You are on page 1of 10

Flow battery

Article Talk
 Language
 Download PDF
 Watch
 Edit
A flow battery, or redox flow battery (after reduction–oxidation), is a type
of electrochemical cell where chemical energy is provided by two chemical
components dissolved in liquids that are pumped through the system on separate sides of a
membrane.[2][3] Ion transfer inside the cell (accompanied by flow of electric current through
an external circuit) occurs through the membrane while both liquids circulate in their own
respective space. Cell voltage is chemically determined by the Nernst equation and ranges, in
practical applications, from 1.0 to 2.43 volts. The energy capacity is a function of the
electrolyte volume and the power is a function of the surface area of the electrodes.[4]

A typical flow battery consists of two


tanks of liquids which are pumped past a membrane held between two electrodes.[1]
Various types of flow batteries have been demonstrated, including inorganic flow
batteries[5] and organic flow batteries.[6] Under each category, flow battery design can be
further classified into full flow batteries, semi-flow batteries, and membraneless flow
batteries. The fundamental difference between conventional and flow batteries is that energy
is stored in the electrode material in conventional batteries, while in flow batteries it is stored
in the electrolyte. Patent Classifications for Flow Batteries have not been fully developed as
of 2021. Cooperative Patent Classification considers RFBs as a subclass of regenerative fuel
cell (H01M8/18), even though it is more appropriate to consider fuel cells as a subclass of
flow batteries.
A flow battery may be used like a fuel cell (where new charged negolyte (a.k.a. reducer or
fuel) and charged posolyte (a.k.a. oxidant) are added to the system) or like a rechargeable
battery (where an electric power source drives regeneration of the reducer and oxidant).
While flow batteries have certain technical advantages over conventional rechargeable
batteries with solid electroactive materials, such as independent scaling of power
( determined by the size of the stack) and of energy (determined by the size of the tanks),
long cycle and calendar life, and potentially lower total cost of ownership, all flow batteries
suffer from low cycle energy efficiency (50–80%). This inferior energy efficiency stems from
the need to operate flow batteries at high (>= 100 mA/cm2) current densities to reduce the
effect of the internal crossover (through the membrane/separator inside each cell) and to
reduce the cost of power (size of stacks). Also, most flow batteries (Zn-Cl2, Zn-Br2 and H2-
LiBrO3 are exceptions) have lower specific energy (heavier weight) compared to lithium-ion
batteries. The heavier weight results mostly from the need to use a solvent (usually-water)
Historyedit

The log number of publication related to electrochemical power sources by year. Also shown
as the magenta line is the inflation-adjusted oil price in US$/liter in linear scale
The zinc–bromine flow battery (Zn-Br2) is the oldest flow battery chemistry,[4] with John
Doyle's patent US224404 filed on September 29, 1879.[8] Zn-Br2 batteries have relatively
high specific energy, and they were demonstrated as power sources for electric cars in the
1970s.[9]
Walther Kangro, an Estonian chemist working in Germany, was the first to demonstrate in
the 1950s flow batteries based fully on dissolved transition metal ions: Ti–Fe and Cr–Fe.
[10] After some initial experimentations with Ti–Fe RFB NASA and some other groups in
Japan and elsewhere selected Cr–Fe chemistry for further development. In order to reduce the
effect of time-varying concentration during RFB cycling, mixed solutions (i.e. comprising
both chromium and iron species in the negolyte and in the posolyte) were used. Among
disadvantages of the Cr–Fe chemistry are: hydrate isomerism (i.e. the equilibrium between
electrochemically active Cr3+ chloro-complexes and inactive hexa-aqua complex, which can
be alleviated by adding chelating amino-ligands) and hydrogen evolution on the negode
(which is mitigated by adding Pb salts for increasing the H2 overvoltage and Au salts for
catalyzing the chromium electrode reaction).[11]
In the late 1980s, Sum, Rychcik and Skyllas-Kazacos[12] at the University of New South
Wales (UNSW) in Australia demonstrated the advantages of all-vanadium RFB chemistry,
such as the existence of four oxidation states within the electrochemical voltage window of
the graphite-aqueous acid interface, and thus the elimination of the mixing dilution,
detrimental in Cr–Fe RFBs. UNSW filed several patents related to VRFBs, that were later
licensed to Japanese, Thai and Canadian corporations, which tried to commercialize this
technology with variable success.[13]
In 2022, Dalian, China began operating a 400 MWh, 100 MW vanadium flow battery, then
the largest of its type.[14]
Designedit
A flow battery is a rechargeable fuel cell in which an electrolyte containing one or more
dissolved electroactive elements flows through an electrochemical cell that reversibly
converts chemical energy directly to electricity. Electroactive elements are "elements in
solution that can take part in an electrode reaction or that can be adsorbed on the
electrode."[15] Additional electrolyte is stored externally, generally in tanks, and is usually
pumped through the cell (or cells) of the reactor, although gravity feed systems are also
known.[16] Flow batteries can be rapidly "recharged" by replacing the electrolyte liquid (in a
similar way to refilling fuel tanks for internal combustion engines) while simultaneously
recovering the spent material for recharging. Many flow batteries use carbon felt electrodes
due to its low cost and adequate electrical conductivity, although these electrodes somewhat
limit power density due to their low inherent activity toward many redox couples.[17][18]
In other words, a flow battery is an electrochemical cell, with the property that the ionic
solution (electrolyte) is stored[19] outside of the cell (instead of in the cell around the
electrodes) and can be fed into the cell in order to generate electricity. The total amount of
electricity that can be generated depends on the volume of electrolyte in the tanks.
Flow batteries are governed by the design principles established by electrochemical
engineering.[20]
Redox flow batteries, and to a lesser extent hybrid flow batteries, have the advantages of:

 Independent scaling of energy (tanks) and power (stack), which allows for a
cost/weight/etc. optimization for each application
 Long cycle and calendar lives (because there are no solid-to-solid phase transitions,
that cause degradation of lithium-ion and related batteries)
 Quick response times
 No need for "equalisation" charging (the overcharging of a battery to ensure all cells
have an equal charge)
 No harmful emissions
 Little to no self-discharge during standing
 Full recycling of electroactive materials

Some types also offer easy state-of-charge determination (through voltage dependence on
charge), low maintenance and tolerance to overcharge/overdischarge.

They are safe because they typically do not contain flammable electrolytes, and electrolytes
can be stored away from the power stack.

Flow batteries have three main disadvantages compared to batteries with solid electroactive
materials[21]
 Low energy density (you need large tanks of electrolyte to store useful amounts of
energy)
 Low charge and discharge rates (compared to other industrial electrode processes).
This means that the electrodes and membrane separators need to be large, which
increases the cost of power.
 Flow batteries have a lower energy efficiency, because they operate at higher current
densities to minimize the effects of the cross-over (internal self-discharge) and to
reduce the cost of power.

Flow batteries typically have a higher energy efficiency than fuel cells, but lower
than lithium-ion batteries.[22]
Traditional flow batteriesedit
The redox (reduction–oxidation) cell is a reversible cell in which redox-active species are in
fluid (liquid or gas) media. Redox flow batteries are rechargeable (secondary) cells.
[23] Because they employ heterogeneous electron transfer rather than solid-state
diffusion or intercalation they are more similar to fuel cells rather than to conventional
batteries (such as lead–acid or lithium-ion). The main reason fuel cells are not considered to
be batteries, is because originally (in the 1800s) fuel cells emerged as a means to produce
electricity directly from fuels (and air) via a non-combustion electrochemical process. Later,
particularly in the 1960s and 1990s, rechargeable fuel cells (i.e. H
2/O

2, such as unitized regenerative fuel cells in NASA's Helios Prototype) were developed.

Examples of redox flow batteries are the vanadium redox flow battery, polysulfide–bromide
battery (Regenesys), iron redox flow battery (IRFB), and uranium redox flow battery.
[24] Redox fuel cells are less common commercially although many systems have been
proposed.[25][26][27][28]
Vanadium redox flow batteries are the most marketed flow batteries at present, due to the
advantages they provide over other chemistries, despite limited energy and power densities.
Since they use vanadium at both electrodes, they do not suffer cross-contamination. The
limited solubility of vanadium salts, however, offsets this advantage in practice. More
importantly for the commercial success of VRFBs is actually an almost perfect match of the
voltage window of carbon/aqueous acid interface with the working voltage range of the
vanadium redox-couples. This assures the durability of the low-cost carbon electrodes and
low-impact of side reactions, such as H2 and O2 evolutions, resulting in record-long calendar
(many years) and cycle(15,000–20,000 cycles) lives, which in turn results in a record
low levelized cost of energy (LCOE, i.e. the system cost divided by the usable energy, the
cycle life, and round-trip efficiency). The long lifetimes of flow batteries allow for the
amortization of their relatively high capital cost (due to vanadium, carbon felts, bipolar
plates, membranes). The levelized cost of energy for VRFBs is in the order of a few tens of $
cents or € cents per kWh, much lower than of solid-state batteries and not so far from the
targets of $0.05 and €0.05, stated by US and EC government agencies.[29] The major
challenges for the broad implementation include: low abundance and high costs of V2O5 (>
$30 / Kg), the raw materials for VRFB; parasite reactions including hydrogen and oxygen
evolution; and precipitation of V2O5 during cycling. It is the major driving force to develop
alternative flow battery technologies.
Traditional flow battery chemistries have both low specific energy (which makes them too
heavy for fully electric vehicles) and low specific power (which makes them too expensive
for stationary energy storage). However a high power of 1.4 W/cm2 was demonstrated for
hydrogen–bromine flow batteries, and a high specific energy (530 Wh/kg at the tank level)
was shown for hydrogen–bromate flow batteries[30][31][32]
In 2022, DARPA SBIR-funded Influit Energy announced a nonflammable, surface-modified
nanoelectrofuel made of a metal oxide suspended in an aqueous solution. The material does
not settle out of the solution, even at high concentration and features and energy density
higher than Lion batteries. Operating temperatures are -40 to 80 °C. They require no
lithium, heavy metals, or rare-earth elements.[33][34]
Hybrid flow batteriesedit
The hybrid flow battery uses one or more electroactive components deposited as a solid layer.
[35] The major disadvantage is the loss decoupled energy and power as seen in full flow
batteries from using a solid state electrode. The cell contains one battery electrode and one
fuel cell electrode. This type is limited in energy by the electrode surface area. Hybrid flow
batteries include the zinc–bromine, zinc–cerium,[36] soluble lead–acid,[37] and iron-salt flow
batteries. Weng et al.[38] reported a vanadium–metal hydride rechargeable hybrid flow battery
with an experimental OCV of 1.93 V and operating voltage of 1.70 V, relatively high values
among rechargeable flow batteries with aqueous electrolytes. This hybrid battery consists of a
graphite felt positive electrode operating in a mixed solution of VOSO
4and H

2SO

4, and a metal hydride negative electrode in KOH aqueous solution. The two electrolytes of

different pH are separated by a bipolar membrane. The system demonstrated good


reversibility and high efficiencies in coulomb (95%), energy (84%), and voltage (88%). They
reported further improvements of this redox couple with achievements of increased current
density, inclusion of larger 100 cm2 electrodes, and the operation of 10 large cells in series.
Preliminary data using a fluctuating simulated power input tested the viability toward kWh
scale storage.[39] In 2016, a high energy density Mn(VI)/Mn(VII)-Zn hybrid flow battery was
proposed.[40]
A prototype zinc–polyiodide flow battery demonstrated an energy density of 167 Wh/L (watt-
hours per liter). Older zinc–bromide cells reach 70 Wh/L. For comparison, lithium iron
phosphate batteries store 233 Wh/L. The zinc–polyiodide battery is claimed to be safer than
other flow batteries given its absence of acidic electrolytes, nonflammability and operating
range of −4 to 122 °F (−20 to 50 °C) that does not require extensive cooling circuitry, which
would add weight and occupy space. One unresolved issue is zinc buildup on the negative
electrode that can permeate the membrane, reducing efficiency. Because of the Zn dendrite
formation, Zn-halide batteries cannot operate at high current density (> 20 mA/cm2) and thus
have limited power density. Adding alcohol to the electrolyte of the ZnI battery can help with
the problem.[41] The drawbacks of Zn/I RFB lie at the high cost of Iodide salts (> $20 / Kg);
limited area capacity of Zn deposition also losing the decoupled energy and power; and Zn
dendrite formation.
When the battery is fully discharged, both tanks hold the same electrolyte solution: a mixture
of positively charged zinc ions (Zn2+
) and negatively charged iodide ion, (I−
). When charged, one tank holds another negative ion, polyiodide, (I−
3). The battery produces power by pumping liquid from external tanks into the battery's stack

area where the liquids are mixed. Inside the stack, zinc ions pass through a selective
membrane and change into metallic zinc on the stack's negative side.[42] To further increase
the energy density of the zinc–iodide flow battery, bromide ions (Br
–) are used as the complexing agent to stabilize the free iodine, forming iodine–bromide ions

(I
2Br

) as a means to free up iodide ions for charge storage.[43]


Proton flow batteries (PFB) integrate a metal hydride storage electrode into a
reversible proton exchange membrane (PEM) fuel cell. During charging, PFB combines
hydrogen ions produced from splitting water with electrons and metal particles in one
electrode of a fuel cell. The energy is stored in the form of a solid-state metal hydride.
Discharge produces electricity and water when the process is reversed and the protons are
combined with ambient oxygen. Metals less expensive than lithium can be used and provide
greater energy density than lithium cells.[44][45]
Organic typesedit
Compared to redox flow batteries that are inorganic, such as vanadium redox flow batteries
and Zn-Br2 batteries, which have been developed for decades, organic redox flow batteries
emerged in 2009. The primary appeal of organic redox flow batteries lies in the tunable redox
properties of the active components. As of 2021, organic RFB experience low durability (i.e.
calendar or cycle life, or both). For this reason, only inorganic RFB have been demonstrated
on a commercial scale.[46]
Organic redox flow batteries can be further classified into aqueous (AORFBs) and non-
aqueous (NAORFBs).[47][48] AORFBs use water as solvent for electrolyte materials while
NAORFBs employ organic solvents. AORFBs and NAORFBs can be further divided into
total and hybrid organic systems. The former use only organic electrode materials, while the
latter use inorganic materials for anode or cathode. In larger-scale energy storage, lower
solvent cost and higher conductivity give AORFBs greater commercial potential, as well as
offering safety advantages from water-based electrolytes. NAORFBs instead provide a much
larger voltage window and occupy less physical space.
pH neutral AORFBsedit
pH neutral AORFBs are operated at pH7 conditions, typically using NaCl as a supporting
electrolyte. At pH neutral conditions, organic and organometallic molecules are more stable
than at corrosive acidic and alkaline conditions. For example, K4[Fe(CN)], a common
catholyte used in AORFBs, is not stable in alkaline solutions but is at pH neutral conditions.
[49]
AORFBs used methyl viologen as an anolyte and 4-hydroxy-2,2,6,6-tetramethylpiperidin-1-
oxyl as a catholyte at pH neutral conditions, plus NaCL and a low-cost anion exchange
membrane. This MV/TEMPO system has the highest cell voltage, 1.25 V, and, possibly,
lowest capital cost ($180/kWh) reported for AORFBs. The aqueous liquid electrolytes were
designed as a drop-in replacement for current systems without replacing existing
infrastructure. A 600-milliwatt test battery was stable for 100 cycles with nearly 100 percent
efficiency at current densities ranging from 20 to 100 mA/cm2, with optimal performance
rated at 40–50 mA, at which about 70% of the battery's original voltage was retained.[50]
[51] Neutral AORFBs can be more environmentally friendly than acid or alkaline AORFBs
while showing electrochemical performance comparable to corrosive RFBs. The
MV/TEMPO AORFB has an energy density of 8.4 Wh/L with the limitation on the TEMPO
side. Viologen-based flow batteries have been mainly developed by Liu's group at Utah State
University. In 2019, the group reported an ultralight sulfonate–
viologen/ferrocyanide AORFB stable for 1000 cycles at an energy density of 10 Wh/L, so far
the most stable, energy dense AORFB.[52]
Acidic AORFBsedit
Quinones and their derivatives are the basis of many organic redox systems.[53][54][55] In one
study, 1,2-dihydrobenzoquinone-3,5-disulfonic acid (BQDS) and 1,4-dihydrobenzoquinone-
2-sulfonic acid (BQS) were employed as cathodes, and conventional Pb/PbSO4 was the
anolyte in a hybrid acid AORFB. Quinones accept two units of electrical charge, compared
with one in conventional catholyte, implying that such a battery could store twice as much
energy in a given volume.
Another quinone 9,10-Anthraquinone-2,7-disulfonic acid (AQDS), has been evaluated.
[56] AQDS undergoes rapid, reversible two-electron/two-proton reduction on a glassy
carbon electrode in sulfuric acid. An aqueous flow battery with inexpensive carbon
electrodes, combining the quinone/hydroquinone couple with the Br
2/Br

redox couple, yields a peak galvanic power density exceeding 6,000 W/m2 at 13,000 A/m2.
Cycling showed > 99% storage capacity retention per cycle. Volumetric energy density was
over 20 Wh/L.[57] Anthraquinone-2-sulfonic acid and anthraquinone-2,6-disulfonic acid on
the negative side and 1,2-dihydrobenzoquinone- 3,5-disulfonic acid on the positive side
avoids the use of hazardous Br2. The battery was claimed to last for 1,000 cycles without
degradation.[58] While this system appears robust, it has a low cell voltage (ca. 0.55 V) and a
low energy density (< 4 Wh/L).
Hydrobromic acid used as an electrolyte has been replaced with a less toxic alkaline solution
(1 M KOH) and ferrocyanide.[59] The higher pH is less corrosive, allowing the use of
inexpensive polymer tanks. The increased electrical resistance in the membrane was
compensated increased voltage. The cell voltage was 1.2 V.[60][61] The cell's efficiency
exceeded 99%, while round-trip efficiency measured 84%. The battery offered an expected
lifetime of at least 1,000 cycles. Its theoretic energy density was 19 Wh/L.[62] Ferrocyanide's
chemical stability in high pH KOH solution without forming Fe(OH)2 or Fe(OH)3 needs to be
verified before scale-up.
Integrating both anolyte and catholyte in the same molecule has been examined. Such
bifunctional analytes[63] or combi-molecules[64] allow the same material to be used in both
tanks. In one tank it is an electron donor, while in the other it is an electron recipient. This has
relevant advantages such as diminishing the effect of crossover.[65] Thus, quinone
diaminoanthraquinone[65] and indigo-based[63] molecules as well as
TEMPO/phenazine[64] combining molecules are potential electrolytes for the development of
symmetric redox-flow batteries (SRFB).
Another approach adopted a Blatter radical as the donor/recipient. It endured 275 charge and
discharge cycles in tests, although it was not water-soluble.[66]
Alkalineedit
Quinone molecules have been used as anolytes in alkaline AROFBs. Another anolyte
candidate is fluorenone, reengineered to increase its water solubility. A
reversible ketone (de)hydrogenation demonstration cell operated continuously for 120 days
over 1,111 charging cycles at room temperature without a catalyst, retaining 97% percent
capacity. The cell offers more than double the energy density of vanadium-based systems.[67]
[68] The major challenge for alkaline AORFBs is the lack of a stable catholyte, holding their
energy densities below 5 Wh/L. All reported alkaline AORFBs use excess potassium
ferrocyanide catholyte because of the stability issue of ferrocyanide in alkaline solutions.
Metal-organic flow batteries use organic ligands to improve the properties of redox-active
metals. The ligands can be chelates like EDTA, and can enable the electrolyte to be in neutral
or alkaline conditions under which metal aquo complexes would otherwise precipitate. By
blocking the coordination of water to the metal, organic ligands can inhibit metal-
catalyzed water-splitting reactions, resulting in higher voltage all-aqueous systems. For
example, the use of chromium coordinated to 1,3-propanediaminetetraacetate (PDTA), gave
cell potentials of 1.62 V vs. ferrocyanide and a record 2.13 V vs. bromine.[69] Metal-organic
flow batteries may be known as coordination chemistry flow batteries, which represents the
technology behind Lockheed Martin's Gridstar Flow technology.[70]
Oligomeredit
Oligomer redox-species RFB have been proposed to reduce the crossover of the electroactive
species, while using low cost membranes. Such redox-active oligomers are known as
redoxymers. One system uses organic polymers and a saline solution with
a cellulose membrane. The prototype underwent 10,000 charging cycles while retaining
substantial capacity. The energy density was 10 Wh/L.[71] Current density reached 100
milliamperes/cm2.[72]
Another oligomer RFB employs viologen and TEMPO redoxymers in combination with low-
cost dialysis membranes. Functionalized macromolecules (similar to acrylic glass
or Styrofoam) dissolved in water are the active electrode material. The size-selective
nanoporous membrane works like a strainer and is produced much more easily and at lower
cost than conventional ion-selective membranes. It retains the big "spaghetti"-like polymer
molecules, while allowing small counterions to pass.[73] The concept may solve the high cost
of traditional Nafion membrane, but the design and synthesis of redox active polymer with
high water solubility is not trivial. So far, RFBs with oligomer redox-species have not
demonstrated competitive area-specific power. It is not clear whether low operating current
density is an intrinsic feature of large redox-molecules or not.
Other typesedit
Other flow-type batteries include the zinc–cerium battery, the zinc–bromine battery, and
the hydrogen–bromine battery.
Membranelessedit
A membraneless battery[74] relies on laminar flow in which two liquids are pumped through a
channel, where they undergo electrochemical reactions to store or release energy. The
solutions stream through in parallel, with little mixing. The flow naturally separates the
liquids, eliminating the need for a membrane.[75]
Membranes are often the most costly and least reliable components of batteries, as they can
be corroded by repeated exposure to certain reactants. The absence of a membrane enables
the use of a liquid bromine solution and hydrogen: this combination is problematic when
membranes are used, because they form hydrobromic acid that can destroy the membrane.
Both materials are available at low cost.[76] The design uses a small channel between two
electrodes. Liquid bromine flows through the channel over a graphite cathode and
hydrobromic acid flows under a porous anode. At the same time, hydrogen gas flows across
the anode. The chemical reaction can be reversed to recharge the battery – a first for any
membraneless design.[76] One such membraneless flow battery published in August 2013
produced a maximum power density of 0.795 mW/cm2, three times as much power as other
membraneless systems— and an order of magnitude higher than lithium-ion batteries.[76]
In 2018, a macroscale membraneless redox flow battery capable of recharging and
recirculation of the same electrolyte streams for multiple cycles has been demonstrated. The
battery is based on immiscible organic catholyte and aqueous anolyte liquids, which exhibits
high capacity retention and Coulombic efficiency during cycling.[77]
Nano-networkedit

Semi-solid flow battery[78]


Lithium–sulfur system arranged in a network of nanoparticles eliminates the requirement that
charge moves in and out of particles that are in direct contact with a conducting plate.
Instead, the nanoparticle network allows electricity to flow throughout the liquid. This allows
more energy to be extracted.[79]
In a Semi-solid flow battery, the positive and negative electrodes are composed of particles
suspended in a carrier liquid. The positive and negative suspensions are stored in separate
tanks and pumped through separate pipes into a stack of adjacent reaction chambers, where
they are separated by a barrier such as a thin, porous membrane. The approach combines the
basic structure of aqueous-flow batteries, which use electrode material suspended in a liquid
electrolyte, with the chemistry of lithium-ion batteries in both carbon-free suspensions and
slurries with conductive carbon network.[1][80][81] The carbon free semi-solid redox flow
battery is also sometimes referred to as Solid Dispersion Redox Flow Battery.[82] Dissolving
a material changes its chemical behavior significantly. However, suspending bits of solid
material preserves the solid's characteristics. The result is a viscous suspension that flows
like molasses.[83]
Flow batteries with redox-targeted solids (ROTS), also known as solid energy
boosters (SEBs), is another recent development.[84][85][86][87][88][89][90] In these batteries either
posolyte or negolyte or both (a.k.a. redox fluids), come in contact with a one or more solid
electroactive materials, stored in tanks outside the power stack. The redox fluids comprise
one or more redox couples, with redox potentials flanking the redox potential of the solid
electroactive material. Such RFBs with Solid Energy Boosters (SEBs) combine the high
specific energy advantage of conventional batteries (such as lithium-ion) with the decoupled
energy-power advantage of flow batteries. SEB(ROTS) RFBs have several advantages
compared to semi-solid RFBs, such as no need to pump viscous slurries, no precipitation
/clogging , higher area-specific power, longer durability, wider chemical design space.
However, because of double energy losses (one in the stack and another in the tank between
the SEB(ROTS) and a mediator), such batteries suffer from a poor energy efficiency. On a
system-level, the practical specific energy of traditional lithium-ion batteries is larger than
that of SEB(ROTS)-flow versions of lithium-ion batteries.[91]
Comparisonedit

Comparison of flow battery compositions

Max. cell Average Average fluid


Couple voltage electrode power energy Cycles
(V) density (W/m2) density

Hydrogen–lithium
1.1 15,000 750 Wh/kg
bromate
Hydrogen–lithium
1.4 10,000 1400 Wh/kg
chlorate
Bromine–hydrogen 1.07 7,950
Iron–tin 0.62 < 200
Iron–titanium 0.43 < 200
Iron–chromium 1.07 < 200
Iron–iron 1.21 < 1000 20 Wh/L 10,000
Organic (2013) 0.8 13,000 21.4 Wh/L 10
Organic (2015) 1.2 7.1 Wh/L 100
MV-TEMPO 1.25 8.4 Wh/L 100
Sulfonate viologen
0.9 > 500 10 Wh/L 1,000
(NH4)4[Fe(CN)6]
Metal-organic–
1.62 2,000 21.7 Wh/L 75
ferrocyanide[69]
Metal-organic–
2.13 3,000 35 Wh/L 10
bromine[69]
Vanadium–vanadium
1.4 ~800 25 Wh/L
(sulphate)
Vanadium–vanadium 50 Wh/L 2,000[2]
Comparison of flow battery compositions

Max. cell Average Average fluid


Couple voltage electrode power energy Cycles
(V) density (W/m2) density

(bromide)
Sodium–bromine
1.54 ~800
polysulfide
Sodium–potassium[92]
Sulfur–oxygen-salt[93]
Zinc–bromine 1.85 ~1,000 75 Wh/kg > 2,000
Lead–acid
1.82 ~1,000
(methanesulfonate)
Zinc–cerium
2.43 < 1,200–2,500
(methanesulfonate)
Zn-Mn(VI)/Mn(VII) 1.2 60 Wh/L[40]

The main focusing point is that we are facing in the following things:

1. How to convert the slurry into the electrolyte.


2. Which is the suitable electrolyte that can produce the electricity with the slurry.
3. How the reaction takes palce between the electrolytes .
4. Can it possible to write the reaction between those without conducting the laboratory
experiment.

You might also like