You are on page 1of 9

Letter

http://pubs.acs.org/journal/aelccp

Pilot-Scale CO2 Electrolysis Enables a Semi-


empirical Electrolyzer Model
Jonathan P. Edwards, Théo Alerte, Colin P. O’Brien, Christine M. Gabardo, Shijie Liu, Joshua Wicks,
Adriana Gaona, Jehad Abed, Yurou Celine Xiao, Daniel Young, Armin Sedighian Rasouli, Amitava Sarkar,
Shaffiq A. Jaffer, Heather L. MacLean, Edward H. Sargent, and David Sinton*
Cite This: ACS Energy Lett. 2023, 8, 2576−2584 Read Online
Downloaded via HUAZHONG UNIV SCIENCE & TECHNOLOGY on October 12, 2023 at 15:45:34 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Carbon dioxide (CO2) electrolysis powered with renewable


electricity can help close the carbon cycle by converting emissions into
chemicals and fuels. Two key advancements are required to bridge the
technological gaps for industrial implementation: pilot plant demonstra-
tions with detailed performance data; and chemical engineering process
models built and tested with lab- and pilot-scale data. Here, we develop a
semi-empirical electrolyzer model in Aspen Custom Modeler which is
trained on a 5 cm2 lab-scale CO2 electrolyzer. We then scale to a pilot-scale
800 cm2 single cell and 10 × 800 cm2 stack and use the results to validate
the model; at 100 mA cm−2, the model can predict six of seven cell performance metrics within 16% absolute error and three
of five stack metrics within 11% absolute error. With the combination of the electrolyzer model and the pilot-scale data, this
work provides the prerequisites for further scaling of CO2 electrolysis.

T he decreasing cost of renewable electricity and


increasing demand for renewable energy storage
motivate the electrochemical conversion of CO2 into
chemicals and fuels as a means to decarbonize chemical
manufacturing while closing the carbon cycle to address global
Industrial CO2 electrolyzers will require integration with
upstream and downstream chemical processes.28−30 To plan
for these deployments, chemical simulations must be able to
accurately predict CO2 electrolyzer behavior at relevant scales.
Multiphysics simulations have been used to estimate reaction
warming.1,2 Of particular interest, multi-carbon products, such kinetics, model the pH distribution, improve CO2 utilization,
as ethylene (C2H4), have high market value and large global identify failure mechanisms, and provide a better fundamental
markets.3,4 Membrane electrode assemblies have been understanding of lab-scale reactors.31−35 Most simulations of
successfully incorporated into zero-gap CO2 electrolyzers, CO2 electrolysis are one-dimensional simulations focused on
yielding stable and efficient conversion of CO2 into multi- single-carbon products with cell sizes <10 cm2 (Table S5).
carbon products at industrially relevant current densities (i.e., Unfortunately, the computational expense of these simulations
>100 mA cm−2).5−8 Two key advancements are required to becomes prohibitive as the reactor grows due to the number of
bridge the scale and technological readiness gaps: pilot-scale physical parameters and scales involved.36,37 Models have been
demonstrations with detailed performance data; and chemical developed for CO2 electrolyzers which are useful in predicting
engineering process models�built and tested with pilot-scale specific elements of electrolyzer performance (e.g., cell
data�to design and optimize larger deployments. voltage) but do not produce a complete mass and energy
Demonstrations of CO2 electrolysis have been limited to the balance.3 A comprehensive prediction of mass and energy
lab-scale, with most cell areas <10 cm2.9 As the cell size balance is crucial to assess the feasibility of CO2 electrolysis in
increases, discrepancies can develop across the active area in the context of competing conventional and emerging
electrode composition,10,11 electrode compression,12 CO2
technologies.
concentration,13−16 pressure,17−19 temperature,20 and current
density/voltage,21,22 all of which impact CO2 electrolysis
selectivity and efficiency.4,23 For CO2 electrolysis into multi- Received: March 22, 2023
carbon products, the largest single cell reported is 25 cm2,24,25 Accepted: May 8, 2023
and the largest stack is 30 cm2 (3 × 10 cm2 cell stack).26 Published: May 11, 2023
Although informative, these demonstrations of CO2 electrol-
ysis are several orders of magnitude too small to inform on
operation at industrial scale (i.e., electrolyzer area >100 m2).27

© 2023 American Chemical Society https://doi.org/10.1021/acsenergylett.3c00620


2576 ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 1. CO2 electrolysis at different scales. (a) Key inputs and outputs for the CO2 electrolyzer model. The key inputs are T (the stack
temperature), PCathode (the cathode pressure), PAnode (the anode pressure), jT (the current density) or VCell (the cell voltage), wChannel (the
width of the flow channel), eChannel (the depth of the flow channel), AGeometric (the active area), Sh (the Sherwood number), θChannel (the
empirical mass transfer coefficient modifier), αi (the product-specific electrochemical transfer coefficient), jo,i (the product-specific exchange
current density), tElectrode (the electrode thickness), εElectrode (the electrode porosity), CathodeIn (the cathode supply rate and composition),
and AnodeIn (the anode supply rate and composition). The key outputs are jT or VCell, EC (the cathode half-cell potential), EA (the anode half-
cell potential), ΔVNer (the Nernstian loss), ΔVOhm (the ohmic loss), FEi (the product-specific Faradaic efficiency), ji (the product-specific
partial current density), QReactor (the heat flux generated by the reactor), Electricity (the electricity required for the reactor), CathodeOut (the
cathode exhaust rate and composition), and AnodeOut (the anode exhaust rate and composition). (b) Lab-scale electrolyzer with an active
area of 5 cm2. (c) Pilot-scale electrolyzer stack having 10 cells each with an active area of 800 cm2.

Semi-empirical models have been developed for more Anodic CO2 Crossover. In CO2 electrolyzers using anion
mature electrochemical technologies, such as alkaline water exchange membranes, reactant CO2 can migrate from the
electrolyzers,38−40 proton exchange membrane water electro- cathode to the anode, so we introduced an anodic CO2 loss
lyzers,41−43 and proton exchange membrane fuel cells.44−46 coefficient, denoted by τ.47−49 The primary contributor to τ is
These models have achieved relatively high accuracies at low (bi)carbonate formation wherein CO2 chemically reacts with
computational expense and have helped to identify optimal hydroxides, travels through the membrane, and is released as
operating conditions (e.g., temperature, pressure, current CO2 upon reaching the locally acidic environment (e.g., pH <
density, electrolyte flow rate) and components (e.g., 4) next to the anode (SI, Anodic CO 2 Crossover
diaphragms, porous transport layers) for use in industrial Modeling).32,50 If hydroxide, carbonate, or bicarbonate is the
applications. A comprehensive semi-empirical model is now sole charge carrier through the membrane, then τ is 0, 5.2, or
needed for the field of CO2 electrolysis. 10.4 μmol C−1, respectively.29 Some of the anodic CO2 can
In this work, we scale up CO2 electrolysis from the lab-scale come from liquid products, which are capable of traveling
to the pilot-scale. We develop a semi-empirical model of a CO2 through anion exchange membranes51,52 and being oxidized
electrolyzer (Figure 1a) that is fit to lab-scale data (single 5 back to CO2 in the presence of an appropriate anode
catalyst.6,53
cm2 cell, Figure 1b) and validated with pilot-scale experimental
We hypothesized that τ is dictated by the ratio between
data (800 cm2 single cell or 10 × 800 cm2 cell stack, Figure 1c,
(bi)carbonates and hydroxides at the cathode, with a greater
Figure S1). The model is based on a zero-gap CO2 electrolyzer fraction of (bi)carbonates leading to more (bi)carbonate
making use of an anion exchange membrane (SI, Electrolyzer transport across the membrane. We assumed that τ could be
Custom Model, Figure S2). The model developed in this work, approximated by a separable function dependent on two
publicly available in an online repository, is written in the variables: the bulk CO2 concentration (the average of the
Aspen Custom Modeler software (Aspen Tech Inc., Bedford, cathode inlet and outlet concentrations) and the rate of
MA) for better integration with upstream and downstream hydroxide generation (i.e., the current density).
process simulations in the Aspen software family.37 When the In the first set of lab-scale experiments, a constant current of
cell voltage (or current density), electrolyzer geometry, 150 mA cm−2 was maintained while the CO2 supply rate to the
electrode properties, and incoming reactant streams are cathode was varied (SI, Anodic CO2 Crossover Measurements,
specified, the model predicts the electrochemical performance, Figure S3a). By lowering the molar flow rate of CO2 to the
energy flows, and product streams (Figure 1a). The model electrolyzer, the bulk CO2 concentration was reduced�
consists of interconnected modules for CO2 transport, anodic experimental data was fed into the electrolyzer model to
CO2 crossover, voltage losses, and electrochemical kinetics. calculate the bulk CO2 concentration (SI, Figure S3b, Model
2577 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 2. Lab-scale experiments and simulations on anodic CO2 crossover and voltage losses. (a) Anode gas flow rate and anodic CO2 loss
coefficient at different bulk CO2 concentrations (150 mA cm−2). Derived from data at five different cathode CO2 supply rates to the cathode:
10, 8, 6, 4, and 2 sccm cm−2 of CO2. (b) Anode gas flow rate and anodic CO2 loss coefficient at different current densities (12 sccm cm−2 of
CO2 supply rate). (c) Simulated cathode/anode pH and Nernstian loss at different CO2 concentrations (100 mA cm−2). (d) Simulated
cathode/anode pH and Nernstian loss at different current densities (39 mol m−3 CO2 concentration). (e) Area-specific ohmic resistance
measured at different current densities. (f) Area-specific ohmic resistance measured with different cathode gas feeds (25 mA cm−2). Error
bars correspond to the standard deviation of three independent measurements.

Inputs, Table S2). As the nominal CO2 supply rate to the nearly linear trend (Figure 2b). This shift in charge carrier
cathode was decreased from 10 to 2 sccm cm−2, there was from carbonate to hydroxide at high current densities was
almost no change in the anode gas flow rate or τ, averaging 9.7 consistent with experiments47 and simulations32,50 performed
mL min−1 and 6.4 μmol C−1, respectively (Figure 2a). The in other reports. The electrochemical rate of hydroxide
results of these experiments suggested that, within the tested generation grew with the current density, eventually generating
range, the bulk CO2 concentration does not influence τ. more hydroxides than could react with CO2 to form
To determine how the current density impacts τ, the lab- (bi)carbonates. At higher current densities, (bi)carbonate
scale electrolyzer was tested at various current densities with a
migration through the membrane plateaued and was
constant CO2 supply (Figure S4). Although varying the current
increasingly complemented with hydroxide migration. The
density also shifts the CO2 concentration, we previously
showed that the CO2 concentration did not strongly affect τ increasing proportion of hydroxides (relative to the total anion
(Figure 2a). As the current density increased 6-fold from 50 to concentration) yielded a shift in the dominant charge carrier
300 mA cm−2, the anode gas flow rate increased only 4-fold, through the membrane at high current densities. Taken
from 3.8 to 16.1 mL min−1 (Figure 2b). If anodic CO2 together, these lab-scale experiments demonstrate that τ was
crossover remained constant, then the anode gas flow rate largely insensitive to bulk CO2 concentration, within the tested
should be directly proportional to the current density. Instead, range, and primarily a linear function of current density (Table
it was calculated that τ decreased with current density in a S3).
2578 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 3. Model validation against 800 cm2 pilot-scale single cell. The experimental results and model predictions are marked with solid and
patterned fill, respectively (left-hand axis). Error is plotted on the right-hand axis. (a) Cell voltage. (b) CO2 in anode tail gas. (c) H2 in
cathode tail gas. (d) C2H4 in cathode tail gas. The molar fractions shown in (b), (c), and (d) are those of the dry tail gas streams, e.g., with
any water content removed. Error bars correspond to the standard deviation of three independent measurements.

Nernstian Losses. The model divides the voltage applied To determine the impact of current density on electrode pH,
to the electrolyzer across four areas: Nernstian loss, ohmic loss, simulations were performed starting at 75 and increased
cathode potential, and anode potential (SI, Voltage Distribu- incrementally to 150 mA cm−2 (Figure 2d, Figure S5b).
tion). Products or reactants which differ from standard Although varying the current density also shifted the CO2
concentrations induce Nernstian losses.3,54 In keeping with concentration, we previously demonstrated that the CO2
other modeling works, our model assumes that hydroxide is concentration only slightly impacted the Nernstian loss (Figure
the only species which differs significantly from standard 2c). The cathode pH increased gradually, starting at 12.7 and
concentration.6,32,50 We assume that the CO2 concentration is reaching 13.0 at the two current density extremes. This
not enough to induce a substantial Nernstian loss (a 10-fold incremental increase in pH was consistent with other
change in CO2 concentration would shift the CO2-to-C2H4 simulations reported in the literature, as the increased current
voltage by only 5 mV when operating at 40 °C). Measuring the density resulted in a local accumulation of hydroxides.32,50 The
pH near the electrodes in situ is challenging,55 so we instead anode pH was less affected by current density, dropping from
estimated the local pH by performing COMSOL simulations 4.79 to 4.73 over the same range. Given the relatively minor
following previously published procedures.8,32 To yield a change in cathode and anode pH with current density, the
conservative estimate of the pH difference, the pH values Nernstian loss increased by only 20 mV as the current was
increased, reaching 510 mV at 150 mA cm−2 (Figure 2d). In
reported for the cathode and anode were the maximum and
these simulations, the difference in pH between the cathode
minimum pH from their respective simulation domains. We
and anode was not very sensitive to CO2 concentration or
hypothesized that the bulk CO2 concentration and the current
current density, so a constant Nernstian loss (510 mV) was
density could be the critical parameters in approximating the assumed for operation at industrially relevant current densities
electrode pH. (Table S3). This Nernstian loss is comparable to that
A set of simulations with varied CO2 concentrations was predicted in other works.6,50
performed to assess how the Nernstian losses were impacted. Ohmic Losses. Resistance measurements of a lab-scale
The CO2 concentration was varied by changing the cathode CO2 electrolyzer were performed at different current densities
boundary condition in the simulation. As the CO2 concen- to quantify the losses associated with the membrane, interfaces,
tration decreased from 39 to 10 mol m−3, the anode pH and electrolyte (SI, Resistance Measurements).6,9 Area-specific
remained constant at 4.74, whereas the cathode pH increased resistance is commonly used in fuel cell and electrolyzer
slightly from 12.8 to 13.1 (Figure 2c, Figure S5a). We showed literature to deconvolute the dependency of resistance on
previously that reduced CO2 concentrations were not enough cross-sectional area.57,58 As the current density increased, there
to affect CO2 crossover (Figure 2a), but the simulations was a decrease in the ohmic resistance from 9.23 Ω cm2 at 25
indicate that reduced CO2 concentrations lessen (bi)carbonate mA cm−2 to 7.41 Ω cm2 at 75 mA cm−2, all at 40 °C (Figure
formation and thus retain slightly more hydroxides near the 2e). Similar decreases in ohmic resistance were observed for
cathode.32,56 Although the pH difference between the cathode the electrolyzer when operated at other temperatures (Figure
and anode did increase at the lowest CO2 concentration, the S6). Simulations of zero-gap CO2 electrolyzers with liquid
resulting change in Nernstian loss was only 20 mV (Figure 2c). anolytes have suggested that, although ohmic voltage drop
2579 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 4. Model validation against 10 × 800 cm2 pilot-scale stack. The experimental results and model predictions are marked with solid and
patterned fill, respectively (left-hand axis). Error is plotted on the right-hand axis. (a) Cell voltage. (b) CO2 in cathode tail gas. (c) H2 in
cathode tail gas. (d) C2H4 in cathode tail gas. The molar fractions shown in (b), (c), and (d) are those of the dry tail gas streams, e.g., with
any water content removed. Error bars correspond to the standard deviation of three independent measurements.

increases monotonically with increasing current density, the 2.1% of the experimental values across all current densities
ohmic resistance can decrease with current density.50 (Figure 3a).
Since hydroxides are much more conductive than (bi)- The model predicted slightly more CO2 in the anode tail gas
carbonates,58,59 the ohmic resistance of the membrane and the (4.2 mol% on average, Figure 3b) and, since the dry anode gas
electrolyte was expected to decrease at higher currents when is a binary O2−CO2 mixture, correspondingly less O2 when
more hydroxides were traveling through the membrane (Figure compared to experimental results (Figure S9a). The model
2b).60,61 To demonstrate this, we switched the cathode feed correctly predicted a decrease in CO2 content at higher current
gas from CO2 (i.e., (bi)carbonate charge carriers) to nitrogen densities and demonstrated an average error of 6.2% in anodic
(i.e., hydroxide charge carriers) and the resistance fell by 0.817 CO2 content across all current densities (Figure 3b).
Ω cm2 (Figure 2f). Increased current densities may also The model accurately predicted the concentration of many
improve conductivity through increased local temperature62 or components of the cathode tail gas. The deviation between the
increased membrane−catalyst contact due to liquid-product- H2 concentration measured experimentally and that predicted
induced membrane swelling.63−65 by the model was small (<1.0 mol% across all currents), with
Based on the experimental data, we incorporated into the an error between 0.4% and 16% (Figure 3c). The C2H4
model a power-law relationship between the ohmic resistance concentration had a maximum deviation of 0.49 mol% at
and the current density for electrolyzer operation at ≤150 mA 150 mA cm−2 (Figure 3d). Since the concentrations of C2H4
cm−2 (Table S3). At higher currents it is likely that the ohmic were relatively small (i.e., 1 to 2 orders of magnitude smaller
resistance will increase, due to flooding of the cathode and/or than the other gas concentrations), they were very sensitive to
membrane dehydration,66−68 and a different relationship will minor inaccuracies in CO2 crossover. As a result, the error
be needed to predict the ohmic resistance.69,70 The cathode reached 55% at 150 mA cm−2, but at slightly lower currents
kinetic constants were determined in situ by coupling the (i.e., 100−125 mA cm−2) the error was <23% (Figure 3d). The
experimental results, Nernstian losses, ohmic losses, and the model predicted much more cathodic CO (Figure S9b) and
anode kinetic parameters (SI, Kinetic Parameter Determi- less cathodic CO 2 (Figure S9c) than was observed
nation, Figures S7 and S8, Tables S1 and S4). experimentally. The experimental selectivity for gas production
Model Validation against Pilot-Scale CO2 Electrolysis. appeared low compared to the lab-scale, suggesting that
CO2 electrolysis was performed with an 800 cm2 pilot-scale increased liquid production was taking place, possibly due to
single cell and the experimental results were compared to the differences in CO2 mass transport within the larger cell (SI,
model predictions (SI, Model Inputs, Table S2). Although the CO/CO2 Concentration Discrepancy, Figure S10). At 100 mA
model predicted all elements of electrolyzer performance, cm−2, the model could predict six of the seven performance
seven of these performance metrics were selected to compare metrics within an absolute error of 16% (Figure S9d).
against experimental results: cell voltage, anodic gas CO2 electrolysis was then performed with a 10-cell stack of
composition (O2 and CO2), and cathodic gas composition 800 cm2 cells for further model validation (SI, Model Inputs,
(H2, C2H4, CO, and CO2). The model underestimated the cell Table S2). Since no anode gas samples were accessible in this
voltage by only 24−79 mV, and was accurate to within 0.7% to pilot-scale test, there were only five performance metrics for
2580 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

comparison: cell voltage and cathodic gas composition (H2, these 800 and 8000 cm2 electrolyzers. Incorporating this semi-
C2H4, CO, and CO2). The modeled voltage was again in empirical electrolyzer model and the associated pilot-scale data
relatively good agreement with the experiments, under- into chemical process simulations can expedite the design and
estimating the voltage by 7% to 11% over the range of current deployment of larger pilot plant demonstrations needed for
densities tested (Figure 4a). The CO2 concentration exhibited industrial CO2 electrolysis.
good agreement between the model and the experimental
results, with an error ranging between 2.6% and 7.2% (Figure
4b). The model accurately predicted the cathodic H 2
■ ASSOCIATED CONTENT
Data Availability Statement
concentrations over a range of current densities, demonstrating The code for the Aspen Custom Model is available at https://
a maximum error of only 11% or 0.79 mol% (Figure 4c). On github.com/jonathanedwards23/CO2-electrolyzer-model.
average, the model predicted 0.19 mol% less C2H4 in the
*
sı Supporting Information
cathode stream than the experiments which, at 100 mA cm−2,
The Supporting Information is available free of charge at
yielded a substantial error of 33% (Figure 4d). Like the single
https://pubs.acs.org/doi/10.1021/acsenergylett.3c00620.
cell results, there was some discrepancy between the CO
concentration predicted by the model and the experimental Experimental methods; electrolyzer custom model
result (Figure S11a). At 100 mA cm−2, the model could predict overview and expanded discussion; modeling studies of
three of the five performance metrics within an absolute error CO2 electrolysis with gas diffusion electrodes; perform-
of 11% and four of the metrics within an absolute error of 34% ance of the lab-scale CO2 electrolyzer at different
(Figure S11b). cathodic CO2 supply rates (150 mA cm−2); performance
Considering the model’s performance at both pilot-scale of the lab-scale CO2 electrolyzer at different current
single cell and stack levels, the discrepancy between the lab- densities (12 sccm cm−2 of cathodic CO2 supply);
scale-trained model and experimental reality increased with simulated pH profile within the MEA electrolyzer; area-
scale. These results indicate both the value of semi-empirical specific ohmic resistance at different temperatures;
reactor models and a key limitation: the need for experimental kinetic parameter determination; model validation
validation at relevant scales. The latest semi-empirical model against 800 cm 2 pilot-scale single cell; CO/CO 2
can aid in the design of the next pilot plant, and the resulting concentration discrepancy; model validation against 10
pilot plant data will further validate and refine the model. The × 800 cm2 pilot-scale stack (PDF)
implication is that the future of CO2 electrolysis is a productive
pairing of reactor modeling and ever-larger pilot plants, each
informing the next. Given the semi-empirical nature of the
■ AUTHOR INFORMATION
Corresponding Author
model, experiments are crucial for training and validation� David Sinton − Department of Mechanical and Industrial
electrolyzer variations (i.e., in electrode structure, operating Engineering, University of Toronto, Toronto, Ontario M5S
conditions, and/or flow field geometry) would require 3G8, Canada; orcid.org/0000-0003-2714-6408;
retraining of the model. A broad set of stakeholders can Email: sinton@mie.utoronto.ca
benefit from this model, including technology developers,
investors performing due diligence, governments evaluating Authors
technologies for funding or making policy decisions, and Jonathan P. Edwards − Department of Mechanical and
systems researchers conducting life cycle or techno-economic Industrial Engineering, University of Toronto, Toronto,
assessments. Ontario M5S 3G8, Canada; orcid.org/0000-0003-4000-
This work details the development of a semi-empirical CO2 5802
electrolyzer model written in Aspen Custom Modeler trained Théo Alerte − Department of Electrical and Computer
on lab-scale data. The current density was demonstrated to be Engineering, University of Toronto, Toronto, Ontario M5S
a critical parameter in approximating the CO2 crossover and 3G4, Canada; TotalEnergies American Services Inc.,
ohmic resistance of the electrolyzer. Simulations suggested that Hopkinton, Massachusetts 01748, United States
Nernstian losses remain constant irrespective of CO 2 Colin P. O’Brien − Department of Mechanical and Industrial
concentration or current density. The model was shown to Engineering, University of Toronto, Toronto, Ontario M5S
accurately predict many aspects of pilot-scale 800 cm2 and 3G8, Canada
8000 cm2 electrolyzers, including the voltage, cathodic H2, Christine M. Gabardo − Department of Mechanical and
C2H4, and CO2 gas compositions, and anodic CO2 and O2 gas Industrial Engineering, University of Toronto, Toronto,
compositions. Further data and refinement to the model can Ontario M5S 3G8, Canada; orcid.org/0000-0002-9456-
improve the accuracy, especially with respect to cathodic CO 6894
concentrations. When supplied with appropriate lab- and pilot- Shijie Liu − Department of Mechanical and Industrial
scale data, the model can be adapted to model the behavior of Engineering, University of Toronto, Toronto, Ontario M5S
a low CO2 crossover system (i.e., utilizing a porous carbonate 3G8, Canada
regeneration layer,48 microchanneled solid electrolyte,49 or Joshua Wicks − Department of Electrical and Computer
bipolar membrane configuration71,72 instead of an anion Engineering, University of Toronto, Toronto, Ontario M5S
exchange membrane). This model presents a framework 3G4, Canada; orcid.org/0000-0001-7819-1167
which can be augmented to consider additional physics, such Adriana Gaona − Department of Civil and Mineral
as liquid product and water transport through the membrane, Engineering, University of Toronto, Toronto, Ontario M5S
to model CO2 electrolyzers producing liquid products. To 1A4, Canada
predict the behavior of larger systems, the model will have to Jehad Abed − Department of Electrical and Computer
be trained on ever-larger electrolysis data, necessitating more Engineering, University of Toronto, Toronto, Ontario M5S
scaled-up demonstrations of CO2 electrolysis which build on 3G4, Canada; orcid.org/0000-0003-1387-2740
2581 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Yurou Celine Xiao − Department of Mechanical and (3) Shin, H.; Hansen, K. U.; Jiao, F. Techno-Economic Assessment
Industrial Engineering, University of Toronto, Toronto, of Low-Temperature Carbon Dioxide Electrolysis. Nat. Sustain 2021,
Ontario M5S 3G8, Canada 4 (10), 911−919.
Daniel Young − Department of Mechanical and Industrial (4) Kibria, M. G.; Edwards, J. P.; Gabardo, C. M.; Dinh, C.;
Engineering, University of Toronto, Toronto, Ontario M5S Seifitokaldani, A.; Sinton, D.; Sargent, E. H. Electrochemical CO2
Reduction into Chemical Feedstocks: From Mechanistic Electro-
3G8, Canada
catalysis Models to System Design. Adv. Mater. 2019, 31 (31),
Armin Sedighian Rasouli − Department of Electrical and No. 1807166.
Computer Engineering, University of Toronto, Toronto, (5) Gu, Z.; Shen, H.; Chen, Z.; Yang, Y.; Yang, C.; Ji, Y.; Wang, Y.;
Ontario M5S 3G4, Canada; orcid.org/0000-0002-0807- Zhu, C.; Liu, J.; Li, J.; Sham, T.-K.; Xu, X.; Zheng, G. Efficient
9468 Electrocatalytic CO2 Reduction to C2+ Alcohols at Defect-Site-Rich
Amitava Sarkar − TotalEnergies EP Research & Technology Cu Surface. Joule 2021, 5 (2), 429−440.
USA, LLC, Houston, Texas 77002, United States; SUNCAT (6) Gabardo, C. M.; O’Brien, C. P.; Edwards, J. P.; McCallum, C.;
Center for Interface Science and Catalysis, Department of Xu, Y.; Dinh, C.-T.; Li, J.; Sargent, E. H.; Sinton, D. Continuous
Chemical Engineering, Stanford University, Stanford, Carbon Dioxide Electroreduction to Concentrated Multi-Carbon
California 94305, United States; orcid.org/0000-0002- Products Using a Membrane Electrode Assembly. Joule 2019, 3 (11),
9038-2510 2777−2791.
Shaffiq A. Jaffer − Department of Mechanical and Industrial (7) Ge, L.; Rabiee, H.; Li, M.; Subramanian, S.; Zheng, Y.; Lee, J. H.;
Engineering, University of Toronto, Toronto, Ontario M5S Burdyny, T.; Wang, H. Electrochemical CO 2 Reduction in
3G8, Canada; TotalEnergies American Services Inc., Membrane-Electrode Assemblies. Chem 2022, 8 (3), 663−692.
(8) Xu, Y.; Edwards, J. P.; Liu, S.; Miao, R. K.; Huang, J. E.;
Hopkinton, Massachusetts 01748, United States; Gabardo, C. M.; O’Brien, C. P.; Li, J.; Sargent, E. H.; Sinton, D. Self-
orcid.org/0000-0001-9311-4469 Cleaning CO2 Reduction Systems: Unsteady Electrochemical Forcing
Heather L. MacLean − Department of Civil and Mineral Enables Stability. ACS Energy Lett. 2021, 6 (2), 809−815.
Engineering, University of Toronto, Toronto, Ontario M5S (9) Chen, Y.; Vise, A.; Klein, W. E.; Cetinbas, F. C.; Myers, D. J.;
1A4, Canada Smith, W. A.; Deutsch, T. G.; Neyerlin, K. C. A Robust, Scalable
Edward H. Sargent − Department of Electrical and Computer Platform for the Electrochemical Conversion of CO2 to Formate:
Engineering, University of Toronto, Toronto, Ontario M5S Identifying Pathways to Higher Energy Efficiencies. ACS Energy
3G4, Canada; orcid.org/0000-0003-0396-6495 Letters 2020, 5 (6), 1825−1833.
(10) Oh, S.; Park, H.; Kim, H.; Park, Y. S.; Ha, M. G.; Jang, J. H.;
Complete contact information is available at:
Kim, S.-K. Fabrication of Large Area Ag Gas Diffusion Electrode via
https://pubs.acs.org/10.1021/acsenergylett.3c00620 Electrodeposition for Electrochemical CO2 Reduction. Coatings 2020,
10 (4), 341.
Notes (11) Kato, N.; Mizuno, S.; Shiozawa, M.; Nojiri, N.; Kawai, Y.;
The authors declare the following competing financial Fukumoto, K.; Morikawa, T.; Takeda, Y. A Large-Sized Cell for Solar-
interest(s): An invention disclosure has been filed with the Driven CO2 Conversion with a Solar-to-Formate Conversion
lead authors host institution protecting the cathode structure, Efficiency of 7.2%. Joule 2021, 5 (3), 687−705.
the electrolyzer design, and the hydrophilic spacer contained in (12) Hoof, L.; Thissen, N.; Pellumbi, K.; junge Puring, K.;
Siegmund, D.; Mechler, A. K.; Apfel, U.-P. Hidden Parameters for
the membrane electrode assembly. J.P.E., C.P.O., C.M.G.,
Electrochemical Carbon Dioxide Reduction in Zero-Gap Electro-
E.H.S., and D.S. have a financial interest in CERT Systems lyzers. Cell Reports Physical Science 2022, 3 (4), No. 100825.
Inc., a company commercializing CO2 electrolysis technology. (13) Kaur, A.; Kim, B.; Dinsdale, R.; Guwy, A.; Yu, E.; Premier, G.
J.P.E., C.P.O., C.M.G., and D.Y. are employees of CERT Challenges in Scale-up of Electrochemical CO2 Reduction to Formate
Systems Inc. Integrated with Product Extraction Using Electrodialysis. J. Chem.
Technol. Biotechnol. 2021, 96 (9), 2461−2471.

■ ACKNOWLEDGMENTS
The authors acknowledge financial support from the Natural
(14) Jung, B.; Park, S.; Lim, C.; Lee, W. H.; Lim, Y.; Na, J.; Lee, C.-
J.; Oh, H.-S.; Lee, U. Design Methodology for Mass Transfer-
Enhanced Large-Scale Electrochemical Reactor for CO2 Reduction.
Sciences and Engineering Research Council of Canada Chemical Engineering Journal 2021, 424, No. 130265.
(NSERC), TotalEnergies SE [TotalEnergies Research & (15) Jeanty, P.; Scherer, C.; Magori, E.; Wiesner-Fleischer, K.;
Technology Feluy (an affiliate of TotalEnergies SE, France)], Hinrichsen, O.; Fleischer, M. Upscaling and Continuous Operation of
the University of Toronto, the Ontario Centre of Innovation Electrochemical CO2 to CO Conversion in Aqueous Solutions on
Silver Gas Diffusion Electrodes. Journal of CO2 Utilization 2018, 24,
(OCI), and the Natural Resources Canada Clean Growth
454−462.
Program. D.S. gratefully acknowledges support from the (16) Hawks, S. A.; Ehlinger, V. M.; Moore, T.; Duoss, E. B.; Beck, V.
Canada Research Chairs Program. The authors thank the A.; Weber, A. Z.; Baker, S. E. Analyzing Production Rate and Carbon
XPRIZE Foundation, NRG COSIA, and the Alberta Carbon Utilization Trade-Offs in CO2RR Electrolyzers. ACS Energy Lett.
Conversion Technology Center (ACCTC) for their support of 2022, 7 (8), 2685−2693.
carbon utilization technologies. (17) Legrand, U.; Lee, J. K.; Bazylak, A.; Tavares, J. R. Product
Crossflow through a Porous Gas Diffusion Layer in a CO 2

■ REFERENCES
(1) Masel, R. I.; Liu, Z.; Yang, H.; Kaczur, J. J.; Carrillo, D.; Ren, S.;
Electrochemical Cell with Pressure Drop Calculations. Ind. Eng.
Chem. Res. 2021, 60 (19), 7187−7196.
(18) Baumgartner, L. M.; Koopman, C. I.; Forner-Cuenca, A.;
Salvatore, D.; Berlinguette, C. P. An Industrial Perspective on Vermaas, D. A. Narrow Pressure Stability Window of Gas Diffusion
Catalysts for Low-Temperature CO2 Electrolysis. Nat. Nanotechnol. Electrodes Limits the Scale-Up of CO2 Electrolyzers. ACS Sustainable
2021, 16 (2), 118−128. Chem. Eng. 2022, 10 (14), 4683−4693.
(2) Jordaan, S. M.; Wang, C. Electrocatalytic Conversion of Carbon (19) Subramanian, S.; Yang, K.; Li, M.; Sassenburg, M.; Abdinejad,
Dioxide for the Paris Goals. Nat. Catal 2021, 4 (11), 915−920. M.; Irtem, E.; Middelkoop, J.; Burdyny, T. Geometric Catalyst

2582 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Utilization in Zero-Gap CO2 Electrolyzers. ACS Energy Lett. 2023, 8 (35) Yang, Z.; Li, D.; Xing, L.; Xiang, H.; Xuan, J.; Cheng, S.; Yu, E.
(1), 222−229. H.; Yang, A. Modeling and Upscaling Analysis of Gas Diffusion
(20) Iglesias van Montfort, H.-P.; Burdyny, T. Mapping Spatial and Electrode-Based Electrochemical Carbon Dioxide Reduction Systems.
Temporal Electrochemical Activity of Water and CO2 Electrolysis on ACS Sustainable Chem. Eng. 2021, 9 (1), 351−361.
Gas-Diffusion Electrodes Using Infrared Thermography. ACS Energy (36) Brée, L. C.; Wessling, M.; Mitsos, A. Modular Modeling of
Lett. 2022, 7 (8), 2410−2419. Electrochemical Reactors: Comparison of CO2-Electolyzers. Comput.
(21) Subramanian, S.; Middelkoop, J.; Burdyny, T. Spatial Reactant Chem. Eng. 2020, 139, No. 106890.
Distribution in CO2 Electrolysis: Balancing CO2 Utilization and (37) Fang, Y.; Cui, H.; Chen, B.; Chen, W.; Ruan, X.; Wu, X.; Cui,
Faradaic Efficiency. Sustainable Energy Fuels 2021, 5 (23), 6040− F.; Guo, M.; He, G. A Kinetics-Mass Transport Model for CO2
6048. Electroreduction Reactor to Investigate Performance Limitation
(22) Hoffmann, H.; Paulisch, M. C.; Gebhard, M.; Osiewacz, J.; Factors. Chem. Eng. Sci. 2022, 262, No. 117996.
Kutter, M.; Hilger, A.; Arlt, T.; Kardjilov, N.; Ellendorff, B.; (38) Sánchez, M.; Amores, E.; Abad, D.; Rodríguez, L.; Clemente-
Beckmann, F.; Markötter, H.; Luik, M.; Turek, T.; Manke, I.; Roth, Jul, C. Aspen Plus Model of an Alkaline Electrolysis System for
C. Development of a Modular Operando Cell for X-Ray Imaging of Hydrogen Production. Int. J. Hydrogen Energy 2020, 45 (7), 3916−
Strongly Absorbing Silver-Based Gas Diffusion Electrodes. J. Electro- 3929.
chem. Soc. 2022, 169 (4), No. 044508. (39) Sánchez, M.; Amores, E.; Rodríguez, L.; Clemente-Jul, C. Semi-
(23) Nitopi, S.; Bertheussen, E.; Scott, S. B.; Liu, X.; Engstfeld, A. K.; Empirical Model and Experimental Validation for the Performance
Horch, S.; Seger, B.; Stephens, I. E. L.; Chan, K.; Hahn, C.; Nørskov, Evaluation of a 15 KW Alkaline Water Electrolyzer. Int. J. Hydrogen
J. K.; Jaramillo, T. F.; Chorkendorff, I. Progress and Perspectives of Energy 2018, 43 (45), 20332−20345.
Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte. (40) de Groot, M. T.; Kraakman, J.; Garcia Barros, R. L. Optimal
Chem. Rev. 2019, 119 (12), 7610−7672. Operating Parameters for Advanced Alkaline Water Electrolysis. Int. J.
(24) Jeng, E.; Qi, Z.; Kashi, A. R.; Hunegnaw, S.; Huo, Z.; Miller, J. Hydrogen Energy 2022, 47 (82), 34773−34783.
S.; Bayu Aji, L. B.; Ko, B. H.; Shin, H.; Ma, S.; Kuhl, K. P.; Jiao, F.; (41) Sartory, M.; Wallnöfer-Ogris, E.; Salman, P.; Fellinger, T.; Justl,
Biener, J. Scalable Gas Diffusion Electrode Fabrication for Electro- M.; Trattner, A.; Klell, M. Theoretical and Experimental Analysis of
chemical CO2 Reduction Using Physical Vapor Deposition Methods. an Asymmetric High Pressure PEM Water Electrolyser up to 155 bar.
ACS Appl. Mater. Interfaces 2022, 14 (6), 7731−7740. Int. J. Hydrogen Energy 2017, 42 (52), 30493−30508.
(25) Qi, Z.; Biener, M. M.; Kashi, A. R.; Hunegnaw, S.; Leung, A.; (42) Espinosa-López, M.; Darras, C.; Poggi, P.; Glises, R.; Baucour,
Ma, S.; Huo, Z.; Kuhl, K. P.; Biener, J. Scalable Fabrication of High P.; Rakotondrainibe, A.; Besse, S.; Serre-Combe, P. Modelling and
Activity Nanoporous Copper Powders for Electrochemical CO2 Experimental Validation of a 46 KW PEM High Pressure Water
Reduction via Ball Milling and Dealloying. Journal of CO2 Utilization Electrolyzer. Renewable Energy 2018, 119, 160−173.
2021, 45, No. 101454. (43) Ojong, E. T.; Kwan, J. T. H.; Nouri-Khorasani, A.;
(26) Lee, W. H.; Lim, C. W.; Lee, S. Y.; Chae, K. H.; Choi, C. H.; Bonakdarpour, A.; Wilkinson, D. P.; Smolinka, T. Development of
Lee, U.; Min, B. K.; Hwang, Y. J.; Oh, H.-S. Highly Selective and an Experimentally Validated Semi-Empirical Fully-Coupled Perform-
Stackable Electrode Design for Gaseous CO2 Electroreduction to ance Model of a PEM Electrolysis Cell with a 3-D Structured Porous
Ethylene in a Zero-Gap Configuration. Nano Energy 2021, 84, Transport Layer. Int. J. Hydrogen Energy 2017, 42 (41), 25831−
No. 105859. 25847.
(27) Smith, W. A.; Burdyny, T.; Vermaas, D. A.; Geerlings, H. (44) Andronie, A.; Stamatin, I.; Girleanu, V.; Ionescu, V.; Buzbuchi,
Pathways to Industrial-Scale Fuel Out of Thin Air from CO2 N. Simplified Mathematical Model for Polarization Curve Validation
Electrolysis. Joule 2019, 3 (8), 1822−1834. and Experimental Performance Evaluation of a PEM Fuel Cell
(28) Greenblatt, J. B.; Miller, D. J.; Ager, J. W.; Houle, F. A.; Sharp, System. Procedia Manufacturing 2019, 32, 810−819.
I. D. The Technical and Energetic Challenges of Separating (45) Wishart, J.; Dong, Z.; Secanell, M. Optimization of a PEM Fuel
(Photo)Electrochemical Carbon Dioxide Reduction Products. Joule Cell System Based on Empirical Data and a Generalized Electro-
2018, 2 (3), 381−420. chemical Semi-Empirical Model. J. Power Sources 2006, 161 (2),
(29) Alerte, T.; Edwards, J. P.; Gabardo, C. M.; O’Brien, C. P.; 1041−1055.
Gaona, A.; Wicks, J.; Obradović, A.; Sarkar, A.; Jaffer, S. A.; MacLean, (46) Barelli, L.; Bidini, G.; Gallorini, F.; Ottaviano, A. Analysis of the
H. L.; Sinton, D.; Sargent, E. H. Downstream of the CO2 Electrolyzer: Operating Conditions Influence on PEM Fuel Cell Performances by
Assessing the Energy Intensity of Product Separation. ACS Energy Means of a Novel Semi-Empirical Model. Int. J. Hydrogen Energy
Lett. 2021, 6 (12), 4405−4412. 2011, 36 (16), 10434−10442.
(30) van Bavel, S.; Verma, S.; Negro, E.; Bracht, M. Integrating CO2 (47) Larrazabal, G. O.; Strøm-Hansen, P.; Heli, J. P.; Zeiter, K.;
Electrolysis into the Gas-to-Liquids−Power-to-Liquids Process. ACS Therkildsen, K. T.; Chorkendorff, I.; Seger, B. Analysis of Mass Flows
Energy Lett. 2020, 5 (8), 2597−2601. and Membrane Cross-over in CO2 Reduction at High Current
(31) Heßelmann, M.; Bräsel, B. C.; Keller, R. G.; Wessling, M. Densities in an MEA-Type Electrolyzer. ACS Appl. Mater. Interfaces
Simulation-based Guidance for Improving CO2 Reduction on Silver 2019, 11 (44), 41281−41288.
Gas Diffusion Electrodes. Electrochemical Science Adv. 2023, 3 (1), (48) O’Brien, C. P.; Miao, R. K.; Liu, S.; Xu, Y.; Lee, G.; Robb, A.;
No. 2100160. Huang, J. E.; Xie, K.; Bertens, K.; Gabardo, C. M.; Edwards, J. P.;
(32) McCallum, C.; Gabardo, C. M.; O’Brien, C. P.; Edwards, J. P.; Dinh, C.-T.; Sargent, E. H.; Sinton, D. Single Pass CO2 Conversion
Wicks, J.; Xu, Y.; Sargent, E. H.; Sinton, D. Reducing the Crossover of Exceeding 85% in the Electrosynthesis of Multicarbon Products via
Carbonate and Liquid Products during Carbon Dioxide Electro- Local CO2 Regeneration. ACS Energy Lett. 2021, 6 (8), 2952−2959.
reduction. Cell Reports Physical Science 2021, 2 (8), No. 100522. (49) Xu, Y.; Miao, R. K.; Edwards, J. P.; Liu, S.; O’Brien, C. P.;
(33) Burdyny, T.; Graham, P. J.; Pang, Y.; Dinh, C.-T.; Liu, M.; Gabardo, C. M.; Fan, M.; Huang, J. E.; Robb, A.; Sargent, E. H.;
Sargent, E. H.; Sinton, D. Nanomorphology-Enhanced Gas-Evolution Sinton, D. A Microchanneled Solid Electrolyte for Carbon-Efficient
Intensifies CO2 Reduction Electrochemistry. ACS Sustainable Chem. CO2 Electrolysis. Joule 2022, 6 (6), 1333−1343.
Eng. 2017, 5 (5), 4031−4040. (50) Weng, L.-C.; Bell, A. T.; Weber, A. Z. Towards Membrane-
(34) Kas, R.; Star, A. G.; Yang, K.; Van Cleve, T.; Neyerlin, K. C.; Electrode Assembly Systems for CO2 Reduction: A Modeling Study.
Smith, W. A. Along the Channel Gradients Impact on the Energy Environ. Sci. 2019, 12 (6), 1950−1968.
Spatioactivity of Gas Diffusion Electrodes at High Conversions (51) Miao, R. K.; Xu, Y.; Ozden, A.; Robb, A.; O’Brien, C. P.;
during CO2 Electroreduction. ACS Sustainable Chem. Eng. 2021, 9 Gabardo, C. M.; Lee, G.; Edwards, J. P.; Huang, J. E.; Fan, M.; Wang,
(3), 1286−1296. X.; Liu, S.; Yan, Y.; Sargent, E. H.; Sinton, D. Electroosmotic Flow

2583 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Steers Neutral Products and Enables Concentrated Ethanol Electro- (69) Weng, L.-C.; Bell, A. T.; Weber, A. Z. A Systematic Analysis of
production from CO2. Joule 2021, 5 (10), 2742−2753. Cu-Based Membrane-Electrode Assemblies for CO2 Reduction
(52) Robb, A.; Ozden, A.; Miao, R. K.; O’Brien, C. P.; Xu, Y.; through Multiphysics Simulation. Energy Environ. Sci. 2020, 13
Gabardo, C. M.; Wang, X.; Zhao, N.; García de Arquer, F. P.; Sargent, (10), 3592−3606.
E. H.; Sinton, D. Concentrated Ethanol Electrosynthesis from CO2 (70) Gerhardt, M. R.; Pant, L. M.; Bui, J. C.; Crothers, A. R.;
via a Porous Hydrophobic Adlayer. ACS Appl. Mater. Interfaces 2022, Ehlinger, V. M.; Fornaciari, J. C.; Liu, J.; Weber, A. Z. Method�
14 (3), 4155−4162. Practices and Pitfalls in Voltage Breakdown Analysis of Electro-
(53) Hasa, B.; Cherniack, L.; Xia, R.; Tian, D.; Ko, B. H.; Overa, S.; chemical Energy-Conversion Systems. J. Electrochem. Soc. 2021, 168
Dimitrakellis, P.; Bae, C.; Jiao, F. Benchmarking Anion-Exchange (7), No. 074503.
Membranes for Electrocatalytic Carbon Monoxide Reduction. Chem. (71) Eriksson, B.; Asset, T.; Spanu, F.; Lecoeur, F.; Dupont, M.;
Catalysis 2023, 3 (1), 100450. Garces-Pineda, F.; Galan Mascaros, J. R.; Cavaliere, S.; Rozière, J.;
(54) Singh, M. R.; Clark, E. L.; Bell, A. T. Effects of Electrolyte, Jaouen, F. Mitigation of Carbon Crossover in CO2 Electrolysis by Use
Catalyst, and Membrane Composition and Operating Conditions on of Bipolar Membranes. J. Electrochem. Soc. 2022, 169 (3), No. 034508.
the Performance of Solar-Driven Electrochemical Reduction of (72) Yang, K.; Li, M.; Subramanian, S.; Blommaert, M. A.; Smith, W.
Carbon Dioxide. Phys. Chem. Chem. Phys. 2015, 17 (29), 18924− A.; Burdyny, T. Cation-Driven Increases of CO2 Utilization in a
18936. Bipolar Membrane Electrode Assembly for CO2 Electrolysis. ACS
(55) Lu, X.; Zhu, C.; Wu, Z.; Xuan, J.; Francisco, J. S.; Wang, H. In Energy Lett. 2021, 6 (12), 4291−4298.
Situ Observation of the PH Gradient near the Gas Diffusion Electrode
of CO2 Reduction in Alkaline Electrolyte. J. Am. Chem. Soc. 2020, 142
(36), 15438−15444.
(56) Jouny, M.; Luc, W.; Jiao, F. High-Rate Electroreduction of
Carbon Monoxide to Multi-Carbon Products. Nature Catalysis 2018,
1 (10), 748−755.
(57) Njodzefon, J.-C.; Klotz, D.; Kromp, A.; Weber, A.; Ivers-Tiffée,
E. Electrochemical Modeling of the Current-Voltage Characteristics
of an SOFC in Fuel Cell and Electrolyzer Operation Modes. J.
Electrochem. Soc. 2013, 160 (4), F313−F323.
(58) Liu, Z.; Sajjad, S. D.; Gao, Y.; Yang, H.; Kaczur, J. J.; Masel, R.
I. The Effect of Membrane on an Alkaline Water Electrolyzer. Int. J.
Hydrogen Energy 2017, 42 (50), 29661−29665.
(59) Liu, Z.; Yang, H.; Kutz, R.; Masel, R. I. CO2 Electrolysis to CO
and O2 at High Selectivity, Stability and Efficiency Using Sustainion
Membranes. J. Electrochem. Soc. 2018, 165 (15), J3371−J3377.
(60) Endrő di, B.; Kecsenovity, E.; Samu, A.; Halmágyi, T.; Rojas-
Carbonell, S.; Wang, L.; Yan, Y.; Janáky, C. High Carbonate Ion
Conductance of a Robust PiperION Membrane Allows Industrial
Current Density and Conversion in a Zero-Gap Carbon Dioxide Recommended by ACS
Electrolyzer Cell. Energy Environ. Sci. 2020, 13 (11), 4098−4105.
(61) Mardle, P.; Cassegrain, S.; Habibzadeh, F.; Shi, Z.; Holdcroft, S. Techno-economic Analysis and Carbon Footprint Accounting
Carbonate Ion Crossover in Zero-Gap, KOH Anolyte CO 2 for Industrial CO2 Electrolysis Systems
Electrolysis. J. Phys. Chem. C 2021, 125 (46), 25446−25454.
(62) Shafaque, H. W.; Lee, J. K.; Krause, K.; Lee, C.; Fahy, K. F.; Tianqi Gao, Jingjing Duan, et al.
Shrestha, P.; Balakrishnan, M.; Bazylak, A. Temperature Enhances the JULY 10, 2023
ENERGY & FUELS READ
Ohmic and Mass Transport Behaviour in Membrane Electrode
Assembly Carbon Dioxide Electrolyzers. Energy Conversion and
Management 2021, 243, No. 114302. An Advanced Guide to Assembly and Operation of CO2
(63) García-Nieto, D.; Barragán, V. M. A Comparative Study of the Electrolyzers
Electro-Osmotic Behavior of Cation and Anion Exchange Membranes Hugo-Pieter Iglesias van Montfort, Thomas Burdyny, et al.
in Alcohol-Water Media. Electrochim. Acta 2015, 154, 166−176. SEPTEMBER 13, 2023
(64) Laín, L.; Barragán, V. M. Swelling Properties of Alkali-Metal ACS ENERGY LETTERS READ
Doped Polymeric Anion-Exchange Membranes in Alcohol Media for
Application in Fuel Cells. Int. J. Hydrogen Energy 2016, 41 (32), Energy, Cost, and Environmental Assessments of Methanol
14160−14170. Production via Electrochemical Reduction of CO2 from
(65) Kutz, R. B.; Chen, Q.; Yang, H.; Sajjad, S. D.; Liu, Z.; Masel, I. Biosyngas
R. Sustainion Imidazolium-Functionalized Polymers for Carbon
Dioxide Electrolysis. Energy Technology 2017, 5 (6), 929−936. Fangfang Li, Xiaoyan Ji, et al.
FEBRUARY 07, 2023
(66) Reyes, A.; Jansonius, R. P.; Mowbray, B. A. W.; Cao, Y.;
ACS SUSTAINABLE CHEMISTRY & ENGINEERING READ
Wheeler, D. G.; Chau, J.; Dvorak, D. J.; Berlinguette, C. P. Managing
Hydration at the Cathode Enables Efficient CO2 Electrolysis at
Commercially Relevant Current Densities. ACS Energy Letters 2020, 5 Toward a Stackable CO2-to-CO Electrolyzer Cell
(5), 1612−1618. Design─Impact of Media Flow Optimization
(67) Wheeler, D. G.; Mowbray, B. A. W.; Reyes, A.; Habibzadeh, F.; Maximilian Quentmeier, Rüdiger-A. Eichel, et al.
He, J.; Berlinguette, C. P. Quantification of Water Transport in a CO2 JANUARY 04, 2023
Electrolyzer. Energy Environ. Sci. 2020, 13 (12), 5126−5134. ACS SUSTAINABLE CHEMISTRY & ENGINEERING READ
(68) Shafaque, H. W.; Lee, C.; Fahy, K. F.; Lee, J. K.; LaManna, J.
M.; Baltic, E.; Hussey, D. S.; Jacobson, D. L.; Bazylak, A. Boosting
Membrane Hydration for High Current Densities in Membrane
Get More Suggestions >
Electrode Assembly CO2 Electrolysis. ACS Appl. Mater. Interfaces
2020, 12 (49), 54585−54595.

2584 https://doi.org/10.1021/acsenergylett.3c00620
ACS Energy Lett. 2023, 8, 2576−2584

You might also like