2019 462 Moesm1 Esm

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

Articles

SUPPLEMENTARY INFORMATION
https://doi.org/10.1038/s41560-019-0462-7

In the format provided by the authors and unedited.

Decoupled hydrogen and oxygen evolution by


a two-step electrochemical–chemical cycle for
efficient overall water splitting
Hen Dotan1,5, Avigail Landman2,5, Stafford W. Sheehan   1,3, Kirtiman Deo Malviya   1,
Gennady E. Shter4, Daniel A. Grave1, Ziv Arzi   1, Nachshon Yehudai4, Manar Halabi4, Netta Gal1,
Noam Hadari1, Coral Cohen1, Avner Rothschild   1* and Gideon S. Grader   4*

1
Department of Materials Science and Engineering, Technion—Israel Institute of Technology, Haifa, Israel. 2The Nancy and Stephen Grand Technion
Energy Program, Technion—Israel Institute of Technology, Haifa, Israel. 3Catalytic Innovations, Fall River, MA, USA. 4Department of Chemical Engineering,
Technion—Israel Institute of Technology, Haifa, Israel. 5These authors contributed equally: Hen Dotan, Avigail Landman.
*e-mail: avner@mt.technion.ac.il; grader@technion.ac.il

Nature Energy | www.nature.com/natureenergy


Supplementary information

Decoupled hydrogen and oxygen evolution by a two-step


electrochemical – chemical cycle for efficient overall water
splitting
Hen Dotan1,§, Avigail Landman2, §, Stafford W. Sheehan1,3, Kirtiman Deo Malviya1, Gennady E. Shter4,
Daniel A. Grave1, Ziv Arzi1, Nachshon Yehudai4, Manar Halabi4, Netta Gal1, Noam Hadari1, Coral
Cohen1, Avner Rothschild1,* and Gideon S. Grader4,*

1
Department of Materials Science and Engineering, Technion—Israel Institute of Technology,
Technion City, Haifa 32000, Israel.
2
The Nancy & Stephen Grand Technion Energy Program (GTEP), Technion—Israel Institute of
Technology, Technion City, Haifa 32000, Israel.
3
Catalytic Innovations, 151 Martine Street, Fall River, Massachusetts 02723, United States.
4
Department of Chemical Engineering, Technion—Israel Institute of Technology, Technion City, Haifa
32000, Israel.
§
Equal contribution
*
To whom correspondence should be addressed: avner@mt.technion.ac.il ; grader@technion.ac.il

1
Supplementary Note 1
Thermodynamic and kinetic considerations
An energy diagram of the reactions in the E-TAC water splitting process and how they compare to
those of purely electrolytic water splitting, at ambient conditions, is presented in Supplementary
Figure 1. In electrolytic water splitting (Supplementary Figure 1A), two electrochemical reactions take
place concurrently: the hydrogen evolution reaction (HER) at the cathode and the oxygen evolution
reaction (OER) at the anode. The HER reversible potential is defined as 0 V versus the reversible
hydrogen electrode (VRHE), and this reaction has a small overpotential of typically1 ~100 mV at a current
density of 10 mA/cm2. The OER reversible potential is 1.23 VRHE, and unlike the HER, this reaction has
a high overpotential, usually1 ranging between 300 and 500 mV at a current density of 10 mA/cm2.
Therefore, the overall cell voltage ranges between 1.6 and 1.8 V at a current density of 10 mA/cm 2.
On the other hand, the E-TAC process is carried out in two consecutive steps (Supplementary Figure
1B). In the first step, HER takes place at the cathode while the Ni(OH)2 anode is electrochemically
charged to NiOOH:

Ni(OH)2 + OH–  NiOOH + H2O + e–

Charging the β-Ni(OH)2 phase to the β-NiOOH phase is generally considered to be a one-electron
reaction wherein Ni oxidizes from a valence of +2 in β-Ni(OH)2 to a valence of +3 in β-NiOOH as protons
are extracted from the lattice 2–4.The overpotential for this one-electron reaction is smaller than that
of the four-electron OER 5–8. This reduces the anodic reaction potential to a level competitive with the
best reported water oxidation catalysts 9–11, enabling water splitting at around the thermoneutral
voltage (1.48 V in standard conditions 12), as demonstrated in this work.

Once the anode is charged (oxidized) to β-NiOOH, it must be discharged (reduced) back to β-Ni(OH)2,
in order to continue reducing water for H2 generation. This can be achieved by facilitating spontaneous
anode reduction (regeneration), which oxidizes water and liberates oxygen 13–18 in a controllable
manner in the second step, according to the following reaction:

4NiOOH + 2H2O  4Ni(OH)2 + O2

This reaction is known as a parasitic process in nickel hydroxide-based batteries 14, as oxygen
generation and anode regeneration are not desired in this case. Here, our E-TAC process uses this
reaction to its advantage to close the cycle in a manner that does not damage the anode, allowing for
repeated cycling. The NiOOH layer is actively involved in the regeneration process, provided that it is
electrochemically accessible to the surrounding electrolyte. In this process, electrons transfer from
adsorbed hydroxide ions to the active NiOOH lattice, reducing it to a lower oxidation state. Oxygen
evolution then proceeds by the combination of two adsorbed OH molecules 14. The surface kinetics of
this process are similar to those of interfacial double-layer discharging at open circuit.

The regeneration reaction can be written as the product of two half-cell reactions:

4NiOOH + 4H2O + 4e– → 4Ni(OH)2 + 4OH– [E⊖ = +1.42 VRHE]

4OH– → O2 + 2H2O + 4e– [E⊖ = +1.23 VRHE]

with 1.42 VRHE being the standard reversible potential (𝐸 ⊖ ) of the Ni(OH)2/NiOOH redox reaction 19–
21
. The anode redox potential must be lower than the potential at which oxygen evolves on this anode,
to avoid oxygen generation during the hydrogen production step. Likewise, it should be higher than
the reversible OER potential in order to achieve spontaneous water oxidation in the second step. The

2
reaction’s driving force is the difference in electromotive force (emf) of the respective half-cell
reactions: emf = 1.42 – 1.23 = 0.19 V.

It is noted that the actual reversible potential (E0) depends on the anode’s composition and state-of-
charge (SOC), as well as the electrolyte concentration, and may be lower or higher than 1.42 VRHE 19,21,
which causes the emf to vary. Supplementary Figure 25A presents the reversible potential of the β-
Ni(OH)2/β-NiOOH couple in 1M KOH solution as a function of Ni(OH)2 mole fraction
(x=Ni(OH)2/[Ni(OH)2+NiOOH]), according to data reported in 21 (orange data series). According to this
data, the reversible potential ranges between 0.416 and 0.470 VHg/HgO (with a delta of 54 mV) for SOC
ranging between 10% and 90%. For our Ni(OH)2 anode (without cobalt) we measured a potential range
of 50 mV for SOC ranging from 10% to 90% (Supplementary Figure 25A, green data series).
Accordingly, the driving force for the anode regeneration reaction (emf) ranges between 0.16 V for a
fully discharged anode and 0.21 V for a fully charged anode. The rate at which the regeneration
reaction takes place depends on parameters such as temperature, time, the SOC of the anode and the
electrolyte's alkali ion concentration. For example, under standard conditions at 25°C, this reaction is
limited by slow kinetics. However, we observe practical regeneration rates at mild temperatures
above 60°C, as discussed in the following (Supplementary Note 2).

Supplementary note 2
The effect of temperature, time and SOC on the anode regeneration reaction
Fast and controllable anode regeneration is important for practical E-TAC water splitting process. The
regeneration reaction is a spontaneous, but kinetically hindered, reaction. Elevated temperature and
sufficient time are required to achieve effective anode regeneration. Additionally, as explained in
Supplementary Note 1, the driving force for the reaction varies with the anode's SOC. Accordingly, the
reaction rate can be enhanced by mitigating mass transfer limitations. Supplementary Figure 6 shows
the regenerated charge (Qregen) as a function of the regeneration time (t) for different temperatures
in the range of 60 to 90°C. For all the temperatures, the regenerated charge increases linearly with
log(t), indicating a fast-initial reaction rate that slows down with time. This behaviour fits the
mechanism proposed by Conway and Bourgault 14,22, describing the self-discharge as a surface
reaction.

The average oxygen evolution reaction rate can be estimated from the number of oxygen moles
evolved during the regeneration step, divided by the regeneration time. According to the regeneration
reaction NiOOH + ½H2O → Ni(OH)2 + ¼O2, the number of oxygen moles evolved during the
regeneration step, 𝑛𝑂2 , equals 0.25𝑛𝑁𝑖(𝑂𝐻)2 , where 𝑛𝑁𝑖(𝑂𝐻)2 is the number of Ni(OH)2 moles
regenerated during this step. Since the Ni(OH)2 moles that are regenerated during the regeneration
step equals the number of Ni(OH)2 moles that are charged during the following charging step in a
stable operation, 𝑛𝑁𝑖(𝑂𝐻)2 can be calculated from the regenerated charge density, Qregen. According to
Faraday's law for the charging reaction, Ni(OH)2 + OH– → NiOOH + H2O + e–,
𝑄𝑟𝑒𝑔𝑒𝑛 ×𝐴 1
𝑛𝑁𝑖(𝑂𝐻)2 = ( )×( ) (Supplementary Equation 1)
𝐹 𝑧

where A is the anode area immersed in the electrolyte solution, F = 96485 C mol-1 is the Faraday
constant and z = 1 is the number of electrons transferred per ion. Therefore,

3
mAh
𝑄𝑟𝑒𝑔𝑒𝑛 ( )×1(cm2 )
cm2
𝑛𝑂2 (𝑚𝑜𝑙) = 0.25𝑛𝑁𝑖(𝑂𝐻)2 = 0.25 ( mAh ) (Supplementary Equation 2)
26801 ( mol )

The average oxygen production rate can therefore be calculated by dividing the total number of
oxygen moles evolved throughout the regeneration step, 𝑛𝑂2 by the total regeneration time, tregen,
mol 𝑛𝑂2 (𝑚𝑜𝑙)
< 𝑟𝑂2 > ( )= (Supplementary Equation 3)
𝑚𝑖𝑛 𝑡𝑟𝑒𝑔𝑒𝑛 (𝑚𝑖𝑛)

Supplementary Figure 7 shows the average oxygen production rate as a function of the total
regeneration time for different regeneration temperatures. The average reaction rate decreases with
regeneration time, i.e., most of the oxygen evolves at the first moments of the regeneration process.
This observation suggests that the reaction initiates at the surface and proceeds into the bulk of the
anode in a diffusion-limited manner, providing additional evidence for the mechanism proposed in 14.
Supplementary Figure 8 shows the same average reaction rate as a function of temperature for
different regeneration times. The average reaction rate increases with temperature for each
regeneration time, and the temperature has the most significant effect on the initial reaction rate, i.e.,
the average rate at short regeneration times. This is, again, in accordance with the previously
proposed reaction mechanism 14. Furthermore, extrapolating the exponential regression curves for all
the plots in Supplementary Figure 8 shows that the average reaction rate becomes negligible at
ambient temperature.

The average incremental reaction rate can be calculated at each time interval according to,
mol ∆𝑛𝑂2 [𝑛𝑂2 (𝑡𝑖𝑟𝑒𝑔𝑒𝑛 )−𝑛𝑂2 (𝑡𝑖−1
𝑟𝑒𝑔𝑒𝑛 )]
< Δ𝑟𝑂𝑖𝑛𝑐𝑟𝑒𝑚𝑒𝑛𝑡𝑟𝑎𝑙, >( )= = (Supplementary equation 4)
2 𝑚𝑖𝑛 ∆𝑡 [𝑡𝑖𝑟𝑒𝑔𝑒𝑛 −𝑡𝑖−1
𝑟𝑒𝑔𝑒𝑛 ](𝑚𝑖𝑛)

𝑖
Where 𝑡𝑟𝑒𝑔𝑒𝑛 is the total regeneration time of the ith increment (i = 1, 2, … ,n for a total of n increments).
Supplementary Figure 9 shows the average incremental reaction rate as a function of the incremental
regeneration time for time increments between [0 – 4 min], [4 – 16 min] and [16 – 31 min]. In the first
increment of [0 – 4 min], the rate was strongly dependent on the temperature, as expected for
thermally-limited reaction kinetics. In the second increment of [4 – 16 min], the regeneration rate
decreases and becomes less temperature-dependent. Finally, in the last increment of [16 – 32 min]
the regeneration rate reaches saturation at approx. 0.5 mol O2/min. The observed trend is consistent
with diffusion-limited kinetics.

Supplementary Figure 10 shows the average incremental regeneration reaction rate as a function of
the anode’s state of charge. This experiment also shows that the first four minutes of regeneration
are the fastest, similar to the results presented in Supplementary Figure 9. Additionally, it shows that
the regeneration rate depends on the anode’s SOC, in agreement with the relationship between the
reaction driving force and the SOC, as discussed in Supplementary Note 1. However, at the last time
increment of [16 – 32 min] the SOC dependence becomes negligible, and the regeneration rate
reaches a saturation value of approx. 0.2 mol O2/min. This further supports the claim that the anode
regeneration (oxygen evolution) process is reaction-limited at first, and then becomes diffusion-
limited.

We can therefore tailor the anodes and the surrounding system to reduce self-discharge and oxygen
evolution during the hydrogen production step, and promote these reactions at a separate step, to
facilitate H2/O2 separation. Here, we achieve this by using customized anodes and operating the
system in a stepwise cycle: first, hydrogen is produced electrochemically at ambient temperature
while the anode is charged from nickel hydroxide to nickel oxyhydroxide. During this step, the anode's

4
SOC is kept well below 100% to suppress oxygen evolution. Then, the anode is disconnected from the
power source, and placed in a hot (95°C) electrolyte to promote oxygen evolution by self-discharge in
a spontaneous chemical reaction, and the anode is regenerated to its initial hydroxide state. This cycle
is illustrated in Figure 1B.

Supplementary Note 3
The effect of surface area on the anode performance
The first proof-of-concept demonstration of the E-TAC water splitting process was achieved with
commercial Ni-MH battery anodes. These anodes are typically prepared by loading an active material
composed of nickel hydroxide particles (with/without additives) with a binding agent into a porous
conductive substrate such as nickel foam 23. Here, we used commercial pasted nickel hydroxide
battery anodes comprised of nickel hydroxide particles coated with cobalt, pressed into a nickel mesh
substrate. Low- and high-magnification SEM images of one of these anodes are presented in
Supplementary Figure 11. Additional material characterizations of these anodes, including SEM, EDS
and XRD, are presented elsewhere 24. While these anodes have relatively high capacities (high energy
density) due to the high loading of active material, their charge and discharge rates are limited by their
small surface area, resulting in low power density and high overpotential. To overcome this limitation,
new anodes were fabricated by electrochemical impregnation (ECI) on nickel foam substrates, as
described in the Methods section. These anodes were tailored to operate at high current densities, as
high as 200 mA/cm2 (see Supplementary Figure 32). This was achieved by using nickel foam substrates
of high surface area, as well as tailoring the anode composition and deposition conditions. Owing to
the increased electrochemical accessibility of the active material 3, the ECI anodes achieve much
higher current densities compared to their commercial battery anode counterparts. However, their
capacity is lower due to the relatively low active material loading as compared to the battery anodes.
Thus, the anodes can be tailored for either high rate (high power density) or high capacity (high energy
density) performance, and one must carefully balance between these properties to meet the target
performance criteria. Here we show a comparison between our anodes and the battery anodes used
in the early proof-of-concept demonstration.

Supplementary Figure 12 compares our ECI anodes (marked as ECI) versus commercial battery anodes
(marked as CB), plotting the maximum charge density that was transferred during charging (Qcharge)
and discharging (Qdischarge) as a function of the current density (current divided by the anode’s area
immersed in the electrolyte solution). It should be noted that this was a preliminary test, carried out
with an unoptimized ECI process that yielded anodes with significantly lower capacity as compared to
the optimized anodes that followed. When charging at a low current density of 20 mA/cm2, the CB
anode displayed a slightly better capacity compared to the ECI anode. However, at higher current
densities of 50 and 100 mA/cm2, the potential of the CB anode rose above the cutoff potential after
only 0.2 s, yielding a capacity of less than 10 mAs/cm2. The ECI anode, on the other hand, yielded
discharge capacities of 1,547 mAs/cm2 and 2,271 mAs/cm2 for charging current densities of 50 and
100 mA/cm2, respectively. These results show that the ECI anode is superior to the CB anode in terms
of its charging capability at high current densities. Namely, the ECI anode demonstrates much higher
power density compared to the CB anode. The anode regeneration performance was also compared
between the ECI and CB anodes. Supplementary Figure 13 presents the regenerated charge for
different regeneration times and initial SOCs measured for the two anodes (CB and ECI). The ECI anode

5
displays faster regeneration rates at lower SOC compared to the CB anode, allowing for better
utilization of the anode's capacity.

Supplementary Note 4
The oxygen evolution reaction on Ni0.9Co0.1(OH)2 and NiFe LDH anodes
To avoid H2/O2 crossover in the E-TAC water splitting process, it is essential that the anode charging
occurs without oxygen generation. Therefore, the anode's charging potential must stay below the OER
potential at the same anode under the same conditions. Furthermore, to compete with benchmark
water oxidation catalysts, the charging potential of our anodes should be lower than the OER potential
of those catalysts. Here, we present chronoamperometric polarization measurements that were
carried out to determine the anode charging potential and OER potential of our Ni0.9Co0.1(OH)2 anodes,
and how they compare to the OER potential of a benchmark water oxidation catalyst based on NiFe
LDH (see Methods section, and Supplementary Figures 27A-B). Supplementary Figure 27C presents
the amperogram measured during constant potential polarization of a fully discharged Ni0.9Co0.1(OH)2
anode at 1.48 VRHE for 2300 s. This test was carried out to evaluate the Faradaic efficiency of the anode
charging reaction at this potential. The steady-state current that was measured at the end of the test
is attributed to the OER. Thus, the OER rate at the Ni0.9Co0.1(OH)2 anode subjected to a potential of
1.48 VRHE was 0.07 mA/cm2. The oxygen generation rate due to spontaneous anode regeneration is
negligible at this temperature, as explained in Supplementary Note 3. Accordingly, at each point in
time during the experiment, t, the total charge, the capacitive charging charge and the OER charge
can therefore be calculated as follows.

The total charge that was passed through the anode, Qtotal, can be calculated by integrating the current
over time,
𝑡
𝑄𝑡𝑜𝑡𝑎𝑙 (𝑡) = ∫0 𝐼(𝑡)𝑑𝑡 (Supplementary equation 5)

The maximal OER charge, QOER, that passed through the anode up to time t, can be calculated by
multiplying the steady-state OER current, i.e., the current obtained at the end of the measurement by
the time,

𝑄𝑂𝐸𝑅 (𝑡) = 𝐼𝑂𝐸𝑅 × 𝑡 (Supplementary equation 6)

QOER is represented by the green dashed area of the inset in Supplementary Figure 27C. This calculation
assumes that the OER occurs at its steady-state rate throughout the entire charging process, whereas
the anode could be charged to a certain degree without oxygen evolution, as discussed below.

The capacitive charge, QC, can be calculated by subtracting the OER charge from the total charge,

𝑄𝐶 (𝑡) = 𝑄𝑡𝑜𝑡𝑎𝑙 (𝑡) − 𝑄𝑂𝐸𝑅 (𝑡) (Supplementary equation 7)

QC is represented by the green dashed area of the inset in Supplementary Figure 27C. The Faradaic
efficiency of the anode charging process, up to time t, can then be calculated by dividing the capacitive
charge by the total charge,
𝑄𝐶 (𝑡)
𝜂𝐹 (𝑡) = (Supplementary equation 8)
𝑄𝑡𝑜𝑡𝑎𝑙 (𝑡)

6
In this test, carried out at 1.48 VRHE, the Faradaic efficiency was greater than 99.7% for the first 2000
s, and greater than 99.6% for the rest of the test. The average anode potential in our proof-of-concept
experiment presented in Figure 3 was 1.42 VRHE at a current density of 50 mA/cm2 for tests of 100 s.
Therefore, the Faradic efficiency for anode charging in our proof-of-concept experiment was greater
than 99.6%. Since the only reaction at a Pt electrode polarized cathodically in 5M KOH alkaline solution
is the HER that produces only H2 gas, the H2 purity can be calculated according to Faraday's law of
electrolysis (Supplementary equation 1),
𝐶
𝑛𝐻2 𝑄𝐻𝐸𝑅 (𝐶)⁄[96485( )×2]
𝑚𝑜𝑙
[𝐻2 , 𝑚𝑜𝑙%] = = 𝐶 𝐶 =
𝑛𝐻2 +𝑛𝑂2 {𝑄𝐻𝐸𝑅 (𝐶)⁄[96485( )×2]}+{𝑄𝑂𝐸𝑅 (𝐶)⁄[96485( )×4]}
𝑚𝑜𝑙 𝑚𝑜𝑙
0.05(𝐴)×100(𝑠)/2
[0.05(𝐴)×100(𝑠)/2]+[0.05(𝐴)×100(𝑠)×(1−0.996)/4]
= 99.8 𝑚𝑜𝑙% (Supplementary equation 9)

Yielding a H2 purity of 99.8 mol%, corresponding to 99.9% vol.%.

To validate this analysis, dissolved oxygen (DO) and gas chromatography (GC) measurements were
carried out during charging the anode as in the hydrogen production step of our E-TAC process.
Supplementary Figure 28 present the DO volume as a function of cumulated charge (Q = I × t) during
chronoamperometric charging of a fully discharged Ni0.9Co0.1(OH)2 anode at 1.48 VRHE. The DO
concentration remained constant and did not rise above the background level for the first 550 C (55
C/cm2). Supplementary Figure 29 presents the DO volume as a function of the cumulated charge (Q =
I × t) during both the E-TAC and water electrolysis control tests. The DO volume during the E-TAC
hydrogen generation step reached 3.8 ± 3.2 µl at the end of the test, corresponding to an oxygen
evolution charge of
𝑔 𝐶
[(3.8±3.2)(𝜇𝑙)]×1.3085×10−6 ( )×96485( )×4
𝜇𝑙 𝑚𝑜𝑙
𝑄𝑂𝐸𝑅 (𝐶) = 𝑔 = (0.06 ± 0.05)(𝐶) (Supplementary equation 10)
32( )
𝑚𝑜𝑙

The Faradaic efficiency of the anode charging reaction is therefore:


𝑄𝑡𝑜𝑡 −𝑄𝑂𝐸𝑅 3(𝐶)−[0.06(𝐶)±0.05(𝐶)]
𝜂𝐹,𝑎𝑛𝑜𝑑𝑒 𝑐ℎ𝑎𝑟𝑔𝑖𝑛𝑔 = = = (98.0 ± 1.7)% (Supplementary equation 11)
𝑄𝑡𝑜𝑡 3(𝐶)

Assuming a 100% Faradaic efficiency for the hydrogen evolution reaction at the cathode 24, the total
hydrogen volume produced as calculated by the charge is
𝑔
3(𝐶)×2( )
𝑚𝑜𝑙
𝑉𝐻2 (𝜇𝑙) = 𝐶 𝑔 = 377.8(𝜇𝑙) (Supplementary equation 12)
96485( )×2×8.23×10−8 ( )
𝑚𝑜𝑙 𝜇𝑙

corresponding to a hydrogen purity of


𝑉𝐻2 3.8(𝑚𝑙)
𝐻2 𝑝𝑢𝑟𝑖𝑡𝑦 = = = (99.9 ± 0.08) 𝑣𝑜𝑙. % (Supplementary equation 13)
𝑉𝐻2 +𝑉𝑂2 3.8(𝑚𝑙)+[(3.8±3.2)(𝜇𝑙)]

Supplementary Figure 30 presents the DO volume obtained during chronoamperometric charging at


a constant potential of 1.48 VRHE carried out after anode regeneration in 5M KOH at 95°C for 120 s.
The DO concentration remained constant and did not rise above the background level for the first 40
C (4 C/cm2). During this time, the current decreased from 25 to 8 mA/cm2. Subsequently, the current
approached a steady-state value of ~6 mA/cm2, and the DO concentration gradually increased,
indicating that OER was taking place at the charged Ni0.9Co0.1(OH)2 anode.

Supplementary Figure 31 presents the gas chromatograms for the purged cell and after the E-TAC
hydrogen generation step, showing no rise in the oxygen peak following the hydrogen generation step.
This result verifies that under these conditions only H2 gas is produced in the cell during the E-TAC
charging steps.

7
Supplementary Note 5
Heat balance for alkaline and E-TAC water splitting
In alkaline water electrolysis, the OER overpotential leads to significant heat release and electric
power loss. Because our anode regeneration step (where O2 is released) is done chemically rather
than electrochemically, the E-TAC water splitting process reduces the polarization heat loss and
improves the power efficiency. In supplementary Figure 1A, depicting alkaline water electrolysis,
arrows 1 and 2 represent the heat that is released due to the OER and HER overpotentials, 𝑞1 =
𝑞(𝜂𝑂𝐸𝑅 ) and 𝑞2 = 𝑞(𝜂𝐻𝐸𝑅 ), respectively. Here, we use 𝜂𝑂𝐸𝑅 = 0.4𝑉 and 𝜂𝐻𝐸𝑅 = 0.1𝑉 as typical
overpotential values1 of the respective reactions at a current density of 10 mA/cm2. Arrow 3
represents the heat that is absorbed by the endothermic cell reaction, 𝑞3 = (∆𝐻⊖ 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑙𝑦𝑠𝑖𝑠 −
∆𝐺 ⊖ 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑙𝑦𝑠𝑖𝑠 ) = 1.48 𝑉 − 1.23 𝑉 = 0.25 𝑉. The overall heat requirement in this case is therefore, 𝑞 =
𝑞1 + 𝑞2 − 𝑞3 = (0.4𝑉 + 0.1𝑉 − 0.25𝑉) × 𝐹 × 𝑛 ≅ 48.2 kJ⁄𝑚𝑜𝑙 𝐻2 where F is Faraday's constant and n is
the number of electrons per mol H2.

In Supplementary Figure 1B, depicting the E-TAC water splitting process, arrow 2 is the same as in
alkaline water electrolysis, 𝑞2 = 𝑞(𝜂𝐻𝐸𝑅 ). Arrow 4 represents the heat that is released due to the anode
charging overpotential, 𝑞4 = 𝑞(𝜂𝑜𝑥 ) = 𝜂𝑜𝑥 × 𝐹 × 𝑛 = 0.1𝑉 × 𝐹 × 𝑛 ≅ 19.3 kJ⁄𝑚𝑜𝑙 𝐻2 , where 𝜂𝑜𝑥 is the
overpotential for the anode charging reaction. Here, the value of 𝜂𝑜𝑥 = 0.1𝑉 was extracted from cyclic-
voltammograms for Ni(OH)2 anodes (without additives) reported elsewhere 5–8. The actual value
depends on the anode's microstructure and chemical composition, and it may deviate significantly
from 0.1 V, as discussed in Supplementary Note 1. Arrow 5 represents the heat absorbed by the
endothermic charging reaction, 𝑞5 = (∆𝐻 ⊖ 𝑠𝑡𝑒𝑝 1 − ∆𝐺 ⊖ 𝑠𝑡𝑒𝑝 1 ) = (1.56𝑉 − 1.42𝑉) × 𝐹 × 𝑛 ≅
27 kJ⁄𝑚𝑜𝑙 𝐻2 , (see Supplementary Table 2). Lastly, arrow 6 represents the heat that is released by the
exothermic anode regeneration reaction, 𝑞6 = (Δ𝐻⊖ step 1 − ∆𝐻⊖ 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑙𝑦𝑠𝑖𝑠 ) = (1.56𝑉 − 1.48𝑉) × 𝐹 ×
𝑛 ≅ 15.3 kJ⁄𝑚𝑜𝑙 𝐻2 . In this case (see Supplementary Figure 1B), the overall heat loss is:
𝑘𝐽 𝑘𝐽 𝑘𝐽
𝑞 = 𝑞2 + 𝑞4 − 𝑞5 + 𝑞6 = (0.1 𝑉 + 0.1 𝑉) × 𝐹 × 𝑛 − 26.4 + 15.3 ≅ 27.5
𝑚𝑜𝑙,𝐻2 𝑚𝑜𝑙,𝐻2 𝑚𝑜𝑙,𝐻2
(Supplementary Equation 14)

In our proof-of-concept experiment, presented in Figure 3, we perform the first (hydrogen generation)
step at a current density of 50 mA/cm2 with an average cell voltage of 1.5 V. Therefore, we calculate
our overall heat loss as:
𝑘𝐽
𝑞 = (1.5 V − 1.48 𝑉) × 𝐹 × 𝑛 ≅ 3.9 (Supplementary Equation 15)
𝑚𝑜𝑙,𝐻2

Thus, our E-TAC water splitting process leads to a greater than tenfold reduction in heat loss
(4 kJ⁄𝑚𝑜𝑙 𝐻2 vs. 48 kJ⁄𝑚𝑜𝑙 𝐻2 ). It is noted that applying an average cell voltage of 1.48 V during the
hydrogen generation step would lead to a unique condition where the endothermic heat requirement
of the hydrogen generation step is equal to the exothermic heat that is released during the anode
regeneration step, yielding a net q = 0, as explained in Supplementary Note 6.

Supplementary Note 6
Energy consumption
In assessing the overall energy efficiency of the E-TAC process, one must consider both the electrical
power and heat consumption involved. Thermodynamically, the total energy requirement for the
hydrogen generation step is given by ΔH = ΔG + TΔS, where ΔH is the enthalpy of the reaction (301.1

8
kJ/mol H2, 1.56 eV per electron), ΔG is the Gibbs free energy that must be supplied by the power
source (274.7 kJ/mol H2, 1.42 eV per electron), and TΔS = Δqin is the heat requirement at a constant
temperature T. The TΔS component may also be supplied by the power source, in which case the cell
operation is adiabatic. Therefore, adiabatic isothermal conditions for the hydrogen generation step
are achieved at Vcell = 1.56 V, the thermoneutral cell voltage (see Supplementary Table 2). At cell
voltages below 1.56 V the process is endothermic, absorbing heat from the environment, and
effectively cooling the system. To keep the cell at ambient temperature, heat must be supplied
according to:
𝑚𝑜𝑙 𝑒 − 𝐶 𝐽 𝑘𝐽
𝑞𝐻2 = (1.56 − 𝑉𝑐𝑒𝑙𝑙 )(𝑉) × 2 ( ) × 96485 ( ) × 1( ) × 10−3 ( ) =
𝑚𝑜𝑙 𝐻2 𝑚𝑜𝑙 𝑒 − 𝐶×𝑉 𝐽
𝑘𝐽
(1.56 − 𝑉𝑐𝑒𝑙𝑙 (𝑉)) × 192.97 ( ) (Supplementary Equation 16)
𝑚𝑜𝑙 𝐻2

If the process is carried out at 1.48 V, the heat that must be supplied to keep the cell at 25°C equals
(1.56 − 1.48) × 192.97 𝑘𝐽⁄𝑚𝑜𝑙 𝐻2 = 15.3 𝑘𝐽⁄𝑚𝑜𝑙 𝐻2 .

On the other hand, the oxygen generation step is a spontaneous (∆𝐺̂ °𝑟𝑥𝑛 < 0) and exothermic reaction,
releasing heat to its environment, equal to the regeneration reaction's enthalpy change, 𝑞𝑂2 = ΔH°rxn=
-15.3 kJ/mol. Neglecting any mechanical work, ∆𝐻 ̂ °𝑟𝑥𝑛 = 𝑞 is the heat released to the environment
during this process at 25°C. Therefore, the power and heat requirements of the E-TAC process depend
on the cell voltage during the hydrogen generation step. The electrical energy consumption at 1.48 V
during the electrochemical (hydrogen generation) step is equal to the total energy demand, since the
heat consumed by hydrogen generation is released during the oxygen generation step. Theoretically,
in a thermally isolated system operating at 1.48 V, the heat released by the regeneration reaction
during oxygen production can be captured and then supplied during the hydrogen production step,
with no additional heat supplied or removed from the system. If the electrochemical step is carried
𝑘𝐽
out below 1.48 V, excess heat in the amount of (1.56 − 𝑉𝑐𝑒𝑙𝑙 (𝑉)) × 192.97 𝑚𝑜𝑙 − 15.3 𝑘𝐽/𝑚𝑜𝑙 must be
supplied to the system. These thermodynamic calculations are presented in Supplementary Figure 37.

In practice, one should account for heat losses arising from heating the anode and the consumed
water in the second step of the E-TAC process. As shown in Figure 3D, the electrical power
consumption required by the E-TAC process is 39.9 kWh per kg of hydrogen, without considering any
heat loss that may arise from cycling between the cold and hot steps. If the system operates at the
thermoneutral water splitting voltage (1.48 V) such that the hydrogen generation step is endothermic
while the oxygen generation and anode regeneration step is exothermic, this is expected to
compensate for some of the heat loss during operation, especially if the containing vessels and pipes
are thermally insulated. However, the heating and cooling of the electrodes as well as water
consumption require careful heat management. The expected heat losses for practical operation are
calculated using literature data 19. First, hot water (95°C) is consumed during the oxygen generation
step of the E-TAC cycle. Without a heat recovery system, heating water from 25°C to 95°C requires
energy investment. This is calculated as follows:
368 𝐾 𝐽 𝑘𝐽 𝑘𝑊ℎ
𝑞 = 𝑛(𝑚𝑜𝑙) ∙ ∫298 𝐾 𝐶𝑝 ( ) 𝑑𝑇 = 4.35 ( ) = 0.6 ( ) (Supplementary Equation 17)
𝑚𝑜𝑙∙𝐾 𝑚𝑜𝑙 𝐻2 𝑘𝑔 𝐻2

𝐽
with the heat capacity of water being 𝐶𝑝,𝑤𝑎𝑡𝑒𝑟 = 52.928 + 47.614 ∙ 10−3 ∙ 𝑇 − 7.238 ∙ 105 ∙ 𝑇 −2 (𝑚𝑜𝑙∙𝐾).
This energy investment is only 1.5% of the electrical energy input, and it can be further reduced by
adding a heat recovery system. Next, the heat required to heat up the active material (nickel
hydroxide) in the anode from room temperature (25°C, used in the hydrogen generation step) to 95°C
(used in the oxygen generation step) is

9
368 𝐾
𝐽
𝑞𝑛𝑖𝑐𝑘𝑒𝑙 ℎ𝑦𝑑𝑟𝑜𝑥𝑖𝑑𝑒 = 𝑁𝑛𝑖𝑐𝑘𝑒𝑙 ℎ𝑦𝑑𝑟𝑜𝑥𝑖𝑑𝑒 (𝑚𝑜𝑙) ∙ ∫ 𝐶𝑝 ( ) 𝑑𝑇
𝑚𝑜𝑙 ∙ 𝐾
298 𝐾

𝐽 𝑘𝐽 𝑘𝑊ℎ
= 2𝑁ℎ𝑦𝑑𝑟𝑜𝑔𝑒𝑛 (𝑚𝑜𝑙) ∙ 70(℃) ∙ 92.33 ( ) = 12.9 ( ) = 1.78 ( ) (Supplementary Equation 18)
𝑚𝑜𝑙∙𝐾 𝑚𝑜𝑙𝐻2 𝑘𝑔 𝐻2

Simulating a fully operational E-TAC water splitting system is outside the scope of this work. We can
only estimate the heat losses caused by the temperature swings in the cell during the E-TAC cycle. This
calculation assumes the following:

- Plug flow in the cell which is reasonable under sufficient electrolyte flow
- Hydrogen generation at 25°C, oxygen generation at 95°C, and leftovers electrolyte at 50°C
- The cell is made of Teflon to reduce heat loss to the environment
- The anode is 3 mm thick and 85 vol% porous, 10 vol% nickel hydroxide and 5 vol% nickel
- The cathode is 0.5 mm thick and 95 vol% porous and 5 vol% nickel
- 23 ml of hydrogen are generated during 10 min hydrogen generation step for each 1 ml
anode which correlates to a current density of 50 mA/cm2 and anode regenerative
capacity of 100 C/ml. Such regenerative capacity means that 47 vol% of the nickel
hydroxide in the anode was functional during the E-TAC cycle
- The oxygen generation step and cell flushing steps with the intermediate solution take 10
min each

Under these assumptions the calculated heat losses are:

- 2.17 kWh/kg H2 consumed by the cell walls


- 1.25 kWh/kg H2 consumed by the anode
- 0.02 kWh/kg H2 consumed by the cathode

Considering the heat required to heat up the water to 95°C (0.6 kWh/kg H2) and the reaction heat
released during oxygen generation (2.15 kWh/kg H2, as calculated in the article), the net heat loss is
estimated to be

Σ𝑞𝐿𝑜𝑠𝑠 = (2.17 + 1.25 + 0.02 + 0.6 − 2.15) 𝑘𝑊ℎ/𝑘𝑔 𝐻2 = 1.9 𝑘𝑊ℎ/𝑘𝑔 𝐻2 (Supplementary
Equation 19)

In summary, this heat loss amounts to about 5% of the ~40 kWh/kg H2 that is required to drive the
hydrogen production step (see Figure 3D). The heat losses can be reduced by installing better wall
insulation, by developing anodes with a smaller heat capacity, and by operating the hydrogen step at
a higher temperature. Such measures will bring the operation closer to the thermo-neutral point.

Supplementary Note 7
Potential cost savings in membraneless electrolysis
Studies comparing membraneless electrolysers with PEM systems emphasize the simple construction
of membraneless systems that alleviates the need for expensive components and complex assembly
25
. While the anode and cathode materials contribute less than 5% of the capital cost of PEM
electrolysers (see Table 1 in 25), other components that could potentially be eliminated in membrane-
less electrolysers contribute much more to the capital cost (bipolar plates – 30.6%, MEA
manufacturing – 6%, membranes – 3%).

10
The materials and manufacturing requirements of these components for PEM electrolysers incur
substantial costs that membraneless systems do not. For example, bipolar plates are a large portion
of the cost of PEM electrolysers, and due to their strict material and manufacturing requirements it is
very challenging to reduce or mitigate their impact on the capital expenditure required for PEM
electrolyser production 26. Accordingly, the components of PEM electrolysers that are not required
for our E-TAC process make up for approximately 40% of the total cost of a standard PEM electrolyzer.

11
Supplementary Figures

Supplementary Figure 1: Energy diagram of the reactions in electrolytic and in E-TAC water splitting. Energy diagram of
the reactions taking place in alkaline water electrolysis (A) and the E-TAC water splitting process with conventional Ni(OH)2
(B) or modified Ni0.9Co0.1(OH)2 anodes (C). The potential values are shown under ambient conditions. The arrows are color-
coded as explained in the legend on the right. The orange and yellow arrows represent the reversible and thermoneutral
voltages that correspond to the maximum amount of free energy (Δ𝐺 ⊖ ) and enthalpy (Δ𝐻 ⊖ ), respectively, available in the
hydrogen combustion reaction. They apply equally to both cases. The light blue arrows represent the cell voltage, Vcell, that
is applied during the electrochemical process in both cases at ambient conditions. Vcell was calculated by (1.23 VRHE – 0 VRHE)
+ 0.4 V + 0.1 V = 1.73 V for the electrolysis case (A), where 𝜂𝑂𝐸𝑅 = 0.4 V and 𝜂𝐻𝐸𝑅 = 0.1 V are typical overpotential values for
the respective reactions, and by (1.42 VRHE – 0 VRHE) + 0.1 V + 0.1 V = 1.62 V for the E-TAC case (B), where 0.1 V is a typical
overpotential value for the Ni(OH)2 oxidation reaction (𝜂𝑜𝑥 ), and 𝜂𝐻𝐸𝑅 = 0.1 V. Since the anode’s SOC changes during the
process, the anode’s potential varies by several tens of mV during operation (Supplementary Figure 25). Thus, the values
displayed in B represent a snapshot around a SOC of 50% that corresponds to the standard reversible potential of a
conventional Ni(OH)2 anode. The red arrows represent heat absorbed (upward arrows) or released (downward arrows)
during the process. In the water electrolysis case (A), arrows 1 and 2 represent the heat that is released due to the OER and
HER overpotentials, respectively, and arrow 3 represents the heat that is absorbed by the endothermic water splitting
reaction. In the E-TAC process (B), arrows 2 and 4 represent the heat that is released due to the HER and anode charging
overpotentials, respectively. Arrow 5 represents the heat that is absorbed by the endothermic anode charging reaction, and
arrow 6 represents the heat that is released by the exothermic anode regeneration reaction. The overall heat loss in water
electrolysis is the sum of arrows 1, 2 and 3, whereas for the E-TAC process it is the sum of arrows 2, 4, 5 and 6. In the E-TAC
process with modified Ni0.9Co0.1(OH)2 anodes (C) the average anode potential and overpotential values are based on our
proof-of-concept experiment (Figure 3). The average anode potential in this experiment is 1.4 VRHE, with an average cell
voltage of 1.5 V. As a result, the overall heat loss is 1.5 V – 1.48 V = 0.02 V = 3.9 kJ/mol, represented by arrow 7. The dark
blue arrows represent the driving force for the anode regeneration reaction, which is the difference between the reversible
potentials of the anode and the OER.

12
Supplementary Figure 2: Ni0.9Co0.1(OH)2 anodes. (A) Photograph of a nickel foam substrate, and (B) photograph of a
Ni0.9Co0.1(OH)2 anode prepared by electrochemical impregnation. The green colour of the deposited material in B was
enhanced for clarity.

Supplementary Figure 3: System for electrochemical impregnation of nickel foam substrates for anode preparation. The
nickel foam substrate was connected as the cathode (working electrode) and placed between two nickel foam anodes. All
electrodes were dipped in an aqueous nitrate solution (nickel nitrate with/without cobalt nitrate), and the deposition was
carried out by applying a constant current of 60 mA.

13
0.59
0.57

Potential [VHg/HgO]
0.55
0.53
0.51 Charging
0.49 Discharging
0.47
0.45
0 5 10 15 20 25 30
Charge [mAh/cm2]

Supplementary Figure 4: Polarization potential as a function of charge for a commercial nickel hydroxide anode. The anode
was charged at a constant current density of 10 mA/cm2 for 0.5 to 2.5 h, and then discharged at 5 mA/cm2 to a potential of
0 VHg/HgO. The polarization potential at the end of each 0.5 h charging step was recorded (blue dots). The anode was then
discharged to various charges from 25 to 5 mAh/cm2, and the polarization potential at the end of each discharge step was
recorded (red dots). There is a linear correlation between the charge and the polarization potential, similar to the results
shown for the electrochemically deposited anode in Supplementary Figure 25. Thus, during constant current measurements,
the anode's SOC can be obtained from its polarization potential.

B
2.5
Average regenerated charge

2
[mAh/cm2]

1.5

0.5

0
Method 1 Method 2
Evaluation method

Supplementary Figure 5: Comparison of the two methods devised to evaluate the anode's regenerated charge, Qregen. (A)
Illustration of the two methods. Method 1 consists of three steps: discharging at a constant current density of 5 mA/cm2 until
a cutoff potential of 1 VRHE, charging at a constant current density of 10 mA/cm2 to a total charge density of 20 mAh/cm2,
and regenerating in 1M NaOH aqueous electrolyte at 80°C for 32 min. Method 2 consists of only two steps of charging and
regeneration, at the same conditions, with a preliminary discharging to the same cutoff potential of 1 VRHE. Each charging
step in Method 2 is carried out to a cutoff potential of 1.52 VRHE, corresponding to a charge density of 20 mAh/cm2. (B) The
regenerated charge obtained by the two methods. The bars represent the average regenerated charge (Qregen) value for three
repeated tests, and the error bars represent the standard deviation. Both methods yield the same values of Qregen. Therefore,
in the following experiments, the regenerated charge density was evaluated according to Method 2.

14
A 1.2
T=60C
B 1.2
T=60C

Regenerated charge density


Regenerated charge density

1 T=70C 1 T=70C
T=80C T=80C
0.8 T=90C 0.8 T=90C
[mAh/cm2]

[mAh/cm2]
0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 10 20 30 40 1 10 100
Regeneration time [min] Regeneration time [min], log scale

Supplementary Figure 6: The regenerated charge density of a commercial nickel hydroxide battery anode as a function of
the regeneration time at different temperatures. (A) Linear time scale; (B) Logarithmic time scale. Each data point
represents the average result for five repeated tests, and the error bars represent the standard deviation.

14
T=60C
Average oxygen production rate

12 T=70C
T=80C
10
[mol O2/min]x106

T=90C
8

0
0 5 10 15 20 25 30 35
Regeneration time [min]

Supplementary Figure 7: The average oxygen production rate of a commercial nickel hydroxide battery anode as a function
of the regeneration time at different regeneration temperatures. Each data point represents the average result for five
repeated tests, and the error bars represent the standard deviation.

15
9

Average oxygen production reaction


8 4 min
7 8 min

rate [mol O2/min]x106


16 min
6
32 min
5
4
3
2
1
0
298 308 318 328 338 348 358 368
Temperature [K]

Supplementary Figure 8: The average oxygen production rate of a commercial nickel hydroxide battery anode as a function
of temperature at different regeneration times. Each data point represents the average result for five repeated tests, and
the error bars represent the standard deviation.

20
generation reaction rate [Δmol

18 T=60C
Average incremental oxygen

16 T=70C
14 T=80C
O2/Δmin]x106

12
10
8
6
4
2
0
0 to 4 min 4 to 16 min 16 to 32 min
Reaction time increment

Supplementary Figure 9: The incremental oxygen evolution rate, ∆𝒏𝑶𝟐 /∆𝒕, of a commercial nickel hydroxide battery
anode, as a function of the incremental time interval, Δt, for different regeneration temperatures. Each data point
represents the average result for five repeated tests, and the error bars represent the standard deviation.

16
2

Average incremental oxygen generation


10 mAh cm -2, 43% SOC

reaction rate [Δmol O2/Δmin]x106


15 mAh cm-2, 65% SOC
20 mAh cm-2, 86% SOC
1.5

0.5

0
0 to 4 min 4 to 16 min 16 to 32 min
Reaction time interval

Supplementary Figure 10: The incremental oxygen evolution rate, ∆𝒏𝑶𝟐 /∆𝒕, of a commercial nickel hydroxide battery
anode, as a function of the incremental time, Δt, for different initial SOCs of the anode. Each data point represents the
average result for five repeated tests, and the error bars represent the standard deviation.

Supplementary Figure 11: SEM micrographs of a commercial battery nickel hydroxide anode. Figures (A) and (B) show low
and high magnifications of a pasted commercial battery anode, presenting a conductive porous substrate filled with the
active material (nickel hydroxide).

17
5000
Charge density [mAs/cm2] ECI - Q,charge ECI - Q,discharge
4000 CB - Q,charge CB - Q,discharge

3000

2000

1000

0
20 50 100
Current density [mA/cm2]

Supplementary Figure 12: Comparison between the energy (charge) and power (current) densities of commercial battery
(CB) and electrochemically impregnated (ECI) nickel hydroxide anodes. The red and blue columns show the charge density
transferred upon polarization of commercial battery (CB) and electrochemically impregnated (ECI) anodes, respectively, at
constant current densities of 20, 50, and 100 mA/cm2 to cutoff potentials of 1.44, 1.46, and 1.52 VRHE, respectively. Each data
point represents the average result for three repeated tests, and the error bars represent the standard deviation.

Supplementary Figure 13: Comparison between the regeneration performance of commercial battery (CB) and
electrochemically impregnated (ECI) nickel hydroxide anodes. The regenerated charge density is presented as a function of
the regeneration duration and the initial SOC at the beginning of the regeneration step for a CB anode (A) and for an ECI
anode (B).

18
Supplementary Figure 14: X-ray diffractograms of a Ni0.9Co0.1(OH)2 anode in the as-deposited state and after aging. The
diffractograms show the transformation from α-Ni(OH)2 in the as-deposited state (black) to β-Ni(OH)2 after aging (green).
The indexed Bragg reflections correspond to α-Ni(OH)2 (PCPDF card# 00-0380715, dashed-dotted blue line), β-Ni(OH)2
(PCPDF card #01-0802855, solid blue line), and nickel (PCPDF card# 04-010-6148, red line). This result shows that the aging
process transforms the anode to the β-hydroxide phase, which is known to be the cyclable phase in rechargeable alkaline
batteries 3,4.

19
Supplementary Figure 15: High-magnification SEM micrographs of a Ni0.9Co0.1(OH)2 anode in the as-deposited state. (A) A
Ni0.9Co0.1(OH)2 layer on the nickel foam substrate, with an average layer thickness of 9±1 μm (B). Figures (C-F) present higher
magnifications of the layer (3.5k to 40k), revealing its nanoflake texture with some cauliflower-like structures at the surface.
The same anode is presented in Supplementary Figure 22, and is imaged here at higher magnification, and after an additional
time delay of ~1 min.

20
Supplementary Figure 16: SEM micrographs of an aged Ni0.9Co0.1(OH)2 anode before charging. The images are shown at
low (A) and high (B) magnifications.

Supplementary Figure 17: SEM micrographs of an aged Ni0.9Co0.1(OH)2 anode taken after sequential charging steps. An
aged Ni0.9Co0.1(OH)2 anode was charged sequentially at a constant current of 40 mA (applied on an anode area of 1 cm 2
immersed in a 5M KOH electrolyte) for 2000, 500, 500, 155, and 1000 seconds (22, 28, 33, 35 and 44 mAh/cm2, respectively).
Figures (A-B) and Figures (C-D) show low and high magnification SEM micrographs taken after 2000 s and after an additional
1155 s (total of 3155 s) charging at a constant current of 40 mA, respectively. The charging process does not lead to visible
changes in the anode’s microstructure as compared to the aged anode before charging (see Supplementary Figure 16).
Increased cracking with time is attributed to the drying shrinkage process, as shown in Supplementary Figures 22-23.

21
Supplementary Figure 18: X-ray diffractograms of a Ni0.9Co0.1(OH)2 anode after sequential charging steps. An aged
Ni0.9Co0.1(OH)2 anode was charged sequentially at a constant current of 40 mA (applied on an anode area of 1 cm2 immersed
in a 5M KOH electrolyte) for 2000, 500, 500, 155, and 1000 seconds (22, 28, 33, 35 and 44 mAh/cm2, respectively). Curves C,
D, E, and F present diffractograms taken after charging at a constant current of 40 mA for 2000 s, 2500 s (total), 3000 s
(total), and 4000 s (total), respectively. For comparison, the XRD patterns of the same anode are shown in the as-deposited
state (A) and after aging (B). The indexed Bragg reflections correspond to α-Ni(OH)2 (PCPDF card# 00-0380715, dashed-
dotted blue line), β -Ni(OH)2 (PCPDF card #01-0802855, solid blue line), β-NiOOH (PCPDF card #- 00-006-0141, dot blue line),
γ-NiOOH (PCPDF card #- 00-006-0075, dash blue line) and nickel (PCPDF card# 04-010-6148, red line). The β-hydroxide phase
undergoes some phase transitions: the β-hydroxide peaks become smaller after charging for 2000 s (22 mAh/cm2, curve C).
The reduction in the β-hydroxide peaks most likely results from amorphization during the charging process. The formation
of the γ-oxyhydroxide phase becomes visible after total charging of 2500 s (28 mAh/cm2) and is more significant after 3000
s (33 mAh/cm2, curves D and E, respectively). Additional charging for 1000 s (total 4000s, 44 mAh/cm2) results in significant
γ-oxyhydroxide formation (curve F). Therefore, charging at a constant current of 40 mA for 2000 s (total charge of 22
mAh/cm2) is considered safe, but charging at the same current for 2500 s (28 mAh/cm2) gives rise to overcharging that results
in partial phase transformation to γ-oxyhydroxide. Accordingly, in the following experiments, the initial charging step was
limited to 22 mAh/cm2 to avoid the formation of the γ-oxyhydroxide phase.

Supplementary Figure 19: X-ray diffractograms of a Ni0.9Co0.1(OH)2 anode measured after different stages of
electrochemical (chare/discharge) cycling. A Ni0.9Co0.1(OH)2 anode was aged for 30 min in 5M KOH at 90°C and pre-charged
for 2000 s at a constant current of 40 mA (applied on an anode area of 1 cm2 immersed in the electrolyte solution). Then,
the anode was used for the electrochemical cycling analysis in 5M KOH at 25°C. The anode was charged and discharged for
100 s at a constant current of 40 mA for a total of ten cycles. Curve (C) presents the diffractogram prior to cycling, after an
initial pre-charging at 40 mA (applied on an anode area of 1 cm2 immersed in the electrolyte solution) for 2000 s and
discharging at the same current for 100 s. Curves (D) and (E) present diffractograms taken after the electrochemical charging
step in the first and last cycles. Curve (F) presents the diffractogram taken after the final discharging step. For comparison,
the XRD patterns of the same anode are shown in the as-deposited state (A) and after aging (B). Electrochemical discharging
does not transform the layer back to the same degree of crystallinity that was observed before the initial charging step (i.e.,
in the aged state prior to cycling).

22
Supplementary Figure 20: XRD diffractograms of a Ni0.9Co0.1(OH)2 anode measured after different stages of E-TAC
(chare/regeneration) cycling. A Ni0.9Co0.1(OH)2 anode was aged for 30 min in 5M KOH at 90°C and pre-charged for 2000 s at
a constant current of 40 mA (applied on an anode area of 1 cm2 immersed in the electrolyte solution). Then, the anode was
used for E-TAC cycling analysis. All electrochemical steps were carried out in 5M KOH at 25°C. The charging steps were
performed at the same conditions (40 mA, 100 s), while the regeneration steps were carried out by immersing the anode in
5M KOH aqueous electrolyte at 90°C for 100 s. A total of ten E-TAC cycles were carried out. Curve (C) presents the
diffractogram prior to cycling, after an initial pre-charging at 40 mA (applied on an anode area of 1 cm2 immersed in the
electrolyte solution) for 2000s and regeneration in 5M KOH at 90°C for 100 s. Curves (D) and (E) present the diffractogram
after the electrochemical charging step in the first and last cycle. Curve (F) presents the diffractogram taken after the final
regeneration step. For comparison, the XRD patterns of the same anode are shown in the as-deposited state (A) and after
aging (B). As in the electrochemical discharging, chemical regeneration does not transform the layer back to the same degree
of crystallinity that was observed before the initial charging step (i.e., in the aged state prior to cycling). Rather, the semi-
amorphous oxyhydroxide phase transforms to a semi-amorphous hydroxide phase. There are no signs of degradation, γ-
phase formation or any other differences before and after ten full cycles.

23
Supplementary Figure 21: SEM micrograph of Ni0.9Co0.1(OH)2 anodes subjected to electrochemical and E-TAC cycles. Two
Ni0.9Co0.1(OH)2 anodes were aged for 30 min in 5M KOH at 90°C and pre-charged for 2000 s at a constant current of 40 mA
(applied on an anode area of 1 cm2 immersed in the electrolyte solution). Then, one anode was used for electrochemical
(charge-discharge) cycling and the second anode was used for E-TAC cycling. All electrochemical steps were carried out at
25°C. For electrochemical cycling, the anode was charged and discharged for 100 s at a constant current of 40 mA. For E-TAC
cycling, the anode was charged at the same conditions (40 mA, 100 s) while the regeneration steps were carried out by
immersing the anode in 5M KOH at 90°C for 100 s. Figures (A) and (B) show low and high magnification images of an anode
subjected to electrochemical (charging-discharging) cycles after ten cycles. Figures (C) and (D) show low and high
magnification images of an anode subjected to E-TAC (charging-regeneration) cycling, after ten cycles. There is no significant
difference between the anode that was subjected to E-TAC cycles compared to the counterpart anode that was subjected to
electrochemical cycles.

24
Supplementary Figure 22: Optical micrographs of a Ni0.9Co0.1(OH)2 anode in the as-deposited state, before and after SEM
imaging. Figures (A) and (B) show low and high magnification optical micrographs of the anode before SEM imaging,
respectively. Figures (C) and (D) show low and high magnification optical micrographs of the same anode after SEM imaging,
respectively. The optical micrographs show conformal coating of the Ni0.9Co0.1(OH)2 layer on the nickel foam substrate. No
cracks are observed before the SEM imaging (figures A and B), whereas after the SEM imaging cracks are distinctly observed
(figures C and D). Since the optical imaging was carried out under wet conditions, the cracks that can be observed only after
SEM imaging must have formed due to drying of the anode under the high vacuum conditions inside the SEM.

Supplementary Figure 23: SEM micrographs showing the drying shrinkage crack formation in a Ni0.9Co0.1(OH)2 anode (as-
deposited) during SEM imaging. To decelerate the drying shrinkage crack formation process, the wet anodes were frozen
prior to imaging in the SEM, so that under the high vacuum conditions inside the SEM the ice in the voids within the nickel
foam could evaporate without melting. Figures (A) and (B) capture the drying shrinkage crack formation process that takes
place under the SEM high vacuum conditions. Figure (A) shows a selected area of the anode at the beginning of the imaging
process, and Figure (B) shows the same area after a time delay of ~1 min. The cracking process is evident in the
transformation of the marked areas, from A, B, and C in figure (A) to A', B' and C' in figure (B). The dashed circled area R in
figures (A) and (B) is an unchanged reference marker in the two micrographs.

25
Supplementary Figure 24: Optical micrographs of an aged Ni0.9Co0.1(OH)2 anode, before and after SEM imaging. The optical
micrographs of a Ni0.9Co0.1(OH)2 anode aged in 5M KOH aqueous solution at 90°C for 20 min. show conformal coating on the
nickel foam substrate after aging, comparable to the same coating observed before aging, as presented in Supplementary
Figure 22. Figures (A) and (B) show low and high magnification micrographs of the anode before SEM imaging, respectively.
Figures (C) and (D) low and high magnification micrographs of the anode after SEM imaging, respectively. The insets in figure
(D) present high magnification micrographs of the selected area in figure (C) showing a peeled-off layer with a thickness of
~10 µm.

26
A 0.5

Reversible potential [VHg/HgO]


0.45

0.4

0.35

0.3
Without cobalt, Barnard et al., 1980
0.25 With cobalt
Without cobalt
0.2
0% 20% 40% 60% 80% 100%
State of Charge [%]

B 1.8

1.7 E,r 10 mA 20 mA 50 mA 100 mA


Potential [VRHE]

1.6

1.5

1.4

1.3

1.2
0% 20% 40% 60% 80% 100%
State of Charge [%]

Supplementary Figure 25: Reversible and polarization potentials of nickel hydroxide anodes. (A) The reversible potentials
of nickel hydroxide anodes with and without cobalt. The orange series represents data extracted from the literature 21 for
the β-Ni(OH)2/β-NiOOH couple in 1M KOH solution. There is an increase in the reversible potential between x = 0 and 0.3,
followed by a constant plateau region between x = 0.3 and 0.75, and a small drop above x = 0.75, which is attributed to γ-
phase formation (the γ-phase has a lower reversible potential than the β-phase). The green and orange series represent
measured data for (our) nickel hydroxide anodes without cobalt (Ni(OH)2, green) and with cobalt (Ni0.9Co0.1(OH)2, blue) as a
function of the anodes' SOC in 5M KOH solution. At a low SOC (< 60%), the reversible potential of the anode with cobalt is
significantly lower (by 20-60 mV) than that of its cobalt-free counterpart anode. At high SOC (> 60%), the difference is
reduced to 10 mV. (B) The polarization potentials of a Ni0.9Co0.1(OH)2 anode were measured for various SOCs at constant
currents of 10 mA (blue), 20 mA (green), 50 mA (orange) and 100 mA (red) in 5M KOH solution. The reversible potential (at
zero current) is also shown for comparison (black X marks). The currents were applied on an anode area of 1 cm 2 immersed
in the electrolyte solution. The overpotential for the Ni(OH)2 → NiOOH charging reaction ranges from 62±10 mV to 250±16
mV for currents ranging from 10 to 100 mA, respectively.

27
1200
1000 With cobalt

Current density [mA/cm2]


800 Without cobalt
600
400
200
0
-200
-400
-600
-800
-1000
-0.4 -0.2 0 0.2 0.4 0.6 0.8
Potential [VHg/HgO]

Supplementary Figure 26: Cyclic voltammograms of discharged nickel hydroxide anodes. Cyclic voltammetry sweeps were
carried out for nickel hydroxide anodes without cobalt (Ni(OH)2, green) and with cobalt (Ni0.9Co0.1(OH)2, blue). Scan rate 50
mV/s.

A 600 B 600
500 500
1.3 V
Current [mA]
Current [mA]

400 1.4 V 400 1.3 V


1.5 V 1.4 V
300 300
1.6 V 1.5 V
200 200 1.6 V
100 100
0 0
0 200 400 600 0 200 400 600
Time [s] Time [s]

C 160
Current density [mA/cm2]

140
120 QC
100 QOER
80
60
40
20
0
0 500 1000 1500 2000 2500
Time (s)

Supplementary Figure 27: Chronoamperometric charging of Ni0.9Co0.1(OH)2 and NiFe LDH anodes at different potentials.
Figures (A) and (B) show the amperogram traces of a NiFe LDH and a Ni0.9Co0.1(OH)2 anode, respectively, measured during
charging at 1.3, 1.4, 1.5, 1.6 and 1.7 VRHE for 600 s. Figure (C) shows the amperomegram trace of an initially discharged
Ni0.9Co0.1(OH)2 anode charged at a constant potential of 1.48 VRHE for 2300 s. All the measurements were carried out in 5M
KOH solution, without heating.

28
30 0.2

25
Current density [mA/cm2]

0.15

DO volume [ml]
20

15 0.1

10
0.05
5

0 0
25 30 35 40 45 50 55 60
Charge [C/cm2]

Supplementary Figure 28: Chronoamperometry and dissolved oxygen measurements during deep charging of a fully
discharged Ni0.9Co0.1(OH)2 anode at a constant potential of 1.48 VRHE. The current density (blue) and DO volume (red) during
chronoamperometric charging of a Ni0.9Co0.1(OH)2 anode at a constant potential of 1.48 VRHE, measured in cold (ambient
temperature) 0.5M KOH solution, from a fully discharged state. In this, and all other Ni0.9Co0.1(OH)2 anode charging tests, an
initial decline in DO concentration was observed, which is attributed to the purging of the electrolyte due to hydrogen
bubbling from the cathode. It is noted that no such decline was observed in the electrolysis test (see inset in Supplementary
Figure 29), where hydrogen was also generated at the cathode at a current of 100 mA. For all the negative DO values, the
result was corrected to zero, and for all the positive values the volume is presented as the difference between the measured
volume at each charge value and the volume at the beginning of the DO rise. The error bars represent the measurement
error of the DO probe.

1.6
1.4
1.2
DO volume [ml]

1
0.8
0.6
0.4 E-TAC hydrogen generation
Electrolysis
0.2
0
-0.2
0 0.5 1 1.5 2 2.5 3
Charge [C/cm2]

Supplementary Figure 29: Dissolved oxygen measurements. The volume of dissolved oxygen (DO) obtained during the
hydrogen generation step of the E-TAC cycle (red) with a Ni0.9Co0.1(OH)2 anode that had been regenerated before the
measurement, and during a control test of water electrolysis (blue) with an uncoated nickel foam anode. The measurements
were carried out in 0.5M KOH solution, at a current of 100 mA (current density of 10 mA/cm2), without heating. For all the
negative DO values, the result was corrected to zero, and for all the positive values the volume is presented as the difference
between the measured volume at each charge value and the volume at the beginning of the DO rise. The error bars represent
the measurement error of the DO probe.

29
30 0.30

Current density [mA/cm2]


25 0.25

DO volume [ml]
20 0.20

15 0.15

10 0.10

5 0.05

0 0.00
0 2 4 6 8
Charge [C/cm2]

Supplementary Figure 30: Chronoamperometry and dissolved oxygen measurements during charging a regenerated
Ni0.9Co0.1(OH)2 anode at a constant potential of 1.48 VRHE. The current density (blue) and DO volume (red) during
chronoamperometric charging of a Ni0.9Co0.1(OH)2 anode at a constant potential of 1.48 VRHE, measured in cold (ambient
temperature) 0.5M KOH solution, following regeneration for 120 s in hot (95°C) 5M KOH solution (as in the regeneration
step in the E-TAC process). For all the negative DO values, the result was corrected to zero, and for all the positive values the
volume is presented as the difference between the measured volume at each charge value and the volume at the beginning
of the DO rise. The error bars represent the measurement error of the DO probe.

4.5 After nitrogen purge


E-TAC hydrogen generation
O2 (area=28.6)
Detector response [AU]

3.5
H2
2.5 O2 (area=29.4)

1.5

0.5

-0.5
0 1 2 3 4
Retention time [min]

Supplementary Figure 31: Gas chromatography. Gas chromatograms of a sample taken from the electrochemical cell after
purging with nitrogen gas (black line) and after hydrogen generation at a constant current density of 50 mA/cm2 for 100 s.
The measurements were carried out in 5M KOH solution, without heating.

30
A 1.60 B 1.70

1.65

Anode Potential [VRHE]


1.55
1.60

Cell Voltage [V]


1.50 1.55

1.45 1.50
200 mA 1.45
100 mA 200 mA
1.40 50 mA 100 mA
25 mA 1.40 50 mA
10 mA 25 mA
10 mA
1.35 1.35
0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0
Charge [C] Chrage [C]

C 1.60 D 1.70

1.65
Anode Potential [VRHE]

1.55
Cell Voltage [V] 1.60
1.50 1.55

1.45 1.50
200 mA
100 mA 1.45 200 mA
1.40 50 mA 100 mA
25 mA 1.40 50 mA
10 mA 25 mA
10 mA
1.35 1.35
0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0
Charge [C] Charge [C]

Supplementary Figure 32: Anode potential and cell voltage during E-TAC cycles carried out at different currents. Figures
(A) and (B) present the anode potential and cell voltage, respectively, during the hydrogen generation steps of E-TAC cycles
carried out at 10, 25, 50, 100 and 200 mA/cm2. Five cycles were performed at each current. Figures (C) and (D) present the
average values for each measurement set presented in Figures (A) and (B), respectively. The solid lines in Figures (C) and (D)
represent the average charging potentials and average cell voltages throughout these cycles, respectively, while the dashed
lines of the same colours represent the average minimum and maximum boundaries, calculated based on the results
presented in Figure (A) and (B).

31
A 1.70
1.65
Anode potential [VRHE]

1.60
1.55
1.50
1.45
1.40
1.35
1.30
0 50 100 150 200
Current density [mA/cm2]

B 1.65

1.60

1.55
Cell voltage [V]

1.50

1.45

1.40

1.35
0 50 100 150 200
Current density [mA/cm2]

Supplementary Figure 33: The anode potential and cell voltage as a function of the current density in the experiments
presented in Supplementary Figure 34. The error bars correspond to the average-minimum and average-maximum values,
and each open circle represents the average anode potential (A) or cell voltage (B) in all ten cycles (at a given current density).
The colour coding matches the colours in Supplementary Figure 32. The full black circles in (A) represent the steady-state
polarization results of the same anode, extracted from the data in Supplementary Figure 27B, and the full black squares in
(A) represent the steady-state polarization results of the NiFe LDH anode, extracted from the data in Supplementary Figure
27A. The average anode potential and cell voltage increased from 1.43 VRHE (potential) and 1.44 V (voltage) at 10 mA/cm2 to
1.53 VRHE (potential) and 1.6 V (voltage) at 200 mA/cm2.

32
1.55

1.525

1.5 1
10
1.475
Potential [VRHE] 20
30
1.45
40
1.425 50
60
1.4 70
80
1.375 90
1.35 100

1.325
0 10 20 30 40 50 60 70
Time [s]

Supplementary Figure 34: Stability and cyclability test. The anode potential was measured as a function of time and is
shown here for every 10 cycles to a total of 100 cycles. There is a small decrease in the anode potential vs. time profile from
the 1st cycle to the 20th cycle, whereas subsequent cycles followed nearly the same behaviour as in the 20th cycle. This
indicates that the anodes are stable, and the process is cyclable for many cycles.

1.52
̶ ̶ 0.6 M K2CO3, 0.4M
1.50
- - - 5M KOH
1.48
Potential [VRHE]

1.46

1.44
1
2
1.42 3
4
1.40 5
6
7
1.38 8
9
10
1.36
0 20 40 60 80 100
Time [s]

Supplementary Figure 35: E-TAC water splitting in a carbonate-bicarbonate electrolyte. The anode potential during the
hydrogen generation steps in ten consecutive 100s/100s E-TAC cycles taking place in carbonate buffer at a constant current
of 25 mA (applied to an anode are of 1 cm2 immersed in the electrolyte solution). The dashed lines are for E-TAC cycles in
5M KOH solution taken from Figure 3.

33
350
300
250
Electrical energy (electrochemical step)

Energy [kJ]
200 Heat (electrochemical step)
150 Energy (electrochemical step)
Heat (total E-TAC)
100 Energy (total E-TAC)
50
0
-50
1.42 1.44 1.46 1.48 1.5 1.52 1.54 1.56
Cell voltage [V]

Supplementary Figure 36: Electrical energy, heat and total energy demands of the E-TAC process as a function of the cell
voltage. The dashed lines are for the hydrogen generation step and the full lines are for the whole cycle. The dashed blue
line represents the electrical energy input of the electrochemical (hydrogen generation) step, which increases linearly with
the cell voltage. The dashed orange line represents the heat consumption during the electrochemical step, which equals the
difference between the thermoneutral voltage and the actual cell voltage, 1.56 V – Vcell. The dashed red line represents the
total energy requirement for the electrochemical step, which is the sum of the electrical energy and the heat during that
step. The full orange line represents the heat that consumption for the entire E-TAC process, considering the heat released
during the chemical (regeneration) step, equal to 15.3 kJ. Finally, the full red line represents the total energy investment for
a full E-TAC cycle, including electrical energy and heat.

34
Supplementary Figure 37: A thermodynamic diagram of the E-TAC water splitting process. The reactants are H2O and
2Ni(OH)2 at 25°C, and the products are H2, ½O2 and 2Ni(OH)2 at 25°C. The reactants may be converted to the products by a
̂ °𝑟𝑥𝑛 1 = 285.5 𝑘𝐽⁄𝑚𝑜𝑙 𝐻2 (arrow 1). In
single electrolytic reaction, H2O → H2 + ½O2, with a standard enthalpy change of ∆𝐻
our E-TAC process, the reactants are first converted to intermediate products, H2O, H2 and 2NiOOH, in an electrochemical
reaction, 2Ni(OH)2 → 2NiOOH + H2, having a standard enthalpy change of ∆𝐻 ̂ °𝑟𝑥𝑛 2 = 301.1 𝑘𝐽⁄𝑚𝑜𝑙 𝐻2 (arrow 2).
Equivalently, this process can be considered as two separate processes: an adiabatic electrochemical reaction having an
enthalpy change of ∆𝐻 ̂𝑟𝑥𝑛 4 during which the system is cooled down to a temperature Tc < 25°C (arrow 4) followed by a
heating step having an enthalpy change of ∆𝐻 ̂ °𝐻𝑒𝑎𝑡𝑖𝑛𝑔 1 (arrow 5). Then, the intermediate product NiOOH undergoes a
spontaneous self-discharge (regeneration) reaction, 2NiOOH + H2O → 2Ni(OH)2 + ½O2, having a standard enthalpy change
̂ °𝑟𝑥𝑛 3 = −15.3 𝑘𝐽⁄𝑚𝑜𝑙 𝐻2 (arrow 3), yielding the same final products (H2, ½O2 and 2Ni(OH)2 at 25°C). Although the
of ∆𝐻
process is spontaneous, the intermediate products must be heated to achieve significant reaction rates. The heat that must
̂ℎ𝑒𝑎𝑡𝑖𝑛𝑔 2 (arrow 6). The reaction then takes place at an elevated temperature, Th, with an enthalpy
be supplied is given by ∆𝐻
5
̂𝑟𝑥𝑛
change of ∆𝐻 (arrow 7) that depends on the temperature. If the products are allowed to cool back to 25°C within the
1
̂ °𝐶𝑜𝑜𝑙𝑖𝑛𝑔 (arrow 8). The sum of these processes (heating the reactants,
system's boundaries, heat is released equal to ∆𝐻
2 5 1
̂ℎ𝑒𝑎𝑡𝑖𝑛𝑔 + ∆𝐻
reaction at Th, and cooling the products) equals the standard enthalpy of reaction, ∆𝐻 ̂𝑟𝑥𝑛 + ∆𝐻
̂ °𝐶𝑜𝑜𝑙𝑖𝑛𝑔 =
3
∆𝐻̂ °𝑟𝑥𝑛 = −15.3 𝑘𝐽⁄𝑚𝑜𝑙 𝐻2 .

35
Supplementary Tables

Supplementary Table 1: Compositional (EDS) analysis of Ni0.9Co0.1(OH)2 anodes after different stages in their fabrication
and operation processes. Two Ni0.9Co0.1(OH)2 anodes were aged for 30 min in 5M KOH at 90°C and pre-charged for 2000 s
at a constant current of 40 mA (applied on an anode area of 1 cm2 immersed in the electrolyte solution). Then, one anode
was used for electrochemical (charge-discharge) cycling and the second anode was used for E-TAC cycling. All electrochemical
steps were carried out at 25°C. For electrochemical cycling, the anode was charged and discharged for 100 s at a constant
current of 40 mA. For E-TAC cycling, the anode was charged at the same conditions (40 mA, 100 s) while the regeneration
steps were carried out by immersing the anode in 5M KOH at 90°C for 100 s. The analysis was performed in six locations for
each sample, collecting 3 spectra at each point (i.e., 18 spectra were collected in total for each sample). Representative SEM
images and EDS spectra are shown in respective columns. The numbered marks on the SEM images show the location for
the EDS spectrum collection. The EDS was performed on peeled off layers to avoid electron beam interactions with the nickel
foam substrate for an accurate compositional analysis of the Ni0.9Co0.1(OH)2 layer, yielding a Ni : Co atomic ratio of 6.7 : 1
(i.e., Ni0.9Co0.1(OH)2), showing no significant variation in layer composition as a result of cycling.

Average atomic
Sample SEM micrograph EDS spectrum composition of
cobalt (at.%)*

Aged 12.6±0.9

Precharged 12.8±0.6

After ten
electrochemical 13.6±0.7
cycles

After ten E-TAC


13.3±0.9
cycles

* [cobalt at. %] / ([cobalt at. %] + [nickel at. %])

36
Supplementary Table 2: Thermodynamic data for the E-TAC water splitting reactions. The standard enthalpy change (∆𝐻 ⊖ )
and Gibbs free energy change (∆𝐺 ⊖ ) of the hydrogen and oxygen generation steps in the E-TAC cycle with undoped
Ni(OH)2/NiOOH anodes. The ∆𝐻 ⊖ and ∆𝐺 ⊖ values for the oxygen generation step were obtained by subtracting the
respective values for the hydrogen generation step from those of the overall water splitting reaction, using data from 27.

∆𝑯⊖ ∆𝑮⊖
Step Reaction (kJ/mol H2) (kJ/mol H2)
(eV/electron) (eV/electron)
301.1 274.7
1 - Hydrogen generation 2Ni(OH)2 → 2NiOOH + H2
(1.56) (1.42)
2 - Oxygen generation 2NiOOH + H2O → 2Ni(OH)2 + ½O2 -15.3 -37.6
285.8 237.2
Total H2O → H2 + ½O2
(1.48) (1.23)

37
Supplementary References
1. McCrory, C. C. L. et al. Benchmarking hydrogen evolving reaction and oxygen evolving
reaction electrocatalysts for solar water splitting devices. J. Am. Chem. Soc. 137, 4347–4357
(2015).
2. Mcbreen, J. Nickel hydroxide. in Handbook of battery materials (eds. Daniel, C. & Besenhard,
J. O.) 149–168 (Wiley-VCH Verlag GmbH & Co, 2011).
3. Hall, D. S., Lockwood, D. J., Bock, C., Macdougall, B. R. & Lockwood, D. J. Nickel hydroxides
and related materials : a review of their structures , synthesis and properties. Proc. R. Soc.
471, 20140792 (2014).
4. Oliva, P., Laurent, J. F. & Leonardi, J. Review of the structure and the electrochemistry of
nickel hydroxides and oxy-hydroxides. J. Power Sources 8, 229–255 (1982).
5. Li, J. et al. Calcium metaborate as a cathode additive to improve the high-temperature
properties of nickel hydroxide electrodes for nickel–metal hydride batteries. J. Power Sources
263, 110–117 (2014).
6. Provazi, K., Giz, M. J., Dall’Antonia, L. H. & Córdoba de Torresi, S. I. The effect of Cd, Co, and
Zn as additives on nickel hydroxide opto-electrochemical behavior. J. Power Sources 102,
224–232 (2001).
7. Aghazadeh, M., Golikand, A. N. & Ghaemi, M. Synthesis, characterization, and
electrochemical properties of ultrafine beta-Ni(OH)2 nanoparticles. Int. J. Hydrogen Energy
36, 8674–8679 (2011).
8. Chen, J., Bradhurst, D. H., Dou, S. X. & Liu, H. K. Nickel hydroxide as an active material for the
positive electrode in rechargeable alkaline batteries. J. Electrochem. Soc. 146, 3606-3612
(1999).
9. Zhou, H. et al. Water splitting by electrolysis at high current densities under 1.6 volts. Energy
Environ. Sci. 11, 2858–2864 (2018).
10. Han, L., Dong, S. & Wang, E. Transition-metal (Co, Ni, and Fe)-based electrocatalysts for the
water oxidation reaction. Adv. Mater. 28, 9266–9291 (2016).
11. Jamesh, M. I. & Sun, X. Recent progress on earth abundant electrocatalysts for oxygen
evolution reaction (OER) in alkaline medium to achieve efficient water splitting – a review. J.
Power Sources 400, 31–68 (2018).
12. LeRoy, R. L. & Bowen, C. T. The thermodynamics of aqueous water electrolysis. J.
Electrochem. Soc. 127, 1954–1962 (1980).
13. Mao, Z. & White, R. E. The self-discharge of the NiOOH/Ni(OH)2 electrode constant potential
study. J. Electrochem. Soc. 139, 1282–1289 (1992).
14. Conway, B. E. & Bourgault, P. L. The electrochemical behavior of the nickel – nickel oxide
electrode: part I. kinetics of self-discharge. Can. J. Chem. 37, 292–307 (1959).
15. Armstrong, G. & Butler, J. A. V. The rate of decay of hydrogen and oxygen overvoltages.
Trans. Faraday Soc. 29, 1261–1266 (1933).
16. Briggs, G. W. D., Jones, E. & Wynne-Jones, W. F. K. The nickel oxide electrode. part 1. Trans.
Faraday Soc. 51, 1433–1442 (1955).

38
17. Briggs, G. & Wynne-Jones, W. The Nickel Oxide Electrode. Part 3. Trans. Faraday Soc. 52,
1272–1281 (1956).
18. Iwakura, C. et al. Self-discharge mechanism of nickel hydrogen batteries using metal hydride
anodes. J. Electrochem. Soc. 136, 1351-1355 (1989).
19. Macdonald, D. D. & Challingsworth, M. L. Thermodynamics of nickel-cadmium and nickel-
hydrogen batteries. J. Electrochem. Soc. 140, 606–609 (1993).
20. Beverskog, B. & Puigdomenech, I. Revised Pourbaix diagrams for nickel at 25-300 C. Corros.
Sci. 39, 969–980 (1997).
21. Barnard, R., Randell, C. F. & Tye, F. L. Studies concerning charged nickel hydroxide electrodes
. I. measurement of reversible potentials. J. Appl. Electrochem. 10, 109–125 (1980).
22. Bourgault, P. L. & Conway, B. E. The electrochemical behavior of the nickel oxide electrode:
part ii. quasi-equilibrium behavior. Can. J. Chem. 38, 1557–1575 (1960).
23. Chang, S., Young, K., Nei, J. & Fierro, C. Reviews on the U.S. Patents regarding nickel/metal
hydride batteries. Batteries 2, 10 (2016).
24. Landman, A. et al. Photoelectrochemical water splitting in separate oxygen and hydrogen
cells. Nat. Mater. 16, 646–651 (2017).
25. Esposito, D. V. Membraneless electrolyzers for low-cost hydrogen production in a renewable
energy future. Joule 1, 651–658 (2017).
26. Lædre, S., Kongstein, O. E., Ødegård, A., Karoliussen, H. & Seland, F. Materials for Proton
Exchange Membrane water electrolyzer bipolar plates. Int. J. Hydrogen Energy 42, 2713–2723
(2017).
27. Meites, L. Handbook of analytical chemistry. (McGraw-Hill, 1963).

39

You might also like