You are on page 1of 9

Journal of The Electrochemical

Society

OPEN ACCESS

A Durable, Inexpensive and Scalable Redox Flow Battery Based on Iron


Sulfate and Anthraquinone Disulfonic Acid
To cite this article: Bo Yang et al 2020 J. Electrochem. Soc. 167 060520

View the article online for updates and enhancements.

This content was downloaded from IP address 81.41.156.37 on 16/04/2020 at 00:10


Journal of The Electrochemical Society, 2020 167 060520

A Durable, Inexpensive and Scalable Redox Flow Battery Based on


Iron Sulfate and Anthraquinone Disulfonic Acid
Bo Yang,=,* Advaith Murali,= Archith Nirmalchandar,=,* Buddhinie Jayathilake,*
G. K. Surya Prakash,** and S. R. Narayanan***,z
Department of Chemistry, Loker Hydrocarbon Research Institute, University of Southern California, Los Angeles, California
90089, United States of America

A new redox flow battery system based on iron sulfate and anthraquinone disulfonic acid (AQDS) is shown here to have excellent
electrical performance, capacity retention, and chemical durability. While these redox couples, iron(II)/iron(III) and AQDS are well
known individually, their combination in a redox flow battery is shown here for the first time to provide unique benefits for large-
scale energy storage. Based on iron sulfate, a waste product of the steel industry, the active materials cost for this battery is
anticipated to be $66/kWh. Cycling studies of over 500 cycles in the symmetric cell configuration show a negligibly low capacity
fade rate of 7.6 × 10−5% per cycle. This symmetric cell also shows a notably high average coulombic efficiency of 99.63%. Using
a graphite felt electrode modified with multi-walled carbon nanotubes (MWCNTs), we could achieve a peak power density of
194 mW cm−2. The major voltage losses are ascribed to the ohmic resistance of the electrode and electrolyte. Despite the lower cell
voltage of the system relative to the vanadium flow battery, the iron–AQDS flow battery system presents a good prospect for
simultaneously meeting the demanding requirements of cost, durability and scalability for large-scale sustainable energy storage.
© 2020 The Author(s). Published on behalf of The Electrochemical Society by IOP Publishing Limited. This is an open access
article distributed under the terms of the Creative Commons Attribution 4.0 License (CC BY, http://creativecommons.org/licenses/
by/4.0/), which permits unrestricted reuse of the work in any medium, provided the original work is properly cited. [DOI: 10.1149/
1945-7111/ab84f8]

Manuscript submitted December 13, 2019; revised manuscript received March 20, 2020. Published April 9, 2020. This was paper
408 presented at the Dallas, Texas, Meeting of the Society, May 26–May 30, 2019.
Supplementary material for this article is available online

The demand for large-scale energy storage.—The exponential friendliness.3 Globally, over 1600 million metric tons of iron (steel) are
growth in solar and wind-based electricity generation has been produced every year.9 There have been previous reports attempting to
driving the need for energy storage systems at an unprecedented use iron-based compounds as the positive electrolyte in redox flow
scale. Even to store 20% of the electricity generated today by solar batteries. In these examples, the iron was deployed as the ferrocyanide/
and wind systems we will need an enormous storage capacity of ferricyanide couple,10,11 ferrocene/ferrocenium couple,12,13 or as iron-
700 GWh.1 While the cumulative world battery production is phenanthroline complexes.14 The specific compositions mentioned
expected to reach about 370 GWh by 2020 due to the demand for here entail significantly higher cost than desirable, and also present
lithium-ion batteries in electric vehicles,2 we can expect further the challenge of long-term chemical stability and durability.15,16
exponential growth because of the global deployment of renewable Although the aqueous “all-iron” flow battery uses only iron(II)
electricity generation systems. At the anticipated “mega” scale of chloride that is relatively inexpensive,17,18 continuous operation has
deployment, the principal requirements for successful storage tech- to contend with multiple challenges leading to high system costs.
nologies are durability, low cost and sustainability of materials.3,4 Firstly, the parasitic evolution of hydrogen at the negative
Rechargeable batteries in general, and redox flow batteries in electrode18,19 requires additional sub-systems and reactors to capture
particular, have many advantages to fulfill this need.5–7 and recombine the hydrogen and maintain the acidity of the solutions.
Also, the kinetics of electrodeposition of iron is relatively slow,
Levelized cost of energy storage.—The levelized cost of energy necessitating operation at low current densities. These characteristics
storage (LCOS) is the metric used to rate the economic competitiveness lead to higher system costs despite the low materials costs.
of energy storage systems. LCOS is calculated by simply dividing the Unfortunately, unlike the vanadium systems, reliable cost data is not
sum of the capital and operating costs of the system by the total energy available for the iron chloride system to make comparisons. Iron-
stored and delivered over the lifetime of the system. To achieve market Chromium redox flow batteries use iron(II) chloride at the positive
competitiveness, the LCOS target for large-scale energy storage electrode,20 but are also faced with the challenge of hydrogen
systems is 2.5 cents kWh−1, as suggested by the US Department of evolution at the chromium electrode.21–23 More recently, Tucker
Energy (DoE).8 To satisfy this LCOS target, a system costing $200/ et al. proposed a low-cost single-use portable battery based on iron
kWh must store 8000 kWh over its lifetime. Assuming one charge/ (III) salts and metallic iron as an inexpensive power source for
discharge cycle per day, such an energy storage system must last at developing countries emphasizing the non-toxic and inexpensive
least for 22 years. Consequently, the system should be based on nature of iron-based battery materials.24 Thus, despite the advantages
inexpensive and abundant materials that are available without geopo- of iron, the large-scale exploitation of iron as an active material in
litical constraints, while also being sustainable and eco-friendly. Thus, redox flow batteries has hitherto been a challenge.
the battery industry is seeking cost-effective alternatives to common
battery materials such as lead, zinc, lithium, vanadium, chromium, etc.
Iron-AQDS flow battery.—We present here for the first time the
Iron-based flow batteries.—Iron is particularly attractive as a properties of a redox flow battery system based on iron that does not
battery material because of its abundance, low cost and environmental suffer from the limitations of the iron-based batteries described above.
This flow battery uses aqueous solutions of iron(II)/iron(III) sulfate at
the positive electrode and a water-soluble organic redox couple,
anthraquinone/anthraquinol disulfonic acid (AQDS/AQDSH2) at the
=
These authors contributed equally to this work. negative electrode. Steel mills generate enormous amounts of iron
*Electrochemical Society Student Member. sulfate as the waste product of scale removal. Hence, iron sulfate is
**Electrochemical Society Member.
***Electrochemical Society Fellow. available for as low as $0.10/kg25 and can be put to good use in this
z
E-mail: sri.narayan@usc.edu battery. We and several others have studied the use of AQDS in flow
Journal of The Electrochemical Society, 2020 167 060520

batteries26–29 and our research has shown that AQDS is one of the sulfuric acid (VWR, ACS grade) and deionized water (18 MOhm-cm)
most robust and efficient redox materials for repeatedly storing and was used in preparing the active material solutions for all the
delivering charge at the negative electrode.30–32 In addition, the large- experiments. Nafion® 212 and 117 proton exchange membranes (Ion
scale manufacturing cost of AQDS is projected to be as low as Power Inc.) were used as purchased. Electrodes (geometric area
$3/kg.33,34 Therefore, the iron sulfate/AQDS system is particularly 25 cm2) were prepared by modifying graphite felt (PAN Graphite Felt,
promising as a low-cost solution for “mega”-scale energy storage. 3 mm thick, Graphite Store) with 10% (w/w) of multi-wall carbon
nanotubes (Flo Tube 9000, Cnano Technology). To prepare such
Operating principle of the iron/AQDS flow battery.—A sche- electrodes, typically 100 mg of carbon nanotubes were dispersed in
matic of the Iron/AQDS cell and the accompanying reversible 100 ml of isopropanol by ultrasonication for 2 h. Nafion® ionomer
reactions at each electrode are shown in Fig. 1. Iron(II) and iron (5 or 10 weight percent of the carbon nanotubes) was added to the
(III) exist as aqua complexes in solution. The positive electrode is mixture of carbon nanotubes and isopropanol before sonication to
based on the interconversion of the iron(II) and iron(III). At the achieve a stable dispersion and also serve as a binder to attach the
negative electrode, AQDS undergoes redox interconversion between carbon nanotubes to the graphite felt. The dispersion of carbon
the quinone and hydroquinone forms and protons are added or nanotubes in isopropanol was then impregnated in the graphite felt
released. Sulfuric acid is the source of protons for this reaction, and to the required loading level by repeatedly dropping small quantities of
is also used on both sides of the cell. A high acid concentration (2 M the dispersion on to a heated graphite felt electrode. When the
sulfuric acid) on the positive side ensures that the aqua complexes of isopropanol evaporated the carbon nanotubes were left behind in the
iron do not hydrolyze. Further, protons also serve to transport the electrode with the Nafion®. The electrodes were then cured at 140 °C
ionic current across the cell. Proton transport occurs through a in air to render the Nafion® insoluble and complete the attachment of
proton exchange membrane. For example, Nafion® 212 is a proton the carbon nanotubes to the graphite felt. Subsequently, the electrodes
exchange membrane that combines high proton conductivity and were then oxidized in a 30% sulfuric acid solution to ensure full
durability. The reversible cell voltage for such an iron/AQDS cell wetting of the interior parts of the electrodes prior to cell assembly.
under standard state conditions is 0.62 V. We determined the charge-
transfer rate constants and the diffusion coefficients for the two Methods of testing and analysis.—Rotating disk voltammetry
redox couples, and found these values to be suitable for achieving was conducted on a glassy carbon electrode (Pine Instruments) with a
high efficiencies and high current densities (Table I). platinum wire counter electrode and a mercury sulfate reference
electrode (Eo = 0.68 V vs NHE). Voltammetry experiments used a
scan rate of 50 mV sec−1 and electrode rotation speeds in the range of
Experimental
400 to 3000 rotations per minute (rpm). Voltammetry and impedance
Materials.—Ferrous sulfate heptahydrate (J. T. Baker, >99%), measurements (over the frequency range of 10 mHz to 100 kHz) were
1 M anthraquinone-2,7-disulfonic acid solution (Riverside Specialty), made using a potentiostat (Versastat 300 with impedance analyzer).

Figure 1. Operating principle of the Iron/AQDS redox flow cell.


Journal of The Electrochemical Society, 2020 167 060520

Table I. Charge-transfer rate constants and diffusion coefficient for iron and AQDS redox couples as measured from linear sweep voltammetry at a
rotating disk electrode.

Diffusion coefficient, 10−6 cm2 s−1

Redox couple Standard rate constant for electron transfer, 10−5 cm s−1 Oxidized species Reduced species
3+ 2+
Fe /Fe 7.3 3.43 5.74
AQDS/AQDSH2 15.2 3.40 1.64

Redox flow cells were assembled using custom-designed cell hard- that are found to crossover through Nafion®.31 Thus, the crossover of
ware (Electrochem Inc.). For flow rates of electrolyte below iron(II) and iron(III) does lead to a capacity fade rendering the
0.3 l min−1, we used a centrifugal magnetic drive pump (BC-2CP- asymmetric cell configuration unsuitable for long-term operation.
MD 12 VOLT DC, March Manufacturing Inc.). For flow rates greater However, the permanent loss of capacity faced in the asymmetric
than 0.3 l min−1, a gear pump (Eclipse pump E02KLVF-X, Baldor configuration can be completely avoided by operating the cell in the
Motor CDP3310, SCR DC control BC140) was used. All full-cell “symmetric” or “mixed electrolyte” configuration.
experiments were carried out at 23 °C. A blanket of argon gas was In the symmetric cell configuration, we used an identical mixture
maintained over the headspace of the reservoirs to avoid reaction with of iron sulfate and AQDS as the positive and negative electrolyte
oxygen. Charge/discharge studies, current-voltage measurements, and solutions (Fig. 2a). At the positive electrode, during charging, only
long-duration cycling were conducted using a battery cycler (Maccor iron(II) in the mixture gets oxidized to iron(III); AQDS present on
4200 with 30 A capability). Charge capacity was determined by the this side cannot be oxidized further at the potential of the positive
coulombs needed to reach a cut-off voltage of 1 V. Discharge capacity electrode (Fig. 2b). At the negative electrode, during charging,
was determined from the coulombs delivered until a cut-off voltage of AQDS is reduced to AQDSH2 as iron(II) cannot be reduced at the
0.001 V. Coulombic efficiency was calculated for every cycle by potential at which AQDS is reduced. Similarly, during discharge, the
dividing the discharge capacity value by the value of charge capacity iron(III) is reduced to iron(II) at the positive electrode while AQDS
immediately preceding the discharge. Therefore, coulombic efficiency stays unaffected. AQDSH2 at the negative electrode is oxidized to
refers to a particular cycle. AQDS, while the iron(II) remains unchanged. At the end of several
cycles of discharge, any imbalance that may arise in the distribution
Permeation studies.—To measure the permeation rate of iron(II) of iron between the two sides can be erased by mixing the
and iron(III) through the proton exchange membrane, a permeation electrolytes and evenly splitting the solutions between the two
cell was set up with two 30 ml glass reservoirs separated by a reservoirs. The effect of the “mix-and-split” protocol can also be
circular piece of Nafion® 212 membrane with a permeation area of achieved by interchanging the leads of the cell periodically so as to
2 cm2. On one side of the permeation cell we used 1 M iron(II) or iron make the previously positive electrode the negative electrode.31,38,39
(III) sulfate in 1 M sulfuric acid, and the other side was a plain 1 M The mixed electrolyte configuration has also the benefit of main-
sulfuric acid solution. Samples were withdrawn at fixed intervals from taining similar osmotic pressures on both sides of the cell thus
the sulfuric acid side and analyzed for their concentration of iron by avoiding water transport across the cell. However, one negative
rotating disk voltammetry at a glassy carbon disk electrode. The mass- aspect of the mixed electrolyte system is the increased viscosity of
transport limited currents at 1500 rpm were converted to concentration the solutions because both the redox couples are present in the same
values. The diffusion coefficient (D) was calculated as per the solution at high concentrations. The additional material required for
following equation (see SI-5) applied to the initial rate. the mixed electrolyte on both sides is less of a concern when the
active materials are as inexpensive as iron sulfate and AQDS.
DCL (t) Dt = ADV-1d -1 (CH ‐CL) In our previous studies on aqueous all-organic flow batteries, we
have shown that the permanent capacity fade caused by crossover
Where CL is the concentration of test species in the “low” side, V is can be avoided by starting with a mixed-electrolyte or symmetric
the volume of the “high” concentration side, A is membrane area, δ is cell configuration.31 In the present study, we started with equal
membrane thickness, and CH is the concentration of test species on the amounts of a mixture of iron sulfate and AQDS on both sides of the
“high” concentration side. The apparent diffusion coefficient of the cell, and repeatedly charged and discharged the cell to full capacity.
permeating species was then obtained from the slope of CL vs time. While the capacity of the “asymmetric” iron/AQDS cell cycled at
200 mA cm−2 showed a rapid decay as expected (Fig. 3a), the
capacity of the “symmetric” cell did not show any noticeable change
Results and Discussion
in capacity even after 500 cycles (Fig. 3b). The average capacity
Cell configuration and overcoming the challenge of cross- fade rate was as low as 7.6 × 10−5% per cycle (or 1.2 × 10−4% per
over.—A conventional redox flow battery operates in an “asym- hour). At this capacity fade rate, we can project a fade of 1.2% in
metric” configuration (Fig. 1) where each type of redox couple is 10,000 h of operation. These highly stable capacity values are a
circulated past the appropriate electrode (positive or negative). In proof of the inherent durability of the iron/AQDS redox battery.
such a case, the redox materials usually crossover from one electrode These results showed that AQDS in a mixed electrolyte environment
to the other through the proton exchange membrane resulting in is stable to oxidation by iron(III). Also, iron(III) that crossed over
fading of the cell’s capacity.31,35–37 We measured the apparent underwent reversible electrochemical reactions with AQDSH2
diffusion coefficients of iron(II) and iron(III) through the Nafion® during the cycling. Repeated cycling did not result in any changes
212 membrane and found them to be comparable to their diffusion to the shape of the charge/discharge curves or produce any new
coefficient in aqueous solutions (Table II). Therefore, it was not discharge plateaus thereby confirming the chemical stability of the
surprising that crossover in the asymmetric cell caused a rapid redox couples (See supplementary information SI-1 is available
decrease of discharge capacity. Although iron(II) and iron(III) online at stacks.iop.org/JES/167/060520/mmedia).
readily crossed over, we found that AQDS did not show any The average coulombic efficiency of the “symmetric” cell was
measurable permeation through the Nafion® 212 membrane, con- 99.63%, over the 500 cycles (Fig. 3b), suggesting that the rate of
sistent with our previous observations in aqueous all-organic redox crossover of iron(II) and iron(III) was extremely low. Since the
flow batteries.30,31 We attributed this inability of AQDS to crossover diffusion coefficient of iron(II) through the membrane is higher
to its relatively larger size compared to the benzoquinone derivatives than that of iron(III), and since AQDS does not crossover, the
Journal of The Electrochemical Society, 2020 167 060520

Table II. Apparent diffusion coefficients through nafion® 212 membrane.

Solutions used for permeation studies Apparent diffusion coefficient cm2 s−1

1 M Iron(II) sulfate 3.33 × 10−6


0.5 M Iron(III) sulfate 1.33 × 10−6
1 M Iron(II) sulfate + 0.5 M AQDS 2.33 × 10−6
0.5 M Iron(III) sulfate + 0.5 M AQDS 1.00 × 10−6

Figure 2. (a) Schematic of the three types of electrolyte configurations studied, and (b) processes during the charging and discharging of symmetric cell with the
mix-and-split protocol.

mix-and-split protocol usually needed to reset the imbalance in equimolar mixture of iron(II) and AQDS solutions as the negative
concentrations in a symmetric cell was not required during this electrolyte (Fig. 2a). Upon extended cycling we noticed that the
cycling test. These advantages arising from the fundamental proper- capacity began to fade albeit at 2.8 × 10−2% per hour (Fig. 3c)
ties of the iron/AQDS system are significant in avoiding additional tending to a lower fade rate with time. Once again, no crossover of
balance-of-plant costs. Subsequent to the repeated cycling experi- AQDS was detected. However the difference in the diffusion
ments, the cell was subject to other tests under various other coefficients of iron(II) in an aqueous solution vs iron(II) in
conditions and even after 1300 cycles of testing in various modes, combination with AQDS solution, and also the difference in the
the cell did not suffer any noticeable change in capacity, further osmotic pressure between the positive and negative sides could
testimony to the stability of the materials. explain the higher fade rate of the semi-symmetric cell compared to
Since AQDS does not crossover through the Nafion® membrane, the symmetric configuration cell (Table II). Such differences in the
we investigated the feasibility of a “semi-symmetric” configuration diffusion coefficients could arise from the higher viscosity of the
in which we used an iron(II) solution as a positive electrolyte and an iron(II)-AQDS solutions compared to the plain iron(II) solutions.
Journal of The Electrochemical Society, 2020 167 060520

Figure 3. (a) Coulombic efficiency and discharge capacity during cycling of asymmetric cell at 200 mA cm−2, with 1 M iron(II) sulfate and 2 M sulfuric acid,
200 ml on the positive side, and 1 M AQDS and 2 M sulfuric acid, 100 ml on the negative side. (b) Coulombic efficiency and discharge capacity during cycling
of symmetric cell at 200 mA cm−2, with mixed electrolyte consisting of 0.67 M iron(II) sulfate, 0.33 M AQDS and 2 M sulfuric, 150 mL in each reservoir, and
(c) Coulombic efficiency and discharge capacity during cycling at 100 mA cm−2 of semi-symmetric cell with 1 M iron(II) sulfate and 2 M sulfuric acid on the
positive side, and 1 M iron(II) sulfate with 2 M sulfuric acid and 0.5 M AQDS on the negative side, with 100 mL on both sides. Flow rate was 0.3 l min−1 in all
the experiments.

Also, some degree of association between iron(II) and AQDS cannot and details of how these values were determined). The charge-
be excluded. transfer resistance decreases with increasing current and mass
For asymmetric cells, the crossover process causes a steady loss transfer resistance contributions do not manifest until the current
in discharge capacity and imbalance of concentration of iron(II). approaches the limiting current for mass transport, while the ohmic
However, with symmetric cells, the iron ions crossover from the resistance is expected to not change much with current. Thus, the
positive electrode to the negative electrode and then diffuse back decrease in cell voltage in range of current being tested, is largely
even during a single charge half-cycle, equalizing the iron concen- governed by ohmic resistance losses. We have plotted the current-
tration on both sides. Therefore, the discharge capacity does not voltage curves after correcting the cell voltage for the ohmic
decline steadily from cycle to cycle, although the crossover process overpotential loss (Fig. 4a). This corrected cell voltage is indicated
lowers the coulombic efficiency of any particular cycle. as the “IR-corrected voltage.” The small difference between the IR-
corrected voltage and the open circuit voltage is consistent with the
Current-voltage performance, ohmic losses and efficiency.—To low contribution of charge-transfer and mass transfer processes to
characterize the electrical performance of the iron/AQDS symmetric the internal resistance of the cell.
cell, we measured the current-voltage curves. The electrodes in these Ohmic losses arise from the use of a thick and porous graphite
cells consisted of graphite felt coated with multi-wall carbon felt electrode (2–3 mm thick) and the Nafion® 117 membrane (about
nanotubes (MWCNT) that were prepared in-house as described in 175 microns thick). While the graphite electrodes modified with
the experimental section. The addition of 10 wt% of MWCNT MWCNT exhibited a lower resistance compared to the unmodified
increased the active surface area of the electrodes by over 150 times felt electrodes, our experiments varying the Nafion® content of these
compared to the plain graphite felt electrode, increased the utiliza- electrodes suggested that there was scope for the optimization of the
tion of active materials, and reduced the ohmic resistance presented MWCNT modified electrodes with respect to the amount of Nafion®.
by the electrodes. Using the “forced flow” configuration for To verify the impact of the ohmic contribution, we prepared
electrolyte movement through the felt electrodes, we could achieve electrodes with a lower content of the Nafion® binder.
current densities as high as 840 mA cm−2 at 100% state-of-charge Specifically, the amount of Nafion® dispersant used in modifying
(Fig. 4a). The maximum power density for this cell was the felts was reduced from 10% to 5%. The use of 10% Nafion was
134 mW cm−2 (Fig. 4b) Electrochemical impedance spectroscopy expected to yield a better dispersion of the carbon nanotubes and
(EIS) measurements on this cell were used to separate the ohmic, more accessible area to the reactant flow compared to the 5% Nafion.
charge-transfer and mass transfer contributions to the polarization This difference was consistent with the differences in the mass
resistance (See supplementary SI-3 for results of EIS measurements transport resistance values. When coupled with a Nafion® 117
Journal of The Electrochemical Society, 2020 167 060520

Figure 4. (a) Current-voltage curves for symmetric cell with Nafion® 117 membrane, 10% Nafion® electrode, 1 M iron(II) sulfate with 0.5 M AQDS and 2 M
sulfuric acid, 100 mL on both sides, (b) power density curve for cell in cell in Fig. 4a. (c) Current voltage curve for semi-symmetric cell with Nafion® 212
membrane, 5% Nafion® electrode, 1 M iron(II) sulfate and 2 M sulfuric acid on the positive side and 1 M iron(II) sulfate with 2 M sulfuric acid and 0.5 M AQDS
on the negative side, and (d) power density of for cell in Fig. 4c.

membrane, the felt with 5% binder showed an ohmic resistance of 17 batteries. The results in Figs. 4c and 4d suggest that achieving higher
mOhms. This value was 6 mOhms lower than the Nafion 117 power density without compromising energy efficiency will require
membrane coupled with the felt that had 10% binder. Furthermore, further reduction of the ohmic losses in the iron/AQDS cell. The
by deploying a Nafion® 212 membrane that is just 50 microns thick, flow field and electrode combinations may also be modified further
we could lower the ohmic resistance by another 4 mOhms, down to to achieve greater utilization of the internal surface area of the
13 mOhms. These observations suggested that the Nafion® binder carbon paper electrodes as demonstrated by Perry et al.40
being an electronic insulator prevented the multi-walled carbon Furthermore, given the inherent stability of the active materials,
nanotubes from being electrically connected to the felt structure the cell may be operated at higher temperatures. With such design
thereby increasing the resistance of the electrode. Thus, the cell with optimization, the power densities can be increased further, although
MWCNT-modified felt electrodes with 5% Nafion® binder resulted such improvements were beyond the scope of the current study.
in a marked increase in power density to 194 mW cm−2 and
reduction of the area specific ohmic resistance from 0.575 Ohm Low materials cost.—The most important features of the iron
cm2 to 0.325 Ohm cm2. The electrodes with 5% of Nafion® and sulfate/AQDS system are the durability and low cost of materials.
N212 membrane could support current densities as high as We have found no measurable change in the capacity of these cells
1.16 A cm−2 compared to the 0.84 A cm−2 by the electrodes with over 500 cycles in symmetric cells and no deterioration of active
10% Nafion® and N117 membrane. Further reductions in the ohmic material across long periods of cycling. Although the thermody-
contributions can be achieved by reducing the thickness of the felt namic cell voltage of 0.62 V is not as high as that for the vanadium
electrodes or by using carbon fiber paper electrodes, and by system, we project an active material cost (including acid), to be
increasing the temperature of the cell. about $54/kWh for the symmetric cell using acidic solutions of iron
The round-trip energy efficiency of the symmetric cell was in the sulfate and AQDS (see Supplementary Information SI-4 for calcula-
range of 70%–75% at a constant current density of 100 mA cm−2 tions and assumptions). For systems that rely on long-duration
and at a utilization of 70% of the theoretical capacity (See operation at low rates of charge and discharge, the cost of redox
Supplementary Information, SI-2). The coulombic efficiency of the materials is the dominant contributor to system cost. In comparison,
iron/AQDS cell being very close to 100% avoids any significant the materials cost is significantly lower than that of the materials cost
reduction in energy efficiency, unlike in the case of vanadium flow of state-of-art vanadium systems at $160–180/kWh,41,42
Journal of The Electrochemical Society, 2020 167 060520

We recognize that the stack costs will be higher for the Fe-AQDS 4. S. R. Narayan, A. Nirmalchandar, A. Murali, B. Yang, L. Hoober-Burkhardt,
system because of the lower voltage. Therefore, we have estimated S. Krishnamoorthy, and G. K. S. Prakash, “Next-generation aqueous flow battery
chemistries.” Curr. Opin. Electrochem., 18, 72 (2019).
the power-related cost at twice that of the VRFB system. With 5. B. Dunn, H. Kamath, and J.-M. Tarascon, “Electrical energy storage for the grid: a
further reduction of the ohmic resistance contributions we can battery of choices.” Science, 334, 928 (2011).
project an operating voltage of 0.5 V for the Fe-AQDS cell. Thus, 6. G. L. Soloveichik, “Flow batteries: current status and trends.” Chem. Rev., 115,
at 0.5 V, the system cost for Fe-AQDS is projected to be 60% of the 11533 (2015).
7. A. Z. Weber, M. M. Mench, J. P. Meyers, P. N. Ross, J. T. Gostick, and Q. Liu,
cost of the VRFB system operating at 1.1 V. Thus, with further “Redox flow batteries: a review.” J. Appl. Electrochem., 41, 1137 (2011).
increase in power densities combined with the attributes of low cost 8. Grid-Scale Rampable Intermittent Dispatchable Storage (GRIDS) (ARPA-E,
and high durability, the iron sulfate/AQDS system could become a Department of Energy (DOE)) DE-FOA-0000290 (2010).
unique and attractive candidate for meeting the LCOS target for 9. Steel Statistical Yearbook 2017 (World Steel Association) (2017).
10. X. Wei, G.-G. Xia, B. Kirby, E. Thomsen, B. Li, Z. Nie, G. G. Graff, J. Liu,
large-scale energy storage systems. V. Sprenkle, and W. Wang, “An aqueous redox flow battery based on neutral alkali
metal ferri/ferrocyanide and polysulfide electrolytes.” J. Electrochem. Soc., 163,
Conclusions A5150 (2016).
11. K. Gong, F. Xu, J. B. Grunewald, X. Ma, Y. Zhao, S. Gu, and Y. Yan, “All-soluble
An iron/AQDS flow battery, based on relatively inexpensive all-iron aqueous redox-flow battery.” ACS Energy Lett., 1, 89 (2016).
redox materials was found to be rechargeable over hundreds of 12. B. Hu, C. DeBruler, Z. Rhodes, and T. L. Liu, “Long-cycling aqueous organic
cycles with an extraordinarily low rate of decline in capacity. Such redox flow battery (AORFB) toward sustainable and safe energy storage.” J. Am.
Chem. Soc., 139, 1207 (2017).
low rates of capacity loss are not encountered with other flow 13. E. S. Beh, D. De Porcellinis, R. L. Gracia, K. T. Xia, R. G. Gordon, and M. J. Aziz,
batteries based on organic redox couples. Using an electrolyte “A neutral pH aqueous organic–organometallic redox flow battery with extremely
mixture of iron sulfate and AQDS in a symmetric cell configuration high capacity retention.” ACS Energy Lett., 2, 639 (2017).
we could achieve a coulombic efficiency of 99.63% and avoid the 14. X. Xing, Y. Zhao, and Y. Li, “A non-aqueous redox flow battery based on tris(1,10-
phenanthroline) complexes of iron(II) and cobalt(II).” J. Power Sources, 293, 778
permanent capacity fade encountered in the asymmetric configura- (2015).
tion. Despite the oxidative properties of iron(III), AQDS was found 15. K. L. Hawthorne, J. S. Wainright, and R. F. Savinell, “Studies of iron-ligand
to be chemically stable in the mixed electrolyte solution. Current complexes for an all-iron flow battery application.” J. Electrochem. Soc., 161,
densities as high as 1.16 A cm−2 could be supported at the A1662 (2014).
16. Y.-W. D. Chen, K. S. V. Santhanam, and A. J. Bard, “Solution redox couples for
electrodes, demonstrating that the electrochemical reaction step electrochemical energy storage.” J. Electrochem. Soc., 128, 1460 (1981).
and the mass transport of reactants were fast. We could achieve 17. L. W. Hruska and R. Savinell, “Investigation of factors affecting performance of the
power densities as high as 194 mW cm−2 with an un-optimized cell iron-redox battery.” J. Electrochem. Soc., 128, 18 (1981).
design. We found that the principal voltage losses arose from the 18. K. L. Hawthorne, T. J. Petek, M. A. Miller, J. S. Wainright, and R. F. Savinell, “An
investigation into factors affecting the iron plating reaction for an all-iron flow
ohmic resistance of the electrode and electrolyte. Thus, the reduction battery.” J. Electrochem. Soc., 162, A108 (2015).
of ohmic resistance by almost 40%, achieved by altering the 19. B. S. Jayathilake, E. J. Plichta, M. A. Hendrickson, and S. R. Narayanan,
composition of the carbon-nanotube modified electrodes resulted “Improvements to the coulombic efficiency of the iron electrode for an all-iron
in an increase of the discharge voltage leading to an increase in the redox-flow battery.” J. Electrochem. Soc., 165, A1630 (2018).
20. N. H. Hagedorn and L. H. Thaller, Power Sources 8, 227 (1981).
power density and energy efficiency. With these characteristics, the 21. L. H. Thaller, Proc. 9th Intersoc. Energy Conv. Eng. Conf., NASA TM X‐71540, 924
iron/AQDS battery overcomes the challenges encountered with other (1974).
iron-based flow batteries. Despite the relatively lower cell voltage 22. R. F. Gahn, N. H. Hagedorn, and J. A. Johnson, “Cycling performance of the iron-
for this system compared to the vanadium flow battery, the chromium redox energy storage system.” NASA TM-87034, NASA, Dept. of Energy,
US (1985).
extraordinary durability of the iron/AQDS battery combined with 23. C. Horne, R. Mosso, T. Smith, and B. Adams, “Demonstration of EnerVault Iron-
the low cost of materials, presents a unique opportunity for further Chromium Redox Flow Battery.” California Energy Commission (2014).
optimization of electrodes, reactant concentrations, and cell config- 24. M. C. Tucker, A. Phillips, and A. Z. Weber, “All-iron redox flow battery tailored for
urations for meeting the requirements of “mega”-scale energy off-grid portable applications.” Chem. Sus. Chem., 8, 3996 (2015).
25. (Taian Health Chemical Co. Ltd.) Ferrous Sulfate Heptahydrate https://alibaba.com/
storage applications. product-detail/Ferrous-Sulfate-Ferrous-sulphate-Iron-sulphate_60413589328.html?
spm=a2700.7724838.2017115.73.841e1e92opchUi.
Acknowledgments 26. B. Yang, L. Hoober-Burkhardt, F. Wang, G. K. Surya Prakash, and S. R. Narayanan,
“An inexpensive aqueous flow battery for large-scale electrical energy storage
The authors thank ARPA-E (Contract #DE-AR0000337) the based on water-soluble organic redox couples.” J. Electrochem. Soc., 161, A1371
Loker Hydrocarbon Research Institute and the Department of (2014).
27. B. Huskinson, M. P. Marshak, C. Suh, S. Er, M. R. Gerhardt, C. J. Galvin, X. Chen,
Chemistry, University of Southern California for financial support. A. Aspuru-Guzik, R. G. Gordon, and M. J. Aziz, “A metal-free organic-inorganic
aqueous flow battery.” Nature, 505, 195 (2014).
Conflict of Interest Statement 28. M. R. Gerhardt, L. Tong, R. Gómez-Bombarelli, Q. Chen, M. P. Marshak,
C. J. Galvin, A. Aspuru-Guzik, R. G. Gordon, and M. J. Aziz, “Anthraquinone
Authors SRN and GKS are principals in Battery Energy Storage derivatives in aqueous flow batteries.” Adv. Energy Mater., 7, 1601488 (2017).
Technology LLC, a licensee of organic redox flow battery tech- 29. T. J. Carney, S. J. Collins, J. S. Moore, and F. R. Brushett, “Concentration-
nology from the University of Southern California. We acknowledge dependent dimerization of anthraquinone disulfonic acid and its impact on charge
storage.” Chem. Mater., 29, 4801 (2017).
the University of Southern California’s financial interest in Battery 30. L. Hoober-Burkhardt, S. Krishnamoorthy, B. Yang, A. Murali, A. Nirmalchandar,
Energy Storage Technology LLC. G. K. S. Prakash, and S. R. Narayanan, “A new michael-reaction-resistant
benzoquinone for aqueous organic redox flow batteries.” J. Electrochem. Soc.,
ORCID 164, A600 (2017).
31. A. Murali, A. Nirmalchandar, S. Krishnamoorthy, L. Hoober-Burkhardt, B. Yang,
S. R. Narayanan https://orcid.org/0000-0002-7259-3728 G. Soloveichik, G. K. S. Prakash, and S. R. Narayanan, “Understanding and
mitigating capacity fade in aqueous organic redox flow batteries.” J. Electrochem.
References Soc., 165, A1193 (2018).
32. B. Yang, L. Hoober-Burkhardt, S. Krishnamoorthy, A. Murali, G. K. S. Prakash,
1. IEA, Renewables 2017, Analysis and Forecasts to 2022 (International Energy and S. R. Narayanan, “High-performance aqueous organic flow battery with
Agency, Paris) (2017). quinone-based redox couples at both electrodes.” J. Electrochem. Soc., 163, A14
2. M. Azevedo, N. Campagnol, T. Hagenbruch, K. Hoffman, A. Lala, and (2016).
O. Ramsbottom, Lithium and Cobalt—A Tale of Two Commodities (McKinsey & 33. J. D. Milshtein, “Cost-targeted design of redox flow batteries for grid storage.”
Company Met. Min., Tokyo) (2018). https://arpa-e.energy.gov/sites/default/files/Panel%202%20-%20Milshtein.pdf
3. S. R. Narayanan, G. K. S. Prakash, A. Manohar, B. Yang, S. Malkhandi, and (2017).
A. Kindler, “Materials challenges and technical approaches for realizing inexpen- 34. V. Dieterich, J. D. Milshtein, J. L. Barton, T. J. Carney, R. M. Darling, and
sive and robust iron–air batteries for large-scale energy storage.” Solid State Ionics, F. R. Brushett, “Estimating the cost of organic battery active materials: a case study
216, 105 (2012). on anthraquinone disulfonic acid.” Transl. Mater. Res., 5, 034001 (2018).
Journal of The Electrochemical Society, 2020 167 060520

35. R. M. Darling, A. Z. Weber, M. C. Tucker, and M. L. Perry, “The influence of 39. K. Dennis, N. Coad, P. Lex, and J. A. Reichard, “Reversible polarity operation and
electric field on crossover in redox-flow batteries.” J. Electrochem. Soc., 163, switching method for ZnBr flow battery when connected to common DC bus.”
A5014 (2016). US9570753B2 (2012).
36. K. W. Knehr, E. Agar, C. R. Dennison, A. R. Kalidindi, and E. C. Kumbur, “A 40. M. L. Perry, R. M. Darling, and R. Zaffou, “High power density redox flow battery
transient vanadium flow battery model incorporating vanadium crossover and water cells.” ECS Trans., 53, 7 (2013).
transport through the membrane.” J. Electrochem. Soc., 159, A16 (2012). 41. S. Ha and K. G. Gallagher, “Estimating the system price of redox flow batteries for
37. D. C. Sing and J. P. Meyers, “Direct measurement of vanadium crossover in an grid storage.” J. Power Sources, 296, 122 (2015).
operating vanadium redox flow battery.” ECS Trans., 50, 61 (2013). 42. M. Zhang, M. Moore, J. S. Watson, T. A. Zawodzinski, and R. M. Counce, “Capital
38. R. A. Potash, J. R. McKone, S. Conte, and H. D. Abruña, “On the benefits of a cost sensitivity analysis of an all-vanadium redox-flow battery.” J. Electrochem.
symmetric redox flow battery.” J. Electrochem. Soc., 163, A338 (2016). Soc., 159, A1183 (2012).

You might also like