You are on page 1of 15

Journal of Power Sources 601 (2024) 234242

Contents lists available at ScienceDirect

Journal of Power Sources


journal homepage: www.elsevier.com/locate/jpowsour

Review article

Towards a high efficiency and low-cost aqueous redox flow battery: A


short review
Zhaoxia Hou a, 1, Xi Chen a, 1, Jun Liu a, Ziyi Huang a, Yan Chen c, Mingyue Zhou b, *, Wen Liu a, **,
Henghui Zhou d
a
State Key Laboratory of Chemical Resource Engineering, College of Chemistry, Beijing University of Chemical Technology, Beijing, 100092, China
b
College of Carbon Neutrality Future Technology, China University of Petroleum, Beijing, 102200, China
c
Beijing Power Equipment Group Co., Ltd., Beijing, 102401, China
d
College of Chemistry and Molecular Engineering, Peking University, 100871, Beijing, China

H I G H L I G H T S

• Performance/cost balance is the key to low cost flow batteries


• The characteristics of various aqueous redox flow batteries are discussed
• Challenges include low energy density, slow kinetics, and instability
• Future trends favor aqueous organic and iron-based chemistries

A R T I C L E I N F O A B S T R A C T

Keywords: The aqueous redox flow battery (ARFB), a promising large-scale energy storage technology, has been widely
Energy storage researched and developed in both academic and industry over the past decades owing to its intrinsic safety and
Aqueous redox flow battery modular designability. However, compared to other technologies (e.g. Li-ion batteries), the relatively low energy
Levelized cost of storage
density, inferior efficiency, and high investment cost of the ARFB limit its further applications. Although
tremendous efforts have been made to improve the performance as well as the cost-effectiveness of ARFB by
various approaches, it is still a challenge to recognize a united and general roadmap for the next commercial-
available ARFB. Here we review the evaluation criteria for the performance of flow batteries and the develop­
ment status of different types of flow batteries. The factors affecting the performance of flow batteries are
analyzed and discussed, along with the feasible means of improvement and the cost of different types of flow
batteries, which is expected to provide useful references for the further development of high-efficiency and low-
cost flow batteries.

1. Introduction challenges for electric grid operators [2,3]. To buffer the impact of the
intermittent problems and improve the quality of power grids, it is
With the growth of population and the continuous consumption of necessary to develop low-cost large-scale electrical energy storage (EES)
fossil fuels, energy crises and environmental problems are becoming systems [4–6]. Lithium-ion batteries (LIBs) have attracted wide atten­
increasingly prominent. Securing sustainable and environmentally tion due to their high energy density and rate capacity, which are of
friendly energy is critical to the development of human society [1]. great importance for various applications from portable electronic
Renewable energy, such as solar, tidal, and wind power, provides sig­ products to electric vehicles [7,8]. However, the relatively high lev­
nificant opportunities to alleviate the potential energy crisis. However, elized cost of storage (LCOS) and security issues have limited their ap­
the inherent intermittence of these renewable sources causes significant plications in large-scale EES systems. The LCOS is defined as the total

* Corresponding author.
** Corresponding author.
E-mail addresses: mingyue.zhou@cup.edu.cn (M. Zhou), wenliu@mail.buct.edu.cn (W. Liu).
1
the co-first author.

https://doi.org/10.1016/j.jpowsour.2024.234242
Received 12 December 2023; Received in revised form 16 February 2024; Accepted 19 February 2024
Available online 10 March 2024
0378-7753/© 2024 Elsevier B.V. All rights reserved.
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

lifetime cost of the energy storage system divided by the cumulated conductivity and selectivity of the membrane, the reaction activity of
released energy of the investment, which is used to evaluate the eco­ active species, and the catalytic activity of the electrode; c) promoting
nomic development prospects of energy storage projects [9–11]. By the cycling stability of the electrolyte, electrode, and membrane
contrast, ARFBs are promising candidates for large-scale energy storage materials.
due to their modular designability, scalability, safety, and potentially In this review, we provide a brief introduction and overview of a low-
low cost on the condition of using low-cost electrolytes and increasing cost ARFB with a variety of active materials, by evaluating the electro­
the storage duration [12–14]. On the power generation and grid side, in chemical performance in terms of efficiency, energy density, power
scenarios where volume and weight requirements are not strict, the density, and cycle stability. The key metrics affecting battery efficiency
disadvantage of low energy density of ARFBs is no longer a limiting are analyzed, followed by mitigation strategies and their benefits.
factor. Its stronger security and scalability are competitive enough and Finally, we provide perspectives on the future direction of the devel­
render it greater promise for the needs of long-duration and grid-scale opment of high-efficiency and low-cost ARFBs.
energy storage than LIBs. Secondly, although the cost of ARFBs is
currently higher than LIBs, the emerging low-cost ARFBs and the effect 2. Performance evaluation method of ARFBs
of large-scale commercial deployment suggest a trend of fast
cost-reducing. Fig. 1 shows the components and working principle of a With the development of grid-scale EES, a variety of ARFBs have
typical redox flow battery (RFB). The conventional RFB consists of the been proposed. However, the electrochemical performance is varied due
stack unit, electrolyte, external storage tanks, circulation pumps, and a to the different nature of the active materials. For example, the limited
management and control unit. The electrode does not undergo redox solubility of K3[Fe(CN)6] in an aqueous medium leads to the relatively
reactions itself, which only provides sites for electrochemical reactions. low energy density of iron-based ARFBs [23,24] while the severe
Generally, the redox species is dissolved in the electrolyte and stored in crossover and slow redox kinetics of polysulfide lead to insufficient ef­
two tanks respectively, which are circulated through a peristaltic pump. ficiency though it is highly soluble in aqueous media [25]. The inno­
Unlike other batteries, this design allows for the separation of electro­ vation of materials and structural optimization has been committed to
chemical reaction sites (electrodes) and the storage of active materials in improving the performance of ARFBs, and a great deal of experimental
space. The battery power rating (electrode size) and capacity (tank size) and simulation data has been demonstrated. It is very important to
are designed to be relatively independent which provides advantages evaluate the performance data of flow cells reliably and accurately.
such as excellent scalability, flexible modular designability, simple Therefore, the establishment of a scientific and reasonable performance
thermal management, and high security [15]. evaluation method is the basis of battery research. Next, we will discuss
The cost-effectiveness of ARFBs depends on the material cost and the the fundamental attributes and performance parameters that are typi­
cycle life cost. The latter depends on the fading rate and maintenance of cally encountered in conventional battery performance testing methods
active species as well as other components [16,17]. Specifically, as and analyze the main influencing factors.
shown in Fig. 1, the cost of ARFB mainly includes three parts that must
be systematically considered for comparison: active materials (energy 2.1. Charging and discharging test
cost), power stack (power cost), and the operation and management
(O&M) cost of the equipment. The sum of energy and power costs is also The charging and discharging test is a common method to evaluate
referred to the initial investment cost, which is mainly determined by the performance of ARFBs [11]. The protocols for charging and dis­
the bills of material cost [18]. Taking the widely used all vanadium charging tests commonly used in batteries are constant current and
redox flow battery (VRFB) as an example, the system with a 4-h constant voltage settings, of which the processes may repeat for multiple
discharge duration has an estimated capital cost of $447 kWh− 1, in charge/discharge cycles. During the test, various performance parame­
which the electrolyte and membrane account for 43% and 27% of the ters can be obtained, which restrict and influence each other, such as
total cost, respectively [19–21]. For electrolyte, the cost between capacity, coulombic efficiency (CE), voltage efficiency (VE), energy ef­
different systems can be reflected by the chemical price of active species, ficiency (EE), capacity utilization (CU), and capacity retention (CR)
e.g. K2S2, ZnO, K4Fe(CN)6•3H2O is 0.5, 1.3, 2 US$ kg− 1, respectively, [26]. The definition, calculation, and influencing factors of various
while the VOSO4•4H2O is 10.2 US$ kg− 1 which greatly increases the performance parameters are listed in Table 1. For the test of constant
initial cost [22]. current, the battery is first fully charged using a constant current. Once
Therefore, the path to reduce the cost of ARFB is mainly considered fully charged, the battery is then discharged at a constant current. Each
from the following aspects: a) developing low-cost chemical materials reduction/oxidation half-cycle ends when a prescribed cell over­
and battery stacks used in the RFB system; b) improving the physical and potential is reached. However, because of high mass transport over­
chemical properties of the components for better efficiency, e.g. the potentials at the extreme state of charge (SOC) values, as well as the time
lag between the attainment of a particular SOC value in the cell and its
attainment in the reservoirs, less than 100% of the actual capacity is
accessed before the switching occurs. For the test of constant voltage,
where each half-cycle ends only after the reduction/oxidation current
has become negligible, at which point close to 100% of the available
charge is expected to have been extracted. This is strictly true if the
chosen potential limits provide large enough overpotentials to access
essentially 100% of the electrolyte capacity and side reactions at those
limits are negligible. In most studies, flow cells comprising two elec­
trolytes separated by an ion exchange membrane are used for per­
forming cycling tests. The reactant of interest usually acts as the
capacity-limiting side (CLS) with a relatively lower initial capacity of
active species.
The capacity of an ARFB is defined as the amount of charge (Ah)
stored in the electrolyte of CLS, as described in Eq (1).
Fig. 1. Schematic illustration of a typical RFB and cost analysis. (abbreviations: nCF ( )
cost of power system (Cp), cost of balance-of-plant components (Cbop), the Cap = Ah L− 1 (1)
3600
addition to the investment cost to reach the system price (Cadd)).

2
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

Where n is the number of transferred electrons involved in a redox to a slow electrochemical reaction of the electrode. Vohm is ohmic in­
process, C is the concentration of redox species, and F is the Faraday’s ternal resistance, which consists of the contact resistance between the
constant. The capacity is typically measured in ampere-hours (Ah) or ion conducting membrane, electrode, and bipolar plate. Therefore, a
milliampere-hours (mAh). CE is the ratio of discharging capacity to membrane with high ion conductivity is in urgent need [29]. Vtranst is
charging capacity which is used to measure the reversibility of the redox the internal resistance of concentration polarization caused by the
reaction. Decomposition, crossover, and dendrite formation of active diffusion speed of products and reactant is far lower than that of a
substances on the CLS will cause capacity loss, which then affects CE and chemical reaction.
CR. In addition, the self-discharge behavior on the CLS and the occur­ The flow battery is mainly composed of two parts: an energy system
rence of irreversible side reactions will reduce the CE [27]. and a power system. In a flow battery, the energy is provided by the
The energy density of the cell, as one of the most significant pa­ electrolyte in external vessels and is decoupled from the power. The
rameters for EES devices, is determined by Eq 2 power density stands for power per unit area that the battery can supply,
which is calculated by Eq 3
Cap U0
Energy Density = (2) IVcell ( )
μv Power Density = mW cm− 2 (3)
A
Where Cap is the capacity on the CLS, U0 is the voltage of a single cell, μv
where I is the discharge current, Vcell is the output potential, and A is the
is the volume factor (μv = 1+smaller capacity/larger capacity of the
active surface area of an RFB, which is normally determined by the
electrolyte). It can be seen that the capacity is related to the concen­
effective area of the electrode. The power density is affected by many
tration of the redox-active species, and U is determined by the redox
complicated factors such as the battery voltage, conductivities of redox
potential of the active species in anolyte and catholyte. So the energy
species and separators together with the redox kinetics of the active
density is determined by the solubility and the redox potential of the
species on electrodes. Notably, operation factors including temperature,
active species.
flow rate, and flow field will also play a great role in the output of power
VE is the ratio of the mean discharging voltage over the mean
[26].
charging voltage at a constant current. EE is the ratio of discharge en­
ergy over charge energy at a certain current density, which also equals
3. Cost analysis and application of low-cost ARFBs
the product of CE and VE. CU is the ratio of achieved capacity over
theoretical capacity. VE, EE, and CU are all affected by polarization and
3.1. Cost analysis of each component of ARFBs
current density. In general, smaller battery resistance and lower current
density will increase VE, EE, and CU. The capacity decay rate is the
The cost of flow batteries is also an important evaluation criterion.
capacity decay in percentage over a total test duration in terms of cycle
LCOS includes all cost elements in the discharge life of the EES system
number or time which is used to evaluate the cycling stability. The CE,
and quantifies the discounted cost per unit of discharge. Therefore,
VE, EE, CU, and CR should be considered comprehensively when
LCOS is often used to evaluate the economic benefits of different EES
comparing the performance of flow batteries.
technologies. LCOS is calculated using eq. (4).

2.2. Polarization test ∑


N ∑
N
Charging cost
investment cost + O&M cost
(1+r)n
+ (1+r)n
+ end(1+r)
of life cost
N+1

LCOS = n n
(4) [10]
VE and power density are affected by battery resistance. So an in­ ∑
N
EDischarged
crease in polarization can cause a decrease in VE and power density, n
(1+r)n

affecting the performance of the battery [28]. Battery polarization


It assumes all investment costs are incurred in the first year and sums
mainly includes electrochemical polarization, concentration polariza­
ongoing costs in each year (n) up to the system lifetime (N), discounted
tion, and ohmic polarization. Vact is the internal resistance of electro­
by the discount rate (r). The investment cost of RFBs mainly includes the
chemical polarization that deviates from the equilibrium potential due

Table 1
The parameters of RFBs to evaluate the electrochemical performance.
Definition Illustration Influence factor

GCD test
Coulombic CE = Qdischarge/Qcharge The ratio of discharge capacity over charge capacity Proton selectivity of ion exchange
Efficiency (conduction) membranes
Voltage VE = ‾Ed/‾Ec The average voltage of the discharge voltage divided by the Internal resistance of ion exchange
Efficiency average voltage of the charging voltage (conduction) membrane and stack
Energy EE = CE × VE The ratio of discharge energy over charge energy at a certain
Efficiency current density. The product of coulombic efficiency and
voltage efficiency
Capacity CU = Qa/Qt The ratio of actual capacity over theoretical capacity Voltage efficiency
Utilization
Capacity decay =(1-QN/Q1)/N Capacity decay in percentage over a total test duration cycle Comprehensive performance
rate number(N) or time(D).
D 1
=(1-Q /Q )/D
Polarization test
V act The potential deviates from the equilibrium potential Low operating current density
due to the slow electrochemical reaction of the
electrode.
V ohm Contact resistance between ion conducting film, Medium operating current density
electrode, and bipolar plate
V trans The diffusion speed of products and reactants is far High operating current density
lower than that of chemical reactions, concentration
polarization
Power density P––(I × Vcell)/A (mW cm− 2) Power output per membrane area at a certain SOC Battery polarization and voltage

3
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

cost of power and energy systems. O&M cost refers to the cost of oper­ Table 3
ation and maintenance of materials and equipment over its lifetime. The The performance and cost of various batteries (iron-chromium redox flow bat­
charging cost is often affected by electricity price over time, which is the tery (ICRFB), zinc-bromine redox flow battery (ZBRFB), polysulfide-
same for different technologies and thus can be set to zero for conve­ polybromide redox flow battery (PSB)).
nience. End of life cost refers to all the costs incurred to treat and recycle EES Energy Power density Cycle life Cost($
at the end of the life cycle of an EES. The energy cost includes the cost of technology density(Wh (W L− 1 KWh− 1)
L− 1)
the active material, salt, solvent, and storage tanks. In aqueous systems,
due to the low cost of solvent and salt, energy cost is mainly determined VRFB 25–35 <2 10000–13000 400
by the active materials as well as the storage tanks. Therefore, the energy ICRFB 13–30 10–20 >10000 620
ZBRFB 30–65 5000–10000 340–1350
cost of flow batteries with different types of active materials varies
<25
PSB 20–30 <2 2000–2500 150–1000
greatly [18]. For example, the cost of vanadium electrolytes is 10.20 US LIB 200–500 1500–10000 1500–4500 200
$ kg− 1, which accounts for about 35% of the total cost of the battery [22, Pb-Acid 50–90 10–400 2000–4500 100–400
30]. Whereas, iron costs only 0.42 US$ kg− 1 [31], zinc costs only 1.90 Ni–Cd 15–150 80–600 2000–2500 800–2400
Na–S 150–300 140–180 2500–4500 300–500
US$ kg− 1 [32], and sulfur costs 0.29 US$ kg− 1, attributed to their nat­
ural abundance. The power cost consists of Cp, Cbop, and Cadd. Cost of
electrodes, membranes, end plate, collector, bolt, and gasket are and ensures long-term durability [34]. There are some issues with
included in Cp. Power conditioning equipment, controls, sensors, VRFBs, although they can offer distinct advantages compared to other
pumps, pipes, fans, valves, and heat exchangers are included in Cbop. flow battery systems. Due to the high cost of vanadium, vanadium-based
Cadd includes factors such as labor, depreciation, overheads, and security flow batteries lack economic advantages. The cost of vanadium elec­
deposits. Table 2 lists the cost of each component. Apparently, the use of trolyte stands at 10.2 US$ kg− 1, constituting approximately 35% of the
inexpensive electrolytes, electrode materials, and membranes can total battery cost. Similarly, the stack cost, encompassing ion exchange
reduce costs on the condition that the low-cost material exhibits similar membrane and electrode materials, accounts for another 35% of the
stability and efficiency. The performance and cost of various ARFBs and overall cost. Moreover, the initial installation cost is substantial,
other commercial batteries are summarised in Table 3. resulting in a total battery cost ranging from 433 to 640 US$ kWh− 1
Costs can also be reduced by improving the performance of RFBs [22]. Furthermore, the development of this technology has been
such as energy density, power density, and cycle life. The use of redox- hampered by solubility issues and suboptimal redox reaction kinetics.
active species with fast kinetics and low viscosity, electrolyte and Therefore, a lot of research has been carried out to reduce the cost of
membrane with high ionic conductivity, current collector with good VRFBs by changing the composition of the electrolyte, membrane
conductivity, and suitable flow field design can reduce battery over­ properties, electrode materials, etc. To improve the ionic conductivity,
potential to improve power density and energy density. As a result, the researchers often use sulfuric acid, hydrochloric acid, phosphoric acid,
size of both the stack and the electrolyte storage tanks can be cut down, and other concentrated acids as the supporting electrolyte. Wei et al.
which in turn reduces the cost of RFBs [22,33]. As for the life-span cost, [35] assembled silica in situ on the nanofiltration membrane surface and
the selection of redox substances, membranes, and electrodes with high pores to improve the ion selectivity and conductivity of the membrane in
stability can also improve the durability of RFBs, thereby reducing life the VRFBs. The ionic conductivity determines the VE of the cell, and the
cycle costs [16]. selectivity of the membrane determines the CE of the cell according to
the analysis of cell polarization behavior. To achieve high selectivity and
3.2. Key redox pairs for the development of low-cost ARFBs high conductivity of active ions, Li’s team [36] carried out the design of
the composite membrane by introducing the polyamide layer in the
3.2.1. Vanadium-based ARFBs porous membrane. The results show that the porous membrane can
VRFBs are the most developed and widely used flow batteries to date, achieve high conductivity and ion selectivity, which improves the CE
with an energy density of about 15–25 Wh L− 1, an energy efficiency of and power density of the cell. At present, graphite felt and carbon felt are
more than 80%, and a cycle life of more than 200,000 cycles [30]. The commonly used in flow batteries [37,38], however, their catalytic ac­
Schematic diagram and electrochemical profile of the ARFBs are shown tivity of electrochemical reaction of vanadium species is limited. Intro­
in Fig. 2. The VRFBs use the same elements in catholyte (VO2+/VO+ 2) ducing catalysts such as carbon materials [38,39], metals [40–42], and
and anolyte (V3+/V2+), which serves to relieve the issue of crossover, metal oxides [43–45] on the electrodes, and acid [46] or heat treatments
[47] have been investigated to improve electrochemical reaction ki­
Table 2 netics of vanadium species. Oxygen-functionalized single-walled carbon
The relevancy of key indicators, components, and costs of RFBs. nanotubes, as a superior electrocatalyst for the positive Br− /Br−3 couple
Criteria Component Essential factor Cost (US$ m− 2)
of the V/Br battery, can facilitate the reaction kinetics due to the for­
mation of active oxygen-containing functional groups on the surface,
Energy Electrolyte Redox potential Energy cost
which increase the effective surface area and catalyze the electron
density solubility
Power Electrode material Current density 10–30 transfer processes [48].
density (carbon fiber felt) Electrical
conductivity 3.2.2. Zinc-based ARFBs
Active site The study of zinc-based batteries can be traced back to the 70s of the
Membrane Selectivity 500
(fluorinated) Ionic conductivity
20th century, which mainly includes zinc-iron (Fe3+/Fe2+, 0.77 V vs
Membrane (fluorine- Stability 25 SHE), zinc-air, zinc-iodine (I− /I3− , 0.536 V vs SHE), zinc-nickel, zinc-
free) bromine (Br2/Br− , 1.087 V vs SHE), zinc-organic RFBs, etc [53–56].
Stack (End plate, × 47 Zinc-based flow batteries have a much lower chemical cost (1.9 US$
collector, bolt)
kg− 1) than VRFBs. Besides inherent safety and stability, the materials in
Pipeline (Piping and flow rate Fixed cost
Pumps) zinc-based ARFBs are abundant, which has always been the focus of
Cycle life × Stability Inverse research on EES technology [30]. Unlike other redox active species, zinc
proportional anodes can be applied over a wide pH range.
Others O&M cost Corrosivity and pH Determined by According to the pH environment, zinc-based ARFBs are divided into
Replacement interval cycle stability the system
alkaline, acidic, and neutral systems. The reactions of the zinc anode in
cost

4
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

Fig. 2. The schematic diagram and electrochemical profile of the various ARFBs. (a) Demonstration (b) CV curve (Reproduced with permission from Ref. [49]) (c)
Voltage profile (Reproduced with permission from Ref. [20]) of VRFB. (d) Demonstration € CV curve (Reproduced with permission from Ref. [50]) (f) Voltage profile
(Reproduced with permission from Ref. [51]) of ZIRFB. (g) Demonstration (h) CV curve (i) Voltage profile (Reproduced with permission from Ref. [31]) of AIRFB (j)
Demonstration (k) CV curve (l) Voltage profile (Reproduced with permission from Ref. [28]) of polysulfide-polyiodide redox flow battery (PSIRFB) (m) Demon­
stration (n) CV curve (o) Voltage profile (Reproduced with permission from Ref. [52]) of DHPS-iron RFB.

different environments are as follows [57]: state, the negative electrode side is oxidized from zinc metal to basic
zincate solution, while the positive side is reduced to ferrocyanide by
2+ −
(5)
ferricyanide. Compared to ferricyanide in alkaline media, Fe2+/Fe3+
Neutral or acidic : Zn + 2e ⇌Zn
and Br− /Br2 in acidic electrolytes have a higher solubility and thus
Alkaline : Zn(OH)2−4 + 2e− ⇌Zn + 4OH− (6)
potentially higher energy density. However, in an acidic zinc-iron redox
In alkaline zinc-based ARFBs, zinc and ferricyanide are generally flow battery (ZIRFB), the acidity of the solution will cause the corrosion
used as active substances in the negative electrolyte and positive elec­ of zinc, the hydrolysis of the Fe2+/Fe3+, and hydrogen evolution re­
trolyte, respectively [58,59], which offer relatively low electrolyte cost actions (HER). For acidic systems with buffer solution (e.g. HAc/NaAc),
and a high open-circuit voltage (OCV) of up to 1.74 V. In the discharged the pH environment of the electrolyte can be maintained between 2–6,

5
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

and high CE can be achieved [60]. The poor reversibility and insufficient stability to prevent penetration by the dendrite [76], improving the
kinetics of Br− /Br2 lead to a low power density although ZBRFB has a service life of the battery. The modification of the electrolyte includes
high energy density and a low cost. In addition, the formation of zinc the optimization of zinc salt and the use of additives, which can reshape
dendrite will limit its cycling stability [61,62]. Compared with alkaline the solvation structure and inhibit the zinc dendrites. Additives such as
and acidic systems, zinc generally forms an ARFB system with organic nicotinamide [32] can reshape the solvation structure of Zn2+, and N, N,
substances in neutral environments, such as zinc-TEMPO, and zinc-poly N′ N’-Tetra(2-hydroxyethyl) ethylenediamine [77] slows down the dy­
(TEMPO), which is mild and has lower requirements for ion membranes namics of zinc deposition to prevent the formation of dendrites.
and other components. The neutral ZIRFB also has been reported. ZnBr2
and FeCl2 were used as active species and Fe2+ inhibits hydrolysis by 3.2.3. Iron-based ARFBs
complexing glycine [63]. Furthermore, the complexation of bromide Due to the natural abundance and low cost of iron (0.42 US$⋅Kg− 1),
ions to zinc ions in neutral electrolytes can greatly enhance the revers­ iron-based flow batteries have received widespread attention in recent
ibility of zinc redox pairs [64]. years [31]. The first iron-based flow battery was proposed in the 70s of
Considering that the area capacity of zinc-based ARFB is limited by the 20th century, with Fe (III)/Fe (II) and Cr (III)/Cr (II) serving as the
electrodes due to the plating/stripping of the zinc, the energy is no positive and negative active components, respectively and HCl as the
longer decoupled from power. In addition, zinc dendrite/accumulation supporting electrolyte, which exhibited the battery voltage of 1.18 V.
is also one of the key problems [65]. In the process of charging, the zinc The historical developments and chemical price comparison of various
dendrite may puncture the membrane and cause the battery short cir­ ARFBs are shown in Fig. 3. However, many factors such as poor reaction
cuit. During the discharging process, the accumulation of unstripped reversibility, HER side reactions, and crossover of active components
zinc on the membrane will increase the cell resistance. The accumula­ severely affect the EE of ICRFBs and pose great safety issues.
tion issue of zinc dendrites in alkaline conditions is more serious than in In recent years, several low-cost and high-stability iron-based redox
acidic and neutral conditions. The increasing concern for the above pairs have been developed including all-iron systems, organic-iron sys­
problems has driven the research on the surface modification of elec­ tems, zinc-iron systems, and sulfur-iron systems, etc. There are three
trode materials, the rational design of membranes, the concentration main trends: First, the use of complex ligands, such as phenanthroline,
optimization of electrolyte solution, and the introduction of additives. triethanolamine, glycine, etc [82–84]; Second, the addition of polar
In terms of the modification of electrode materials, the direct contact solvents that increase the solubility of iron complexes or reshape the
between the bulky electrolyte and the active zinc anode is prevented to solvation structure [31]; Third, manipulation of the anions to regulate
increase the over-potential of HER and block the charge transfer path the solubility and kinetics [85]. The use of suitable complexing agents
from metal zinc to water molecules. To eliminate the zinc dendrites, we can effectively overcome the problem of iron ions precipitating as ferric
should promote the transfer of ions and electrons on the anode surface, hydroxide at pH > 2. At the same time, the use of organic complex li­
achieve a uniform electric field and concentration distribution, reduce gands can change the redox potential and the solubility of the active
the local current density, and reduce the concentration or diffusion species. A variety of organic ligands have been studied. For instance,
polarization. Based on these ideas, the mitigation strategies related to Robert F. Savinell [86] studied the ratio of glycine to Fe3+/Fe2+ as a
the anode mainly include the construction of protective layer on the function of the pH of solution, showing that the optimized ratio of
surface of the electrode material and the structural design of the zinc glycine to Fe3+ is 1:1, at which high ligand solubility and a high redox
metal anode [66,67]. The protective layer can inhibit the direct contact potential can be achieved at pH 2. Kwon [87] reported 2,2-bis(hydrox­
of the electrolyte with the zinc anode and inhibit the corrosion of the ymethyl)2,2,2-hyponitrotriethanol (BIS-TRIS) as an organic ligand of Fe
zinc group, especially in the weak acidic or high-concentration alkaline (BIS-TRIS) compound and constitutes an AIRFB coupled with [Fe
electrolyte. At present, the reported protective layer materials mainly (CN)6]3− catholyte, which does not exhibit any side reactions after 250
include metal (In and Pd) [66,68,69], metal compound (ZnS, TiO2, cycles, exhibiting a capacity of 11.7 Ah L− 1, CE and CR of 99.8 and
CaCO3, ZrO2) [70,71], carbon material (activated carbon, carbon 99.9%, respectively.
nanotubes, MOFs) [67,72], etc. In terms of ionic membranes, it mainly Among all the ligands, cyanide exhibits great properties with
aims at modifying the current commercial membrane or inducing Zn inherent chemical structure stability and facile kinetics, which makes
uniform deposition to circumvent the challenges of zinc dendrites [65, ferricyanide one of the most popular redox species for catholytes.
73–75]. In addition, the composite film has a stronger mechanical However, the energy density of ferricyanide-based batteries is relatively

Fig. 3. The historical developments and chemical price comparison of various ARFBS. Reproduced with permission from Refs. [78–81]. (Abbreviations: AIRFB for all
iron redox flow battery).

6
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

low due to the limited solubility of ferricyanide (<0.8 M for K4[Fe(CN)6] 3.2.4.1. Liquid-gas polysulfide-based hybrid ARFB. Li et al. [92] sum­
and <0.6 M for Na4[Fe(CN)6]). Based on the redox-targeting reaction of marized the existing low-cost flow batteries. Among the known redox
[Fe(CN)6]4-/3- and Prussian blue (PB), Wang Qing’s team [88] designed active substances, sulfur has the lowest cost per stored charge, which is
a redox-targeted flow battery with [Fe(CN)6]4-/3- as the redox mediator only higher than water and air. This group then coupled oxygen re­
and PB as a solid energy storage material to break the solubility limi­ actions (oxygen evolution reaction (OER)/oxygen reduction reaction
tation of ferricyanide, which greatly improve the capacity of the system. (ORR)) with polysulfide chemistry to construct an exceptionally
In addition, the solubility of ferricyanide can be improved by cation low-cost system. With a solid electrolyte (LiSICON) used as the separator
exchange strategies. For instance, the solubility of Li4[Fe(CN)6] and of the flow battery, an acid-alkaline hybrid sulfur-air system was
Li3[Fe(CN)6] is 2.32 M and 2.70 M, and the solubility of (NH4)4[Fe investigated with 0.5 M Li2SO4/0.5 M H2SO4 and 1 M Li2S4/1 M LiOH as
(CN)6] and (NH4)3[Fe(CN)6] is1.60 M and 1.92 M, respectively. catholyte and anolyte, respectively [92]. Based on the different con­
centrations of sulfur species, the energy density of the battery is 30~145
3.2.4. S-based ARFBs Wh L− 1 with the chemical cost of the active material is ~1 US$⋅kWh− 1,
Sulfur is abundant on the earth, usually in the form of elemental which is the lowest level.
sulfur, sulfide, and sulfate. Among the existing active materials, sulfur
has a lower cost (0.29 US$ kg− 1). Elemental sulfur can react with soluble 3.2.4.2. Solid-liquid polysulfide-based hybrid flow batteries. To improve
polysulfides in aqueous solution to generate other polysulfides S2− x (x = the energy density, some alkali metals with lower potential were paired
2–8). Sulfur species, i.e. polysulfides can achieve reasonably high with a highly soluble negative electrolyte and separated by a glass-
theoretical energy densities with low chemical costs due to their excel­ ceramic membrane to avoid crossover of the active species. Lu’s [93]
lent solubility in water and organic solvents, which can satisfy the group explored the electrochemical interactions of polysulfides and io­
fundamental requirements of large-scale energy storage devices. In dides, reducing the crossover of polyiodide ions and removing the use of
addition, the redox process between various polysulfides is a multi- ion-selective ceramic membranes. In addition, the team developed a
electron reaction, thus delivering high theoretical capacity. Combined highly reversible low-cost sulfur/manganese ARFB. The positive active
with the above application advantages, polysulfides have been widely species of Mn2+/MnO2(s) has a high theoretical potential and a high
used in energy storage systems, including metal-sulfur batteries, solar theoretical capacity (616 mAh g− 1) [94]. The battery used low-cost
cells, and RFBs. active materials and circumvented the problem of zinc dendrites in
All-liquid polysulfide-based ARFBs. The earliest research on the Zn/MnO2 battery. The cycling stability under high areal capacity
polysulfide-based flow batteries dates back to the 1980s [89]. Poly­ (50–100 mAh⋅cm− 2) is greatly improved with the capacity retention rate
sulfide was paired with bromine, which has a high open-circuit voltage of 98% after 75 cycles at 50 mA cm− 2, which is much higher than that of
(1.35 V). The chemical reaction and standard potential were shown in Zn/MnO2 flow batteries.
equations (7) and (8). During the charging process, Br− was oxidized to
Br2 and complexed simultaneously to form Br−3 , while polysulfide was 3.2.5. Organic-based ARFBs (AORFBs)
reduced to a species with lower valence. The discharge process is the Electroactive organic molecules take advantage of natural renew­
opposite. ability, flexible designability, and sustainable production compared to
present commercial inorganic materials with uneven resource distribu­
Cathode: Br−3 + 2e− ⇆ 3Br− E⊝ = 1.09 V vs. SHE (7)
tions [26]. The quinones were first reported as organic electroactive
Anode: S2−
4

+ 2e ⇆ 2S22 ⊝
E = − 0.48 V vs. SHE (8) material in an aqueous solution by Smith et al., in 2007 [95]. Afterward,
a series of organic molecules began to be used in ARFBs, mainly
However, the self-discharge behavior caused by the crossover of the including hydroquinone and quinone based on carbonyl (C– –O) bond
active species is more serious, thus, the capacity decay is relatively fast. reaction, alloxazine [96], and phenazine [96,97] based on C– –N reac­
The occurrence of HER and the precipitation of sulfur seriously affect tion, viologens [98] and nitroxide radicals [99] based on doping reac­
cycle stability. In addition, the solubility of Br− is low (0.21 M), as well tion. In recent years, soluble organic electroactive materials have been
as the toxicity of bromine pollutes the environment, which has great widely regarded as promising candidates for low-cost and sustainable
safety risks. Compared to Br2/Br− redox pairs, highly soluble I− /I−3 pairs ARFBs due to their facile molecular engineering with various functional
have a faster and highly reversible kinetic process [90], and the reaction groups to adjust the solubility and redox potential [26].
process is shown in Equation (9). However, the energy density of AORFBs is limited by the solubility
I−3 + 2e− ⇆ 3I− E⊝ = 0.54 V vs. SHE (9) and redox potential of organic species in an aqueous solution, which has
been the main obstacle limiting practical application. By strategically
Polysulfide/polyiodide flow batteries exhibit lower initial invest­ modifying the structures of molecules with hydrophilic hydroxyl or
ment costs (polysulfide/iodine: 85.4 US$ kWh− 1) under the same energy sulfonate groups, the solubility of organic redox-active molecules in
density. The PSIRFB was assembled using high-concentration KI and aqueous electrolytes can be significantly promoted [52,100,101]. For
K2S2. The battery achieved high capacity utilization. Without capacity instance, 2,6-dihydroxyanthraquinone (2,6-DHAQ) exhibits a solubility
decay, 50 stable cycles were realized [80]. over 0.6 M in 1 M potassium hydroxide (KOH) solution and 7,8-dihy­
Wei Wang et al. [25] reported a polysulfide/potassium ferricyanide droxyphenazine-2-sulfonic acid at its near-saturation concentration
flow battery with an open-circuit voltage of 0.91 V. The neutral ferri­ (1.4 M) exhibits an operating voltage of 1.4 V with a reversible anolyte
cyanide and polysulfide were used as catholyte and anolyte, respec­ capacity of 67 Ah L− 1 [52]. It is worth noting that the substituting
tively. Cheap redox materials and battery components with longer location, number of functional groups, and the species of its
service life can reduce investment costs. Organic compounds have the counter-cation contribute to different solubility [102,103]. Alloxazine
advantages of structural diversity and flexible structural regulation, modified by carboxylic acid (COOH) groups is proven to increase the
posing great promise for energy storage applications. Taking the mo­ aqueous solubility and deliver a desirable shift in the redox potential
lecular of viologen as an example [91], the molecule shows strong [104]. With the further introduction of carboxylic acid-doped carbon
electrostatic attraction to polysulfides and inhibits the growth of lithium nanotube (CA-CNT) catalyst for increasing the reactivity, the redox
dendrites, which helps to improve performance. During the redox pro­ reactivity of alloxazine-COOH can be improved greatly because of the
cess, the molecule undergoes two single-electron reactions that provide role of COOH within alloxazine-COOH as a proton donor, the fortified
high energy density. hydrophilic attribute of alloxazine-COOH, and the increased number of
active sites. Additionally, the use of additives, such as ethylene glycol

7
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

[105] and choline hydroxide [102] is another effective strategy to is shown in Fig. 4.
improve solubility. Different from the abovementioned redox organic materials involved
In terms of the redox potential governed by the energy levels of their in the protonation process in redox equilibrium, viologens are used
frontier highest occupied molecular orbital (HOMO) and lowest unoc­ preferentially in neutral media thus the redox potential is independent
cupied molecular orbital (LUMO), it can be effectively tuned through of pH. Radical species are involved as intermediates or products of the
structural functionalization. Specifically, the addition of electron- reduction reaction. Methyl viologen (MV) dichloride, considered as
donating groups leads to a negative shift of the redox potential while archetypical 4,4′ substituted bipyridine, was first reported by Wang and
the addition of electron-withdrawing groups leads to a positive shift Liu in RFBs with high solubility in a neutral aqueous solution (3 M in 1.5
[106]. The redox potential of the organic molecules can also be influ­ M NaCl) [111]. When paired with nitroxide radicals or ferrocene de­
enced by the solution environment due to the variation of the pH and rivatives as catholytes, the constructed neutral AORFBs exhibit excellent
supporting electrolytes. For example, 2,6-DHAQ was reported by Lin electrochemical performance [112,113]. Liu et al. [13] designed a
et al. used as the negative active material coupled with K3[Fe neutral AORFB composed of MV and ferrocyanide to achieve integrated
(CN)6]/K4[Fe(CN)6] catholyte in basic KOH supporting electrolyte to salt water desalination and energy storage. This technology offers op­
replace the toxic Br2/Br− in previously reported system. These OH− portunities to address both renewable energy production and water
groups on anthraquinone core structure are deprotonated to promote scarcity through one device. TEMPO is a stable radical that can undergo
solubility and get greater electron donation capacity, which contributes fast reversible redox transformation to corresponding oxoammonium
to a 47% increase in OCV over previously reported systems, higher salt. But it still suffers from severe crossover due to the smaller size than
solubility (>0.6 M at room temperature) and comparable kinetic rate in the ion-conductive channels in ion exchange membranes [111,114,
1 M KOH to that in the previously reported acid solvent environment. 115]. Recently, Feng et al. [116] synthesized a class of molecularly
However, a key deficiency of most reported organic chemistries is engineered ionic liquid-mimicking TEMPO dimers to address this chal­
the utterly impractical capacity decay rates (>0.1 %/day) for RFB in­ lenge, one of which coupled with viologen derivatives demonstrated
stallations to reach the projected value due to the decomposition of their unobservable capacity decay at 4 M electron concentration over 96 days
organic constituents. Thus, the most challenging but imminent object in benefitted from membrane compatibility and a strong water coordina­
developing AORFBs is the designation of highly stable redox-active tion environment.
compounds at both discharged and charged states even at high con­ Last but not least, considering the self-decomposition characteristics
centrations (>0.5 M). The degradation of organic active species can be of organic molecules, physical organic chemistry, and advanced char­
attributed to both the type and structure of organic molecules and acterizations that can get fundamental and practical insight at the mo­
exogenous factors (pH [107], temperature [108], oxygen tightness lecular level is urgently required to further understand the degradation
[109] et al.). After the introduction of the ether-linked alkyl chains with mechanism of redox active molecules, thus developing higher-lasting
solubilizing carboxylate functional groups onto an anthraquinone core, devices for large scale energy storage [117,118].
4,4’-((9,10-anthraquinone-2,6-diyl) deoxy) dibutyrate was six times
more soluble than 2,6-DHAQ at pH value of 12 and affords the lowest 4. Main factors of loss and the mitigation strategies
ever reported capacity fade (<0.01%/day) for RFBs without rebalancing
[81]. Moreover, solvation regulation can also be a facile strategy to Electrolytes, membranes, and electrodes serve as pivotal constituents
extend the lifetime of organic active species. Xu et al. [110] incorporated for RFBs. Their intrinsic characteristics significantly determine the
tetramethylammonium cations in the supporting electrolytes to interfere overall battery performance (Fig. 5). Consequently, interest in the
with the solvation structure of DHAQ2− /DHAHQ4− anions, thereby investigation of these materials has been steadily escalating. This section
deactivating the chemical or electrochemical reduction of DHAHQ4− provides a comprehensive assessment of the distinctive properties of
and subsequent side reactions. The status of various redox flow batteries these components and an overview of current methods employed for the
improvement of CE and VE.

4.1. Key materials that affect CE

Electrolyte and CE: The CE can be affected by the self-


decomposition caused by the instability of the active substance and
the side reactions in the electrolyte. Self-decomposition is particularly
common in organic-based flow batteries, such as the bimolecular
degradation of viologen and its derivatives. The mechanism of self-
decomposition of organic species includes disproportionation, nucleo­
philic substitution, disproportionation, dimerization, and polymeriza­
tion, etc. The self-decomposition of active substances may lead to
permanent loss of active substances, resulting in capacity decay and
reduction of CE. Liu et al. [119] reported a highly stable AORFB using 1,
1′-bis[3-(trimethylammonium)propyl]-4,4′-bipyridinium tetrachloride
((NPr)2V) as the anolyte and 4-trimethylammonium-TEMPO chloride
(NMe-TEMPO) as a catholyte, with a high voltage of 1.38V and a capacity
retention rate of 99.995% per cycle. The viologen radicals can undergo a
dimeric degradation process. When compared to previously reported
methyl viologen (MV), [(NPr)2V]+ exhibited extraordinary stability
than [MV]+, which is explained by reduced dimerization caused by
higher charge repulsion from two extra positively charged propyl
ammonium pendant arms, as shown in Fig. 6c. Thus, its capacity
retention rate per cycle can reach 99.995%, and CE is close to 100%.
Fig. 4. The status of various RFBs (VRFB, ICRFB, AIRFB, ZIRFB, PSIRFB) based Side reactions and decomposition of solvents are also a major chal­
on radar diagram of key indicators. (Concentration, Power cost, Energy cost, lenge, namely HER and OER in aqueous systems. For active substances
Open circuit voltage, Energy density). with a potential exceeding the voltage window of water splitting, the

8
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

Fig. 5. Analysis of capacity loss mechanism in terms of efficiency in an RFB. (a) Main factors affecting CE and attenuation analysis; (b) Main factors affecting VE and
attenuation analysis.

Fig. 6. Various strategies to deal with low CE (a) The ammonium chloride-supported zinc-iodine redox flow battery (AC-ZIFB) using ammonium-based electrolyte.
Reproduced with permission from Ref. [53]. (b) Design concept of the bifunctional electrocatalysts. Reproduced with permission from Ref. [120]. (c) Proposed
favorable dimerization of [MV]+• and unfavorable dimerization of [(NPr)2V]+•. The red double arrows indicate charge repulsion between the pendent ammonium
groups. Reproduced with permission from Ref. [119]. (d) Charge-reinforced ion-selective membrane repels the negatively charged active materials (S2−
x and Ix ) from

crossover and alleviates free water migration. Reproduced with permission from Ref. [22]. (e) Microstructure of Paper Reinforced sulfonated poly ether ketone
(SPEEK) Membrane before and after Thermal Cross-Linking. Reproduced with permission from Ref. [121]. (f) The structural superiority of mixed matrix membrane
with highly anti-alkali microporous hollow spheres (denoted as DM-HM) under alkaline electrolyte. Reproduced with permission from Ref. [29]. (g) Schematic
illustration of Zn deposition induction mechanism. Reproduced with permission from Ref. [122]. (h) Multifunctional carbon felt-based electrode (NTCF) with
abundant N-rich defects for high-power density and long-cycle life ZBRFBs. Reproduced with permission from Ref. [62]. (i) The use of non-solvent-induced phase
separation (NIPS) with easily controllable parameter to create varieties of porous microstructures. Reproduced with permission from Ref. [123].

HER and the OER exert considerable influence, leading to reduced CE material and causes a decrease in CE. This problem can be resolved with
and alterations in pH levels. In response to these challenges arising the proper design of electrolytes. Chen et al. [53] developed an AC-ZIFB
within the electrolyte, a substantial array of enhancements has been based on NH4I/NH4I3 redox pair design. The design of AC-ZIFB is shown
undertaken to ameliorate its performance. For instance, Jang et al. [120] in Fig. 6a. The decoupling design allows electrolyte and counterions to
reported a promising electrocatalyst consisting of Ketjenblack (KB) exist in anolyte and catholyte, respectively. NH+ 4 can reduce the for­
carbon with embedded bismuth nanoparticles (Bi–C). As shown in mation of zinc dendrites and promote the reaction kinetic process of
Fig. 6b, Bi nanoparticles are uniformly incorporated into KB carbon by a electrodes. The anionic properties of I− , I−3 , and I2Cl− prevent the
simple reduction process, and due to the synergistic effect of Bi and KB, migration of cations through the cation exchange membrane. A constant
the formation of BiHx delays the reduction of H+ to inhibit hydrogen current cycle test with a charge-discharge current of 20 mA cm− 2 was
evolution and provides an active site to enhance the Cr2+/Cr3+ redox performed for both AC-ZIFB and conventional zinc-iodine redox flow
reaction. ICRFB with Bi–C catalyst as the negative electrode exhibited an battery (ZIFB). The CE of AC-ZIFB remains at 99% over 100 cycles, while
average CE of 97.4% during a charge-discharge cycle at room conventional ZIFB has only 90% CE in about 50 cycles.
temperature. Membrane and CE: Crossover and damage of the membrane may
In hybrid flow cells, the formation of dendrites depletes the active cause the reduction of CE. Crossover mainly refers to the transport of

9
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

active substances through the membrane leading to self-discharge or strong competition to the market’s present fiber-based electrodes.
irreversible reaction. Efforts have been made to improve the CE by
weakening the crossover of active species through the membrane. For 4.2. Key materials that affect VE
instance, Lu’s group [22] has created a charge-enhanced ion-selective
membrane that allows polyvinylidene fluoride (PVDF)-bound KB to VE is the ratio between the average discharging voltage and the
enter the Nafion membrane and reduce expansion and water absorption. average charging voltage and it is mainly affected by battery resistance
It delivered a strong polysulfide/polyiodide absorption capability to and current density. Voltage hysteresis between the charge and
capture and accumulate negatively charged anions, which electrostati­ discharge caused by the potential difference between the two redox
cally repels active anions in the electrolyte, and thus decreases crossover molecules inevitably deteriorates the VE of RFB. The lower battery
and migration of water/OH− , as shown in Fig. 6d. PSIRFB exhibited a resistance and smaller current densities usually deliver higher VE.
low capacity decay rate (0.0004% per cycle) over 1200 cycles, as well as Additionally, the state of charge, operating temperature, and state of
CE>99.9% benefitted from the high hydrophobicity of PVDF, the high health of the battery have been identified as the main parameters that
absorption strength, and high conductivity of KB for influence the VE [124]. Furthermore, VE can also be significantly
polysulfides/polyiodides. improved through the redox targeting reactions [125], the use of cata­
SPEEK membranes are considered potential candidates to replace lysts, and the design of the flow fields [126].
expensive Nafion membranes due to their high selectivity of protons. To Electrode and VE: The electrode provides the active sites for redox
overcome the poor mechanical/chemical stability of SPEEK membranes reactions that occur in RFBs and can alter the kinetics of redox pairs. In
in VRFB, Xi et al. [121] adopted a novel method by employing com­ addition, the electrode structure can affect the concentration over­
mercial paper (copy paper, rice paper, and filter paper) as the SPEEK potential, which in turn affects the VE. Insufficient redox reaction ki­
scaffold and subsequently crosslinking the SPEEK by heat treatment. The netics on the electrode lead to activation polarization, which further
fibers of these commercial papers contain abundant hydroxyl groups, reduces VE. The main strategy for increasing electrode activity is to
which can interact with the sulfonic acid group of SPEEK to produce decorate the electrode with carbon-based or metal-based [127] de­
hydrogen bonds (Fig. 6e), thereby enhancing the mechanical properties rivatives that can increase the active surface area or provide some cat­
of the membrane. Compared to the CE of the benchmark Nafion 115 alytic activity. Based on semiconductor theory, Zai et al. [128] designed
membrane (92.8%), the optimized rice paper-enhanced SPEEK mem­ CoS2/CoS heterojunction nanoparticle materials using the hydrothermal
brane demonstrates an improved CE of 98.3% even at a high current method. The heterojunction material contacting interface forms two
density of 120 mA cm− 2. Li et al. [29] developed a cost-effective inverted spatial charge regions and excites a built-in electric field. CV
DM-HM. The daramic membrane used is a polyethylene (PE) porous data shows that heterojunction materials (CoS2/CoS) exhibit minimal
membrane, which can not only trap the highly anti-alkali inorganic redox peak voltage separation (0.10 V), compared to graphite felts and
components in the PE matrix but also maintain the porous structure of monometallic sulfides. The uneven distribution of the charges within the
the PE polymer, thus, improving the wettability of the electrolyte. As a material enhances the charge transfer activity, and the reaction revers­
result, DM-HM has strong mechanical strength to resist the growth of ibility of the redox active pairs can be significantly increased by the
zinc dendrite and excellent chemical and mechanical stability. At 80 mA adsorption of charged ions and the improvement of the electron-transfer
cm− 2, the alkaline ZIRFB displayed 500 stable cycles with a CE of 98.6%. process. The PSIRFB using CoS2/CoS heterojunction materials as elec­
Electrode and CE: For the hybrid RFB, dendrite formation and loss trodes was found to achieve stable cycling performance at 10 mA cm− 2
of solids are the two main problems that account for the low CE. The with a CR rate of up to 84.5%. The peak power of the CoS2/CoS electrode
formation and detachment of dendrites consume active substances, battery reaches 86.2 mW cm− 2, which is much higher than the peak
which in turn leads to low CE, capacity attenuation, and even battery power of the graphite felt electrode (8.9 mW cm− 2). The nitrogen-doped
lapse. Insufficient dissolution of solids resulting from redox reactions graphene (N-G) reported by Barun et al. [129] also revealed the sur­
involving solid-phase formation can cause electrode passivation and prisingly beneficial effects on three types of Hybrid RFBs, namely
reduced CE. At present, the main improvement approach for electrodes hydrogen/vanadium cells, hydrogen/manganese cells (RHMnFC), and
includes thermal/chemical treatments, active site modifications, and polysulfide/air cells. Fig. 7b shows the polarization and power density
three-dimensional design. For instance, a multifunctional tin modified curves of RHMnFC. Using advanced materials such as nanomaterials can
three-dimensional carbon felt anode host material (SH) was engineered also effectively improve the efficiency of ARFB. Park et al. [130] re­
as an anode by Li et al. [122]. The adsorption energy of Sn to zinc atoms ported an aqueous mixed zinc-manganese ARFB with a
is higher than that of the pristine carbon felt host, which can provide double-membrane-based three-electrolyte using an alkaline zinc redox
sufficient zinc nucleation sites and low nucleation overpotential for zinc couple (Zn/[Zn(OH)4]2− ) and an acidic manganese redox couple
deposition. The strong interaction between Sn and Zn can effectively (Mn2+/MnO2(s)). The CV curves of Mn2+/MnO2 on different electrodes
avoid the aggregation of deposited Zn on the surface of the Sn layer and indicated that the Ipa/Ipc values on the bismuth nanoparticle embedded
avoid the fall-off of Zn from the electrode. With SH electrode design, carbon felt (BCF) with the metal ionic catalysts (MIC) (BCF-MIC) elec­
ZBRFB steadily demonstrated 120 cycles with an average CE of 97.2%. trode increase significantly, evidencing the enhancement of electro­
Similarly, Li et al. [62] prepared a NTCF, which shows high catalytic catalyst on the redox kinetics of Mn2+/MnO2 and exhibiting an EE of
activity for the Br2/Br− reaction for ZBRFB. At the same time, the 89.8% at a current density of 15 mA cm− 2 (Fig. 7c). Additionally,
reduced zinc atoms tend to be deposited uniformly in N-rich defects until N-doped vertical graphene is in-situ grown on graphite felt via a
these defects are filled, which results in stable operation of ZBRFB at an metal-free chemical vapor deposition method, which exhibits a high
unprecedented high current density of 180 mA cm− 2 with a CE of specific surface area and remarkable catalytic activity [131]. The
97.25% by using NTCF electrodes (Fig. 6h). In addition, under the high vertical-standing nanostructure promotes the accessibility of vanadium
area capacity of 40 mAh cm− 2, the zinc symmetrical flow battery can ions to electrode/electrolyte interfaces, effectively decreasing the mass
steadily operate for 140 cycles at 80 mA cm− 2 exhibiting a CE of 98.93%. transport resistance of active species. The EE is significantly higher than
As a general technique for creating tunable porous materials suited for that of the original electrode.
RFB electrodes, NIPS has been proposed [123]. It is shown that elec­ Membrane and VE: The selectivity of the membrane is highly
trodes with distinct microstructures and porosities can be produced with correlated with the CE of the battery, while the conductivity of the
different relative concentrations of scaffold-forming polyacrylonitrile to membrane is related to VE. The low conductivity of the membrane leads
pore-forming poly(vinylpyrrolidone). NIPS-derived electrodes made of to a decrease in VE, while the increase in conductivity often leads to a
continuous, interconnected porous networks show great promise for decrease in selectivity. Thus, the trade-off between conductivity and
RFBs and provide a solid foundation for material innovation, offering selectivity is especially important when improving the properties of

10
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

Fig. 7. Various strategies to deal with low VE. (a) The CV of the catalysts deposited on glassy carbon in 4 mM NaI + 0.5 M NaCl solution and 0.06 M Na2S + 0.02 M S
+ 0.5 M NaCl solution at a scan rate of 50 mV s− 1. Reproduced with permission from Ref. [128]. (b) Polarization and power density curves using the same electrolytic
conditions with the exception of a higher hydrogen flow rate of 100 mL min− 1. Reproduced with permission from Ref. [129]. (c) CV of the catalytic effect of the
treated carbon felt with various methods. Reproduced with permission from Ref. [130]. (d) Schematic illustration of the membrane containing a hydrophobic poly
(vinylidene fluoride-total hexafluoropropylene) (PVDF-HFP) impregnated into a bacterial cellulose (BC) aerogel scaffold and interconnected proton conduction
pathways of the BC/PVDF-HFP membrane. Reproduced with permission from Ref. [132]. (e) The proposed degradation mechanisms of anions exchange membranes
(AEMs). Reproduced with permission from Ref. [133]. (f) Interleaved flow field (IFF) used in flow batteries. Reproduced with permission from Ref. [134]. (g) Ef­
ficiencies (CE and VE) of the PSIRFB cell under flow-mode in comparison with the VE achieved under the static mode at the same current density Reproduced with
permission from Ref. [80].

membrane. Improving the conductivity and chemical stability of the the effectiveness of higher flow rates on battery performance.
membrane is of great significance to improve VE. Zhu et al. [132] re­ In addition, temperature may also impact the efficiency of ARFBs.
ported a highly conductive and selective RFBs membrane containing a Temperature affects the thermodynamics and kinetics of electro­
PVDF-HFP impregnated into a BC aerogel scaffold. The high proton chemical reactions. Taking VRFB for example, higher temperature can
conductivity of the membrane stems from the intrinsic highly hydro­ accelerate the kinetics of electrochemical reactions, reduce the viscosity
philic network of BC fibrils, which serves as an osmotic pathway for of electrolytes, enhance the H+ transmembrane transport, and improve
proton conduction. Meanwhile, its selectivity originates from the hy­ the solubility of V2+, V3+, and VO2+. In addition, high temperature can
drophobic lamination on the BC scaffold, which functions as a barrier to reduce the OCV, boost the crossover, speed up the HER, and promote the
prevent the migration of vanadium ions and minimize crossover. In formation of V2O5 precipitation. Except for VRFBs, Lu’s team [135]
addition, both BC/PVDF-HCP showed higher conductivity by comparing performed the test of the Sn(OH)2− 6 negolyte at 60 C. By increasing the

with Nafion 115 membranes at different temperatures. At a current temperature, the sluggish kinetics was improved. Furthermore, as the
density of 100 mA cm− 2, VRFBs using BC/PVDF-HFP membranes can temperature increased, the resistance and internal resistance of the
operate 300 cycles with an average CE of 97.56%, VE of 81.56%, and EE battery decreased [136]. Ding et al. [137] conducted redox targeting
of 79.49%. To balance the ionic conductivity and vanadium perme­ flow battery tests at different temperatures and uncovered that the uti­
ability, Chen, etc. [133] have designed polysulfide-based side-­ lization of solid materials increased with the increase of the activity of
chain-type AEMs (the side-chain-type AEMs having PTA groups and the redox species at high temperature.
side-chain-type AEMs (PSf-c-PTA) with QA moieties) with a lower area Based on the potential cost-effectiveness and the development stage,
resistance and a relatively higher vanadium permeability, due to its the roadmap of ARFBs is described in Fig. 8. The most maturely devel­
higher water uptake. The VRFB with PSf-c-PTA membrane delivered a oped is VRFB, which is currently in the stage of commercialization.
CE of 98.4% and an EE of 84.3% at a current density of 120 mA cm− 2. However, the economic benefit of VRFB is poor because of the high cost
Device design and VE: The device of the battery, such as flow of vanadium. The industrialization of ICRFB is still in infancy and has
channel design, and flow rate will also affect VE. Typical flow field de­ just entered the demonstration and verification stage of commerciali­
signs used in RFBs are the serpentine flow field (SFF) and IFF [134]. The zation. ZIRFB is promising considering the cost and performance, which
structure of IFF on a porous electrode is shown in Fig. 7f. Flow rate also is in the stage of system integration. In order to achieve the commer­
has a great influence on VE. Li et al. [80] prepared a high-energy and cialization of ZIRFBs, further breakthroughs are still needed in material,
low-cos PSIRFB. The performance of the battery at different flow rates chemistry, and battery design. The problem of poor power capability
(Fig. 7g) was tested, and the results showed that when the flow rate and cycle stability of PSB has hindered its development, which is still in
increased within a certain range, the VE value increased because the the stage of engineering optimization. Despite the advantages of high
electrolyte flow reduced the mass transport loss of redox substances on energy density and low cost, AORFBs are currently in the stage of
the electrode surface. However, once the flow rate exceeds a certain fundamental R&D. With breakthroughs in material modifications and
value, the increase of the flow rate does not improve VE, as the balance improvements in mass production, as well as the gradual improvement
between mass transport resistance and the rate of redox reactions limits of domestic and international industry chains, the LCOS for AORFBs is

11
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

Fig. 8. The roadmap of ARFBs.

expected to decline significantly. This could lead to further technolog­ most optimistic about ZIRFB in the short term, and AORFB and
ical breakthroughs and make them suitable for commercialization and AIRFB in the long term.
widespread application. The AIRFB is primarily still in the exploration (4) The factors affecting the performance. The factors affecting
stage. Although it technically possesses the same intrinsic system the performance of ARFBs are varied. We present a comprehen­
simplicity as VRFBs and has a low material cost, its development is sive overview of these determinants and corresponding
relatively immature when considering the battery performance, which is enhancement strategies, focusing on CE and VE. Factors affecting
mainly due to the lack of suitable active species. CE encompass the reversibility of the electrochemical reaction,
the stability of active materials, and the membrane crossover
5. Conclusion and perspective effect. VE, on the other hand, is subject to limitations arising from
sluggish reaction kinetics, membrane conductivity, and design-
To advance the widespread adoption of ARFBs, substantial research related issues within the flow channel configuration.
efforts have been devoted to enhancing their performance while
simultaneously reducing costs. We comprehensively review key per­ Despite the existing challenges in ARFBs, there is promise for their
formance metrics and cost assessment methodologies associated with utilization in large-scale energy storage applications. To facilitate their
ARFBs. Achieving cost advantages is contingent upon the judicious broader adoption, it is imperative to establish a rigorous and systematic
utilization of cost-effective materials, stack configurations, and the im­ framework for evaluating ARFB performance. Moreover, ongoing efforts
provements of battery performance. are directed toward ameliorating current challenges in active materials
and optimizing various components, including electrodes, membranes,
(1) The balance of the cost and performance. In practical appli­ and other devices, with the overarching goal of enhancing ARFB per­
cations, it is essential to adopt a holistic perspective that balances formance and cost-effectiveness.
cost considerations with battery performance. Based on the
standpoint of lifecycle levelized cost, energy cost-effectiveness, CRediT authorship contribution statement
power cost-effectiveness, and long-term stable cycling perfor­
mance are integral elements, rendering low-cost constituents Zhaoxia Hou: Writing – original draft. Xi Chen: Writing – original
advantageous only when coupled with all three aspects. draft. Jun Liu: Writing – original draft. Ziyi Huang: Writing – original
(2) The challenges of ARFBs and paradigm-shift innovations. A draft. Yan Chen: Writing – original draft. Mingyue Zhou: Writing –
detailed cost and performance analysis of various ARFB variants original draft. Wen Liu: Writing – original draft. Henghui Zhou:
reveals persistent challenges, including low energy density, Writing – original draft.
sluggish kinetics, and instability of low-cost redox-active mate­
rials. To overcome those challenges, paradigm-shift innovations
such as new chemistries and new materials are pressingly needed Declaration of competing interest
to go beyond conventional thinking.
(3) The characteristics of various ARFBs. Among current ARFB The authors declare that they have no known competing financial
configurations, the VRFB demonstrates the most stable perfor­ interests or personal relationships that could have appeared to influence
mance; however, its rapid expansion is hindered by the high cost the work reported in this paper.
of active materials. Zinc-based ARFBs and iron-based ARFBs have
the advantage of low cost, but their development is seriously Data availability
hindered by dendrite formation and hydrogen evolution. The use
of organic ligands can alleviate the current problems of iron- No data was used for the research described in the article.
based ARFBs. Sulfur-based ARFBs have the advantages of low
cost and high solubility, but the electrochemical activity and ki­ Acknowledgment
netic reversibility of polysulfides are poor. AORFBs have attrac­
ted attention for their natural renewability and flexible This work was supported by Beijing Municipal Natural Science
designability, but their low solubility and instability in aqueous Foundation-Xiaomi Joint Project (L223011), National Natural Science
solutions are the biggest problems at present. From the perspec­ Foundation of China (51967020, 22108149). In addition, the work was
tive of material cost differences and kinetic expectations, we are also supported by the Research and Development program in Shanxi
Province and Shanxi Energy Internet Research Institute. Mingyue Zhou
would like to thank the support from "the Fundamental Research Funds

12
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

for the Central Universities" (2462023QNXZ016). [31] Y. Song, H. Yan, H. Hao, Z. Liu, C. Yan, A. Tang, Simultaneous regulation of
solvation shell and oriented deposition toward a highly reversible Fe anode for
all-iron flow batteries, Small 18 (2022) e2204356.
References [32] J. Yang, H. Yan, H. Hao, Y. Song, Y. Li, Q. Liu, A. Tang, Synergetic modulation on
solvation structure and electrode interface enables a highly reversible zinc anode
[1] Y. Ding, C. Zhang, L. Zhang, Y. Zhou, G. Yu, Molecular engineering of organic for zinc–iron flow batteries, ACS Energy Lett. 7 (2022) 2331–2339.
electroactive materials for redox flow batteries, Chem. Soc. Rev. 47 (2018) [33] W. Yu, W. Shang, P. Tan, B. Chen, Z. Wu, H. Xu, Z. Shao, M. Liu, M. Ni, Toward a
69–103. new generation of low cost, efficient, and durable metal–air flow batteries,
[2] Z. Yang, J. Zhang, M.C. Kintner-Meyer, X. Lu, D. Choi, J.P. Lemmon, J. Liu, J. Mater. Chem. A 7 (2019) 26744–26768.
Electrochemical energy storage for green grid, Chem Rev 111 (2011) 3577–3613. [34] H. Chen, X. Zhang, s. zhang, S. Wu, F. Chen, J. Xu, A comparative study of iron-
[3] C.A. Machado, G.O. Brown, R. Yang, T.E. Hopkins, J.G. Pribyl, T.H. Epps, Redox vanadium and all-vanadium flow battery for large scale energy storage, Chem.
flow battery membranes: improving battery performance by leveraging Eng. J. 429 (2022) 132403.
structure–property relationships, ACS Energy Lett. 6 (2020) 158–176. [35] H. Zhang, H. Zhang, X. Li, Z. Mai, W. Wei, Silica modified nanofiltration
[4] R. Yan, Q. Wang, Redox-targeting-based flow batteries for large-scale energy membranes with improved selectivity for redox flow battery application, Energy
storage, Adv Mater 30 (2018) e1802406. Environ. Sci. 5 (2012) 6299–6303.
[5] M. Park, J. Ryu, W. Wang, J. Cho, Material design and engineering of next- [36] Q. Dai, Z. Liu, L. Huang, C. Wang, Y. Zhao, Q. Fu, A. Zheng, H. Zhang, X. Li, Thin-
generation flow-battery technologies, Nat. Rev. Mater. 2 (2016). film composite membrane breaking the trade-off between conductivity and
[6] T. Xuan, L. Wang, Eutectic electrolyte and interface engineering for redox flow selectivity for a flow battery, Nat. Commun. 11 (2020) 13.
batteries, Energy Storage Mater. 48 (2022) 263–282. [37] Y. Liu, Y. Shen, L. Yu, L. Liu, F. Liang, X. Qiu, J. Xi, Holey-engineered electrodes
[7] J.-Y. Liu, X.-X. Li, J.-R. Huang, J.-J. Li, P. Zhou, J.-H. Liu, X.-J. Huang, Three- for advanced vanadium flow batteries, Nano Energy 43 (2018) 55–62.
dimensional graphene-based nanocomposites for high energy density Li-ion [38] L. Wu, Y. Shen, L. Yu, J. Xi, X. Qiu, Boosting vanadium flow battery performance
batteries, J. Mater. Chem. A 5 (2017) 5977–5994. by Nitrogen-doped carbon nanospheres electrocatalyst, Nano Energy 28 (2016)
[8] S. Yuan, T. Kong, Y. Zhang, P. Dong, Y. Zhang, X. Dong, Y. Wang, Y. Xia, 19–28.
Advanced electrolyte design for high-energy-density Li-metal batteries under [39] L. Liu, X. Zhang, D. Zhang, K. Zhang, S. Hou, S. Wang, Y. Zhang, H. Peng, J. Liu,
practical conditions, Angew Chem. Int. Ed. Engl. 60 (2021) 25624–25638. C. Yan, Regulating the N/B ratio to construct B, N co-doped carbon nanotubes on
[9] V. Jülch, Comparison of electricity storage options using levelized cost of storage carbon felt for high-performance vanadium redox flow batteries, Chem. Eng. J.
(LCOS) method, Appl. Energy 183 (2016) 1594–1606. (2023) 473.
[10] O. Schmidt, S. Melchior, A. Hawkes, I. Staffell, Projecting the future levelized cost [40] B. Li, M. Gu, Z. Nie, Y. Shao, Q. Luo, X. Wei, X. Li, J. Xiao, C. Wang, V. Sprenkle,
of electricity storage technologies, Joule 3 (2019) 81–100. W. Wang, Bismuth nanoparticle decorating graphite felt as a high-performance
[11] C.S. Lai, M.D. McCulloch, Levelized cost of electricity for solar photovoltaic and electrode for an all-vanadium redox flow battery, Nano Lett. 13 (2013)
electrical energy storage, Appl. Energy 190 (2017) 191–203. 1330–1335.
[12] B. Hu, M. Hu, J. Luo, T.L. Liu, A stable, low permeable TEMPO catholyte for [41] H.-M. Tsai, S.-J. Yang, C.-C.M. Ma, X. Xie, Preparation and electrochemical
aqueous total organic redox flow batteries, Adv. Energy Mater. (2021) 12. activities of iridium-decorated graphene as the electrode for all-vanadium redox
[13] C. Debruler, W. Wu, K. Cox, B. Vanness, T.L. Liu, Integrated saltwater flow batteries, Electrochim. Acta 77 (2012) 232–236.
desalination and energy storage through a pH neutral aqueous organic redox flow [42] J. Shen, S. Liu, Z. He, L. Shi, Influence of antimony ions in negative electrolyte on
battery, Adv. Funct. Mater. 30 (2020). the electrochemical performance of vanadium redox flow batteries, Electrochim.
[14] Y. Zhao, Y. Ding, Y. Li, L. Peng, H.R. Byon, J.B. Goodenough, G. Yu, A chemistry Acta 151 (2015) 297–305.
and material perspective on lithium redox flow batteries towards high-density [43] H. Zhou, Y. Shen, J. Xi, X. Qiu, L. Chen, ZrO2-Nanoparticle-Modified graphite felt:
electrical energy storage, Chem. Soc. Rev. 44 (2015) 7968–7996. bifunctional effects on vanadium flow batteries, ACS Appl. Mater. Interfaces 8
[15] C. Zhang, L. Zhang, Y. Ding, S. Peng, X. Guo, Y. Zhao, G. He, G. Yu, Progress and (2016) 15369–15378.
prospects of next-generation redox flow batteries, Energy Storage Mater. 15 [44] A.W. Bayeh, D.M. Kabtamu, Y.-C. Chang, G.-C. Chen, H.-Y. Chen, G.-Y. Lin, T.-
(2018) 324–350. R. Liu, T.H. Wondimu, K.-C. Wang, C.-H. Wang, Synergistic effects of a
[16] F.R. Brushett, M.J. Aziz, K.E. Rodby, On lifetime and cost of redox-active organics TiNb2O7–reduced graphene oxide nanocomposite electrocatalyst for high-
for aqueous flow batteries, ACS Energy Lett. 5 (2020) 879–884. performance all-vanadium redox flow batteries, J. Mater. Chem. A 6 (2018)
[17] W. Wang, Q. Luo, B. Li, X. Wei, L. Li, Z. Yang, Recent progress in redox flow 13908–13917.
battery research and development, Adv. Funct. Mater. 23 (2012) 970–986. [45] C. Yao, H. Zhang, T. Liu, X. Li, Z. Liu, Carbon paper coated with supported
[18] R.M. Darling, K.G. Gallagher, J.A. Kowalski, S. Ha, F.R. Brushett, Pathways to tungsten trioxide as novel electrode for all-vanadium flow battery, J. Power
low-cost electrochemical energy storage: a comparison of aqueous and Sources 218 (2012) 455–461.
nonaqueous flow batteries, Energy Environ. Sci. 7 (2014) 3459–3477. [46] B. Sun, M. Skyllas-Kazacos, Chemical modification of graphite electrode materials
[19] C. Minke, U. Kunz, T. Turek, Techno-economic assessment of novel vanadium for vanadium redox flow battery application—part II. Acid treatments,
redox flow batteries with large-area cells, J. Power Sources 361 (2017) 105–114. Electrochim. Acta 37 (1992) 2459–2465.
[20] Y.K. Zeng, T.S. Zhao, L. An, X.L. Zhou, L. Wei, A comparative study of all- [47] K.V. Greco, A. Forner-Cuenca, A. Mularczyk, J. Eller, F.R. Brushett, Elucidating
vanadium and iron-chromium redox flow batteries for large-scale energy storage, the nuanced effects of thermal pretreatment on carbon paper electrodes for
J. Power Sources 300 (2015) 438–443. vanadium redox flow batteries, ACS Appl. Mater. Interfaces 10 (2018)
[21] V. Viswanathan, A. Crawford, D. Stephenson, S. Kim, W. Wang, B. Li, G. Coffey, 44430–44442.
E. Thomsen, G. Graff, P. Balducci, M. Kintner-Meyer, V. Sprenkle, Cost and [48] X. Rui, A. Parasuraman, W. Liu, D.H. Sim, H.H. Hng, Q. Yan, T.M. Lim, M. Skyllas-
performance model for redox flow batteries, J. Power Sources 247 (2014) Kazacos, Functionalized single-walled carbon nanotubes with enhanced
1040–1051. electrocatalytic activity for Br-/Br3- redox reactions in vanadium bromide redox
[22] Z. Li, Y.-C. Lu, Polysulfide-based redox flow batteries with long life and low flow batteries, Carbon 64 (2013) 464–471.
levelized cost enabled by charge-reinforced ion-selective membranes, Nat. Energy [49] N. Roznyatovskaya, J. Noack, H. Mild, M. Fühl, P. Fischer, K. Pinkwart, J. Tübke,
6 (2021) 517–528. M. Skyllas-Kazacos, Vanadium electrolyte for all-vanadium redox-flow batteries:
[23] J. Luo, B. Hu, C. Debruler, Y. Bi, Y. Zhao, B. Yuan, M. Hu, W. Wu, T.L. Liu, the effect of the counter ion, Batteries 5 (2019).
Unprecedented capacity and stability of ammonium ferrocyanide catholyte in pH [50] S. Huang, Z. Yuan, M. Salla, X. Wang, H. Zhang, S. Huang, D.G. Lek, X. Li,
neutral aqueous redox flow batteries, Joule 3 (2019) 149–163. Q. Wang, A redox-mediated zinc electrode for ultra-robust deep-cycle redox flow
[24] X. Wang, M. Zhou, F. Zhang, H. Zhang, Q. Wang, Redox targeting of energy batteries, Energy Environ. Sci. 16 (2023) 438–445.
materials, Curr. Opin. Electrochem. (2021) 29. [51] Z. Chen, W. Yu, Y. Liu, Y. Zeng, Q. He, P. Tan, M. Ni, Mathematical modeling and
[25] X. Wei, G.-G. Xia, B. Kirby, E. Thomsen, B. Li, Z. Nie, G.G. Graff, J. Liu, numerical analysis of alkaline zinc-iron flow batteries for energy storage
V. Sprenkle, W. Wang, An aqueous redox flow battery based on neutral alkali applications, Chem. Eng. J. (2021) 405.
metal ferri/ferrocyanide and polysulfide electrolytes, J. Electrochem. Soc. 163 [52] A. Hollas, X. Wei, V. Murugesan, Z. Nie, B. Li, D. Reed, J. Liu, V. Sprenkle,
(2015) A5150–A5153. W. Wang, A biomimetic high-capacity phenazine-based anolyte for aqueous
[26] J. Luo, B. Hu, M. Hu, Y. Zhao, T.L. Liu, Status and prospects of organic redox flow organic redox flow batteries, Nat. Energy 3 (2018) 508–514.
batteries toward sustainable energy storage, ACS Energy Lett. 4 (2019) [53] M. Mousavi, G. Jiang, J. Zhang, A.G. Kashkooli, H. Dou, C.J. Silva, Z.P. Cano,
2220–2240. Y. Niu, A. Yu, Z. Chen, Decoupled low-cost ammonium-based electrolyte design
[27] Y. Yao, J. Lei, Y. Shi, F. Ai, Y.-C. Lu, Assessment methods and performance for highly stable zinc–iodine redox flow batteries, Energy Storage Mater. 32
metrics for redox flow batteries, Nat. Energy 6 (2021) 582–588. (2020) 465–476.
[28] C. Hu, H. Chen, Y. Xie, L. Fang, J. Fang, J. Xu, H. Zhao, J. Zhang, Alleviating [54] M. Mousavi, H. Dou, H. Fathiannasab, C.J. Silva, A. Yu, Z. Chen, Elucidating and
polarization by designing ultrasmall Li2S nanocrystals encapsulated in N-rich tackling capacity fading of zinc-iodine redox flow batteries, Chem. Eng. J. (2021)
carbon as a cathode material for high-capacity, long-life Li–S batteries, J. Mater. 412.
Chem. A 4 (2016) 18284–18288. [55] C. Liu, X. Chi, Q. Han, Y. Liu, A high energy density aqueous battery achieved by
[29] N. Chang, Y. Yin, M. Yue, Z. Yuan, H. Zhang, Q. Lai, X. Li, A cost-effective mixed dual dissolution/deposition reactions separated in acid-alkaline electrolyte, Adv.
matrix polyethylene porous membrane for long-cycle high power density alkaline Energy Mater. 10 (2020).
zinc-based flow batteries, Adv. Funct. Mater. 29 (2019). [56] M. Park, E.S. Beh, E.M. Fell, Y. Jing, E.F. Kerr, D. De Porcellinis, M.A. Goulet,
[30] H. Zhang, C. Sun, M. Ge, Review of the research status of cost-effective zinc–iron J. Ryu, A.A. Wong, R.G. Gordon, J. Cho, M.J. Aziz, A high voltage aqueous
redox flow batteries, Batteries (2022) 8. zinc–organic hybrid flow battery, Adv. Energy Mater. 9 (2019).
[57] Z. Yuan, Y. Yin, C. Xie, H. Zhang, Y. Yao, X. Li, Advanced materials for zinc-based
flow battery: development and challenge, Adv Mater 31 (2019) e1902025.

13
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

[58] Z. Yuan, X. Li, Perspective of alkaline zinc-based flow batteries, Science China [87] M. Shin, S. Oh, H. Jeong, C. Noh, Y. Chung, J.W. Han, Y. Kwon, Aqueous redox
Chemistry 67 (2024) 260–275. flow battery using iron 2,2-bis(hydroxymethyl)-2,2′,2′-nitrilotriethanol complex
[59] D. Yu, L. Zhi, F. Zhang, Y. Song, Q. Wang, Z. Yuan, X. Li, Scalable alkaline zinc- and ferrocyanide as newly developed redox couple, Int. J. Energy Res. 46 (2022)
iron/nickel hybrid flow battery with energy density up to 200 Wh L− 1, Adv. 8175–8185.
Mater. 35 (2023) 2209390. [88] Y. Chen, M. Zhou, Y. Xia, X. Wang, Y. Liu, Y. Yao, H. Zhang, Y. Li, S. Lu, W. Qin,
[60] Z. Xie, Q. Su, A. Shi, B. Yang, B. Liu, J. Chen, X. Zhou, D. Cai, L. Yang, High X. Wu, Q. Wang, A stable and high-capacity redox targeting-based electrolyte for
performance of zinc-ferrum redox flow battery with Ac− /HAc buffer solution, aqueous flow batteries, Joule 3 (2019) 2255–2267.
J. Energy Chem. 25 (2016) 495–499. [89] H. Zhou, H. Zhang, P. Zhao, B. Yi, A comparative study of carbon felt and
[61] Z. Xu, Q. Fan, Y. Li, J. Wang, P.D. Lund, Review of zinc dendrite formation in zinc activated carbon based electrodes for sodium polysulfide/bromine redox flow
bromine redox flow battery, Renew. Sustain. Energy Rev. 127 (2020). battery, Electrochim. Acta 51 (2006) 6304–6312.
[62] W. Lu, P. Xu, S. Shao, T. Li, H. Zhang, X. Li, Multifunctional carbon felt electrode [90] W. Wang, A membrane with repelling power, Nat. Energy 6 (2021) 452–453.
with N-rich defects enables a long-cycle zinc-bromine flow battery with ultrahigh [91] M. Kathiresan, B. Ambrose, N. Angulakshmi, D.E. Mathew, D. Sujatha, A.
power density, Adv. Funct. Mater. (2021) 31. M. Stephan, Viologens, A versatile organic molecule for energy storage
[63] C. Xie, Y. Duan, W. Xu, H. Zhang, X. Li, A low-cost neutral zinc-iron flow battery applications, J. Mater. Chem. A 9 (2021) 27215–27233.
with high energy density for stationary energy storage, Angew Chem. Int. Ed. [92] Z. Li, M.S. Pan, L. Su, P.-C. Tsai, A.F. Badel, J.M. Valle, S.L. Eiler, K. Xiang, F.
Engl. 56 (2017) 14953–14957. R. Brushett, Y.-M. Chiang, Air-breathing aqueous sulfur flow battery for ultralow-
[64] M. Yang, Z. Xu, W. Xiang, H. Xu, M. Ding, L. Li, A. Tang, R. Gao, G. Zhou, C. Jia, cost long-duration electrical storage, Joule 1 (2017) 306–327.
High performance and long cycle life neutral zinc-iron flow batteries enabled by [93] W. Huang, Q. Zou, Y.C. Lu, Ion-selective membrane-free dual sulfur-iodine
zinc-bromide complexation, Energy Storage Mater. 44 (2022) 433–440. catholyte for low-cost and high-power flow battery applications, Batteries &
[65] J. Hu, C. Yuan, L. Zhi, H. Zhang, Z. Yuan, X. Li, In situ defect-free vertically Supercaps 2 (2019) 941–947.
aligned layered double hydroxide composite membrane for high areal capacity [94] J. Lei, Y. Yao, Y. Huang, Y.-C. Lu, A highly reversible low-cost aqueous
and long-cycle zinc-based flow battery, Adv. Funct. Mater. 31 (2021) 2102167. sulfur–manganese redox flow battery, ACS Energy Lett. 8 (2022) 429–435.
[66] M. Cui, Y. Xiao, L. Kang, W. Du, Y. Gao, X. Sun, Y. Zhou, X. Li, H. Li, F. Jiang, [95] M. Quan, D. Sanchez, M.F. Wasylkiw, D.K. Smith, Voltammetry of quinones in
C. Zhi, Quasi-isolated Au particles as heterogeneous seeds to guide uniform Zn unbuffered aqueous solution: reassessing the roles of proton transfer and
deposition for aqueous zinc-ion batteries, ACS Appl. Energy Mater. 2 (2019) hydrogen bonding in the aqueous electrochemistry of quinones, J. Am. Chem.
6490–6496. Soc. 129 (2007) 12847–12856.
[67] C. Shen, X. Li, N. Li, K. Xie, J.-g. Wang, X. Liu, B. Wei, Graphene-boosted, high- [96] J. Xu, S. Pang, X. Wang, P. Wang, Y. Ji, Ultrastable aqueous phenazine flow
performance aqueous Zn-ion battery, ACS Appl. Mater. Interfaces 10 (2018) batteries with high capacity operated at elevated temperatures, Joule 5 (2021)
25446–25453. 2437–2449.
[68] J.F. Parker, E.S. Nelson, M.D. Wattendorf, C.N. Chervin, J.W. Long, D.R. Rolison, [97] S. Pang, X. Wang, P. Wang, Y. Ji, Biomimetic amino acid functionalized
Retaining the 3D framework of zinc sponge anodes upon deep discharge in Zn–air phenazine flow batteries with long lifetime at near-neutral pH, Angew Chem. Int.
cells, ACS Appl. Mater. Interfaces 6 (2014) 19471–19476. Ed. Engl. 60 (2021) 5289–5298.
[69] J. Hu, J. Ding, Z. Du, H. Duan, S. Yang, Zinc anode with artificial solid electrolyte [98] X.-L. Lv, P. Sullivan, H.-C. Fu, X. Hu, H. Liu, S. Jin, W. Li, D. Feng, Dextrosil-
interface for dendrite-free Ni-Zn secondary battery, J. Colloid Interface Sci. 555 Viologen, A robust and sustainable anolyte for aqueous organic redox flow
(2019) 174–179. batteries, ACS Energy Lett. 7 (2022) 2428–2434.
[70] L. Kang, M. Cui, F. Jiang, Y. Gao, H. Luo, J. Liu, W. Liang, C. Zhi, Nanoporous [99] Y. Zhang, F. Li, T. Li, M. Zhang, Z. Yuan, G. Hou, J. Fu, C. Zhang, X. Li, Insights
CaCO3 coatings enabled uniform Zn stripping/plating for long-life zinc into an air-stable methylene blue catholyte towards kW-scale practical aqueous
rechargeable aqueous batteries, Adv. Energy Mater. 8 (2018) 1801090. organic flow batteries, Energy Environ. Sci. 16 (2023) 231–240.
[71] J. Hao, B. Li, X. Li, X. Zeng, S. Zhang, F. Yang, S. Liu, D. Li, C. Wu, Z. Guo, An in- [100] V. Singh, S. Kim, J. Kang, H.R. Byon, Aqueous organic redox flow batteries, Nano
depth study of Zn metal surface chemistry for advanced aqueous Zn-ion batteries, Res. 12 (2019) 1988–2001.
Adv. Mater. 32 (2020) 2003021. [101] B. Huskinson, M.P. Marshak, C. Suh, S. Er, M.R. Gerhardt, C.J. Galvin, X. Chen,
[72] A. Xia, X. Pu, Y. Tao, H. Liu, Y. Wang, Graphene oxide spontaneous reduction and A. Aspuru-Guzik, R.G. Gordon, M.J. Aziz, A metal-free organic-inorganic aqueous
self-assembly on the zinc metal surface enabling a dendrite-free anode for long- flow battery, Nature 505 (2014) 195–198.
life zinc rechargeable aqueous batteries, Appl. Surf. Sci. 481 (2019) 852–859. [102] J. Cao, M. Tao, H. Chen, J. Xu, Z. Chen, A highly reversible anthraquinone-based
[73] Z. Yuan, X. Liu, W. Xu, Y. Duan, H. Zhang, X. Li, Negatively charged nanoporous anolyte for alkaline aqueous redox flow batteries, J. Power Sources 386 (2018)
membrane for a dendrite-free alkaline zinc-based flow battery with long cycle 40–46.
life, Nat. Commun. 9 (2018) 3731. [103] S. Huang, H. Zhang, M. Salla, J. Zhuang, Y. Zhi, X. Wang, Q. Wang, Molecular
[74] W. Lu, T. Li, C. Yuan, H. Zhang, X. Li, Advanced porous composite membrane engineering of dihydroxyanthraquinone-based electrolytes for high-capacity
with ability to regulate zinc deposition enables dendrite-free and high-areal aqueous organic redox flow batteries, Nat. Commun. (2022) 13.
capacity zinc-based flow battery, Energy Storage Mater. 47 (2022) 415–423. [104] W. Lee, B.W. Kwon, Y. Kwon, Effect of carboxylic acid-doped carbon nanotube
[75] D. Kong, C. Yuan, L. Zhi, Q. Zheng, Z. Yuan, X. Li, N-CNTs-Based composite catalyst on the performance of aqueous organic redox flow battery using the
membrane engineered by A partially embedded strategy: a facile route to high- modified alloxazine and ferrocyanide redox couple, ACS Appl. Mater. Interfaces
performing zinc-based flow batteries, Adv. Funct. Mater. 33 (2023) 2301448. 10 (2018) 36882–36891.
[76] Y. Xia, X. Hou, X. Chen, F. Mu, Y. Wang, L. Dai, X. Liu, Y. Yu, K. Huang, W. Xing, [105] W. Lee, A. Permatasari, B.W. Kwon, Y. Kwon, Performance evaluation of aqueous
Z. Xu, Membrane with horizontally rigid zeolite nanosheet arrays against zinc organic redox flow battery using anthraquinone-2,7-disulfonic acid disodium salt
dendrites in zinc-based flow battery, Chem. Eng. J. 465 (2023) 142912. and potassium iodide redox couple, Chem. Eng. J. 358 (2019) 1438–1445.
[77] L. Zhi, T. Li, X. Liu, Z. Yuan, X. Li, Functional complexed zincate ions enable [106] B. Huskinson, M.P. Marshak, C. Suh, S. Er, M.R. Gerhardt, C.J. Galvin, X. Chen,
dendrite-free long cycle alkaline zinc-based flow batteries, Nano Energy 102 A. Aspuru-Guzik, R.G. Gordon, M.J. Aziz, A metal-free organic–inorganic aqueous
(2022) 107697. flow battery, Nature 505 (2014) 195–198.
[78] W. Lu, C. Zhang, H. Zhang, X. Li, Anode for zinc-based batteries: challenges, [107] T.J. Carney, S.J. Collins, J.S. Moore, F.R. Brushett, Concentration-dependent
strategies, and prospects, ACS Energy Lett. 6 (2021) 2765–2785. dimerization of anthraquinone disulfonic acid and its impact on charge storage,
[79] Z. Yuan, Y. Duan, T. Liu, H. Zhang, X. Li, Toward a low-cost alkaline zinc-iron Chem. Mater. 29 (2017) 4801–4810.
flow battery with a polybenzimidazole custom membrane for stationary energy [108] R. Chen, Redox flow batteries: electrolyte chemistries unlock the thermodynamic
storage, iScience 3 (2018) 40–49. limits, Chem. Asian J. 18 (2023) e202201024.
[80] Z. Li, G. Weng, Q. Zou, G. Cong, Y.-C. Lu, A high-energy and low-cost polysulfide/ [109] T. Kong, J. Liu, X. Zhou, J. Xu, Y. Xie, J. Chen, X. Li, Y. Wang, Stable operation of
iodide redox flow battery, Nano Energy 30 (2016) 283–292. aqueous organic redox flow batteries in air atmosphere, Angew. Chem. Int. Ed. 62
[81] D.G. Kwabi, K. Lin, Y. Ji, E.F. Kerr, M.-A. Goulet, D. De Porcellinis, D.P. Tabor, D. (2023) e202214819.
A. Pollack, A. Aspuru-Guzik, R.G. Gordon, M.J. Aziz, Alkaline quinone flow [110] K. Peng, Y. Li, G. Tang, Y. Liu, Z. Yang, T. Xu, Solvation regulation to mitigate the
battery with long lifetime at pH 12, Joule 2 (2018) 1894–1906. decomposition of 2,6-dihydroxyanthraquinone in aqueous organic redox flow
[82] M. Shin, C. Noh, Y. Chung, Y. Kwon, All iron aqueous redox flow batteries using batteries, Energy Environ. Sci. 16 (2023) 430–437.
organometallic complexes consisting of iron and 3-[bis (2-hydroxyethyl)amino]- [111] T. Liu, X. Wei, Z. Nie, V. Sprenkle, W. Wang, A total organic aqueous redox flow
2-hydroxypropanesulfonic acid ligand and ferrocyanide as redox couple, Chem. battery employing a low cost and sustainable methyl viologen anolyte and 4-HO-
Eng. J. (2020) 398. TEMPO catholyte, Adv. Energy Mater. 6 (2015).
[83] W. Ruan, J. Mao, S. Yang, Q. Chen, Communication—tris(bipyridyl)iron [112] B. Hu, C. DeBruler, Z. Rhodes, T.L. Liu, Long-cycling aqueous organic redox flow
complexes for high-voltage aqueous redox flow batteries, J. Electrochem. Soc. battery (AORFB) toward sustainable and safe energy storage, J. Am. Chem. Soc.
(2020) 167. 139 (2017) 1207–1214.
[84] A. Sy, A.I. Bhatti, F. Hamidouche, O. Le Bacq, L. Lecarme, J.C. Lepretre, [113] T. Janoschka, N. Martin, M.D. Hager, U.S. Schubert, An aqueous redox-flow
Correlation of electrochemical and ab initio investigations of iron poly-bipyridine battery with high capacity and power: the TEMPTMA/MV system, Angew. Chem.
coordination complexes for battery applications: impact of the anionic Int. Ed. 55 (2016) 14427–14430.
environment and the local geometries of the redox complexes on the [114] Y. Liu, M.-A. Goulet, L. Tong, Y. Liu, Y. Ji, L. Wu, R.G. Gordon, M.J. Aziz, Z. Yang,
electrochemical response, Phys. Chem. Chem. Phys. 22 (2020) 24077–24085. T. Xu, A long-lifetime all-organic aqueous flow battery utilizing TMAP-TEMPO
[85] M.C. Tucker, A. Phillips, A.Z. Weber, All-iron redox flow battery tailored for off- radical, Chem 5 (2019) 1861–1870.
grid portable applications, ChemSusChem 8 (2015) 3996–4004. [115] M. Pan, L. Gao, J. Liang, P. Zhang, S. Lu, Y. Lu, J. Ma, Z. Jin, Reversible redox
[86] K.L. Hawthorne, J.S. Wainright, R.F. Savinell, Studies of iron-ligand complexes chemistry in pyrrolidinium-based TEMPO radical and extended viologen for high-
for an all-iron flow battery application, J. Electrochem. Soc. 161 (2014) voltage and long-life aqueous redox flow batteries, Adv. Energy Mater. (2022) 12.
A1662–A1671.

14
Z. Hou et al. Journal of Power Sources 601 (2024) 234242

[116] X.-L. Lv, P.T. Sullivan, W. Li, H.-C. Fu, R. Jacobs, C.-J. Chen, D. Morgan, S. Jin, [128] D. Ma, B. Hu, W. Wu, X. Liu, J. Zai, C. Shu, T. Tadesse Tsega, L. Chen, X. Qian, T.
D. Feng, Modular dimerization of organic radicals for stable and dense flow L. Liu, Highly active nanostructured CoS(2)/CoS heterojunction electrocatalysts
battery catholyte, Nat. Energy 8 (2023) 1109–1118. for aqueous polysulfide/iodide redox flow batteries, Nat. Commun. 10 (2019)
[117] Y. Jing, E.W. Zhao, M.-A. Goulet, M. Bahari, E.M. Fell, S. Jin, A. Davoodi, 3367.
E. Jónsson, M. Wu, C.P. Grey, R.G. Gordon, M.J. Aziz, In situ electrochemical [129] B.K. Chakrabarti, J. Feng, E. Kalamaras, J. Rubio-Garcia, C. George, H. Luo,
recomposition of decomposed redox-active species in aqueous organic flow Y. Xia, V. Yufit, M.M. Titirici, C.T.J. Low, A. Kucernak, N.P. Brandon, Hybrid
batteries, Nat. Chem. 14 (2022) 1103–1109. redox flow cells with enhanced electrochemical performance via binderless and
[118] E.W. Zhao, T. Liu, E. Jónsson, J. Lee, I. Temprano, R.B. Jethwa, A. Wang, electrophoretically deposited nitrogen-doped graphene on carbon paper
H. Smith, J. Carretero-González, Q. Song, C.P. Grey, In situ NMR metrology electrodes, ACS Appl. Mater. Interfaces 12 (2020) 53869–53878.
reveals reaction mechanisms in redox flow batteries, Nature 579 (2020) 224–228. [130] M. Kim, S. Lee, J. Choi, J. Park, J.-W. Park, M. Park, Reversible metal ionic
[119] B. Hu, Y. Tang, J. Luo, G. Grove, Y. Guo, T.L. Liu, Improved radical stability of catalysts for high-voltage aqueous hybrid zinc-manganese redox flow batteries,
viologen anolytes in aqueous organic redox flow batteries, Chem. Commun. 54 Energy Storage Mater. 55 (2023) 698–707.
(2018) 6871–6874. [131] J. Guo, L. Pan, J. Sun, D. Wei, Q. Dai, J. Xu, Q. Li, M. Han, L. Wei, T. Zhao, Metal-
[120] Y. Ahn, J. Moon, S.E. Park, J. Shin, J. Wook Choi, K.J. Kim, High-performance free fabrication of nitrogen-doped vertical graphene on graphite felt electrodes
bifunctional electrocatalyst for iron-chromium redox flow batteries, Chem. Eng. J. with enhanced reaction kinetics and mass transport for high-performance redox
(2021) 421. flow batteries, Adv. Energy Mater. 14 (2024) 2302521.
[121] D. Mu, L. Yu, L. Liu, J. Xi, Rice paper reinforced sulfonated poly(ether ether [132] A. Mukhopadhyay, Y. Yang, Z. Cheng, P. Luan, A. Natan, H. Zhu, Proton-
ketone) as low-cost membrane for vanadium flow batteries, ACS Sustain. Chem. conductive membranes with percolated transport paths for aqueous redox flow
Eng. 5 (2017) 2437–2444. batteries, Materials Today Nano (2021) 13.
[122] Y. Yin, S. Wang, Q. Zhang, Y. Song, N. Chang, Y. Pan, H. Zhang, X. Li, Dendrite- [133] Y. Xing, L. Liu, C. Wang, N. Li, Side-chain-type anion exchange membranes for
free zinc deposition induced by tin-modified multifunctional 3D host for stable vanadium flow battery: properties and degradation mechanism, J. Mater. Chem.
zinc-based flow battery, Adv Mater 32 (2020) e1906803. A 6 (2018) 22778–22789.
[123] C.T. Wan, R.R. Jacquemond, Y.M. Chiang, K. Nijmeijer, F.R. Brushett, A. Forner- [134] X. Ke, J.M. Prahl, J.I.D. Alexander, J.S. Wainright, T.A. Zawodzinski, R.
Cuenca, Non-solvent induced phase separation enables designer redox flow F. Savinell, Rechargeable redox flow batteries: flow fields, stacks and design
battery electrodes, Adv Mater 33 (2021) e2006716. considerations, Chem. Soc. Rev. 47 (2018) 8721–8743.
[124] P. Meister, H. Jia, J. Li, R. Kloepsch, M. Winter, T. Placke, Best practice: [135] W. Zhou, M. Song, P. Liang, X. Li, X. Liu, H. Li, T. Zhang, B. Wang, R. Zhao,
performance and cost evaluation of lithium ion battery active materials with Z. Zhao, W. Li, D. Zhao, D. Chao, High-energy Sn–Ni and Sn–air aqueous batteries
special emphasis on energy efficiency, Chem. Mater. 28 (2016) 7203–7217. via stannite-ion electrochemistry, J. Am. Chem. Soc. 145 (2023) 10880–10889.
[125] F. Pan, Q. Huang, H. Huang, Q. Wang, High-energy density redox flow lithium [136] P. Schröder, N. Aguiló-Aguayo, D. Obendorf, T. Bechtold, Near to neutral pH all-
battery with unprecedented voltage efficiency, Chem. Mater. 28 (2016) iron redox flow battery based on environmentally compatible coordination
2052–2057. compounds, Electrochim. Acta (2022) 430.
[126] Z. Huang, A. Mu, Flow field design and performance analysis of vanadium redox [137] S. Yan, S. Huang, H. Xu, L. Li, H. Zou, M. Ding, C. Jia, Q. Wang, Redox targeting-
flow battery, Ionics 27 (2021) 5207–5218. based neutral aqueous flow battery with high energy density and low cost,
[127] J. Liu, L. Ren, Y. Wang, X. Lu, M. Zhou, W. Liu, A highly-stable bifunctional ChemSusChem (2023) 16.
NiCo2S4 nanoarray@carbon paper electrode for aqueous polysulfide/iodide
redox flow battery, J. Power Sources (2023) 561.

15

You might also like