated. In addition, the A steady state condition would be attained that should allow on-
SOC of the testing electrolyte will also have a strong impact on going electro-activation with less chance of deterioration during
the results observed. For example, the presence of small amounts charge–discharge cycling.
of V(IV) in the V(V) solution can significantly vary the ther- In an early patent by Skyllas-Kazacos [3], the use of electrolyte
mal stability of the V(V) electrolyte due to interactions between additives with kinetic enhancing ions and/or hydrogen evolution
the ions. inhibitors was described. Salts of Au, Mn, Pt, Ir, Ru, Os, Re, Rh, Sb,
Another important consideration is that in an operating cell, Te, Pb and/or Ag can be added to the electrolyte and the deposition
diffusion of any additives across the membrane is inevitable. Any of the metal or metal oxides on the electrode surface is induced
additive used to stabilise the positive or negative half-cell elec- in-situ, thus providing active sites for the vanadium redox couple
trolytes in the VRFB must not have any detrimental effect in reactions. The addition of Pb, Bi, Tl, Hg, Cd, In, Ag, Be, Ga, Sb, As,
the other half-cell. For example, K2 SO4 can improve the stability Zn, Ca, and/or Mg ions into the anolyte can be used to increase
of V(IV) solution but decreased that of V(V) ions [49,51]. Some the hydrogen overpotential and inhibit H2 evolution during charg-
reducing additives such as simple organic species with C–OH, ing of the VRFB. In a later study, the cyclic voltametric behaviour
C=O, COOH, etc. may be suitable for the negative half-cell elec- of graphite fibre electrodes in vanadium electrolyte containing the
trolyte, but can be oxidized by the V(V) ions in the positive so- metal ions Pt4+ , Pd2+ , Au4+ , Mn2+ , Te4+ , In3+ or Ir3+ [74], irid-
lution, resulting in CO2 evolution, which can interfere with the ium ions were shown to have the best electro-catalytic effect for
normal operation of the VRFB. On the other hand, oxidative ad- the vanadium redox reactions. Mn2+ , Te4+ and In3+ exhibited simi-
ditives such as Na2 WO4 and H2 WO4 [62] will oxidize the V(II) lar enhancement in the electroactivity but Pt, Pd and Au salts were
ions, leading to reduced coulombic efficiency. Reductants and oxi- found to catalyse the hydrogen evolution which will decrease the
dants are therefore not appropriate stabilising agents for the VRFB coulombic efficiency of the VRFBs. Recently, Cao et al reported that
electrolytes adding MoO4 2− into the electrolyte can achieve a similar improve-
After a comparison of the results ammonium phosphate is seen ment as that involving the modification of carbon paper electrodes
to be the most promising stabilising agent for the VRFB, with with MoO3 , leading to an increase in the cell voltage efficiency by
SHMP appearing to be a promising alternative, but further testing more than 5% over the current density range 100–150 mA cm−2
in both static and flowing cells over a wider range of temperatures [80].
and electrolyte compositions is still needed to verify the practical PNNL researchers also investigated the effect of Bi3+ additive in
suitability of these additives. the 2 M V + 5 M Cl− system. The redox peak potential separation
for the V(III)/V(II) redox couple reaction was found to be 90 mV
3.2. Immobilizing agents less with 0.01 M BiCl3 than for the cell without any additives while
no obvious change was observed in the electrochemical response
An alternative strategy proposed by UNSW to ensure high con- of the positive half-cell couple. It was believed that Bi metal is
centration of vanadium ions and prevent them from precipitat- electrodeposited onto the surface of graphite fibre since there was
ing is to add immobilizing agents such as gels, gums, starch, al- a peak located at 46 mV vs Ag/AgCl corresponding to the stan-
kaline pectate and operate the cell in static mode [46,47]. This dard potential for the Bi/Bi(III) redox couple. A further comparison
idea stems from the concept of gelled electrolytes and suspensions with Bi-coated graphite fibre also confirmed this assumption. The
which are efficient ion conductors without compromising stability largest improvement in voltage efficiency was obtained for the Bi3+
under different thermal conditions. The gelled electrolytes can also addition of 0–0.02 M, and the cell containing 0.01 M Bi3+ gave the
eliminate problems of electrolyte leakage, however, the increased highest discharge capacity. Moreover, no obvious decrease in en-
energy density and reduced leakage obtained from this approach ergy efficiency was observed over 50 charge–discharge cycles [81].
come at the expense of losing the distinctive feature of flow bat- A research group from Central South University further inves-
teries, i.e., independence of power and energy. Although the practi- tigated the effect of In3+ additives in the positive vanadium elec-
cal applications of a static vanadium redox battery would be more trolyte through a series of cyclic voltametric (CV) measurements,
limited, the use of such a device in frequency regulation and volt- steady-state polarization and constant current charge–discharge
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
tests. The results showed that the addition of In3+ can improve the • Some metals such as Cu and Ni, could deposit on the negative
rate constant of the V(V)/V(IV) redox couple reaction by 42% and electrode during charging, leading to increased hydrogen evolu-
the best addition amount was only 10 mM [82]. A similar investiga- tion
tion was carried out on Sb3+ ions and it was shown that Sb can be • SiO2 and NH4 + were reported to deposit in the graphite felt
deposited onto the electrode surface during the charge–discharge electrode and reduce flow-rate [91–93]
operation and catalyse the V(III)/V(II) reaction [83].
Apart from these metal-ion containing electrolyte additives, or- Certain impurities may also deposit within the pores of the
ganic compounds containing different functional groups have also membrane, leading to fouling and increased cell resistance.
been reported to provide electrocatalytic properties when added In a patent on preparation of high purity vanadium electrolyte
to the vanadium battery electrolyte. In the investigation of precip- [91], impurities such as Na2 SiO3 were found to undergo gela-
itation inhibitors discussed above, some organic additives showed tion and produce H2 SiO3 gel in the acidic solution, and it was
an improvement in electrochemical behaviour to varying extents therefore recommended that they be removed by filtration. It was
[53–60,69–71,84]. Since thermal treatment and chemical modifica- also suggested that ammonium ions be eliminated by heating
tion have been confirmed as effective methods for catalysing the the vanadium compound within the temperature range 400 °C–
vanadium redox reactions because of the introduction of certain 690 °C for 1–4 h to avoid the precipitation of vanadium. Sumit-
functional groups on the electrode that provide active sites for the omo Electric Industries further looked into the impurity effects of
redox couple reactions, Jia and co-workers added hydroxyl groups Si and NH4 + in the electrolyte and disclosed that the VRFBs can
into the electrolyte. By adding 1 vol% of glycerine and n-propyl give stable performance and precipitation in the carbon electrodes
alcohol to the electrolyte, an improvement in the kinetics of the can be avoided by limiting the NH4 + content below 20 ppm and
V(IV)/V(V) redox couple reaction in terms of the CV and electro- Si under 5 ppm m3 /m2 (impurity concentration × electrolyte vol-
chemical impedance spectroscopy (EIS) results was reported, with ume/electrode area) [93]. Burch from University of Tennessee also
the greater improvement seen for glycerine [85]. These results fail reported the negative effect of Si impurity and confirmed that the
to mention any long-term behaviour however, especially given that filtration method can reduce the Si level in electrolyte [94].
both of these organic compounds are unstable and undergo oxida- Given Cr3+ and Mn2+ are frequently found in the V2 O5 prod-
tion in V(V) as discussed above. This highlights a major problem uct of Pangang Group, Fei and co-workers from Sichuan Univer-
with many electrochemical studies of vanadium redox couples in sity decided to study the concentration effect of Mn2+ and Cr3+
the literature. Researchers often report small differences in peak on the electrochemical behaviour of the V(V)/V(IV) couple [95,96].
potential separation as being significant when in fact these dif- The test result did not show any side reactions with the Mn2+ and
ferences more often fall within experimental error. When study- Cr3+ impurities. Additionally, when these ions were in the range of
ing electrocatalysis, it is therefore vital to conduct several repeat 0.04–0.13 g L−1 and 0–0.30 g L−1 , the reversibility of V(V)/V(IV) re-
experiments in order to establish repeatability and reproducibility, action was enhanced and the diffusion coefficient of vanadium ions
especially given that the same electrode can give major differences was improved. But adverse effects on the impedances and diffusion
in peak current and peak potential separation from one experiment coefficients were discovered when the concentrations of Mn2+ and
to the next after polishing. The same can be said for cell cycling Cr3+ exceeded 0.13 g L−1 and 0.30 g L−1 .
tests that can exhibit variations in voltage and coulombic efficiency As Na+ , K+ and NH4 + are also common impurities in vanadium
up to 5% from one cycle to the next in the same experiment and electrolyte, Wen and co-workers studied the effect of those con-
from one cell to another using exactly the same materials and elec- taminants by adding K2 SO4 , Na2 SO4 and (NH4 )2 SO4 deliberately
trolyte composition. into vanadium electrolytes [61]. They found that the addition of
those additives into 2 M V(IV) in 5 M H2 SO4 negatively influenced
the stability of V(IV). When the concentration of H2 SO4 was de-
3.4. Electrolyte impurities creased to 4 M, only 0.05 mol% addition of K+ was found to in-
crease the stability of V(IV) slightly while the addition of Na+ and
In the preparation of vanadium electrolyte, vanadium pentox- NH4 + both accelerate the precipitation of V(IV) ions. Moreover, the
ide (V2 O5 ) is the preferred raw material due to its greater cost- addition of the three kinds of additives also showed an adverse
effectiveness [86,87]. However, during the production or recovery effect in extending the precipitation time. In the 2 M V(V) in 4 M
process of V2 O5 from sources such as ores, concentrates, iodide H2 SO4 solution with or without additives, however, no precipita-
thermal decomposition products, metallurgical slags, petroleum tion was observed at 25 °C for 44 days so they claimed the addi-
residues, fly ash and other wastes [88,89], other elements includ- tives had negligible effect on V(V).
ing but not limited to Fe, Si, K, Na, Ca, Mg, Cr, Cu, Mo, Zn, Mg, Ni More intriguingly, the researchers [61] also found that the con-
[61,86,90,91] may be introduced or remain in the vanadium oxide tent of impurities increased both in the positive and negative elec-
and subsequently enter the electrolyte. While some of these im- trolyte after charge–discharge cycling. Among these, the amount
purities have no detrimental effect in the VRFB, others could lead of Al, Ca, K increased the most remarkably by 100–50 ppm while
to operational problems and must be avoided. Unfortunately there the Mg, Fe, Na and B levels were elevated by 38–15 ppm and Si
have been few studies on the effects of impurities on the perfor- changed less than 1.5 ppm. After eliminating the possible source of
mance of the VRFB, so as a precaution, vanadium battery manu- graphite felt by experiment, they speculated the rise in impurity
facturers have tended to specify high purity V2 O5 for electrolyte levels was mainly from the framework materials and pipelines.
production, despite the fact that many excluded impurities may Potential impurities in the VRFB electrolyte were also investi-
have no effect at all. Given that a small increase in purity level gated by the UNSW group using cyclic voltammetry and their re-
above 98.5% can have a major impact on electrolyte costs, an un- sults are presented in Table 5. As expected, Ni, Cu and Sn increased
derstanding of impurities in the VRFB is vital for future cost reduc- hydrogen evolution when the electrode was scanned to the V(II)
tion [86,92]. peak potential, but other additives only enhanced hydrogen evolu-
Although there have been very few published papers on elec- tion if the electrode was scanned to more negative potentials cor-
trolyte impurity effects on VRFB performance, it is known that cer- responding to overcharging in the VRFB [86].
tain impurities could have detrimental effects and therefore need Although the above results provide some insight into the poten-
to be minimised. These impurities and their potential effects in- tial effects of certain electrolyte impurities on VRFB performance, it
clude: should be stressed that these are only preliminary studies and fur-
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
Table 5. Effect of selected metal cations on hydrogen evolution and vanadium redox couple reactions at glassy carbon electrode in 2 M VOSO4 + 2 M H2 SO4 .
ther detailed analysis is needed on a wider range of impurities and and human health. The use of bromine complexing agents previ-
impurity mixtures in order to establish a broad electrolyte specifi- ously employed in the Zn/Br battery was thus proposed for the
cation for VRFB producers. V/Br system. Agents including polyethylene glycol, tetrabutylam-
monium bromide, N-methyl-N-ethyl pyrrolidinium bromide (MEP)
3.5. Reducing agents and N-methyl-N-ethyl morpholinium bromide (MEM) and their
mixtures were evaluated in terms of the inhibiting effect on bro-
Chemical reductants can be added into the solution to assist mide vapour production [100]. Among these, MEP and MEM can
the dissolution and reduction of V(V) compounds in the process supress the Br2 vapour effectively but will deposit an organic layer
of electrolyte preparation [97]. They can also be used to rebalance onto the membrane, thus leading to reduced conductivity of the
the electrolyte when a variation between the positive and negative membrane and a decline of the total energy efficiency [101,102].
SOCs occurs, as caused by H2 evolution at the negative electrode Non-aqueous vanadium redox flow batteries are another branch
during charging or air oxidation of V(II) in the anolyte [98,99]. of RFBs based on the redox reactions between vanadium cou-
Skyllas-Kazacos and co-workers proposed the use of specific ples. Liu and co-workers first proposed the concept with vana-
chemical reductants as rebalancing agents for the VRFB. The de- dium acetyacetonate (V(acac)3 ) in acetonitrile as the initial elec-
sirable agents should be able to reduce the V(V) species without trolyte and tetraethylammonium tetrafluoroborate (TEABF4 ) as the
introducing any new impurities into the electrolyte. One example supporting electrolyte in 2009 [103]. The redox reactions be-
is adding oxalic acid to the V(V) solution to partially reduce the tween V(III)/V(II) and V(III)/V(IV) in the 0.01 M V(acac)3 and 0.05 M
V(V) ions to V(IV) so as to balance the SOCs of the two half-cell TEABF4 solution can produce a cell potential of 2.2 V (while the
solutions [97,98] : theoretical cell potential is 1.26 V for vanadium in sulphuric acid
system under standard conditions). However, even when the cell
2VO2 + + H2 C2 O4 ↔ 2VO2 + + H2 O + CO2 ↑ (15) was charged at 1 mA and discharged at 0.1 mA, the overpoten-
tial in the demonstrated cell was very large and the coulombic
A similar effect can also be achieved by adding ethanol, ethy- efficiency was only around 47%. The poor performance was at-
lene glycol, and many other simple organic species with C–OH, C– tributed to the low conductivity of the electrolyte and side reac-
O–C, C = O, COOH [99]. The preferred compounds should not intro- tions. Later the supporting electrolyte was changed to ionic liquids
duce undesirable elements to the electrolyte and ideally be fully 1-ethyl-3-methyl imidazolium hexafluorophosphate (EMIPF6 ) and
oxidised to CO2 and water, thereby leaving no residual impurities tetrabutylammonium hexafluorophosphate (TEAPF6 ) and the cell
that could accumulate and produce detrimental effects on cell per- potential increased to 2.4 V and 2.3 V respectively [104]. However,
formance [99]. Although a number of organic chemical rebalancing neither the cell polarizations nor the coulombic efficiencies (45%
agents have been used in laboratory studies to successfully rebal- for the EMIOF6 and 55% for TEAPF6 ) showed any obvious improve-
ance the VRFB electrolytes, there has been no investigation of any ment. Herr and co-workers replaced acetonitrile with various or-
intermediates that might be produced during the oxidation reac- ganic solvents including tetrahydrofuran (THF), 1,3-dioxolane (1,3-
tion that could potentially impact on long-term performance. This DO), acetylacetone (Hacac) and dimethyl sulfoxide (DMSO) while
is an area that deserves further investigation. still keeping TEAPF6 as the supporting electrolyte [105]. All sys-
tems can offer cell potentials in the range of 2.21–2.61 V and the
3.6. Additives in other vanadium redox flow systems best energy efficiency was achieved from the cell with DMSO as
solvent, but this was only 43.6% at the highest current density of
Apart from the vanadium/sulphuric acid electrolyte system, 0.1 mA cm−2 .
Skyllas-Kazacos and co-workers also proposed and developed the
vanadium bromide/chloride mixed electrolyte cell, known as the 4. Summary and future perspective
vanadium polyhalide or V/Br redox flow battery [4]. This technol-
ogy utilizes V(II)/V(III) as the redox couple in the negative elec- The vanadium electrolyte is the main component of the VRFB,
trolyte and Br− /Br3 − or Br− /ClBr2 − couple in the positive half- serving both ionic conduction and electrochemical energy conver-
cell. Up to 4 M vanadium bromide solution can be used as the sion. The use of a single metal element in both half-cells eliminates
electrolyte, giving an energy density of 70 Wh L−1 which is much the cross-contamination problems suffered by other redox flow cell
higher than that of the original VRFB. However, this system tends chemistries and thus offers an indefinite electrolyte life. The solu-
to suffer from Br2 evolution in the positive half-cell during charg- bility of vanadium ions not only determines the achievable energy
ing, which can impose detrimental effects on battery components density but also limits the operational conditions. Moreover, the
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
cost of the electrolyte occupies a significant portion of the overall [6] A. Parasuraman, T. Mariana, C. Menictas, M. Skyllas-kazacos, Electrochim. Acta
VRFB system cost, especially in applications requiring more than 4 101 (2013) 27–40.
[7] G. Wei, X. Fan, J. Liu, C. Yan, New Carbon Mater. 29 (2014) 272–279.
hours of storage capacity. After providing a brief introduction on [8] M. Park, J. Ryu, J. Cho, Chem. Asian J. 10 (2015) 2096–2110.
the chemistry of vanadium ions of different valence state in sul- [9] L. Cao, M. Skyllas-Kazacos, D.-W. Wang, Energy Storage Sci. Technol. 4 (2015)
phuric acid solutions and their precipitation mechanism, this paper 433–457.
[10] M. Kazacos, M. Cheng, M. Skyllas-Kazacos, J. Appl. Electrochem. 20 (1990)
has reviewed the literature on electrolyte additives in relation to 463–467.
increasing energy density, improving electrochemical performance, [11] J.E. Carpenter, J. Am. Chem. Soc. 56 (1934) 1847–1850.
effects of impurities and chemical rebalancing agents in the VRFB. [12] S.C. Furman, C.S. Garner, J. Am. Chem. Soc. 72 (1950) 1785–1789.
[13] I. Nils, F. Brito, Acta Chem. Scand. 13 (1959) 1971–1996.
It can be seen that although only a few groups are investigating
[14] H. Strehlow, H. Wendt, Inorg. Chem. 2 (1963) 6–10.
electrolyte additives, controversies still remain in identifying the [15] C. Dijkgraaf, Spectrochim. Acta 21 (1965) 1419–1421.
most effective stabilizing agents that will allow an increased en- [16] C. Dijkgraaf, Spectrochim. Acta 21 (1965) 769–773.
[17] O.W. Howarth, R.E. Richards, J. Chem. Soc. 0 (1965) 864–870.
ergy density and operating temperature range in practical systems.
[18] J.V. Hatton, Y. Saito, W.G. Schneider, Can. J. Chem. 43 (1965) 47–56.
It should be noted that any additive employed to enhance the ther- [19] W.P. Griffith, T.D. Wickins, J. Chem. Soc. A Inorg. Phys. Theor. 0 (1966)
mal stability should be compatible with all vanadium ions because 1087–1090.
the electrolyte will undergo changes in oxidation state in each half- [20] P. Blanc, C. Madic, J.P. Launey, Inorg.Chem. 21 (1982) 2923–2928.
[21] C.E. Ophardt, S. Stupgia, J. Chem. Educ. 61 (1984) 1102–1103.
cell during the charging and discharging processes and the additive [22] C. Madic, G.M. Begun, R.L. Hahn, J.P. Launay, W.E. Thiessen, Inorg. Chem. 23
will eventually diffuse into the other half-cell and interact with (1984) 469–476.
other vanadium ions. Furthermore, stability tests conducted only [23] F. Hardcastle, I. Wachs, Solid State Ion. 45 (1991) 201–213.
[24] K. Kanamori, Y. Ookubo, K. Ino, K. Kawai, H. Michibatat, Inorg. Chem. 30
in static solutions are insufficient since they do not consider any (1991) 3832–3836.
disturbance in the flow battery that might vary the precipitation [25] G.J.J. Leigh, J.S.S. de Souza, Coord. Chem. Rev. 154 (1996) 71–81.
behaviour and to lead to different results. Since disturbances could [26] F. Rahman, M. Skyllas-Kazacos, J. Power Sources 189 (2009) 1212–1219.
[27] F. Rahman, (Ph.D. Thesis), University of New South Wales, Australia. (1998).
include the introduction of small carbon fibres into the electrolyte [28] F. Rahman, M. Skyllas-Kazacos, J. Power Sources 72 (1998) 105–110.
from the carbon felt electrodes that could act as nucleation sites [29] M. Vijayakumar, L. Li, G. Graff, J. Liu, H. Zhang, Z. Yang, J.Z. Hu, J. Power
for the precipitation reaction. The transfer of water and sulphate Sources 196 (2011) 3669–3672.
[30] M. Vijayakumar, Z. Nie, E. Walter, J. Hu, J. Liu, V. Sprenkle, W. Wang,
species from one half-cell to the other in the case of cells employ-
ChemPlusChem 80 (2015) 428–437.
ing anion exchange membranes can also lead to significant compo- [31] M. Skyllas-Kazacos, Electrochem. Solid State Lett. 2 (1999) 121.
sition changes that will affect precipitation as observed by Skyllas- [32] A. Mousa, PhD Thesis, Univ. New South Wales, Aust. (2003).
[33] A. Mousa, M. Skyllas-Kazacos, ChemElectroChem 4 (2017) 130–142.
Kazacos and co-workers [33,51]. Longer operation under different
[34] N. Kausar, R. Howe, M. Skyllas-Kazacos, J. Appl. Electrochem. 31 (2001)
temperature conditions and varying SOC ranges should also be fur- 1327–1332.
ther evaluated. Vanadium electrolytes in mixed acids seemed to be [35] X. Lu, Electrochim. Acta 46 (2001) 4281–4287.
promising in increasing the thermal stability, energy density and [36] A.A. Ivakin, Zhurnal Prikladoni Khimii 39 (1966) 277–284.
[37] F. Sepehr, S.J. Paddison, J. Phys. Chem. A 119 (2015) 5749–5761.
achieve great electrochemical performance, however more evalu- [38] N. Kausar, (Ph.D. Thesis), University of New South Wales, Australia (2002).
ation tests are required to ensure the stability and safety of the [39] A.A. Ivakin, E.M. Voronova, Russ. J. Inorg. Chem. 18 (1973) 956.
newly developed systems. Further studies should be conducted on [40] F.K. Cameron, J.F. Breazeale, J. Phys. Chem. 7 (1903) 571–577.
[41] W.L. Marshall, E.V. Jones, J. Phys. Chem. 70 (1966) 4028–4040.
electrolyte impurities to understand the effect of each element on [42] M. Vijayakumar, S.D. Burton, C. Huang, L. Li, Z. Yang, G.L. Graff, J. Liu, J. Hu,
cell performance and determine their maximum content so as to M. Skyllas-Kazacos, J. Power Sources 195 (2010) 7709–7717.
offer the possibility of lower purity vanadium raw materials for [43] A. Mousa, M. Skyllas-Kazacos, Chemeca 99 Chemcial Eng. Solut. Chang. Envi-
ron. (1999) 1147–1152 [Barton, A. C. T.] Institute Engineering.
cost reduction. Chemical reductants are not suitable for use as pre- [44] M.A. Samad, M. Skyllas-Kazacos, M. Kazacos, Pat. WO1995012219 A1 (1995).
cipitation inhibitors but can be used in the preparation of vana- [45] M. Skyllas-Kazacos, M. Kazacos, Pat. US652514 B1 (20 0 0).
dium electrolyte as well as the rebalancing the vanadium elec- [46] M. Kazacos, M. Skyllas-Kazacos, Pat. WO1996035239 A1 (1996).
[47] M. Kazacos, M. Skyllas-Kazacos, Pat. US7078123 B2 (2002).
trolyte. The possible build-up of oxidation intermediates in the
[48] S. Roe, C. Menictas, M. Skyllas-Kazacos, J. Electrochem. Soc. 163 (2016)
electrolyte during long-term use is not understood however and A5023–A5028.
needs further investigation. Similarly, impurities introduced during [49] F. Rahman, M. Skyllas-Kazacos, J. Power Sources 340 (2017) 139–149.
[50] N. Kusar, A. Mousa, M. Skyllas-Kazacos, ChemElectroChem 3 (2016) 276–282.
electrolyte production, either from the use of chemical reductants
[51] J. Zhang, L. Li, Z. Nie, B. Chen, M. Vijayakumar, S. Kim, W. Wang, B. Schwenzer,
or from electrode materials used during suspended powder elec- J. Liu, Z. Yang, J. Appl. Electrochem. 41 (2011) 1215–1221.
trolysis, need detailed investigation. For example, some companies [52] Z. Nie, W. Wang, X. Wei, B. Li, J. Liu, V.L. Sprenkle, Pat. US2016/0099480 A1
employ dimensionally stable anodes (DSA) in the electrolysis cells (2016).
[53] X.W. Wu, S.Q. Liu, K.L. Huang, J. Inorg. Mater. 25 (2010) 641–646.
used to produce the 50/50 V3+ /V4+ solution used as the initial [54] S. Li, K. Huang, S. Liu, D. Fang, X. Wu, D. Lu, T. Wu, Electrochim. Acta 56
electrolyte for the VRFB. DSAs have a limited life and require reg- (2011) 5483–5487.
ular maintenance because the noble metal coating can slowly dis- [55] X. Wu, S. Liu, N. Wang, S. Peng, Z. He, Electrochim. Acta 78 (2012) 475–482.
[56] Z. He, J. Liu, H. Han, Y. Chen, Z. Zhou, S. Zheng, W. Lu, S. Liu, Z. He, Elec-
solve in the anolyte during electrolysis. This could lead to the in- trochim. Acta 106 (2013) 556–562.
troduction of noble metal impurities into the electrolyte that could [57] S. Peng, N. Wang, C. Gao, Y. Lei, X. Liang, S. Liu, Y. Liu, Int. J. Electrochem. Sci.
not only increase hydrogen evolution during charging, but more 7 (2012) 2440–2447.
[58] Y. Lei, S.-Q. Liu, C. Gao, X.-X. Liang, Z.-X. He, Y.-H. Deng, Z. He, J. Electrochem.
significantly could catalyse the oxidation of V(II) by hydrogen ions Soc. 160 (2013) A722–A727.
in the negative half-cell, leading to hydrogen evolution even dur- [59] X. Liang, S. Peng, Y. Lei, C. Gao, N. Wang, S. Liu, D. Fang, Electrochim. Acta 95
ing stand-by. These are all potential issues that need to be further (2013) 80–86.
[60] F. Chang, C. Hu, X. Liu, L. Liu, J. Zhang, Electrochim. Acta 60 (2012) 334–338.
investigated to ensure the stable operation and long cycle life of
[61] Y. Wen, Y. Xu, J. Cheng, G. Cao, Y. Yang, Electrochim. Acta 96 (2013) 268–273.
the VRFB in commercial installations. [62] A. Mousa, M. Skyllas-Kazacos, ChemElectroChem 2 (2015) 1742–1751.
[63] L. Li, S. Kim, W. Wang, M. Vijayakumar, Z. Nie, B. Chen, J. Zhang, G. Xia, J. Hu,
References G. Graff, J. Liu, Z. Yang, Adv. Energy Mater. 1 (2011) 394–400.
[64] C. Menictas, M. Skyllas-Kazacos, SERDF Grant, NSW Dep. Energy (2018).
[1] M. Skyllas-Kazacos, M. Rychcik, R.G. Robins, aG. Fane, J. Electrochem. Soc. 133 [65] M. Vijayakumar, W. Wang, Z. Nie, V. Sprenkle, J. Hu, J. Power Sources 241
(1986) 1057–1058. (2013) 173–177.
[2] M. Skyllas-Kazacos, WO1989005526A1 (1988). [66] S. Kim, M. Vijayakumar, W. Wang, J. Zhang, B. Chen, Z. Nie, F. Chen, J. Hu,
[3] M. Skyllas-Kazacos, J. Power Sources 124 (2003) 299–302. L. Li, Z. Yang, Phys. Chem. Chem. Phys. 13 (2011) 18186.
[4] M. Skyllas-Kazacos, Pat. US7320844 B2 (2002). [67] S. Kim, E. Thomsen, G. Xia, Z. Nie, J. Bao, K. Recknagle, W. Wang,
[5] W. Wang, Q. Luo, B. Li, X. Wei, L. Li, Z. Yang, Adv. Funct. Mater. 23 (2013) V. Viswanathan, Q. Luo, X. Wei, A. Crawford, G. Coffey, G. Maupin, V. Spren-
970–986. kle, J. Power Sources 237 (2013) 300–309.
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007
JID: JECHEM
ARTICLE IN PRESS [m5G;May 2, 2018;17:12]
[68] L. Li, S. Kim, Z. Yang, W. Wang, N. Zimin, B. Chen, J. Zhang, G. Xia, Pat. [89] R.R. Moskalyk, A.M. Alfantazi, Miner. Eng. 16 (2003) 793–805.
US9077011 B2 (2014). [90] Y. Tanaka, K. Horikawa, M. Mita, N. Tokuda, M. Kubata, US6613298B2 (2001).
[69] C. Tang, D. Zhou, Electrochim. Acta 65 (2012) 179–184. [91] M. Nakajima, T. Akahoshi, M. Sawahata, Y. Nomura, K. Sato, Pat. US5587132
[70] Z. He, Z. Li, Z. Zhou, F. Tu, Y. Jiang, C. Pan, S. Liu, J. Renew. Sustain. Energy 5 (1995) 2–6.
(2013) 23130. [92] J. St. John, Imergy uses recycled vanadium to cut materials costs for
[71] S. Peng, N. Wang, C. Gao, Y. Lei, X. Liang, S. Liu, Y. Liu, Int. J. Electrochem. Sci. flow batteries, Greentech media, 2014. https://www.greentechmedia.com/
7 (2012) 4388–4396. articles/read/imergys- slag- to- energy- storage- vanadium- recipe#gs.mZBWQtQ
[72] B. Sun, M. Skyllas-Kazacos, Electrochim. Acta 37 (1992) 2459–2465. (accessed 06-12-2017).
[73] B. Sun, M. Skyllas-Kazacos, Electrochim. Acta 37 (1992) 1253–1260. [93] M. Kubata, H. Nakaishi, N. Tokuda, Pat. US7258947 B2 (2002).
[74] B. Sun, M. Skyllas-Kazakos, Electrochim. Acta 36 (1991) 513–517. [94] A.W. Burch, (Master Thesis), University of Tennessee (2015).
[75] W.H. Wang, X.D. Wang, Electrochim. Acta 52 (2007) 6755–6762. [95] F. Huang, G. Wang, K. Yan, D. Luo, Chin. J. Inorg. Chem. 28 (2012) 898–904.
[76] D.J. Suárez, Z. González, C. Blanco, M. Granda, R. Menéndez, R. Santamaría, [96] F. Huang, Q. Zhao, C.H. Luo, G.X. Wang, K.P. Yan, D.M. Luo, Chin. Sci. Bull. 57
ChemSusChem 7 (2014) 914–918. (2012) 4237–4243.
[77] Y. Shao, X. Wang, M. Engelhard, C. Wang, S. Dai, J. Liu, Z. Yang, Y. Lin, J. Power [97] M. Kazacos, R.J.C. Mcdermott, M. Skyllas-Kazacos, Pat. WO1989005363A1
Sources 195 (2010) 4375–4379. (1988).
[78] J. Jin, X. Fu, Q. Liu, Y. Liu, Z. Wei, K. Niu, J. Zhang, ACS Nano 7 (2013) [98] M. Skyllas-Kazacos, B.G. Maddern, M. Kazacos, J. Jaqui, Pat. WO1990 0 03666A1
4764–4773. (1990) 1–18.
[79] L. Shi, S. Liu, Z. He, J. Shen, Electrochim. Acta 138 (2014) 93–100. [99] M. Keshavarz, G. Zu, Vanadium Flow Cell (2013) US2013/0095362A1.
[80] L. Cao, M. Skyllas-Kazacos, D.-W. Wang, ChemElectroChem 4 (2017) [100] M. Skyllas-Kazacos, G. Kazacos, G. Poon, H. Verseema, Int. J. Energy Res. 34
1836–1839. (2010) 182–189.
[81] B. Li, M. Gu, Z. Nie, Y. Shao, Q. Luo, X. Wei, X. Li, J. Xiao, C. Wang, V. Sprenkle, [101] G. Poon, A. Parasuraman, T.M. Lim, M. Skyllas-Kazacos, Electrochim. Acta 107
W. Wang, Nano Lett. 13 (2013) 1330–1335. (2013) 388–396.
[82] Z. He, L. Chen, Y. He, C. Chen, Y. Jiang, Z. He, S. Liu, Ionics 19 (2013) [102] S. Winardi, G. Poon, M. Ulaganathan, A. Parasuraman, Q. Yan, N. Wai, T.M. Lim,
1915–1920. M. Skyllas-Kazacos, ChemPlusChem 80 (2015) 376–381.
[83] J. Shen, S. Liu, Z. He, L. Shi, Electrochim. Acta 151 (2015) 297–305. [103] Q. Liu, A.E.S. Sleightholme, A.A. Shinkle, Y. Li, L.T. Thompson, Electrochem.
[84] S. Peng, N.F. Wang, X.J. Wu, S.Q. Liu, D. Fang, Y.N. Liu, K.L. Huang, Int. J. Elec- Commun. 11 (2009) 2312–2315.
trochem. Sci. 7 (2012) 643–649. [104] D. Zhang, Q. Liu, X. Shi, Y. Li, J. Power Sources 203 (2012) 201–205.
[85] Z. Jia, B. Wang, S. Song, X. Chen, J. Electrochem. Soc. 159 (2012) A843–A847. [105] T. Herr, J. Noack, P. Fischer, J. Tübke, Electrochim. Acta 113 (2013) 127–133.
[86] M. Skyllas-Kazacos, in: Proceedings of the Canadian Institute of Mining, Met- [106] N.V. Roznyatovskaya, V.A. Roznyatovsky, C.-C. Höhne, M. Fühl, T. Gerber,
allurgy and Petroleum, 2014. M. Küttinger, J. Noack, P. Fischer, K. Pinkwart, J. Tübke, J. Power Sources 363
[87] Www.sigmaaldrich.com 2017 (2017) 234–243.
[88] Y.S. Yusfin, T.N. Bazilevich, P.I. Chernousov, A.L. Petelin, V.I. Gubanov,
A.Y. Travyanov, Metallurgist 42 (1998) 209–211.
Please cite this article as: L. Cao et al., A review of electrolyte additives and impurities in vanadium redox flow batteries, Journal of
Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.04.007