You are on page 1of 30

Sustainable

Energy & Fuels


View Article Online
REVIEW View Journal | View Issue

Current understanding of chemical degradation


mechanisms of perfluorosulfonic acid membranes
Cite this: Sustainable Energy Fuels,
2017, 1, 409 and their mitigation strategies: a review
M. Zatoń, J. Rozière and D. J. Jones*

This article provides a comprehensive and up-to-date perspective of the understanding of


Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

perfluorosulfonic acid (PFSA) fuel cell membrane degradation phenomena, reviews key concepts for
the mitigation of membrane degradation, appraises the effectiveness of these strategies by discussing
their benefits and drawbacks, and identifies remaining challenges and research priorities for fuel cell
membranes with increased longevity. Identification of stressors and improved understanding of how
scavenging reactions proceed are essential for accelerated development of new fuel cell membrane
materials with antioxidant properties and enhanced durability. Fuel cells convert the chemical energy of
a fuel such as hydrogen into electricity and heat and have the potential to deliver environmental and
economic benefits across various sectors, including transportation, power generation, industrial
equipment, military power, and consumer electronics. The development of Proton Exchange
Membrane Fuel Cell (PEMFC) membranes with increased durability is of crucial importance since it
directly impacts fuel cell system lifetime and, therefore, large scale implementation of fuel cell
technology. The PFSA proton conducting polymer (ionomer) at the heart of the membrane electrode
assembly of a PEMFC is subject to chemical degradation as a result of the attack on the polymer chains
by reactive oxygen species generated electrochemically in the fuel cell. For the first time, this article
Received 20th January 2017
Accepted 13th March 2017
reviews the literature both on mechanisms by which this chemical degradation of the ionomer
membrane occurs and the sites on the polymer susceptible to free radical attack, as well as how this
DOI: 10.1039/c7se00038c
learning and increased understanding have been used to build and develop successful mitigation
rsc.li/sustainable-energy strategies designed to annihilate the harmful effect of oxidative species on membrane integrity.

electrolyte, an acid or an alkaline fuel cell can be distin-


1. Introduction guished. According to the operation temperature, fuel cells
Fossil fuels are currently the main source of energy. The may be grouped as low (up to 100  C), intermediate (up to 600

growing concern for the shrinking availability of the world's C) and high temperature (up to 1000  C). This review is solely
oil and gas supplies and the detrimental impact on the envi- focused on Low Temperature Proton Exchange Membrane
ronment associated with the use of such sources of energy Fuel Cells (LTPEMFC). PEMFC are known to have interesting
have motivated the search for alternative, clean energy tech- properties such as high power density, easy scale-up and fast
nologies. Fuel Cells (FC) are devices that convert the chemical start-up capability. These features make the device an excel-
energy stored in fuels into electricity via electrochemical lent candidate for automotive and small stationary applica-
reactions. Since fuel cells can generate electricity continuously tions and portable electronics. However the direction of
with high efficiency and low or zero pollution emission at the PEMFC development currently poised to make a real impact
point of use as long as the fuel and oxidant are supplied, they on emissions reduction is in transportation applications.
represent a promising substitute to heat and electricity Prototype Fuel Cell Electric Vehicles (FCEVs) have already been
generation from fossil fuels, and can be used for a wide range presented by most automakers and were commercialized in
of applications from micro power sources to multi – MW 2015 Toyota, and in 2016 by Hyundai and Honda. Furthermore
power plants. There are many types of fuel cells that can be PEM fuel cells are already used in certain stationary, portable
classied according to the ion transport medium and/or to the and transport applications in most developed countries.1–4 A
operation temperature. Regarding the nature of the lot of research is being done in this area due to the economic
and environmental benets that such systems can offer.
However there are still challenges that are hindering the global
ICGM – Aggregates, Interfaces and Materials for Energy, UMR 5253, Université de
Montpellier 2, Place Eugène Bataillon, 34095 Montpellier, France. E-mail: Deborah.
applications of fuel cells. The main issue is the practical long-
Jones@umontpellier.fr term operation of the cell components.5–7 There are several

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 409
View Article Online

Sustainable Energy & Fuels Review

losses that limit the long-term stability and the performance of Excellent articles by Borup et al.20 and by de Bruijn et al.26
these devices.8 Those losses are due to the degradation of the reviewed the literature on degradation and durability of fuel cell
bipolar plates, the gas diffusion layer, the catalysts and the components up to 2008 while that of Wu et al.9 additionally
membranes.9 discussed possible mitigation approaches. In 2014 the article
The loss of the active surface area of the catalyst layer, Dubau et al. provided an updated overview of degradation of
which results in the decrease of the performance of the fuel fuel cell components.27 None of these reviews was focussed on
cell, is caused by platinum oxidation, dissolution and corro- membrane degradation phenomena. Other authors summa-
sion of the carbon material, which can lead to aggregation of rized and classify accelerated stress test (AST) protocols,28,29
Pt particles and their detachment from the support.10 The high aiming to provide practical tools for the study of fuel cell
potential delivered by the system as well as the oxidative durability. The article from Rodgers et al. reviewed different
environment favour the growth of Pt particles. Platinum accelerated stress test and lifetime tests and correlated them to
dissolution is not only accelerated by potential or load cycling membrane decay and fuel cell lifetime.30 Ishimoto et al. pub-
but as well as the elevated temperatures used in these lished a survey of literature on the mechanism of chemical
devices.11–13 Migration of the dissolved ionic Pt and further degradation of PFSA membranes from an atomistic point of
precipitation in the membrane can be in certain conditions view.31 Despite the signicant number of publications on the
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

considered as one of the factors contributing to membrane topics of PFSA membrane degradation and mitigation strategies
degradation.14,15 Oxidation of the carbon support is initiated to reduce this degradation that have been published over the
by high-electrochemical potentials caused either by fuel/air past decade, there is no article providing a comprehensive and
boundary passing over the active area during the start-up/ detailed review of the eld.
shut-down processes or local hydrogen starvation.16 Carbon The goal here is to review the substantial literature on how
corrosion results in the thinning of the electrode16 and growth increased understanding of factors that lead to the degradation
of the platinum particles due to detachment and agglomera- of PFSA membranes has led to the development of mitigation
tion.17–19 GDL degradation is related to loss of hydrophobicity strategies and improvement of membrane and MEA lifetime, in
of the macro and microporous layers due to the oxidation of particular since the ground-breaking work introducing effective
the carbon and the decomposition of the PTFE component. radical scavengers by Endoh et al. at Asahi Glass Company and
These microstructural changes affect the gas transport and by Coms et al. at General Motors.
therefore lower the cell performance.20
Degradation of the membrane is one of the major reasons 2. PFSA polymer types
that can lower the durability of a fuel cell. Membrane
decomposition can be induced by many factors for instance The fuel cell electrolyte must have high proton conductivity in
mechanical degradation which is an effect of non-uniform order to maximize the efficiency. It should also be mechanically
contact pressure or fatigue from stress during shrinkage and chemically robust during fuel cell operation to ensure
and swelling of the membrane during humidity cycling. a long lifetime and nally needs to be impermeable to the direct
Reduction of the membrane mechanical strength results in transfer of reactant gases and electronically insulating.
the formation of pinholes and cracks, which can further Although many types of FC membranes have been developed
contribute to the failure of the MEA.21 Another factor affecting over the past two decades32–35 the commonly used proton
the stability of the membrane is the contaminant ions origi- exchange membranes are peruorosulfonic acid (PFSA)
nated from corrosion of the different cell components.22 membranes. The rst widely used PFSA membrane is Naon®
Released cations such as Fe and Ni catalyse the decomposi- developed by DuPont.36 The synthesis of Naon® is based on
tion of hydrogen peroxide, which is the main source of free the copolymerisation of tetrauoroethylene (TFE) with a per-
radicals. Moreover the counter ions formed tend to associate uorinated vinyl ether co-monomer and sulfonyl acid uoride.
with the sulfonic acid sites of the polymer, which causes The product obtained from the copolymerisation is hydrolysed
partial drying of the membrane and results in loss of with NaOH and then converted to the acid form. The Naon®
conductivity.23,24 Chemical degradation crucially affects polymer structure comprises two well dened components: the
membrane durability. Chemical decomposition is initiated by hydrophobic backbone (polytetrauoroethylene – PTFE) and
free radicals, highly reactive oxygen species, which are formed the hydrophilic side chain terminated with sulfonic acid
during the chemical and electrochemical reaction of cross- groups. The PTFE part provides mechanical strength and rela-
over gases. Those species attack the polymer structure tively high stability in very harsh chemical environments, while
causing an “unzipping” reaction on the main chain and the the conduction properties are due to the presence of hydrated
scissions of the side chain, which nally result in thinning of sulfonic acid groups.37 Commercially available polymers with
the membrane. equivalent structure are Flemion® (trade mark of Asahi Glass
All these damaging effects oen occur simultaneously, Company), Aciplex® (produced by Asahi Kasei) and Fumion® F
thereby creating a very complex mechanism of the overall fuel (developed by FuMA-Tech). This type of composition is now
cell degradation.25 Since amelioration of durability is crucial for known as the long-side-chain (LSC) structure; correspondingly
technology development, understanding the fundamental ionomers with a shorter pendant side chain are referred as to as
degradation mechanisms of the fuel cell components is short-side-chain (SSC) ionomers. The chemical structures of
primordial. PFSAs with different side chain lengths are presented in Fig. 1.

410 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

degradation may be distinguished: chemical, mechanical and


thermal.24

3. Mechanical and thermal


degradation of PFSA membranes
Mechanical degradation is an effect of the mechanical forces
affecting the integrity of the membrane. Long term fuel cell
testing shows a time dependent deformation – creep and micro-
crack fracture of Naon®. These phenomena are related to the
Fig. 1The chemical structure of LSC – Nafion® structure (a) and SSC repeated stress applied to the membrane between bipolar plates
perfluorosulfonic ionomer membranes (b) and (c). Reprinted from and strain from local changes of the relative humidity. The most
Polymer Electrolyte Membrane and Direct Methanol Fuel Cell Tech-
important factors inducing micro and macro defects of the
nology, vol. 1, D. J. Jones, Polymer Electrolyte Membrane and Direct
Methanol Fuel Cell Technology, 27–56.38 Copyright (2012), with membrane are: relative humidity cycling,36,50,51 transients at
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

permission from Elsevier. open circuit voltage (OCV),52 temperature cycling51 and poten-
tial cycling. Some of those stressors have a bigger impact on the
polymer mechanical durability than others, for example cycling
of the temperature was reported to have less inuence on
In the SSC polymer structures there is no –O–CF2CF(CF3)– membrane breakdown than RH cycling50,53 probably due to the
group. Furthermore, the length of the peruoro vinylether side higher strain oscillation amplitude.54 Repeated swelling and
chain varies between 2 and 4 carbons for the two commercially shrinking of the membrane causes the gradual reduction of the
available SSC ionomers i.e. Aquivion® and 3M™ respectively. membrane strength that nally results in membrane dimen-
Pioneering studies on the synthesis of a SSC polymer date back sional changes, poor interface contact between membrane and
to 1981 at Dow.39 Early investigations of Dow SSC membranes electrode,55–57 and defects such as pinhole formation.57,58
demonstrated excellent balance between their stability and Pinhole formation, due to the local stress applied to the
transport properties.40–42 However a complex and expensive membrane is one of the main reasons for the failure of PFSA
synthesis route impeded their wide commercialisation. More membranes in PEM fuel cells.59 Accelerated gas crossover of
recently, a low cost synthesis route developed by Solvay Solexis hydrogen and oxygen to the opposite side of the electrolyte
resulted in new benchmark Hyon® Ion membrane,43 through pinholes causes combustion reaction on the catalyst
renamed in 2010 as the Aquivion® membrane. The parameter surface and generates local hot spots. The use of high temper-
that reects the transport properties of the PFSA membrane ature and relative humidity increase gas crossover.60–62
and its mechanical integrity, within a given class of PFSA Membrane hydration is a very important factor affecting
types, is the equivalent weight (EW). EW is dened as dry mass mechanical and thermal properties of PFSA membranes.31,63,64
of the polymer in the acid form in grams per mole of Kundu et al. reported that the transition temperatures of
exchangeable groups. Another important feature of PFSA Naon® decrease with increasing membrane hydration.65
membranes in relation to the EW is the ion exchange capacity Majsztrik et al. described that high hydration of Naon® at low
(IEC). The relationship between these parameters is expressed temperature plasticises its structure,66 whereas at elevated
by: 1000/EW. The IEC of the membrane can be measured temperature opposite behaviour is observed.67 This implies that
experimentally by acid/base titration. Differences in side chain at elevated temperature for a highly hydrated membrane,
length leads to differences in the molecular weight of the mechanical creep occurs much faster than for a dry polymer.
polymer repeat unit and translates into specic properties Recently Luo et al. reported signicantly higher permeability of
such as higher degree of crystallinity of SSC membranes44 degraded Naon NR211 following 48 h exposure to Fenton
compared to LSC Naon® for a given polymer equivalent reagent compared to the pristine membrane.68 Increased water
weight, and higher glass transition temperature (Tg).45 Higher content and proton mobility, despite lower IEC and proton
ion exchange capacity46 SSC polymers can be used to higher conductivity of degraded membranes, were explained by the
temperature due to these properties.47–49 In general PFSA authors as resulting from enhanced water sorption of
membranes possesses good mechanical properties, high a damaged polymer membrane structure. Water remaining in
proton conductivity and electronic resistance and low gas non-ionic cavities formed through degradation does not
permeability. All these properties are necessary in order to contribute to ion conduction but can cause hygrothermal stress
have an efficient barrier for gas separation and to withstand and fracture development.69 Also Shi et al. interrogate hetero-
strong oxidative and reductive environments and severe geneity in permeability of ex situ aged Naon membranes.70
mechanical stresses imparted by hydration–dehydration Venkatesan reported high water uptake of degraded catalyst
cycles compounded by the specic electrochemical environ- coated membrane (CCM) subjected to combined mechanical/
ment of the fuel cell. Such harsh conditions nevertheless chemical accelerated stress test.71 Authors demonstrated that
initiate decomposition or degradation of the ionomer or voids regions in severely degraded material, depleted of F and C
membrane, of which three main types of membrane are susceptible to micro crack formation. PEM fuel cells require

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 411
View Article Online

Sustainable Energy & Fuels Review

to be operated over a wide working temperature range. There- materials or humidiers, catalyse the decomposition of H2O2 to
fore it is important to study the stability of PFSA membranes at produce hydroxyl radicals as shown in eqn (1).93,94 Pozio et al.
sub-freezing temperature. The existence of three72,73 or even four performed experiments where stainless steel plates were
different “states” of water in Naon® termed non-freezing replaced with iron-free aluminium alloy plates.89 The
water, bound freezing water and free water has been re- membrane degradation aer 1200 h under such conditions was
ported.74 The chemically bound water does not freeze down to signicantly reduced, which clearly shows correlation between
120  C, whereas it is probable that the water which is non- uoride emission rate and trace amounts of transition metal
bonded to the polymer chain freezes below 0  C.73 Phase ions. Moreover the amount of iron can be as low as 1 part per 5–
transformation as well as difference in densities of water and ice 25 parts of H2O2 for the reaction to occur for a concentration of
can cause the degradation of PFSA membranes.75 However hydrogen peroxide of <10–25 mg L1.95
McDonald et al. reported that aer 385 cycles in the tempera-
ture range from 40  C to 80  C no catastrophic failure was H2O2 + Mz+ / M(z+1)+ + HOc + OH (1)
observed.76 On the other hand, Plazanet et al. observed ice
formation outside of the membrane with simultaneous increase The hydroxyl radical produced via eqn (1) can further
of concentration of hydronium ions in membrane.77 This undergo reaction (eqn (2)) with hydrogen peroxide to form
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

phenomenon was related to desorption of water upon cooling hydroperoxyl radical (more detail in Section 7).
and contraction of the membrane. In such a scenario, the
HOc + H2O2 / H2O + HOOc (2)
formation of ice or frost at the membrane electrode interface is
detrimental for the catalyst layer structure. The impact of M(z+1)+ + H2O2 / Mz+ + HOOc + H+ (3)
mechanical degradation can be signicantly decreased by
optimizing the design of the fuel cell components i.e. by Furthermore, the hydroperoxyl radical can also be produced
choosing appropriate materials and operating conditions.78 On by the reaction of transition metal ions with hydrogen peroxide,
the other hand, chemical degradation appears more complex eqn (3). In situ hold at open circuit voltage (OCV hold
and difficult to mitigate as the process has a larger impact on testing)96–98 is the accepted means to accelerate membrane
the integrity of the whole polymer.79 chemical degradation, while outside of the fuel cell environ-
ment, the Fenton reagent comprising hydrogen peroxide and
ferrous ions is frequently used.70,99–102 In both cases accelerated
4. Chemical degradation of PFSA membrane decomposition is due to high concentration of
Chemical degradation of the PFSA membrane is the major issue hydroxyl radicals that attack the most vulnerable sites in poly-
considered in this paper and it will therefore be discussed mer structure: carboxylic acid end groups; C–S bonds; tertiary
starting from the structure of the membrane at the microscopic carbon atoms and ether groups. These four main mechanisms
level. The mechanism of membrane degradation at the meso of radical attack on the polymer structure are discussed. The
and macroscopic levels is also fully described followed by the rst, generally referred to as the “unzipping reaction” is
classication of free radicals generated under fuel cell operating a radical reaction occurring in non-so-called “stabilised” PFSA
conditions. Finally a summary of mitigation strategies that have polymers, and starting on weak terminal –COOH groups of the
been proposed is discussed at the end of this review. polymer main chain (Fig. 2). Such terminal carboxylic acid
groups can also be generated through transformation of non-
peruorinated groups via hydroxyl radical attack.103 As
4.1 Radical attack on polymer structure
described in eqn (4)–(6) end carboxylic groups react with
Chemical degradation of PFSA membranes involves radical- hydroxyl radicals to produce CO2 and HF, and reform a terminal
induced decomposition of the polymer structure. The process –COOH group.104 Several propagation steps of this reaction
of chemical degradation is not yet fully understood, however the brings about decomposition of the uorocarbon main chain by
damage caused to the membrane and its properties are evident, consistent loss of CF2 units.105 When the polymer unzipping
for instance a decrease in the ion exchange capacity followed by reaches a junction with a side chain, cleavage of the overall side
conductivity losses, as well as uoride emission and subsequent chain groups generates peruoro(3-oxa-5-methyl)pentane-1-
membrane thinning (subsection 3.2). The current generally sulfonic-5-carboxylic diacid, HCOO–CF(CF3)–O–CF2CF2–
accepted mechanism for chemical degradation of per- SO3H.105 This species can diffuse out of the membrane or can
uorosulfonic acid polymers is by radical attack on the polymer undergo decomposition by the unzipping reaction, as a carbox-
main chain and side chain.80–85 Reactive oxygen species (ROS) ylic acid end group is also present in its molecular structure.
can be generated directly via chemical or electrochemical This mechanism was the rst to be identied, following seminal
reaction of crossover gases over the electrocatalyst surface,86–88 work at General Motors and DuPont, and this greater under-
however their short lifetime limits to a large extent their diffu- standing of what was understood at that time to represent
sion length.80 Another pathway for radical formation occurs a primary cause of polymer degradation led directly at DuPont
through homolysis of hydrogen peroxide (H2O2).23,89 H2O2 itself (and other producers of PFSAs) to post-uorination of terminal
is not a strong enough oxidant to damage the polymer structure carboxylic acid groups with the aim of stabilising them against
substantially.90–92 However, iron or other multivalent metal ions the unzipping reaction.
like Cu2+ or Ti3+, originated from corrosion of the cell or stack

412 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

rst experimental conrmation of the unzipping mechanism.108


Interestingly, for Aquivion® and 3M™ ionomer dispersions, the
magnetic parameters of the DMPO/CCR adduct were identical
and also indicated an unzipping process whereas for Naon®
dispersion those parameters were totally different which could
mean rather side chain cleavage. It was reported that the low
concentration of hydrogen containing terminal groups in PFSA
membranes is disproportionate to the observed level of released
Fig. 2 Unzipping mechanism – hydroxyl radical attack on carboxylic HF,109,110 moreover Naon® membranes that have been chem-
acid groups. ically stabilised by post-uorination of the carboxylic groups
show reduced, but not eliminated, mass loss (see Fig. 3).104,111
Other studies demonstrated that Naon® converted to the
4.2 Mechanism I – “unzipping mechanism” Na+, Cs+ or Li+ forms shows two orders of magnitude lower
uoride emission rate compared to the acid form membrane.
Rf–CF2COOH + HOc / Rf–CF2c + CO2 + H2O (4) Such a signicant change in the FER values suggest that radi-
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

cals also attacked the sulfonic acid groups (Fig. 4).91


Rf–CF2c + HOc / Rf–CF2OH / Rf–COF + HF (5)

Rf–COF + H2O / Rf–COOH + HF (6) 4.3 Mechanism II – attack on C–S bond


A second possible mechanism of membrane degradation is the
attack by free radicals on C–S bonds following the reactions
From a practical point of view some conclusions can be drawn shown below:113
regarding the decomposition model described above. Experi-
mentally, the signicant mass loss of the membrane (the main Rf–CF2SO3H + HOc / Rf–CF2c + SO3 + H2O (7)
degradation product of the unzipping mechanism is HF) can be
correlated to the experimentally determined uoride emission Rf–CF2c + HOc / Rf–CF2OH / Rf–COF + HF (8)
rate (FER). Fluoride emission can be measured using an ion
selective electrode (ISE) or by high performance liquid chro- Rf–COF + H2O / Rf–COOH + HF (9)
matography (HPLC) of exhaust water. The FER depends on the
The radical attack begins at the end of the side chain on the
in situ experimental conditions known to cause macroscopic
C–S bond (eqn (7)). Aer this step the polymer decomposition
membrane degradation such as relative humidity, temperature,
gas type, ow and pressure and cell voltage. Moreover FER is
a good method for investigation of the kinetics of the radical
reaction and it is a useful tool to estimate the life time of the
fuel cell. For a “pure” “unzipping” mechanism i.e. in the
absence of any other degradation mechanisms, the FER should
be constant during the stress test without any signicant
acceleration, and the ratio of –COOH groups to CF2 should
remain unchanged. This observation has been made for ex situ
accelerated ageing i.e. using the aqueous Fenton test.106
Schwiebert et al. questioned the possibility of radical attack
other than on the terminal carboxylic acid groups. Their studies
of sulfonic acid and carboxylic acid functionalised model
compounds such as CF3CF(COOH)OCF2CF(CF3)OCF2CF2SO3H
Fig. 3 Fluorine emission rates as a function of the content of
and CF3CF(COOH)(CF2CF2)nSO3H support the unzipping carboxylic groups in Nafion® structure. Reproduced by permission of
degradation mechanism. Moreover these authors suggested The Electrochemical Society from ref. 112.
that ionomers with a linear side chain (Aquivion®, 3M™ ion-
omer) have similar stability to that with a branched per-
uoroether side chain (Naon, Fumion, Flemion).107 However
using a spin trapping ESR technique and 5,5-dimethylpyrroline-
N-oxide (DMPO) as a spin trap for unstable species, Danilczuk
et al. provided different evidence regarding the stability of LSC
and SSC membranes against radical attack, and concluded that
3M and Aquivion® polymers were signicantly more stable than
the Naon® polymer. Furthermore a DMPO/CCR (carbon cen-
tred radical) adduct formed through hydroxyl radical attack on
the carboxylic acid groups was identied, which constituted the Fig. 4 Hydroxyl radical attack on C–S bond.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 413
View Article Online

Sustainable Energy & Fuels Review

proceeds with progressive shortening of the side chain and HF This author pointed out the weakening effect of sulfonyl
emission (eqn (8) and (9)). Eqn (8) and (9) seem to be a repeti- radicals on the C–S bond. In other words, SO3c facilitates the
tion of the steps in the unzipping mechanism described in eqn C–S cleavage and formation of the peruororadical as shown in
(5) and (6). This conclusion is partially right, the incorporation mechanism II.
of a new weak –COOH group accelerates the degradation and
therefore the unzipping reaction occurs on both the side and 4.4 Mechanism III – attack on ether groups
main chains. However as the Rf have different chemical struc- The above discussion brings us to a third mechanism which
tures, the decomposition products of the side chain have concerns hydroxyl radical attack on the O–C bond.
different nature. The detection of –O–CF2–CF2–SO3c and –O– It was postulated that ether groups next to a C–S bond
CF2–CF2c radicals aer a UV – induced Fenton test supports the represent the preferred site for the radical attack in the side
above degradation mechanism.114 This type of membrane chain structure.120,121 This nding might support the above
decomposition results in the simultaneous loss of weight and results of Danilczuk et al. where Naon® ionomer exhibited 20
conduction properties of the polymer. Recently Kurniawan et al. times lower stability than respective SSC ionomers.108 This is
studied the PFSA degradation mechanism using quantum probably due to the absence of OCF2CF(CF3) in the SSC polymer
chemical calculations (QCC) based on density functional theory structure. On the other hand, according to studies of Ghas-
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

(DFT).115 Triuoromethanesulfonate (TFMS) was used as semzadeh et al. the rst point of HOc attack is located near the
a model of the side chain structure. For these authors, the sulfonic acid group (a-OCF2) rather than near the main chain (b-
calculated bond dissociation energies (BDEs) and bond lengths OCF2).118,119 Aer cleaving the a-OCF2 the polymer decomposi-
indicate that radical attack on the C–S bond is favourable due to tion continues as shown in Fig. 6(a) and involves the b-OCF2
its lower dissociation energy in comparison to the other struc- and CF units. These results are in agreement with theoretical
tural bonds. They further pointed out the behaviour of uorine studies of Ishimoto et al.122
atoms as another important factor that can affect the radical
attack on the polymer structure, i.e. the electronegative uorine 4.5 Mechanism IV – attack on the tertiary carbon
atoms efficiently shield the carbon backbone from radical
reaction. Tokumasu et al. also investigated the BDE of a per- A further potential site for radical attack leading to PFSA poly-
uorosulfonic acid (PFSA) molecule by DFT.116 Similarly to the mer degradation can occur at the tertiary carbon according to
work of Kurnawian et al. the C–S bond appeared as the weakest the reaction given in Fig. 7.117 Two tertiary carbons are present
bond in the side chain structure of the neutral molecule, thus in the Naon® structure; both are adjacent to ether bonds.
most susceptible to hydroxyl radical attack. However further Radical attack on the CF of the side chain initiates side chain
analysis showed that the ionisation of the sulfonic group makes decomposition, whereas attack on the main chain CF results in
this C–S bond stronger while the C–O bond becomes weaker. chain scission. This mode of degradation is characterised by
From thermochemical and kinetic analysis, Coms deduced relatively low uoride emission but with an important increase
further possible route of side chain scissions.117 The mecha-
nism is initiated by sulfonyl radicals, which are formed under
dry conditions, where the sulfonic acid groups are not fully
ionised. Sulfonyl radicals can be formed directly through
hydrogen abstraction from –SO3H by hydroxyl radical or indi-
rectly through reaction with hydrogen peroxide where bissul-
fonyl peroxide is an intermediate product, which further
undergoes homolysis to give sulfonyl radicals (Fig. 5).

Fig. 6 Mechanism of side chain scission by HOc radical attack on C–O


bond (a) direct attack Reprinted with permission from (J. Am. Chem.
Soc., 2013, 135, 15923–15932).118 Copyright (2013) American Chemical
Society. (b) Attack as a consequence of bond weakening after C–S
bond cleavage. Reprinted with permission from (J. Am. Chem. Soc.,
2013, 135, 8181–8184).119 Copyright (2013) American Chemical
Fig. 5 Mechanism of sulfonyl radical formation. Society.

414 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

bands were observed at depths of 22 mm and 82 mm from the


cathode. The presence of the bands at 22 mm from the cathode
was associated with the location of Pt particles precipitated in
the membrane, however larger damage was observed near the
anode interface at 82 mm where these bands showed highest
intensity.

Fig. 7 Mechanism of chain scission by radical attack at the tertiary


carbon. Reprinted with permission from (J. Am. Chem. Soc., 2013, 135,
15923–15932)118. Copyright (2013) American Chemical Society. 4.6 Summary of radical attack on PFSA type polymers
A graphical summary of degradation mechanisms proposed in
the literature is provided in Fig. 8.
in COOH groups that can then be degraded by the unzipping The “unzipping” mechanism, in this work labelled as
mechanism. The vapour phase Fenton test, OCV hold test and mechanism I, proposed by Curtin et al.104 and Healy et al.124 is
low power demand fuel cell test were reported to accelerate the most frequently cited origin of membrane degradation.
Naon® degradation according to this mechanism.109 Tertiary However it should be emphasised that the main role of the
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

carbons atoms are assumed to be most fragile to radical attack unzipping mechanism in chemical decomposition of PFSA
among the uorocarbons, due to different thermodynamic ionomers changed with the evolution in the ionomer post-
stabilities.117 Strong dependence of the C–F bond dissociation uorination process. A new generation of chemically post-
energies on abstraction of the uorine atoms, gives the treated PFSA membranes with reduced number of COOH
respective order: tertiary > secondary > primary, with tertiary groups show considerably improved durability.108,109 Moreover
carbon as the least stable position.116,117 this improvement signicantly increases with elimination of
Schiraldi remarked that thermodynamic assumptions the –OCF2CF(CF3) fragment from the polymer structure.108 This
should be re-evaluated in relation to the complex morphology of behaviour clearly shows the contribution of radical attack on
Naon®, which is likely to increase the kinetic factor in the the side chain to overall membrane decomposition. It can be
degradation process.111 According to studies using model concluded that all the mechanisms discussed above most
compounds, in ex situ accelerated degradation tests a carboxylic probably play a role in the chemical degradation of the
acid terminated compound demonstrates 300 times faster membrane during fuel cell operation, where the working
generation of uoride than the corresponding sulfonic acid
terminated compound.111 However in their studies on Naon®
degradation by applying in depth proling by micro FTIR,
Danilczuk et al. recently proposed mechanism III as a degrada-
tion pathway of membrane decomposition.123 The micro FTIR
spectra of Naon® 115 membrane cross section were recorded
aer OCV hold testing for 180 h at 90  C and 30% RH. Two
bands C]O and C–H characteristic of the degraded membrane
appeared at the same prole depths which indicated their
generation through a single mechanism, which is uorine atom Fig. 9 Degradation of the Nafion® structure by radical attack and
abstraction from tertiary carbon atom by Hc. C]O and C–H unzipping of the side and main chains.

Fig. 8 Summary of the mechanisms of radical attack on the Nafion® polymer structure.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 415
View Article Online

Sustainable Energy & Fuels Review

conditions determine the contribution of each of them. Very platinum at the anode, cathode or present in the membrane
likely, a membrane weakened by one type of radical attack is following degradation of the electrode catalyst layer are all
more susceptible to undergo a second type of radical reaction.119 possible.60,86,136
According to recent reports, hydroxyl radical attack on the
susceptible bond in the side chain initiates “unzipping” along H2 / 2Hc (12)
the side chain110,117,125 and then main chain structure126,127 as
displayed in Fig. 9. Therefore the term “unzipping” cannot be Hc + O2 / HOOc (13)
removed in the discourse of degradation; however, it must be
HOOc + H+ + e / H2O2 (14)
approached with greater attention.
Finally two possible sites of the radical attack in the side Regardless of the exact location of hydrogen peroxide
chain; C–S bond110,113,117,128 and a-OCF2 unit118,119 are suggested formation, the platinum catalyst is a constant factor in that
in the literature. There is no agreement concerning the domi- process. Pt or Pt alloy is considered as being the best catalyst for
nance of the radical attack on one of these sites however in both the hydrogen oxidation reaction and the cathodic oxygen
scenarios the whole moiety OCF2CF2SO3 is lost. According to reduction reaction for PEMFC application.137 However, besides
Ishimoto et al. the impact of the relative humidity on attack of the four-electron pathway for ORR leading to production of
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

hydroxyl radical on either ether or sulfonic sides has a great water as displayed in eqn (11), a partial two-electron reduction
importance.129 Based on DFT calculations these authors reaction to H2O2 also occurs (eqn (10)). Many experiments were
concluded that HOc attack the ether group under high humidity developed to simulate hydrogen peroxide generation on Pt
conditions whereas sulfonic acid groups are more susceptible to particles.138–140 Ex situ evaluation of this mechanism using
radical attack at low relative humidity. a rotating ring disc electrode (RRDE) demonstrates an increase
in H2O2 formation with decreasing electrochemical surface area
(ECSA) of the catalyst.138,141 ECSA corresponds to the catalyst
5. Hydrogen peroxide formation in active fraction for the electrode reactions and its decrease is due
membrane-electrode assemblies to platinum nanoparticles agglomeration, dissolution or inac-
cessibility. Therefore the elevated formation of H2O2 was
(MEA) explained by the lower probability of further reduction reaction
The understanding of the different possibilities for radical of hydrogen peroxide re-adsorbed on the sparsely distributed
attack at weakest points in the polymer chain is essential in active sites of the Pt particles. Moreover faster decrease in ECSA
order to decrease membrane decomposition by chemically and further increase of the H2O2 fraction was reported for
stabilising the vulnerable groups in the polymer structure. catalysts with low Pt loading.138,141,142 Adsorption of anions such
However it is also important to identify the fragile points of the as Cl, Br or of CO on the catalyst surface was reported to
complete membrane electrode assembly (MEA) with regard to accelerate H2O2 formation.139,143 This is in agreement with DFT
chemical degradation on a macroscopic level. For that reason calculations where the generation of H2O2 and free radicals is
many researchers have attempted to localise the formation of expected to occur when the ORR is blocked by an adsorbed
highly oxidative species in an operating fuel cell. As stated species. Four-electron ORR is kinetically promoted for oxygen
above, it is believed that the main source of hydroxyl radicals is molecules adsorbed at two Pt sites, in which case any hydrogen
hydrogen peroxide, which is generated in the fuel cell and peroxide or hydroxyl radical generation is energetically unfav-
which has been observed in the membrane,87 exhaust water ourable. However if O2 bonds with only one Pt site, and acces-
from the anode and cathode sides and in cathode outlet sibility to other active sites is hindered by other adsorbed
gas.60,130 The concentration of the H2O2 depends largely on the species, both H2O2 and reactive oxygen species can be easily
operating conditions as well as on the membrane thickness. generated.144,145
During in situ experiments 8–10 mg cm3 of hydrogen peroxide H2O2 generation during PEMFC operation can be affected by
was determined in the membrane aer 2 h of operation of the additional factors impossible to mimic ex situ. One such factor
fuel cell at 65  C with H2/O2 as reactants.131 H2O2 can be formed that needs to be considered regarding H2O2/radical formation
within an operating fuel cell at the cathode side by direct and degradation mechanism is the relation of the gas crossover
generation during the oxygen reduction reaction (ORR) when to membrane thickness and gas partial pressure.146,147 High
the potential is below 0.696 V. oxygen partial pressure leads to high oxygen crossover through
the membrane, which elevates the hydrogen peroxide genera-
O2 + 2H+ + 2e / H2O2 (10) tion at the anode catalyst according to LaConti. This nding was
conrmed experimentally by many research groups,60,113,148,149
O2 + 4H+ + 4e / 2H2O (11) which have reported negligible membrane degradation in the
absence of O2 crossover. The inuence of membrane thickness
Formation of H2O2 at the anode side was rst described by on gas crossover was investigated by Chen et al., who studied
LaConti et al.132 who proposed that crossover oxygen can react different membranes with various thicknesses ranging from
with hydrogen which was chemisorbed on the Pt catalyst as 150 to 280 mm and revealed that with the increase in the
displayed in eqn (12)–(14).133,134 Another mechanism is the membrane thickness the concentration of H2O2 in the
chemical combination of crossover gas,135 where reaction on

416 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

membrane decreases.131 Also relative humidity plays an the ferrous (Fe2+) at OCV condition also differ from those in
important role in the formation of hydrogen peroxide: normal working mode.159 At OCV, hydrogen peroxide produc-
Sethuraman et al., described a decrease of the concentration of tion at the cathode side can be neglected due to high cathode
H2O2 in high water activity conditions (i.e. high humidity) in potential and lack of faradic current.80
studies of the oxygen reduction reaction on the RRDE.150 Chen
et al. reported an increase of H2O2 emission with decrease of
relative humidity on both anode and cathode side, which he 6. Gas crossover and membrane
explained by a low membrane water content and an increase of degradation
gas partial pressure that led to higher gas crossover.131 Moreover
higher concentration of H2O2 was observed on the anode site. Gas crossover is an inevitable process in an operating PEMFC
For those reasons low humidity conditions are commonly used as typically applied electrolyte membranes have a thickness of
to accelerate the polymer degradation rate.78 20–50 mm. Greater thickness could decrease the permeation of
Decomposition of hydrogen peroxide is one of the possible the reactants across the membrane,87,131 however it simulta-
mechanisms for radical formation in an operating fuel cell. neously increases the membrane area specic resistance. It
Taking into account only this route to generate reactive oxygen should be mentioned that besides the fact of the inevitability
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

species (ROS), already three different possible areas of chem- of gas crossover, this process is much more pronounced under
ical attack need to be considered: anode side, cathode side and open circuit conditions as there is no consumption of H2 and
membrane bulk. Studies carried out in this area show con- O2 through the electrochemical reactions.60 The key role of gas
icting conclusions. Some researchers propose propagation of crossover in membrane degradation was demonstrated in
the chemical attack from the anode side towards the series of ex situ experiments. Independently, using different
membrane to the cathode side.151 Others observed that the experimental set-ups and methods of evaluation of membrane
predominant degradation is at the anode103,123,152–154 or the degradation, several researchers drew the conclusion that the
cathode side,89,98,155,156 while others, did not nd any signi- polymer decomposition occurs only when O2, H2 and Pt
cant difference between anode or cathode membrane–elec- coexist.113,135,148 That initiated further studies on hydrogen
trode interfaces in term of chemical degradation.157 It is very crossover investigated in an operating fuel cell. Inaba et al., in
difficult to distinguish the degradation phenomena as those their work on membrane degradation at OCV, demonstrated
arising from mechanical and chemical processes respec- the incontestable increase of hydrogen crossover with simul-
tively.25 To facilitate the analysis of areas of membrane taneous drop of OCV and increase of uoride emission.60 It
degradation, a bilayer catalyst coated membrane method was was pointed out that measurement of the hydrogen perme-
proposed.98,156 With such a conguration, post-test analysis ation across the electrolyte might be a reliable indicator of
could be performed on the detachable membrane and the membrane degradation.60,160 Other authors have investigated
MEA centre could be easily identied. This method, resulting the inuence of cell temperature, humidication (refer to
in membrane thickness much higher than usually applied, can Fig. 10) and gas pressure on hydrogen crossover. Clearly the H2
be replaced by using commercially available thin reinforced crossover current density increases with increasing tempera-
PFSA membranes such as Naon XL or GORE-SELECT. The ture (from 40 to 80  C), relative humidity (here from 40 to 80%)
middle of those membranes is clearly determined by the and (not included on Fig. 8) gas pressure. This nding was
ePTFE reinforcement. Still the difficulty in obtaining answers further conrmed by Baik et al. in similar range of tempera-
to the questions on where exactly the chemical attack ture and relative humidity.161 Moreover a particular design of
occurred, and which mechanism dominates in PEM the bipolar plate allowed the measurement of the hydrogen
membrane failure, is due to the morphology of the Naon® crossover rate over a local area. The local variation of crossover
membrane, the short life time and the low concentration of
the radical species, but also, more importantly, due to the
different conditions applied for accelerating the degradation
process. Accelerated stress tests (AST) are commonly used in
order to avoid thousands of hours of in situ life test to deter-
mine membrane durability; however that usually introduces
some incertitude as to the relevancy of the methods employed.
As an example, open circuit voltage (OCV) hold test is,
currently, the most applied in situ stress test to assess the
chemical stability and provide better understanding of fuel
cell failure mechanisms. It was reported that OCV hold
accelerates to greater or lesser extent, depending on the
conditions used, such parameters as ECSA losses, voltage
Fig. 10 Effect of temperature and relative humidity on H2 crossover
decay and gas crossover in comparison to long duration
current density at atmospheric pressure. Reprinted from (vol. 51, M.
operation.158 The degradation rates calculated from acceler- Inaba et al., Gas crossover and membrane degradation in polymer
ated stress tests tend to be higher than those obtained from electrolyte fuel cells, 5746–5753).60 Copyright (2006), with permission
a lifetime test. Moreover the H2O2 and concentration prole of from Elsevier.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 417
View Article Online

Sustainable Energy & Fuels Review

(higher H2 crossover was measured near the inlet of the gas ux vs. the in situ radical formation rate for all conditions
than near the outlet) was attributed to the partial pressure investigated.
gradient of H2. This experiment validates the assumption of a high
As stated above, high temperature162 and gas pressure lead to concentration of reactive oxygen species at low relative humidity
elevated crossover rates and accelerate membrane degradation. and conrms the role of the high temperature and cathode
For this reason, those parameters are typically applied in AST. potential as stressors for membrane degradation.
However the inuence of relative humidity on the mechanism The mechanism of the membrane decomposition depends
of membrane chemical decomposition is much more complex. largely on the experimental conditions; however it always
According to the previous discussion, H2 permeation increases results in scission or cross-linking133 of polymer chains. Recent
with increasing relative humidity. This observation was inter- publications in this eld show that attention is increasingly
preted using a multi-structural model of Naon®, comprising paid to chemical analysis of exhaust gas and water, since this
an ion cluster region, an interfacial region and a uorocarbon analysis can provide information about the presence of three
phase.163 Based on the assumption that H2 permeates through general groups of species considered from the perspective of
exible, amorphous interfacial uorocarbon regions,164 it was membrane degradation. The rst group consists of products of
proposed that water absorbed by the membrane increases the membrane decomposition such as uoride ion,169,170 sulfate ion
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

size of the ion clusters, plasticizes the interfacial regions60 and or polymer fragments.109,113,124,170 The second group are pollut-
enlarges the free volume165 where gas permeability can occur. ants171,172 which act as degradation catalyst. Contaminants in
Keeping in mind the relationship between H2 crossover and a fuel cell set up originate from gas or water supplied to the
membrane degradation, the observation that low relative system, from the cell hardware or the alloy catalysts. Gaseous
humidity166 or humidity cycling167 accelerates the polymer impurities such as HCOOH, CO, NO2, SO2, H2S, Cl2 can cause
decomposition to larger extent than high relative humidity is blocking of the catalyst active sites, affect the ORR mechanism
rather surprising. This discrepancy clearly shows that for the or accelerate Pt dissolution.173,174 Hydrated cations such as Fe2+,
mechanism of radical formation and membrane degradation, Fe3+, Cu2+ or Al3+ accelerate membrane degradation and cause
the degree of humidication itself is one of the critical factors. performance losses.175,176 Also certain degradation compounds
Inaba et al. argued that it might be related to a higher such as: in case of Naon – peruoro(2-methyl-3-oxa-5-sulfonic
concentration of hydrogen peroxide under low RH,60 no exper- pentanoic) acid and 3M membranes – peruoro(4-sulfonic
imental conrmation was provided. Some more light on this butanoic) acid may migrate and adsorb into the catalyst layer
issue is given by research performed recently by Prabhakaran leading to a loss of ECSA and ORR activity.177 The last group of
et al.168 These authors investigated formation of the reactive species consists of free radicals, which are initiators but oen
oxygen species (ROS) in an operating fuel cell using in situ also products of chemical degradation or of the decomposition
uorescence spectroscopy. Again a clear correlation between of hydrogen peroxide.
generation of ROS and PFSA membrane degradation was
demonstrated under different fuel cell operating conditions.
ROS formation was found to be enhanced at lower relative
humidity (50% > 75% > 95%), elevated cell temperature (100  C
7. Influence of Pt band on membrane
> 80  C > 60  C > 40  C) and nally at higher cathode potential degradation
(0.8 V > 0.6 V > 0.4 V). An excellent graphical summary of these
The above discussion concludes that there are different
results is shown in Fig. 11, which plots the uoride emission
contributions of the cathode and anode catalyst to the overall
chemical degradation, however recently more attention has
been focused on the inuence of platinum in the membrane
(PTIM) on the acceleration/diminution of its chemical degra-
dation.136 One of the rst reports concerning Pt band formation
in PFSA membrane appears in 2005, when Ferreira observed the
existence of a Pt band in the membrane aer 2000 h of OCV
hold test,12 and further studies revealed that a Pt band can be
formed already aer the rst 50 h of OCV hold.136 The Pt catalyst
is unstable in the acid environment of the fuel cell at high
voltage and can dissolve at the cathode due to electrochemical
oxidation to Ptz+ ions178,179 as displayed in eqn (15):180

Pt / Pt2+ + 2e (15)

Then Pt ions migrate into the membrane due to the


Fig. 11 Correlation between in situ reactive oxygen species ROS
concentration gradient and undergo reduction in the presence
formation rate and membrane degradation expressed as fluoride of the crossover hydrogen according to the eqn (16).
emission rate. Reproduced from ref. 168 with permission of the PCCP
Owner Societies. Pt2+ + H2 / Pt0 + 2H+ (16)

418 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

Pt atoms formed on that way (Pt0) can further diffuse and across the membrane.184 These results were interpreted as
agglomerate in the membrane in order to minimise their high formation of electronic short-circuits due to accumulation of Pt
surface energy forming Pt particles.180 particles. Studies of Liu and co-workers revealed an increased
degradation when a membrane exposed to H2O2 solution was Pt
(Pt)n + Pt0 / (Pt)n+1 (17) catalyst coated in comparison to a bare membrane in an H2O2
ow cell experiment.135 Hydrogen peroxide formed at the elec-
Pt ions can also precipitate electrochemically at low trode surface can potentially decompose to hydroxyl radicals on
potentials. the active Pt sites,186 and free radicals can also be directly
generated from the crossover gases on the platinum particles
(Pt)n + Pt2+ / (Pt)n+12+ (18)
precipitated in the membrane (Fig. 12).137,187 Evidence for such
(Pt)n+12+ + H2 / (Pt)n+1 + 2H+ (19) a mechanism was provided by the work of Kim et al. where
formation of carbon radicals on the Pt particles was conrmed
Such dynamic oxidation/dissolution and reduction/ by using electron spin resonance.187 Ohguri also demonstrated
deposition processes results in Pt band formation. The Pt the formation of HOc in the presence of dispersed Pt particles83
particle shape, size and distribution is strongly inuenced by however it is necessary to keep in mind that metallic platinum is
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

the composition and partial pressure of the crossover oxygen a less active Fenton catalyst than iron and therefore direct
and hydrogen gases.10,180 Kim et al. studied the location of the Pt formation of HOc radical is much less effective.188
band in a Naon® membrane with 25 mm thickness and Pt(II) species present within the membrane can lead to the
a ePTFE reinforcement incorporated at the centre.181 A Pt band generation of free radicals136 and can affect the membrane
was observed between 1 and 10 mm from the cathode membrane physical properties due to the interaction with the per-
interface, where oxygen is typically depleted. These authors re- uorosulfonic groups resulting in loss of conductivity and
ported a greater deposition of Pt nanoparticles in the polymer membrane stiffening.189 The lowest uoride emission rate153
structure under OCV hold test conditions than under constant and no Pt band formation was observed for MEAs with air
current operation. Similarly OCV hold was reported to accel- supplied to the cathode;14,190 on the contrary to induce PTIM
erate polymer degradation. It was proposed that the presence formation in the membrane pure oxygen is commonly used as
and location of the platinum band might correlate with the the reactant gas.156 All those ndings conrm the destructive
formation of highly reactive oxygen species.182 Hatanaka et al. effect of the Pt particles deposited in the membrane on its
carried out studies with pristine Naon® and with a membrane degradation.
where a Pt-band had been intentionally introduced.183 Aer the
OCV hold testing, the FER of the membrane containing plat-
inum was clearly higher than that of the pristine membrane.
Ohma et al. employed micro-Raman spectroscopy to investigate
the degradation prole of in situ aged membranes and found
enhanced degradation of the polymer around the Pt band.14,15
Also post-mortem SEM study of an aged MEA revealed
membrane thinning at the cathode site in presence of PTIM and
conductive AFM measurements showed electronic connections

Fig. 12 Formation of free radicals on Pt and the Pt dissolution


mechanism. Reprinted from (vol. 195, D. Zhao, B. L. Yi, H. M. Zhang and Fig. 13 The effect of membrane thickness and Pt particle size on FER
M. Liu, The effect of platinum in a Nafion® membrane on the durability for different platinum loadings (B) and spacing (A) in the membrane.
of the membrane under fuel cell conditions, 4606–4612).185 Copyright Reproduced by permission of The Electrochemical Society from ref.
(2010), with permission from Elsevier. 186.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 419
View Article Online

Sustainable Energy & Fuels Review

Table 1 Overview of reactions involving formation and decay of intermediates in a fuel cell environment. Reproduced by permission of The
Electrochemical Society from ref. 81. Listed rate constants were estimated at room temperature and low pH. Reactions of hydroxyl radical with
polymer structure (reactions 1 & 2); reaction of formation, decomposition and consumption of hydrogen peroxide (reactions 3–8), Fenton
reaction and reaction of hydrogen peroxide and radicals catalysed by iron impurities

Reaction Rate constant (M1 s1)

1 HOc + RfCF2SO3H / unzipping —


2 HOc + RfCF2COOH / unzipping <106
3 H2O2 / 2HOc 1.2  107
4 HOc + H2O2 / HOOc + H2O 2.7  107
5 HOOc + H2O2 / HOc + H2O + O2 #1 (reasonable upper limit value)
6 HOc + H2 / Hc + H2O 4.3  107
7 Hc + O2 / HOOc 1.2  1010
8 2HOOc / H2O2 + O2 8.6  105
9 Fe2+ + H2O2 + H+ / Fe3+ + HOc + H2O 63
10 Fe2+ + HOc + H+ / Fe3+ + H2O 2.3  108
11 Fe2+ + HOOc + H+ / Fe3+ + H2O2 1.2  106
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

12 Fe3+ + HOOc / Fe2+ + O2 + H+ 2  104


13 Fe3+ + H2O2 / Fe2+ + HOOc + H+ 4  105

On the other hand, some studies show no correlation membrane samples, enabled the deactivation of the generated
between the presence of the PTIM and membrane degrada- radicals prior to the polymer decomposition.196 Finally, non-
tion152 or even a decrease of PFSA ionomer degradation (lower platinum-group-metal ORR catalysts are being increasingly
uoride emission rate) of Pt-containing membranes in investigated as an option for fuel cell application. However,
comparison to pristine polymer, due to the scavenging of catalysts with Fe or Co and N species anchored on carbon
hydrogen peroxide and hydroxyl radicals by Pt particles.148,152,191 substrates (Fe/Co–N–C) have to be considered in the context of
Macauley et al. reported enhanced stability of a eld-operated PFSA membrane degradation. Fe/Co–N–C catalysts have low
MEA with a PTIM compared to a freshly prepared MEA. The stability197 and suffer from dissolution/leaching of (Fenton
pristine membrane showed higher OCV decay and uoride active) metal, oxidative attack of hydrogen peroxide and
emission rate aer 10 hours of AST cycle as well as a decrease in protonation of active sites on neighbouring N species.198 All
IEC aer exposure to Fenton's reagent.191 This group developed three mechanisms not only cause a strong decrease in catalyst
an Accelerated Membrane Durability Test (AMDT) protocol to activity, but also threaten the integrity of PFSA membranes.199
mimic membrane degradation during heavy duty fuel cell Surprisingly little work has been carried out on membrane
operation.192 Once again PTIM prolonged MEA life-time under degradation in the presence of non-noble metal catalysts, and
these AMDT conditions. Peron et al. suggested that this well designed in situ accelerated stress testing is required to
discrepancy arises from coexistence in the membrane of develop proper understanding.
different Pt species such as Pt(0), Pt(II) or Pt(IV). While metallic
platinum might act as radical quencher, Pt(II) catalyses the
hydrogen peroxide decomposition and radical formation. Later
8. Radical types, reactivity and
Gummalla et al. explained these nonlinear trends by two lifetime
competitive reactions which can occur on the Pt surface:
Aer examining the degradation mechanisms in the context of
production and deactivation of free radicals.186 In that context
the polymer structure susceptible to radical attack, as well as
such parameters like: the Pt particle size,186 shape,193 loading
H2O2 formation area and conditions, in this section we will
and spacing186,194,195 (see Fig. 13) dene its potential radical
consider the radical types and their generation in a fuel cell in
scavenging or generating properties.
more detail, up to this point the membrane degradation was
Furthermore those parameters combined with temperature,
described mainly as hydroxyl radical HOc attack on polymer
level of hydration or concentration of crossover gases appear as
structure. However in a working fuel cell Hc and HOOc radicals
crucial key factors impacting polymer decomposition. The
are present in concentrations much higher than those of HOc
proposed model nds conrmation in recently performed
which, once generated from hydrogen peroxide, is further
studies, on the degradation of a Pt-containing membrane.195
consumed to produce HOOc or scavenged by crossover gases. An
Rodgers et al. studied the effect of precipitated platinum with
excellent summary of reactions likely taking place in a fuel cell
concentrations of 0, 10, 30 and 50 mol% on membrane degra-
(Table 1) was given by Gubler et al.81
dation. The highest FER with similar magnitude to that
This inventory of relevant mechanisms helped to create
observed with Pt/C catalyst coated on the polymer, was found
simulation of possible radical attack. Simulation based on the
for a membrane containing 10 mol% Pt. In contrast, samples
reaction (2) represents the chemical attack of the hydroxyl
containing 0, 30 and 50 mol% Pt gave very low FER. It was
radical on the terminal –COOH groups, i.e. the “unzipping
proposed that a dense distribution of Pt particles in the two last
mechanism” discussed in the previous section. In their studies

420 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

Table 2 Overview of reduction potentials of free radicals (pH ¼ 0).


Reproduced by permission of The Electrochemical Society from ref.
118

Standard electrode
Half-cell reaction potential (V)

1 HOc + H+ + e / H2O 2.59


2 Hc + H+ + e / H2 2.32 (ref. 80)
3 HOOc + H+ + e / H2O2 1.48
4 H2O2 + 2H+ + 2e / 2H2O 1.74

of the kinetics of membrane degradation, Gubler et al. did not


take into consideration the radical attack on the side chain
(cited previously as mechanisms II and III and represented in
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

the table by reaction (1)) due to the lack of information on the Scheme 1 Factors impacting PFSA fuel cell membrane degradation.
kinetics of these reactions in the literature. Two sets of reactions
used by Gubler et al. in their studies can be distinguished with
hydrogen peroxide as the precursor for the radical experimental values of FER on OCV hold indicates the
intermediates. substantial role of these factors on the chemical decomposition
Reactions (3)–(8) describe the homolytic dissociation of of the ionomer membrane in the MEA.
H2O2 (reaction (3)) and consumption of free radicals origi- Another important issue is the oxidative strength of free
nating from H2O2 in the absence of iron catalyst. Reaction (3) radicals. The most cited radical with regard to chemical
occurs with very low rate constant approximately 1.2  107 degradation is the hydroxyl radical HOc. This is due to the
M1 s1. As displayed in eqn (4), HOc reacts with hydrogen ranking of the reactivity of the oxidative species, where the
peroxide, to form HOOc, while reaction (5) follows a similar hydroxyl radical takes a rst position: HOc > Hc > HOOc >
pathway but in this case HOOc is the reactant for the genera- H2O2.80 High oxidising power of HOc is reected in higher
tion of hydroxyl radicals. Furthermore the reaction of Hc and reduction potential of HOc in comparison to the hydroperoxyl
HOc radicals with crossover gases in a fuel cell follows the eqn radical and hydrogen peroxide (Table 2).
(6) and (7). Finally hydrogen peroxide can be regenerated as Similarly, the difference in the rate of hydrogen abstraction
described by eqn (8) due to HOOc disproportionation. However can be estimated based on the experimental bond strength.117
in real working fuel cell conditions, reactions (9)–(13) are the As the O–H bond in water is very strong, the hydroxyl radical has
most relevant, since H2O2 decomposition is catalysed by Fe ion strong thermodynamic driving force (497.9 kJ mol1) to abstract
impurities. Reaction (9) is widely known as the Fenton reac- the hydrogen and form an H2O molecule. In contrast it is very
tion and it is the primary source of hydroxyl radicals, where unlikely that hydrogen peroxide or HOOc will cause hydrogen
Fe2+ is the reducing agent regenerated aer the reaction of abstraction with direct detrimental effect on the PFSA
Fe3+ with H2O2 and HOOc. The rate constant of reaction (9) membrane. The main role of those species is therefore as
(here 63 M1 s1) is much higher that previously cited in the a source of much more reactive HOc.
homolytic dissociation of hydrogen peroxide, which further
means higher rate of HOc generation.
Indeed the simulation studies of the radical attack per- 9. Impact of operating conditions on
formed by Gubler showed higher rate of the ionomer attack PFSA membrane chemical degradation
under ex situ, Fenton reagent conditions, in comparison to that
observed in an OCV hold test.81 It is most likely due to a higher The evaluation of membrane durability in an operating fuel cell
concentration of hydroxyl radicals. Moreover, based on the is challenging since many factors including operation temper-
kinetic model the theoretically estimated uoride emission rate ature, pressure, humidity, operating voltage and current, start-
aer Fenton aging was in good agreement with FER reported in up/shut-down conditions, the composition and design of
the literature.100,150,200 However a similar FER estimation for in other cell components and the effect of contaminants have to be
situ accelerated tests were 2–3 orders of magnitude lower than considered. The synergetic effect of all these factors in oper-
the literature values.91,188 The differences between the in situ ating fuel cell make identication of the local phenomena
measurement and the kinetic simulation might be due to the leading to membrane chemical decomposition quite complex.
absence in the provided model of reaction (1), which describes Test protocols have been designed to limit certain phenomena
the side chain decomposition. Moreover two other factors were and exacerbate others, and furthermore harmonized methods
not included in Gubler's model i.e. the inuence of the direct provide the basis for comparing results and facilitate assess-
formation of free radicals on Pt nanoparticles in the membrane ment of technology status. In real operation these conditions
and the inuence of the electrode surface on membrane are rarely separated and performance losses can arise from
degradation. A large difference between simulated and pinhole formation201 following mechanical/chemical

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 421
View Article Online

Sustainable Energy & Fuels Review

membrane degradation cycling, while RH uctuations, irradiation has also been investigated.218 In general new cross
temperature and loading cycling and start-up, shut-down can all linked materials show reduced dimensional swelling and
cause membrane dimensional change, and lead to plastic higher mechanical stability in comparison to non-reinforced
deformation, cracks and pinhole formation. Following pinhole PFSA membranes. However a similar effect can be achieved
formation, high hydrogen crossover causes accelerated radical using physical reinforcements, in which the membrane prepa-
formation and further polymer decomposition. The degrada- ration is further simplied. To fabricate a reinforced composite
tion pathway continues until the membrane, thinned and membrane, a highly mechanically stable organic or inorganic
weakened by radical attack membrane, can no longer withstand matrix is impregnated or otherwise associated with the PFSA
mechanical stress when, regardless of the decomposition mode dispersion. A large spectrum of membrane reinforcements has
by chemical attack or shear forces, the consequence is a MEA been developed, including expanded porous PTFE
failure (Scheme 1). sheets204,219–221 or brils as well as nanobre mats prepared by
electrospinning of polymers such as poly(vinylidene uoride)
(PVDF),222,223 poly(vinyl alcohol) (PVA),223,224 polybenzimidazole
10. Mitigation of membrane (PBI),225 poly(phenyl sulfone), UHMWPE (ultrahigh molecular
degradation weight polyethylene)226 or nanober network silicon carbide
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

SiC.227 Among these, low cost expanded PTFE (ePTFE) supports


The foregoing discussion shows that degradation of PFSA
have been particularly successful. ePTFE reinforced membranes
membranes is currently considered to involve complex inter-
are usually prepared by soak or spray methods.228 The ionomer
connected mechanical and chemical mechanisms. Due to the
lls the pores of the thin ePTFE lm and forms a layer of ion-
complexity of these factors contributing to the overall degra-
omer on either side of PTFE mat. The advantages of using PTFE
dation, many different approaches have been developed to
as the membrane support include the reduction of membrane
minimise polymer cleavage processes. The main mitigation
cost228 and improved dimensional stability and durability of
strategies proposed in the literature include development of
composite membrane229 when subjected to mechanical accel-
new reinforced or composite membranes, stabilisation of most
erated tests such as freeze/thaw or, humidity cycling230,231 as
vulnerable PFSA polymer groups, incorporation of a hydrogen
PTFE suppresses swelling of PFSA electrolyte within the porous
peroxide decomposition catalyst or radical quencher in the
matrix. However, PTFE/PFSA membranes demonstrate higher
membrane/electrode, development of new electrode/GDL
proton transport resistivity than the respective non modied
material.
membranes.228 This negative effect can be minimized by
applying ultrathin PTFE mats with high porosity,232,233 or by
10.1 Development of new reinforced or composite hydrophilic treatment of hydrophobic PTFE structure.230 Such
membranes treatment improves the wettability and adhesion of more
Preparation of a reinforced material was proposed to improve hydrophilic polymer which in turn allows good interfacial
the durability of the PFSA polymer.107,202–205 In general two compatibility. However according to Kundu et al. the ePTFE
widely investigated routes to achieve this goal can be distin- reinforced PFSA membranes undergo accelerated chemical
guished: chemical modication of the polymer structure and degradation at low relative humidities 20–50%.166 Thus
physical reinforcement.203 Chemical modications include increased mechanical durability of the membrane might extend
membrane annealing206–212 uniaxial stretching213,214 and chem- lifetime but not suppress chemical failure. Inorganic llers
ical cross linking of the polymer,203,215 these were used to reduce resulting in nanocomposite membranes like Naon® –
swelling of the membrane and improve mechanical durability. SiO2,234–238 Naon® – TiO2,236,238–242, Naon® – SnO2,243–246
Annealing of PFSA polymers leads to more efficient chain Naon® – ZrO2 (ref. 236, 238, 247 and 248) or Naon® –
packing and higher polymer crystallinity. Thermal treatment is ZrP248–251 deserved special attention. Incorporation of inorganic
extremely important, especially for a solvent cast membrane, as particles can not only enhance the mechanical properties of the
the annealing procedure improves intrinsic properties of the PFSA membranes, but also increase the water retention, which
polymer and makes them similar to those of an extruded allows higher performance at elevated operating temperatures
membrane. Treatment by uniaxial stretching has been and low relative humidity.249 Furthermore inorganic particles
demonstrated to effectively increase Young's modulus, reduce can be successfully applied as a hydrogen peroxide decompo-
area swelling and slightly increase proton conductivity. Other sition catalyst or radical scavenger, as will be discussed in the
chemical modications such as polymer cross linking can be later section of this review. PFSA-inorganic nanoparticles
carried out by conversion of the sulfonic acid side chain composite membranes can be prepared either from a disper-
however such method leads to inevitably high conductivity sion of oxide particles in the ionomer solution prior to
losses and therefore was not widely studied. Incorporation in membrane casting or by in situ generation of inorganic particles
the polymer structure of new cross-linkable functional groups within the membrane/ionomer through a sol–gel-like process.
appears as the more appropriate method and was performed The properties of such prepared composite membranes depend
with sulfonyl uoride and sulfonamide216,217 or sulfonyl uoride to large extent on the type of the polymer, the particle size, their
and 3-aminopropyltriethoxysilane (APrTEOS) as reactants. A dispersion through the membrane and any preferential orien-
different method based on pendant alkyl bromide group, which tation, in the case of inorganic particles with large aspect
can be cross linked by thermal treatment or electron beam ratio.252

422 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

10.2 Stabilisation of most vulnerable PFSA polymer groups – carbon–tungsten oxide (C/WO3) or carbon–phosphotungstic
development of SSC PFSA membrane acid (Pt/C-PTA).271,272 Metal particles such as Ce,108,273–277
Mn,274,276,277 Pd, Ag, Au or Pt,278 metal oxides CeO2,279–289
Stabilisation of the most vulnerable groups of the PFSA polymer
MnO2,290,291 SnO2 (ref. 292 and 293) and terephthalic acid
is another method to mitigate membrane degradation. It was
(TPA)294 also demonstrate ability to mitigate chemical
discussed previously in connection with different variants of
degradation.
chemical decomposition and should be considered in relation
The main role of HPDC is to lower the H2O2 concentration
to the mechanism of free radical attacks on the polymer struc-
ture. Naon® with chemically stabilised carboxylic acid end and indirectly decrease the formation of free radicals. In
groups – the most susceptible point in polymer – shows reduced contrast introduction of a regenerative radical scavenger into
MEA components causes direct reduction of the concentration
but not eliminated mass loss. Lower uoride emission of
of already created highly oxidative species. For these reasons
chemically stabilised Naon® is due to limited participation of
two potential areas for HPDC/radical scavenger incorporation
the unzipping mechanism in the overall decomposition reac-
may be distinguished: the polymer membrane and the elec-
tion, however weight decrease is still noticeable and it is related
trode. In the rst case, preparation of a composite membrane
to the continuously occurring side chain degradation with
was accomplished by mixing the HPDC/radical scavenger
tertiary carbon, ether bonded carbon and/or sulfonic acid
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

groups as the points of initial attack as these are the most nanoparticles or chemical compounds with the ionomer
vulnerable points. Although the development of a per- suspension prior to formation of the membrane. In this
approach, hydrogen peroxide can be neutralised only aer its
uorinated ionomer with a shorter pendant side chain than
diffusion into the polymer membrane. In the work of Xiao et al.,
Naon®, the so-called short side chain (SSC) ionomer pre-dates
zirconia particles were prepared by a hydrothermal method and
understanding of membrane degradation, the short side
incorporated in the Naon® membrane. The resulting ZrO2
structure is expected to bring several advantages regarding
composite membrane was examined under OCV hold (80  C;
membrane durability. Elimination from the side chain struc-
50% RH; H2/O2) and Fenton's test conditions. Accelerated stress
ture of the tertiary carbon (–CF) and one of the ether bonded
carbon atoms (–O–CF2) decreases the susceptibility of the SSC tests show reduced FER (by up to an order of magnitude) and
membrane to radical attack, a nding that was conrmed in thinning, relative to native Naon®.200 As zirconium oxide
enhances the membrane durability (due to its radical scav-
several studies.100,108,149 More recently attention has been drawn
enging ability) and water retention at high temperature and low
to low EW multi-acid side chain proton exchange membranes
relative humidity, development of composite ZrO2/PFSA
developed by 3M.253–256 In these new materials high number of
membranes is an interesting approach to mitigate mechanical
tetrauoroethylene units in the backbone structure has been
and chemical degradation. A similar concept to reduce both
combined with side chains which carry more than one acid site.
mechanical and chemical degradation was investigated by Patil
3M peruoro-imide acid (PFIA) membranes have high acid
content while preserving excellent mechanical stability when et al. by using a TiO2/Naon® composite membrane.265 A titania
compared to similar EW PFSA membranes.253 Although funda- “quasi-network” in the polymer structure was developed by in
situ sol–gel polymerisation of titanium isopropoxide. The EW of
mental studies regarding mechanical, conductive properties of
the modied membrane remains unaltered, however the
3M PFIA membranes have been carried out no reports have
hydration capacity decreased in relation to the pristine
been published so far on PFIA durability. Only conclusion
Naon®. The authors postulated that the incorporation of
which might be drawn today is that both the SSC (including
small (nano size) and evenly dispersed particles, placed along
PFIA) and long side chain (LSC) membranes have a sulfonic acid
the gas diffusion pathways, potentially increase tortuosity and
as functional ion transporting groups, therefore they both can
suffer the decomposition which begins under certain condi- reduce fuel crossover. Additionally, if the titania particles form
tions at the C–S bonds. To mitigate degradation of the side a slightly interconnected network, due to the presence of weak
metal oxide bonds, they can constitute a mechanical membrane
chain, incorporation of a free radical scavenger as a chemical
reinforcement.252 Indeed the modication provides signicant
trap for highly reactive oxygen species was proposed.
dimensional membrane stability (higher modulus and lower
creep) and minimised voltage loss at OCV hold test.239 It should
10.3 Incorporation into membrane/electrode hydrogen be noted that use of ZrO2 and TiO2 as additives for PFSA
peroxide decomposition catalyst (HPDC) or radical quencher membranes was carried out with the aim of improving their
In this manner we come to the highly effective approach to water-retention and mechanical properties, and the radical
minimize the chemical degradation, which involves incorpora- scavenging effect of those compounds was discovered later.
tion, into the membrane electrode assembly, of species with Similarly heteropolyacid/PFSA membranes were initially devel-
peroxide decomposition or radical scavenging properties. This oped to maintain proton conductivity of membrane at elevated
concept has been investigated already before 2005. Convincing temperature and low relative humidity conditions.295–298
results in this area have led to the increased interest of many Haugen et al., investigated the durability of 3M membrane
researchers and several patents.257–262 Examples of hydrogen modied by the addition of various heteropolyacids (HPAs) as
peroxide decomposition catalysts (HPDC) applied for mitiga- HPDC.267 Of these, l-H3P2W18O62, H6P2W21O71 and H4SiW12O40
tion of membrane degradation are: metal oxides like signicantly reduced the amount of uorine released under
TiO2,239,263–265 MnO2 (ref. 266) and ZrO2,200 heteropoly acids,267–270 accelerated stress test conditions. Incorporation of the HPAs

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 423
View Article Online

Sustainable Energy & Fuels Review

enhanced performance of the MEA in a narrow concentration displays a possible mechanism of the reaction of TPA with
range of HPA. The efficacy of the mitigation properties of metal a hydroxyl radical. The generated intermediate, hydrox-
nanoparticles such as Au, Ag, Pt and Pd in Naon® was studied ycyclohexadienyl radical can either undergo a disproportion-
by Trogadas et al.278 The idea of using those elements as radical ation reaction to form hydroxylated terephthalate and
scavengers was rst provided by their known activities in regenerate TPA, or can be catalysed by an Fe3+ ion to form the
biology as antioxidant species. Composite membranes were hydroxylated terephthalate.
prepared by casting dispersions with 3 wt% metals loading. A composite TPA/PFSA membrane aged under ex situ Fenton
Such a metal content lowers the conductivity of Naon® due to test conditions shows higher durability than pristine PFSA
dissolution of the metal particles above a certain potential i.e. membrane, which indicates HOc scavenging ability. However
>0.4 V vs. SHE for Ag and >0.8 V vs. SHE for Pd, and the these promising results need to be further conrmed in situ in
formation of ionic species that exchange with protons. Plat- a fuel cell in order to verify the performance, stability and
inum particles were an exception, where in situ water produc- impact on other MEA components of this newly developed
tion on Pt surface compensates metal dissolution and ion approach.
exchange with the sulfonic acid groups. Regarding the radical The second strategy involving the addition of peroxide
scavenging properties of the incorporated metals, the addition decomposition catalysts or radical scavengers in the catalyst
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

of Ag, Pt, Pd and Au nanoparticles decreased the uoride layers has the advantage that the H2O2 can be decomposed
emission rate by 35%, 60%, 80% and 90% respectively. To already at the catalyst surface. Typically a catalyst ink includes:
minimise the impact of metal dissolution on membrane carbon support, metal catalyst and ionomer. The ionomer can
conductivity or further performance, metal particles were sup- comprise an HPDC associated with anion groups in the ion-
ported on silica, SiO2. Indeed, the conductivity of the composite omer salt form or suspended as particles in the catalyst ink.
membrane with SiO2-supported metal particles increased Trogadas et al. validated the second approach by preparing
signicantly and reached values similar to the reference various hybrid catalysts such as C/WO3, C-PTA271 or Pt/C–
Naon®. The impact of silica incorporation was especially MnO2,266 which were studied for their efficacy of hydrogen
visible for SiO2/Pt Naon® membrane. However a lower peroxide decomposition and mitigation of membrane degra-
concentration and simultaneously larger size of the supported dation. RRDE measurements showed signicant reduction in
metal nanoparticles brought considerable FER increase. a- the concentration of H2O2 produced during ORR on these
Tocopherol (a-TOH) is other natural antioxidant (vitamin E catalysts. Also in situ OCV hold tests of the MEAs with HPDC
component) which, when embedded into Naon® prevented additives revealed decrease of the rate of membrane degrada-
membrane degradation and performance losses.299 a-TOH acts tion which was reected in lower uoride emission rate. The
as a trap for hydroxyl and peroxyl radicals once oxidized, a-TOH best results in H2O2 decomposition were given by the incorpo-
can be further reduced by crossover hydrogen during fuel cell ration of MnO2 particles with mixed valence state whereby they
operation. The reversibility of a-TOc/a-TOH system assures its can undergo relatively easy reversible redox reaction with H2O2,
high efficiency as a radical scavenger. However the long-term moreover the free hydroxyl radicals – HOc can be scavenged at
stability of a-TOH incorporated into electrolyte is unknown. the same time. On the other hand, the performance of an MEA
Recently Zhu et al. reported a new method to increase the containing an MnO2 hybrid catalyst was inadequate, due to low
chemical stability of PFSA membrane by the incorporation of an stability of MnO2 in acidic media at high potentials. Recently
organic radical scavenger – terephthalic acid (TPA).294,300 Fig. 14 developed MnO2 nanotubes and nanowires in small content did
not affect the available catalyst area.290 The performance of
manganese oxide doped catalyst layer measured at 100% RH
was similar to that with a conventional catalyst material.
Interestingly for lower RH values, incorporation of 1D MnO2
decreased the sensitivity of the fuel cell to variation in hydration
level. Moreover, OCV hold degradation tests have shown much
lower OCV decay with an MnO2 composite catalyst in compar-
ison to commercial material especially if incorporated on the
anode side (see Table 3) Tacconi et al. developed a Pt/C–TiO2
composite catalyst by using photocatalytic synthesis,263 and
found that the incorporation of TiO2 in the catalyst layer brings
signicant increase to the electrode and membrane durability
with comparable performance of a Pt/C–TiO2 composite for
oxygen reduction reaction to that with a commercial catalyst.
Finally, Brooker et al.268 incorporated adsorbed HPAs on high
Fig. 14 Mechanism of HOc radical trapping process by TPA proposed surface area carbon at the membrane–electrode interface. Aer
by Zhu. Reprinted from (vol. 432, Y. Zhu, S. Pei, J. Tang, H. Li, L. Wang,
OCV hold testing, a cell with a sub-layer enriched with phos-
W. Z. Yuan and Y. Zhang, Enhanced chemical durability of per-
fluorosulfonic acid membranes through incorporation of terephthalic photungstic acid (PW12O43) showed decreased OCV loss and
acid as radical scavenger, 66–72).294 Copyright (2013), with permission uoride release of around 50% over that of the non-HPA sub-
from Elsevier. layer cell. However the inclusion of the carbon sub-layers

424 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

Table 3 Overview of hydrogen decomposition catalyst and radical scavenger incorporated in MEA components

Accelerated durability test

REF Form/location Content Ex situ ageing In situ ageing

271
Hybrid catalyst 15 wt% RRDE OCV: 90  C 50% RH O2/H2
C/WO3 30–40% reduction in the amount FER: (ppm)
of H2O2 produced for 15 wt%
C/WO3 and C/PTA versus Pt/C.
C/PTA Anode:
C – 1.3
C/WO3 – 0.5
C/PTA–0.7
Cathode:
C – 0.55
C/WO3 – 0.38
C/PTA–0.36
OCV: 90  C 50% RH Air/H2
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

266
Hybrid catalyst 5 wt% RRDE
FER: (ppm)
CATHODE 50–60% reduction in the amount of Anode:
Pt/C/MnO2 H2O2 produced for 5 wt% Pt/C/MnO2 Pt/C1.1
versus Pt/C. Pt/C/MnO20.22
Cathode:
Pt/C0.84
Pt/C/MnO20.59
290
Hybrid catalyst – OCV: 60  C 75% RH Air/H2 decay: (mV min1)
CATHODE/ ANODE Pt/C51.2
Pt/C/MnO2 Cathode : Pt/C/MnO2  nanowire  39.4
Cathode : Pt/C/MnO2  nanotubes  38.8
Anode : Pt/C/MnO2  nanowire  17
Anode : Pt/C/MnO2  nanotubes  18.6
283
Hybrid catalyst 20 wt% – OCV: 90  C 30% RH O2/H2 decay: (mV)
CATHODE Pt/C  0.72
Pt/C/CeO2 Cathode : Pt/C/MnO2  0.079
326
Hybrid catalyst TiO2 – FER: (mmol cm2 h1)
CATHODE 0 wt% ETEK (Pt 5 wt%)  0.03
Pt/C/TiO2 5 wt% SIDCAT (Pt 5 wt%)  0.3
5 wt% SIDCAT (Pt 10 wt%)  0.05
0 wt% TKK (Pt 50 wt%)  0.75
5 wt% SIDCAT (Pt 50 wt%)  0.06
10 wt% SIDCAT (Pt 50 wt%)  0.02
294
Composite membrane 0.5 wt% Fenton test F (%) –
TPA/PFSA  12% (48h)
TPA/PFSA PFSA  14% (48h)
TPA/PFSA  21% (120h)
PFSA  29% (120h)
278
Nanocomposite membrane 3 wt% – OCV: 90  C 30% RH O2/H2
FER:(mmol cm2 h1)
Ag Naon  0.41
Pt Naon/Ag  0.275
Pd Naon/Pt  0.18
Au Naon/Au  0.05
Naon/Pd  0.08
280
Nanocomposite membrane – – OCV: 90  C 30% RH O2/H2
– FER: (mmol cm2 h1)
CeO2/Naon® CeO2: Anode:
0 wt% Naon®  0.2
0, 5 wt% Naon®/CeO2  0.007
1 wt% Naon®/CeO2  0.006
1 wt% Naon®/CeO2(nc)  0.005 (nc – non commercial)
3 wt% Naon®/CeO2  0.008
– Cathode:
0 wt% Naon®  0.08
0,5 wt% Naon®/CeO2  0.006
1 wt% Naon®/CeO2  0.005

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 425
View Article Online

Sustainable Energy & Fuels Review

Table 3 (Contd. )

Accelerated durability test

REF Form/location Content Ex situ ageing In situ ageing

1 wt% Naon®/CeO2(nc)  0.003


(nc – non commercial)
3 wt% Naon®/CeO2  0.007
286
Nanocomposite membrane 1 wt% – OCV: 90  C 30% RH Air/H2 FER: (mmol)
PFSA  3800
CeO2/Naon® Ceria(nc)  340 (nc – non commercial)
Ceria(com.)  33 (com commercial)
287
Nanocomposite membrane CeO2: Fenton test F (mmol g1) –
0 wt% PFSA: 2600(LF) 2600(GF)
CeO2/Naon® 0.5 wt% Ceria(nc): 500(LF) 1100(GF)
1 wt% Ceria(nc): 400(LF) 800(GF)
2 wt% Ceria(nc): 200(LF) 250(GF)
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

0.5 wt% Ceria(com.): 400(LF) 1600(GF)


1 wt% Ceria(com.): 350(LF) 700(GF)
2 wt% Ceria(com.): 100(LF) 350(GF)
(LF) – Fenton reaction in solution phase
(GF) – Fenton reaction in gas phase
282
Nanocomposite membrane CeO2: Fenton test F (mg h1) OCV: 95  C RH cycling Air/H2 decay: (mV s1)
0 wt% Naon  55.78 Naon  11.7  104
CeO2/Naon® 1 wt% Ceria(self assembled)  43.05 –
3 wt% Ceria(self assembled)  8.67 –
5 wt% Ceria(self assembled)  6.01 Ceria(self assembled)  1.13  104
10 wt% Ceria(self assembled)  4.47 –
5 wt% Ceria(sol–gel)  11.64 Ceria(sol–gel)  5.78  104
281
Nanocomposite membrane – – OCV: 90  C 30% RH O2/H2
Pt: FER: (mmol cm2 h1)
Pt/CeO2/Naon® 0 wt% Naon®  0.6
– Naon®/CeO2  0.01
0.5 wt% Naon®/Pt/CeO2  0.007
1 wt% Naon®/Pt/CeO2  0.007
2 wt% Naon®/Pt/CeO2  0.007
– Naon®/MnO2  0.008
327
Nanocomposite membrane 1 wt% Fenton test F (mmol/gh) OCV: 80  C 50% RH O2/H2
Naon®: 2.25 FER: (mmol cm2 h1)
CsxH3xPW12O40/CeO2 Naon®/CeO2: 0.6 Anode:
H3xPW12O40/CeO2: 0.5 Naon®  0.6
Naon®/CeO2  0.4
H3xPW12O40/CeO2  0.025
Cathode:
Naon®  0.8
Naon®/CeO2  0.5
H3xPW12O40/CeO2  0.1
264
Nanocomposite membrane 20 wt% – OCV: 90  C 30% RH O2/H2 FER: (mg cm2 h1)
Anode:
TiO2/Naon® Naon®  3.681
TiO2/Naon®  0.354
Cathode:
Naon®  4.98
TiO2/Naon®  0.1755
265
Nanocomposite membrane 20 wt% – OCV: 100  C 25% RH O2/H2
FER: (mg cm2 h1)
Anode:
TiO2/Naon® Naon®  1.43
TiO2/Naon®  0.07
Cathode:
Naon®  1.46
TiO2/Naon®  0.12

426 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

Table 3 (Contd. )

Accelerated durability test

REF Form/location Content Ex situ ageing In situ ageing

200
Nanocomposite membrane 3 wt% Fenton test OCV: 80  C 50% RH O2/H2
FER (mmol g1 h1) FER: (mmol cm2 h1)
ZrO2/Naon® Anode:
Naon® < 85 Naon®  0.6
ZrO2/Naon® < 75 ZrO2/Naon®  0.06
Cathode:
Naon®  0.8
ZrO2/Naon®  0.02
291
Nanocomposite membrane 3 wt% Fenton test OCV: 90  C 50% RH Air/H2
FER (mmol g1 h1) FER: (mmol cm2 h1)
MnO2/SiO2–SO3H Anode:
Naon®  0.8 Naon®  0.8
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

MnO2/Naon®  0.04 MnO2/Naon®  0.04


MnO2/SiO2/Naon®  0.05 MnO2/SiO2/Naon®  0.05
Cathode:
Naon®  1.1
MnO2/Naon®  0.2
MnO2/SiO2/Naon®  0.08
279
CeO2 free particles in 1 mM, Fenton test – EPR STUDIES –
fenton solution 10 mM Higher radical scavenging properties
of CeO2 with lower (10 mM) concentration
267
HPAs doped membranes – – OCV: 90  C Air/H2 FER: (mg per cm2 per day)
HP2W18 10 wt% 0.6
HP2W21 10 wt% 0.8
HSiW 20 wt% 0.84
275
Ce ion exchange membrane Ce: – OCV: 90  C 30% RH Air/H2 FER: (g cm2 h1)
0 wt% Naon®  1  106
5 wt% Naon®/Ce  1  109
10 wt% Naon®/Ce  1.6  1010
– – OCV: cold start conditions FER: (mmol)
0 wt% Naon(50) 0.45
(Naon50 – aer 50 OCV cycles)
10 wt% Naon(50) z 0
276
Ion exchange membrane (mmol) Hydrogen peroxide ow cell experiments OCV: 95  C 50% RH O2/H2 FER: (g cm2 h1)
Mn2+  anode 2.2 Co-doping of Ce into membrane containing 2  107
Mn2+  cathode 2.2 25 ppm of iron eliminates the accelerating 3  107
Ce3+ – anode 2.2 impact of Fe catalyzed hydroxyl 5  108
Ce3+ – anode/cathode 2.8 radical generation. 3  108
Ce3+ – membrane 45 3  108
268
HPA-carbon membrane- mg cm2 – OCV: 90  C 30% RH Air/H2 FER: (mmol cm2)
electrode sublayers
PW12O403 – PTA 0.13 No sublayer  40
PW11O404 – PTA-1V No HPA  48
SiW11O405 – STA-1V PW12O403  27
PW11O404  35
SiW11O405  40

signicantly increased the electrode ohmic and diffusion losses properties of PTA and low degradation of ionomer in catalyst
as well as OCV losses in comparison to the cell with no sub- layer.
layer. To overcome these problems Chen et al. impregnated Finally stabilisation301 or development of new ORR electro-
catalyst layers of CCM with phosphotungstic acid (PTA).272 catalysts can also be considered as a mitigation strategy as
Composite PTA-CCM exhibited better performance (especially elimination of potential catalytically active sites for hydrogen
at high current density) and stability than a standard non peroxide and radical generation is of great importance in
modied CCM. The power density of PTA-CCM decreased by increasing membrane durability. Platinum alloys are promising
14% in comparison to 33% of non-modied CCM aer 100 h candidates for PEMFC application due to their lower cost and
aging, and the authors related these results to scavenging high performance.302 Rodgers et al. carried out studies on the
inuence of a PtCo/C catalyst on membrane durability.303 They

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 427
View Article Online

Sustainable Energy & Fuels Review

showed that higher stability of the PtCo/C material in compar- with manganese. Both cations undergo easy redox reaction with
ison to the Pt/C catalyst is translated into lower membrane HOc to form water as displayed in eqn (20) and (21). These
degradation which results in reduced FER, lower hydrogen reactions have higher rate constants than those of hydrogen
crossover and voltage decay. Ramaswamy et al. also reported abstraction reactions by hydroxyl radicals, which results in the
higher durability of a membrane associated with an electro- scavenging properties of Ce3+ and Mn2+ ions. However, the
catalyst enriched with cobalt using novel segmented cell design, studies of Coms et al. demonstrated that for an equal molar
where membrane decay is correlated to losses in its ionic basis, cerium was much more effective at reducing FER. The
conduction.95 Based on RRDE measurement they concluded authors explained such behaviour by a higher rate constant of
that such behaviour is due to lower adsorption of oxygenated radical reduction reaction by cerium ions k ¼ 3  108 M1 s1. A
species on the Pt in the presence of Co, thereby diminishing similar rate constant of Ce3+ with HOc was reported Danilczuk
H2O2 formation. et al.273 based on the competitive kinetics approach.
Both cerium and manganese in ionic or oxide form have
been commonly applied to increase MEA durability by their HOc + Ce3+ + H+ / H2O + Ce4+, k ¼ 3  108 M1 s1 (20)
incorporation into the membrane or electrode. Due to the high
effectiveness in mitigation of the membrane degradation and HOc + Mn2+ + H+ / H2O + Mn3+, k ¼ 4  107 M1 s1 (21)
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

exceptional mechanism of the radical decomposition they


An excellent overview of all reaction involving Ce and Mn
deserve special attention. The idea of incorporation in the
cations was described by Gubler et al. in the framework of their
polymer matrix of cerium and manganese ions by partial ion
kinetic modelling studies.81 Gubler conrmed the correlation
exchange, was developed by Endoh et al.274,304
presented by Coms between reaction rate constants of HOc with
The concept proposed by Endoh assumed that when the
radical quenchers and their scavenging efficiency. As an
ionomer is ionically cross-linked with a radical scavenger
example, the concentration of 1% of Mn cations with respect to
cation, it can show improved mechanical properties and higher
the sulfonic acid groups scavenges 46% of HOc whereas the
chemical stability (Fig. 15). The in situ durability studies of
same concentration of Ce cations scavenges 89%. Further
newly developed material called NPC – New Polymer Composite
studies in this area underline the signicant impact of the ratio
membrane performed by Endoh under the following fuel cell
of the oxidation states of Ce and Mn on their mitigation prop-
conditions: OCV hold test, 20% RH and temperature 120  C,
erties. Only Ce3+ in the Ce3+/Ce4+ redox couple is able to react
demonstrated excellent stability of the NPC of over 1000 h with
with hydroxyl radicals, therefore high dynamic equilibrium of
FER in range of 2  108 g cm2 h1.
Ce3+/Ce4+ is necessary to keep high efficiency of the system.
Further research of Coms et al. in this area shed more light
Cerium has advantage to be self-regenerative, especially in an
on the scavenging properties of both Mn and Ce cations.276
acidic medium. Fast reduction of Ce4+ via reaction with H2O2
Coms et al. concluded that there is no impact of the initial
ensures rapid reactivation of Ce3+. It was reported that Ce4+ can
location of the radical quencher on the membrane stability
be also reduced by H2, H2O or HOOc on the Pt catalyst layer273,276
based on in situ degradation test of pristine and cation-
but, as argued by Gubler, the reaction of Ce4+ with H2O2 is fast
containing MEAs, where an ion-exchanged ionomer was used
enough to convert essentially all cerium ions to Ce3+ form.
for preparation of the catalyst ink. In other words, doping of the
Similarly to cerium ions, Mn2+ is also effectively restored via
anode or cathode electrode with radical scavenger gave a similar
reaction with H2O2 or HOOc.81 However the Mn2+/Mn3+ system
effect on the MEA durability. Coms argued that the hot-pressing
is much more complex than that of cerium. Species with
procedure promoted cation migration from the catalyst layer to
different oxidation states such as Mn+; Mn2+; Mn3+; and MnO2+
a low concentration area – here the membrane – and that such
need to be considered.81,305 Indeed Tanuma et al. provided
migration occurred irrespective of which electrode was initially
conrmation for this through their studies where the X-ray
doped. These authors did not nd evidence of cation leakage
absorption spectroscopy (XAFS) was employed to measure
out of the system however. Another important conclusion con-
a shi in absorption energy in the XANES (X-ray adsorption near
cerned the higher efficiency of cerium cation in comparison
edge structure) spectra of Ce and Mn ion exchanged
membranes before and aer OCV hold.305 No difference was
found in the XANES spectra of Ce exchanged membrane before
and aer accelerated degradation, whereas in the case of the
Mn exchanged membrane, a shi in the XANES spectrum to
higher energy aer OCV hold conrmed presence of Mn atoms
with valence state higher than +2.

HOOc + Ce4+ / O2 + Ce3+ + H+ (22)

H2O2 + Ce4+ / HOOc + Ce3+ + H+ (23)

HOOc + Mn3+ / O2 + Mn2+ + H+ (24)


Incorporation of radical quencher – concept of highly durable
Fig. 15
PFSA membrane developed by Endoh.274 H2O2 + Mn3+ / MnO2+ + 2H+ (25)

428 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

Experimental evidence of the Ce3+/Ce4+ redox couple was that Ce3+ ions can mitigate the membrane degradation under
provided by Danilczuk et al.93 by in situ monitoring of radical repeated cold-starts.
formation at the anode and cathode sides of pristine (MEA/H) However besides obvious advantages of the incorporation of
and Ce ion exchanged (MEA/Ce) Naon® 117 membrane. The Ce or Mn ions in MEAs, there are some disadvantages mainly
5,5-dimethylpyrroline-N-oxide (DMPO) was applied as a spin a decrease of membrane conductivity resulting in fuel cell
trap. performance losses. Another important problem is the mobility
Both MEAs types examined showed completely different of the ions. As was reported by Coms et al., cerium ions incor-
behaviour under OCV hold testing. For baseline MEA/H the porated in the catalyst layer aer hot pressing migrated into the
dominant adducts were DMPO/OH and DMPO/CCR – carbon PFSA membrane.276 It can be expected that scavengers can
centred radicals derived from Naon®, whereas for MEA/Ce the migrate under fuel cell operation conditions from an initially
main adduct detected aer membrane aging was DMPO/OOH. ion exchanged membrane to the electrodes as well, and further
The absence of hydroxyl radicals as well as the absence of be washed out with exhaust water. Decrease of the scavenger
carbon centred radicals conrmed the scavenging properties of concentration can be reected in the simultaneous decrease of
Ce3+. Moreover, the presence of DMPO/OOH indicates reaction mitigating efficiency, which makes this strategy inappropriate
of hydrogen peroxide with Ce4+ to restore the Ce3+. Another for long-term use. According to Gubler et al. the relation
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

study of the same research group performed with an ionomer between the membrane degradation rate and ions concentra-
dispersion demonstrated a concentration of OH and CCR tion can be expressed by the Fig. 16.81
adducts lower by a factor of 12 for a dispersion enriched with Cheng et al. postulated scavenger migration from membrane
cerium ions than for a pristine polymer dispersion.108 These to electrode layers. He demonstrated accelerated performance
studies also indicate scavenging properties of cerium, but also degradation aer AST for MEAs with Ce and Mn ions added to
point out the role of membrane morphology in the radical the membrane.277 This author assumed that this higher
attack as different degradation products were found for performance degradation might be due to increased cathode
membrane and polymer dispersion. catalyst layer (CCL) ionic losses. However further studies
Endoh et al. investigated the inuence of the incorporation showed no impact of scavengers on Pt dissolution, agglomera-
of Ce3+ ions in PFSA membranes on their degradation under the tion, carbon corrosion or oxidation.
low humidity subzero conditions of 30  C.275 Aer 50 Trogadas and co-workers developed an alternative approach
temperature cycles from 70  C to 30  C, a 10% Ce3+ doped to include cation radical scavengers, namely incorporation of
MEA showed very low 1.6  1010 g cm2 h1 FER in contrast to cerium oxide in the membrane.280 Ceria particles, commonly
1  106 g cm2 h1 with pristine Naon®. This result indicates known as a crucial component in three-way catalysts, emerged
as an efficient hydrogen peroxide decomposition catalyst.306 A
CeO2 – Naon® composite membrane showed decrease in the
FER of greater than an order of magnitude aer accelerated
degradation testing.280 This observation was attributed to
quenching of the highly oxidative species by the non-
stoichiometric CeO2. This assumption nds a conrmation in
further studies of Prabhakaran et al. with a uorescent molec-
ular probe in conjunction with in situ uorescence spectroscopy
to investigate the membrane chemical degradation and CeO2
mitigation properties in an operating fuel cell.284,285 This is
probably due to the particular electron structure and mixed

Fig. 16 Mitigation of ionomer degradation as a function of total Ce-


ion content (Ce3+ + Ce4+). (a) HOc concentration and rate of PFSA
ionomer attack, expressed as fluoride emission rate (FER) assuming
a membrane thickness of 50 mm, at a carboxylic end-group concen-
tration of [–COOH] = 18 mM. (b) Fraction of HOc reacting with the
ionomer and Ce3+. Reproduced by permission of The Electrochemical Fig. 17 Scavenging mechanism of CeO2 and regeneration of Ce3+
Society from ref. 81. active sites. Reprinted with permission from ref. 284.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 429
View Article Online

Sustainable Energy & Fuels Review

valence state of CeOx.307–309 It was proposed that scavenging lower performance than the baseline MEA. On the other hand,
properties of ceria are mediated at oxygen vacancies or defects performance of the cell with CeO2 incorporated on the cathode
in the lattice structure.310 Thus the hydroxyl radical can be side was slightly higher than that of baseline MEA. The author
scavenged by Ce3+ ions, which are further oxidised to Ce4+. linked this result to the oxygen storage ability of CeO2.
Oxidised Ce3+ is regenerated in acidic media through the Increased local concentration of oxygen on the catalyst surface
mechanism displayed in Fig. 17. This regeneration of the Ce3+ might improve the electrocatalytic reaction. In another study,
form appears as important factor for CeO2 mitigation ability. the enhanced performance observed with a catalyst based on
To tune the Ce3+/Ce4+ ratio, Trogadas et al. proposed two self-assembled mesoporous carbon with ceria nanoparticles has
strategies.309 A rst approach was based on particle size control, been explained by improved water retention of the modied
as the concentration of oxygen vacancies and Ce3+ formation catalyst.313 The idea of using CeOx as an active and durable
might depend on particle size. The second strategy included catalyst support was followed by Lei et al.314 These authors re-
doping of ceria particles with Zr4+ cations in order to improve ported lower chemical degradation of Naon binder in catalyst
the stability and oxygen storage capacity of the CeO2 micro- layers during CV-cycles in the high voltage region when using
structure. To investigate the inuence of ceria particle size and CeO2 nanocubes–graphene oxide (CeO2–GO) catalyst support.
Zr incorporation, Ce0.25Zr0.75O2 nanoparticles and ceria parti- They observed a slightly increased performance of a single cell
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

cles with varying Ce3+ surface concentrations were synthesised with CeO2–GO when CeO2 loading was lower than 8 wt%. Other
incorporated in Naon, and operated in a fuel cell MEA. Aer studies based on ex situ electrochemical characterisation
accelerated in situ degradation testing, the MEAs with revealed that incorporation of a low content (between 2 and 4
composite membranes containing CeO2 or Ce0.25Zr0.75O2 wt%) of amorphous CeO2 in a Pt catalyst enhances its tolerance
demonstrated one order of magnitude lower uorine emission to SO2 and CO.315 It was reported that the presence of nano-scale
rate compared to that of an MEA with a pristine membrane. ceria facilitates the oxygen transfer thus assists CO oxidation
Furthermore, the Ce3+ surface concentration of incorporated and inhibits SO2 oxidation on Pt surfaces.
particles was reected in FER values. The general conclusion Weissbach et al. investigated stability of PFSA membranes
can be drawn from this study is that a higher Ce3+ content co-casted with CeO2, ZrO2 and yttria-stabilized zirconia (YSZ)
lowers the FER from OCV hold-tested MEAs. Following this exposed to Fenton reagent.316 It was found that incorporation of
concept, Trogadas et al.281 performed another interesting CeO2 into the membrane signicantly reduced FER, mass loss
investigation by platinum incorporation in the polymer and loss of sulfonic groups, while PFSA-ZrO2 and PFSA-YSZ
membrane. Naon® doped with CeO2 – supported platinum membranes did not present improved stability. Naon – CeO2
demonstrated over an order of magnitude reduction of uoride nanocomposite membranes prepared through a “self-assem-
emission rate. This system showed increased scavenging ability bled” route, where ceria nanoparticles were incorporated into
of cerium oxide nanoparticles independently of the Pt concen- the polymer structure through in situ sol–gel process, were
tration (0.5–2 wt%). This observation was explained in terms of studied by Wang et al.282 The composite material demonstrated
enhanced rate of Ce4+ to Ce3+ reduction in the presence of Pt. slightly lower proton conductivity at 100% RH, whereas at
Wang et al. also studied the durability and performance of a Pt/ relative humidity below 75%, it showed higher values than
CeO2-Naon composite membrane.311 This author pointed out pristine Naon®. Both in situ and ex situ accelerated degrada-
the higher performance of the Pt/CeO2-Naon composite tion tests revealed superior durability of self-assembled
membrane in comparison to NRE-211 membrane at 35% RH Naon® – CeO2 membranes in comparison to the non-
that can be due to a self-hydrating function of Pt particles, modied membrane, as well as to Naon® – CeO2
a concept rst introduced by Watanabe.312 Furthermore, the membranes prepared by recasting from CeOx – Naon colloidal
composite material demonstrated stable OCV over 180 h of AST suspension. D'Urso et al. studied the effect of the presence of
performed at 70  C/35% RH and H2/O2 as reactants. Aer the silica supported CeOx in an ePTFE reinforced membrane on its
pioneering studies of Trogadas with CeO2 as radical scavenger, chemical stability.221 The authors did not provide any FER or
several authors investigated membrane durability with CeO2 OCV decay measurement to assess the efficiency of their system.
additives in the MEA components. Similarly to the above However they reported seven times longer lifetime of MEA
described Ce and Mn ion radical scavengers, CeO2 can be enriched with CeOx in comparison to the non-modied
incorporated in both the membrane and electrode structure. membrane. Also Pearman et al. performed in situ and ex situ
Wang and co-workers reported the reduction of ionomer membrane degradation tests on composite materials consisting
degradation through incorporation of CeO2 in the catalyst of porous PTFE supported cerium oxide impregnated with
layer.283 A newly prepared catalyst was applied either on the Naon.286,287 Mechanical reinforcement of the membrane,
anode or the cathode side. MEAs containing CeO2, regardless of enhanced with radical scavenging ability of ceria resulted in
the initial position of the radical scavenger, showed similar high durability over 500 h in an OCV hold test. Moreover the
improvement of the membrane durability aer in situ OCV hold inuence of the CeO2 incorporation on Pt band formation was
testing performed at 90  C and 30% RH in comparison to the reported. It was found that Pt formed larger particles deposited
baseline MEA. Aer 200 hours of accelerated testing, no further into the polymer membrane. The author connected this
membrane thinning or pinhole development was observed for behaviour with the presence of the ceria particles which affected
the MEA with anode or cathode cerium-rich catalyst. The MEA the potential prole through the membrane thus altered the Pt
with CeO2 incorporated in the anode demonstrated a slightly band. Contrary to reports by Trogadas et al., on the inuence of

430 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

size of ceria particles on FER, Pearman et al. did not nd scavenger loading, nature of the introduced species, or their
evidence for a lower FER for small CeOx in comparison to larger, location on the membrane durability can be compared using
commercial ceria particles. A recent study of Schlick et al. shed the uoride emission rate.
some new light on the subject.317 These authors emphasize that
the effectiveness of Ce nanoparticles may depend not only on
the initial particle size or oxidation state but also on the extent 11. Summary and outlook
of particle dispersion and agglomeration.
It might be concluded that ceria is very efficient radical The objective was to review and analyse the reaction mecha-
scavenger. However it should be noted that recent studies on nisms leading to PFSA membrane degradation and failure, and
MEAs containing either cerium ions or cerium oxide clearly to examine methodologies and means to reduce this degrada-
showed migration of cerium ions through the MEA compo- tion in an operating fuel cell. As discussed in Section 3,
nents,287,318–320 which indicates dissolution of ceria in acid membrane degradation is not a single isolated reaction and
medium and creates a new question regarding the stability of cannot be classied by any known model. It is a complex
ceria during fuel cell operation. Prabhakaran et al. reported loss combination of different factors including intrinsic properties
of the efficiency of ceria already aer 7 hours of an accelerated of the material used for MEA preparation and the operating
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

stress test, based on in situ uorescence spectroscopy conditions applied. This work has reviewed the factors that can
studies.284,285 However the authors related this observation to promote membrane degradation and which, once clearly iden-
a decreased concentration of Ce3+ reactive centers rather than to tied, provide the basis for design of materials with enhanced
dissolution of ceria particles. In order to improve its regenera- robustness. Beyond that, this review has also provided the rst
tive abilities, ceria nanoparticles have been doped with nitrogen survey of mitigation strategies designed to annihilate the
by annealing in nitrogen-rich atmosphere.321 Aer treatment harmful effect of oxidative species on membrane integrity and
the N-tuned CeOx demonstrated high efficiency in radical has comprehensively appraised their effectiveness by discus-
scavenging over 90 h of testing which was associated with sing the benets and drawbacks of each strategy. By identifying
increased number of Ce3+ active clusters aer exposure to the N- stressors and through improved understanding of how scav-
rich atmosphere. The possibility of ceria dissolution and cerium enging reactions proceed can certainly accelerate development
ion migration through MEA components was not considered. of new materials with antioxidant properties and enhanced
A detailed analysis of the effect of positioning of ceria in the durability. To meet fuel cell durability and lifetime targets over
MEA was performed by Zatoń et al.319,322 The authors used the widest range of temperature, pressure and relative humidity
electrospinning to develop a thin protective composite layer of operating conditions requires membranes that are highly
PFSA nanobers enriched with cerium oxide particles – NFCeOx heterogeneous, comprising not only a reinforcing component,
that was incorporated into the MEA at the desired anode/ or designed architecture at nano- to micro-metric length scales,
cathode interface. In the previous reports incorporation of but also an advanced radical scavenger component and/or
CeOx has been realized through preparation of a ceria enriched hydrogen peroxide decomposition catalyst. The choice of
membrane281,287 or catalyst layer.318,323–325 The presence of CeOx radical scavenger will be dictated by its efficiency and stability
in the membrane bulk can restrict the transport pathways and in operating conditions. Controlled release of a radical scav-
increase tortuosity and thus decrease proton conductivity. enger or hydrogen peroxide decomposition catalysts in
Moreover such an approach does not allow studies of the response to a chemical signal from a repository within the
impact of ceria particles on membrane degradation when membrane would represent a paradigm shi by providing the
located at respective anode and cathode sides. The incorpora- requisite groundwork for the elaboration of self-healing fuel cell
tion of CeOx in the catalyst layer gives such possibility, however membranes. Further development of the other cell compo-
affects the MEA performance as the charge transfer of ceria is nents: GDL, catalyst and bipolar plates will also have, of course,
rather low. Zaton et al. demonstrated that the lifetime of MEAs signicant inuence on membrane stability. This is especially
with anode or cathode sides protected by NFCeOx was eight salient in the context of emerging non platinum group metals
times longer relative to an unmitigated MEA and FER and OCV cathode catalysts of which the most active are iron or cobalt
decay were signicantly reduced. The authors also reported that based. Chemical stabilisation of PFSA membranes in an MEA
additional protective interlayer was more effective when incor- with a catalyst based on Fe2+ ions will require the development
porated on the anode side, possibly due to the reductive envi- of radical scavengers of magnied efficacy, to which early
ronment that accelerates regeneration of active Ce3+ scavenging consideration should be given in support of a future generation
centers, as well as a result of partial dissolution of CeOx at the of fuel cell catalysts using earth-abundant metals.
anode interface and migration of cerium ions through
membrane. In contrast, cerium species created aer dissolution
of CeOx on the cathode side might be easily leached from the Acknowledgements
MEA.
Table 3 summarises the results obtained aer AST on various The research leading to these results has received funding from
composite materials with scavenging properties. This provides the European Community's Seventh Framework Programme
a clearer view on the state of the art for mitigation of membrane (FP7/2010-2013) for the Fuel Cells and Hydrogen Joint Under-
degradation. Moreover the effectiveness of parameters like taking under grant agreement IMMEDIATE no. 3034.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 431
View Article Online

Sustainable Energy & Fuels Review

25 A. Sadeghi Alavijeh, M. A. Goulet, R. M. H. Khorasany,


References
J. Ghataurah, C. Lim, M. Lauritzen, E. Kjeang, G. G. Wang
1 A. J. Martin, A. Hornes, A. Martinez-Arias and L. Daza, and R. K. N. D. Rajapakse, Fuel Cells, 2015, 15, 204–213.
Recent advances in fuel cells for transport and stationary 26 F. A. de Bruijn, V. A. T. Dam and G. J. M. Janssen, Fuel Cells,
applications, Elsevier B.V., Amsterdam, 2013. 2008, 8, 3–22.
2 J. Garche and L. Jorissen, Electrochem. Soc. Interface, 2015, 27 L. Dubau, L. Castanheira, F. Maillard, M. Chatenet,
24, 3–43. O. Lottin, G. Maranzana, J. Dillet, A. Lamibrac,
3 Y. Wang, K. S. Chen, J. Mishler, S. C. Cho and X. C. Adroher, J.-C. Perrin, E. Moukheiber, A. ElKaddouri, G. De Moor,
Appl. Energy, 2011, 88, 981–1007. C. Bas, L. Flandin and N. Caqué, Wiley Interdiscip. Rev.:
4 T. Wilberforce, A. Alaswad, A. Palumbo, M. Dassisti and Energy Environ., 2014, 3, 540–560.
A. G. Olabi, Int. J. Hydrogen Energy, 2016, 41, 16509–16522. 28 S. Zhang, X. Yuan, H. Wang, W. Merida, H. Zhu, J. Shen,
5 W. Schmittinger and A. Vahidi, J. Power Sources, 2008, 180, S. Wu and J. Zhang, Int. J. Hydrogen Energy, 2009, 34, 388–
1–14. 404.
6 V. Andrea, P. d. S. P. Oliveira, E. I. Santiago and M. Linardi, 29 X.-Z. Yuan, H. Li, S. Zhang, J. Martin and H. Wang, J. Power
ECS Trans., 2016, 71, 233–238. Sources, 2011, 196, 9107–9116.
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

7 A. de Frank Bruijn and G. J. M. Janssen, PEM Fuel Cell 30 M. P. Rodgers, L. J. Bonville, H. R. Kunz, D. K. Slattery and
Materials: Costs, Performance and Durability, Springer New J. M. Fenton, Chem. Rev., 2012, 112, 6075–6103.
York, New York, NY, 2013. 31 T. Ishimoto and M. Koyama, Membranes, 2012, 2, 395–414.
8 M. K. Debe, Nature, 2012, 486, 43–51. 32 M. A. Hickner, H. Ghassemi, Y. S. Kim, B. R. Einsla and
9 J. Wu, X. Z. Yuan, J. J. Martin, H. Wang, J. Zhang, J. Shen, J. E. McGrath, Chem. Rev., 2004, 104, 4587–4612.
S. Wu and W. Merida, J. Power Sources, 2008, 184, 104–119. 33 J. Roziere and D. J. Jones, Annu. Rev. Mater. Res., 2003, 33,
10 W. Bi, G. E. Gray and T. F. Fuller, Electrochem. Solid-State 503–555.
Lett., 2007, 10, B101–B104. 34 D. J. Jones and J. Rozière, Inorganic/organic composite
11 R. M. Darling and J. P. Meyers, J. Electrochem. Soc., 2003, membranes, John Wiley & Sons, Ltd, Chichester, UK, 2010.
150, A1523–A1527. 35 H. Zhang and P. K. Shen, Chem. Rev., 2012, 112, 2780–2832.
12 P. J. Ferreira, G. J. la O', Y. Shao-Horn, D. Morgan, 36 K. A. Mauritz and R. B. Moore, Chem. Rev., 2004, 104, 4535–
R. Makharia, S. Kocha and H. A. Gasteiger, J. Electrochem. 4585.
Soc., 2005, 152, A2256–A2271. 37 C. Wang, G. Duscher and S. J. Paddison, RSC Adv., 2015, 5,
13 K. Yasuda, A. Taniguchi, T. Akita, T. Ioroi and Z. Siroma, 2368–2373.
Phys. Chem. Chem. Phys., 2006, 8, 746–752. 38 D. J. Jones, 2-Membrane materials and technology for low
14 A. Ohma, S. Yamamoto and K. Shinohara, ECS Trans., 2007, temperature fuel cells, Woodhead Publishing, Cambridge,
11, 1181–1192. 2012.
15 A. Ohma, S. Yamamoto and K. Shinohara, J. Power Sources, 39 The Dow Chemical Company, EP Pat., EP0041735A1, 1981.
2008, 182, 39–47. 40 Y. M. Tsou, M. C. Kimble and R. E. White, J. Electrochem.
16 C. A. Reiser, L. Bregoli, T. W. Patterson, J. S. Yi, J. D. Yang, Soc., 1992, 139, 1913–1917.
M. L. Perry and T. D. Jarvi, Electrochem. Solid-State Lett., 41 G. A. Eisman, J. Power Sources, 1990, 29, 389–398.
2005, 8, A273–A276. 42 A. Ghielmi, P. Vaccarono, C. Troglia and V. Arcella, J. Power
17 M. Watanabe, K. Tsurumi, T. Mizukami, T. Nakamura and Sources, 2005, 145, 108–115.
P. Stonehart, J. Electrochem. Soc., 1994, 141, 2659–2668. 43 V. Arcella, C. Troglia and A. Ghielmi, Ind. Eng. Chem. Res.,
18 D. Spernjak, J. D. Fairweather, T. Rockward, R. Mukundan 2005, 44, 7646–7651.
and R. Borup, ECS Trans., 2011, 41, 741–750. 44 J. Li, M. Pan and H. Tang, RSC Adv., 2014, 4, 3944–3965.
19 Z. Y. Liu, B. K. Brady, R. N. Carter, B. Litteer, M. Budinski, 45 Q. Zhao and J. Benziger, J. Polym. Sci., Part B: Polym. Phys.,
J. K. Hyun and D. A. Muller, J. Electrochem. Soc., 2008, 155, 2013, 51, 915–925.
B979–B984. 46 X. Luo, S. Holdcro, A. Mani, Y. Zhang and Z. Shi, Phys.
20 R. Borup, J. Meyers and B. Pivovar, Chem. Rev., 2007, 107, Chem. Chem. Phys., 2011, 13, 18055–18062.
3904–3951. 47 A. S. Aricò, A. Di Blasi, G. Brunaccini, F. Sergi, G. Dispenza,
21 F. Nandjou, J. P. Poirot-Crouvezier, M. Chandesris, L. Andaloro, M. Ferraro, V. Antonucci, P. Asher, S. Buche,
J. F. Blachot, C. Bonnaud and Y. Bultel, J. Power Sources, D. Fongalland, G. A. Hards, J. D. B. Sharman, A. Bayer,
2016, 326, 182–192. G. Heinz, N. Zandonà, R. Zuber, M. Gebert, M. Corasaniti,
22 T. Okada, Effect of ionic contaminants, John Wiley & Sons, A. Ghielmi and D. J. Jones, Fuel Cells, 2010, 10, 1013–1023.
Ltd, Chichester, UK, 2003. 48 A. Stassi, I. Gatto, E. Passalacqua, V. Antonucci, A. S. Arico,
23 N. Zamel and X. Li, Prog. Energy Combust. Sci., 2011, 37, L. Merlo, C. Oldani and E. Pagano, J. Power Sources, 2011,
292–329. 196, 8925–8930.
24 A. Collier, H. Wang, X. Zi Yuan, J. Zhang and 49 M. Marrony, D. Beretta, S. Ginocchio, Y. Nedellec,
D. P. Wilkinson, Int. J. Hydrogen Energy, 2006, 31, 1838– S. Subianto and D. J. Jones, Fuel Cells, 2013, 13, 1146–1154.
1854. 50 H. Tang, S. Peikang, S. P. Jiang, F. Wang and M. Pan, J.
Power Sources, 2007, 170, 85–92.

432 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

51 R. M. H. Khorasany, E. Kjeang, G. G. Wang and 77 M. Plazanet, F. Sacchetti, C. Petrillo, B. Deme, P. Bartolini


R. K. N. D. Rajapakse, J. Power Sources, 2015, 279, 55–63. and R. Torre, J. Membr. Sci., 2014, 453, 419–424.
52 X. Huang, M. Rodgers, W. Yoon, B. Li and N. Mohajeri, ECS 78 S. D. Knights, K. M. Colbow, J. St-Pierre and
Trans., 2008, 16, 1573–1579. D. P. Wilkinson, J. Power Sources, 2004, 127, 127–134.
53 Y. Xiao and C. Cho, Energies, 2014, 7, 6401–6411. 79 F. Wang, H. Tang, M. Pan and D. Li, Int. J. Hydrogen Energy,
54 R. M. H. Khorasany, A. Sadeghi Alavijeh, E. Kjeang, 2008, 33, 2283–2288.
G. G. Wang and R. K. N. D. Rajapakse, J. Power Sources, 80 L. Gubler, S. M. Dockheer and W. H. Koppenol, J.
2015, 274, 1208–1216. Electrochem. Soc., 2011, 158, B755–B769.
55 N.-I. Kim, Y. Seo, K. B. Kim, N. Lee, J.-H. Lee, I. Song, 81 L. Gubler and W. H. Koppenol, J. Electrochem. Soc., 2011,
H. Choi and J.-Y. Park, J. Power Sources, 2014, 253, 90–97. 159, B211–B218.
56 R. Lin, F. Xiong, W. C. Tang, L. Técher, J. M. Zhang and 82 N. Ohguri, A. Y. Nosaka and Y. Nosaka, Electrochem. Solid-
J. X. Ma, J. Power Sources, 2014, 260, 150–158. State Lett., 2009, 12, B94–B96.
57 L. Dubau, L. Castanheira, M. Chatenet, F. Maillard, J. Dillet, 83 N. Ohguri, A. Y. Nosaka and Y. Nosaka, J. Power Sources,
G. Maranzana, S. Abbou, O. Lottin, G. De Moor, A. El 2010, 195, 4647–4652.
Kaddouri, C. Bas, L. Flandin, E. Rossinot and N. Caqué, 84 Y. Nosaka, K. Ohtaka, N. Ohguri and A. Y. Nosaka, ECS
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

Int. J. Hydrogen Energy, 2014, 39, 21902–21914. Trans., 2010, 33, 899–905.
58 Y. Tang, A. M. Karlsson, M. H. Santare, M. Gilbert, 85 Y. Nosaka, K. Ohtaka, N. Ohguri and A. Y. Nosaka, J.
S. Cleghorn and W. B. Johnson, Mater. Sci. Eng., A, 2006, Electrochem. Soc., 2011, 158, B430–B433.
425, 297–304. 86 E. Endoh, S. Terazono, H. Widjaja and Y. Takimoto,
59 B. Wu, M. A. Parkes, L. de Benedetti, A. J. Marquis, Electrochem. Solid-State Lett., 2004, 7, A209–A211.
G. J. Offer and N. P. Brandon, J. Appl. Electrochem., 2016, 87 W. Liu and D. Zuckerbrod, J. Electrochem. Soc., 2005, 152,
46, 1157–1162. A1165–A1170.
60 M. Inaba, T. Kinumoto, M. Kiriake, R. Umebayashi, 88 H. S. Casalongue, S. Kaya, V. Viswanathan, D. J. Miller,
A. Tasaka and Z. Ogumi, Electrochim. Acta, 2006, 51, 5746– D. Friebel, H. A. Hansen, J. K. Nørskov, A. Nilsson and
5753. H. Ogasawara, Nat. Commun., 2013, 4, 2817.
61 H. S. Sodaye, P. K. Pujari, A. Goswami and S. B. Manohar, J. 89 A. Pozio, R. F. Silva, F. M. De and L. Giorgi, Electrochim.
Polym. Sci., Part B: Polym. Phys., 1997, 35, 771–776. Acta, 2003, 48, 1543–1549.
62 H. S. Sodaye, P. K. Pujari, A. Goswami and S. B. Manohar, 90 V. O. Mittal, H. Russell Kunz and J. M. Fenton, Electrochem.
Radiat. Phys. Chem., 2000, 58, 567–570. Solid-State Lett., 2006, 9, A299–A302.
63 M. N. Silberstein and M. C. Boyce, J. Power Sources, 2010, 91 V. O. Mittal, H. R. Kunz and J. M. Fenton, J. Electrochem.
195, 5692–5706. Soc., 2007, 154, B652–B656.
64 M. Ozmaian and R. Naghdabadi, PCCP, 2014, 16, 3173– 92 X. Fang, P. K. Shen, S. Song, V. Stergiopoulos and
3186. P. Tsiakaras, Polym. Degrad. Stab., 2009, 94, 1707–1713.
65 S. Kundu, L. C. Simon, M. Fowler and S. Grot, Polymer, 93 M. Danilczuk, F. D. Coms and S. Schlick, J. Phys. Chem. B,
2005, 46, 11707–11715. 2009, 113, 8031–8042.
66 P. W. Majsztrik, A. B. Bocarsly and J. B. Benziger, Rev. Sci. 94 A. Panchenko, H. Dilger, J. Kerres, M. Hein, A. Ullrich,
Instrum., 2007, 78, 103904. T. Kaz and E. Roduner, PCCP, 2004, 6, 2891–2894.
67 F. Bauer, S. Denneler and M. Willert-Porada, J. Polym. Sci., 95 N. Ramaswamy, N. Hakim and S. Mukerjee, Electrochim.
Part B: Polym. Phys., 2005, 43, 786–795. Acta, 2008, 53, 3279–3295.
68 X. Luo, L. Ghassemzadeh and S. Holdcro, Int. J. Hydrogen 96 S. Xiao, H. Zhang, C. Bi, Y. Zhang, Y. Zhang, H. Dai, Z. Mai
Energy, 2015, 40, 16714–16723. and X. Li, J. Power Sources, 2010, 195, 5305–5311.
69 S. v. Venkatesan, C. Lim, E. Rogers, S. Holdcro and 97 M. Zhao, W. Shi, B. Wu, W. Liu, J. Liu, D. Xing, Y. Yao,
E. Kjeang, Phys. Chem. Chem. Phys., 2015, 17, 13872–13881. Z. Hou, P. Ming and Z. Zou, Electrochim. Acta, 2015, 153,
70 W. Shi and L. A. Baker, RSC Adv., 2015, 5, 99284–99290. 254–262.
71 S. v. Venkatesan, C. Lim, S. Holdcro and E. Kjeang, J. 98 S. Zhang, X.-Z. Yuan, R. Hiesgen, K. A. Friedrich, H. Wang,
Electrochem. Soc., 2016, 163, F637–F643. M. Schulze, A. Haug and H. Li, J. Power Sources, 2012, 205,
72 Z. Lu, G. Polizos, D. D. Macdonald and E. Manias, J. 290–300.
Electrochem. Soc., 2008, 155, B163–B171. 99 S. Mu, C. Xu, Q. Yuan, Y. Gao, F. Xu and P. Zhao, J. Appl.
73 Y. S. Kim, L. Dong, M. A. Hickner, T. E. Glass, V. Webb and Polym. Sci., 2013, 129, 1586–1592.
J. E. McGrath, Macromolecules, 2003, 36, 6281–6285. 100 L. Ghassemzadeh, K. D. Kreuer, J. Maier and K. Muller, J.
74 Z. Lu, G. Polizos, E. Manias and D. Macdonald, ECS Trans., Power Sources, 2011, 196, 2490–2497.
2010, 28, 81–89. 101 S. Kundu, L. C. Simon and M. W. Fowler, Polym. Degrad.
75 F. Teocoli, A. Paolone, O. Palumbo, M. A. Navarra, Stab., 2008, 93, 214–224.
M. Casciola and A. Donnadio, J. Polym. Sci., Part B: Polym. 102 M. Danilczuk, A. Bosnjakovic, M. K. Kadirov and S. Schlick,
Phys., 2012, 50, 1421–1425. J. Power Sources, 2007, 172, 78–82.
76 R. C. McDonald, C. K. Mittelsteadt and E. L. Thompson, 103 A. A. Shah, T. R. Ralph and F. C. Walsh, J. Electrochem. Soc.,
Fuel Cells, 2004, 4, 208–213. 2009, 156, B465–B484.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 433
View Article Online

Sustainable Energy & Fuels Review

104 D. E. Curtin, R. D. Lousenberg, T. J. Henry, P. C. Tangeman 134 V. A. Sethuraman, J. W. Weidner, A. T. Haug, S. Motupally
and M. E. Tisack, J. Power Sources, 2004, 131, 41–48. and L. V. Protsailo, J. Electrochem. Soc., 2007, 155, B50–B57.
105 T. Xie and C. A. Hayden, Polymer, 2007, 48, 5497–5506. 135 H. Liu, H. A. Gasteiger, A. Laconti and J. Zhang, ECS Trans.,
106 S. Hommura, K. Kawahara, T. Shimohira and Y. Teraoka, J. 2006, 1, 283–293.
Electrochem. Soc., 2008, 155, A29–A33. 136 J. Peron, Y. Nedellec, D. J. Jones and J. Roziere, J. Power
107 K. E. Schwiebert, K. G. Raiford, G. Escobedo and Sources, 2008, 185, 1209–1217.
G. Nagarajan, ECS Trans., 2006, 1, 303–311. 137 H. A. Gasteiger, S. S. Kocha, B. Sompalli and F. T. Wagner,
108 M. Danilczuk, A. J. Perkowski and S. Schlick, Appl. Catal., B, 2005, 56, 9–35.
Macromolecules, 2010, 43, 3352–3358. 138 K. Ono, Y. Yasuda, K. Sekizawa, N. Takeuchi, T. Yoshida
109 C. Zhou, M. A. Guerra, Z.-M. Qiu, T. A. Zawodzinski and and M. Sudoh, Electrochim. Acta, 2013, 97, 58–65.
D. A. Schiraldi, Macromolecules, 2007, 40, 8695–8707. 139 U. A. Paulus, T. J. Schmidt, H. A. Gasteiger and R. J. Behm, J.
110 N. E. Cipollini, ECS Trans., 2007, 11, 1071–1082. Electroanal. Chem., 2001, 495, 134–145.
111 D. A. Schiraldi, Polym. Rev., 2006, 46, 315–327. 140 K. Ke, T. Hatanaka and Y. Morimoto, Electrochim. Acta,
112 G. Escobedo, K. Raiford, G. S. Nagarajan and 2010, 56, 2098–2104.
K. E. Schwiebert, ECS Trans., 2006, 1, 303–311. 141 M. Inaba, H. Yamada, J. Tokunaga and A. Tasaka,
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

113 L. Ghassemzadeh, K.-D. Kreuer, J. Maier and K. Muller, J. Electrochem. Solid-State Lett., 2004, 7, A474–A476.
Phys. Chem. C, 2010, 114, 14635–14645. 142 A. Bonakdarpour, T. R. Dahn, R. T. Atanasoski, M. K. Debe
114 A. Bosnjakovic, M. Kadirov and S. Schlick, Res. Chem. and J. R. Dahn, Electrochem. Solid-State Lett., 2008, 11,
Intermed., 2007, 33, 677–687. B208–B211.
115 D. Kurniawan, H. Arai, S. Morita and K. Kitagawa, 143 N. M. Markovic, H. A. Gasteiger, B. N. Grgur and P. N. Ross,
Microchem. J., 2013, 106, 384–388. J. Electroanal. Chem., 1999, 467, 157–163.
116 T. Tokumasu, I. Ogawa, M. Koyama, T. Ishimoto and 144 V. Atrazhev, S. F. Burlatsky, N. E. Cipollini, D. A. Condit and
A. Miyamoto, J. Electrochem. Soc., 2011, 158, B175–B179. N. Erikhman, ECS Trans., 2006, 1, 239–246.
117 F. D. Coms, ECS Trans., 2008, 16, 235–255. 145 V. Atrazhev, E. Timokhina, S. F. Burlatsky, V. Sultanov,
118 L. Ghassemzadeh, T. J. Peckham, T. Weissbach, X. Luo and T. Madden and M. Gummalla, ECS Trans., 2008, 6, 69–74.
S. Holdcro, J. Am. Chem. Soc., 2013, 135, 15923–15932. 146 F. N. Buchi, B. Gupta, O. Haas and G. G. Scherer,
119 L. Ghassemzadeh and S. Holdcro, J. Am. Chem. Soc., 2013, Electrochim. Acta, 1995, 40, 345–353.
135, 8181–8184. 147 M. Marrony, R. Barrera, S. Quenet, S. Ginocchio,
120 A. M. Dreizler and E. Roduner, Fuel Cells, 2012, 12, 132–140. L. Montelatici and A. Aslanides, J. Power Sources, 2008,
121 M. Danilczuk, F. D. Coms and S. Schlick, Fuel Cells, 2008, 8, 182, 469–475.
436–452. 148 M. Aoki, H. Uchida and M. Watanabe, Electrochem.
122 T. Ishimoto, T. Ogura and M. Koyama, ECS Trans., 2011, 35, Commun., 2005, 7, 1434–1438.
1–6. 149 L. Merlo, A. Ghielmi, L. Cirillo, M. Gebert and V. Arcella, J.
123 M. Danilczuk, L. Lancucki, S. Schlick, S. J. Hamrock and Power Sources, 2007, 171, 140–147.
G. M. Haugen, ACS Macro Lett., 2012, 1, 280–285. 150 V. A. Sethuraman, J. W. Weidner, A. T. Haug, S. Motupally
124 J. Healy, C. Hayden, T. Xie, K. Olson, R. Waldo, and L. V. Protsailo, J. Electrochem. Soc., 2008, 155, B50–B57.
M. Brundage, H. Gasteiger and J. Abbott, Fuel Cells, 2005, 151 C. Huang, K. Seng Tan, J. Lin and K. Lee Tan, Chem. Phys.
5, 302–308. Lett., 2003, 371, 80–85.
125 C. Lim, L. Ghassemzadeh, F. Van Hove, M. Lauritzen, 152 E. Endoh, S. Hommura, S. Terazono, H. Widjaja and
J. Kolodziej, G. G. Wang, S. Holdcro and E. Kjeang, J. J. Anzai, ECS Trans., 2007, 11, 1083–1091.
Power Sources, 2014, 257, 102–110. 153 M. Bodner, B. Cermenek, M. Rami and V. Hacker,
126 K. H. Wong and E. Kjeang, J. Electrochem. Soc., 2014, 161, Membranes, 2015, 5, 888.
F823–F832. 154 G. De Moor, C. Bas, N. Charvin, J. Dillet, G. Maranzana,
127 M. Ghelichi, P.-E. A. Melchy and M. H. Eikerling, J. Phys. O. Lottin, N. Caqué, E. Rossinot and L. Flandin, Int. J.
Chem. B, 2014, 118, 11375–11386. Hydrogen Energy, 2016, 41, 483–496.
128 T. H. Yu, Y. Sha, W.-G. Liu, B. V. Merinov, P. Shirvanian and 155 J. Yu, B. Yi, D. Xing, F. Liu, Z. Shao, Y. Fu and H. Zhang,
W. A. Goddard, J. Am. Chem. Soc., 2011, 133, 19857–19863. PCCP, 2003, 5, 611–615.
129 T. Ishimoto, R. Nagumo, T. Ogura, T. Ishihara, B. Kim, 156 W. Yoon and X. Y. Huang, J. Electrochem. Soc., 2010, 157,
A. Miyamoto and M. Koyama, J. Electrochem. Soc., 2010, B599–B606.
157, B1305–B1309. 157 B. Mattsson, H. Ericson, L. M. Torell and F. Sundholm,
130 K. Teranishi, K. Kawata, S. Tsushima and S. Hirai, Electrochim. Acta, 2000, 45, 1405–1408.
Electrochem. Solid-State Lett., 2006, 9, A475–A477. 158 M. P. Rodgers, R. P. Brooker, N. Mohajeri, L. J. Bonville,
131 C. Chen and T. Fuller, ECS Trans., 2007, 11, 1127–1137. H. R. Kunz, D. K. Slattery and J. M. Fenton, J. Electrochem.
132 A. B. LaConti, M. Hamdan and R. C. McDonald, in Soc., 2012, 159, F338–F352.
Handbook of Fuel Cells, John Wiley & Sons, Ltd, 2003. 159 K. H. Wong and E. Kjeang, ChemSusChem, 2015, 8, 1072–1082.
133 J. Qiao, M. Saito, K. Hayamizu and T. Okada, J. Electrochem. 160 S. A. Vilekar and R. Datta, J. Power Sources, 2010, 195, 2241–
Soc., 2006, 153, A967–A974. 2247.

434 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

161 K. D. Baik, I. M. Kong, B. K. Hong, S. H. Kim and M. S. Kim, 185 D. Zhao, B. L. Yi, H. M. Zhang and M. Liu, J. Power Sources,
Appl. Energy, 2010, 101, 560–566. 2010, 195, 4606–4612.
162 J. Zhang, Y. Tang, C. Song, J. Zhang and H. Wang, J. Power 186 M. Gummalla, V. V. Atrazhev, D. Condit, N. Cipollini,
Sources, 2006, 163, 532–537. T. Madden, N. Y. Kuzminyh, D. Weiss and S. F. Burlatsky,
163 H. L. Yeager and A. Steck, J. Electrochem. Soc., 1981, 128, J. Electrochem. Soc., 2010, 157, B1542–B1548.
1880–1884. 187 T. Kim, H. Lee, W. Sim, J. Lee, S. Kim, T. Lim and K. Park,
164 Z. Ogumi, Z. Takehara and S. Yoshizawa, J. Electrochem. Korean J. Chem. Eng., 2009, 26, 1265–1271.
Soc., 1984, 131, 769–773. 188 H. Liu, J. Zhang, F. Coms, W. Gu, B. Litteer and
165 L. Liu, A. Chakma and X. Feng, J. Membr. Sci., 2008, 310, 66– H. A. Gasteiger, ECS Trans., 2006, 3, 493–505.
75. 189 C. Iojoiu, E. Guilminot, F. Maillard, M. Chatenet,
166 S. Kundu, M. W. Fowler, L. C. Simon, R. Abouatallah and J. Y. Sanchez, E. Claude and E. Rossinot, J. Electrochem.
N. Beydokhti, J. Power Sources, 2010, 195, 7323–7331. Soc., 2007, 154, B1115–B1120.
167 K. Panha, M. Fowler, X.-Z. Yuan and H. Wang, Appl. Energy, 190 A. Ohma, S. Suga, S. Yamamoto and K. Shinohara, J.
2012, 93, 90–97. Electrochem. Soc., 2007, 154, B757–B760.
168 V. Prabhakaran, C. G. Arges and V. Ramani, PCCP, 2013, 15, 191 N. Macauley, L. Ghassemzadeh, C. Lim, M. Watson,
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

18965–18972. J. Kolodziej, M. Lauritzen, S. Holdcro and E. Kjeang,


169 T. Madden, D. Weiss, N. Cipollini, D. Condit, ECS Electrochem. Lett., 2013, 2, F33–F35.
M. Gummalla, S. Burlatsky and V. Atrazhev, J. 192 N. Macauley, A. S. Alavijeh, M. Watson, J. Kolodziej,
Electrochem. Soc., 2009, 156, B657–B662. M. Lauritzen, S. Knights, G. Wang and E. Kjeang, J.
170 M. Takasaki, Y. Nakagawa, Y. Sakiyama, K. Tanabe, Electrochem. Soc., 2015, 162, F98–F107.
K. Ookubo, N. Sato, T. Minamide, H. Nakayama and 193 N. Macauley, K. H. Wong, M. Watson and E. Kjeang, J.
M. Hori, J. Electrochem. Soc., 2013, 160, F413–F416. Power Sources, 2015, 299, 139–148.
171 K. Matsuoka, S. Sakamoto, K. Nakato, A. Hamada and 194 S. Helmly, B. Ohnmacht, P. Gazdzicki, R. Hiesgen,
Y. Itoh, J. Power Sources, 2008, 179, 560–565. E. Gülzow and K. A. Friedrich, J. Electrochem. Soc., 2014,
172 D. Imamura and E. Yamaguchi, ECS Trans., 2009, 25, 813– 161, F1416–F1426.
819. 195 S. Helmly, B. Ohnmacht, R. Hiesgen, E. Gluzow and
173 M. Chen, C. Du, J. Zhang, P. Wang and T. Zhu, J. Power K. A. Friedrich, ECS Trans., 2013, 58, 969–990.
Sources, 2011, 196, 620–626. 196 M. P. Rodgers, B. P. Pearman, L. J. Bonville, D. A. Cullen,
174 X. Zhang, H. M. Galindo, H. F. Garces, P. Baker, X. Wang, N. Mohajeri and D. K. Slattery, J. Electrochem. Soc., 2013,
U. Pasaogullari, S. L. Suib and T. Molter, J. Electrochem. 160, F1123–F1128.
Soc., 2010, 157, B409–B414. 197 C. H. Choi, C. Baldizzone, J.-P. Grote, A. K. Schuppert,
175 M. Sulek, J. Adams, S. Kaberline, M. Ricketts and F. Jaouen and K. J. J. Mayrhofer, Angew. Chem., Int. Ed.,
J. R. Waldecker, J. Power Sources, 2011, 196, 8967–8972. 2015, 54, 12753–12757.
176 H. Li, K. Tsay, H. Wang, J. Shen, S. Wu, J. Zhang, N. Jia, 198 C. H. Choi, C. Baldizzone, G. Polymeros, E. Pizzutilo,
S. Wessel, R. Abouatallah, N. Joos and J. Schrooten, J. O. Kasian, A. K. Schuppert, N. Ranjbar Sahraie,
Power Sources, 2010, 195, 8089–8093. M.-T. Sougrati, K. J. J. Mayrhofer and F. Jaouen, ACS
177 J. M. Christ, K. C. Neyerlin, H. Wang, R. Richards and Catal., 2016, 6, 3136–3146.
H. N. Dinh, J. Electrochem. Soc., 2014, 161, F1481– 199 D. Banham, S. Ye, K. Pei, J.-i. Ozaki, T. Kishimoto and
F1488. Y. Imashiro, J. Power Sources, 2015, 285, 334–348.
178 V. Berejnov, Z. Martin, M. West, S. Kundu, D. Bessarabov, 200 S. Xiao, H. Zhang, C. Bi, Y. Zhang, Y. Ma, X. Li, H. Zhong
J. Stumper, D. Susac and A. P. Hitchcock, Phys. Chem. and Y. Zhang, J. Power Sources, 2010, 195, 8000–8005.
Chem. Phys., 2012, 14, 4835–4843. 201 M. Hu and G. Cao, Int. J. Hydrogen Energy, 2014, 39, 7940–
179 J. Xie, D. L. Wood, K. L. More, P. Atanassov and R. L. Borup, 7954.
J. Electrochem. Soc., 2005, 152, A1011–A1020. 202 H.-C. Chien, L.-D. Tsai, C.-M. Lai, J.-N. Lin, C.-Y. Zhu and
180 S. F. Burlatsky, M. Gummalla, V. V. Atrazhev, F.-C. Chang, J. Power Sources, 2013, 226, 87–93.
D. V. Dmitriev, N. Y. Kuzminyh and N. S. Erikhman, J. 203 S. Subianto, M. Pica, M. Casciola, P. Cojocaru, L. Merlo,
Electrochem. Soc., 2011, 158, B322–B330. G. Hards and D. J. Jones, J. Power Sources, 2013, 233, 216–
181 L. Kim, C. G. Chung, Y. W. Sung and J. S. Chung, J. Power 230.
Sources, 2008, 183, 524–532. 204 J. A. Kolde, B. Bahar, M. S. Wilson, T. A. Zawodzinski and
182 J. Peron, D. Jones and J. Roziere, ECS Trans., 2007, 11, 1313– S. Gottesfeld, Proc.–Electrochem. Soc., 1995, 95–23, 193–201.
1319. 205 B. Bahar, C. Cavalca, S. Cleghorn, J. Kolde, D. Lane,
183 T. Hatanaka, T. Takeshita, H. Murata, N. Hasegawa, M. Murthy and G. Rusch, J. New Mater. Electrochem. Syst.,
T. Asano, M. Kawasumi and Y. Morimoto, ECS Trans., 1999, 2, 179–182.
2008, 16, 1961–1965. 206 G. Alberti, R. Narducci, M. L. Di Vona and S. Giancola, Fuel
184 S. Helmly, R. Hiesgen, T. Morawietz, X.-Z. Yuan, H. Wang Cells, 2013, 13, 42–47.
and K. Andreas Friedrich, J. Electrochem. Soc., 2013, 160, 207 G. Alberti, R. Narducci and M. Sganappa, J. Power Sources,
F687–F697. 2008, 178, 575–583.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 435
View Article Online

Sustainable Energy & Fuels Review

208 M. K. Hassan, A. Abukmail and K. A. Mauritz, Eur. Polym. J., 235 D. Yuan, Z. Liu, S. W. Tay, X. Fan, X. Zhang and C. He,
2012, 48, 789–802. Chem. Commun., 2013, 49, 9639–9641.
209 L. Maldonado, J.-C. Perrin, J. Dillet and O. Lottin, J. Membr. 236 N. H. Jalani, K. Dunn and R. Datta, Electrochim. Acta, 2005,
Sci., 2012, 389, 43–56. 51, 553–560.
210 A. Kusoglu, S. Savagatrup, K. T. Clark and A. Z. Weber, 237 C.-C. Ke, X.-J. Li, Q. Shen, S.-G. Qu, Z.-G. Shao and B.-L. Yi,
Macromolecules, 2012, 45, 7467–7476. Int. J. Hydrogen Energy, 2011, 36, 3606–3613.
211 X. Wang, H. Tang and M. Pan, J. Membr. Sci., 2011, 379, 238 K. T. Adjemian, R. Dominey, L. Krishnan, H. Ota,
106–111. P. Majsztrik, T. Zhang, J. Mann, B. Kirby, L. Gatto,
212 J. Li, X. Yang, H. Tang and M. Pan, J. Membr. Sci., 2010, 361, M. Velo-Simpson, J. Leahy, S. Srinivasan, J. B. Benziger
38–42. and A. B. Bocarsly, Chem. Mater., 2006, 18, 2238–
213 J. Lin, P. H. Wu, R. Wycisk, A. Trivisonno and 2248.
P. N. Pintauro, J. Power Sources, 2008, 183, 491–497. 239 Y. Patil, S. Kulkarni and K. A. Mauritz, J. Appl. Polym. Sci.,
214 W. Zhang, R. Wycisk, D. L. Kish and P. N. Pintauro, J. 2011, 121, 2344–2353.
Electrochem. Soc., 2014, 161, F770–F777. 240 V. Baglio, A. S. Arico, A. D. Blasi, V. Antonucci,
215 H. Ghassemi, T. Zawodzinski, D. Schiraldi and S. Hamrock, P. L. Antonucci, S. Licoccia, E. Traversa and F. S. Fiory,
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

ACS Symp. Ser., 2012, 1096, 201–220. Electrochim. Acta, 2005, 50, 1241–1246.
216 Y. M. Zhang, L. Li, J. Tang, B. Bauer, W. Zhang, H. R. Gao, 241 H. Uchida, Y. Ueno, H. Hagihara and M. Watanabe, J.
M. Taillades-Jacquin, D. J. Jones, J. Roziere, N. Lebedeva Electrochem. Soc., 2003, 150, A57–A62.
and R. Mallant, ECS Trans., 2009, 25, 1469–1472. 242 T. Jian-hua, G. Peng-fei, Z. Zhi-yuan, L. Wen-hui and
217 N. Uematsu, N. Hoshi, T. Koga and M. Ikeda, J. Fluorine S. Zhong-qiang, Int. J. Hydrogen Energy, 2008, 33, 5686–
Chem., 2006, 127, 1087–1095. 5690.
218 L. Sauguet, B. Ameduri and B. Boutevin, J. Polym. Sci., Part 243 C. F. Nørgaard, U. G. Nielsen and E. M. Skou, Solid State
A: Polym. Chem., 2006, 44, 4566–4578. Ionics, 2012, 213, 76–82.
219 J. Shim, H. Y. Ha, S.-A. Hong and I.-H. Oh, J. Power Sources, 244 F. Chen, A. D'Epifanio, B. Mecheri, E. Traversa and
2002, 109, 412–417. S. Licoccia, ECS Trans., 2009, 25, 1935–1941.
220 P. Xiao, J. Li, H. Tang, Z. Wang and M. Pan, J. Membr. Sci., 245 S. Brutti, R. Scipioni, M. A. Navarra, S. Panero, V. Allodi,
2013, 442, 65–71. M. Giarola and G. Mariotto, Int. J. Nanotechnol., 2014, 11,
221 C. D'Urso, C. Oldani, V. Baglio, L. Merlo and A. S. Aricò, J. 882–896.
Power Sources, 2014, 272, 753–758. 246 R. Scipioni, D. Gazzoli, F. Teocoli, O. Palumbo, A. Paolone,
222 Utc Power Corporation, WO Pat., WO2012099582A1, N. Ibris, S. Brutti and M. A. Navarra, Membranes, 2014, 4,
2012. 123–142.
223 B. S. Pivovar, Y. Wang and E. L. Cussler, J. Membr. Sci., 1999, 247 A. Saccà, I. Gatto, A. Carbone, R. Pedicini and
154, 155–162. E. Passalacqua, J. Power Sources, 2006, 163, 47–51.
224 N. W. DeLuca and Y. A. Elabd, J. Power Sources, 2006, 163, 248 S. Subianto, A. Donnadio, S. Cavaliere, M. Pica, M. Casciola,
386–391. D. J. Jones and J. Roziere, J. Mater. Chem. A, 2014, 2, 13359–
225 Johnson Matthey Fuel Cells Limited, Centre National de la 13365.
Recherche Scientique, Université Montpellier 2, WO Pat., 249 M. Casciola, P. Cojocaru, A. Donnadio, S. Giancola,
WO2016020668A1, 2016. L. Merlo, Y. Nedellec, M. Pica and S. Subianto, J. Power
226 M. H. Yildirim, D. Stamatialis and M. Wessling, J. Membr. Sources, 2014, 262, 407–413.
Sci., 2008, 321, 364–372. 250 G. Alberti and M. Casciola, Membranes for medium
227 T.-E. Kim, S. M. Juon, J. H. Park, Y.-G. Shul and K. Y. Cho, temperature PEFC based on Naon lled with layered metal
Int. J. Hydrogen Energy, 2014, 39, 16474–16485. phosphates and phosphonates, Wiley-VCH Verlag GmbH &
228 T.-C. Jao, G.-B. Jung, S.-C. Kuo, W.-J. Tzeng and A. Su, Int. Co. KGaA, 2008.
J. Hydrogen Energy, 2012, 37, 13623–13630. 251 M. Casciola, D. Capitani, A. Donnadio, V. Frittella, M. Pica
229 H. Tang, M. Pan, F. Wang, P. K. Shen and S. P. Jiang, J. Phys. and M. Sganappa, Fuel Cells, 2009, 9, 381–386.
Chem. B, 2007, 111, 8684–8690. 252 D. Jones and J. Rozière, Advances in the Development of
230 J. Park, L. Wang, S. G. Advani and A. K. Prasad, Inorganic–Organic Membranes for Fuel Cell Applications,
J. Electrochem. Soc., 2012, 159, F864–F870. Springer, Berlin, Heidelberg, 2008.
231 H. L. Tang, M. Pan and F. Wang, J. Appl. Polym. Sci., 2008, 253 M. S. Schaberg, J. E. Abulu, G. M. Haugen, M. A. Emery,
109, 2671–2678. S. J. O'Conner, P. N. Xiong and S. Hamrock, ECS Trans.,
232 X. Zhu, H. Zhang, Y. Liang, Y. Zhang, Q. Luo, C. Bi and B. Yi, 2010, 33, 627–633.
J. Mater. Chem., 2007, 17, 386–397. 254 J. K. Clark Ii and S. J. Paddison, Electrochim. Acta, 2013, 101,
233 T.-C. Jao, G.-B. Jung, S.-T. Ke, P.-H. Chi and S.-H. Chan, Int. 279–292.
J. Energy Res., 2011, 35, 1274–1283. 255 N. J. Economou, A. M. Barnes, A. J. Wheat, M. S. Schaberg,
234 J. Yuan, H. Pu and Z. Yang, J. Polym. Sci., Part A: Polym. S. J. Hamrock and S. K. Buratto, J. Phys. Chem. B, 2015, 119,
Chem., 2009, 47, 2647–2655. 14280–14287.

436 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017
View Article Online

Review Sustainable Energy & Fuels

256 L. Puskar, E. Ritter, U. Schade, M. Yandrasits, 281 P. Trogadas, J. Parrondo and V. Ramani, Chem. Commun.,
S. J. Hamrock, M. Schaberg and E. F. Aziz, PCCP, 2017, 2011, 47, 11549–11551.
19, 626–635. 282 Z. Wang, H. Tang, H. Zhang, M. Lei, R. Chen, P. Xiao and
257 J. A. Leistra, N. E. Cipollini, W. R. Schmidt, J. B. Hertzberg, M. Pan, J. Membr. Sci., 2012, 421–422, 201–210.
C. H. Paik, T. D. Jarvi, T. W. Patterson and S. Tulyani, US 283 L. Wang, S. G. Advani and A. K. Prasad, Electrochim. Acta,
Pat., US20050095355A1, 2005. 2013, 109, 775–780.
258 Ballard Power Systems Inc., US Pat., US20050136308A1, 284 V. Prabhakaran, C. G. Arges and V. Ramani, PNAS, 2011,
2005. 109, 1029–1034.
259 3M Innovative Properties Company, US Pat., 285 V. Prabhakaran, C. G. Arges and V. Ramani, ECS Trans.,
US20070099053A1, 2007. 2011, 41, 1347–1357.
260 3M Innovative Properties Company, US Pat., 286 B. P. Pearman, N. Mohajeri, R. P. Brooker, M. P. Rodgers,
US20080160380A1, 2008. D. K. Slattery, M. D. Hampton, D. A. Cullen and S. Seal, J.
261 GM Global Technology Operations, Inc., DE Pat., Power Sources, 2013, 225, 75–83.
DE102007048872A1, 2008. 287 B. P. Pearman, N. Mohajeri, D. K. Slattery, M. D. Hampton,
262 GM Global Technology Operations, Inc., US Pat., S. Seal and D. A. Cullen, Polym. Degrad. Stab., 2013, 98,
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

US20120122016A1, 2012. 1766–1772.


263 N. R. de Tacconi, C. R. Chenthamarakshan, K. Rajeshwar, 288 H. F. Xu and X. L. Hou, Int. J. Hydrogen Energy, 2007, 32,
W.-Y. Lin, T. F. Carlson, L. Nikiel, W. A. Wampler, 4397–4401.
S. Sambandam and V. Ramani, J. Electrochem. Soc., 2008, 289 A. M. Baker, L. Wang, W. B. Johnson, A. K. Prasad and
155, B1102–B1109. S. G. Advani, J. Phys. Chem. C, 2014, 118, 26796–26802.
264 Y. Patil and K. A. Mauritz, J. Appl. Polym. Sci., 2009, 113, 290 M. Lei, T. Z. Yang, W. J. Wang, K. Huang, Y. C. Zhang,
3269–3278. R. Zhang, R. Z. Jiao, X. L. Fu, H. J. Yang, Y. G. Wang and
265 Y. Patil, S. Sambandam, V. Ramani and K. Mauritz, J. W. H. Tang, J. Power Sources, 2013, 230, 96–100.
Electrochem. Soc., 2009, 156, B1092–B1098. 291 D. Zhao, B. L. Yi, H. M. Zhang and H. M. Yu, J. Membr. Sci.,
266 P. Trogadas and V. Ramani, J. Power Sources, 2007, 174, 2010, 346, 143–151.
159–163. 292 M. T. Taghizadeh and M. Vatanparast, RSC Adv., 2016, 6,
267 G. M. Haugen, F. Meng, N. V. Aieta, J. L. Horan, M.-C. Kuo, 56819–56826.
M. H. Frey, S. J. Hamrock and A. M. Herring, Electrochem. 293 M. T. Taghizadeh and M. Vatanparast, J. Mater. Sci.: Mater.
Solid-State Lett., 2007, 10, B51–B55. Electron., 2017, 28, 778–786.
268 P. R. Brooker, L. J. Bonville and D. K. Slattery, J. Electrochem. 294 Y. Zhu, S. Pei, J. Tang, H. Li, L. Wang, W. Z. Yuan and
Soc., 2013, 160, F75–F80. Y. Zhang, J. Membr. Sci., 2013, 432, 66–72.
269 A. M. Herring, H. M. Gregory, M. Fanqin, N. V. Aieta, 295 F. Meng, N. V. Aieta, S. F. Dec, J. L. Horan, D. Williamson,
J. L. Horan, M. H. Frey, S. J. Hamrock and M. C. Kuo, ECS M. H. Frey, P. Pham, J. A. Turner, M. A. Yandrasits,
Trans., 2006, 3, 551–559. S. J. Hamrock and A. M. Herring, Electrochim. Acta, 2007,
270 J. Rajeswari, Z. Ziegler, G. M. Haugen, S. J. Hamrock and 53, 1372–1378.
A. M. Herring, ECS Trans., 2011, 41, 1561–1565. 296 V. Ramani, H. R. Kunz and J. M. Fenton, J. Membr. Sci.,
271 P. Trogadas and V. Ramani, ECS Trans., 2007, 11, 949–960. 2005, 266, 110–114.
272 G.-Y. Chen, C. Wang, Y.-J. Lei, J. Zhang, Z.-Q. Mao, 297 V. Ramani, H. R. Kunz and J. M. Fenton, Electrochim. Acta,
J.-W. Guo and J.-L. Wang, Int. J. Hydrogen Energy, 2016, 2005, 50, 1181–1187.
41, 16167–16172. 298 Y. Zhou, J. Yang, H. Su, J. Zeng, S. P. Jiang and
273 M. Danilczuk, S. Schlick and F. D. Coms, Macromolecules, W. A. Goddard, J. Am. Chem. Soc., 2014, 136, 4954–4964.
2009, 42, 8943–8949. 299 Y. Yao, J. Liu, W. Liu, M. Zhao, B. Wu, J. Gu and Z. Zou,
274 E. Endoh, Highly durable PFSA membranes, John Wiley & Energy Environ. Sci., 2014, 7, 3362–3370.
Sons, Ltd, Chichester, UK, 2010. 300 Y. Zhu, J. Mai, H. Li, J. Tang, W. Z. Yuan and Y. Zhang,
275 E. Endoh, N. Onoda, Y. Kaneko, Y. Hasegawa, S. Uchiike, Polym. Degrad. Stab., 2014, 107, 106–112.
Y. Takagi and T. Take, ECS Electrochem. Lett., 2013, 2, 301 P. Parthasarathy and A. V. Virkar, J. Power Sources, 2011,
F73–F75. 196, 9204–9212.
276 F. D. Coms, H. Liu and J. E. Owejan, ECS Trans., 2008, 16, 302 S. Chen, H. A. Gasteiger, K. Hayakawa, T. Tada and Y. Shao-
1735–1747. Horn, J. Electrochem. Soc., 2010, 157, A82–A97.
277 T. T. H. Cheng, S. Wessel and S. Knights, J. Electrochem. 303 M. P. Rodgers, N. Mohajeri, L. J. Bonville and D. K. Slattery,
Soc., 2013, 160, F27–F33. J. Electrochem. Soc., 2012, 159, B564–B569.
278 P. Trogadas, J. Parrondo, F. Mijangos and V. Ramani, J. 304 E. Endoh, ECS Trans., 2006, 3, 9–18.
Mater. Chem., 2011, 21, 19381–19388. 305 T. Tanuma and T. Itoh, J. Power Sources, 2016, 305, 17–
279 S. Babu, A. Velez, K. Wozniak, J. Szydlowska and S. Seal, 21.
Chem. Phys. Lett., 2007, 442, 405–408. 306 T. Montini, M. Melchionna, M. Monai and P. Fornasiero,
280 P. Trogadas, J. Parrondo and V. Ramani, Electrochem. Solid- Chem. Rev., 2016, 116, 5987–6041.
State Lett., 2008, 11, B113–B116.

This journal is © The Royal Society of Chemistry 2017 Sustainable Energy Fuels, 2017, 1, 409–438 | 437
View Article Online

Sustainable Energy & Fuels Review

307 C. Korsvik, S. Patil, S. Seal and W. T. Self, Chem. Commun., 318 S. M. Stewart, D. Spernjak, R. Borup, A. Datye and
2007, 1056–1058. F. Garzon, ECS Electrochem. Lett., 2014, 3, F19–F22.
308 Y. Xue, Q. Luan, D. Yang, X. Yao and K. Zhou, J. Phys. Chem. 319 M. Zaton, D. Jones and J. Rozière, ECS Trans., 2014, 61, 15–
C, 2011, 115, 4433–4438. 23.
309 P. Trogadas, J. Parrondo and V. Ramani, ACS Appl. Mater. 320 A. M. Baker, R. Mukundan, D. Spernjak, E. J. Judge,
Interfaces, 2012, 4, 5098–5102. S. G. Advani, A. K. Prasad and R. L. Borup, J. Electrochem.
310 F. Esch, S. Fabris, L. Zhou, T. Montini, C. Africh, Soc., 2016, 163, F1023–F1031.
P. Fornasiero, G. Comelli and R. Rosei, Science, 2005, 309, 321 V. Prabhakaran and V. Ramani, J. Electrochem. Soc., 2014,
752–755. 161, F1–F9.
311 L. Wang, S. G. Advani and A. K. Prasad, ECS Electrochem. 322 M. Zatoń, J. Roziere and D. Jones, J. Mater. Chem. A, 2017, 5,
Lett., 2014, 3, F30–F32. 5390–5401.
312 M. Watanabe, H. Uchida, Y. Seki, M. Emori and 323 L. Pino, A. Vita, M. Cordaro, V. Recupero and M. S. Hegde,
P. Stonehart, J. Electrochem. Soc., 1996, 143, 3847–3852. Appl. Catal., A, 2003, 243, 135–146.
313 M. Lei, T. Z. Yang, W. J. Wang, K. Huang, R. Zhang, X. L. Fu, 324 D. Banham, S. Ye, T. Cheng, S. Knights, S. M. Stewart,
H. J. Yang, Y. G. Wang and W. H. Tang, Int. J. Hydrogen M. Wilson and F. Garzon, J. Electrochem. Soc., 2014, 161,
Published on 05 April 2017. Downloaded on 2/26/2021 7:59:49 AM.

Energy, 2013, 38, 205–211. F1075–F1080.


314 M. Lei, Z. B. Wang, J. S. Li, H. L. Tang, W. J. Liu and 325 D. Banham, S. Ye, T. Cheng, S. Knights, S. M. Stewart and
Y. G. Wang, Sci. Rep., 2014, 4, 7415. F. Garzon, ECS Trans., 2013, 58, 369–380.
315 F. Xu, R. Xu and S. Mu, Electrochim. Acta, 2013, 112, 304–309. 326 N. R. deTacconi, C. R. Chenthamarakshan, K. Rajeshwar,
316 T. Weissbach, T. J. Peckham and S. Holdcro, J. Membr. W.-Y. Lin, T. F. Carlson, L. Nikiel, W. A. Wampler,
Sci., 2016, 498, 94–104. S. Sambandam and V. Ramani, J. Electrochem. Soc., 2008,
317 S. Schlick, M. Danilczuk, A. R. Drews and R. S. Kukreja, J. 155, B1102–B1109.
Phys. Chem. C, 2016, 120, 6885–6890. 327 D. Zhao, B. L. Yi, H. M. Zhang, H. M. Yu, L. Wang, Y. W. Ma
and D. M. Xing, J. Power Sources, 2009, 190, 301–306.

438 | Sustainable Energy Fuels, 2017, 1, 409–438 This journal is © The Royal Society of Chemistry 2017

You might also like