Home Understanding methods of preparation and characterization of pore-filling polymer composites for proton exchange membranes: a beginner’s guide
Article Publicly Available

Understanding methods of preparation and characterization of pore-filling polymer composites for proton exchange membranes: a beginner’s guide

  • Robert Gloukhovski

    Robert Gloukhovski is currently a researcher in the Department of Mechanical Engineering at Technion – Israel Institute of Technology. He received his PhD in Chemical Engineering from Technion under the direction of Assoc. Prof. Yoed Tsur and Assoc. Prof. Viatcheslav Freger in 2017. His graduate research was focused on preparation and characterization of pore-filling polymer composites for proton exchange membranes.

    , Viatcheslav Freger

    Viatcheslav Freger is currently a faculty at the Department of Chemical Engineering of Technion – Israel Institute of Technology in Haifa, Israel. His main research interests are in the areas of membrane technology for water and energy, fundamentals and modeling of membranes, polymer science, electrochemistry, colloidal and surface science and advanced characterization of materials. He has published over 80 papers and book chapters, has given numerous invited talks at international meetings and holds several patents and patent applications in these fields.

    EMAIL logo
    and Yoed Tsur

    Yoed Tsur is currently a faculty at the Department of Chemical Engineering of Technion – Israel Institute of Technology in Haifa, Israel. He is the director of the Grand Technion Energy Program (GTEP). His main research interests are in the areas of point defect chemistry of oxide materials, fuel cells and supercapacitors, and advanced EIS characterization and analysis. He has published over 50 papers and has given numerous invited talks at international meetings.

    EMAIL logo
Published/Copyright: September 13, 2017
Become an author with De Gruyter Brill

Abstract

Composite membranes based on porous support membranes filled with a proton-conducting polymer appear to be a promising approach to develop novel proton exchange membranes (PEMs). It allows optimization of the properties of the filler and the matrix separately, e.g. for maximal conductivity of the former and maximal physical strength of the latter. In addition, the confinement itself can alter the properties of the filling ionomer, e.g. toward higher conductivity and selectivity due to alignment and restricted swelling. This article reviews the literature on PEMs prepared by filling of submicron and nanometric size pores with Nafion and other proton-conductive polymers. PEMs based on alternating perfluorinated and non-perfluorinated polymer systems and incorporation of fillers are briefly discussed too, as they share some structure/transport relationships with the pore-filling PEMs. We also review here the background knowledge on structural and transport properties of Nafion and proton-conducting polymers in general, as well as experimental methods concerned with preparation and characterization of pore-filling membranes. Such information will be useful for preparing next-generation composite membranes, which will allow maximal utilization of beneficial characteristics of polymeric proton conductors and understanding the complicated structure/transport relationships in the pore-filling composite PEMs.

Abbreviations
AAO

Anodized aluminum oxide

BI

Benziimidazole

CLPE

Cross-linked polyethylene

dH2

O Deionized water

DMAC

N,N-dimethyl acetamide

DMF

N,N-dimethyl formamide

DMFC

Direct methanol fuel cell

DMSO

N,N-dimethyl sulfoxide

DVB

Divynyl benzene

ePTFE

Expanded (porous) polytetrafluoroethylene

IPA

Isopropyl alcohol (isopropanol)

MEA

Membrane-electrode assembly

MeOH

Methanol

MCO

Methanol crossover

NMP

N-Methyl-2-pyrrolidone

PAN

Polyacrylonitrrile

PCTE

Polycarbonate track-etched (membrane)

PEM

Proton exchange membrane

PEMFC

PEM fuel cell

PEMWE

PEM water electrolysis

PFSA

Perfluorinated sulfonic acid

PI

Polyimide

PLGA

Poly(lactic-co-glycolic acid)

PSSA

Poly(styrene sulfonic acid)

PSS-b-PMB

Polystyrenesulfonate-block-polymethylbutylene

PTABS

Poly(acrylamide-tert-butylsulfonic acid)

PTFE

Polytetrafluoroethylene

PVA

Polyvinyl alcohol

PVB

Poly(vinyl butyral)

PVDF

Polyvinylidene fluoride

PVDP

Poly(vinylidene fluoride-co-hexafluoropropylene)

RH

Relative humidity

SAXS

Small-angle x-ray scattering

SEM

Scanning electron microscopy

SHSBS

Sulfonated and hydrogenated styrene/butadiene block copolymer

SPAES

Sulfonated poly(arylene ether sulfone)

SPEEK

Sulfonated poly(ether ether ketone)

SPEEKK

Sulfonated poly(ether ether ketone)

SPFCB

Sulfonated perfluorocyclobutane

SPES

Sulfonated poly(ether sulfone)

SPI

Sulfonated polyimide

SPS

Sulfonated polystyrene

SPSU

Disulfonated poly(arylene ether sulfone)

SSC

Short side chain

S-SIBS

Sulfonated poly(styrene-isobutylene-styrene)

UHMWPE

Ultra-high molecular weight polyethylene

1 Introduction: proton exchange membranes in fuel cells and water electrolysis

The world-wide concern for environment protection requires the replacement of fossil fuel combustion with alternative energy sources. Fuel cells (FC) are among the most promising technologies for potential replacement, due to their high efficiency (about 60%), low air and noise pollution and low maintenance costs. FCs convert energy produced in a redox reaction to electricity directly by separating the half-reactions via a selective medium. This medium conducts only one charge carrier (either positive or negative) and is virtually impermeable to the reactants. The circuit is closed through an external load that, depending on the FC design and size, can vary from battery chargers to bus motors.

Hydrogen oxidation by air is the most common reaction in FCs, although fuels such as methanol and formic acid are also used. Depending on the selective medium material, FCs are categorized into solid oxide (SOFC), alkaline (AFC), molten carbonate (MCFC), phosphoric acid (PAFC) and proton exchange membrane fuel cells (PEMFC) – among which the latter category is currently the dominating one (The Fuel Cell Industry Review 2015).

In PEMFCs, as shown in Figure 1, hydrogen is fed to the anode compartment and air is fed to the cathode compartment (Wycisk et al. 2014). Each gas diffuses through a gas diffusion layer (GDL) toward a PEM (known also as a polymer electrolyte membrane). It is coated on each side by a catalyst (mainly Pt) mixed with carbon nanoparticles. The GDLs, the membrane and the catalyst layers are sandwiched together by hot-pressing to form a membrane-electrode assembly (MEA). Hydrogen is oxidized by the catalyst into protons and electrons. The protons diffuse toward the cathode through the PEM, while the electrons travel through the polar plates enclosing the FC body and are transferred to the external load, producing work and heat. The reactants are usually prehumidified to increase RH within the cell. Thus, water goes out from the cathode with the remaining oxygen and excess nitrogen. Humidified hydrogen also comes from the anode. Eventually, protons, electrons and oxygen from the air recombine together on the cathode catalyst layer, producing water, which is the only byproduct (Mekhilef et al. 2012). This process is exploited in reverse in PEM water electrolysis (PEMWE) for high-purity hydrogen production. Ultrapure water is decomposed in PEMWE by external voltage, and protons travel through the PEM to recombine with electrons and form H2 (Zoulias et al. 2004). PEMWE yields higher current densities (up to 2 A/cm2), compared to the more traditional alkaline water electrolysis, thus allowing more compact and versatile electrolyzers (Carmo et al. 2013).

Figure 1: A schematic representation of a hydrogen-fed proton exchange membrane fuel cell (PEMFC).
Figure 1:

A schematic representation of a hydrogen-fed proton exchange membrane fuel cell (PEMFC).

The theoretical maximum voltage produced by the hydrogen-oxygen reversible reaction is 1.23 V. However, in practice, the voltage produced by FC decreases with the current density; this dependence is described by the generalized polarization curve shown in Figure 2.

Figure 2: A sample polarization curve of a single proton exchange membrane fuel cell (PEMFC).
Figure 2:

A sample polarization curve of a single proton exchange membrane fuel cell (PEMFC).

There are several factors causing the drop from the maximum voltage, also called overvoltage or overpotential (Jiao and Li 2011).

  1. When no current density flows in or out of FC (i=0), the current densities of the forward and the backward reactions in the 2H2+O2↔2H2O system are equal. This current density is called the exchange current density (i0), and the overvoltage required to overcome it (i>i0) causes the open circuit voltage (OCV) to be <1.2 V. An additional loss at this stage occurs because the PEM is not ideal and allows some hydrogen and electron crossover (this is referred to as the mixed potential).

  2. At current densities just above the i0, the reaction rate on the MEA Pt electrodes is the limiting step, causing kinetic (or activation) loss.

  3. At current densities between 100 mA/cm2 and 800 mA/cm2, there is a linear fall in voltage (“Ohmic losses”) as a result of resistance to the proton and electron transfer. High PEM conductivity is crucial for minimizing the Ohmic losses.

  4. At very high current densities (>800 mA/cm2), the H2 and O2 consumption and water production rates are high, creating concentration gradients across the GDL. This mass transfer (or concentration) loss drives the voltage rapidly to zero.

As it can be concluded from the polarization curve in Figure 2, the materials and working conditions of an ideal PEMFC should provide it with the following principal characteristics: high protonic and low electronic conductivity of PEM, low reactant crossover, fast oxidation-reduction kinetics and reactant mass transfer. In addition, physical and chemical PEM stability at elevated temperatures is required. However, imperfections of real-life devices, e.g. kinetics and mass-transfer limitations, membrane Ohmic resistance and reactant crossover, result in a series of overpotentials, lowering the output voltage in FCs and increasing the input voltage in WEs. In direct methanol fuel cells (DMFCs), where a water-methanol mixture is fed on the anode side, the methanol crossover (MCO) through the membrane is a serious problem due to methanol having good miscibility with water, thus requiring especially thick membranes (Neburchilov et al. 2007).

Per-fluorinated sulfonic acid polymers (PFSAs) remain today the standard materials for PEMs. Nafion®, the first PFSA, was commercialized by Du Pont in the late 1960s. Dow membrane from Dow Chemical, Flemion® membrane from Asashi Glass Corporation and Aciplex® membrane from Asahi Chemicals are also known today (Tian 2004). Aromatic sulfonated membranes manufactured by Ballard and Dais Analytics, the only known commercial alternatives to PFSAs, are of limited use (Hickner et al. 2004).

Nafion remains the most widely used material for PEM due to its high proton conductivity and good mechanical and chemical stability (Dupuis 2011). This high proton conductivity (about 10 S m−1 at room temperature) originates from a random 3D network of channels formed by the sulfonic groups and water molecules (Elliott et al. 2011). Environment of high relative humidity (RH) is required to keep high water content in the membrane, λ, up to 22 water molecules per sulfonic group (Kraytsberg and Ein-Eli 2014). This requirement for high hydration, as well as deterioration of Nafion mechanical stability at elevated temperatures, poses several limitations on devices operated with Nafion PEMs (Devanathan 2008).

In non-pressurized hydrogen-fed PEMFCs, the operational temperature range is limited to 60–80°C in order to prevent water loss due to decreasing RH (Alberti et al. 2008). Increasing the gas pressure to 3 bar would raise the water boiling point to 135°C; however, this poses several disadvantages, e.g. water management problems and membrane damage due to fuel impurities (Tian 2004, Bose et al. 2011). At the same time, operating non-pressurized PEMFCs at temperatures above 100°C is highly desirable for increased efficiency as well as lower manufacturing and operating costs. For example, a significant amount of heat needs to be removed from a PEMFC operating at 80°C with an efficiency of 40–50%. Increasing the operational temperature would simplify the cooling systems, e.g. reduce the front area of radiators, which is important for automobile applications. Conditions of low RH would make the water management easier. Also, absence of liquid water will increase the exposed surface area, improving interlayer reactant diffusion in MEA. Higher operational temperatures should yield higher oxygen reaction rates, so the expensive Pt catalyst loading on the cathode side can be reduced. It may even be feasible to use nonprecious metal catalysts at temperatures above 120°C. It would also significantly mitigate the problem of catalyst poisoning by carbon monoxide (CO), which otherwise requires usage of expensive high-purity hydrogen (Steele and Heinzel 2001). With the catalyst poisoning problem solved, relatively cheap hydrogen produced by steam-methane reforming (and thus containing residual CO) can be used as fuel in PEMFCs. For example, the allowed CO concentration in PEMFC will increase from 20 ppm to 1000 ppm as the operating temperature increases from 80°C to 130°C (Devanathan 2008).

The same concerns of kinetics and catalyst poisoning apply to DMFCs, where CO is formed during methanol decomposition. This requires high Pt loading on the anode side, too. As DMFCs are fed with aqueous methanol solution, the PEM remains hydrated; however, the need for thick Nafion membranes to reduce the methanol crossover causes higher Ohmic losses and larger manufacturing costs (Hickner et al. 2004). Also, methanol permeability increases with temperature, requiring PEMs with increased selectivity (i.e. high ratio of proton conductivity to methanol permeability) for elevated operational temperatures (DeLuca and Elabd 2006).

In a PEMWE, operating at 50–80°C, the conductivity loss is also not critical, as the membrane remains fully hydrated (Smith et al. 2013). However, due to the high current densities, the membrane resistance still contributes up to 50% of the PEMWE overvoltage. In addition, the overvoltage required for water splitting decreases with temperature (Carmo et al. 2013). Thus, increasing PEM conductivity and operational temperature can significantly reduce power consumption in PEMWEs (Choi et al. 2004, Ayers et al. 2010).

These considerations, together with the high cost of Nafion membrane ($700–1500/m2, depending on thickness), stimulated a vast amount of research aiming to develop a cost-effective PEMs (Peighambardoust et al. 2010). The target set by the US Department of Energy (DoE) for PEMFC commercialization in automobile applications is conductivity of 10 S/m at 120°C and RH 50% (Kim et al. 2015). The numerous existing approaches may be categorized into the following groups (Alberti and Casciola 2003, Smitha et al. 2005, DeLuca and Elabd 2006, Zhang and Shen 2012a, Díaz et al. 2014):

  1. Modification of existing PFSAs, e.g. by altering the content and location of sulfonic groups or by chemical crosslinking to increase the protonic conductivity and mechanical and chemical stability.

  2. Synthesis of cheaper sulfonated aromatic membranes with different architectures to improve selectivity to methanol. These materials can be blended with Nafion or other polymers, too, for optimized performance.

  3. Development of PEMs based on solid acids or suitable oxygenated acids (e.g. phosphoric or sulfuric acids) immobilized in a polymeric matrix bearing basic groups (e.g. polybenzimidazole) for high conductivity under anhydrous conditions.

  4. Incorporation of hygroscopic inorganic fillers (e.g. silica, zirconia, etc.) and proton-bearing fillers (e.g. heteropolyacids) into Nafion to increase water retention and available proton concentration at elevated temperatures.

  5. Development of PEMs based on ionic liquids, which have the potential for use at elevated temperatures above 100°C.

  6. Filling non-conductive porous matrixes of high strength and stability with ionomers or inorganic particles of high-proton conductivity.

The focus of this review is concerned with the last approach, i.e. with filling porous matrix membranes and nanofiber mats with proton-conducting polymers. We also review here the background knowledge on structural and transport properties of Nafion and proton-conducting polymers in general, as well as experimental methods concerning membrane preparation and characterization. We believe that this background is essential in order to comprehend the complicated structure/transport relationships within the pore-filling composite PEMs.

2 Nafion structure

The unique properties of Nafion come from its chemical and morphological structure. The Nafion molecule consists of a perfluorinated backbone with side chains terminated by –SO3H groups (Figure 3). The value of n is related to the equivalent weight, EW, i.e. mass of polymer in grams per 1 mole of –SO3H groups. EW is coded into the names of commercial Nafion membranes; for example, N117 means EW 1100, thickness is 7 mil (≈150 μm). EW 1100 (n=6.5–6.6) is the most widely used in commercial membranes. The value of x depends on Nafion molecular weight, which is estimated to be in the range of 105–106 Da (Mauritz and Moore 2004). The perfluorinated backbone provides Nafion with chemical stability, while the flexible side chains allow formation of an effective proton conduction network. Due to the electronegativity of the perfluorinated backbone, Nafion is a superacid with pKa≈−6, so the protons from the sulfonic groups are fully dissociated, if sufficient amount of water is available. As a result, Nafion proton conductivity increases with water content, reaching a room temperature maximum of about 10 S m−1 when fully hydrated (Dupuis 2011).

Figure 3: Chemical formula of Nafion.
Figure 3:

Chemical formula of Nafion.

The understanding of the Nafion structure can be dated back to 1970, when Eisenberg (1970) developed a theory, according to which clusters of ionic groups in ionomers are dispersed within a matrix of low ion content. This nano-phase separation results from incompatibility of the ionic groups with the organic matrix, whose dielectric constant is low. It represents the balance between the electrostatic forces of the ionic groups and the elastic forces of the stretched polymer chains. Later, Gierke (1977) and his co-workers developed a phenomenological cluster network model, according to which the ionic domains in water-swollen Nafion are considered spherical inverted micelles connected by short narrow channels with diameters of 4 and 1 nm, respectively (Gierke et al. 1981, Hsu and Gierke 1983).

While the original Gierke model is still the most popular, small-angle X-ray scattering (SAXS) studies clearly demonstrate that ionic clusters have elongated rather than spherical structure (Devanathan 2008). Nafion membrane’s structure and its evolution during the water uptake are continuously being discussed and re-examined (Mauritz and Moore 2004, Schmidt-Rohr and Chen 2008, Elliott et al. 2011, Kreuer and Portale 2013). However, there seems to be a general agreement about a few key features:

  1. The hydrophobic region is not uniform and is divided into amorphous and crystalline domains. The latter provide the membrane with physical “cross-linking”, responsible for mechanical strength, resistance to solvents and low gas permeability. The crystallinity in Nafion of EW 1100 is estimated in the range of 5–20%. Temperature well above the Nafion glass transition temperature, Tg=105°C, must be applied during the membrane preparation to allow crystalline domain formation (an annealing procedure).

  2. Ionic clusters are elongated rod-like or lamella-like aggregates with a hydrophilic core and hydrophobic shell, and they are packed in bundles with a short-range order (a few hundreds of nanometers).

  3. When exposed to water (in either liquid or gas phase), these ionic clusters, which are initially small and isolated in dry Nafion, increase in size and coalesce upon water uptake, resulting in percolation at hydrophilic phase fraction, ϕh≈0.1 (Kirkpatrick 1973, Hsu et al. 1980). This percolation threshold value is lower than found for randomly dispersed spherical clusters (around 0.3), indicating the elongated geometry of the ionic clusters in Nafion (Kirkpatrick 1973, Wu et al. 2011). The situation when very few isolated clusters remain and a cluster-channel network is formed is achieved at ϕh of about 0.2 (Gebel 2000). For Nafion of EW 1100, this corresponds to about λ=4, i.e. completion of the first hydration layer of the sulfonic groups (Zhao et al. 2011).

  4. When boiled in water at 100°C, Nafion expands to a state with ϕh≈0.4, which corresponds to λ≈19–22 H2O molecules per –SO3H group. Upon water uptake, the hydrophilic core increases in size, forming a network of water channels with diameter of about 5 nm, which can vary along the length of a given channel. This network allows fast transport of protons and polar molecules, e.g. water and methanol, across the Nafion membrane.

  5. At very high hydrations ϕh>0.5, which do not occur under normal conditions and require boiling the Nafion membrane in a pressure vessel at T>120°C, inversion of the structure occurs and the membranes assume the structure of a connected network of rod-like polymer aggregates. The proton conduction network expands to become a dilute continuous aqueous phase, and with further increase in ϕh, the specific conductivity decreases due to dilution of the sulfonic groups and protons.

3 Transport in Nafion

Proton conduction, as well as transport of other water-soluble species (e.g. salts and alcohols) in Nafion and other ionomers, occurs almost exclusively through the water channels. Therefore, the transport properties of a PEM strongly depend on its hydration (Steele and Heinzel 2001); however, other factors like charge density, tortuosity and ordering are also important (Kirkpatrick 1973, Elabd et al. 2003). This section briefly describes the transport mechanisms in Nafion.

3.1 Mechanisms of proton conduction in Nafion

Proton conduction in Nafion occurs through the channels formed by sulfonic groups and filled with water. Therefore, λ has a crucial effect on the proton conductivity, σ. For example, in Nafion 112 at 80°C, it decreases from 9 to 2 S m−1 when RH is lowered from 90% to 50% (c). The effect of λ on the conductivity originates from three mechanisms that are responsible for proton transfer in Nafion (Kraytsberg and Ein-Eli 2014), as explained below.

At low hydration (λ<5), the state of water in Nafion is defined as “non-freezable” (“bound water”, i.e. tightly bound by the electrostatic interactions to the sulfonic groups forming the walls of ionic channels) (Devanathan 2008, Wu 2008). The bound water region exists within 3–4Å near the pore surface, which is roughly the size of a water molecule (Sahu et al. 2009). Only slow hopping of protons between the sulfonic groups occurs in this regime, referred to as “surface diffusion mechanism.” As hydration increases, the ionic channels expand, reaching an average diameter of 4–6nm. The water accumulating in the central part of the ionic channels is defined as “freezable” (“bulk water”). Protons travel through the bulk water the same way as through aqueous acid solutions, i.e. by “en masse” (“vehicle”) diffusion of H3O+ ions and by “structural diffusion” (“Grotthuss mechanism”) (Choi 2004, Choi et al. 2005). The structural diffusion means that a proton passes from a H3O+ ion to a neighboring H2O molecule through a series of spatial rotations involving breaking of hydrogen and covalent bonds of a hydrogen atom with adjacent water oxygen. At full hydration, the contributions to overall conductivity are roughly 80% Grotthuss and 20% vehicle, while the surface hopping contribution is negligible. As the total number of protons is constant and equal to the number of sulfonic groups, the increase in conductivity with hydration occurs through the increase of proton fraction found in the bulk water. In addition, the width and tortuosity of the proton path through the 3D network decrease as the ion channels expand, thus the resistance to the proton transport decreases.

The conductivity contributions of these three types of protons and their diffusion in the water channels to the conductivity can be described by the modified Nernst-Einstein equation (Choi et al. 2005):

(1)σH+=ϕhτF2RT(DH+SCH+S+DH+GCH+B+DH+ECH+B)

where F, R and T are the Faraday constant, the universal gas constant and the absolute temperature; DH+ and CH+ are the proton diffusion coefficient and concentration in the hydrophilic phase; ϕh and τ are the water volume fraction and the tortuosity factor, both dependent on λ; and the indices B, S, G and E stand for “bulk”, “surface”, “Grotthuss” and “en masse”, respectively.

Recent research has revealed that the Grotthuss mechanism can be subdivided into two mechanisms. One is the structural diffusion described above, for which water fluctuation is indispensable. The second is a “packed-acid mechanism,” which occurs in materials containing highly concentrated (packed) acids (Ogawa et al. 2014). In contrast to structural diffusion, the packed-acid mechanism can lead to conduction of protons without water movement or fluctuation. This type of conductivity is used in solid-acid PEMs, which demonstrate high conductivity under anhydrous conditions (Goni-Urtiaga et al. 2012).

3.2 Methanol transport in Nafion

In general, the transport of small molecules across a dense (nonporous) polymer membrane follows a solution-diffusion mechanism, i.e. sorption of solutes into the membrane (on the high concentration side), diffusion across the membrane driven by electrochemical potential gradient and desorption of solutes out of the membrane (on the low concentration side) (DeLuca 2004). The resulting profile is represented schematically in Figure 4.

Figure 4: A schematic description of a solution-diffusion mechanism in a polymer membrane separating two volumes of solution with high and low solute concentrations.
Figure 4:

A schematic description of a solution-diffusion mechanism in a polymer membrane separating two volumes of solution with high and low solute concentrations.

The transport rate of the solute in the membrane J [mol m−2 s−1] under concentration gradient and without additional external fields is defined by the following expression:

(2)J=DΔCmd,

where D [m2 s−1] is the diffusion coefficient, ΔCm =Cm1Cm2 is the concentration drop in the membrane [mol m−3], and d is the membrane thickness [m]. As the solute concentrations in the membrane surface layers Cm1, Cm2 are not easily measurable, the transport rate is usually expressed through the specific permeability P [m2 s−1]:

(3)J=PΔCsd,
(4)P=KD

where ΔCs=Cs1Cs2 is the concentration drop between the high concentration and the low concentration solutions, and the partitioning coefficient K=Cm /Cs is the ratio of solute concentrations in the membrane and in the solution at the membrane-solution interface.

When phase-separated ionomers, such as Nafion, and water-miscible solutes, e.g. methanol (MeOH) or ethanol, are considered, K depends on their solubility in the hydrophilic and hydrophobic phases of the ionomer. As was shown by Zhao et al. (2012), sorption of pure water, methanol and ethanol in Nafion occurs essentially by the same mechanism, i.e. solvation of the sulfonic groups, so the contribution of the hydrophobic phase is minor. The first solvation shell consists of four strongly bound molecules; the second solvation shell and beyond are less strongly bound. However, with water-alcohol solutions, the first solvation shell, i.e. the bound water, is formed exclusively by water molecules due to their higher affinity to the sulfonic groups. As a result, alcohol molecules are excluded from the bound water region. The effect is known as “salting out” by fixed ionic groups and obeys the Setschenow equation (Freger et al. 1997a):

(5)Kexp(βCAm),

where β is the salting-out coefficient and CAm is the concentration of the fixed anion groups in the swollen ionomer membrane.

In the simplest case, methanol permeation through Nafion can be viewed as diffusion through the bulk water region, where MeOH concentration and diffusion coefficient are essentially the same as in the external MeOH solution (Lin et al. 2008a,b). The overall permeability, therefore, is controlled by the hydrophilic volume fraction ϕh, or more specifically, by the fraction of bulk water in the membrane, as K increases with ϕh (i.e. decreasing CAm) (Wu 2008). As the proton and the methanol transport in Nafion and other sulfonated ionomer systems follow essentially the same routes, σ and PMeOH are coupled and change simultaneously with ϕh, which makes improving selectivity difficult (DeLuca and Elabd 2006). ϕh, however, is not the only parameter determining permeability. The morphology and microstructure of the hydrophilic phase may also be very important (Freger et al. 1997b, Freger et al. 1999). For example, Schmidt-Rohr and Chen (2008) concluded that the protonic channels in Nafion, in comparison to stiffer sulfonated aromatic polymers, are wider and less branched due to flexible side chains. This results in higher water diffusion coefficients and higher methanol permeation in Nafion. MeOH concentration can influence the permeability to some extent, too, e.g. an increase in PMeOH from 1.5×10−10 to 2.5×10−10m2 s−1 was reported for a Nafion N117 membrane when the methanol concentration increased from 1 m to 10 m (Yildirim et al. 2008).

Methanol crossover poses a serious problem to the operation of DMFCs. While in hydrogen-operated PEMFCs, thin Nafion membranes (N112) are used to minimize Ohmic losses, in DMFCs, the losses due to methanol crossover are even more significant than the Ohmic ones, requiring thicker membranes (N117) (Hickner et al. 2004). Still, even using thicker membranes, MeOH concentrations in the DMFCs have to be low (1–2 m; 4–8vol.%) to mitigate the crossover problem. If methanol crossover in DMFCs was not an issue, a higher methanol concentration could be used, which would yield a significantly higher cell voltage. Current DMFCs have an open circuit voltage of 0.7V, which is approximately half of the reversible “no-loss” cell voltage (1.2V). However, a 66% reduction in voltage at an optimal operating current density and a decrease in open circuit voltage were observed when the MeOH concentration in the feed was increased from 2 to 6 m (DeLuca and Elabd 2006).

3.3 Salt transport in Nafion

Nafion membranes were originally developed as ion-selective barriers in chlor-alkali applications (Schlick 1996). Nafion selectivity, i.e. exclusion of free anions, results from the effect of fixed anionic groups, which is explained as follows. When uncharged polymer film is immersed in a monovalent salt solution (M+X), salt sorption occurs by a simple partitioning mechanism, whereby for each equivalent of uptaken cations, an equivalent of anions has to be uptaken to maintain electro-neutrality within the film. In a polymer with fixed anions A, however, salt sorption follows both a simple partitioning and an ion-exchange mechanism, whereby cations M+ act as counter ions for both the free (X) and the fixed (A) anions. As electro-neutrality must be maintained, the resulting concentration of X in the membrane is lower than that of M+. This phenomenon whereby the difference in free anion and cation concentrations in the polymer and in the solution is balanced electrostatically by the polymer’s fixed charge groups is referred to as Donnan exclusion (Geise 2012).

For a phase-separated ionomer like Nafion, salt is almost exclusively limited to the hydrophilic phase; the partition coefficient for the free salt in the membrane can be given by the following expression (Geise 2012):

(6)KCXmCXs=ϕh{[(CAh2CMXs)2+(γ±sγ±h)2]0.5CAh2CMXs},

where the superscript indices s, m and h represent the solution, the membrane and the hydrophilic phase of the membrane, respectively. The γ± terms are the mean activities of the ions, and their ratio K0γ±s/γ±h represents all the non-Donnan contributions to the partitioning (Bason et al. 2010). At very low salt concentrations CMXsCAh2K0, the partition coefficient is nearly proportional to the inverse of the fixed anion concentration: KϕhK02CMXsCAh. At very high salt concentrations CMXsCAh/2K0, the Donnan contribution becomes negligible and the partition coefficient is determined only by the non-Donnan contributions: KϕhK0.

Prediction of K is, however, complicated. Even in the simplest case of Nafion membrane immersed in NaCl solution, the presence of salt affects K in several non-linear ways. First, the exchange of protons for sodium ions lowers λ and thus ϕh due to the lower hydration number of Na+. In addition, λ further decreases with increasing CNaCls (so-called osmotic de-swelling), resulting in higher CAh, narrower ion channels, increased tortuosity and stiffening of the polymer chains (Narebska et al. 1984, Cwirco and Carbonell 1992a,b, Lehmani et al. 1997). The dielectric constant in the membrane will change, too, affecting K0.

When considering a more complicated system, e.g. Nafion membrane separating two compartments, one filled with concentrated NaCl solution and the other, with pure water, λ should increase across the membrane from the solution side to the pure water side, consequently affecting K. From a practical point of view, however, the permeability of NaCl in Nafion and other sulfonated polymers under such conditions follows a simple empirical log-log linear correlation with the bulk NaCl concentration (Geise et al. 2014). For a Nafion N115 membrane, for example, PNaCl=5×10−12CNaCl0.427 when a membrane is pre-treated by soaking in water at room temperature and PNaCl=2×10−11CNaCl0.452 when a membrane is pre-treated by boiling in water (the units of PNaCl and CNaCl are m2 s−1 and mol l−1, respectively). PNaCl in both cases increases with CNaCl concentration due to the suppression of Donnan exclusion, and the higher PNaCl of a boiled membrane results from its higher λ.

4 Characterization of PEM transport properties

4.1 In situ PEMFC characterization by electrochemical impedance spectroscopy

PEM transport properties can be studied in situ and ex situ, i.e. as a MEA performance in a PEMFC or as isolated membrane properties. In the former case, the contributions of the catalyst kinetics, membrane resistance and mass transfer to the overall overpotential are studied at different current densities. The MEA permeability to the reactants can be assessed as a half-cell limiting current density under external potential, e.g. when the anode side is fed with H2 (Tang et al. 2007a) or MeOH (Choi et al. 2008) solution and the cathode side is fed with N2. The contributions mentioned above can be separated and quantified by means of electrochemical impedance spectroscopy (EIS) (Scribner Associates – Tutorial). In this technique, a small AC voltage or current bias is applied to the FC over a broad range of frequencies (f) to acquire the system impedance response Z (spectrum). The response at the high end of the frequency range is generally associated with the dielectric effects and resistance to charge transfer in the bulk electrolyte, while the electrode effects (reaction, adsorption, double layer response) are typically observed at very low frequencies (Macdonald 1992).

The system impedance can be analyzed by building an equivalent circuit (EC) out of several elements connected in parallel and/or in series, so the frequency response of the EC fits as close as possible to the system response (Latham 2004). The advantage of the EC approach is being relatively fast and simple; however, one of the problems is that ECs are not unique. Several different ECs can result in the same spectrum, leading to completely different physical interpretations (Baltianski and Tsur 2003). Thus, other modern analysis techniques, notably Impedance Spectroscopy Genetic Programming (ISGP) (Hershkovitz et al. 2011, Hershkovitz et al. 2012, Drach et al. 2016, Oz et al. 2016, Oz et al. 2017), analyzing distribution of relaxation times in the system, and mechanistic models (Amphlett et al. 1995, Springer et al. 1996, Franco et al. 2007, Reshetenko and Kulikovsky 2017), are being also employed.

4.2 Ex situ techniques for measurement of PEM transport properties

While an in situ membrane testing provides the most adequate information on the membrane’s performance, this method requires considerable time and skill, as well as expensive specialized equipment (Cooper 2009). Moreover, the hot-pressing procedure during the MEA preparation can alter the PEM properties. As a result, PEM transport characteristics are often measured ex situ in electrochemical and/or diffusion cells. For example, H2 permeability was measured when a membrane was placed between two vessels, one filled with pressurized H2 and the other one initially kept under vacuum (Lin et al. 2004). In the setup employed, the pressure change in the second vessel was monitored over time (nearly 60 h). In a similar way, methanol permeability is commonly measured in diffusion cells, in which the membrane separates two stirred compartments. One compartment is filled with MeOH, and the other, with deionized water (dH2O) (Lin et al. 2005, Yildirim et al. 2008, Zhang et al. 2009, Lin et al. 2010, Gloukhovski et al. 2016). The change in MeOH concentration in the second compartment is measured with gas chromatograph (GC) or density meter using a suitable calibration curve.

The literature suggests that, while the permeability measurements are relatively standardized and straightforward, ex situ conductivity measurements appear to be more controversial. The measurement of the through-plane membrane conductivity under controlled temperature and RH conditions, a mode that best represents the actual mode of operation in PEMFC, requires long equilibration (approximately 2 h) prior to testing (Cooper 2010, Cooper 2011). In addition, poor membrane-electrode contact due to the surface roughness of the experimental membranes can impair the measurements (Cho et al. 2004). Consequently, numerous methods for conductivity measurement in liquid at ambient conditions were proposed, resulting in a wide scatter of reported conductivity values, often significantly different from 10 S m−1 (Slade et al. 2002, Silva et al. 2004). As a result, comparing conductivity and, correspondingly, selectivity values of different PEMs can be problematic due to limitations and uncertainties of the measurement techniques. Gloukhovski et al. (2016) presented a detailed discussion of the subject, which may be summarized as follows.

The first very common approach is to measure in-plane conductivity of a PEM, when the resistance of a long wet membrane stripe is measured in the tangential direction by ac impedance (typically below 100 kHz) with two or four electrodes placed along the stripe. This method is simple and reliable, as the contact and the electrode resistances are negligible relative to the membrane resistance and results in σ≈6–9 S m−1 for Nafion, depending on λ. However, as the actual “working” direction in a PEM is through-plane (normal to the surface), this method is adequate only for isotropic membranes. An anisotropic experimental PEM where some alignment is expected in either in-plane or through-plane direction requires measuring conductivity in the through-plane direction.

In an often used second approach, the through-plane resistance is deduced from ac impedance measurements when the membrane is equilibrated with dH2O. High frequencies up to 10 MHz are typically required to eliminate effects other than the bulk electrical conductivity. Yet the resistance of the thin water layer in the region of membrane-electrode contact still makes a significant relative contribution to the overall resistance, especially with thin membranes. For Nafion membranes, the conductivity measured by this method was reported to be σ≈1–3 S m−1, typically with lower values for thinner membranes and higher ones for thicker films, strongly suggesting a significant contribution of contact and solution film resistances.

The contact resistance can be eliminated by equilibrating the membrane with strong acids of concentration above 1 m rather than with dH2O (Gloukhovski et al. 2016). Again, the proton conduction in pore-filling composite PEMs discussed later can result from Nafion presumably deposited in the pores (specific transport) as well as from free transport of acid ions through the unblocked pores (non-specific transport). The distinction can be made with a complementary measurement of PNaCl, as Cl transport will be much slower in the Nafion-filled pores than in the empty pores due to Donnan exclusion. The advantage of this method is that it requires only a simple real-time monitoring of cell conductivity or impedance, and thus, it is faster and simpler compared to PH2 and PMeOH measurements. Thereby, it can be used in combination with σ measurements for the initial screening of numerous experimental membranes, before more technically demanding characterization techniques are applied (Gloukhovski et al. 2017).

5 Structure/transport relationships in Nafion

The major drawback of Nafion membranes is loss of conductivity in vapor at elevated temperatures. In general, conductivity of Nafion in vapor rises with temperature up to ≈70°C and then starts to deteriorate. This behavior stems from a combination of two effects varying oppositely with the temperature: increasing water mobility and decreasing λ. Park et al. (2007) demonstrated that, in contrast with Nafion, in polystyrenesulfonate-block-polymethylbutylene (PSS-b-PMB), both conductivity and λ increase continuously up to 90°C. The suggested reason for decreasing λ in Nafion was the instability of ionic channels and the deterioration of their 3D network as the chain mobility in Nafion, which is a random copolymer, increases with temperature. The more structured block-copolymer PSS-b-PMB, meanwhile, demonstrated higher stability of the ionic channels and thus remained hydrated and conductive at elevated temperatures.

On the other hand, Alberti et al. (2008, 2013), Alberti and Narducci (2009) demonstrated that at temperatures ≥120°C and RH close to 100% (either in vapor or liquid), Nafion membranes suffer from irreversible excessive swelling (λ>35) leading to loss of tensile strength. Additionally, under these conditions, the expansion of the membrane clamped between the electrodes is restricted in the direction of thickness, resulting in anisotropic swelling. This realigns the proton-conducting channels in parallel to the surface, lowering the through-plane conductivity. These structural properties of Nafion lead to a “deadlock” whereby a Nafion membrane at temperatures of 120°C and above inevitably suffers from conductivity loss either through dehydration (at low RH) or through excessive hydration (at high RH).

Alternatively, deliberate alignment of Nafion bundles resulting in enhanced conductivity was demonstrated by several methods as an increase in SAXS “ionomer peak” intensity in the direction normal to the alignment (Gierke et al. 1981, Hsu and Gierke 1983, Truffier-Boutry et al. 2007). The bundles can be aligned in parallel-to-draw direction by uniaxial drawing of the Nafion film (van der Heijden et al. 2004). Proton conductivity, which is isotropic under no-draw conditions, increased by 30% in the draw direction and decreased by 50% in the two normal-to-draw directions under 4:1 elongation ratio (Cable et al. 1995, Park et al. 2011b).

It was shown by means of grazing incidence small-angle X-ray scattering (GISAXS) and small-angle neutron scattering (SANS) that alignment of Nafion micelles in water-alcohol solutions may be induced near external surfaces (Bass et al. 2010) or in parallel to the interface with a solid material upon film deposition (Dura et al. 2009, Bass et al. 2011). The alignment is most significant in the layer a few tens of nanometers from the surface, with more distant layers increasingly disordered. Parallel-to-axis alignment of ionic aggregates was demonstrated in Nafion nanofibers, where the longitudinal conductivity remarkably increased from 20 S m−1 in fibers with diameter of 2.5 μm to 150 S m−1 in fibers with diameter of 0.4 μm (Dong et al. 2010). The increase in conductivity was especially sharp for the fibers with diameter <1 μm, supporting the previous conclusion that the surface-induced alignment propagates to a limited distance of the order of tens to a few hundred nanometers into the fiber. Therefore, the desirable properties of Nafion membranes may be enhanced through proper manipulation of its structure.

6 Alternative polymer and composite systems for PEMs

A vast amount of research is dedicated to the development of more stable, conductive and selective PEMs. In this review, we are concerned primarily with PEMs prepared by filling micro- and nanoporous support membranes with Nafion and other proton-conductive polymers. PEMs based on alternative perfluorinated and non-perfluorinated polymer systems and incorporation of fillers will be briefly discussed first, as they share in common some structure/transport relationships with pore-filling PEMs. Other approaches, e.g. polymer blending and crosslinking (DeLuca and Elabd 2006, Subianto et al. 2013), anhydrous PEMs (Neburchilov et al. 2007, Dupuis 2011, Goni-Urtiaga et al. 2012, Kraytsberg and Ein-Eli 2014, Wycisk et al. 2014) and porous PEMs filled with inorganic ion-conductive particles (Alberti and Casciola 2003), will not be discussed here. Proton conductivity, permeability of some other species i (e.g. H2, MeOH) and their ratio (selectivity) will be referred to below as σ, Pi and ηi=σ/Pi, respectively.

6.1 Perfluorinated and partially fluorinated PEMs

Short side-chain (SSC) PFSA membranes seem to be a promising strategy in search for PEMs with higher conductivity, as shorter side chains allow for a reduction in the EW of the membrane and an increase in the density of acid groups. On the other hand, they are known for problems of durability and mechanical integrity due to increased swelling. Dow Chemical Company demonstrated an SSC PEM with EW of 800–900, which achieved four times the power of Nafion membrane (EW 1100) at the same operating voltage (Savadogo 1998) due to higher conductivity provided by higher density of sulfonic groups. However, the Dow membrane advantage in conductivity over Nafion diminishes with decreasing hydration, and at λ<5, the conductivity of Nafion is superior, probably due to longer and thus more flexible side chains. Other SSC membranes like Flemion (Asahi Glass Co. Ltd.), Aciplex (Asahi Chemical Co. Ltd.) and 3M membrane (3M Inc.) are known to allow higher operational temperatures due to higher crystallinity at a given EW and thus a higher glass transition temperature, as compared to Nafion. Aquivion (Solvay Solexis) membranes are also currently marketed and appear to outperform Nafion with their lower methanol crossover and superior water management (Devanathan 2008). However, in hydrogen-fed FC applications, they are less effective, as their conductivity at 100°C decreases from 11 S m−1 at RH 90% to 7 S m−1 at RH 50%.

An interesting approach to overcome excessive swelling was proposed by 3M Company (Wycisk et al. 2014). Their membranes with EW 580 demonstrated conductivity of 10 S/m at 80°C and RH 50%, yet, below EW 800, polymer crystallinity disappeared, resulting in membrane excessive swelling and loss of mechanical strength in a water vapor environment. In liquid water, the polymer became soluble at elevated temperatures. Later, 3M Company synthesized perfluorinated and partially fluorinated ionomers with side chains containing multiple protogenic groups. This allowed longer runs of backbone CF2 groups between side chains for the same EW, rendering the membrane with crystallinity sufficient to be insoluble in liquid water. For example, such membranes with EW 625 demonstrated proton conductivity larger than 10 S m−1 at 80°C and RH 50%, similar to a PFSA film of EW 700, yet the water uptake was only half that of PFSA.

Ballard Power Systems developed a partially fluorinated PEM with aromatic side-chains based on trifluorostyrene compositions (Ballard Advanced Materials 3rd Generation, BAM3G) (Tian 2004, Devanathan 2008). It was claimed that in FC applications, BAM3G demonstrated a superior performance compared with Nafion and Dow membranes, as well as excellent durability. EW of 509 appears to be optimal from the conductivity standpoint (17 S m−1 at room temperature and λ=33). Significantly lower conductivities were found for lower EW because of acid dilution by an excessive water uptake (λ≈100) and for higher EW because of poor water uptake (λ≈10). As the styrenic side-chains of Ballard membranes are much more rigid than Nafion side-chains, phase separation is not well developed and sulfonic acid groups are distributed uniformly throughout the membrane. The aggregation of ionic sites is inhibited, and a large fraction of water in the membrane is not associated with ionic aggregates, so the proton conductivity mechanism in BAM membranes may be different from Nafion.

Sulfonated perfluorocyclobutane (sPFCB) ionomer developed by General Motors Company demonstrated good proton conductivity and gas crossover lower than that of PFSA films of similar thickness. The insufficient mechanical stability of this membrane material could be overcome by blending it with polyvinylidene fluoride (PVDF) or using a porous reinforcing mat of expanded polytetrafluoroethylene (ePTFE) (Kraytsberg and Ein-Eli 2014). The latter approach will be elaborated in Section 7.1.

6.2 Non-fluorinated PEMs

Ionomers based on sulfonated or phosphonated hydrocarbon polymers appear to be an attractive replacement to Nafion due to their low production cost and convenience of recycling (Peighambardoust et al. 2010). The main disadvantage of hydrocarbon PEMs is that very high acid group concentrations are required to achieve proton conductivities comparable with Nafion. This leads to very high water content, resulting in poor mechanical properties and even the risk of membrane dissolution in water. Phosphonic groups, being weaker acids than sulfonic groups, induce lower water uptake; however, their conductivity is lower, too. Another undesirable consequence is that loss of water and conductivity at low RH in these membranes is more severe than in Nafion, which makes them less attractive for PEMFC applications. As was demonstrated with sulfonated poly(ether ether ketone) (SPEEK), this difference arises from the backbone rigidity and lack of side-chains, as the sulfonic groups are attached directly to the backbone (Hickner et al. 2004, Dupuis 2011, Park et al. 2011a). The distances between the sulfonic groups are then greater and water channels are narrower, with more branches and dead-ends, leading to poorer connectivity. Also, the acidity of the sulfonic groups in hydrocarbon ionomers is lower due to lower electronegativity of the backbone, as compared with Nafion. Yet even given that, SPEEK and sulfonated poly(ether ether ketone ketone) (SPEEKK) demonstrated selectivity, expressed as the ratio of conductivity to methanol permeability(ηMeOH), seven times higher than Nafion.

Another limitation of hydrocarbon polymers is their inherently lower resistance to oxidative attack as compared with PFSA. In sulfonated polyimides (SPI), for example, this lower resistance arises from greater susceptibility of the sulfonated polymer backbones to chemical attack, as compared with non-sulfonated PI (DeLuca and Elabd 2006), or from their low molecular weight due to the low reactivity of sulfonated monomers. Sulfonated polysulfones, on the other hand, demonstrate excellent oxidative stability in an operating FC due to exceptionally low H2 and O2 gas permeability, resulting in low concentration of oxidizing species in the membrane (Wycisk et al. 2014).

Many attempts were made to overcome the limitations of conductivity and oxidative stability via synthesis of copolymers with desirable polymer architectures (e.g. block copolymers, high-free volume copolymers, grafted/branched copolymers etc.) (Zhang and Shen 2012b, Awang et al. 2015, Kim et al. 2015). The aforementioned PSS-b-PMB demonstrated a high degree of phase separation, stability of the ion-conducting channels and effective maintenance of hydration and conductivity at elevated temperatures and low RH (Park et al. 2007). Increasing system order results in faster percolation for proton transport due to lower percolation threshold ϕh,c and critical exponent n values, as defined by Hsu et al. (1980): σ=σ0(ϕhϕh,c )n. For example, for the triblock copolymer ionomer, sulfonated poly(styrene-isobutylene-styrene) (S-SIBS), ϕh,c =0.077 and n=0.76 were reported for proton conductivity (Elabd et al. 2003), lower than Nafion (0.1 and 1.5, respectively), which is a random copolymer. Interestingly, the same research reported higher values of ϕh,c and n for methanol permeability, 0.095 and 1.15, respectively. This indicates that increasing the ordering of the system can improve selectivity, which was about five times higher for SSIBS as compared to Nafion. The good connectivity between hydrophilic domains, however, also provides an effective transportation channel for free radicals, which can decrease the durability of these multi-block copolymer membranes (Zhang and Shen 2012a).

6.3 Incorporation of hygroscopic and proton-conductive fillers into PEMs

Water retention at low RH in Nafion and other sulfonated ionomers can be improved by incorporation of hygroscopic inorganic fillers (oxides of aluminum, silicon, titanium, zirconium, zirconium and cesium phosphates, zeolites, clays) (Li et al. 2003, Cele and Ray 2009, Ahmad et al. 2010, Peighambardoust et al. 2010, Patil et al. 2011, Thiam et al. 2011, Tripathi and Shahi 2011, Kim et al. 2015). These particles also improve thermal and mechanical stability of the parent polymer. Despite the increased water retention, the conductivity of the composites is generally on par with the unfilled polymer or lower, even when high filler content is used. The probable reason is that significant amount of water retained by the filler is not available for deprotonation of the sulfonic groups, and filler particles partially block ion-conducting pathways, increasing tortuosity. The composite membranes also become brittle at high filler concentration (Viswanathan and Helen 2007). Methanol permeability, however, also decreases due to pathway blockage and reduced water mobility in the composite membranes. Significant reduction of methanol permeability (>10 times) was reported for hygroscopic particles [e.g. zirconium phosphates (Alberti and Casciola 2003) and clays like laponite and montmorillonite (DeLuca and Elabd 2006)] that undergo exfoliation into thin sheets oriented in parallel to the membrane surface. The effect can be attributed to the orientation of the exfoliated sheets, which is normal to the direction of methanol transport, so the methanol diffusion pathways are significantly extended. An additional advantage of that geometry is a high aspect ratio of the filler particles, resulting in large organic/inorganic interfaces, which help avoid generation of cavities (Zhang and Shen 2012b). Curiously, only a minor decrease in proton conductivity was measured in the composites. This result, however, can be misleading as conductivity is often measured along the membrane surface, while methanol permeability is measured in the direction normal to the surface.

A feasible solution for the problem of poor conductivity mentioned above is using bifunctional fillers, which are both hydrophilic and conductive. They can replace the bulk water in the ionomer channels, thus significantly reducing methanol transport, while still participating in the proton transport. Nafion membranes filled with sulfonated silanes demonstrated ηMeOH three times higher than the pristine Nafion N117 membrane (Ladewig et al. 2007). A hybrid membrane based on SPI and sulfonated molecular sieve Si-MCM-41 showed significantly higher PEMFC performance than that of pure SPI or Nafion membranes (Zhang and Shen 2012b).

7 Pore-filling composite PEMs

The approaches to improving PEM reviewed above aim primarily at developing membranes with enhanced proton conductivity at elevated temperatures and low RH and/or reduced permeability to reactants. These approaches are described in hundreds of publications and are regularly and thoroughly reviewed, as summarized above. However, another important limitation of Nafion is loss of mechanical strength. Nafion, being a rubbery material with low modulus but high elongation-to-break, shows reasonably robust mechanical properties up to 80°C. This is advantageous for MEA fabrication; however, it poses a problem for long-term performance in high-temperature environments. Because of its low Tg, hydrated Nafion undergoes a viscoelastic relaxation causing the membrane to develop pinholes when operated at temperatures near or above 100°C (Hickner et al. 2004). The need for mechanical strength precludes reducing the membrane thickness (and thus resistance and cost) below some critical level. This limitation can be solved by means of a reinforced composite PEM, in which a highly proton-conductive ionomer is filled into a highly porous matrix membrane with good mechanical stability. In addition, selectivity in this type of membranes is expected to increase with the decreasing pore size as a result of increasingly suppressed swelling of the filling polymer, as was proposed by Yamaguchi et al. (1991).

The most known composite of this type is Nafion-filled commercial porous polytetrafluoroethylene (PTFE) film (Gore-Tex). It was first reported by Penner and Martin (1985) and then claimed by Gore and Associates, Inc., under the Gore Select trademark (Savadogo 1998). The Gore Select membranes are very thin (12–20 μm), which allows lower resistance and higher power densities in FC tests, even though the hydrophobic PTFE matrix significantly lowers conductivity and hydraulic permeability, as compared with Nafion. The tensile strengths of dry Nafion and Gore Select membranes were found to be similar; however, strength loss under hydration in the composite membrane was significantly lower. The main advantage of the Gore Select membranes over Nafion was found in significantly increased lifetime, which was attributed to the PTFE matrix acting as a barrier to propagation of defects (Savadogo 2004).

In contrast with the other approaches mentioned before, the literature on pore-filling composite PEMs is quite scarce and, to the best of our knowledge, was reviewed only once (Subianto et al. 2013). A discussion on the existing literature on the subject is, therefore, presented below. It is divided into two principal categories by the support type, as shown in Figure 5:

Figure 5: A schematic illustration of the porous support types: (A) isotropic microporous membrane, (B) anisotropic nanoporous membrane with parallel pores oriented normal to the surface.
Figure 5:

A schematic illustration of the porous support types: (A) isotropic microporous membrane, (B) anisotropic nanoporous membrane with parallel pores oriented normal to the surface.

  1. Membranes with isotropic microporous support, where the properties of pore-filling ionomer are expected to be influenced mainly by swelling suppression and physical reinforcement provided by the support.

  2. Membranes with anisotropic support bearing highly ordered parallel pores oriented normal to the surface. Membranes of this type are the main subject of this work, as pore orientation and size are expected to increase proton conductivity due to alignment of ion channels, as was explained in Section 5.

7.1 Isotropic pore-filling membranes

The material most widely used for support membranes is porous ePTFE due to its high porosity, mechanical and chemical stability and good compatibility with Nafion. The basic technique for preparation of an ePTFE-Nafion composite membrane is pouring Nafion solution on a pre-cleaned ePTFE film, followed by drying and annealing. The annealing of recast Nafion at temperatures above Nafion Tg=105°C is essential to allow the backbone chain to reorganize and render the film insoluble (Zook and Leddy 1996). It should also improve adherence between the PTFE domains in Nafion backbone and porous matrix due to Van der Waals forces. Liu et al. (2003) used this technique with a≈4.5wt.% in-house made Nafion solution in three kinds of alcohol+N,N-dimethylformamide (DMF) and a PTFE film (0.3–0.5-μm pore size) pre-cleaned with ethanol. The composite membrane demonstrated very effective pore impregnation with Nafion and, similarly to the Gore Select membrane, smaller loss of tensile strength upon hydration and smaller water uptake, which was attributed to higher hydrophobicity due to the ePTFE matrix. The composite membrane performance in H2/O2 FC (80°C) was slightly better than the N115 membrane, especially at current density above 1000 mA cm−2. The oxygen permeability coefficient, however, was 15% higher for the composite, which could be attributed to higher intrinsic PTFE permeability to O2 as well as to non-ideal pore plugging.

The group of Huang investigated impregnation of ePTFE film (pore size 0.5 μm, porosity 85%) pre-cleaned by boiling in acetone with 5 wt.% Nafion solution in a mixture of water, propanol, ethanol, methanol and unspecified ethers (DuPont) (Lin et al. 2004, Yu et al. 2004, Lin et al. 2005). They demonstrated that pore impregnation is poor with the original DuPont solution but becomes better with the addition of Triton X-100 surfactant (up to 5 wt.%), which both improves compatibility between the solvent mixture and the PTFE matrix and reduces the size of Nafion aggregates in the solution (Lin et al. 2004). The composite membrane demonstrated σ and PH2, respectively, by 3 and 10 times lower than Nafion and outperformed the N112 membrane in H2/O2 FC (70°C) at current density above 1000 mA cm−2. The aggregation in Nafion solutions in alcohol-water mixtures (4:1 wt. ratio) also decreased in the following order: methanol→ethanol→isopropanol (Lin et al. 2005), as the solubility parameter δ of isopropanol (IPA) is the closest to that of Nafion. Also, ePTFE shrinks under evaporation-induced stress, which is highest for IPA (Zhengbang et al. 2011). This property might be beneficial for Nafion/ePTFE membrane preparation, leading to decrease in pore size during the impregnation and thus tighter pore-filling. The composite prepared from this solution demonstrated properties similar to those found in the previous research, i.e. effective pore impregnation, σ and PN2 lower than those of Nafion and better performance in H2/O2 FC (70°C) (Yu et al. 2004). A similar trend was observed for PMeOH and performance in a DMFC; moreover, performance deterioration with increasing MeOH concentration was less significant for the composite membrane than for the N112 membrane (Lin et al. 2005).

Ramya et al. (2006) showed that adding ethylene glycol, a high-boiling-point solvent, could probably further improve impregnation and tighten the composite membrane, while adding n-hexane, a low-boiling-point solvent, results in fissure formation. Teng et al. (2013) repeated the methods of Liu et al. (2003) and presented a membrane prepared from a 5 wt.% Nafion/DMF solution, which looked essentially the same as a Nafion 212 membrane both to the naked eye and in scanning electron microscopy (SEM) pictures. The composite was completely transparent, in contrast to the milky-white original PTFE film, which was taken as evidence of tight and complete pore-filling.

Wang et al. (2011) concluded that Tg of PFSA films increased upon exchange of H+ to alkali ions Li+, Na+, K+ reaching 290°C for Na+, which is the highest among the four and the closest to the Tg of ePTFE (340°C). Consequently, the recast PFSA films and composite PFSA/ePTFE prepared from Nafion 5 wt.% DuPont solution (original or alkali-exchanged), annealed at the corresponding Tg and finally converted to H+ form demonstrated a change in various properties. For the membranes prepared from the H+-PFSA solution, the crystallinity, the degree of PTFE pore-filling, σ and the PEMFC performance were the lowest, while those prepared from the Na+-PFSA solution were the best in terms of all the above properties. In contrast, H2 crossover and solubility in alcohol were the highest for H+-PFSA and the lowest for Na+-PFSA. It can be concluded that Na+-PFSA solutions are more suitable for preparation of PTFE pore-filling membranes; however, this study did not rule out the possibility that all the differences were caused mainly by the increase in the annealing temperature.

The aggregation of Nafion in pore-filling solutions appears to become the dominant problem with decreasing pore size, requiring special precautions to achieve effective impregnation. For example, Yang et al. (2013) demonstrated that pore-filling using ~4wt.% Nafion solution in IPA-water mixture (4:1 wt. ratio), which was good for 0.3–0.5-μm pore size (Lin et al. 2005), was poor for 0.1–0.2-μm pores. The filling was, however, successful using a similar solution of Nafion precursor, perfluorosulfonyl fluoride (PFSF), followed by alkali hydrolysis and conversion to sulfonyl chlorides SO2F to sulfonic acid groups SO3H. The advantage of the method is that unhydrolyzed -SO2F groups in PFSF, in contrast to SO3H groups in Nafion, do not induce microphase separation and aggregation in water-alcohol solutions. The resulting performance of the composite membrane PEMFC was very similar to that of the Nafion 211 membrane.

Tang and co-workers investigated impregnation of chemically modified ePTFE film (pore size, 0.1–0.2 μm; 85% porosity) with 5 wt.% Nafion (DuPont) (Tang et al. 2007a,b,c, 2008). The impregnation technique was based on soaking the ePTFE film in the ionomer solution followed by drying, which was repeated three times. The pore-filling of the as-received ePTFE film with Nafion solution and 5% Triton X-100 added was incomplete, again, in contrast to the effective filling of larger pores with the same solution (Yu et al. 2004). However, it improved significantly with pre-hydrophilization of the PTFE film via treatment with sodium naphthalene and N-methyl acrylamide (Tang et al. 2007a). Comparing with a Nafion 211 membrane, the composite demonstrated lower water uptake and conductivity, similar H2 crossover and slightly better PEMFC performance. It must be noted, however, that the composite membranes were only briefly annealed (5 min at 120°C) and probably could be partially dissolved upon immersion in IPA, which was required to remove Triton X100 after Nafion deposition. It was claimed that lowering the initial pressure to 500Pa before applying the Nafion+Triton X100 solution to the as-received ePTFE film significantly improved pore impregnation (Tang et al. 2007a). It was also demonstrated that the use of ePTFE treated with sodium naphthalene for impregnation with Nafion solution in the Na+ form followed by annealing at 270°C (Tang et al. 2007c) resulted in a superior resistance to Fenton oxidation compared with Nafion 111 membrane. The DMFC performance of such a membrane, however, was very poor (Tang et al. 2008). A possible explanation is that the annealing at 270°C for 2 min was still too brief to induce sufficient degree of crystallinity and thus reduce PMeOH.

Improper choice of the solvent, Nafion concentration, matrix material, pore size and pore-filling method frequently lead to poor pore impregnation. For instance, the following four examples proved to be unsuccessful combinations:

  • porous polyethylene, polypropylene and PTFE supports (pore size, 0.1–0.5 μm) and 18wt.% Nafion solution in ethanol (Nouel and Fedkiw 1998);

  • polysulfone and microfiber glass fleeces (pore size, 7–15 μm) and Aldrich 5 wt.% Nafion solution (Haufe and Stimming 2001);

  • porous polyethylene, polypropylene and PTFE supports (pore size, 0.1–10 μm) and 20–25 wt.% Nafion solution in DMF (Rodgers et al. 2008); and

  • ePTFE support (pore size 0.075 μm) and DuPont 5 wt.% Nafion solution applied by solution spraying (Xing et al. 2013).

Apart from Nafion, some other ion-conductive polymers were used to impregnate ePTFE films. Shin et al. (2005) prepared sulfonated polystyrene (SPS)/PTFE membranes by direct polymerization of styrene and DVB (divinylbenzene, the crosslinker) in the ePTFE pores, followed by sulfonation. The degree of sulfonation decreased with the crosslinker concentration, as it became more difficult for the reactants to access the polymer chains. As expected, water content, methanol permeability and proton conductivity dropped, as the degree of sulfonation decreased. In comparison with Nafion 117, however, ion exchange capacity (IEC, measured in milliequivalents of acid groups per gram of dry membrane), water content and conductivity were higher while methanol permeability was lower at all degrees of crosslinking. As a result, ηMeOH increased with the crosslinking degree up to a value of about eight times higher than ηMeOH in Nafion for the styrene/DVB ratio of 85/15. It is unclear, though, how the ePTFE matrix influenced the membrane properties, as no comparison with bulk SPS film was made.

In another study, SPEEK/PTFE membranes (pore size, 0.3–0.5 μm; porosity, 85%) prepared from SPEEK solutions in N-methyl pyrrolidone (NMP) (Xing et al. 2005) and N,N-dimethylacetamide (DMAC) (Wei et al. 2012) demonstrated lower σ and water uptake and higher oxygen permeability, as compared with recast SPEEK films. The PEMFC performance was below that of both recast SPEEK and Nafion. Moreover, the composite membranes included a thick layer of bulk SPEEK above the SPEEK/PTFE region, thus concealing the “true composite” properties.

A similar result was obtained with Na+-Aquivion solution in IPA+Triton X-100 (Xiao et al. 2013). Composite membrane conductivity and PEMFC performance, due to thick surface layers and apparently poor pore-filling, were nearly identical to the recast Aquivion film. Using the same ePTFE support, Zhu et al. (2007) prepared composite membranes from 5 wt.% solutions of disulfonated poly(arylene ether sulfone) (SPSU) and Nafion in a mixture of dimethyl sulfoxide (DMSO) and n-butanol (1:1 v/v). Interestingly, the SPSU/PTFE membrane prepared from the solution in DMSO only was brownish and opaque, as compared with the transparent and almost colorless SPSU/PTFE membrane prepared from the DMSO/n-butanol solvent. This indicated that addition of n-butanol increased compatibility between the SPSU and PTFE. The SPSU/PTFE water uptake was smaller than for the recast SPSU, while in Nafion/PTFE and recast Nafion, it was about the same. The conductivities of the composites were about 50% of the corresponding bulk polymer. Curiously, the PEMFC performance increased in the order SPSU→Nafion→Nafion/PTFE→SPSU/PTFE. This order can be explained by higher intrinsic conductivity of SPSU and lower thickness of the composite membranes. However, it is also possible that the choice of solvents was more compatible with SPSU than with Nafion, resulting in better pore impregnation and higher relative increase in performance.

Some research was dedicated to Nafion filling into electrospun nanofiber mats, which provide relatively large and open pores and thus are easily impregnated. The conductivity of Nafion-impregnated polyvinylidene fluoride (PVDF) electrospun mats demonstrated an increase in σ with Nafion content (Choi et al. 2008). It was still, however, an order of magnitude lower than for bulk Nafion due to the insulating properties of PVDF. The DMFC performance, though, was better than for Nafion 115 membrane of commensurate thickness due to lower methanol crossover. The problem of immiscibility of Nafion and PVDF was addressed by preparation of Nafion-modified PVDF nanofiber mats, impregnated then with Nafion (Li and Liu 2014). PEMFC performance of the resulting composite membrane was better than for Nafion 212 and Nafion/non-modified PVDF composite. DMFC performance was better than for Nafion 117 membrane, due to higher σ and lower PMeOH. It must be noted, however, that the conductivity was measured in parallel to the surface, which is also the prevalent orientation of the nanofibers, while the direction of proton conduction in FCs is normal to the surface.

Similar results were obtained for Nafion-filled mats made of PVDF grafted with poly(styrene sulfonic acid) (PSSA) (Li et al. 2014). The composite demonstrated higher water uptake and σ due to the additional sulfonic groups in the nanofibers, and lower PMeOH, as compared to Nafion 212 membrane. PEMFC performance of the composite was notably better than for Nafion, although DMFC performance was only slightly better, as the increase in selectivity to methanol was moderate. An interesting result was obtained with Nafion-filled SPEEK nanofibers mats, in which conductivity was higher than in both Nafion and recast SPEEK over a wide range of temperatures and RH values (Xu et al. 2015). However, similar to the above report by Li and Liu (2014), conductivity was measured in parallel to the surface and may not represent correctly the conductivity in the direction normal to the surface. Moreover, this modification resulted in only moderate increase in ηMeOH, as the intrinsic methanol permeability of SPEEK is of the same order of magnitude as that of Nafion.

Wang and Lin (2014) studied Nafion-filled mats made from poly(vinylidene fluoride-co-hexafluoropropylene) (PVDP) or PVDP blended with polybenzimidazole (BI) to improve the compatibility with Nafion. Nafion content was 90 wt.%. The conductivity of the Nafion/PVDP-BI composite was on par with Nafion, but PMeOH was about 200 times lower. Both conductivity and permeability decreased with the BI content due to BI interaction with the sulfonic groups, thereby the resulting selectivity for all the composite membranes was about 200 times higher than for Nafion. Addition of the minimal fraction of BI (5 wt.%) was found to be optimal, resulting in the highest selectivity and best DMFC performance.

Several studies dealt with Nafion/poly(vinyl alcohol) (Zhang et al. 2009, Lin et al. 2010, Molla and Compan 2011, Molla et al. 2011, Lin and Wang 2014) and poly(lactic-co-glycolic acid) (Wang et al. 2013) electro-spun fiber composite membranes. Due to the high hydrophilicity of PVA and PLGA fibers, the composite membranes showed only moderate decrease in conductivity and methanol permeability, while the water uptake could be on par with Nafion or even higher (Molla and Compan 2011, Lin and Wang 2014). DMFC and PEMFC performance of such membranes was also on par with Nafion or only slightly better, indicating that PVA and PLGA nanofibers induce little change in membrane properties, except for significant increase in tensile strength for PVA (Molla et al. 2011). Similar results were obtained with triple-layer composite membranes, in which a sulfonated polyether sulfone (SPES) fiber mat filled with Nafion was sandwiched between two layers of plain Nafion (Hasani-Sadrabadi et al. 2011, Shabani et al. 2011). The composite DMFC performance was on par with N117 membrane at methanol concentration 1 m and notably better at 5 m.

Apart from Nafion, sulfonated poly(arylene ether sulfone) (SPAES) was filled into nonwoven polyacrylonitrile (PAN) nanofiber mat (Yu et al. 2013). The recast SPAES membranes demonstrated higher water uptake, conductivity and PEMFC performance at RH 100%, as compared with Nafion 212 membrane. However, typically for hydrocarbon PEMs (Section 6.2), conductivity and performance deteriorated faster for SPAES with decreasing RH. The SPAES-PAN composite membrane characteristics were below those of both Nafion and recast SPAES, except for tensile strength, which increased significantly due to the PAN reinforcement. Membranes made from SPEEK-PVA blends, recast or impregnated into mats of electrospun SPEEK-PVB (polyvinyl butyral) demonstrated significant decrease in water uptake, σ and PMeOH with increasing annealing temperature (from 110 to 140°C) (Molla and Compan 2015). As compared with recast SPEEK-PVA films, water uptake and σ were lower in the composite membranes while PMeOH remained almost the same. However, both conductivity and selectivity were significantly lower for all the experimental membranes than for Nafion.

Yamaguchi et al. (2003) noted that swelling suppression, especially in the in-plane dimension, is important in PEMs for DMFCs not only to reduce methanol crossover but also to prevent detachment of carbon electrodes, which do not swell. Using porous supports filled with poly(acrylamide-tert-butylsulfonic acid) (PTABS) by direct polymerization in the pores, they showed that composites based on cross-linked high-density polyethylene (CLPE) and polyimide (PI) demonstrate significantly lower areal swelling in the surface dimension in water than those based on ePTFE supports. In comparison to Nafion 117, methanol permeability of PTABS/CLPE composites was many times lower and increased to a significantly lesser degree with methanol concentration, while proton conductivity was only about 50% lower (Yamaguchi et al. 2005). As a result, the composite DMFC performance was only slightly better than N117 at methanol concentration of 8 wt.% but significantly better at 32wt.%.

Yildirim et al. (2008) reported on composite membranes made by filling of ultra-high molecular weight polyethylene (UHMWPE, pore size 0.7 of μm) with Nafion (Yildirim et al. 2008) and sulfonated poly(phthalazinone ether ketone) (SPPEK) (Yildirim et al. 2009). Proton conductivity and methanol permeability of the recast SPPEK film were about 20 and 10 times lower than for the N117 membrane. Nafion/UHMWPE swelling in area dimension was negligible and independent on methanol concentration, while N117 swelling increased. The total swelling by weight, however, was significantly higher for the composite and increased much stronger with methanol concentration, approaching that of recast Nafion. Proton conductivity and methanol permeability of the composite were about six and two times lower, respectively, than for the N117. In SPPEK/UHMWPE composite, both types of swelling decreased by about half while conductivity and methanol permeability decreased about three times, as compared to the recast SPPEK film. Therefore, in both the Nafion and SPPEK cases, selectivity of the composite membrane was notably lower than in N117. The most probable reason could be the incomplete filling of the porous support (≈60% out of ≈90% pore volume fraction) and effect of the bulk ionomer layers deposited on the composite membrane surfaces.

Navarro et al. (2008) filled the UHMWPE supports (pore size, 0.3 μm) from a 5% solution of sulfonated and hydrogenated styrene/butadiene block copolymer (SHSBS) in chloroform/ethanol mixture. Several plasma treatments were tested on the supports to improve the pore-filling. All the composite membranes demonstrated significantly lower methanol permeability as compared with N117, while conductivity only slightly decreased. Electron cyclotron resonance (ECR) treatment in argon provided the best results, increasing selectivity 1.6 and 10 times, as compared to untreated SHSBS/UHMWPE and N117, respectively. Interestingly, liquid uptake in water-methanol solutions decreased with methanol concentration for the composite membranes, in contrast with N117, where it increased.

Porous polyimide (PI) is probably the most effective material for suppressing PEM swelling, because it is rigid and mechanically very strong, as was shown by Yamaguchi et al. (2007) using a PI support (pore size, 0.1 μm; porosity, 55%), filled with SPAES by direct polymerization in the pores. As compared with N117, composite conductivity was three times lower while methanol permeability was 300 times lower and did not increase with methanol concentration or temperature. Later, the same group presented a PI support (pore size, 0.1–0.3 μm; porosity, 50%) filled with SPAES from 10 wt.% solution in NMP (Hara et al. 2009). The increase in water content with IEC in the composite was significantly lower than in the recast SPAES film. Conductivity in SPAES/PI was an order of magnitude lower than in SPAES for low IEC values but closed the gap significantly as IEC increased. This is because conductivity in SPAES reached a plateau due to sulfonic group dilution with bulk water, while restricted swelling in SPAES/PI composite resulted in the absence of bulk (“free”) water, so conductivity increased through the entire IEC range. The authors attributed the extremely low methanol permeability of SPAES/PI to the absence of bulk water and suggested that the proton conduction mechanisms in pore-filling PEMs with high swelling restriction may be quite different from bulk PEMs. It can be related to the “packed-acid mechanism” (Ogawa et al. 2014), which does not require water movement and thus can possibly operate through closely located acid groups and bound water in the PEMs with suppressed swelling.

Similar results were obtained by Nguyen for sulfonated poly(styrene-ran-ethylene) (SPSE)/PI (Nguyen et al. 2009) and Nafion/PI (Nguyen and Wang 2010) composite membranes (support porosity, 80%). In both cases, water uptake in the composite membrane was about one-third of the bulk ionomer, while proton conductivity remained almost the same, and methanol permeability decreased by two orders of magnitude. With the sulfonated polyimide (SPI)/PI membranes, however, the increase in selectivity, as compared with recast SPI, was moderate, although swelling was reduced by half (Wang et al. 2012).

7.2 Anisotropic pore-filling membranes

The results presented in the previous section demonstrate that confinement of ionomers in isotropic rigid and porous supports results in proton conductivity at best on par and generally lower than of the bulk ionomer, even for highly porous supports (80–85% porosity). The reason is that supports of that type increase the tortuosity of proton-conducting paths and, apparently, induce no preferable orientation in the pore-filling ionomer phase. The only exception was demonstrated with ionomer-filled mats of nanofibers bearing sulfonic groups (Li and Liu 2014, Xu et al. 2015) when conductivity was measured in parallel to the surface. As this is also the prevalent orientation of the nanofibers and significant conductivity enhancement along the nanofiber was demonstrated (Dong et al. 2010), the “true” through-plane conductivity of these PEMs might be substantially lower.

Therefore, there is significant evidence (see also Section 5) that conductivity of Nafion and other proton conducting polymers can be increased by alignment. However, to increase “trough-plane” conductivity, alignment normal to the membrane surface is required. This could be achieved by filling an array of highly ordered parallel pores with Nafion or other proton-conducting polymer, i.e. to use a support membrane that would provide a properly aligned solid interface, as shown in Figure 5B.

Anodized aluminum oxide (AAO) membranes and polycarbonate track-etched (PCTE) membranes are the most popular architectures for making anisotropic structures aligned normal to the surface. For instance, an increased electric conductivity of nanowires prepared within their pores is reported, as compared with the same bulk polymerized materials (Malinauskas et al. 2005). However, the examples of PEM preparation with AAO and PCTE membranes are even scarcer than with the isotropic supports reviewed earlier. Probably, the first publications by Leddy and Vanderborgh (1987) and Fang and Leddy (1995) presented PCTE membranes filled with Nafion from an in-house-made ~3wt.% water-ethanol solution (50:50) by soaking and drying. A 20-fold increase in cation flux with pore diameter decreasing from 600 to 30 nm was reported, and increasing ordering of Nafion within the pores was suggested as an explanation for the phenomenon. However, no SEM visualization or structural study results were provided.

Later, Raghav (2005) investigated pore-filling of PCTE (pore diameters are 10, 50, 100nm) and AAO (pore diameters are 20, 100nm) membranes with Nafion and sulfonated polystyrene (SPS). Applying the methods of Leddy and Vanderborgh (1987) to commercial Nafion solutions (Sigma-Aldrich, Ion Power) failed entirely to show any filling of the PCTE membrane pores. Several other impregnation methods, including vacuum filtration, also failed to fill the PCTE and AAO membranes with Nafion. Vacuum filtration of 5 wt.% SPS solution in acetone, however, resulted in the formation of well-defined SPS nanorods in the PCTE pores and incomplete, although still noticeable filling of the AAO pores. These results were attributed to poor substrate wetting by the Nafion water-alcohol solutions, as opposed to very good wetting by the SPS/acetone solution.

Bocchetta et al. (2007) used prolonged soaking of AAO membranes (pore size, 200nm) in a commercial Nafion solution; however, pore-filling was mostly next to the AAO membrane surface without significant penetration into the pores. A work based on the same support filled with Nafion solutions (5 and 20wt.%) under low vacuum (0.1–0.3atm) offered some interesting results (Rayon et al. 2006). While no pore-filling was achieved with the as-received materials, widening the AAO pores to 500nm resulted in some Nafion deposition, and neutralization of Nafion into Na+ form allowed the formation of nanorods in the 200-nm pores, as verified by SEM. No conductivity/selectivity results were provided, though. The only example of successful pore-filling with Nafion combined with electrochemical measurements could be found with straight-hole PI support (Guo et al. 2012). However, very large pore diameters (50–200 μm) were employed. As a result, composite PEM conductivity increased with pore size, mostly due to the increasing Nafion fraction and with RH, similarly to the Nafion 211 membrane. PEMFC performance under various conditions was, at best, on par with Nafion 211.

Available data suggest that the main obstacle for Nafion-H+ impregnation into pores of nanometric size was the aggregation of Nafion in the solutions. Indeed, recent reports indicate that there is a threshold concentration C*, depending on the solvent, below which Nafion micelles break up to single molecules of estimated ~5-nm width and ~100-nm length (Aldebert et al. 1988) and hydrodynamic radius (Rh) of 20–40 nm (Lee et al. 2004, Zhang et al. 2008). C* was found to be about 1 g l−1 in water and water-methanol mixtures (Jiang et al. 2001, Lee et al. 2004). On the other hand, in 10:9 v/v IPA/water mixture, a better solvent for the perfluorinated backbone, C* increases to ~4.5 g l−1 (Zhang et al. 2008). As previous attempts to fill PCTE or AAO pores with Nafion (Leddy and Vanderborgh 1987, Fang and Leddy 1995, Raghav 2005, Bocchetta et al. 2007) were carried out with solutions containing ≥30 g l−1 Nafion in solvents containing ≥50% water and <15% IPA, Nafion molecules must have aggregated to form large particles (>500 nm) (Jiang et al. 2001, Zhang et al. 2008) and were unlikely to enter 100–200-nm pores. This is also consistent with the aggregation problems during the pore-filling of PTFE membranes described in the previous section.

Recently, Gloukhovski et al. (2016) demonstrated successful filling of the AAO membrane (pore diameter, 200 nm) by through-the-membrane evaporation of Nafion solution (~1.2 g l−1 in IPA) under vacuum. Through-plane σ and ηNaCl of the pore-filling Nafion were two and four times higher, respectively, than for bulk recast Nafion. Later, the same group also reported on Nafion impregnation into PCTE (pore diameter, 600–1000 nm) membranes (Gloukhovski et al. 2017). The pore-filling Nafion demonstrated increased proton conductivity, presumably due to the lower tortuosity and better connectivity of water channels aligned by the pore walls. Methanol and sodium chloride permeability, on the other hand, decreased significantly, presumably due to the suppressed swelling, resulting in narrower water channels and stronger “salting-out” and Donnan exclusion effects. Presumably, confinement could also encourage formation of regions of closely packed sulfonic groups, which facilitate proton transfer but strongly exclude other solutes. The alignment of the pore-filling Nafion, however, could not be demonstrated by SAXS, presumably due to the insufficient amount of Nafion in the experimental membranes. Also, no evidence of enhanced structural stability at elevated temperatures could be found.

Examples of successful filling of PCTE and AAO pores with other polymers (not necessarily in FC context) also exist in the literature, e.g. pore-filling of PCTE (pore diameters, 30–400nm) with polyethylene oxide (Vorrey and Teeters 2003) and of AAO (pore diameters, 20–200nm) with poly(phenylene sulfonic acid) (White et al. 2008). Direct attachment of sulfonic groups to PCTE pore walls (pore diameter, 10–2000 nm) by graft polymerization was also reported (Chen et al. 2006). In all these examples, proton conductivity increased with decreasing pore diameter; however, even maximal conductivity was low and selectivity was not reported.

Oleksandrov et al. (2009) attached sulfonic groups to the pores of a commercial AAO membrane (Whatman Anodisc; pore size, 20nm) through silanization. The composite membrane demonstrated conductivity three times higher and methanol permeability of about 10 times lower than Nafion 117 membrane under the same conditions. However, no SEM visualization was provided. It must also be noted that the bulk section of a 20-nm Anodisc membrane is identical to a 200-nm membrane, with 20-nm pores present only in an additional extra thin layer on top of the bulk membrane. This is because the main application of Whatman AAO membranes is filtration, for which the thin top layer is sufficient for size separation (Zhang 2006).

Another example of silanization + sulfonation, with porous silica substrate (pore diameter, 5–7nm; mouth diameter, 2 nm) showed overall conductivity of about 10 S m−1, almost independent of RH (Moghaddam et al. 2010). This weak dependence on RH, in contrast with Nafion, might arise from the stability of proton-conducting channels (Park et al. 2007), in which their structure is determined by support pore geometry and obviously cannot be affected by RH. It could also be related to capillary condensation of water within the nanopores, which were not filled with dense polymer, as in the above example, and evaporation of pore-filling water could be prevented by high curvature of the liquid-vapor interface at the pore mouth down to very low RH. It is hard to evaluate the intrinsic conductivity of the channels in the last two examples, though, as the pore fraction was not reported.

8 Summary and outlook

Ionomers and, in particular, Nafion remain the most successful and widely used ion-conductive material for commercial PEMs due to their unique combination of high proton conductivity, chemical stability and physical strength and flexibility. However, a common downside is high fuel permeability, especially of methanol, and loss of conductivity and structural instability at elevated temperatures. Confinement of ionomers in rigid porous matrices may efficiently overcome these limitations and enhance mechanical strength, conductivity, selectivity, stability and durability. The beneficial effects mainly come from suppressed swelling and preferential alignment of conductive pathways.

Isotropic porous matrices have been most common for preparing pore-filling composites. Membranes composed of supporting ePTFE matrix filled with Nafion were extensively studied and demonstrated significant benefits. Even greater potential was demonstrated using more rigid matrices, e.g. porous polyimides. Enhanced mechanical properties of the composite allow a thinner membrane and smaller Ohmic losses, which is most beneficial for the hydrogen PEMFCs. Suppressed swelling also reduces methanol crossover and improves proton-fuel selectivity, which is important for DMFCs.

Further gains in performance, especially in conductivity, may be achieved in anisotropic matrices through the alignment of ionic channels. A microscopic alignment may occur next to solid surfaces, up to several hundred nanometers from the interface. Another microscopic effect comes from the anisotropic stress developed in a swollen ionomer confined within a pore, promoting alignment along the pore axis. The last effect may come from the macroscopic orientation of the pores themselves. All such effects may be realized in anisotropic matrices with straight submicron pores oriented normal to the membrane surface. By using commercial AAO and PCTE substrates possessing such pore structure, Nafion-filled anisotropic composites were successfully prepared and demonstrated both enhanced proton conductivity and suppressed fuel crossover. Potentially, confinement in a solid matrix may also help stabilize the ionomer microstructure and thus improve thermal stability.

Despite successful proof-of-concept and much promise, there are still challenges and bottlenecks that one needs to overcome toward making pore-filling composite a viable alternative to presently used membranes, especially for anisotropic composites. Among these most notable are the following:

  • Availability of rigid and robust supports, which must be also affordable, upscalable and thermally and chemically stable. This requirement is particularly critical for anisotropic matrices, as presently available substrates are expensive and have poor mechanical characteristics.

  • Efficient filling of the substrate nanopores with ionomer, which is often difficult due to poor wetting and adhesion, ionomer aggregation and slow filling kinetics. Thorough attention to the choice of solvent, ionomer concentration in solution, use of surfactants, substrate type, chemical and physical pretreatments, prevention of pore blocking (e.g. by applying vacuum to prevent it), and filling procedure are critical for successful pore-filling.

  • Poor understanding of the confinement and alignment and their impact on performance, which calls for development of suitable structural and transport characterization methods. In particular, this requires accurate measurements of through-plane conductivity for anisotropic membranes and facile assessment of selectivity. For the latter, salt permeability is suggested as an easily measurable indicator of selectivity.

About the authors

Robert Gloukhovski

Robert Gloukhovski is currently a researcher in the Department of Mechanical Engineering at Technion – Israel Institute of Technology. He received his PhD in Chemical Engineering from Technion under the direction of Assoc. Prof. Yoed Tsur and Assoc. Prof. Viatcheslav Freger in 2017. His graduate research was focused on preparation and characterization of pore-filling polymer composites for proton exchange membranes.

Viatcheslav Freger

Viatcheslav Freger is currently a faculty at the Department of Chemical Engineering of Technion – Israel Institute of Technology in Haifa, Israel. His main research interests are in the areas of membrane technology for water and energy, fundamentals and modeling of membranes, polymer science, electrochemistry, colloidal and surface science and advanced characterization of materials. He has published over 80 papers and book chapters, has given numerous invited talks at international meetings and holds several patents and patent applications in these fields.

Yoed Tsur

Yoed Tsur is currently a faculty at the Department of Chemical Engineering of Technion – Israel Institute of Technology in Haifa, Israel. He is the director of the Grand Technion Energy Program (GTEP). His main research interests are in the areas of point defect chemistry of oxide materials, fuel cells and supercapacitors, and advanced EIS characterization and analysis. He has published over 50 papers and has given numerous invited talks at international meetings.

Nomenclature

EW

equivalent weight/mol polymer mol−1 SO3H

K

partition coefficient

P

specific permeability, m2 s−1

Tg

glass transition temperature, °C

λ

hydration, mol H2O mol−1 SO3H

σ

proton conductivity, S m−1

η

selectivity

ϕh

hydrophylic phase volume fraction in ionomer

Acknowledgments

This research was carried out in the framework of the Grand Technion Energy Program (GTEP) and supported by the Adelis foundation for renewable energy research. We also acknowledge research support by the Israel National Research Center for Electrochemical Propulsion (INREP) of the Israel Science Foundation (ISF, grant no. 2797/11).

References

Ahmad H, Kamarudin SK, Hasran UA, Daud WRW. Overview of hybrid membranes for direct-methanol fuel-cell applications. Int J Hydrogen Energy 2010; 35: 2160–2175.10.1016/j.ijhydene.2009.12.054Search in Google Scholar

Alberti G, Casciola M. Composite membranes for medium-temperature PEM fuel cells. Annu Rev Mater Res 2003; 33: 129–154.10.1146/annurev.matsci.33.022702.154702Search in Google Scholar

Alberti G, Narducci R. Evolution of permanent deformations (or memory) in Nafion 117 membranes with changes in temperature, relative humidity and time, and its importance in the development of medium temperature PEMFCs. Fuel Cells 2009; 4: 410–420.10.1002/fuce.200800148Search in Google Scholar

Alberti G, Narducci R, Sganappa M. Effects of hydrothermal/thermal treatments on the water-uptake of Nafion membranes and relations with changes of conformation, counter-elastic force and tensile modulus of the matrix. J Power Sources 2008; 178: 575–583.10.1016/j.jpowsour.2007.09.034Search in Google Scholar

Alberti G, Narducci R, Di Vona ML, Giancola S. More on Nafion conductivity decay at temperatures higher than 80°C: preparation and first characterization of in-plane oriented layered morphologies. Ind Eng Chem Res 2013; 52: 10418–10424.10.1021/ie303628cSearch in Google Scholar

Aldebert P, Dreyfus B, Gebel G, Nakamura N, Pineri M, Volino F. Rod like micellar structures in perfluorinated ionomer solutions. Journal de Physique 1988; 49: 2101–2109.10.1051/jphys:0198800490120210100Search in Google Scholar

Amphlett JC, Baumert RM, Mann RF, Peppley BA, Roberge PR, Harris TJ. Performance modeling of the Ballard Mark IV solid polymer electrolyte fuel cell I. Mechanistic model development. J Electrochem Soc 1995; 142: 1–8.10.1149/1.2043866Search in Google Scholar

Awang N, Ismail AF, Jaafar J, Matsuura T, Junoh H, Othman MHD, Rahman MA. Functionalization of polymeric materials as a high performance membrane for direct methanol fuel cell: a review. React Funct Polym 2015; 86: 248–258.10.1016/j.reactfunctpolym.2014.09.019Search in Google Scholar

Ayers KE, Anderson EB, Capuano CB, Carter BD, Dalton LT, Hanlon G, Manco J, Niedzwiecki M. Research advances towards low cost, high efficiency PEM electrolysis. ECS Trans 2010; 33: 3–15.10.1149/1.3484496Search in Google Scholar

Baltianski S, Tsur Y. Analysis of impedance spectroscopy data – finding the best system function. J Electroceramics 2003; 10: 89–94.10.1023/A:1025639604939Search in Google Scholar

Bason S, Kaufman Y, Freger V. Analysis of ion transport in nanofiltration using phenomenological coefficients and structural characteristics. J Phys Chem B 2010; 114: 3510–3517.10.1021/jp911615nSearch in Google Scholar PubMed

Bass M, Berman A, Singh A, Konovalov O, Freger V. Surface structure of Nafion in vapor and liquid. J Phys Chem B 2010; 114: 3784–3790.10.1021/jp9113128Search in Google Scholar PubMed

Bass M, Berman A, Singh A, Konovalov O, Freger V. Surface-induced micelle orientation in Nafion films. Macromolecules 2011; 44: 2893–2899.10.1021/ma102361fSearch in Google Scholar

Bocchetta P, Conciauro F, Di Quarto F. Nanoscale membrane electrode assemblies based on porous anodic alumina for hydrogen–oxygen fuel cell. J Solid State Electrochem 2007; 11: 1253–1261.10.1007/s10008-007-0280-xSearch in Google Scholar

Bose S, Kuila T, Nguyen TXH, Kim NH, Lau K, Lee JH. Polymer membranes for high temperature proton exchange membrane fuel cell: recent advances and challenges. Prog Polym Sci 2011; 36: 813–843.10.1016/j.progpolymsci.2011.01.003Search in Google Scholar

Cable KM, Mauritz KA, Moore RB. Anisotropic ionic conductivity in uniaxially oriented perfluorosulfonate ionomers. Chem Mater 1995; 7: 1601–1603.10.1021/cm00057a002Search in Google Scholar

Carmo M, Fritz DL, Mergel J, Stolten D. A comprehensive review on PEM water electrolysis. Int J Hydrogen Energy 2013; 38: 4901–4934.10.1016/j.ijhydene.2013.01.151Search in Google Scholar

Cele N, Ray SS. Recent progress on Nafion-based nanocomposite membranes for fuel cell applications. Macromol Mater Eng 2009; 294: 719–738.10.1002/mame.200900143Search in Google Scholar

Chen H, Palmese GR, Elabd YA. Membranes with oriented polyelectrolyte nanodomains. Chem Mater 2006; 18: 4875–4881.10.1021/cm061422wSearch in Google Scholar

Cho KY, Eom JY, Jung HY, Choi NS, Lee YM, Park JK, Choi JH, Park KW, Sung YE. Characteristics of PVdF copolymer/Nafion blend membrane for direct methanol fuel cell (DMFC). Electrochim Acta 2004; 50: 583–588.10.1016/j.electacta.2004.03.063Search in Google Scholar

Choi P. Investigation of thermodynamic and transport properties of proton-exchange membranes in fuel cell applications’, Ph.D. Thesis, MA, USA: Worcester Polytechnic Institute, 2004.Search in Google Scholar

Choi P, Bessarabov DG, Datta R. A simple model for solid polymer electrolyte (SPE) water electrolysis. Solid State Ionics 2004; 175: 535–539.10.1016/j.ssi.2004.01.076Search in Google Scholar

Choi P, Jalani NH, Datta R. Thermodynamics and proton transport in Nafion II. Proton diffusion mechanisms and conductivity. J Electrochem Soc 2005; 152: E123–E130.10.1149/1.1859814Search in Google Scholar

Choi SW, Fu Y-Z, Ahn YR, Jo SM, Manthiram A. Nafion-impregnated electrospun polyvinylidene fluoride composite membranes for direct methanol fuel cells. J Power Sources 2008; 180: 167–171.10.1016/j.jpowsour.2008.02.042Search in Google Scholar

Cooper KR. Progress towards accurate through-plane membrane resistance and conductivity measurement. ECS Trans 2009; 25: 995–1007.10.1149/1.3210653Search in Google Scholar

Cooper KR. Progress toward accurate through-plane ion transport resistance measurement of thin solid electrolytes. J Electrochem Soc 2010; 157: B1731–B1739.10.1149/1.3481561Search in Google Scholar

Cooper KR. Characterizing through-plane and in-plane ionic conductivity of polymer electrolyte membranes. ECS Trans 2011; 41: 1371–1380.10.1149/1.3635668Search in Google Scholar

Cwirco EH, Carbonell RG. Ionic equilibria in ion-exchange membranes: a comparison of pore model predictions with experimental results. J Membr Sci 1992a; 67: 211–226.10.1016/0376-7388(92)80026-GSearch in Google Scholar

Cwirco EH, Carbonell RG. Interpretation of transport coefficients in Nafion using a parallel pore model. J Membr Sci 1992b; 67: 227–247.10.1016/0376-7388(92)80027-HSearch in Google Scholar

DeLuca NW. Nafion blend membranes for the direct methanol fuel cell’, Ph.D. Thesis, Philadelphia, PA, USA: Drexel University, 2004.Search in Google Scholar

DeLuca NW, Elabd YA. Polymer electrolyte membranes for the direct methanol fuel cell: a review. J Polym Sci B Polym Phys 2006; 44: 2201–2225.10.1002/polb.20861Search in Google Scholar

Devanathan R. Recent developments in proton exchange membranes for fuel cells. Energy Environ Sci 2008; 1: 101–119.10.1039/b808149mSearch in Google Scholar

Díaz M, Ortiz A, Ortiz I. Progress in the use of ionic liquids as electrolyte membranesin fuel cells. J Membr Sci 2014; 469: 379–396.10.1016/j.memsci.2014.06.033Search in Google Scholar

Dong B, Gwee L, Salas-de la Cruz D, Winey KI, Elabd YA. Super proton conductive high-purity Nafion nanofibers. Nano Lett 2010; 10: 3785–3790.10.1021/nl102581wSearch in Google Scholar

Drach Z, Hershkovitz S, Ferrero D, Leone P, Lanzini A, Santarelli M, Tsur Y. Impedance spectroscopy analysis inspired by evolutionary programming as a diagnostic tool for SOEC and SOFC. Solid State Ionics 2016; 288: 307–310.10.1016/j.ssi.2016.01.001Search in Google Scholar

Dupuis AC. Proton exchange membranes for fuel cells operated at medium temperatures: materials and experimental techniques. Prog Mater Sci 2011; 56: 289–327.10.1016/j.pmatsci.2010.11.001Search in Google Scholar

Dura JA, Murthi VS, Hartman M, Satija SK, Majkrzak CF. Multilamellar interface structures in Nafion. Macromolecules 2009; 42: 4769–4774.10.1021/ma802823jSearch in Google Scholar

Eisenberg A. Clustering of ions in organic polymers. A theoretical approach. Macromolecules 1970; 3: 147–154.10.1021/ma60014a006Search in Google Scholar

Elabd YA, Napadensky E, Sloan JM, Crawford DM, Walker CW. Diffusion and interfacial transport of water in Nafion. J Memb Sci 2003; 217: 227–242.10.1016/S0376-7388(03)00127-3Search in Google Scholar

Elliott JA, Wu D, Paddison SJ, Moore RB. A unified morphological description of Nafion membranes from SAXS and mesoscale simulations. Soft Matter 2011; 7: 6820–6827.10.1039/c1sm00002kSearch in Google Scholar

Fang Y, Leddy J. Surface diffusion in microstructured, ion-exchange matrices: Nafion/neutron track-etched polycarbonate membrane composites. J Phys Chem 1995; 99: 6064–6073.10.1021/j100016a049Search in Google Scholar

Franco AA, Schott P, Jallut C, Maschke B. A multi-scale dynamic mechanistic model for the transient analysis of PEFCs. Fuel Cells 2007; 7: 99–117.10.1002/fuce.200500204Search in Google Scholar

Freger V, Korin E, Wisniak J, Korngold E. Preferential sorption in ion-exchange pervaporation membranes: sorption of water-ethanol mixture by sodium polyethylene sulphonate. J Membr Sci 1997a; 128: 151–162.10.1016/S0376-7388(96)00306-7Search in Google Scholar

Freger V, Korin E, Wisniak J, Korngold E. Transport mechanism in ion-exchange pervaporation membranes: dehydration of water-ethanol mixture by sodium polyethylene sulphonate membranes. J Membr Sci 1997b; 133: 255–267.10.1016/S0376-7388(97)00080-XSearch in Google Scholar

Freger V, Korin E, Wisniak J, Korngold E, Ise M, Kreuer KD. Diffusion of water and ethanol in ion-exchange membranes: limits of the geometric approach. J Membr Sci 1999; 160: 213–224.10.1016/S0376-7388(99)00084-8Search in Google Scholar

Gebel G. Structural evolution of water swollen perfluorosulfonated ionomers from dry membrane to solution. Polymer 2000; 41: 5829–5838.10.1016/S0032-3861(99)00770-3Search in Google Scholar

Geise GM. Water and salt transport structure/property relationships in polymer membranes for desalination and power generation applications’, Ph.D. Thesis, Austin-Texas, USA: The University of Texas at Austin, 2012.Search in Google Scholar

Geise GM, Paul DR, Freeman BD. Fundamental water and salt transport properties of polymeric materials. Prog Polym Sci 2014; 39: 1–42.10.1016/j.progpolymsci.2013.07.001Search in Google Scholar

Gierke TD. 152nd Meeting of the Electrochemical Society Extended Abstracts, Atlanta, GA, USA, Abstract No. 438, 1977.Search in Google Scholar

Gierke TD, Munn GE, Wilson FC. The morphology in Nafion perfluorinated membrane products as determined by wide- and small-angle x-ray studies. J Polym Sci 1981; 19: 1687–1704.10.1002/pol.1981.180191103Search in Google Scholar

Gloukhovski R, Freger V, Tsur Y. A novel composite Nafion/anodized aluminum oxide proton exchange membrane. Fuel Cells 2016; 16: 434–443.10.1002/fuce.201500166Search in Google Scholar

Gloukhovski R, Tsur Y, Freger VA. Nafion-filled polycarbonate track-etched composite membrane with enhanced selectivity for direct methanol fuel cells. Fuel Cells 2017; 17: 56–66.10.1002/fuce.201600154Search in Google Scholar

Goni-Urtiaga A, Presvytes D, Scott K. Solid acids as electrolyte materials for proton exchange membrane (PEM) electrolysis: review. Int J Hydrogen Energy 2012; 37: 3358–3372.10.1016/j.ijhydene.2011.09.152Search in Google Scholar

Guo W, Tang H, Sun M, Yang H, Pan M, Duan J. Physically stable proton exchange membrane with ordered electrolyte for elevated temperature PEM fuel cell. Int J Hydrogen Energy 2012; 37: 9782–9791.10.1016/j.ijhydene.2012.03.119Search in Google Scholar

Hara N, Ohashi H, Ito T, Yamaguchi T. Rapid proton conduction through unfreezable and bound water in a wholly aromatic pore-filling electrolyte membrane. J Phys Chem B 2009; 113: 4656–4663.10.1021/jp810575uSearch in Google Scholar

Hasani-Sadrabadi MM, Shabani I, Soleimani M, Moaddel H. Novel nanofiber-based triple-layer proton exchange membranes for fuel cell applications. J Power Sources 2011; 196: 4599–4603.10.1016/j.jpowsour.2011.01.002Search in Google Scholar

Haufe S, Stimming U. Proton conducting membranes based on electrolyte filled microporous matrices. J Membr Sci 2001; 185: 95–103.10.1016/S0376-7388(00)00637-2Search in Google Scholar

Hershkovitz S, Baltianski S, Tsur Y. Harnessing evolutionary programming for impedance spectroscopy analysis: a case study of mixed ionic-electronic conductors. Solid State Ionics 2011; 188: 104–109.10.1016/j.ssi.2010.10.004Search in Google Scholar

Hershkovitz S, Baltianski S, Tsur Y. Electrochemical impedance analysis of SOFC cathode reaction using evolutionary programming. Fuel Cells 2012; 12: 77–85.10.1002/fuce.201100126Search in Google Scholar

Hickner MA, Ghassemi H, Kim YS, Einsla BR, McGrath JE. Alternative polymer systems for proton exchange membranes (PEMs). Chem Rev 2004; 104: 4587–4612.10.1021/cr020711aSearch in Google Scholar

Hsu WY, Gierke TD. Ion transport and clustering in Nafion perfluorinated membranes. J Membr Sci 1983; 13: 307–326.10.1016/S0376-7388(00)81563-XSearch in Google Scholar

Hsu WY, Barkley JR, Meakin P. Ion percolation and insulator-to-conductor transition in Nafion perfluorosulfonic acid membranes. Macromolecules 1980; 13: 198–200.10.1021/ma60073a041Search in Google Scholar

Jiang S, Xia KQ, Xu G. Effect of additives on self-assembling behavior of Nafion in aqueous media. Macromolecules 2001; 34: 7783–7788.10.1021/ma010124bSearch in Google Scholar

Jiao K, Li X. Water transport in polymer electrolyte membrane fuel cells. Prog Energy Combustion Sci 2011; 37: 221–291.10.1016/j.pecs.2010.06.002Search in Google Scholar

Kim DJ, Jo MJ, Nam SY. A review of polymer–nanocomposite electrolyte membranes for fuel cell application. J Ind Eng Chem 2015; 21: 36–52.10.1016/j.jiec.2014.04.030Search in Google Scholar

Kirkpatrick S. Percolation and conduction. Rev Mod Phys 1973; 45: 574–588.10.1103/RevModPhys.45.574Search in Google Scholar

Kraytsberg A, Ein-Eli Y. Review of advanced materials for proton exchange membrane fuel cells. Energy Fuels 2014; 28: 7303–7330.10.1021/ef501977kSearch in Google Scholar

Kreuer KD, Portale G. A critical revision of the nano-morphology of proton conducting ionomers and polyelectrolytes for fuel cell applications. Adv Funct Mater 2013; 23: 5390–5397.10.1002/adfm.201300376Search in Google Scholar

Ladewig BP, Knott RB, Hill AJ, Riches JD, White JW, Martin DJ, Dinizda Costa JC, Lu GQ. Physical and electrochemical characterization of nanocomposite membranes of Nafion and functionalized silicon oxide. Chem Mater 2007; 19: 2372–2381.10.1021/cm0628698Search in Google Scholar

Latham RA. Algorithm development for electrochemical impedance spectroscopy diagnostics in PEM fuel cells’, M.Sc. Thesis, University of Victoria, Saanich and Oak Bay, Greater Victoria, British Columbia, Canada, 2004.Search in Google Scholar

Leddy J, Vanderborgh NE. Microstructural enhancement of mass transport through Nafion+nucleopore composite membranes. Electroanal Chem 1987; 235: 299–315.10.1016/0022-0728(87)85215-4Search in Google Scholar

Lee SJ, Yu TL, Lin H-L, Liu W-H, Lai C-L. Solution properties of Nafion in methanol/water mixture solvent. Polymer 2004; 45: 2853–2862.10.1016/j.polymer.2004.01.076Search in Google Scholar

Lehmani A, Turq P, Perie M, Perie J, Simonin JP. Ion transport in Nation 117 membrane. J Electroanal Chem 1997; 428: 81–89.10.1016/S0022-0728(96)05060-7Search in Google Scholar

Li HY, Liu YL. Nafion-functionalized electrospun poly(vinylidene fluoride) (PVDF) nanofibers for high performance proton exchange membranes in fuel cells. J Mater Chem A 2014; 2: 3783–3793.10.1039/C3TA14264GSearch in Google Scholar

Li Q, He R, Jensen JO, Bjerrum NJ. Approaches and recent development of polymer electrolyte membranes for fuel cells operating above 100°C. Chem Mater 2003; 15: 4896–4915.10.1021/cm0310519Search in Google Scholar

Li HY, Lee YY, Lai JY, Liu YL. Composite membranes of Nafion and poly(styrene sulfonic acid)-grafted poly(vinylidene fluoride) electrospun nanofiber mats for fuel cells. J Membr Sci 2014; 466: 238–245.10.1016/j.memsci.2014.04.057Search in Google Scholar

Lin HL, Wang SH. Nafion/poly(vinylalcohol) nano-fiber composite and Nafion/poly(vinylalcohol) blend membranes for direct methanol fuel cells. J Membr Sci 2014; 452: 253–262.10.1016/j.memsci.2013.09.039Search in Google Scholar

Lin HL, Yu TL, Shen KS, Huang LN. Effect of Triton-X on the preparation of Nafion/PTFE composite membranes. J Membr Sci 2004; 237: 1–7.10.1016/j.memsci.2004.01.021Search in Google Scholar

Lin HL, Yu TL, Huang LN, Chena LC, Shen KS, Jung GB. Nafion/PTFE composite membranes for direct methanol fuel cell applications. J Power Sources 2005; 150; 11–19.10.1016/j.jpowsour.2005.02.016Search in Google Scholar

Lin J, Wu PH, Wycisk R, Trivisonno A, Pintauro PN. Direct methanol fuel cell operation with pre-stretched recast Nafion. J Power Sources 2008a; 183: 491–497.10.1016/j.jpowsour.2008.05.051Search in Google Scholar

Lin J, Wu PH, Wycisk R, Pintauro PN, Shi Z. Properties of water in prestretched recast Nafion. Macromolecules 2008b; 41: 4284–4289.10.1021/ma800194zSearch in Google Scholar

Lin HL, Wang SH, Chiu CK, Yu TL, Chena LC, Huang C-C, Cheng TH, Lin JM. Preparation of Nafion/poly(vinyl alcohol) electro-spun fiber composite membranes for direct methanol fuel cells. J Membr Sci 2010; 365: 114–122.10.1016/j.memsci.2010.08.045Search in Google Scholar

Liu F, Yi B, Xing D, Yu J, Zhang H. Nafion/PTFE composite membranes for fuel cell applications. J Membr Sci 2003; 212: 213–223.10.1016/S0376-7388(02)00503-3Search in Google Scholar

Macdonald JR. Impedance spectroscopy. Ann Biomed Eng 1992; 20: 289–305.10.1016/B0-12-227410-5/00333-1Search in Google Scholar

Malinauskas A, Malinauskien J, Ramanavicius A. Conducting polymer-based nanostructurized materials: electrochemical aspects. Nanotechnology 2005; 16: R51–R62.10.1088/0957-4484/16/10/R01Search in Google Scholar PubMed

Mauritz KA, Moore RB. State of Understanding of Nafion. Chem Rev 2004; 104: 4535–4585.10.1021/cr0207123Search in Google Scholar

Mekhilef S, Saidurb R, Safari A. Comparative study of different fuel cell technologies. Renewable and Sustainable Energy Reviews 2012; 16: 981–989.10.1016/j.rser.2011.09.020Search in Google Scholar

Moghaddam S, Pengwang E, Jiang Y-B, Garcia AC, Burnett DJ, Binker CJ, Masel RI, Shannon MA. An inorganic–organic proton exchange membrane for fuel cells with a controlled nanoscale pore structure. Nat Nanotechnol 2010; 5: 230–236.10.1038/nnano.2010.13Search in Google Scholar

Molla S, Compan V. Polyvinyl alcohol nanofiber reinforced Nafion membranes for fuel cell applications. J Membr Sci 2011; 372: 191–200.10.1016/j.memsci.2011.02.001Search in Google Scholar

Molla S, Compan V. Nanocomposite SPEEK-based membranes for direct methanol fuel cells at intermediate temperatures. J Membr Sci 2015; 492: 123–136.10.1016/j.memsci.2015.05.055Search in Google Scholar

Molla S, Compan V, Gimene E, Blazquez A, Urdanpilleta I. Novel ultrathin composite membranes of Nafion/PVA for PEMFCs. Int J Hydrogen Energy 2011; 36: 9886–9895.10.1016/j.ijhydene.2011.05.074Search in Google Scholar

Narebska A, Koter S, Kujavski W. Ions and water transport across charged Nafion membranes. Irreversible thermodynamics approach. Desalination 1984; 51: 3–17.10.1016/0011-9164(84)85048-1Search in Google Scholar

Navarro A, Del Rio C, Acosta JL. Pore-filling electrolyte membranes based on microporous polyethylene matrices activated with plasma and sulfonated hydrogenated styrene butadiene block copolymer. Synthesis, microstructural and electrical characterization. J Polym Sci B Polym Phys 2008; 46: 1684–1695.10.1002/polb.21505Search in Google Scholar

Neburchilov V, Martin J, Wang H, Zhang J. A review of polymer electrolyte membranes for direct methanol fuel cells. J Power Sources 2007; 169: 221–238.10.1016/j.jpowsour.2007.03.044Search in Google Scholar

Nguyen T, Wang X. Multifunctional composite membrane based on a highly porous polyimide matrix for direct methanol fuel cells. J Power Sources 2010: 195: 1024–1030.10.1016/j.jpowsour.2009.08.049Search in Google Scholar

Nguyen TH, Wang C, Wang X. Pore-filling membrane for direct methanol fuel cells based on sulfonated poly(styrene-ran-ethylene) and porous polyimide matrix. J Membr Sci 2009; 342: 208–214.10.1016/j.memsci.2009.06.042Search in Google Scholar

Nouel KM, Fedkiw PS. Nafion-based composite polymer electrolyte membranes. Electrochim Acta 1998; 43: 2381–2387.10.1016/S0013-4686(97)10151-7Search in Google Scholar

Ogawa T, Kamiguchi K, Tamaki T, Imai H, Yamaguchi T. Differentiating Grotthuss proton conduction mechanisms by nuclear magnetic resonance spectroscopic analysis of frozen samples. Anal Chem 2014; 86: 9362–9366.10.1021/ac5021485Search in Google Scholar PubMed

Oleksandrov S, Lee J-W, Jang J-H, Haam S, Chung C-H. Proton-conductive nanochannel membrane for fuel-cell applications. J Nanosci Nanotechnol 2009; 9: 1551–1554.10.1166/jnn.2009.C200Search in Google Scholar

Oz A, Hershkovitz S, Belman N, Tal-Gutelmacher E, Tsur Y. Analysis of impedance spectroscopy of aqueous supercapacitors by evolutionary programming: finding DFRT from complex capacitance. Solid State Ionics 2016; 288: 311–314.10.1016/j.ssi.2015.11.008Search in Google Scholar

Oz A, Gelman D, Goren E, Shomrat N, Baltianski S, Tsur Y. A novel approach for supercapacitors degradation characterization. J Power Sources 2017; 355: 74–82.10.1016/j.jpowsour.2017.04.048Search in Google Scholar

Park MJ, Downing KH, Jackson A, Gomez ED, Minor AW, Cookson D, Weber AZ, Balsara NP. Increased water retention in polymer electrolyte membranes at elevated temperatures assisted by capillary condensation. Nano Lett 2007; 7: 3547–3552.10.1021/nl072617lSearch in Google Scholar PubMed

Park CH, Lee CH, Guiver MD, Lee MY. Sulfonated hydrocarbon membranes for medium-temperature and low-humidity proton exchange membrane fuel cells (PEMFCs). Prog Polym Sci 2011a; 36: 1443–1498.10.1016/j.progpolymsci.2011.06.001Search in Google Scholar

Park JK, Li J, Divoux GM, Madsen LA, Moore RB. Oriented morphology and anisotropic transport in uniaxially stretched perfluorosulfonate ionomer membranes. Macromolecules 2011b; 44: 5701–5710.10.1021/ma200865pSearch in Google Scholar

Patil MB, Bhagat SL, Sapkal RS, Sapkal VS. A review on the fuel cells development. Sci Revs Chem Commun 2011; 1: 25–41.Search in Google Scholar

Peighambardoust SJ, Rowshanzamir S, Amjadi M. Review of the proton exchange membranes for fuel cell applications. Intern J Hydrogen Energy 2010; 35; 9349–9384.10.1016/j.ijhydene.2010.05.017Search in Google Scholar

Penner RM, Martin CR. Ion transporting composite membranes I: Nafion-impregnated Gore-Tex. J Electrochem Soc 1985; 132: 514–515.10.1149/1.2113875Search in Google Scholar

Raghav S. Polymer-filled nanoporous membranes’, M. Sc. Thesis, Philadelphia-CA, USA: Drexel University, 2005.Search in Google Scholar

Ramya K, Velayutham G, Subramaniam CK, Rajalakshmi N, Dhathathreyan KS. Effect of solvents on the characteristics of Nafion/PTFE composite membranes for fuel cell applications. J Power Sources 2006; 160: 10–17.10.1016/j.jpowsour.2005.12.082Search in Google Scholar

Rayon E, Gonzalez V, Matveeva E, Parkhutik V. Pore filling membrane based on porous anodic alumina and Nafion. Porous Semiconductors Science and Technology Conference PSST-2006, At Sitges, Spain, 2006.Search in Google Scholar

Reshetenko T, Kulikovsky A. Impedance spectroscopy study of the PEM fuel cell cathode with nonuniform nafion loading. J Electrochem Soc 2017; 164: E3016–E3021.10.1149/2.0041711jesSearch in Google Scholar

Rodgers MP, Berring J, Holdcroft S, Shi Z. The effect of spatial confinement of Nafion in porous membranes on macroscopic properties of the membrane. J Membr Sci 2008; 321: 100–113.10.1016/j.memsci.2008.01.007Search in Google Scholar

Sahu AK, Pitchumani S, Sridar P, Shukla AK. Nafion and modified-Nafion membranes for polymer electrolyte fuel cells: an overview. Bull Mater Sci 2009; 32: 285–294.10.1007/s12034-009-0042-8Search in Google Scholar

Savadogo O. Emerging membranes for electrochemical systems: (I) Solid polymer electrolyte membranes for fuel cell systems. J New Mater Electrochem Syst 1998; 1: 47–66.10.1002/chin.199847334Search in Google Scholar

Savadogo O. Emerging membranes for electrochemical systems Part II. High temperature composite membranes for polymer electrolyte fuel cell (PEFC) applications. J Power Sources 2004; 127: 135–161.10.1016/j.jpowsour.2003.09.043Search in Google Scholar

Schlick S. Ionomers: characterization, theory, and applications. Boca Raton, FL: CRC Press – Science, 1996, pp. 84.Search in Google Scholar

Schmidt-Rohr K, Chen Q. Parallel cylindrical water nanochannels in Nafion fuel-cell membranes. Nat Mater 2008; 7: 75–83.10.1142/9789814317665_0033Search in Google Scholar

Scribner Associates. Tutorial, http://www.scribner.com/files/tech-papers/Scribner%20Associates%20-%20Electrochemical%20Impedance%20Spectroscopy%20for%20Fuel%20Cell%20Research.pdf.Search in Google Scholar

Shabani I, Hasani-Sadrabadi MM, Haddadi-Asl V, Soleimani M. Nanofiber-based polyelectrolytes as novel membranes for fuel cell applications. J Membr Sci 2011; 368: 233–240.10.1016/j.memsci.2010.11.048Search in Google Scholar

Shin J-P, Chang B-J, Kim J-H, Lee S-B, Suh DH. Sulfonated polystyrene/PTFE composite membranes. J Membr Sci 2005; 251: 247–254.10.1016/j.memsci.2004.09.050Search in Google Scholar

Silva RF, De Francesco M, Pozio A. Tangential and normal conductivities of Nafion membranes used in polymer electrolyte fuel cells. J Power Sources 2004; 134: 18–26.10.1016/j.jpowsour.2004.03.028Search in Google Scholar

Slade S, Campbell SA, Ralph TR, Walsh FC. Ionic conductivity of an extruded Nafion 1100 EW series of membranes. J Electrochem Soc 2002; 149: A1556–A1564.10.1149/1.1517281Search in Google Scholar

Smith DW, Oladoyinbo FO, Mortimore WA, Colquhoun HM, Thomassen MS, Odegard A, Guillet N, Mayousse E, Klicpera T, Hayes W. A microblock ionomer in proton exchange membrane electrolysis for the production of high purity hydrogen. Macromolecules 2013; 46: 1504–1511.10.1021/ma3026145Search in Google Scholar

Smitha B, Sridhar S, Khan AA. Solid polymer electrolyte membranes for fuel cell applications – a review. J Membr Sci 2005; 259: 10–26.10.1016/j.memsci.2005.01.035Search in Google Scholar

Springer TE, Zawodzinski TA, Wilson MS, Gottesfeld S. Characterization of polymer electrolyte fuel cells using AC impedance spectroscopy. J Electrochem Soc 1996; 143: 587–599.10.1149/1.1836485Search in Google Scholar

Steele BC, Heinzel A. Materials for fuel-cell technologies. Nature 2001; 414: 345–352.10.1142/9789814317665_0031Search in Google Scholar

Subianto S, Pica M, Casciola M, Cojocaru P, Merlo L, Hards G, Jones DJ. Physical and chemical modification routes leading to improved mechanical properties of perfluorosulfonic acid membranes for PEM fuel cells. J Power Sources 2013; 233: 216–230.10.1016/j.jpowsour.2012.12.121Search in Google Scholar

Tang H, Pan M, Jiang SP, Wang X, Ruan Y. Fabrication and characterization of PFSI/ePTFE composite proton exchange membranes of polymer electrolyte fuel cells. Electrochim Acta 2007a; 52: 5304–5311.10.1016/j.electacta.2007.02.031Search in Google Scholar

Tang H, Pan M, Wang F, Shen PK, Jiang SP. Highly durable proton exchange membranes for low temperature fuel cells. J Phys Chem B Lett 2007b; 111: 8684–8690.10.1021/jp073136tSearch in Google Scholar PubMed

Tang H, Wang X, Pan M, Wang F. Fabrication and characterization of improved PFSA/ePTFE composite polymer electrolyte membranes. J Membr Sci 2007c; 306: 298–306.10.1016/j.memsci.2007.09.008Search in Google Scholar

Tang H, Pan M, Zhaohui W. Improvement of the Nafion–polytetrafluoroethylene membranes for potential direct methanol fuel cell use by reduction of the methanol crossover. J Appl Polym Sci 2008; 110: 2227–2233.10.1002/app.28438Search in Google Scholar

Teng X, Dai J, Su J, Zhu Y, Liu H, Song Z. A high performance polytetrafluoroethene/Nafion composite membrane for vanadium redox flow battery application. J Power Sources 2013; 240: 131–139.10.1016/j.jpowsour.2013.03.177Search in Google Scholar

The Fuel Cell Industry Review. http://www.fuelcellindustryreview.com/FCIRarchive.html, 2015.Search in Google Scholar

Thiam HS, Daud WRW, Kamarudin SK, Mohammad AB, Kadhum AAH, Loh KS, Majlan EH. Overview on nanostructured membrane in fuel cell Applications. Int J Hydrogen Energy 2011; 36: 3187–3205.10.1016/j.ijhydene.2010.11.062Search in Google Scholar

Tian H. New composite membranes based on modified Nafion or Flemion for PEM fuel cell application, Ph.D. Thesis, Montreal, Quebec, Canada: Université de Montréal, 2004.Search in Google Scholar

Tripathi BP, Shahi VK. Organic–inorganic nanocomposite polymer electrolyte membranes for fuel cell applications. Prog Polym Sci 2011; 36: 945–979.10.1016/j.progpolymsci.2010.12.005Search in Google Scholar

Truffier-Boutry D, De Geyer A, Guetaz L, Diat O, Gebel G. Structural study of zirconium phosphate-nafion hybrid membranes for high-temperature proton exchange membrane fuel cell applications. Macromolecules 2007; 40: 8259–8264.10.1021/ma0706576Search in Google Scholar

van der Heijden PC, Rubatat L, Diat O. Orientation of drawn Nafion at molecular and mesoscopic scales. Macromolecules 2004; 37: 5327–5336.10.1021/ma035642wSearch in Google Scholar

Viswanathan B, Helen M. Is Nafion the only choice?’ Bull Catal Soc India 2007; 6: 50–56.Search in Google Scholar

Vorrey S, Teeters D. Study of the ion conduction of polymer electrolytes confined in micro and nanopores. Electrochim Acta 2003; 48: 2137–2141.10.1016/S0013-4686(03)00196-8Search in Google Scholar

Wang SH, Lin HL. Poly (vinylidene fluoride-co-hexafluoropropylene)/polybenzimidazole blend nanofiber supported Nafion membranes for direct methanol fuel cells. J Power Sources 2014; 257: 254–263.10.1016/j.jpowsour.2014.01.104Search in Google Scholar

Wang X, Tang H, Pan M. Durable perfluorosulfonic polymer electrolyte membranes prepared from alkaline-ion- assisted heat treatment. J Membr Sci 2011; 379: 106–111.10.1016/j.memsci.2011.05.058Search in Google Scholar

Wang T, Sun F, Wang H, Yang S, Fan L. Preparation and properties of pore-filling membranes based on sulfonated copolyimides and porous polyimide matrix. Polymer 2012; 53: 3154–3162.10.1016/j.polymer.2012.05.049Search in Google Scholar

Wang Z, Tang H, Li J, Jin A, Wang Z, Zhang H, Pan M. Balancing dimensional stability and performance of proton exchange membrane using hydrophilic nanofibers as the supports. Int J Hydrogen Energy 2013; 38: 4725–4733.10.1016/j.ijhydene.2013.01.176Search in Google Scholar

Wei W, Zhang H, Li X, Mai Z, Zhang H. Poly(tetrafluoroethylene) reinforced sulfonated poly(ether ether ketone) membranes for vanadium redox flow battery application. J Power Sources 2012; 208: 421–425.10.1016/j.jpowsour.2012.02.047Search in Google Scholar

White S, Granados-Focil S, Woudenberg RC, Yavuzcetin O, Tuominen M, Coughlin EB. Proton conducting nano-composite membrane for high-temperature polymer electrolyte membrane fuel cells. Polym Prepr 2008; 49: 1145–1146.Search in Google Scholar

Wu PH. Prestretched recast Nafion for direct methanol fuel cells, M.Sc. Thesis, Cleveland, OH, USA: Case Western Reserve University, 2008.Search in Google Scholar

Wu X, Wang X, He G Benziger J. Differences in water sorption and proton conductivity between Nafion and SPEEK. J Polym Sci B Polym Phys 2011; 49: 1437–1445.10.1002/polb.22326Search in Google Scholar

Wycisk R, Pintauro PN, Park JW. New developments in proton conducting membranes for fuel cells. Curr Opin Chem Eng 2014; 4: 71–78.10.1016/j.coche.2014.01.012Search in Google Scholar

Xiao P, Li J, Tang H, Wang Z, Pan M. Physically stable and high-performance Aquivion/ePTFE composite membrane for high-temperature fuel cell application. J Membr Sci 2013; 442: 65–71.10.1016/j.memsci.2013.04.014Search in Google Scholar

Xing D, Yi BC, Liu FQ, Fu YZ, Zhang HM. Characterization of sulfonated poly(ether ether ketone)/polytetrafluoroethylene composite membranes for fuel cell applications. Fuel Cells 2005; 5: 406–411.10.1002/fuce.200500089Search in Google Scholar

Xing D, He G, Hou Z, Ming P, Song S. Properties and morphology of Nafion/polytetrafluoroethylene composite membrane fabricated by a solution-spray process. Int J Hydrogen Energy 2013; 38: 8400–8408.10.1016/j.ijhydene.2013.04.084Search in Google Scholar

Xu X, Li L, Wang H, Li X, Zhuang X. Solution blown sulfonated poly(ether ether ketone) nanofiber–Nafion composite membrane. RSC Adv 2015; 5: 4934–4940.10.1039/C4RA10898ASearch in Google Scholar

Yamaguchi T, Nakao S, Kimura S. Plasma-graft filling polymerization: preparation of a new type of pervaporation membrane for organic liquid mixtures. Macromolecules 1991; 24: 5522–5527.10.1021/ma00020a006Search in Google Scholar

Yamaguchi T, Miyata F, Nakao S. Polymer electrolyte membranes with a pore-filling structure for a direct methanol fuel cell. Adv Mater 2003; 15: 1198–1201.10.1002/adma.200304926Search in Google Scholar

Yamaguchi T, Kuroki H, Miyata F. DMFC performances using a pore-filling polymer electrolyte membrane for portable usages. Electrochem Commun 2005; 7: 730–734.10.1016/j.elecom.2005.04.030Search in Google Scholar

Yamaguchi T, Zhou H, Nakazawa S, Hara N. An extremely low methanol crossover and highly durable aromatic pore-filling electrolyte membrane for direct methanol fuel cells. Adv Mater 2007; 19: 592–596.10.1002/adma.200601086Search in Google Scholar

Yang L, Li H, Ai F, Chen X, Tang J, Zhu Y, Wang C, Yuan WZ, Zhang Y. A new method to prepare high performance perfluorinated sulfonic acid ionomer/porous expanded polytetrafluoroethylene composite membranes based on perfluorinated sulfonyl fluoride polymer solution. J Power Sources 2013; 243: 392–395.10.1016/j.jpowsour.2013.06.056Search in Google Scholar

Yildirim MH, Stamatialis D, Wessling M. Dimensionally stable Nafion–polyethylene composite membranes for direct methanol fuel cell applications. J Membr Sci 2008; 321: 364–372.10.1016/j.memsci.2008.05.013Search in Google Scholar

Yildirim MH, Schwarz A, Stamatialis DF, Wessling M. Impregnated membranes for direct methanol fuel cells at high methanol concentrations. J Membr Sci 2009; 328: 127–133.10.1016/j.memsci.2008.11.051Search in Google Scholar

Yu TL, Lin H-L, Shen K-S, Huang L-N, Chang Y-C, Jung G-B, Huang JC. Nafion/PTFE composite membranes for fuel cell applications. J Polym Res 2004; 11: 217–224.10.1023/B:JPOL.0000043408.24885.c6Search in Google Scholar

Yu DM, Yoon S, Kim T-H, Lee JY, Lee J, Hong YT. Properties of sulfonated poly(arylene ethersul fone)/electrospun nonwoven polyacrylonitrile composite membrane for proton exchange membrane fuel cells. J Membr Sci 2013; 446: 212–219.10.1016/j.memsci.2013.06.028Search in Google Scholar

Zhang L. Fabrication of novel nanomaterials for PEMFCs and self-cleaning applications’, Ph.D. Thesis, Chapel Hill, North Carolina, USA: The University of North Carolina at Chapel Hill, 2006.Search in Google Scholar

Zhang H, Shen PK. Advances in the high performance polymer electrolyte membranes for fuel cells. Chem Soc Rev 2012a; 41: 2382–2394.10.1039/c2cs15269jSearch in Google Scholar PubMed

Zhang H, Shen PK. Recent development of polymer electrolyte membranes for fuel cells. Chem Rev 2012b; 112: 2780–2832.10.1021/cr200035sSearch in Google Scholar PubMed

Zhang H, Pan J, He X, Pan M. Zeta potential of nafion molecules in isopropanol-water mixture solvent. J Appl Polym Sci 2008; 107: 3306–3309.10.1002/app.27473Search in Google Scholar

Zhang Y, Cui Z, Liu C, Xing W, Zhang J. Implantation of Nafion ionomer into polyvinyl alcohol/chitosan composites to form novel proton-conducting membranes for direct methanol fuel cells. J Power Sources 2009; 194: 730–736.10.1016/j.jpowsour.2009.06.022Search in Google Scholar

Zhao Q, Majsztrik P, Benziger J. Diffusion and interfacial transport of water in Nafion. J Phys Chem B 2011; 115: 2717–2727.10.1021/jp1112125Search in Google Scholar PubMed

Zhao Q, Carro N, Ryu HY, Benziger J. Sorption and transport of methanol and ethanol in H+-Nafion. Polymer 2012; 53: 1267–1276.10.1016/j.polymer.2012.01.050Search in Google Scholar

Zhengbang W, Haolin T, Junsheng L, Mu P. Morphology change of biaxially oriented polytetrafluoroethylene membranes caused by solvent soakage. J Appl Polym Sci 2011; 121: 1464–1468.10.1002/app.33670Search in Google Scholar

Zhu X, Zhang H, Liang Y, Zhang Y, Luo O, Bi C, Yi B. Challenging reinforced composite polymer electrolyte membranes based on disulfonated poly(arylene ether sulfone)-impregnated expanded PTFE for fuel cell applications. J Mater Chem 2007; 17: 386–397.10.1039/B611690FSearch in Google Scholar

Zook LA, Leddy J. Density and solubility of Nafion: recast, annealed, and commercial films. Anal Chem 1996; 68: 3793–3796.10.1021/ac960604eSearch in Google Scholar PubMed

Zoulias E, Varkaraki E, Lymberopoulos N, Christodoulou CN, Karagiorgis CN. A review on water electrolysis. TCJST 2004; 4: 41–58.Search in Google Scholar

Received: 2016-12-15
Accepted: 2017-06-04
Published Online: 2017-09-13
Published in Print: 2018-07-26

©2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 1.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/revce-2016-0065/html?lang=en
Scroll to top button