You are on page 1of 11

HardwareX 6 (2019) e00072

Contents lists available at ScienceDirect

HardwareX
journal homepage: www.elsevier.com/locate/ohx

Open source all-iron battery for renewable energy storage


Nicholas Yensen, Peter B. Allen ⇑
University of Idaho, United States

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

Article history: The price of renewable energy is dropping rapidly. Energy storage will be needed to take
Received 13 March 2019 full advantage of abundant but intermittent energy sources. Even with economies of scale,
Received in revised form 4 July 2019 the price is prohibitively high for a lithium-ion battery pack capable of storing tens of kilo-
Accepted 9 July 2019
watts of energy for many consumers. A more abundant and less expensive material is nec-
essary. All-iron chemistry presents a transformative opportunity for stationary energy
storage: it is simple, cheap, abundant, and safe. All-iron batteries can store energy by
Keywords:
reducing iron (II) to metallic iron at the anode and oxidizing iron (II) to iron (III) at the cath-
Electrochemistry
Non-toxic
ode. The total cell is highly stable, efficient, non-toxic, and safe. The total cost of materials
Low cost is $0.1 per watt-hour of capacity at wholesale prices. This battery may be a useful compo-
Eco-friendly nent of open source hardware projects that require a safe and ecologically friendly battery.
Iron This is also one of the few battery chemistries that can be built safely in a DIY setting.
Ó 2019 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Specifications table

Hardware name Open Source All-Iron Battery


Subject area  Chemistry and Biochemistry
Hardware type  Electrical engineering and computer science
Open Source License CC-BY-3.0
Cost of Hardware $1 per 50 g cell; cost of set-up $300
Source File Repository https://osf.io/4jvr9/

1. Hardware in context

The all-iron battery is an electrochemical cell for powering an electronic device. It contains two chemical reagents, one of
which is oxidized and the other is reduced. The result is current flow through a connected electrical load. The all-iron cell is
similar to historical electrochemical cells like the Edison cell (iron-nickel, first developed in 1901). Commercial rechargeable
batteries use a nickel cathode and a metal hydride cathode (NiMH type) or an iron phosphate cathode and a lithiated gra-
phite anode (Lithium-ion type) [1]. Iron batteries have lower specific energy (watt-hours per kilogram) than these commer-
cial cells but have low-cost reagents and present opportunities for simpler construction and experimentation. Lithium and

⇑ Corresponding author.
E-mail address: pballen@uidaho.edu (P.B. Allen).

https://doi.org/10.1016/j.ohx.2019.e00072
2468-0672/Ó 2019 Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
2 N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072

metal hydrides are toxic, flammable, and react violently with water and corrode in air. Iron is comparatively non-toxic and is
only slightly reactive with water and air (i.e., it rusts).
Iron batteries have been published previously and are being deployed commercially (e.g., ESS, inc [2]). Various iron che-
mistries have been used in flow batteries [3]. Flow batteries use solutions of the redox-active materials pumped through an
electrochemical cell. They are flexible: the energy capacity (limited by the size of the chemical storage tank) is independent
of the power capacity (limited by the size of the cell in which the chemical reactions happen). An example of an all-iron flow
battery includes a soluble flow battery by Yan and co-workers [4]. Another flow battery uses an iron powder slurry as the
anode chemistry [5]. One flow battery was designed for use in off-grid settings [6]. Flow batteries have the disadvantage that
they require pumps and plumbing to bring the stored chemistry into an electrochemical flow cell to permit charging or dis-
charging. A non-reversible iron battery was presented in 2016 as a possible source of energy for recharging phones or other
devices in remote locations [7]. A non-flow ‘‘scrap metal cell,” also known as ‘‘The Vanderbilt Battery,” was also presented in
2016 based on iron and brass (commonly found in scrapyards) [8].
We present here a rechargeable all-iron battery with an iron metal anode and an iron (III) sulfate cathode. It is based on
aqueous chemistry and so is not flammable. The chemical and hardware components are inexpensive and are simpler to
build and operate than a flow battery. Once assembled, it has a neutral pH and is very safe and environmentally friendly.
Our all-iron battery still considerably more practical than zinc/copper cells (e.g., available as an educational kit by hyLyte
Education [9]) or the typical ‘‘potato cell” batteries that are sometimes used in classrooms. Neither demonstration cell is
rechargeable as the oxidation of zinc is not reversible in water.

2. Hardware description.

The all-iron galvanic electrochemical cell discharges and liberates energy (Fig. 1A). During discharge, iron oxidizes at the
anode and reduces an iron salt at the cathode. Our design uses steel wool (anode) and a precipitated ferric iron salt (cathode)
plus carbon felt current collectors and graphite electrodes. At the most basic level, the half reactions were designed as
follows, at the anode:

Fe ! Fe2þ þ 2e ð1Þ


And at the cathode:

e þ Fe3þ ! Fe2þ ð2Þ


In practice, the reactions are more complex. The open cell potential of this cell is 0.5–0.7 V (depending on the state of
charge). This electrochemical potential is consistent with a number of possible iron redox reactions. One hypothesis for
the reactions of the cell is as follows.
At the anode, iron is oxidized to soluble Fe2+. The half-reaction standard oxidation potential of this reaction is 0.45 V:

Fe ! Fe2þ þ 2e ð3Þ


At the cathode, ferric iron oxide is reduced to ferrous iron hydroxide with a standard reduction potential of 0.16 V:

Fe2 O3 þ 2e þ 4Hþ ! FeOHþ þ H2 O ð4Þ


This gives a cell potential of 0.61 V at standard conditions (per CRC Handbook [10]) in rough agreement with our results.
However, we cannot yet conclude that these specific reactions are occurring. Any of several precipitates with mixed oxida-
tion states could be involved (e.g., Fe2O3, Fe3O4, Fe(OH)2, FeSO4, and Fe2(SO4)3 to name a few).
The sodium polyacrylate [CH2–CH(CO2Na)–]n separator membrane allows positive ions (sodium and potassium) to pass
through to maintain a neutral electrical charge. All components are assembled into a plastic pouch (‘‘pill pouch”) that can be

Fig. 1. (A): The all-iron battery schematic, showing all major components that interact. (B): A demonstration of a single cell and it’s voltage after charging.
N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072 3

sealed for both ease of construction and safety. This pouch allows easy connection to other cells, whether in series or in
parallel.

2.1. Iron-sulfate salt chemistry

We found an iron and sulfate solution to be a stable and reliable salt chemistry for the all-iron battery. Iron chloride was
mixed with a saturated potassium sulfate solution and then pH was adjusted. This generated a precipitate. Iron (II) chloride
was used to produce the anode electrolyte. Iron (III) chloride was used as the cathode electrolyte. Although iron (IV) can be
produced in principle, the safety concerns associated with high oxidation state materials precluded its use [11]. Other iron
salt solutions were tested as well, including solutions containing sucrose, fructose, EDTA, and ferricyanide. In solution, these
alternatives behaved in a similar manner, but a sulfate precursor had the best available coulombic capacity. The sulfate sus-
pension also produces the least loss of iron from the metallic iron after repeated discharges and charges. This indicates highly
reversible chemistry (discussed below under Validation and Characterization).
The iron anode is oxidized in the reaction, liberating electrons at the negative side of the cell. Other cells have used iron
anodes including the classic Edison cell (with a nickel cathode) and a more recent disposable iron battery for low-resource
settings [7]. In our cell, the electrons are taken up by an iron (III) cathode electrolyte at the positive side of the cell. In the
fully discharged state, iron (II) is present throughout. We found that abundant iron (II) in the anode electrolyte helped to
prevent irreversible loss of the iron and allowed for greater reversibility. This design is robust to leakage and mixing. If
the anode electrolyte and cathode electrolyte mix, the battery is discharged, but not damaged by contamination.

2.2. Sodium polyacrylate membrane

All galvanic cells must separate their two half-reactions. The purpose of a separator membrane is to maintain charge bal-
ance by passing ions while preventing electronic conductivity which discharges the cell. In other cell types, the separator
also serves to keep the anode electrolyte and cathode electrolyte from cross-contaminating each other. A good membrane
does not permit large molecules to pass through (like the iron-sulfate salt found in our all-iron battery). The all-iron battery
has consistent chemistry throughout and so cross-contamination does not permanently degrade the cell. However, the dif-
fusion of the iron species through the membrane does discharge the cell and should be minimized.
The membrane with the best cost-performance ratio was the sodium polyacrylate membrane. It features polyacrylate,
which is found in diapers as a super water absorber. Sodium polyacrylate absorbs many times its weight in water. To make
a viable membrane from sodium polyacrylate, we dried it onto a folded piece of cellulose chromatography paper. We tested
multiple membrane materials, including agar, Nafion, polyacrylic acid, nanoporous polycarbonate, and others. Most were
either too low in performance or too high in cost for performance. The commercial grade Nafion performed best, but it
was prohibitively expensive. However, there are other commercially available membranes as options, aside from the mem-
branes we tested (see Fig. 4). Fumatech produces ion exchange membranes that may be suitable (e.g., sulfonated polyether
ether ketone membrane E360). The synthesis of other membranes can be found in the literature [12,13]. While such mem-
branes may present advantages, their synthesis is challenging and presents significant safety hazards.
All membrane tests were done in a concentration cell. The premise of the test was to set-up the membrane to separate
two sides of an acrylic chamber. The anode electrolyte and cathode electrolyte were put on either side of the membrane, and
the potential was measured until both sides had completely equilibrated through the membrane. The concentration cell con-
sisted of several pieces of clear acrylic plastic sealed together. This method of testing allowed less variation between tests, as
the enclosure would not change, and the same amount of anode electrolyte and cathode electrolyte could be added. Addi-
tionally, this concentration cell was easy to clean and was inert to the iron battery chemistry. Plans for constructing the
membrane testing cell are provided (see Optimization of Membrane, below).

2.3. Carbon-felt electrode

The purpose of the electrode in any battery is to act as the interface between the reaction and the load. The carbon-felt
electrode acts as a high surface area interface between the cathode electrolyte and the graphite current collector. The cath-
ode electrolyte is able to permeate the void spaces in the carbon felt, which reduces the path length over which electrons
must travel in the cathode electrolyte. The felt is made of carbon in the form of graphite, which is a good conductor and
resists corrosion. Corrosion resistance adds to the longevity of the battery. Other carbon-based electrodes are available
(e.g., rods, powder), but the pouch-style cell is best suited to a sheet of material. Flat graphite sheets alone had a low surface
area that limited reaction rates; carbon felt improved performance.

2.4. Graphite sheet current collector

The current collector serves as the ‘‘wires” for the battery. The best current collector for the all-iron battery was a thin,
flexible graphite foil. It has low resistance and is simple to connect between cells in series. Additionally, graphite is less prone
to corrode than metal wires. Even corrosion resistant chrome-nickel wires corroded after extended use in the all-iron
battery.
4 N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072

2.5. Possible applications

Iron batteries will be useful in stationary applications. For projects where an environment is very sensitive or where fire
hazards are prohibited, the iron battery may be a good match. We envision projects such as open source wind turbines
(e.g., the Zoetrope [14]) and open source charge controllers (e.g., Libre Solar [15,16]) could be integrated with an open source
battery. This could reduce the barriers to entry for innovative business models in renewable energy and energy storage. The all-
iron battery could replace lithium batteries where cost and fire risk are more important than specific energy. Lithium chemistry
has a high specific energy and power density. It is perfect for power-demanding mobile applications where high power and low
weight are required. In applications where weight is less important, cheaper and safer chemistry may be preferred.
An open source all-iron battery may also find application in scientific research and education. In environmentally sensi-
tive or fire-sensitive locations, this chemistry might be more appropriate than lithium to power data loggers or remote cam-
eras. Building an open source all-iron battery is also a chance for students to learn about chemistry and energy storage. This
system is accessible to an undergraduate-level chemist (unlike lithium batteries which require specialized equipment and
facilities). It is still considerably more practical than the typical ‘‘potato cell” batteries that are sometimes used in classrooms,
which are not rechargeable and tend to spoil or dry out quickly.

3. Design files

3.1. Bill of materials

The following is for a 3 V battery, consisting of 6 cells.

Designator Component Number Cost per unit - Total cost - Source of Material
currency currency materials type
FeSalt1 Ferrous Chloride (FeCl2) 17.4 g $0.425/g $42.50 1. Amazon Inorganic/
$0.0031/g $0.31* 2. Alibaba Metal
$3.25/g $81.20* 3. Sigma Aldrich
(372870-25G)
FeSalt2 Ferric Chloride (FeCl3) 19.2 g $0.097/g $10.99 1. Amazon Inorganic/
$0.0005/g $0.50* 2. Alibaba Metal
$0.405/g $40.50* 3. Sigma Aldrich
(157740-100G)
Salt Potassium Sulfate 20.9 g $0.11/g $12.67 1. Amazon 1 Inorganic
(K2SO4) $0.02/g $8.99* 2. Amazon 2
$0.01/g $10.00* 2. Alibaba
$0.22/g $55.00* 3. Sigma Aldrich
(P0772-250G)
Base Sodium Hydroxide 15.0 g $0.062/g $6.99 1. Amazon Inorganic
(NaOH) $0.0007/g $0.73* 2. Alibaba
$1.58/g $39.50 3. Sigma Aldrich
(221465-25G)
Acid **Hydrochloric Acid <5 mL $0.09/mL $44.95 1. Amazon Inorganic
(HCl) $0.30/g $300.00* 2. Alibaba
$0.14/mL $68.00* 3. Sigma Aldrich
(320331–500 mL)
Poly Sodium polyacrylate 0.3 g $0.12/g $6.00 1. Amazon Polymer
(C3H3NaO2)n $0.0035/g $3.50* 2. Alibaba
$0.45/g $44.50* 3. Sigma Aldrich
(447013-100G)
FeMetal Steel Wool (Fe) 6.7 g $0.064/g $6.10 1. Amazon Metal
$0.45/roll $0.45* 2. Alibaba
Pch1 Pill Pouches 6 $0.11/piece $5.47* 1. Amazon Composite
Pch2 Chromatography Paper 6 $0.00177/cm2 $23.55* 1. Amazon Biomaterial
(10x10) cm2
Pch3 **Polyethylene Sheet 6 $0.03/cm2 $24.48* 1. Amazon Composite
(5  4) cm2
Pch4 Clear Plastic Sheet 6 $0.012/cm2 $10.58 1. Amazon Composite
(5  4) cm2
N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072 5

(continued)

Designator Component Number Cost per unit - Total cost - Source of Material
currency currency materials type
Ele1 Graphite Felt (5  4) 6 $0.03/cm2 $11.99* 1. Alibaba Nanomaterial
cm2
Ele2 Graphite Foil (25  25) 1 $0.019/cm2 $17.50 1. Amazon Nanomaterial
cm2 $10/piece $100.00* 2. Alibaba
$0.108/m2 $47.79* 3. Other
Totals (Lowest): $7.48/6 cells $83.57
Totals (Highest): $115.1/6 cells $742.87
Totals (Recommended): $35.63/6 cells $223.77
*Does not include shipping and handling costs. For Sigma Aldrich, the freight shipping cost is $45 per entire order.
**Recommended but not necessary.

4. Build instructions

Short build instructions are presented here. We highly recommend using the complete build instructions in the Supple-
mental material for greater detail and to avoid mistakes in preparation.

4.1. Chemical solutions

There are five solutions that must be prepared: 1 M potassium sulfate, or salt of potash, (K2SO4), 10 M sodium hydroxide,
or lye, (NaOH), 10 M hydrochloric acid (HCl), cathode electrolyte, and anode electrolyte. The first three solutions listed above
are needed to make the anode electrolyte and cathode electrolyte. This means that it is best to make the 1 M K2SO4 first, then
the 10 M NaOH and 10 M HCl, and then the anode electrolyte and cathode electrolyte. Reading the supplemental build
instructions is highly recommended.

4.1.1. Anode electrolyte


The anode electrolyte is the solution directly in contact with the anode (iron metal) electrode. Approximately 60 mL is
needed for the six-cell battery. Anode electrolyte solution consists of 2 M iron (II) chloride (FeCl2) (17.4 g in 60 mL of 1 M
K2SO4) at pH 7.5. To prepare the anode electrolyte, measure 60 mL of 1 M K2SO4 solution. Dispense powdered FeCl2 [FeSalt1]
into the beaker with 1 M K2SO4 solution granule-wise, avoiding clumping of the FeCl2 (Iron (II) chloride, FeCl2) [FeSalt1].
Caution: Care when handling must be exercised. Gloves are required, as well as safety glasses or goggles, and a lab-coat or other
form of full body personal protective equipment (PPE). A solution of 10 M NaOH [Base], (24.0 g in 60 mL of water) is used to
adjust to pH 7.5 by adding small amounts, around 0.5 mL at a time. The pH measurements should be made after every
addition of NaOH. Approximately 12 mL of NaOH is required.

4.1.2. Cathode electrolyte


The cathode electrolyte is the solution directly in contact with the cathode (graphite felt) electrode. The cathode elec-
trolyte solution preparation is similar to the anode electrolyte solution preparation. This solution consists of 2 M iron (III)
chloride (FeCl3) [FeSalt2] (19.2 g in 60 mL of 1 M K2SO4) at pH 7.5. Preparation of the cathode electrolyte solution proceeds
identically to the anode electrolyte solution. It is worth noting that the FeCl3 [FeSalt2] powder is harder to manipulate than
the FeCl2 [FeSalt1] because the powder adheres to itself. A small spatula can be used to dispense the FeCl3 [FeSalt2] into the
solution. Caution: Care when handling must be exercised. Gloves are absolutely necessary, as well as closed goggles, and a lab-coat
or other form of full body PPE. Iron (III) chloride (ferric chloride) is highly corrosive. When adjusting to pH 7.5, the solution
becomes very viscous, to the point where it is difficult to stir. Manual stirring with a glass rod is recommended, along with
a magnetic stir-bar. Approximately 45 mL of NaOH are required to adjust to pH 7.5.

4.2. Battery cell preparation

Once the anode electrolyte and cathode electrolyte solutions are made, the cell can be prepared. This includes the mem-
brane, steel wool anode [FeMetal], graphite felt cathode [Ele1], and the graphite foil electrodes [Ele2]. Altogether, this will
create the parts to assemble the cell. See Fig. 2A–H for visual reference.

4.2.1. Membrane preparation


The membrane consists of a folded piece of cellulose chromatography paper [Pch2], approximately 10 cm  10 cm. It
should be handled on clean surfaces and with gloves, to prevent contaminating the surface. The paper is folded into an envel-
ope to fit into the plastic pouch [Pch1], see Fig. 2B. Sodium polyacrylate solution is prepared to a concentration of 150 mg in
1.5 L of solution (100 parts per million, ppm). This is enough for a six-cell battery. A polyethylene spacer [Pch3] was placed
6 N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072

Fig. 2. A visual reference for the assembly of the pouch and battery.

inside the pocket of the folded cellulose paper. This prevents the two sides from sticking together when wet. Next, half of the
100-ppm sodium polyacrylate solution is poured over the folded pieces of cellulose paper, completely wetting them. The
water must be evaporated, leaving behind the sodium polyacrylate dispersed throughout the fibers of the paper. This can
be accomplished by drying in air over 2–3 days or drying for 6 h in a warm oven heated to 80 °C. The paper should be gently
flipped over while moist. Any remaining solution should be added to the paper. Any torn pouches should be discarded. Store
membranes in a dry place and handle gently to avoid scraping off the polyacrylate crystals.

4.2.2. Steel wool preparation


Steel wool [FeMetal] is used for the anode of the cell. For a six-cell battery, 6.7 g of steel wool [FeMetal] are needed, which
constitutes roughly 1.11 g per cell. Fine steel wool can be used without further preparation.

4.2.3. Graphite felt preparation


Graphite felt [ELE 1] is used as the cathode of the cell. Thin felt (5 mm thick) works best. Six pieces of the graphite felt [ELE
1] should be cut to dimensions: 5 cm wide, 4 cm tall. Caution: The graphite felt can leave micro slivers if touched with bare
hands, which can cause discomfort. Gloves are highly recommended.

4.2.4. Graphite foil preparation


The graphite foil [ELE 2] acts as the current collector for the anode and the cathode. The graphite foil [ELE 2] can be pur-
chased as a 25 cm  25 cm sheet or as a roll of foil. It should be cut with a sharp utility knife or scissors. Cut twelve strips
with dimensions of 6.25 cm  5 cm and four strips of dimensions 12.5 cm  5 cm. It is necessary to cut extra strips, as the foil
is delicate and ripped strips should be discarded.

4.3. Battery cell dry assembly

Using all the materials that have been prepared, the dry part of the cell can be assembled (excluding the anode electrolyte
and cathode electrolyte). Use Fig. 2 for reference. Slide the membrane envelope into the pill pouch [PCH 1]. Following this,
orient the membrane envelope correctly. The side of the membrane envelope with folded flaps is the ‘‘non-working” side.
The opposite side of the membrane envelope with the smooth side is the ‘‘working” side. Next, a piece of steel wool with
mass 1.11 g (or slightly greater) and the pre-cut graphite foil is inserted into the membrane envelope. Next, the pre-cut gra-
phite felt, and graphite foil is also slipped into the pouch touching the working side of the membrane envelope. This is the
finished, dry part of the cell. The procedure should be repeated five more times. Two ‘‘end” cells (the first and last in the
array) should be constructed with a longer strip of graphite foil. At one end of the array, a long foil is used for the cathode.
At the far end of the array, a long strip of graphite foil is used for the anode. Longer strips may be reinforced with a sheet of
plastic backing to avoid tearing the sheet.

4.4. Add electrolyte to complete battery of cells

Anode electrolyte and cathode electrolyte are added to the assembled dry cells with syringes. Each cell should be filled
with 8 mL of anode electrolyte and 8 mL of cathode electrolyte. It is best to keep the same volume between both sides.
The cell should be opened by pinching the edges. Caution: Contact with both the anode electrolyte and cathode electrolyte is
possible here. Gloves are recommended, as well as safety glasses or goggles. An initial 5 mL of anode electrolyte is aspirated
and slowly injected to the anode side of the cell using a 5 mL syringe. The second volume of 3 mL is then added. The goal
is to make sure all the steel wool is saturated in anode electrolyte, and that it is in direct contact with the membrane. Next, a
fresh syringe is used to add cathode electrolyte to the cathode (carbon felt) side of the cell in a similar manner. Once both
N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072 7

anode electrolyte and cathode electrolyte are injected, the cell is finished, and care must be taken not to allow the graphite
foil electrodes to touch, as this will short the battery and diminish performance. The other 5 cells are prepared likewise.

4.5. Assemble battery

Once the cells are completed, the battery can be assembled and connected. The objective is a 3 V battery, made of six cells.
The open cell potentials of each cell should be measured with a digital multimeter. Cell potentials should stabilize after about
10 to 15 min. Any cells below 300 mV should be considered defective and should be discarded. Caution: Contact with both the
anode electrolyte and cathode electrolyte is possible here. Gloves and safety glasses or goggles are recommended. Following cell
testing, obtain a plastic enclosure for the battery. The cells must always stay upright inside the enclosure. Once all cells were
inside the enclosure, arrange the cells such that the extra-long graphite foils are at the ends of the cell array. All the cells are
electrically connected in series, cathode to anode, with small binder clips. A damp paper towel should be placed away from
the graphite foil electrodes but still inside the enclosure. The paper towel should be checked and periodically wetted with
water to help slow evaporation.

4.6. Sealing pouches for long term use

Hot glue can also be used to seal the pouches and prevent evaporation. A clear plastic sheet is placed between the gra-
phite foil electrodes to keep the electrodes from touching. Two toothpicks are taped to the bag’s closure mechanism to hold
the bag flat while upright. The pouch may contain a small air space. Then, the outside of the pouch is glued shut with a thin
bead of hot glue. The two electrodes and the plastic clear sheet are left outside.

5. Operation instructions

5.1. Operation tips and safety concerns

Once the battery is completely built, it is safe to touch the enclosure and graphite electrodes without gloves, safety
glasses, or goggles. Care must still be exercised to ensure that the cells and battery remain upright. Otherwise, the cells
may leak or mix the anode electrolyte and cathode electrolyte. If the battery is tilted too much or dropped, it is likely it will
no longer work as desired and may have very diminished performance. However, this problem could be solved by recharging
the cell. If anything must be modified inside the battery, gloves and goggles are necessary. Extremes in temperature are
expected to have a detrimental effect on the battery. When handling and connecting wires to the graphite sheet electrodes,
much care must be taken not to rip the strips. Ripping the strips will call for a difficult repair or construction of a new cell.

5.2. Operational instructions

Batteries are generally operated in a circuit, and the operation will depend on the application. With the exception of its
unique geometry, the battery operates like any commercial battery. The anode electrode should be connected on the nega-
tive, or ground, terminal and the cathode should be connected on the positive, or hot, terminal. It was found that the best
way to connect the terminals was with alligator clips, as they gripped the graphite sheets well, but did not destroy them.
Self-adhesive copper tape can be applied to reinforce the graphite prior to connection with clips.

5.3. Charging instructions

A cell is constructed fully charged according to the instructions above. After use, it has the capability of being charged by
two methods, which both require a direct current source. In our experiments, we used a benchtop laboratory regulated DC
power supply (TP3005T Tekpower, Montclair, CA), or a WaveNow potentiostat (Pine Research Instrumentation, Durham, NC).
To charge the six-cell battery using our regulated DC power supply, the leads were connected to the battery, with positive to
the cathode and negative to the anode. The voltage was set to 3.0 V and was then ‘‘locked.” The current was then adjusted to
3 mA maximum, and the battery charged until current read 1 mA on the power supply. A single cell should not be charged
at a higher applied voltage than 1.2 V (7.2 V for the 6 cells in series). At higher voltages, hydrogen and oxygen production by
the electrolysis of water may be significant and can present an explosion hazard. A WaveNow potentiostat was also used to
charge the device. The potentiostat leads were connected, and the battery was charged using a constant current potentiom-
etry experiment in the WaveNow software. We charged at 1 mA until the battery potential reached 3.00 V.

6. Validation and characterization

The basic parts to the all-iron battery are outlined in Fig. 1A, and for each of these components, there are many options.
We analyzed the effect of modifying each component on the performance of this electrochemical cell. We chose pH 7.5 based
on performance losses at higher and lower pH; the cell is not tolerant of a pH change of 0.5 pH units or larger. The separator
8 N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072

was optimized for a balance of cost and performance. The choice of carbon electrodes was also explored. While many com-
binations of components of the all-iron battery could function, some combinations offer greater advantages.

6.1. Choice of iron-sulfate

The iron-sulfate salt was one of five iron salts that were tested. The experimental set-up was identical for each of the five
salts, the only variable being the salt itself. The set up was simple, as it consisted of a very basic galvanic cell, with each half
reaction chamber being a separate beaker. As a membrane, a polyacrylamide, 0.1 M sodium chloride (NaCl) salt bridge was
used. The polyacrylamide helped sustain the salt in place, and still permitted the balancing of charge. As opposed to the
pouch, the beaker cells did not use graphite foil as a current collector, but rather anti-oxidative chromel wire. The source
of iron metal was 99.9% iron wire, coiled to fit the beaker. Like the pouch, carbon felt was used as the cathode. All 5 salts
were tested at pH 8, and in approximately 10 mL beakers.
There were 4 metrics by which the iron salts were evaluated. 1. Overall coulombic discharge capacity. This test was done
by setting a constant current discharge at 1 mA until the cell potential became less than 10 mV. 2. Stability and repeatability
in 12 cycles of charging and discharging. This was performed by setting the cell to discharge for 1 h at 1 mA and then charg-
ing for 1 h at 1 mA. 3. The open cell potential upon assembly of the cell. 4. The percent of iron lost to the solution. As the
battery cycles through discharges and charges, it was possible that some iron might not be reduced back onto the surface
of the wire caused by charging.
The iron-sulfate salt yielded the best performance out of the 5 salts tested. Fig. 3 shows that the open cell potential was
the best, along with the most coulombs that were discharged until failure. While all the iron-salt configurations contain the
same amount of iron that can be used to discharge the battery, the iron-sulfate salt clearly made more iron accessible than
any of the other salts. The iron anode loss as a percent was determined to be negative, which indicates that the precipitate
more easily allowed the reverse reaction during the charging cycles to occur.
Given the possibility that precipitated iron sulfate is the optimal electrode material, we also tested iron sulfate directly.
We dissolved 1 M iron (II) sulfate in 1 M sodium sulfate (for the anode electrolyte) and 1 M iron (III) sulfate in 1 M sodium
sulfate (for the cathode electrolyte). The iron sulfate cell made in this manner was not rechargeable (80% capacity loss after
one cycle). The dissolved iron sulfate is highly acidic (pH < 2) and presents similar hazards to iron chloride. The specific pre-
cipitated product from neutralization of iron chloride in the presence of sulfate is important to the performance of the cell.

6.2. Choice of pH

Initially, the five salts were tested at pH 8. The pH then varied from 7.5 to 8.5. This was done by making the same solution
for the 3 trials but changing the amount of base added during the pH adjustment step in the making of the anode electrolyte

Fig. 3. Schematic and picture of galvanic cell experimental set-up. A table outlining the selection of the iron-salt chemistry. 3 additional graphs for
reference in the choice of pH.
N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072 9

and cathode electrolyte. No pH below 7 was tested because the generation of hydrogen becomes significant and is not
desired (due to safety and performance considerations). As Fig. 3 shows, pH 7.5 had a stable cyclic chronopotentiogram with
roughly the same potential being reached every time on both the discharge and the charge cycles. The cell was unstable at
pH 8 but still functional. However, even though pH 8.5 seems to be stable with similar voltages on the charge and discharge
cycles, it was found that the leads had corroded. Additionally, there is another advantage to pH 7.5. The suspensions were
even more viscous at pH 8 and pH 8.5 than pH 7.5, making them harder to prepare. We also discovered that at 2 M iron con-
centration, the anode electrolyte and cathode electrolyte viscosity increase. At higher pH, it is likely that a more concentrated
solution, such as 3 M, would be too difficult, or impossible, to make.

6.3. Optimization of membrane

We tested many different membranes: 1. Nafion in paper; 2. a pure Nafion membrane; 3. nano-porous polycarbonate; 4.
agarose in paper; 5. agarose on a plastic mesh; 6. polyacrylic acid-co-polyacrylamide in paper; 7. sodium polyacrylate in
paper; 8. paper (only). The paper used was Whatman cellulose chromatography paper, as found in Bill of Materials. All mem-
brane tests can be compared to a paper only control, which highlights the effect of selective ion materials, such as Nafion or
polyacrylate. Selective ion materials are polymers decorated with immobilized anions. They resist conducting anions and are
more conductive to cations. The pure Nafion and polyacrylate-treated paper show significantly higher initial cell potentials
and slower self-discharge than paper alone. We analyzed the cost/performance ratio for each membrane tested. Many were
unsuited either because their performance was unacceptable (e.g., agar gel on paper) while others were too costly (e.g., pure
Nafion membrane). The best balance of low cost and acceptable performance was the sodium polyacrylate in paper. The
sodium polyacrylate membrane material was relatively easy to create and form into a pouch. It showed a self-discharge
in a concentration cell with a half-voltage of 6 h, with a maximum voltage of 110 mV, as seen in Fig. 4D.
To measure the self-discharge performance, we designed and built a concentration cell. A concentration cell produces
a small voltage based on the difference in concentration in otherwise identical compartments. It is split into two cham-
bers, and each will hold one of the electrolytes. The plans for the cell are available online (see Source File Repository). In
this case, both compartments contain iron sulfate but at very different average oxidation states. One half-cell was pre-
pared from Fe2+ and the other from Fe3+. The membrane is held between the compartments. Over time, electron and ion
transfer through the membrane allows the concentration of iron at the different oxidation states to equilibrate. This pro-
cess results in a slow decay of the measured voltage. The results are shown in Fig. 4C. Our concentration cell (Fig. 4A
and B) was designed in our laser cutter’s software. It was also cut using the same laser cutter. The layers, except for the
Viton layer, were adhered together using a clear silicone sealant. Viton should not be cut with a laser cutter due to the
production of toxic and corrosive gasses. Viton can be cut with utility scissors or a sharp knife. Machine screws held
everything in place during testing.

Fig. 4. Membrane self-discharge testing. (A) A schematic shows the construction of the concentration cell. (B) Digital photographs show the constructed
concentration cell. (C) A graph shows the self-discharge plot of measured voltage over time. (D) A summary table of max potential and discharge metric
(time to half-max potential).
10 N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072

6.4. Relevant use case

The final cell design is a result of the optimization strategy above. We present a very simple application to illustrate a
working version of the battery. Six cells were combined in series to make a battery to power a light emitting geode
(LED). The 3 V battery performance was tested using a Vernier data logger to record both current and voltage over time.
The LED was powered for approximately 350 continuous hours. Once the potential was too low for the LED, a simple resistive
load was applied to measure the remaining capacity. Fig. 5C shows the discharge of 600 mAh over two weeks.
In a different experiment, the reliability of a single cell was tested by performing a cyclic discharge and charge test. Each
complete cycle, a discharge followed by a charge, was done at 1 mA, and the voltage was recorded, as seen in Fig. 5A. In
Fig. 5B, the overall charge of the cell can be seen. At the end of the 8th cycle, there is a change in the stability of the battery,
but it is still functional. This could be due to the age of the cell, and its direct exposure to air. This iteration was not sealed.

6.5. Capabilities and limitations

Our iron battery has sufficient capabilities for practical use in low power devices and projects. The cell’s internal resis-
tance is high, and so the discharge rate is limited. There are several possible sources of this high resistance: poor ionic con-
ductivity of the membrane, poor charge transfer rates from the carbon electrode to the Fe3+ in the cathode due to the low
electrical conductivity of the electrolyte, or poor electrical contact among the components. In future work, this could be
addressed by comparing the effects of better membranes, conductive additives to the electrolyte suspensions (e.g., carbon
black), or adding compression to help bring the components into better contact. The pouch cell as described, can provide
1 mA of current. At the typical operating voltage of 0.5 v, each cell can provide 0.5 mW. The cell is robust to at least
ten cycles of charge and discharge without noticeable loss in capacity. An array of 6 cells is thus sufficient for low current
electronics for sensing applications (such as low power microcontrollers like Texas Instruments MSP430). More cells can
always be added in series or parallel to increase the available power and energy storage capacity.
The cell will degrade if water is lost to evaporation or electrolysis. At pH 7.5, the hydrogen reduction overpotential is
approximately 0.4 V. The activation overpotential for the iron surface is 0.15 V. If the potential applied to the iron elec-
trode during charging is more strongly reducing than –0.55 v, hydrogen gas may be generated at a significant rate. Because
the total cell potential is low, charging can be accomplished with a modest applied voltage such that the potential on the iron
electrode does not meet this threshold. Water losses due to evaporation will degrade the battery, as will the slow reaction of
the internal iron with atmospheric oxygen. If the cell is sealed with a thermoplastic, or hot glue, and placed in a secondary
container, then these issues will be minimized. Oxygen and humidity can diffuse slowly through the bag. Minimizing the

Fig. 5. The iron cell charges and discharges and a battery of cells can be connected in series. (A) Potential over time shows stability to repeated charge-
discharge cycles. (B) Charge (in coulombs) over during the same repeated cycle test. (C) Photograph shows a green LED powered by a battery of six cells in
series. (D) The potential over time graph during discharge through the LED. (E) Current over time graph during discharge through the LED. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
N. Yensen, P.B. Allen / HardwareX 6 (2019) e00072 11

airspace around the bag will limit degradation. Pure water can also be added to make up for evaporative losses through the
plastic.

Declaration of Competing Interest

None.

Acknowledgments

We gratefully acknowledge our generous crowdfunding supporters, Jim Kasper, Coeur Power, Thomas Golson, Prahasith
Veluvolu, Dwight Irving, John Palin, Magnus Svantesson, Greg Thomson, Rod Farlee, and Michael Clark. We also acknowledge
the generous support of Avista Utilities. The content is solely the responsibility of the authors and does not necessarily rep-
resent the views of our supporters or Avista Utilities.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ohx.2019.e00072.

References

[1] M. Winter, R.J. Brodd, Chem. Rev. 104 (2004) 4245.


[2] ‘‘Battery Systems - Behind the Meter Storage - Iron Flow Battery - IFB|ESS,” can be found under https://www.essinc.com/energy-storage-products/,
(n.d.).
[3] K.L. Hawthorne, J.S. Wainright, R.F. Savinell, J. Electrochem. Soc. 161 (2014) A1662.
[4] K. Gong, F. Xu, J.B. Grunewald, X. Ma, Y. Zhao, S. Gu, Y. Yan, ACS Energy Lett. 1 (2016) 89.
[5] T.J. Petek, N.C. Hoyt, R.F. Savinell, J.S. Wainright, J. Power Sources 294 (2015) 620.
[6] M.C. Tucker, A. Phillips, A.Z. Weber, ChemSusChem 8 (2015) 3996.
[7] M. C. Tucker, D. Lambelet, A. Phillips, M. Oueslati, B. Williams, W.-C. Wang, A. Z. Weber, ECS Meet. Abstr. 2016, MA2016-02, 37.
[8] N. Muralidharan, A.S. Westover, H. Sun, N. Galioto, R.E. Carter, A.P. Cohn, L. Oakes, C.L. Pint, ACS Energy Lett. 1 (2016) 1034.
[9] HiLyte-Educ. 2018.
[10] D.R. Lide, CRC Handbook of Chemistry and Physics, 89th ed., Taylor & Francis, 2008.
[11] S. Licht, B. Wang, S. Gosh, J. Li, V. Naschitz, Electrochem. Commun. 1 (1999) 522.
[12] A.R. Kim, C.J. Park, M. Vinothkannan, D.J. Yoo, Compos. Part B Eng. 155 (2018) 272.
[13] X. Li, C. Zhao, H. Lu, Z. Wang, H. Na, Polymer 46 (2005) 5820.
[14] ‘‘The Zoetrope - Applied Sciences,” can be found under http://www.applied-sciences.net/library/zoetrope/, (n.d.).
[15] M. Jäger, ‘‘Free your energy,” can be found under http://libre.solar/, 2016.
[16] ‘‘Libre Solar,” can be found under https://github.com/LibreSolar, (n.d.).

You might also like