You are on page 1of 23

Hindawi

International Journal of Energy Research


Volume 2023, Article ID 3183108, 23 pages
https://doi.org/10.1155/2023/3183108

Review Article
A Review of Accelerated Stress Tests for Enhancing MEA
Durability in PEM Water Electrolysis Cells

Eveline Kuhnert , Viktor Hacker , and Merit Bodner


Institute of Chemical Engineering and Environmental Technology, Graz University of Technology, Inffeldgasse 25/C, 8010,
Graz, Austria

Correspondence should be addressed to Eveline Kuhnert; eveline.kuhnert@tugraz.at

Received 26 September 2022; Revised 28 November 2022; Accepted 29 November 2022; Published 4 February 2023

Academic Editor: Palaniappan Subramanian

Copyright © 2023 Eveline Kuhnert et al. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is
properly cited.

During the past decades, a significant amount of excellent scientific results has been generated in the field of polymer electrolyte
membrane water electrolysis (PEMWE). Compared to current state-of-the-art technologies, PEMWE offers the opportunity to
produce green hydrogen with zero carbon emissions. However, the membrane electrode assembly (MEA), whose price is still
high for a rather limited lifetime, needs further improvement in terms of performance, cost, and durability. In order to
efficiently process novel materials, accelerated stress tests (ASTs) can be implemented to provoke and investigate cell ageing
processes and assess failure modes under real-life conditions. In this review, the different accelerated stressors of the main
components of the MEA are discussed, and recent publications of ASTs in the study of PEMWE cell durability are
summarized. Furthermore, a concise review of the degradation mechanisms for the individual MEA components depicted in
recent publications is presented. The different aspects identified in this review serve as a roadmap to further advance the
durability of novel stack materials.

1. Introduction time of 60.000 to 100.000 hours [3] and can achieve current
densities of >3 A cm–2, with commercial offerings at 1–
In the efforts to create a renewable and sustainable energy 3 A cm–2 [3–5]. Yet this relatively new and emerging tech-
supply chain, hydrogen plays a key role. It can be produced nology faces several challenges. Improved understanding of
from various feedstocks, stored, and efficiently transformed the performance, cost, and durability trade-offs under pre-
back into energy through fuel cells or be used in the chemi- dicted future dynamic operating modes using CO2-free elec-
cal industry. However, the current hydrogen production tricity is necessary to meet the hydrogen shot clean
(i.e., approximately 70 Mt/a) is limited compared to, for hydrogen cost target of $1/kg H2 by 2030 (and an interim
example, the annual global energy demand in the transport target of $2/kg H2 by 2025) [6]. Furthermore, new materials
sector alone, which amounts to roughly 1012 Mt/a hydrogen are slow to make their way into actual products, as qualifica-
equivalent (neglecting efficiency losses) in 2020. Further- tion and validation are time-consuming processes with little
more, the vast majority of hydrogen is produced from natu- harmonization in testing procedures [5]. The examination
ral gas and coal [1]. Combined with the increasing need to and understanding of performance losses will become even
efficiently store energy from renewable sources (e.g., wind, more critical with a shift to low-cost applications and inter-
solar, water, and tidal), electrolysis is capable of contributing mittent power sources [7, 8].
a valuable share to filling this gap. Among the available elec- Another well-known key issue of PEMWE is the scalabil-
trolysis technologies, polymer electrolyte membrane water ity from microscopic phenomena in laboratory cells to stack
electrolysis (PEMWE) is –up to date– the most capable of or system level [9]. Different measurement set-ups and cell
coping with the relatively dynamic energy supply pattern geometries influence the measurement results and their
renewable sources exhibit [2]. PEMWEs have a quoted life- reproducibility. A variety of MEA materials and preparation
2 International Journal of Energy Research

methods exist, that additionally influence their comparabil- in PEMWE and summarized the alleviating strategies.
ity. Therefore, it is important to introduce harmonized test- Khatib et al. [15] recently published a summary of the deg-
ing procedures to rate the comparability of different cell sizes radation of components in PEMWE and investigated the
assembled with similar materials. effects of degradation on the overall performance of the elec-
In PEM fuel cells (PEMFC), the reverse process of trolyzer. Shirvanian and Berkel [16] introduced a survey of
PEMWE, accelerated stress tests (ASTs) have experienced PEMWE status and reviewed state-of-the-art components
much attention, as they are capable of accelerating the devel- and limitations in performance and durability. In other
opment and increasing the understanding of lifetime- recent publications, the focus was set on specific compo-
limiting mechanisms [7]. The principle of an AST is to apply nents in the PEMWE, such as electrocatalysts [17, 18],
different degradation accelerating stressors, either via in situ membranes [19–21], PTLs [22–24], and BPPs [25]. Further-
testing or ex situ measurements, to project the durability of more, EU-harmonized protocols for testing of several low-
the whole cell or a desired component. In order to design temperature water electrolysis systems are summarized in
ASTs for PEMWE that are capable of portraying real-life rel- Joint Research Centre (JRC) reports by the Fuel Cells and
evant ageing, a fundamental understanding of dominating Hydrogen Joint Undertaking (FCH JU) [26–29]. A summary
degradation mechanisms as well as their dependency on of some articles related to the field of PEMWE degradation
stressors during operation is critical. Degradation mecha- is given in Table 1.
nisms in PEMWEs are primarily dominated by irreversible These works provide a general overview of potential deg-
efficiency losses. Reversible degradation has been reported radation mechanisms and testing procedures in PEMWE;
as well, although it is understood to a far lesser extent [8]. however, to the authors’ knowledge, no comprehensive
Problems related to the higher voltage in PEMWEs are per- reviews on AST strategies for PEMWE exist so far. For
formance degradation and material deterioration and bring PEMFCs, a technology facing similar challenges, various
a challenge for their durability. Prolonged use under certain AST protocols at component stack and system level have
operation conditions can lower the electrochemical stability already been reviewed [30–32]. In this review, aspects of
and accelerate the degradation of the materials. The most degradation and durability of the key components of the
critical components in the PEMWE cell are represented by PEMWE are brought in context with their individual accel-
the membrane electrode assembly (MEA). A typical MEA erating conditions, aiming to gather the information neces-
is composed of a polymer electrolyte membrane (PEM), sary for the qualification and evaluation of state-of-the-art
two catalyst layers (CLs), a carbon-based gas diffusion layer and novel materials for PEMWE. The accelerating factors
(GDL) on the cathode side, and a metal-based porous trans- responsible for PEMWE performance decay are summarized
port layer (PTL) on the anode side of the cell. Figure 1 shows for the membrane, catalysts, and diffusion media (GDL and
a schematic of the components in a PEMWE cell and the PTL), respectively. Finally, AST protocols for lifetime pre-
half-cell reactions at the anodic and cathodic compartment. diction based on the functional properties of the materials
At the anode, the oxygen evolution reaction (OER) takes and operational conditions are discussed for the individual
place where water is electrochemically split into oxygen components.
(O2), electrons (e–), and protons (H+). These protons diffuse
through a proton-conducting membrane to the cathode side, 2. Membrane Degradation and
where they are recombined with electrons to evolve as gas- Accelerated Stressors
eous hydrogen (H2). On the cathode, platinum (Pt) is pre-
dominantly used as catalyst, whereas on the anode-side Due to the simultaneous presence of hydrogen on one side
iridium (Ir) is typically used for the OER due to its relatively and pure oxygen on the other side, gas separation is critical
high activity and stability [10–13]. for the safe operation of an electrolyzer. The state-of-the-art
PEMs, such as those used in electrolysis cells, are typi- materials used to separate the gaseous reaction products
cally made from perfluorosulfonic acid (PFSA) polymers (oxygen and hydrogen) are PFSA-based polymer mem-
such as Nafion™ with a thickness of about 100 μm. Efforts branes (Nafion™). Nafion™ comprises of a PTFE backbone
are being made to develop cheaper and more durable mem- with sulfonic acid end-groups that provide protonic conduc-
branes with higher proton conductivity. However, as of now, tivity (Figure 2) [16]. Historically, Nafion™ membranes are
developments in one direction typically limit the perfor- identified with a three-digit number in which the first two
mance in other directions as compared to commercial mem- digits refer to the equivalent weight and the third digit refers
branes [14]. The PTLs act as current collectors and are to the membrane thickness in 25.4 μm-units (e.g., Nafion™
located between the current collecting flow field and the 115: equivalent weight = 1100 and thickness = 127 μm).
reactive sites in the CL. The GDL on the cathode side of Applying thick electrolytic membranes to PEMWE cells
the catalyst coated membrane (CCM) consists of a standard improve the physical and chemical durability and reduce
carbon cloth. On the anode side of the CCM, metallic mate- safety risks in the system [33]. However, the driving needs
rials like Ti-based meshes are used as PTLs due to the high to reduce the overall transport resistance results in increas-
potential of the OER. ingly thinner PFSA membranes which tend to accelerate
Extensive reviews of the durability and degradation gas crossover in PEMWE cells [20].
mechanisms of PEMWE cells and its components exist. Regarding long-term performance, the PFSA membrane
For instance, Feng et al. [8] presented a comprehensive is the weakest component in PEMWE systems, as perfor-
review on the degradation mechanisms of key components mance deterioration is mostly attributed to membrane
International Journal of Energy Research 3

Membrane Carbon GDL


PTL
BPP

Anode: H2O → 2H+ + ½ O2 + 2e−

O2 out H2 out Cathode: 2H+ + 2e− → H2

Overall reaction: 2H2O → H2 + ½ O2

H2O in

Pt-Catalyst
Ir-Catalyst

Figure 1: Schematic illustration of PEM water electrolysis (PEMWE).

Table 1: Highlighted publications related to PEM water electrolyzer degradation and standardized testing.

Technology Topic Article type Ref.


PEMWE Degradation of key components Review [8, 15, 16]
PEMWE Mitigation strategies Review [8]
PEMWE Catalyst properties Review [17, 18, 21]
PEMWE Membrane properties Review [19, 20]
PEMWE Porous transport layer (PTL) properties Review [22–24]
PEMWE Bipolar plate (BPP) properties Review [25]
PEMFC Durability test protocols Review [30, 31]
PEMWE, AWE, and AEMWE EU-harmonized test procedures for water electrolysis Report [27–29]
PEMWE, AWE, and AEMWE EU-harmonized protocols for electrolyzer testing Protocol [26]

(CF2 CF2)X (CF CF2) Table 2: Reactions involving free radicals in PEM membranes [36].
N
Rate
(OCF2 CF) O (CF2 CF2) SO3–H+ # Reaction
constant
CF3 1 HO• + Rf CF2 COOH ⟶ products <106 M-1 s-1
• −7
2 H2 O2 ⟶ 2HO 1:2 · 10 s-1
Figure 2: Chemical structure of Nafion™, x∗ 6.5.
3 HO• + H2 O2 ⟶ HOO• + H2 O 2:7 · 107 M-1 s-1
4 HOO• + H2 O2 ⟶ HO• + H2 O + O2 <1 M-1 s-1
pollution or chemical degradation [8]. Apart from Nafion™, • •
5 HO + H2 ⟶ H + H2 O 4:3 · 107 M-1 s-1
a few other membranes can be used in hydrogen produc-
• •
tion processes [34]. For instance, polybenzimidazoles 6 H + O2 ⟶ HOO 1:2 · 10 10
M-1 s-1
(PBI), sulfonated polyether ether ketones (SPEEK), polyox- 7 2HOO• ⟶ H2 O2 + O2 8:6 · 105 M-1 s-1
adiazole, polysulfone (PSF), or polyimides have recently
8 Fe2+ + H2 O2 + H+ ⟶ Fe3+ + HO• + H2 O 63 M-1 s-1
been considered as cheaper and more effective alternatives
in membrane-based electrolytic systems [34].

2.1. Membrane Degradation Mechanisms. In this section, the The reactions 1-8 that are listed in Table 2 show that the
supposed degradation mechanisms refer to PFSA mem- hydroxyl radical is the most aggressive intermediate as its
branes, unless otherwise stated. Apart from safety and oper- reaction with carboxylic end-groups in the main chain of
ating issues, one issue that affects the stability of the the PFSA ionomer can lead to a gradual “unzipping” of the
membrane and thus accelerates chemical degradation is the PFSA main chain (reaction 1) [37]. Primary sources of
crossover of gases. The crossover of product gases (hydrogen HO•, which leads to the chain scission, are the slow homol-
and oxygen) can result in the formation of hydrogen perox- ysis of H2O2 (reaction 2) and the Fenton reaction (reaction 8).
ide or radicals at the catalysts [16, 35]. An overview of the At high concentrations of H2O2, the “quenching of” HO•
reactions involving free radicals in PEM membranes as pre- becomes significant (reaction 3). The reaction of HOO• with
sented by Gubler et al. [36] is given in Table 2. H2O2 is thus a source of HO• at high concentrations of
4 International Journal of Energy Research

Table 3: Performance of different Nafion™ membranes in PEM technology [43, 44]. The thicknesses are listed at begin of testing (BoT) and
end of testing (EoT), and the degradation rate is given in mV h–1.

Thickness BoT Thickness EoT Degradation


Operating conditions Research findings Ref.
(μm) (μm) (mV h–1)
NR211 25.4 7.5 0.26
NR212 44 14 0.18 AST under idle conditions The performance of thin membranes
[43, 44]
N115 124 78 0.09 for 1000 h degrades much faster
N117 179 142 0.09

H2O2 (reaction 4). Hydrogen radicals are produced in the PEMWE [20]. Furthermore, enhanced operation conditions
reaction of H2 with HO• (reaction 5) that are further con- like high temperature [45–47], current density [40, 46], and
verted into hydroperoxyl radicals (HOO•) in reaction 6. In operation pressure [48–50] lead to an increase of the gas-
reaction 8, highly reactive hydroxyl radicals (HO•) are formed crossover rate. High-pressure operation results in significant
through the produced H2O2. This reaction is also known as stress on the membrane, and the differential pressure in par-
the so-called Fenton’s reaction and strongly catalyzed by ticular favors permeation. Therefore, the high-operation
Fe2+-ions [38]. If no ferrous iron-ion impurities are present, pressure that is targeted in PEMWE requires the use of
HOO• mainly decays via disproportionation (reaction 7) thicker materials, because high pressure driven by the pres-
[36]. The radical attack of the perfluorinated backbone and sure gradient increases the probability of H2 or O2 gas diffu-
perfluorosulfonic side chains of the PEM can cause degrada- sion across the membrane [14, 25, 42].
tion resulting in fluoride and sulphur release and thinning of Additional stress may originate from preexisting defects
the membrane [8]. The proposed reaction is as follows: and foreign objects introduced during manufacturing or
from components such as the PTLs and current collectors.
ðCF2 Þj COOH + 2OH• ⟶ ðCF2 Þj−1 COOH + CO2 + 2HF: Punctures, tears, and cracks are known to function as start-
ing points for further degradation, which inevitably result in
ð1Þ a reduced lifetime. In the presence of pure and usually pres-
surized oxygen, the thereby posed risk is, however, a multi-
In reaction 9, the carboxylic acid end-groups of the PEM tude higher than in, e.g., fuel cells, where typically ambient
membrane react with two hydroxyl radicals. As a result, two air is fed. Even though the combustion of hydrogen and oxy-
HF molecules are released and one CO2 and one CF2 molecule gen is in the presence of catalyst favored over explosion in
are formed [39]. Therefore, in durability tests, the fluoride-ion the case of crossover between the electrodes, severe destruc-
release rate (FRR) of the PEM electrolyzer exhaust can be seen tion of the materials can occur rapidly [51, 52]. Therefore, a
as indicator for the rate of chemical degradation of the sufficient water supply to the anode is necessary to prevent
membrane. hot spots and local drying of the membrane [51].
Another degradation mechanism that poses a critical
safety issue is the diffusion of evolved gases (H2 and O2) in 2.2. Accelerated Stressors for Membrane Degradation
water through the membrane to the opposite compartment
of the electrolyzer [40]. This phenomenon is driven by dif- 2.2.1. Hydration Conditions and Elevated Temperature.
ferences in the dissolved gas concentrations in the liquid Compared to PEMFC, where the membrane has to be
water phases across the membrane [19, 40, 41]. humidified by gases, the PEMWE membrane is in contact
The membrane thickness strongly correlates with the gas with liquid water during the whole electrolyzer operation
crossover and therefore affects the lifetime of the electroly- [14, 19]. Several researchers found, that drying at elevated
zer. Although very thin membranes offer a reduced trans- temperatures (>100°C) has a negative impact on the pro-
port resistance, they cannot guarantee safe operation due tonic conductivity and the water-uptake properties of PEM
to higher diffusion of H2 and O2 through the PFSA mem- membranes [53, 54]. The water-uptake properties are
brane. Therefore, relative thick membranes (120–200 μm) affected by the irreversible reorganization of the membrane
like Nafion™ 115 and Nafion™ 117 are used in PEMWE to structure, and any decrease in water-uptake results in the
avoid the crossover of gases [42]. An overview of the perfor- degradation of the proton conductivity [19]. During electro-
mance of different PFSA membranes in respect to their lyzer operation, the water not only functions as reactant but
thickness from comparable PEMFC studies [43, 44] is pre- also is used for cooling purposes. Consequently, water flow-
sented in Table 3. rate and temperature are important parameters that should
The major reason for the larger performance decay for not be neglected. These findings agree with a study by
thin membranes can be explained by higher permeation Selamet et al. [55], where the effect of water flowrate on
effects across the MEA. However, thicker membranes exhibit the performance of a 10-cell PEM electrolyzer stack
a higher thickness loss, as they have more material to (100 cm−2 active area MEAs) is investigated. The PEMWE
degrade [43]. Chemical degradation processes like mem- stack is operated in a range of flowrates from 750
brane thinning or pinhole formation can increase the overall to1500 mL/min at 1.35 A cm−2 for 2000 hours. According
permeation and subsequently reduce the efficiency of the to the authors, the flowrate has no major influence on the
International Journal of Energy Research 5

performance of the PEM electrolyzer. However, if the flow- decreases with the increase of gas permeability caused by
rate of water is too low, the cell temperature rises uncon- membrane thinning, crack, and pinhole development [58, 59].
trollably, which affects the efficiency of the PEMWE. Cationic contamination has been reported as an accel-
Therefore, the flow rate is generally chosen for cooling pur- erating factor for membrane degradation. While improp-
poses at an operating current density [55]. The researchers erly treated, feed water is the major cause of PEMWE
also tested the influence of other parameters during failures in this field [60, 61]. Weiß et al. [62] demon-
PEMWE operation and found that the operating tempera- strated the ion exchange of metal impurities into the
ture is the most important factor affecting the performance. membrane/ionomer phase during an OCV-AST. The cor-
At high operating temperatures, the activation barrier is rosion processes at low potentials resulted in the forma-
lowered, which results in enhanced membrane properties tion of metal cations which not only reduced the
like faster reaction kinetics, improved mass transport membrane/ionomer conductivity but also increased the
(MT), and higher conductivity [55]. high-frequency resistance (HFR) through the displacement
Temperatures above 90°C accelerate the rapid degrada- of protons in the ionomeric membrane [62].
tion of the membrane, and the increase of chemical degrada-
tion has reported to be directly proportional to the 2.2.3. Load Cycling and Constant Current Operation.
temperature increase [56]. According to Kusoglu and Weber Another lifetime-limiting factor for membrane ageing in
[20], the thermal decomposition of the PFSA ionomer PEMWE is the operating current. Chandesris et al. [35]
occurs in three stages: highlighted the impact of current density on the chemical
degradation rate in PEMWE membranes. They investigated
(1) Loss of (residual) water (from 100 to 200 ± 20°C), chemical membrane degradation in a 25 cm2 single cell set-
ending with breakup of H3O+ and protonation of up at various temperatures and current densities. According
SO3 to SO3H as well as partial decomposition of S to the researchers, at low-current densities, two competing
O3 H ⟶ SO2 + • OH close to 200°C HO• radical-consuming reactions take place: a first reaction
that requires hydrogen peroxide, and the second reaction
(2) Cleavage of the C–S bond (280 ± 30°C to 400 ± 20°C)
that accelerates the membrane attack. As the peroxide con-
resulting in sulfonate-group degradation ending with
centration is quite high at low current densities, the first
oxidation of the side-chain terminus, and
reaction becomes predominant, and the degradation rate is
(3) The final decomposition (oxidation) of the perfluori- reduced. High-current density operation results in a
nated matrix (400 ± 20°C to 600 ± 40°C) [20] decrease of the degradation rate, since the molar percentage
of oxygen at the cathode is reduced, and peroxide and free
Therefore, PEMWE operation at higher temperatures radicals are formed [35].
of above 90°C can limit the lifetime of the membrane sig- These trends in low-current density operation have
nificantly. According to Frensch et al. [57], a temperature also been confirmed by more recent research conducted
increase from 60 to 80°C results in a moderate increase by Li et al. [63]. The researchers designed AST protocols
of the degradation rate from 1.2 to 3.0 μV h−1, while comprising of dynamic load profiles of 0–0.5 A cm−2 and
operating at 90°C has a significantly more detrimental 1.2–2 A cm−2, respectively. The experiments revealed that
effect of 183.8 μV h−1. Furthermore, operation at elevated low-current cycling had a far greater influence on the volt-
temperatures significantly increases the FRR and mem- age degradation, whereas the ohmic resistance decreased
brane thinning over time. Chandesris et al. [35] devel- significantly during the high-load cycles, which could
oped a model that investigates the oxygen crossover and imply membrane thinning and failure.
temperature-induced membrane degradation in PEMWE. In a recent study, Fouda-Onana et al. [64] measured the
The authors revealed that the lifetime of an electrolyzer FRR as indicator for the membrane state-of-health. The
cell is significantly decreased at high temperatures. Only researchers performed in situ measurements of PEMWE
8.700 h operation at 353 K were necessary to thin 50% single cells that consisted of an alternating current profile
of the membrane [35]. between 0 and 1 A cm-2. A set of electrochemical measure-
ments was performed at the end of each current sequence
2.2.2. OCV. It has been reported that idle or open-circuit to measure the performance deterioration of the cell. The
voltage (OCV) conditions cause largely uniform degradation researchers additionally investigated the effect of tempera-
across the MEA active area in PEMFC and accordingly lead ture on the MEA. Therefore, the same ageing test (load
to near uniform thinning of the membrane [30]. Singh et al. cycling between 0 and 1 A cm–2) was carried out on a
[58] showed that in situ hold at OCV causes chemical degra- MEA at 60°C and at 80°C. The FRR shows an extremum
dation of PEMFC membranes. The results were verified via between 0.2 and 0.4 A cm–2. The extremum can be
infrared (IR) and scanning electron microscopy (SEM) explained by the mentioned competitive reaction between
images that show that the pure chemical membrane degra- radical (HO•) formations and the two reactions that con-
dation proceeds generally uniform across the whole active sume those radicals. In contrast to ageing at 60°C, ageing
region of MEA. Therefore, in situ hold at OCV is the at 80°C has a more severe impact on the membrane ageing
accepted means to accelerate membrane chemical degrada- (see Table 4). The FRR shows a 5-fold increase at 80°C
tion of PEM cells, while the change of OCV is a useful index compared to that one measured at 60°C (operation at
to evaluate the membrane health, because the OCV 1 A cm-2). These findings indicate that the temperature has
6 International Journal of Energy Research

Table 4: Fluoride release rate (FRR) in μg (F–)/h/cm2 from the face morphologies of Nafion™ membranes after exposure to
cathode water exhaust during ageing tests at 60°C and 80°C, Fenton’s reagent by two different test methods (solution and
adapted from [64]. exchange method).
A modelling framework for PEM degradation in elec-
0.2 A cm–2 0.4 A cm–2 1 A cm-2
°
trolysis mode was simulated by Frensch et al. [38]. The
Ageing at 60 C 0.5 0.5 0.3 researchers developed a Fenton model in which all major
Ageing at 80 C °
1.3∗ 2.2∗ 1.5∗ involved electrochemical reactions were visualized. Addi-

: mean values. tionally, experiments were performed that revealed a high
dependence of the FRR on the interaction of hydrogen per-
oxide and Fe impurities in the system. The iron impurities
much more severe effect on the membrane degradation were found to catalyze the reaction, while hydrogen peroxide
than the operating current [64]. is necessary for the formation of destructive hydroxyl radi-
Lettenmeier et al. [65] performed a study on the degra- cals [38]. Further studies showed that Fenton-like metal
dation mechanisms of a commercially available 8-cell catalysts can replace Fe2+ ions to produce OH• radicals
120 cm2 stack electrolyzer. The researchers designed a proto- [67, 68]. Laconti et al. [69] proposed an order for the rate
col of measurements comprising of three setpoints (T1–T3). of chemical degradation caused by common contaminants:
After an initial activation for 300 h, at nominal condition Fe2+ > > Cu2+ > > TiO2+ > Co2+ > Pt2+ > Ni2+, where Fe2+
setpoint with 2 A cm−2 was applied for more than 400 h has the highest impact on the degradation rate and Ni2+
(T1). Thereafter, the electrolyzer operated dynamically for the lowest. In state-of-the-art PEMWE cells, titanium is the
50 h between 2 A cm−2 and 0.15 A cm−2 (T2). At setpoint material of choice for anodic components such as BPPs
(T3), the electrolyzer was operated at 4 A cm−2 for 250 h. and PTL [38].
Operation at high current densities resulted in a reduction
of the ohmic drops, while longer operation gradually deacti- 2.3. Summary Membrane Degradation. The Nafion™ mem-
vated the anode. After the test sequences, the researchers brane represents a core component of the MEA. Membrane
performed ex situ measurements of the MEAs (SEM and failure is mainly accelerated by undesirable operation condi-
AFM) and water resin (XPS). The results indicate that the tions such as high temperature, OCV, and mechanical stress
loss of ionomer and catalyst material in the anode is depen- and results in changes of thickness (membrane thinning)
dent on the set current [65]. Interestingly, no lowering of the and the formation of defects (e.g., cracks or perforation).
performance of the PEM electrolyzer was observed over The deterioration of these parameters can be measured as
time, despite the degradation of the MEA components that a result of the FRR, voltage drop, or changes in the HFR in
was reported [65]. membrane specific failure analysis [59]. An overview of
membrane failure in lifetime tests is given in Table 5.
2.2.4. Fenton’s Test. Fenton’s reagent is commonly used to Other AST strategies than the experiments presented in
study the ex situ chemical degradation of membranes in Table 5 exist focus on the mechanical degradation of the
PEMFC research [30]. Although PEMWE is the reverse membrane. For instance, Borgardt et al. [70] tested the influ-
technology of PEMFC, the degradation mechanisms of the ence of the clamping pressure on the performance of differ-
membrane are the same, since the state-of-the-art mem- ent PEMWE cell designs. To minimize the negative impact
brane material used in both technologies is Nafion™. In gen- of membrane degradation on the cell performance, novel
eral, Fenton’s test is carried out in a solution of hydrogen approaches like the reinforcement or modification of mem-
peroxide (H2O2) with ferrous iron (Fe2+ ions) as a catalyst: branes exist. Although the reduction of membrane thickness
in PEMWE is an important route to achieve proper effi-
Fe2+ + H2 O2 + ⟶ Fe3+ + OH• + OH− : ð2Þ ciency at high-current density, thinner membranes can
bring the risk of an increase of gas crossover and therefore
It is commonly agreed that trace Fe2+ or other contami- compromise the mechanical stability of the MEA. Recently,
nants accelerate the membrane attack in the presence of Siracusano et al. [71] demonstrated a reinforced short-side-
H2O2 [38]. In practice, these metallic impurities come from chain PTFE-based membrane that enables stable electrolysis
the balance of plant [8]. The interaction of Fe2+ ions with the operation at 4 A cm−2. Still, at the present stage, novel mate-
PEMWE system is not well documented, i.e., the ions may rials to replace PFSA membranes need to undergo durability
cross the membrane, stay within it, or be flushed out at studies for the evaluation of their long-term performance.
one of the outlets. The radicals formed through the Fenton’s
reaction attack the membrane and produce further radicals, 3. Catalyst Degradation and
such as R-CF2 in the case of Nafion™ [38]. A systematic Accelerated Stressors
investigation of the degradation of Nafion™ membranes
using two variations of Fenton’s accelerated ageing experi- The harsh operation conditions, especially at the anodic
ments was conducted by Kundu et al. [66]. A solution side of the PEMWE, limit the choice of electrocatalysts to
method where Fe2+ ions and peroxide exist together in solu- a small range of precious metal compounds. For monome-
tion prior to the addition of Nafion™, and an exchange tallic catalysts during the OER, a link between the activity
method where Nafion™ in the Fe2+ form is exposed to and stability has been established in which the most active
hydrogen peroxide. Figure 3 shows a comparison of the sur- oxides (Au ≪ Pt < Ir < Ru ≪ Os) show the lowest stability
International Journal of Energy Research 7

100 𝜇m 100 𝜇m

(a) (b)

Figure 3: Surface morphology of Nafion™ membranes subjected to Fenton’s reagent ASTs via (a) solution method 100x magnification and
(b) exchange method 100x magnification. Reprinted from [66] with permission from Elsevier.

Table 5: Summary of ASTs for the evaluation of PEMWE membrane stability.

AST mode Operating conditions Research findings Ref.


Membrane weight loss over the testing time was over 20% of
Ex situ degradation Fenton’s reagent [66]
the original weight and accompanied by fluoride release
The fluoride emission is highly dependent on the iron
Ex situ degradation Fenton’s reagent [38]
concentration, which catalyzes the reaction
Feed water flowrate between 750 The water flowrate indirectly influences the performance because
Hydration conditions [55]
and 1500 sccm it is also used for cooling purposes in the system
Operation at 1 A cm−2 for 80.700 h Membrane chemical degradation (fluoride emission and membrane
Thermal stress [35]
at 353 K and for 380.500 h at 333 K thinning) is strongly affected by the operating temperature
Cycling between OCV and operating Cationic contamination of the membrane is a possible reason
OCV-AST [62]
potentials (up to 3 A cm−2) for the increase of the high-frequency resistance (HFR)
The temperature was found to have a more critical effect
Load cycling Current between 0 and 2 A cm−2 on the membrane degradation compared to the current [64]
density in PEMWE operation
Dynamic load of 0–0.5 A cm−2 and Dynamic conditions may lower the cell performance by
Load cycling [63]
1.2–2 A cm-2 for 350 h accelerated membrane thinning effects
Constant current of 2 A cm−2 for 400 h High current density operation has no immediate negative effect
Constant current [65]
and 4 A cm−2 for 250 h on the cell performance, although MEA components degrade
The current density has a minor effect on the degradation
Simulation of different constant current
High constant current −2 rate. It follows a complex mechanism, and a maximum [35]
densities up to 10 A cm
degradation rate is observed at quite low-current densities

(Au ≫ Pt > Ir > Ru ≫ Os) [61]. Therefore, the most popular 3.1. Catalyst Degradation Mechanisms. Catalyst degradation
materials that are currently used for the OER in commer- occurs on both anode and cathode catalysts, whereas the fast
cial anodes are Ir-based and Ru-based electrocatalysts kinetics of the HER on widely used Pt catalysts in PEMWE
[10–12, 72]. For the hydrogen-evolution reaction (HER) cathodes allow loading reduction down to 0.025 mg cm−2
that occurs on the cathode side of the PEMWE, carbon- and have a minor impact on the performance [73]. There-
supported platinum (Pt/C) is recognized as the state-of- fore, the main challenge remains the stability of the anode
the-art material [16]. The performance of the catalysts is catalyst. According to Spöri et al. [74], the degradation of
largely credited to the morphology, particle size, crystal the CLs during the harsh PEMWE operating conditions
structure, and substrate materials [5]. Various material includes catalyst dissolution, catalyst migration and agglom-
combinations have been tested, especially for the OER to eration, CL detachment, and support passivation [74]. Cata-
reduce the amount of catalyst loading and enhance the sta- lyst degradation can also be related to the instability of the
bility of the materials. In this review, the focus is set on cat- PFSA membrane and PTLs. During long-term electrolyzer
alysts that comprise of Pt-based metals as cathode and IrOx operation, the instability of the PFSA-membrane highly
as anode catalysts. affects the anode and cathode CLs. Grigoriev et al. [60]
8 International Journal of Energy Research

reported membrane thinning during PEMWE long-term Oxygen evolution Dissolution


operation as a result of interactions with oxidizing species
formed at the electrodes. The stability issue of the PTLs
½ O2 IrO2OH HIrO2 Ir(III)
has also been studied. According to Rakousky et al. [75],
the Ti in the PTL corrodes and causes contamination of
the anode CL, leading to a decrease in the anodic exchange Slow IrO2 (hydr.) Fast
current density. In addition, a low-conductive Ti-oxide layer
is formed between the anode and the PTL that increases the Figure 4: Possible pathways of iridium dissolution during the
ohmic resistance of the cell [75, 76]. Other mechanisms for oxygen evolution reaction (OER) proposed by Cherevko et al. [85].
catalyst degradation are related to mechanical damage (i.e.,
uneven contact pressure or improper membrane swelling).
For instance, bubble formation on the catalyst surface is or several unstable Ir(III) complexes which rapidly dissolve
reported to induce particle loss or layer detachment and dis- in an electrochemical process. The tentative mechanism, as
ruption of the catalyst-membrane interface [8]. reported by the authors, is represented in the following
chemical reactions:
3.1.1. Catalyst Agglomeration and Migration. Nanoparticles
(NPs) tend to agglomerate into bigger particles with respect 2IrO2 ðOHÞ + 2H2 O ⟶ 2IrðOHÞ3 + O2 ,
to time due to their high-specific surface energy [8, 15]. As a ð3Þ
result, the active catalyst area available for the electrochem- 2IrO2 ðOHÞ ⟶ 2HIrO2 :
ical reaction is reduced, resulting in a decrease of the
catalytic activity [77]. The interaction between catalyst The suggested mechanism for the OER is shown in
nanoparticles (NPs) is a common phenomenon in PEMFC, Figure 4. The whole pathway for the second reaction can
and there are several mechanisms that are believed to drive be rewritten as following:
the Pt surface area loss with time: (i) coalescence via crystal
migration, (ii) growth via modified Ostwald ripening, and HIrO2 ⟶ IrO2 + H+ + e− ,
(iii) reprecipitation at other nucleation sites [74, 78]. These
particle growth mechanisms may also be applied to IrO2 + H2 O ⟶ IrO2 ðOHÞ + H+ + e− , ð4Þ
PEMWE, because changes in the morphology of the CLs like 2IrO2 ðOHÞ ⟶ 2HIrO2 + O2 :
migration, agglomeration, coalescence, or detachment of
catalyst NPs have been reported at the cathode and anode
during the long-term operation of PEMWE cells [8, 15, Degradation of IrOx through the evolution of oxygen
79]. For instance, IrOx agglomeration is accelerated under from the lattice strongly depends on the structure and
harsh operation conditions and contributes to the drop of morphology of the electrocatalyst [86]. While crystalline,
mass activity toward the OER [79]. Paciok et al. [80] showed rutile-type iridium oxides are metallic conductors with a
that the migration of Pt particles is correlated to the applied high degree of stoichiometry; their amorphous counter-
overpotential at the cathode. In their study, XPS analysis parts are permeable to water and more vulnerable to cor-
revealed that with a more negative overpotential, the hydro- rosion in the OER potential region [86, 87]. Kasian et al.
gen coverage rate of the Pt particles increases, which results [83] found that lattice oxygen evolution may occur in both
in a bigger gap between the Pt particles and the carbon sup- materials; however, amorphous hydrous IrOx exhibits a
port. Thus, the weakened Van der Waals forces allow an higher number of defects, and the porosity and the pres-
increase of the mobility of Pt particles, which further results ence of -Ir(III)OOH- groups in the CL increase the prob-
in their agglomeration [80]. ability of the formation of molecular oxygen from the
Yu et al. [81] confirmed that catalysts undergo degra- lattice and the dissolution of catalyst [83].
dation through agglomeration during the long-term opera- 3.2. Accelerated Stressors for Catalyst Degradation
tion of PEMWE. Pt was profiled via EDX mapping, and
the results show the particle agglomeration and shape 3.2.1. Start-Up/Shut-Down Cycling and OCV. Factors such as
transformation from spherical to dendritic. Additionally, operational currents and the frequency of changes in load,
Claudel et al. [79] revealed that simulated PEMWE anode temperature, and pressure make real-world operation of
operating conditions result in a loss of OER activity due to PEMWEs challenging [88]. Many stressors are involved,
the growth of catalyst particles and oxidation of species and coupling effects make it difficult to distinguish between
along with the decrease in the number of individual NPs their impacts on material degradation [89]. According to
concomitantly [79]. recent studies, AST profiles for the investigation of PEMWE
catalyst durability frequently mimic operation from renew-
3.1.2. Dissolution of IrOx. Regardless of its high activity, dis- able energy, like wind, solar, or water sources. In this case,
solution of IrOx during the OER has been reported [82–84]. the fluctuating power supply results in PEMWE operation
The dissolution rate enhances with an increase of potential, between periods of hydrogen generation, when power is
and stable IrO2 starts to dilute at approximately 1.8 V [8]. available, and idle periods, where no electricity is supplied
Cherevko et al. [85] assume that dissolution processes at (OCV) [73]. For instance, Weiß et al. [62] evaluated the cat-
high potentials are a consequence of the formation of one alyst stability for such realistic operating conditions. The
International Journal of Energy Research 9

[88] reported open-circuit Pt dissolution during intermittent


3 operation of a PEMWE for the first time. The researchers
3 A.cm–2geo. found that the underlying mechanism of Pt dissolution at
i (A.cm–2geo.)

the cathode is similar to cathode degradation in PEMFCs,


2
where the passivation of catalyst surfaces gradually hinders
the Pt dissolution process. During a 90 h OCV, hold test
1 152 ng cm−2 or 0.005% of the total catalyst loading was lost
[88] in their study.
0.1 A.cm–2geo. 0 A.cm–2geo.
The effect of CL degradation at low catalyst loading and
0 dynamic operation was observed by Alia et al. [90]. The
OCV-AST researchers analyzed various test variables including the
Reference test stressor pattern/severity, the cycling frequency, and potential
(a)
ramping [90].
The dynamic operation leads to thinning of the Ir CLs
which significantly contributed to the overall durability loss
1.5 [90]. Life tests of an electrolysis cell that were carried out
by Grigoriev et al. [60] showed that current reversals can
occur during shut-down procedures in PEMWE. In these
Ecell (V)

1.0
shut-down periods, operation in FC mode is possible by
the consumption of the remaining electrolysis gases (H2
0.5 and O2). Consequently, Pt may be carried away from the
cathodic electrocatalytic layer into the membrane. The
0.0 decreased Pt concentration and the loss of cohesion of Pt
0 5 10 15 20 25 30 NPs result in a lowered activity of the cathodic CL. Besides
t (min) Pt, several other contaminants (Ti, Ir, Fe, Ni, and Si) from
feed water or, e.g., the corrosion of pipes that have lower
OCV-AST
mobility compared to hydrogen is deposited in the mem-
Reference test brane’s near-cathode region, leading to an increase of the
(b) electrolysis potential over the course of durability tests [60].
Figure 5: Accelerated stress test (AST) profile imitating an
3.2.2. Load Cycling and Constant Current Operation. The
intermittent power supply during PEMWE. Current profiles (a)
and associated potential profiles (b) at 80°C during the first cycle effect of constant current operation and load cycling on
of the OCV-test. Adapted from [62] (CC BY 4.0). catalyst degradation in PEMWE was studied by Rakousky
et al. [91]. The researchers investigated a PEMWE cell at
80°C and found that high-static current densities (up to
proposed test protocol is shown in Figure 5. According to 3 A cm−2) resulted in the largest performance losses.
the researchers, an AST that cycles between OCV and oper- Siracusano et al. [92] investigated PEMWE catalyst dissolu-
ating potential has a much severe effect on the degradation tion at high-current densities. Galvanostatic operation at
rate compared to experiments where no OCV periods are 1 A cm−2 leads to a quite low-degradation rate of 5 μV h−1,
applied. whereas operation at 3 A cm−2 caused a three-fold higher
The study also revealed that IrOx can easily be reduced degradation rate of 15 μV h−1 [92]. In particular, the anodic
to metallic Ir when IrOx-based MEAs are held at OCV. dissolution of IrOx has been reported at high current den-
Crossover H2 from the cathode side between ∼0 V and sities [14, 17, 30]. A case study involving anode dissolution
∼1.6 V during the OCV-AST reduces the surface of the IrOx has been performed by Grigoriev et al. [93]. In order to
catalyst at the anode to a state closer to less stable hydrous maintain a higher cell voltage and anode potential in their
IrOx [62]. Considering that hydrous IrOx is much more degradation study, Pt was used as the electrocatalyst at both
likely to corrode in the OER potential region compared to anode and cathode CL. The researchers found that a corro-
dry, crystalline IrOx, these results represent a critical degra- sion process in the CL is the first MEA degradation mech-
dation mechanism in the study of PEMWE [86]. anism followed by thinning of the Nafion™ membrane.
These findings are supported by a recent study con- After 100 h of activation the current density was set to
ducted by Rheinländer et al. [86]. The researchers have 1 A cm−2; afterwards, on/off galvanostatic cycles (between
tested events similar to operation interrupts in a PEMWE 0 and 1 A cm−2) for approximately 5500 h of operation were
cell at 80°C. The AST profile consisted of three steps: in applied. Additionally, SEM, TEM, and EDX analyses were
the first step, the cathode pressure was set to 5.5 bara and performed to detect the observed performance decay.
a current of 1 A cm−2 was applied. In the second step, the TEM pictures representing cross-sections of the anodic area
operation was interrupted by switching off the power supply, after the experiments are shown in Figure 6. The results
waterflow, and cell heating on the anode side. In the third indicate that the dissolution and precipitation of anode cat-
step, the cell was held at OCV for 2 h in a transient period. alyst particles have a major impact on to the overall degra-
After these steps, the cell was restarted [86]. Dodwell et al. dation of the cell performance [93].
10 International Journal of Energy Research

(a) (b)
−2
Figure 6: TEM micrographs of the cross-sectioned MEA 5500 h operation between 0 and 1 A cm . A 5 μm thick Pt layer is visible on (a). A
clear representation of Pt particles on the micrograph is presented in (b). Adapted from [93] with permission from Elsevier.

3.2.3. Potential Cycling. Potential cycling as an AST in experiments have a lower impact on the degradation in
PEMWE durability studies has been analyzed by Brightman single cell PEMWE measurements compared to the RDE
et al. [94]. The researchers used a reference electrode to measurements in their study. This might be due to the
study the catalyst degradation in an operating PEMWE. higher temperature of 80°C in CCM vs. ∼25°C in RDE
The results revealed that the cathode contributes more to experiments and higher catalyst loading in CCMs. Still,
changes in OCV than the anode during shut-down cycling. the researchers observed a higher degree of mass-activity
The researchers also demonstrated that potential cycling losses in transient compared to static CCM measurements.
has a more severe effect on the catalyst durability than cur- Additionally, the results imply that static constant current
rent cycling. The Pt surface area barely decreased during or constant potential stability measurements (CCM-CA)
the current cycling procedure, whereas a significant loss of are insufficient to test the catalyst stability of state-of-
the ECSA (electrochemically active surface area) was the-art PEMWE applications [95].
observed during potential cycling. Therefore, potential
cycling can be considered as a useful AST for the evaluation 3.2.4. Cationic Contamination. The accumulation of metallic
of catalyst stability in Pt-based cathode CLs in PEMWE [94]. cations like Fe3+ [96], Cu2+ and Al3+ [97], Ca2+ [98], and
In a recent study, Spöri et al. [95] presented a transient Na+ [99] in the MEA can range from being a contributing
ADT (accelerated degradation test) protocol for RDE tests factor to being the dominating mechanism in PEM electrol-
that consist of potential square-wave cycles between 0.05 ysis performance degradation. The foreign cationic impuri-
and 2 V. Additionally, the researchers ran static CA (chron- ties may originate from impurities in stack component
oamperometric) tests at the same potentials for comparison materials, impurities of feed water, and other sources [96,
reasons. A much higher electrode potential dependence of 100]. According to Cheng et al. [100], even trace amounts
catalyst dissolution was observed during the transient ADT of impurities present in either fuel or air streams of PEMFCs
protocols compared to the static CA protocols [95]. In order can have a severely poisoning effect on MEA components,
to verify the proposed ADTs for their applicability in including the anode and cathode CL [100]. This effect is
PEMWE systems, the researchers also ran PEMWE single anticipated to be identical, if not more severe in electrolyzer
cell tests. Therefore, CCMs were subjected to (i) 24 h poten- operation if impurities enter the stack with the feed water.
tiostatic operation at 1.75 V; (CCM-CA), (ii) 24 h power The distribution of dissolved contaminants is promoted by
cycling between 0 A cm−2 and a potential of 1.75 V (CCM- the over stoichiometric feed rate (λ > 4) of liquid water
PC), and (iii) 15.000 cycles between 0 and 1.75 V (CCM- [101] in PEMWE. Cationic contamination from Fe3+,
ADT) [95]. Figure 7 represents the ADT protocols (i)–(iii) Cu2+, and Al3+ has been reported as a result from degrada-
and the polarization curve during (iii) [95]. According to tion of materials and corrosion of pipes in the system [96,
Spöri et al. [95], transient fast potential cycling and static 97]. Li et al. [102] investigated the effect of long-term
International Journal of Energy Research 11

2.0 1.90

1.9 1.85
1.80
1.8
1.75
1.7 1.70
1.6 1.65
1.60
1.5
1.55
1.4 1.50
1.3 1.45
1.40
1.2
0 2 4 6 8 10 12 14 16 18 20 22 24
t (h)

Ecell

(a)
2.00 1.90
1.75 1.85
1.80
1.50
1.75
1.25 1.70
1.00 1.65
0.75 1.60
1.55
0.50
1.50
0.25 1.45
0.00 1.40
40 41 42 43 44 45 46
t (min)

Ecell

(b)
2.0 2.00

1.5 1.75

1.0 1.50

0.5 1.25

0.0 1.00

–0.5 0.75

–1.0 0.50

–1.5 0.25

–2.0 0.00
2535 2540 2545 2550 2555 2560
t (s)

Ecell

(c)

Figure 7: Continued.
12 International Journal of Energy Research

2.1
Standard loading CCM - NR-212 - 1.69 mglr cm–2
2.0 ADT 0.0 V - 1.75 V
Ecell: closed symbols
EiR: open symbols
1.9

1.8

Ecell (V)
1.7

1.6

1.5

1.4
0.0 0.5 1.0 1.5 2.0 2.5 3.0
jgeo. (A cm ) –2

2.5k 10.0k
5.0k 12.5k
7.5k 15.0k
(d)

Figure 7: Current voltage characteristics of the three ADT protocols: (a) CCM-CA, (b) CCM-PC, and (c) CCM-ADT with the
corresponding polarization curve (d). Reprinted from [95] (CC BY 4.0).

H+available H+ depleted region of CCL

½O2 + 2H+ + 2e− → H2O C + 2H2O → CO2 + 4H+ + 4e−

½ O2 + 2H+ + 2e− → H2O


Proton occupancy (𝛾H+)

Pt
Pt Ca2+
H+ H+ Ca2+
Pt Pt Pt
Pt Pt C
C Pt
H+ H+ Pt
Pt Pt C
Pt Pt
Ca2+ Pt Pt
Ca2+ Ca2+
+ + +
H H H
2+ Pt + Ca2+
Ca H
Pt Pt Pt
C
Pt Pt
C
Pt Pt
Pt Pt

PEM GDL →
0.0 Normalized distance from membrane 1.0

Figure 8: Hypothesized carbon corrosion process of the cathode CL in a PEMFC.

contamination with iron ions on the performance of a and associated performance breakdown can already occur
PEMWE single cell. The researchers observed a significant at low concentrations of 3 ppm Al3+ on the cathode side
performance decay in which the degradation rate reached [97]. Cathode CL thinning due to the accumulation of
to 128.9 μV h−1 after the 829 h contamination test [102]. In Ca2+ ions in PEMFC cells has been investigated by Banas
a more recent publication, Li et al. [97] investigated the effect et al. [103]. Results showed that the accumulation of Ca2+
of further cationic impurities (Fe3+, Cu2+, and Al3+ ions) in ions can lead to proton depletion in the cathode CL [97,
single PEMWE cells and found that all impurity ions have 103]. As a result, carbon corrosion can occur, which then
a severe impact on the cell performance. Cell poisoning accelerates cathode CL thinning (Figure 8) [97].
International Journal of Energy Research 13

As shown in Figure 8, foreign cations accumulate in the PEMWE, since the state-of-the-art Ir-based anode materials
cathode side due to the balance of migration and diffusion. are still affected by the harsh reaction environment [108].
Lack of proton access accelerates the carbon corrosion [97].
On the cathode CL, the same state-of-the-art materials are 4. PTL/GDL Degradation and
used in PEMWE and PEMFC cells. Therefore, the effect Accelerated Stressors
of cationic impurities on the performance of the PEMWE
cell is expected to be even more detrimental due to the In PEMFCs, usually carbon materials are used as current
higher potentials [97]. The contamination effect of Na+ collectors on anode and cathode side (GDL). Due to the
on PEMWE has been studied by Zhang et al. [104]. Differ- harsh conditions on the anode in PEMWEs, carbon mate-
ent poisoning modes (anode poisoning and cathode poi- rials are in most cases only used on the cathode, where
soning) were compared, and the results imply that Na+ hydrogen is formed. The carbon corrosion and consumption
poisoning has a more severe effect on the anodic than on due to the high potential of the oxygen electrode result in the
the cathodic cell components. In the experiments, the rapid deterioration of interfacial contact, which has a major
anode potential increased by about 0.160 V because protons impact on the performance and efficiency of the cell [109].
were replaced by Na+ ions at the anode and membrane Therefore, metallic materials like Ti-based meshes, foams,
with increasing cell voltage. At the cathode, the potential or sintered powders are usually used as PTLs at the anode
decreased with decreasing proton concentration [104]. side of the electrolyzer. In this section, the term “GDL” is
Generally, reactions (5) and (6) occur at anode and cathode used for the material on the cathode side, and the term
in PEMWE systems. However, if there are not enough pro- “PTL” is used for the material on the anode side of the elec-
tons for reaction (6), the reaction (7) takes place at the trolyzer. In a recent review, Doan et al. [22] summarized
cathode. As a result, the cathode potential declines signifi- some of the state-of-the-art PTL and GDL materials used
cantly due to the lack of protons. Additionally, Na+ ions in PEM technology (Table 7).
and OH− can enter the feed water, which increases the Currently, only a limited number of studies have focused
pH at the cathode compartment [104]. on the lifetime-limiting factors for the carbon-based GDL
and even less for the Ti-PTL on the anode side of the
Anode : H2 O ⟶ 2H+ + ½O2 + 2e−  φ0 = 1:229 V, PEMWE [30]. PTL/GDL degradation can generally be
divided into two groups: on the one hand, chemical degrada-
ð5Þ
tion that is mainly caused by corrosion and on the other
hand, mechanical degradation that in most cases originates
Cathode : 2H+ + 2e− ⟶ H2  φ0 = 0:000 V, ð6Þ from compression force, dissolution, and erosion by hydro-
thermal effects [8].
Cathode : 2H2 O + 2e− ⟶ H2 + 2OH−  φ0 = −0:828 V: 4.1. Gas Diffusion Layer Degradation Mechanisms. The GDL
ð7Þ is placed on the cathode side of the CCM between the cur-
rent collecting flow field and the reactive sites of the Pt cat-
3.3. Summary Catalyst Degradation. The main failure alyst. It provides mechanical support to the MEA and is
modes for catalyst degradation, primarily OCV, dynamic responsible for the electron and gas transport to and from
operation, load cycling, and cationic contamination can be the CL with minimum voltage, current, thermal, interfacial,
identified by various AST methods. Considering the fact and fluidic losses [89]. Mechanical stress that results from
that during the lifetime of an electrolyzer operation inter- the high compression of the cell components strongly affects
rupts can be expected to occur frequently, AST profiles that the physical properties of the GDL. Carbon fibers can break
represent realistic operation conditions are needed. In or even be displaced under compression, which can change
Table 6 a summary of lifetime tests for PEMWE catalysts, the structure morphology, and affect the current transport
including test protocols simulating an intermittent power and permeability of reactants [8, 122]. As a result, the
supply, are presented. obstruction of diffusion pathways can lead to ineffective
Coupling of an electrolyzer unit with fluctuating renew- gas transport or saturation of the pores, decreasing the over-
able energy sources results in intermittent operation with all performance of the cell [8]. Compared to the anode com-
idle periods when no power is available and hydrogen gener- partment, the cathodic overpotential (approximately 0.1 V
ation, when enough current is supplied [62]. This recurring SHE) is insufficient to cause rapid carbon corrosion of the
transition between reducing and oxidizing conditions signif- GDL. Although carbon-based GDLs are used in the cathode
icantly affects the performance of the PEMWE system [62]. side of the cell, further research is needed to gain a better
According to the currently understood decay mechanisms, understanding of the performance of these materials under
the sluggish kinetics and instability of the OER catalyst real-life conditions over time [14].
materials constitute an obstacle toward commercialization
of PEMWE technology [16]. Therefore, current research 4.1.1. Accelerated Stressors for Gas Diffusion Layer
focuses on lowering the Ir loading and improving the degra- Degradation. In PEMFC research, various AST protocols
dation resistance [105–107]. In particular, understanding the exist, targeting for instance the chemical or mechanical deg-
degradation mechanism on the catalyst surface during OER radation of the GDLs [123]. GDL degradation can occur
is a cornerstone in the development of novel catalysts for rapidly under accelerated conditions like immersion in
14

Table 6: Summary of ASTs for the evaluation of PEMWE catalyst stability.

AST mode Operating conditions Research findings Ref.


Cycling between OCV and operating potentials leads to Ir dissolution
OCV-AST Cycling between OCV and operating potentials (up to 3 A cm−2) [62]
and an increased contact resistance between the electrode and the PTL
Succession of on/off galvanostatic cycles (ranging between 0 A corrosion process in the CL is the first MEA degradation mechanism
OCV-AST [93]
and 1 A cm−2) followed by thinning of the Nafion™ membrane
5 cm−2 electrolyzer is switched from average duty point Operation transients during PEMWE can induce conversion of IrO2 in
OCV-AST [86]
(1 A cm−2) to idle mode (OCV) the anode CL into a more hydrous form
−2
1 h 1 A cm pre-OCV followed by OCV for 72 h and 1 h During operation, the Pt dissolution is constant, whereas at idle periods,
OCV hold test [88]
operation at 1 A cm−2 post-OCV the Pt dissolution depends on the potential and duration of OCV
4000 h life tests (6 h operation everyday (shut-down over night When the electrolysis cell is switched off, current reversal and operation
Shut-down procedures [60]
and for weekends and holidays) in FC mode can cause degradation of the cathode CL
Evaluation of various stressor profiles (potential hold, triangle/ Less severe performance losses are reported for wind and solar AST
Dynamic operation [90]
square wave, and Sawtooth up/down profiles) models compared to triangle and square-wave potential cycling
2 A cm−2 up to 1100 h, 2.5 A cm−2 up to 1450 h, and 3 A cm−2 The performance stability is independent on the operation mode—there
High and elevated current [91]
up to 1940 h is no difference between constant and dynamic current density operation
Cell operation at 3 A cm−2 shows three times increase of corrosion rate
Constant current Operation at 3 A cm−2 for 1000 h [92]
compared to operation at 1 A cm−2
1000 square-wave cycles with a 30 s hold a period between Potential cycling is a more severe test of catalyst durability than current
Potential cycling [94]
1.75 V and 0.3 V cycling
An enhanced electrode potential dependence of catalyst dissolution was
Transient operation Square-wave cycles between 0.05 V and 2 V [95]
observed in transient compared to static ADT protocols
+
Anode poisoning: 1000 mL 0.05 mol L−1 Na2SO4 to anode and Na contaminants in the anode compartment have more severe effects
Na+ contamination 1000 mL DI water to cathode; operation at 0.5 A cm−2 on the cell performance than the contamination of cathode feed in [104]
(cathode poisoning v. v.) PEMWE
2+ Soaking PEMFC CCMs in cationic solution of 0.9 mM of CaSO4 Contaminated CCM had significant and uniform CL thinning caused by
Ca contamination in PEMFC [103]
and 29.1 mM of H2SO4 carbon mass loss by a carbon oxidation mechanism
3+ Solution containing 1 ppm Fe3+ fed to a single cell for ~800 h at
Fe contamination The cell performance degrades severely with 1 ppm Fe3+ contamination [102]
0.5 A cm−2
Fe3+ ions have a higher degrading effect than Cu2+ ions. Al3+ ion
Contaminated feed solution containing 5 ppm of Fe2(SO4)3,
Fe3+, Cu2+, and Al3+ contaminations poisoning results in severe cathode CL degradation even at low [97]
CuSO4, or Al2(SO4)3 in DI water, respectively
concentration of 3 ppm
International Journal of Energy Research
International Journal of Energy Research 15

Table 7: Summary of some PTL/GDL materials used in PEM pressure experiments at 3.4 MPa showed a more severe effect
technology [22]. during the initial cycles [123, 129].
PTL/GDL material PEMWE PEMFC 4.2. Porous Transport Layer Degradation Mechanisms. The
Titanium material PTL, which is located between the anode CL and flow
Iridium coated PTL [110] field, is responsible for the transport of the reactant water
Titanium powder [111–113] and removal of the produced oxygen gas in the PEMWE
Titanium fiber [112] cell. Furthermore, it acts as a current collector to ensure
Titanium felt [114]
stable electrochemical performance [24]. Generally, mate-
rials like Ti powders, foams, films, or meshes are used as
Carbon material PTL in PEMWE cells [130]. The metal-based PTLs can
Carbon cloth [115, 116] undergo hydrogen embrittlement or surface passivation
Carbon fiber paper [115, 117] and are therefore often coated with precious metals such
Carbon felt [115] as platinum or gold with obvious effects on the capital
IrO2 deposited on carbon paper [118]
costs. The resistance for hydrogen embrittlement is given
by Lu and Srinivasan [131] and follows the following
Other metals material order Ti = Ta > Nb > Zr > graphite (with Ti and Ta being
Stainless steel [119] the most prone to hydrogen embrittlement and graphite
Copper foil [119–121] as the least) [131].
PTL properties like surface roughness have lately
received much attention because of the effect on the Ir-
based anode CL which is in close contact with the PTL mate-
H2O2 solution or subjection to cyclic or static compression rial [22]. The formation and growth of oxygen bubbles
tests. H2O2 is a strong oxidizing agent and can cause severe inside the PTL pores of the anode of a PEMWE cell can
degradation across most PEMWE components. Several stud- block active catalyst sites and decrease the effective ECSA.
ies have reported GDL performance loss after H2O2 immer- Approaches exist which aim to reduce the MT losses by
sion tests, and the MPL was found to be even more open to modification of the PTL morphology to decrease the number
attack from the oxidizing species [124, 125]. For instance, of coalescence host sites [132]. The choice of PTL material
Liu et al. [125] performed a carbon corrosion test in H2O2 influences the performance of PEMWE, for example, for sin-
solution and analyzed the effect of a MPL on commercial tered Ti powder, the morphology and surface properties can
carbon-based GDLs. For a pristine GDL with MPL at be the most important factors, while porosity is the most
2.5 A cm−2, the potential decreased by 59% and an increased important factor for thin/tunable Ti films (TT PTL) [24].
water content of up to 44% was observed [125].
The GDL is the component with the highest porosity in 4.2.1. Accelerated Stressors for Porous Transport Layer
the PEMWE, and its cellular structure leaves it vulnerable to Degradation. Borgardt et al. [111] investigated the effect of
structural changes and permanent degradation under pressure on the mechanical properties of Ti PTLs. In situ
mechanical stress [123, 126]. In the study of Chun et al. and ex situ durability tests were carried out on PTLs made
[127], the mechanical durability of different GDLs with from two differently shaped powders. Samples made of
commercial and home-made crack-free MPL has been tested spherical powder were in a porosity range of 10–33%, while
in a dummy cell without CL and only air supply to avoid samples of irregular-shaped powder were in the range of 10–
electrochemical degradation. The results imply that the 55%. According to the researchers, high porosity and thin
mechanical degradation during long-term operation can be thickness lower the physical stability of the PTL but exhibit
significantly reduced by the preparation of crack-free MPL high durability under high pressure. Furthermore, a porosity
[127]. Dotelli et al. [128] investigated the effect of compres- of at least 25% and 500 μm thickness enabled the PTL to
sion pressure on the GDL, when the assembly of GDLs and resist 50 bar differential pressure [111]. Pushkarev et al.
CCMs was compressed to 30% and 50% of its initial thick- [133] investigated MEA combinations of different PTLs in
ness. The higher compression ratio at 50% resulted in a PEMWE single cells. The PTL structure had only a minor
reduced cell performance. This effect was partially mitigated impact on the activation and ohmic overpotential. However,
when the operating temperature was raised to 80°C. The at current densities of 2 A cm−2 and higher, structure-
presence of the MPL in the carbon cloth resulted in a dependent MT losses were reduced due to the higher volume
reduced contact resistance and therefore proved extremely of void phase and flow pore size. According to the
beneficial for the operations, especially at high current den- researchers, PTLs with low porosity and small pore size
sity [128]. Radhakrishnan and Haridoss [129] performed could be responsible for MT limitations due to the worsened
five compression cycles on a Toray carbon paper GDL at contact between the flow field channels and the PTL. In
two different pressure setpoints (1.4 MPa and 3.4 MPa). Figure 9, two SEM images of PTLs with different bulk poros-
The cyclic compression experiments resulted in changes in ities tested in this study are shown [133].
the structure like cracks of fibers and a decreased porosity Passivation that has been reported on the anode side of
and hydrophobicity. Furthermore, a gradual loss of PTFE the PEMWE for Ti-based BPPs will also influence the Ti-
was observed in course of the experiments. The elevated PTL if not coated with a protective layer [89]. A study
16 International Journal of Energy Research

(a) (b)

Figure 9: Initial SEM images of a PTL with bulk porosity of 51.9% (a) and 25.4% (b). Reprinted from [133] with permission from Elsevier.

Table 8: Investigated operation modes in a PEM water electrolysis cheaper, more abundant materials like carbon fibers in short
degradation study. Fresh cell assemblies were used for each term experiments [24].
experiment, and a break-in of at least 500 h was performed. All Results from Young et al. [134] show that the stability of
cycling tests were performed at 80°C. From [57] with permission conventional carbon-based fuel cell GDLs is sufficient for
from Elsevier. cell-conditioning and break-in procedures as well as the ini-
tial characterization of PEMWE cells [134]. The perfor-
Operation mode (name) Test condition
mance of the carbon-based materials was tested in a 10 h
Constant current at 80 (cc) 2.0 A cm−2 cell-conditioning procedure followed by twelve polarization
Constant voltage at 60 (T60) 2.0 V curve measurements (2.5 h each). Allowable operation of
Constant voltage at 80 (T80) 2.0 V 40 h was reported that includes 15 h of operation between
Constant voltage at 90 (T90) 2.0 V 2 V and 3 V [134]. Becker et al. [135] designed PEMWE sin-
100 s current cycling (cyc100s) 0−2.0 A cm−2, 100 s dwell time gle cell tests for anode PTLs made out of carbon-coated
10 s current cycling (cyc100s) 0−2.0 A cm−2, 10 s dwell time
316 L stainless steel. The substituted material was tested in
situ for 30 days at 2 A cm−2 and potentials up to 1.2 V vs.
Solar PV profile (solar) 0−2.0 A cm−2, 60 s dwell time
RHE [135]. These results show that Ti-based PTLs are not
necessarily required for initial performance characterization
procedures in PEMWE.
performed by Rakousky et al. [75] shows a performance loss
at constant current density (2 A cm−2) during long-term 4.3. Summary PTL/GDL Degradation. Studies on the impact
operation of a PEMWE due to semiconducting TiOx that is of different PTL/GDL materials on PEMWE performance
formed on the anode Ti-PTL. Coating of the Ti-PTLs with are still under investigation, and the high cost of titanium
Pt achieved a substantially reduced degradation rate of only is still a limiting factor that makes the development of
12 μV h−1, compared to 194 μV h−1 when no coating was commercialized materials challenging. To date, no effective
applied [75]. characterization protocols exist for PEMWE, although
The passivation of Ti components in PEMWEs was also much attention has been paid to the design and screening
observed in lifetime studies performed by Frensch et al. [57]. of PEMWE materials more recently. In Table 9, some of
They investigated several MEAs with the same specifications the available studies performed on GDL durability are
for 500 h at different operation modes, including variations summarized.
in the operating temperature and different current cycling According to the currently understood decay mecha-
modes (Table 8). Titanium-sintered PTLs (1 mm, 30% nisms in GDLs, effects such as carbon corrosion and CL
porosity) were used on the cathode and anode sides, respec- interaction under compression lead to a performance decay
tively. In an attempt to separate the total ohmic resistance during PEMWE operation. Successful mitigation strategies
into its components, the researchers identified the passiv- so far have relied on scarce and expensive metal coatings
ation of the Ti-PTL and the accompanying positive effect that contribute to the already high capital cost. To date, only
of MEA adjustments as main PTL contributions [57]. a limited number of degradation protocols for the PTL on
Although the use of corrosion-resistant coatings for PTL the anodic side of the electrolyzer exist. Some of the available
materials is advantageous for long-term PEMWE operation, lifetime tests for PTL degradation are listed in Table 10.
it makes small scale research hardware expensive and diffi- Recently, great promise for PTL design optimization has
cult to source and is hardly sustainable in industrial applica- been shown by Stiber et al. [136]. The researchers presented
tions. Despite the fact that carbon materials are generally not a PTL for PEMWE that enables operation at up to 6 A cm−2,
used in anodic PTLs, several studies approve the use of 90°C, and 90 bar H2 output pressure. The PTL consisted of a
International Journal of Energy Research 17

Table 9: Summary of ASTs for the evaluation of GDL stability in PEM technology.

AST mode Operating conditions Research findings Ref.


Mechanical degradation in a PEMFC dummy Puddle-shaped defects decrease the performance
Mechanical degradation [127]
cell for 28 days at 65°C due to water accumulation
A higher compression ratio results in lower cell
Mechanical compression of GDLs to 30% and
performance. Introduction of a MPL onto the
Static compression 50% of its initial thickness at 60°C and 80°C, [128]
carbon cloth is extremely beneficial for the
respectively
operation, especially at high current density
The impact of cyclic compression on the
Compression of GDL samples between a pair of
hydrophobicity of the GDL is seen to be
Cyclic compression graphite plates, which are sandwiched between [129]
significantly higher than that reported elsewhere
aluminum end plates at 1.7 MPa and 3.4 MPa
due to electrochemical effects
Submersion of pristine GDLs in a 35 wt% H2O2 Water accumulation within the GDL is attributed
Accelerated carbon corrosion solution at 90°C for 12 h and further soaking in to the loss of MPL hydrophobicity caused by a [125]
DI water for 24 h carbon-corrosion-based degradation mechanism

Table 10: Summary of ASTs for the evaluation of PTL stability in PEMWEs.

AST mode Operating conditions Research findings Ref.


Durability tests under a load of up Highly porous samples (>30%) exhibit brittle behavior.
Ex situ and in situ
to 300 bar in ex situ and up to Samples with low sintering temperature cannot [111]
pressure tests
50 bar in in situ tests guarantee sufficient fatigue strength
Passivation of the anode Ti-PTL results in an increased
Constant current operation of
Constant current ohmic cell resistance. Pt-coated PTL reduces [75]
2 A cm−2 for 1000 h
degradation by 89%
Longevity test of Carbon-based PTL material sets are sufficient for
40 h operation between 2 V and 3 V [134]
carbon materials initial PEMWE performance characterization
Ex situ corrosion testing at 70 °C for 2.5 h
Longevity test of Substitution of Pt-Ti with cheaper PTL materials
in acidic environment, and in situ
carbon-coated shows potential for significant cost savings in [135]
degradation testing for 30 days
steel materials PEMWE stack design
at 2 A cm−2
Various operation Mitigation strategies are necessary to prevent Ti
See Table 8 [57]
modes/MEA passivation from reducing the cell’s efficiency

Ti porous-sintered layer (PSL) with a low-cost Ti mesh. membrane and ionomer degrade when the electrolyzer
Interestingly, this approach did not require a flow field in operates at elevated temperatures and at OCV. For the
the BPP [136]. Other studies suggest that traditional Ti-based PTL, surface corrosion and passivation are of con-
carbon-based GDL materials are sufficient for electrolyzer cern. This is triggered by the high current density operation
initial performance assessments [134, 137]. To facilitate of the PEMWE, and once a passive film is formed on the
low-cost PTL material and component screening and devel- surface, the contact resistance increases and the perfor-
opment, standardized procedures for various types of mate- mance declines. A precious metal coating is therefore com-
rials and experimental designs are required [24]. monly applied to protect the PTL from corrosion and
passivation of the metal surface. To conclude, novel mate-
5. Concluding Remarks rials and cell components for PEMWE are still slow to
make their way into the market due to the limited access
This document reviews and summarizes different AST to standardized test equipment and procedures and the
protocols in PEMWE technology with a focus on the mem- high cost of precious metal components. The implementa-
brane electrode assembly. The component specific stressors tion of AST protocols has been successful to some extent;
under different operating conditions are outlined for the however, a harmonized database needs to be created to
PFSA-membrane, catalyst layers, and GDL and PTL, move technology from lab scale to commercial deployment.
respectively. Regarding material deterioration in catalyst This complementary investigation of accelerated stressors
degradation research, the current focus is set on the disso- in PEMWE cells provides a basis for the development of
lution of the Ir-catalyst during the OER, whereas on the a variety of ASTs, which can help the research community
cathode-side Pt/C corrosion is a major issue, particularly to quickly assess component durability and efficiently pro-
during the dynamic operation of the PEMWE. The PFSA- cess novel materials.
18 International Journal of Energy Research

Abbreviations mechanical stress and surface passivation. (Supplementary


Materials)
AEMWE: Anion exchange membrane water electrolyzer
ADT: Accelerated degradation test References
AST: Accelerated stress test
AWE: Alkaline water electrolyzer [1] International Energy Agency, “The future of hydrogen-
BPP: Bipolar plate seizing today’s opportunitiesReport prepared by the IEA for
CA: Chronoamperometric the G20, JapanAccessed: Jul. 02, 2022 https://www.iea.org/
CCM: Catalyst coated membrane reports/the-future-of-hydrogen.
CL: Catalyst layer [2] U. Babic, M. Suermann, F. N. Büchi, L. Gubler, and T. J.
DI: Deionized Schmidt, “Critical review—identifying critical gaps for poly-
ECSA: Electrochemically active surface area mer electrolyte water electrolysis development,” Journal of
FRR: Fluoride release rate the Electrochemical Society, vol. 164, no. 4, pp. F387–F399,
GDBL: Gas diffusion backing layer 2017.
GDL: Gas diffusion layer [3] A. Buttler and H. Spliethoff, “Current status of water electrol-
HER: Hydrogen evolution reaction ysis for energy storage, grid balancing and sector coupling via
power-to-gas and power-to-liquids: a review,” Renewable and
HFR: High frequency resistance
Sustainable Energy Reviews, vol. 82, pp. 2440–2454, 2018.
MEA: Membrane electrode assembly
MPL: Microporous layer [4] S. Shiva Kumar and V. Himabindu, “Hydrogen production
by PEM water electrolysis - a review,” Materials Science for
MT: Mass transfer
Energy Technologies, vol. 2, no. 3, pp. 442–454, 2019.
NP: Nanoparticle
[5] T. Wang, X. Cao, and L. Jiao, “PEM water electrolysis for
OCV: Open circuit voltage
hydrogen production: fundamentals, advances, and pros-
OER: Oxygen evolution reaction pects,” Carb Neutrality, vol. 1, no. 1, p. 21, 2022.
PBI: Polybenzimidazoles
[6] Office of Energy Efficiency & Renewable Energy, “DOE
PEMFC: Proton exchange membrane fuel cell
research focuses on overcoming challenges in hydrogen pro-
PEMWE: Polymer electrolyte water electrolysis duction,” Access: Aug. 26, 2022. https://www.energy.gov/
PFSA: Perfluorosulfonic acid eere/fuelcells/hydrogen-production-electrolysis.
PSF: Polysulfone [7] K. Ayers, “The potential of proton exchange membrane-
PTL: Porous transport layer based electrolysis technology,” Current Opinion in Electro-
SEM: Scanning electron microscopy chemistry, vol. 18, pp. 9–15, 2019.
SOEC: Solid oxides electrolysis cell [8] Q. Feng, X.−. Z. Yuan, G. Liu et al., “A review of proton
SPEEK: Sulfonated polyether ether ketones. exchange membrane water electrolysis on degradation mech-
anisms and mitigation strategies,” Journal of Power Sources,
Data Availability vol. 366, pp. 33–55, 2017.
[9] C. Immerz, M. Paidar, G. Papakonstantinou et al., “Effect of
The data that support the findings of this study are available the MEA design on the performance of PEMWE single cells
within the article. with different sizes,” Journal of Applied Electrochemistry,
vol. 48, no. 6, pp. 701–711, 2018.
Conflicts of Interest [10] H. Kim, J. Kim, J. Kim et al., “Dendritic gold-supported irid-
ium/iridium oxide ultra-low loading electrodes for high-
The authors declare that they have no conflicts of interest. performance proton exchange membrane water electrolyzer,”
Applied Catalysis B: Environmental, vol. 283, article 119596,
Acknowledgments 2021.
[11] A. Hartig-Weiss, M. Miller, H. Beyer et al., “Iridium oxide
The financial support by the Austrian Research Promotion
catalyst supported on antimony-doped tin oxide for high
Agency (FFG) through the 8th call of “COMET-Projects” oxygen evolution reaction activity in acidic media,” ACS
is gratefully acknowledged. Applied Nano Materials, vol. 3, no. 3, pp. 2185–2196, 2020.
[12] S. Chatterjee, X. Peng, S. Intikhab et al., “Nanoporous iridium
Supplementary Materials nanosheets for polymer electrolyte membrane electrolysis,”
Advanced Energy Materials, vol. 11, no. 34, article 2101438,
“Graphical Abstract”: membrane electrode assembly (MEA) 2021.
accelerated stressors responsible for degradation of PEM
[13] G. C. da Silva, K. J. J. Mayrhofer, E. A. Ticianelli, and
water electrolysis cells. A series of complex mechanical, S. Cherevko, “Dissolution stability: the major challenge in
chemical, and thermal stressors is responsible for MEA fail- the regenerative fuel cells bifunctional catalysis,” Journal of
ure during PEMWE operation. At the membrane, thinning the Electrochemical Society, vol. 165, no. 16, pp. F1376–
processes and pinhole formation can occur as a result from F1384, 2018.
interactions with oxidizing species formed at the electrodes. [14] M. Carmo, D. L. Fritz, J. Mergel, and D. Stolten, “A compre-
Catalyst accelerated stressors are carbon corrosion, Ostwald hensive review on PEM water electrolysis,” International
ripening, and agglomeration/dissolution processes. The gas Journal of Hydrogen Energy, vol. 38, no. 12, pp. 4901–4934,
diffusion media (GDL and PTL) are mainly affected by 2013.
International Journal of Energy Research 19

[15] F. N. Khatib, T. Wilberforce, O. Ijaodola et al., “Material deg- [30] S. Zhang, X. Yuan, H. Wang et al., “A review of accelerated
radation of components in polymer electrolyte membrane stress tests of MEA durability in PEM fuel cells,” Interna-
(PEM) electrolytic cell and mitigation mechanisms: a review,” tional Journal of Hydrogen Energy, vol. 34, no. 1, pp. 388–
Renewable and Sustainable Energy Reviews, vol. 111, pp. 1– 404, 2009.
14, 2019. [31] R. Petrone, D. Hissel, M. C. Péra, D. Chamagne, and
[16] P. Shirvanian and F. van Berkel, “Novel components in pro- R. Gouriveau, “Accelerated stress test procedures for PEM
ton exchange membrane (PEM) water electrolyzers fuel cells under actual load constraints: state-of-art and pro-
(PEMWE): status, challenges and future needs. A mini posals,” International Journal of Hydrogen Energy, vol. 40,
review,” Electrochemistry Communications, vol. 114, article no. 36, pp. 12489–12505, 2015.
106704, 2020. [32] X.-Z. Yuan, H. Li, S. Zhang, J. Martin, and H. Wang, “A
[17] S. Cherevko, “Stability and dissolution of electrocatalysts: review of polymer electrolyte membrane fuel cell durability
building the bridge between model and "real world" systems,” test protocols,” Journal of Power Sources, vol. 196, no. 22,
Current Opinion in Electrochemistry, vol. 8, pp. 118–125, pp. 9107–9116, 2011.
2018. [33] S. Y. Baek, J. Y. Lee, T. K. Lee, H. J. Kim, and S. Y. Lee, “Asym-
[18] C. V. Pham, D. Escalera-López, K. Mayrhofer, S. Cherevko, metric porous membranes with improving mass transfer for
and S. Thiele, “Essentials of high performance water electro- proton exchange membrane water electrolysis (PEMWE),”
lyzers–from catalyst layer materials to electrode engineering,” ECS Meeting Abstracts, vol. MA2020-02, no. 38, pp. 2485–
Advanced Energy Materials, vol. 11, no. 44, article 2101998, 2485, 2020.
2021. [34] M. F. Ahmad Kamaroddin, N. Sabli, T. A. Tuan Abdullah
[19] H. Ito, T. Maeda, A. Nakano, and H. Takenaka, “Properties of et al., “Membrane-based electrolysis for hydrogen produc-
Nafion membranes under PEM water electrolysis condi- tion: a review,” Membranes, vol. 11, no. 11, p. 810, 2021.
tions,” International Journal of Hydrogen Energy, vol. 36, [35] M. Chandesris, V. Médeau, N. Guillet, S. Chelghoum,
no. 17, pp. 10527–10540, 2011. D. Thoby, and F. Fouda-Onana, “Numerical modelling of
[20] A. Kusoglu and A. Z. Weber, “New insights into perfluori- membrane degradation in PEM water electrolyzer: influence
nated sulfonic-acid ionomers,” Chemical Reviews, vol. 117, of the temperature and current density,” International Jour-
no. 3, pp. 987–1104, 2017. nal of Hydrogen Energy, vol. 40, pp. 1–8, 2015.
[21] K. Ayers, “High efficiency PEM water electrolysis: enabled by [36] L. Gubler, S. M. Dockheer, and W. H. Koppenol, “Radicals in
advanced catalysts, membranes, and processes,” Current fuel cell membranes: mechanisms of formation and ionomer
Opinion in Chemical Engineering, vol. 33, article 100719, attack,” ECS Transactions, vol. 41, no. 1, pp. 1431–1439,
2021. 2011.
[22] T. L. Doan, H. E. Lee, S. S. H. Shah et al., “A review of the [37] D. Curtin, R. Lousenberg, T. Henry, P. Tangeman, and
porous transport layer in polymer electrolyte membrane M. Tisack, “Chemical Degradation of Perfluorinated Sulfonic
water electrolysis,” International Journal of Energy Research, Acid Membranes,” in Polymer Electrolyte Fuel Cell Durabil-
vol. 45, no. 10, pp. 14207–14220, 2021. ity, Power Sources, pp. 57–69, Springer, 2004.
[23] C. H. Lee, R. Banerjee, F. Arbabi, J. Hinebaugh, and [38] S. H. Frensch, G. Serre, F. Fouda-Onana et al., “Impact of iron
A. Bazylak, “Porous transport layer related mass transport and hydrogen peroxide on membrane degradation for poly-
losses in polymer electrolyte membrane electrolysis: a mer electrolyte membrane water electrolysis: computational
review,” in ASME 2016 14th International Conference on and experimental investigation on fluoride emission,” Jour-
Nanochannels, Microchannels, and Minichannels, pp. 1–11, nal of Power Sources, vol. 420, pp. 54–62, 2019.
Washington, DC, USA, 2016. [39] P. Marocco, K. Sundseth, T. Aarhaug et al., “Online measure-
[24] X.-Z. Yuan, N. Shaigan, C. Song et al., “The porous transport ments of fluoride ions in proton exchange membrane water
layer in proton exchange membrane water electrolysis: per- electrolysis through ion chromatography,” Journal of Power
spectives on a complex component,” Sustainable Energy & Sources, vol. 483, article 229179, 2021.
Fuels, vol. 6, no. 8, pp. 1824–1853, 2022. [40] A. Martin, P. Trinke, B. Bensmann, and R. Hanke-Rauschen-
[25] H. Teuku, I. Alshami, J. Goh, M. S. Masdar, and K. S. Loh, bach, “Hydrogen crossover in PEM water electrolysis at cur-
“Review on bipolar plates for low-temperature polymer rent densities up to 10 a cm−2,” Journal of the Electrochemical
electrolyte membrane water electrolyzer,” International Society, vol. 169, no. 9, article 094507, 2022.
Journal of Energy Research, vol. 45, no. 15, pp. 20583– [41] P. Trinke, G. P. Keeley, M. Carmo, B. Bensmann, and
20600, 2021. R. Hanke-Rauschenbach, “Elucidating the effect of mass
[26] G. Tsotridis and A. Pilenga, EU Harmonized Protocols for transport resistances on hydrogen crossover and cell perfor-
Testing of Low Temperature Water Electrolysis, Publications mance in PEM water electrolyzers by varying the cathode
Office of the European Union, Luxembourg, 2021. ionomer content,” Journal of the Electrochemical Society,
[27] T. Malkow, A. Pilenga, and G. Tsotridis, EU Harmonised Test vol. 166, no. 8, pp. F465–F471, 2019.
Procedure: Electrochemical Impedance Spectroscopy for Water [42] P. Holzapfel, M. Bühler, C. van Pham et al., “Directly coated
Electrolysis Cells, Publications Office, 2018. membrane electrode assemblies for proton exchange mem-
[28] T. Malkow, EU Harmonised Cyclic Voltammetry Test Method brane water electrolysis,” Electrochemistry Communications,
for Low-Temperature Water Electrolysis Single Cells, Publica- vol. 110, article 106640, 2020.
tions Office, 2018. [43] X.-Z. Yuan, S. Zhang, S. Ban et al., “Degradation of a PEM
[29] T. Malkow, EU Harmonised Polarisation Curve Test Method fuel cell stack with Nafion® membranes of different thick-
for Low-Temperature Water Electrolysis, Publications Office, nesses. Part II: ex situ diagnosis,” Journal of Power Sources,
2018. vol. 205, pp. 324–334, 2012.
20 International Journal of Energy Research

[44] X.-Z. Yuan, S. Zhang, H. Wang et al., “Degradation of a poly- in fuel cell membranes,” Journal of the Electrochemical Soci-
mer exchange membrane fuel cell stack with Nafion® mem- ety, vol. 164, no. 13, pp. F1331–F1341, 2017.
branes of different thicknesses: part I. in situ diagnosis,” [59] Y. Xing, H. Li, and G. Avgouropoulos, “Research progress of
Journal of Power Sources, vol. 195, no. 22, pp. 7594–7599, proton exchange membrane failure and mitigation strate-
2010. gies,” Materials, vol. 14, no. 10, p. 2591, 2021.
[45] M. Schalenbach, M. A. Hoeh, J. T. Gostick, W. Lueke, and [60] S. A. Grigoriev, D. G. Bessarabov, and V. N. Fateev, “Degra-
D. Stolten, “Gas permeation through Nafion. Part 2: resistor dation mechanisms of MEA characteristics during water elec-
network model,” Journal of Physical Chemistry C, vol. 119, trolysis in solid polymer electrolyte cells,” Russian Journal of
no. 45, pp. 25156–25169, 2015. Electrochemistry, vol. 53, no. 3, pp. 318–323, 2017.
[46] P. Trinke, B. Bensmann, and R. Hanke-Rauschenbach, “Cur- [61] N. Danilovic, R. Subbaraman, K. C. Chang et al., “Activity–
rent density effect on hydrogen permeation in PEM water stability trends for the oxygen evolution reaction on mono-
electrolyzers,” International Journal of Hydrogen Energy, metallic oxides in acidic environments,” Journal of Physical
vol. 42, no. 21, pp. 14355–14366, 2017. Chemistry Letters, vol. 5, no. 14, pp. 2474–2478, 2014.
[47] S. S. Kocha, J. Deliang Yang, and J. S. Yi, “Characterization of [62] A. Weiß, A. Siebel, M. Bernt, T.-H. Shen, V. Tileli, and H. A.
gas crossover and its implications in PEM fuel cells,” AICHE Gasteiger, “Impact of intermittent operation on lifetime and
Journal, vol. 52, no. 5, pp. 1916–1925, 2006. performance of a PEM water electrolyzer,” Journal of the
[48] A. Martin, P. Trinke, M. Stähler et al., “The effect of cell com- Electrochemical Society, vol. 166, no. 8, pp. F487–F497, 2019.
pression and cathode pressure on hydrogen crossover in [63] N. Li, S. S. Araya, and S. K. Kær, “Investigating low and high
PEM water electrolysis,” Journal of the Electrochemical Soci- load cycling tests as accelerated stress tests for proton
ety, vol. 169, no. 1, article 014502, 2022. exchange membrane water electrolysis,” Electrochimica Acta,
[49] D. Ye, Z. Tu, Y. Yu et al., “Hydrogen permeation across vol. 370, article 137748, 2021.
super-thin membrane and the burning limitation in low- [64] F. Fouda-Onana, M. Chandesris, V. Médeau, S. Chelghoum,
temperature proton exchange membrane fuel cell,” Interna- D. Thoby, and N. Guillet, “Investigation on the degradation
tional Journal of Energy Research, vol. 38, no. 9, pp. 1181– of MEAs for PEM water electrolysers part I: effects of testing
1191, 2014. conditions on MEA performances and membrane proper-
[50] P. Trinke, B. Bensmann, S. Reichstein, R. Hanke-Rauschen- ties,” International Journal of Hydrogen Energy, vol. 41,
bach, and K. Sundmacher, “Impact of pressure and tempera- no. 38, pp. 16627–16636, 2016.
ture on hydrogen permeation in PEM water electrolyzers [65] P. Lettenmeier, R. Wang, R. Abouatallah et al., “Durable
operated at asymmetric pressure conditions,” ECS Transac- membrane electrode assemblies for proton exchange mem-
tions, vol. 75, no. 14, pp. 1081–1094, 2016. brane electrolyzer systems operating at high current densi-
[51] P. Millet, A. Ranjbari, F. de Guglielmo, S. A. Grigoriev, and ties,” Electrochimica Acta, vol. 210, pp. 502–511, 2016.
F. Auprêtre, “Cell failure mechanisms in PEM water electro- [66] S. Kundu, L. C. Simon, and M. W. Fowler, “Comparison of
lyzers,” International Journal of Hydrogen Energy, vol. 37, two accelerated Nafion™ degradation experiments,” Polymer
no. 22, pp. 17478–17487, 2012. Degradation and Stability, vol. 93, no. 1, pp. 214–224, 2008.
[52] M. Bodner, C. Hochenauer, and V. Hacker, “Effect of pinhole [67] S. Goldstein, D. Meyerstein, and G. Czapski, “The Fenton
location on degradation in polymer electrolyte fuel cells,” reagents,” Free Radical Biology and Medicine, vol. 15, no. 4,
Journal of Power Sources, vol. 295, pp. 336–348, 2015. pp. 435–445, 1993.
[53] R. W. Kopitzke, C. A. Linkous, H. R. Anderson, and G. L. [68] Z. Liu, T. Wang, X. Yu, Z. Geng, Y. Sang, and H. Liu, “In situ
Nelson, “Conductivity and water uptake of aromatic-based alternative switching between Ti4+ and Ti3+ driven by H2O2
proton exchange membrane electrolytes,” Journal of the Elec- in TiO2 nanostructures: mechanism of pseudo-Fenton reac-
trochemical Society, vol. 147, no. 5, p. 1677, 2000. tion,” Materials Chemistry Frontiers, vol. 1, no. 10,
[54] M. Doyle, M. E. Lewittes, M. G. Roelofs, S. A. Perusich, and pp. 1989–1994, 2017.
R. E. Lowrey, “Relationship between ionic conductivity of [69] A. Laconti, H. Liu, C. Mittelsteadt, and R. McDonald, “Poly-
perfluorinated ionomeric membranes and nonaqueous sol- mer electrolyte membrane degradation mechanisms in fuel
vent properties,” Journal of Membrane Science, vol. 184, cells - findings over the past 30 years and comparison with
no. 2, pp. 257–273, 2001. electrolyzers,” ECS Transactions, vol. 1, no. 8, pp. 199–219,
[55] Ö. F. Selamet, M. C. Acar, M. D. Mat, and Y. Kaplan, “Effects 2006.
of operating parameters on the performance of a high- [70] E. Borgardt, L. Giesenberg, M. Reska et al., “Impact of clamp-
pressure proton exchange membrane electrolyzer,” Interna- ing pressure and stress relaxation on the performance of dif-
tional Journal of Energy Research, vol. 37, no. 5, pp. 457– ferent polymer electrolyte membrane water electrolysis cell
467, 2013. designs,” International Journal of Hydrogen Energy, vol. 44,
[56] G. S. Ogumerem and E. N. Pistikopoulos, “Parametric opti- no. 42, pp. 23556–23567, 2019.
mization and control for a smart proton exchange membrane [71] S. Siracusano, F. Pantò, S. Tonella, C. Oldani, and A. S. Aricò,
water electrolysis (PEMWE) system,” Journal of Process Con- “Reinforced short-side-chain Aquivion® membrane for pro-
trol, vol. 91, pp. 37–49, 2020. ton exchange membrane water electrolysis,” International
[57] S. H. Frensch, F. Fouda-Onana, G. Serre, D. Thoby, S. S. Journal of Hydrogen Energy, vol. 47, no. 35, pp. 15557–
Araya, and S. K. Kær, “Influence of the operation mode on 15570, 2022.
PEM water electrolysis degradation,” International Journal [72] J. Shan, T. Ling, K. Davey, Y. Zheng, and S. Qiao, “Transition-
of Hydrogen Energy, vol. 44, no. 57, pp. 29889–29898, 2019. metal-doped RuIr bifunctional nanocrystals for overall water
[58] Y. Singh, F. P. Orfino, M. Dutta, and E. Kjeang, “3D failure splitting in acidic environments,” Advanced Materials,
analysis of pure mechanical and pure chemical degradation vol. 31, no. 17, article 1900510, 2019.
International Journal of Energy Research 21

[73] M. Bernt, A. Siebel, and H. A. Gasteiger, “Analysis of voltage [88] J. Dodwell, M. Maier, J. Majasan et al., “Open-circuit dissolu-
losses in PEM water electrolyzers with low platinum group tion of platinum from the cathode in polymer electrolyte
metal loadings,” Journal of the Electrochemical Society, membrane water electrolysers,” Journal of Power Sources,
vol. 165, no. 5, pp. F305–F314, 2018. vol. 498, article 229937, 2021.
[74] C. Spöri, J. T. H. Kwan, A. Bonakdarpour, D. P. Wilkinson, [89] P. Aßmann, A. S. Gago, P. Gazdzicki, K. A. Friedrich, and
and P. Strasser, “The stability challenges of oxygen evolving M. Wark, “Toward developing accelerated stress tests for
catalysts: towards a common fundamental understanding proton exchange membrane electrolyzers,” Current Opinion
and mitigation of catalyst degradation,” Angewandte Chemie, in Electrochemistry, vol. 21, pp. 225–233, 2020.
International Edition, vol. 56, no. 22, pp. 5994–6021, 2017. [90] S. M. Alia, S. Stariha, and R. L. Borup, “Electrolyzer durability
[75] C. Rakousky, U. Reimer, K. Wippermann, M. Carmo, at low catalyst loading and with dynamic operation,” Journal
W. Lueke, and D. Stolten, “An analysis of degradation phe- of the Electrochemical Society, vol. 166, no. 15, pp. F1164–
nomena in polymer electrolyte membrane water electrolysis,” F1172, 2019.
Journal of Power Sources, vol. 326, pp. 120–128, 2016. [91] C. Rakousky, G. P. Keeley, K. Wippermann, M. Carmo, and
[76] C. Rakousky, U. Reimer, K. Wippermann et al., “Polymer D. Stolten, “The stability challenge on the pathway to high-
electrolyte membrane water electrolysis: restraining degrada- current-density polymer electrolyte membrane water electro-
tion in the presence of fluctuating power,” Journal of Power lyzers,” Electrochimica Acta, vol. 278, pp. 324–331, 2018.
Sources, vol. 342, pp. 38–47, 2017. [92] S. Siracusano, N. Hodnik, P. Jovanovic et al., “New insights
[77] P. Paciok, C. Rakousky, and C. Marcelo, “Sensivity analysis of into the stability of a high performance nanostructured cata-
a PEM electrolyser cathode with respect to the platinum and lyst for sustainable water electrolysis,” Nano Energy, vol. 40,
Nafion loading,” ECS Meeting Abstracts, vol. MA2015-01, pp. 618–632, 2017.
no. 26, 2015. [93] S. A. Grigoriev, K. A. Dzhus, D. G. Bessarabov, and P. Millet,
[78] S. Shrestha, Y. Liu, and W. E. Mustain, “Electrocatalytic activ- “Failure of PEM water electrolysis cells: case study involving
ity and stability of Pt clusters on state-of-the-art supports: a anode dissolution and membrane thinning,” International
review,” Catalysis Reviews, vol. 53, no. 3, pp. 256–336, 2011. Journal of Hydrogen Energy, vol. 39, no. 35, pp. 20440–
[79] F. Claudel, L. Dubau, G. Berthomé et al., “Degradation mech- 20446, 2014.
anisms of oxygen evolution reaction electrocatalysts: a com- [94] E. Brightman, J. Dodwell, N. van Dijk, and G. Hinds, “In situ
bined identical-location transmission electron microscopy characterisation of PEM water electrolysers using a novel ref-
and X-ray photoelectron spectroscopy study,” ACS Catalysis, erence electrode,” Electrochemistry Communications, vol. 52,
vol. 9, no. 5, pp. 4688–4698, 2019. pp. 1–4, 2015.
[80] P. Paciok, M. Carmo, W. Lüke, M. Heggen, and D. Stolten, [95] C. Spöri, C. Brand, M. Kroschel, and P. Strasser, “Accelerated
“Investigation of Pt/C catalysts using identical location trans- degradation protocols for iridium-based oxygen evolving cat-
mission electron microscopy under PEM electrolysis condi- alysts in water splitting devices,” Journal of the Electrochemi-
tions,” ECS Meeting Abstracts, vol. MA2016-02, no. 38, 2016. cal Society, vol. 168, no. 3, article 034508, 2021.
[81] H. Yu, L. Bonville, J. Jankovic, and R. Maric, “Microscopic [96] X. Wang, L. Zhang, G. Li, G. Zhang, Z.-G. Shao, and B. Yi,
insights on the degradation of a PEM water electrolyzer with “The influence of ferric ion contamination on the solid poly-
ultra-low catalyst loading,” Applied Catalysis B: Environmen- mer electrolyte water electrolysis performance,” Electrochi-
tal, vol. 260, article 118194, 2020. mica Acta, vol. 158, pp. 253–257, 2015.
[82] J. Knöppel, M. Möckl, D. Escalera-López et al., “On the limi- [97] N. Li, S. S. Araya, X. Cui, and S. K. Kær, “The effects of cat-
tations in assessing stability of oxygen evolution catalysts ionic impurities on the performance of proton exchange
using aqueous model electrochemical cells,” Nature Commu- membrane water electrolyzer,” Journal of Power Sources,
nications, vol. 12, no. 1, p. 2231, 2021. vol. 473, article 228617, 2020.
[83] O. Kasian, S. Geiger, T. Li et al., “Degradation of iridium oxi- [98] A. Uddin, J. Park, U. Pasaogullari, and L. J. Bonville, “Cation
desviaoxygen evolution from the lattice: correlating atomic contamination in polymer electrolyte fuel cells: impacts,
scale structure with reaction mechanisms,” Energy & Envi- mechanisms, and mitigation,” ECS Meeting Abstracts, vol. -
ronmental Science, vol. 12, no. 12, pp. 3548–3555, 2019. MA2015-03, no. 3, 2015.
[84] T. Naito, T. Shinagawa, T. Nishimoto, and K. Takanabe, [99] E. Price, “Durability and degradation issues in PEM electrol-
“Recent advances in understanding oxygen evolution reac- ysis cells and its components,” Johnson Matthey Technology
tion mechanisms over iridium oxide,” Inorganic Chemistry Review, vol. 61, no. 1, pp. 47–51, 2017.
Frontiers, vol. 8, no. 11, pp. 2900–2917, 2021. [100] X. Cheng, Z. Shi, N. Glass et al., “A review of PEM hydrogen
[85] S. Cherevko, S. Geiger, O. Kasian, A. Mingers, and K. J. J. fuel cell contamination: impacts, mechanisms, and mitiga-
Mayrhofer, “Oxygen evolution activity and stability of irid- tion,” Journal of Power Sources, vol. 165, no. 2, pp. 739–756,
ium in acidic media. Part 2. - electrochemically grown 2007.
hydrous iridium oxide,” Journal of Electroanalytical Chemis- [101] H. Sato, H. Matsushima, M. Ueda, and H. Ito, “Effect of water
try, vol. 774, pp. 102–110, 2016. stoichiometry on deuterium isotope separation by anion
[86] P. J. Rheinländer and J. Durst, “Transformation of the OER- exchange membrane water electrolysis,” International Jour-
active IrOxspecies under transient operation conditions in nal of Hydrogen Energy, vol. 46, no. 68, pp. 33689–33695,
PEM water electrolysis,” Journal of the Electrochemical Soci- 2021.
ety, vol. 168, no. 2, article 024511, 2021. [102] N. Li, S. S. Araya, and S. K. Kær, “Long-term contamination
[87] V. Pfeifer, T. E. Jones, J. J. Velasco Vélez et al., “The electronic effect of iron ions on cell performance degradation of proton
structure of iridium and its oxides,” Surface and Interface exchange membrane water electrolyser,” Journal of Power
Analysis, vol. 48, no. 5, pp. 261–273, 2016. Sources, vol. 434, article 226755, 2019.
22 International Journal of Energy Research

[103] C. J. Banas, M. A. Uddin, J. Park, L. J. Bonville, and cells,” Journal of Fuel Cell Science and Technology, vol. 1,
U. Pasaogullari, “Thinning of cathode catalyst layer in poly- no. 1, pp. 2–9, 2004.
mer electrolyte fuel cells due to foreign cation contamina- [117] P. M. Wilde, M. Mändle, M. Murata, and N. Berg, “Structural
tion,” Journal of the Electrochemical Society, vol. 165, no. 6, and physical properties of GDL and GDL/BPP combinations
pp. F3015–F3023, 2018. and their influence on PEMFC performance,” Fuel Cells,
[104] L. Zhang, X. Jie, Z.-G. Shao, Z.-M. Zhou, G. Xiao, and B. Yi, vol. 4, no. 3, pp. 180–184, 2004.
“The influence of sodium ion on the solid polymer electrolyte [118] S. Song, H. Zhang, X. Ma, Z.-G. Shao, Y. Zhang, and B. Yi,
water electrolysis,” International Journal of Hydrogen Energy, “Bifunctional oxygen electrode with corrosion-resistive gas
vol. 37, no. 2, pp. 1321–1325, 2012. diffusion layer for unitized regenerative fuel cell,” Electro-
[105] D. Kulkarni, A. Huynh, P. Satjaritanun et al., “Elucidating chemistry Communications, vol. 8, no. 3, pp. 399–405, 2006.
effects of catalyst loadings and porous transport layer mor- [119] G. Chen, H. Zhang, H. Zhong, and H. Ma, “Gas diffusion
phologies on operation of proton exchange membrane water layer with titanium carbide for a unitized regenerative fuel
electrolyzers,” Applied Catalysis B: Environmental, vol. 308, cell,” Electrochimica Acta, vol. 55, no. 28, pp. 8801–8807,
article 121213, 2022. 2010.
[106] Y. N. Regmi, E. Tzanetopoulos, G. Zeng et al., “Supported [120] F.-Y. Zhang, S. G. Advani, and A. K. Prasad, “Performance of
oxygen evolution catalysts by design: toward lower precious a metallic gas diffusion layer for PEM fuel cells,” Journal of
metal loading and improved conductivity in proton exchange Power Sources, vol. 176, no. 1, pp. 293–298, 2008.
membrane water electrolyzers,” ACS Catalysis, vol. 10, no. 21,
pp. 13125–13135, 2020. [121] F.-Y. Zhang, A. K. Prasad, and S. G. Advani, “Investigation of
a copper etching technique to fabricate metallic gas diffusion
[107] E. Cruz Ortiz, F. Hegge, M. Breitwieser, and S. Vierrath,
media,” Journal of Micromechanics and Microengineering,
“Improving the performance of proton exchange membrane
vol. 16, no. 11, pp. N23–N27, 2006.
water electrolyzers with low Ir-loaded anodes by adding
PEDOT: PSS as electrically conductive binder,” RSC [122] F. Lapicque, M. Belhadj, C. Bonnet, J. Pauchet, and
Advances, vol. 10, no. 62, pp. 37923–37927, 2020. Y. Thomas, “A critical review on gas diffusion micro and
macroporous layers degradations for improved membrane
[108] O. Kasian, J.-P. Grote, S. Geiger, S. Cherevko, and K. J. J.
fuel cell durability,” Journal of Power Sources, vol. 336,
Mayrhofer, “The common intermediates of oxygen evolution
pp. 40–53, 2016.
and dissolution reactions during water electrolysis on irid-
ium,” Angewandte Chemie, International Edition, vol. 57, [123] Y. Pan, H. Wang, and N. P. Brandon, “Gas diffusion layer
no. 9, pp. 2488–2491, 2018. degradation in proton exchange membrane fuel cells: mecha-
[109] S. Steen and F.-Y. Zhang, “Optimization of gas diffusion nisms, characterization techniques and modelling
layers with high-corrosion resistance for PEMWE,” ECS approaches,” Journal of Power Sources, vol. 513, article
Meeting Abstracts, vol. MA2014-01, no. 14, 2014. 230560, 2021.
[110] C. Liu, M. Carmo, G. Bender et al., “Performance enhance- [124] M. G. George, H. Liu, D. Muirhead et al., “Accelerated degra-
ment of PEM electrolyzers through iridium-coated titanium dation of polymer electrolyte membrane fuel cell gas diffu-
porous transport layers,” Electrochemistry Communications, sion layers,” Journal of the Electrochemical Society, vol. 164,
vol. 97, pp. 96–99, 2018. no. 7, pp. F714–F721, 2017.
[111] E. Borgardt, O. Panchenko, F. J. Hackemüller et al., “Mechan- [125] H. Liu, M. G. George, N. Ge et al., “Microporous layer degra-
ical characterization and durability of sintered porous trans- dation in polymer electrolyte membrane fuel cells,” Journal of
port layers for polymer electrolyte membrane electrolysis,” the Electrochemical Society, vol. 165, no. 6, pp. F3271–F3280,
Journal of Power Sources, vol. 374, pp. 84–91, 2018. 2018.
[112] O. Panchenko, E. Borgardt, W. Zwaygardt et al., “In-situ two- [126] A. M. Dafalla and F. Jiang, “Stresses and their impacts on pro-
phase flow investigation of different porous transport layer ton exchange membrane fuel cells: a review,” International
for a polymer electrolyte membrane (PEM) electrolyzer with Journal of Hydrogen Energy, vol. 43, no. 4, pp. 2327–2348,
neutron spectroscopy,” Journal of Power Sources, vol. 390, 2018.
pp. 108–115, 2018. [127] J. H. Chun, D. H. Jo, S. G. Kim, S. H. Park, C. H. Lee, and S. H.
[113] J. Mo, R. R. Dehoff, W. H. Peter, T. J. Toops, J. B. Green, and Kim, “Improvement of the mechanical durability of micro
F.-Y. Zhang, “Additive manufacturing of liquid/gas diffusion porous layer in a proton exchange membrane fuel cell by
layers for low-cost and high- efficiency hydrogen produc- elimination of surface cracks,” Renewable Energy, vol. 48,
tion,” International Journal of Hydrogen Energy, vol. 41, pp. 35–41, 2012.
no. 4, pp. 3128–3135, 2016. [128] G. Dotelli, L. Omati, P. Gallo Stampino, P. Grassini, and
[114] C. M. Hwang, M. Ishida, H. Ito et al., “Influence of properties D. Brivio, “Investigation of gas diffusion layer compression
of gas diffusion layers on the performance of polymer by electrochemical impedance spectroscopy on running poly-
electrolyte-based unitized reversible fuel cells,” International mer electrolyte membrane fuel cells,” Journal of Power
Journal of Hydrogen Energy, vol. 36, no. 2, pp. 1740–1753, Sources, vol. 196, no. 21, pp. 8955–8966, 2011.
2011. [129] V. Radhakrishnan and P. Haridoss, “Effect of cyclic compres-
[115] S. Escribano, J.-F. Blachot, J. Ethève, A. Morin, and sion on structure and properties of a gas diffusion layer used
R. Mosdale, “Characterization of PEMFCs gas diffusion in PEM fuel cells,” International Journal of Hydrogen Energy,
layers properties,” Journal of Power Sources, vol. 156, no. 1, vol. 35, no. 20, pp. 11107–11118, 2010.
pp. 8–13, 2006. [130] J. O. Majasan, F. Iacoviello, J. I. S. Cho et al., “Correlative
[116] V. Mishra, F. Yang, and R. Pitchumani, “Measurement and study of microstructure and performance for porous trans-
prediction of electrical contact resistance between gas diffu- port layers in polymer electrolyte membrane water electroly-
sion layers and bipolar plate for applications to PEM fuel sers by X-ray computed tomography and electrochemical
International Journal of Energy Research 23

characterization,” International Journal of Hydrogen Energy,


vol. 44, no. 36, pp. 19519–19532, 2019.
[131] P. W. T. Lu and S. Srinivasan, “Advances in water electrolysis
technology with emphasis on use of the solid polymer elec-
trolyte,” Journal of Applied Electrochemistry, vol. 9, no. 3,
pp. 269–283, 1979.
[132] A. Nouri-Khorasani, E. Tabu Ojong, T. Smolinka, and D. P.
Wilkinson, “Model of oxygen bubbles and performance
impact in the porous transport layer of PEM water electroly-
sis cells,” International Journal of Hydrogen Energy, vol. 42,
no. 48, pp. 28665–28680, 2017.
[133] A. S. Pushkarev, I. V. Pushkareva, M. A. Solovyev et al., “On
the influence of porous transport layers parameters on the per-
formances of polymer electrolyte membrane water electrolysis
cells,” Electrochimica Acta, vol. 399, article 139436, 2021.
[134] J. L. Young, Z. Kang, F. Ganci, S. Madachy, and G. Bender,
“PEM electrolyzer characterization with carbon-based hard-
ware and material sets,” Electrochemistry Communications,
vol. 124, article 106941, 2021.
[135] H. Becker, E. J. F. Dickinson, X. Lu et al., “Assessing potential
profiles in water electrolysers to minimise titanium use,”
Energy & Environmental Science, vol. 15, no. 6, pp. 2508–
2518, 2022.
[136] S. Stiber, H. Balzer, A. Wierhake et al., “Porous transport
layers for proton exchange membrane electrolysis under
extreme conditions of current density, temperature, and pres-
sure,” Advanced Energy Materials, vol. 11, no. 33, article
2100630, 2021.
[137] J. Schröder, V. A. Mints, A. Bornet et al., “The gas diffusion
electrode setup as straightforward testing device for proton
exchange membrane water electrolyzer catalysts,” JACS Au,
vol. 1, no. 3, pp. 247–251, 2021.

You might also like