You are on page 1of 12

WWW.POLYMERPHYSICS.

ORG REVIEW

Water-Mediated Transport in Ion-Containing Polymers

Michael A. Hickner
Department of Materials Science and Engineering, The Pennsylvania State University, University Park, Pennsylvania 16802
Correspondence to: M. A. Hickner (E-mail: hickner@matse.psu.edu)

Received 3 October 2011; accepted 3 October 2011; published online 17 October 2011
DOI: 10.1002/polb.22381

ABSTRACT: Water-mediated ion conduction enables high con- polymer associations can be exploited to tune the transport
ductivity in hydrated polymer membranes commonly used in and mechanical property tradeoffs in these polymers. Measure-
electrochemical devices. Understanding the coupling of the ments of water motion provide important criteria for assessing
absorbed water with the polymer matrix and the dynamics of the factors that control the performance of these types of mate-
water inside the polymer network across the full range of rials. This review article discusses current understanding
length scales in the membrane is important for unraveling the of water behavior in ion-containing polymers and the rela-
structure–property relationships in these materials. By consid- tionship between water motion and ion and molecular
ering the water behavior in ion-containing polymers, next-gen- transport. VC 2011 Wiley Periodicals, Inc. J Polym Sci Part B:

eration fuel cell membranes are being designed that exceed Polym Phys 50: 9–20, 2012
the conductivity of the state-of-the-art materials and have opti-
mized conductivity and permeability that may be useful in KEYWORDS: charge transport; diffusion; hydrogen bonding; ion-
other types of devices such as redox flow batteries. Water– omers; morphology

INTRODUCTION Polymer membranes are important func- cient of the ions that give the conductivity response. For pro-
tional components of low-temperature fuel cells and electro- ton exchange membranes (PEMs), which are single ion
lyzers. The scientific and engineering push for alternative conductors, the proton concentration in the membrane can be
energy conversion and storage technologies has intensified expressed as cHþ. In most cases, the activity of the protons or
efforts to develop new ion-containing membranes that dis- the concentration of free protons in the membrane is not
play high conductivity (101 S cm1 or greater) with low or known;1 therefore, the analytical proton concentration in the
no water absorption. However, water is an excellent medium membrane, as determined by titration of the sulfonate groups,
for fast ion conduction; thus, the most successful approaches is usually sufficient as an approximation to cHþ to estimate the
to new materials that display ionic conductivities high effective proton diffusion coefficient. For a membrane with a
enough for application in devices have water uptakes of usu- proton conductivity of 101 S cm1 and a proton concentra-
ally 10 wt % or greater. The water behavior in ion-conduct- tion of about 1 M, the effective proton diffusion coefficient is
ing membranes underpins their transport properties, espe- on the order of 3  105 cm2 s1, which can be compared to a
cially in tradeoff relationships between ion conductivity and proton diffusion coefficient of 104 cm2 s1 in liquid water.
molecular or ion permeability and permselectivity. In many The diffusion or mobility of protons in these materials is not
cases, if higher conductivity is desired, more water must be quite as high as in bulk water, but most measurements of the
absorbed into the membrane which promotes larger molecu- water behavior and ion and small molecule diffusion coeffi-
lar and ion diffusion coefficients. cients are indicative of liquid-like dynamics.2

In these types of materials, the functionalized polymer intro- The polymer, in essence, provides a framework for the
duces excess ions into the water phase in the membrane. absorbed water. Thus, understanding the water sorption
The concentration of ions in the membrane contributes to properties of these materials and studying the water–poly-
the conductivity (ri) by: mer interactions provides basic mechanistic information on
the behavior of the membranes. The fundamental studies of
ci Di z2i F 2 water motion in these types of polymers help to rationalize
ri ¼ (1) the transport properties of materials with new polymer com-
RT
positions or ionic domain structures. This understanding can
where ci is the concentration of ions that contribute to the then be used to boost the performance of next-generation
conductivity, zi is their charge, and Di is the diffusion coeffi- materials for application in devices. Instead of designing and

V
C 2011 Wiley Periodicals, Inc.

WWW.MATERIALSVIEWS.COM JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20 9
REVIEW WWW.POLYMERPHYSICS.ORG

Michael Hickner is the Virginia S. and Philip L. Walker, Jr. Faculty Fellow and an Assistant
Professor of Materials Science and Engineering at The Pennsylvania State University.
Hickner’s work has been recognized by Office of Naval Research and Army Research Office
Young Investigator Awards and a Presidential Early Career Award in Science and
Engineering (PECASE). He has co-authored five US and international patents and over 70
peer-reviewed publications with more than 3200 citations.

optimizing materials toward a nebulous goal of ‘‘higher con- influence the material’s transport properties. Different con-
ductivity,’’ considering the interplay between polymer chemi- ceptual frameworks have been used to describe the water–
cal structure, adsorbed water, and ion transport may give polymer associations in a range of hydrophilic polymers.
researchers in this area an additional tool for predictive Terms such as ‘‘state of water’’ or ‘‘free, loosely bound, and
design of these systems. tightly bound’’ help to describe the distribution of physical
and chemical environments that water experiences in a
The archetypal ion-containing polymer used as a PEM in fuel
hydrophilic polymer network. In practice, an observable
cells and other types of electrochemical devices is NAFIONV
R

property of the water such as its T1 relaxation time, self-dif-


(a product of DuPont) and its related structures, which are
fusion coefficient, or vibrational frequency of a characteristic
generally classified as poly(perfluorosulfonic acid)s (PFSAs).
mode is needed to quantify how water is held within the
These perfluorinated polymers have tethered sulfonic acid
polymer and in many cases the water inside a polymer mem-
groups that render the materials hydrophilic and increase
brane shows different signatures than bulk water.
the proton concentration in the membrane to achieve high
conductivity. Much research has been devoted to exploring The preceding discussion of morphology and water dynamics
the details of the arrangement of hydrophilic and hydropho- in water-absorbing, ion-containing polymers is decidedly
bic components of PFSAs in the solid state, particularly in separate from polymer-mediated ion transport, for example,
the context of deciphering their proton conduction proper- in Li-ion battery polymer electrolytes, where the ion motion
ties. This ionic domain morphology is thought to give rise to is coupled tightly to the polymer segmental dynamics.12
NAFIONV’s high ionic conductivity, high water permeability,
R
Water as the medium of ion transport leads to thermally
and other interesting properties. The nanophase morphologi- activated transport where the temperature-dependent con-
cal structure in fuel cell membranes is indeed an important ductivity can be described by an Arrhenius-type activation
factor in their performance as the ionic nanophases provide energy. For polymer-mediated ion conductors, the conductiv-
the ‘‘pipes’’ through which water and ions move. However, ity is often described by a Vogel-Fulcher-Tammann depend-
the connectivity and size of these ionic domains does not ence on temperature. In many cases, the conductivity of
fully explain the ion conduction behavior of these materials.
Many of the new classes of ion-containing polymers designed
to supplant PFSAs are based on aromatic backbones. There
are distinct differences in the ionic domain structure of
sulfonated poly(aromatic) membranes compared to PFSAs.
Foremost, the ability of the ions to cluster in the stiffer-
chained poly(aromatic) membranes is poor which leads to
smaller ionic domains than what are observed in PFSAs, Fig-
ure 1. This domain structure and the chemical properties of
the polymer have a direct influence on the absorbed water
mobility as evidenced by the self and Fickian diffusion coeffi-
cients of water and the deuterium T1 relaxation times meas-
ured as a function of hydration for each type of material.3,5,6
There have been many studies of NAFIONV’s ionic domain
R

structure,7–10 and comparisons to aromatic polymers have


revealed some of the structural origins of the differences in
conductivity between these two classes of ion-conducting
polymers.11 However, there is no direct ionic domain struc-
ture to conductivity relationship when materials with differ-
ent polymer backbones are considered. Interesting insights
to explain the differences in transport properties between
PFSAs and aromatic polymers have arisen from exploring FIGURE 1 Backbone chain schematics, transmission electron
how the water rotational and translational motion, or in micrographs, ion concentrations, and water self-diffusion co-
short, the water binding and dynamics, in the membrane efficients for PFSA and sulfonated aromatic PEMs.3,4

10 JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20


WWW.POLYMERPHYSICS.ORG REVIEW

sion coefficients. However, when sufficient hydration is present,


the proton diffusion coefficient is much greater than diffusion
coefficient of the surrounding water medium and proton ‘‘hop-
ping’’ is evident. This high hydration case in a NAFIONV mem-
R

brane is similar to what is observed in bulk liquid water where


the excess proton diffusion coefficient exceeds the self-diffusion
coefficient of water by factor of 5.11
As water is added to a membrane, the signatures of water
rotational and translational mobility increase. Measurements
of the 2H T22
1 and the self-diffusion coefficient of water by
PFG-NMR11,20 show that water is more mobile for higher
water to acid group ratios. The water:acid molar ratio is
termed the hydration number (k or wo). It is tempting to
invoke hopping-type transport when the rotational dynamics
of water are slow. However, the above experimental meas-
FIGURE 2 Proton (n) and water (h) diffusion coefficients as a urements demonstrate that fast water rotation is required to
function of hydration number. Adapted from Ref. 20. achieve hopping. Otherwise, at low hydration numbers, the
hydronium or H2nþ1On (hydronium plus excess waters) must
diffuse through the water matrix and hopping is minimized.
polymer-based ion conductors can be scaled by the Tg of the Water plays a special role in the context of proton conduc-
material, which collapses data from polymers with different tion because of its amphotericity. The dual role of water as a
Tgs onto a characteristic curve.13 Deviations from Tg scaling
proton donor (as hydronium) and proton acceptor coupled
in polymer-mediated ion conductors can be due to increased
with its fast rotational dynamics imparts high ionic conduc-
free ion content,14 phase separation and morphology,15
tivities to water-absorbing polymers as long as the materials
or increased density of ion donors or acceptors.16 Similar
remain hydrated and there exists sufficient acid concentra-
observations have been made for the temperature depend-
tion to create a number of excess protons. Other amphoteric
ence of ionic liquids where the long-range molecular
molecules such as phosphoric acid23 and triazole/imidaz-
dynamics controls the temperature-dependent conductivity
ole16,24,25 conduct protons with a hopping-type mechanism,
response.17–19
but their larger molecular structures and thus lower rota-
The purpose of this review is to highlight the important tional dynamics and higher viscosities lead to significant
insights and tools that have facilitated our understanding of depressions in their proton conductivity compared to water.
ion transport in water-absorbing membrane and further the The fast dynamics of proton transport in water is contrasted
discussion of water–polymer interactions in materials that with the slower molecular dynamics of the polymer mole-
rely on the presence of water for high transport rates. The cules themselves. Paddison and Zawodzinski26 studied the
relationship between water mobility and proton conductivity side chain of NAFIONV and concluded that although the sul-
R

will be discussed. Different length-scale measurements for fonate group can freely rotate, the rest of the side chain is
water mobility will be described and sections are included rather stiff. Also, the perfluoroether side chain of NAFIONV
R

on the effect that water has on molecular transport proper- is hydrophobic and likely exists in a coiled configuration in
ties and mechanical properties. Finally, recommendations are
the hydrated polymer. Thus, the flexibility and molecular
given for how to take the fundamental knowledge developed
motion of the side chain likely plays a minor role in the pro-
for proton transport in fuel cell membranes and extend
ton conduction process. However, the topology and flexibility
these concepts to new types of ion-containing membranes
of the ion-containing polymer determines the ability of the
and applications.
ionic nanophase to self-assemble into ordered domains. The
difficulty in deconvoluting the morphology is in part due to
WATER MOTION AND PROTON CONDUCTIVITY the subtle changes in morphology that can occur with even
minor variations in polymer chemical structure.
Seminal studies on PEM fuel cell (PEMFC) materials have
revealed the basic phenomena that are important to consider In many applications, ions must move through a 10–150-lm-
when dealing with proton conductivity in water-absorbing sys- thick membrane, which is at least three orders of magnitude
tems. The diffusion coefficient of protons, calculated from the greater than the 1–10-nm-size ionic domains observed for
Nernst–Einstein relationship and measured conductivity values, many systems. Clearly, ionic conduction across a membrane
is greater than the self-diffusion coefficient of water, deter- is a multilength-scale problem from the atomic-scale hydro-
mined by pulsed-field gradient nuclear magnetic resonance gen bonding dynamics between water molecules, to the ionic
(PFG-NMR) self-diffusion coefficient measurements, at high lev- domains that may contain thousands of water molecules, to
els of hydration (Fig. 2).20,21 When minimally hydrated, less the micron length scales that the protons must traverse.
than three to four water molecules per excess proton, the water Thus, a variety of techniques are required to probe the water
and protons move in concert as evidenced by their similar diffu- motion over a range of length and time scales to more

WWW.MATERIALSVIEWS.COM JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20 11
REVIEW WWW.POLYMERPHYSICS.ORG

the character of each type of water or the population of envi-


ronments occupied by water changes as the membrane is
hydrated. This assertion is supported by poor single expo-
nential or two component fits (bulk water and k ¼ 1 water)
to the relaxation data and shows that the distribution of
water across multiple hydrogen bonding environments
changes with hydration. These changes in the water–polymer
interactions could be due to the character of the water–
NAFIONV contacts at the hydrophilic/hydrophobic interface
R

evolving as a function of hydration.


NMR relaxation measurements such as T1 probe the local
rotational mobility of water within a sample. 2H T1 values
between 5 and 200 ms have been measured in a range of
PEMs imbibed with 2H2O, which correlate to water rota-
tional rate constants on the order of 107 to 1010 s1, simi-
lar to values found in supercooled liquid water.35,36 These
FIGURE 3 Water self-diffusion coefficient for NAFIONV R (~),
values compare to an 2H T1 of 400 ms for ambient pressure
random poly(aromatic) (h), and block poly(aromatic) (n) PEMs. liquid 2H2O at 292 K with a rotational rate constant of 2.8
Adapted from Ref. 39.  1011 s1.37 2H T1 measurements show a strong correla-
tion to the hydration number and conductivity for
NAFIONV.22 Lee et al.5 showed that sulfonated aromatic
R
completely understand the role that water plays in mediating
ion transport in these materials. polymers showed lower 2H T1 values than those observed
in PFSAs at equivalent hydration numbers. For instance at a
Vibrational spectroscopy has been used to probe the hydro-
k ¼ 5 for NAFIONV 117 the 2H T1 was 80 ms and for a sul-
R

gen bond dynamics of water bound in NAFIONV.27,28 In


R

fonated Radel sample at k ¼ 5 the 2H T1 was 20 ms which


Falk’s work, the DAO stretch in hyrogen-oxygen-deuterium
resulted in depressed conductivity for sulfonated Radel.
(HOD) was used to quantify the hydrogen bonding environ-
These authors hypothesized that the ionic domain structure
ment of water in Na-form NAFIONV.29 This vibrational signa-
R

had a significant influence on the observed results.


ture is convenient as the DAO stretch appears in a relatively
featureless region of the mid-IR spectrum at about 2509 PFG-NMR measurements have been used extensively to mea-
cm1, and there are no complications with interpretation of sure water self-diffusion on the 50 ms and longer time
the DAO stretch as compared to HAO due to lack of Fermi scales. In a 1D experiment, the length scale over which the
resonance of DAO bonds. Moilanen et al.30 observed a high diffusion coefficient is measured is given by:
frequency shoulder for DAO in NAFIONV at a frequency of
R

2708 cm1. The water associated with this stretch was pffiffiffiffiffiffiffiffi
l¼ 2Dt : (2)
interpreted to be in strong contact with the hydrophobic
portion of the NAFIONV membrane. The main DAO peak fre-
R

Typical water self-diffusion coefficients in PEMs are on the


quency was dependent on the hydration of the membrane
order of 1  1010 to 1  108 cm2 s1, which gives diffu-
and shifted from 2590 cm1 at a k ¼ 1 to 2553 cm1 at a k
sion lengths of 1–10 lm. This size scale is clearly larger than
¼ 9 as result of the relative populations of water changing
the nanoscale features of the ionic domain morphology in
as the membrane became more hydrated. This type of mea-
PFSAs or the ionic domains in a block copolymer-based
surement yields interesting information on the distribution
PEMs; however, good correspondence between chemical
of water environments, but a vibrational frequency is diffi-
cult to correlate to rotational or translational diffusion coeffi- structure and the water self-diffusion coefficient has been
cients, which may be more relevant to transport processes. observed.
Similar insights into NAFIONV–water interactions have been
R
Roy et al.38,39 reported that block copolymers display higher
gained from fluorescence spectroscopy and UV–vis adsorp- water self-diffusion coefficients than their random analogs
tion of fluorescent probe dyes.31–33 (Fig. 3). In random systems, the self-diffusion coefficient
Ultrafast spectroscopy has been used to make measurements measured by PFG-NMR declined steadily with increasing dif-
of the multiple hydrogen bond environments of water in fusion time. At long diffusion times, the self-diffusion coeffi-
cients in NAFIONV and block copolymers were not a function
R

NAFIONV membranes.30,34 The vibrational lifetimes in


R

NAFIONV membranes showed two components with values of diffusion length. The high conductivity in block copoly-
R

between 2 and 11 ps for hydration numbers between 1 and mers is thought to be due to their higher concentration
7.5, which compares to a vibrational lifetime of HOD in bulk of ionic groups in the functionalized domains and the
water of 1.7 ps with a single exponential decay. Interestingly, connectivity of the phases. However, the block copolymer
the vibrational lifetime experiments show that the character- architecture also appears to have an influence on the water
istic relaxation times of fast and slow relaxing water both behavior, which is intimately tied to their proton conduction
change as a function of hydration. This result indicates that properties.

12 JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20


WWW.POLYMERPHYSICS.ORG REVIEW

Li et al.40 showed that orientation in extruded NafionV 112


R
throughout the membrane for accurate determination of the
and dispersion cast NAFION (NRE)-212 membranes could be diffusion coefficient at a known hydration value.
detected through measurements of 2H NMR quadrupole cou-
Siu et al.45 showed that the water associations with the poly-
pling. The residual quadrupole splittings are related to the
mer may be beneficial in promoting conductivity under
orientational order parameter where the motion of water in
freezing conditions. On freezing PEMs to 40  C, the conduc-
the ionic domains is biased by the orientation of the ionic
tivity dropped by four orders of magnitude, and the activa-
domains. Biaxial orientation in the plane of the film was
tion energy of proton conduction increased from 0.15 to 0.4
observed in the extruded membranes, but the dispersion
eV. Appreciable conductivity in the frozen samples was
cast samples had uniaxial orientation in the thickness direc-
thought to stem from the water that was still tightly associ-
tion of the sample. Hou et al.41 used both 2H quadrupole
ated with the acid groups in the samples and did not show
splitting and pulsed field gradient diffusion measurements to
any thermal transitions around 0  C in DSC experiments. Ma
explore the anisotropy of block copolymer PEMs over differ-
et al.46 observed a water self-diffusion coefficient of 4 
ent length scales. The 2H splitting measurements probe
1011 m2 s1 at 240 K using PFG-NMR for membranes ini-
order on about the 0.4 lm length scale, where diffusion
tially equilibrated at 98% RH. This value compares to a
measurements probe larger length scales as discussed above.
water self-diffusion coefficient of 2  1010 m2 s1 at 273
These results demonstrate that the nature and number of
K. Although this work demonstrates that absorbed water in
defects and the local ordering of the material over multiple
PFSAs was moving more slowly in the membranes below
length scales can affect different measurements of water
0  C compared to membranes at higher temperatures, there
motion. More combined studies of structural order including
was enough mobility to promote proton transport. Other
defect type and density and water and ion mobility will help
studies of water mobility at low temperatures showed step
to clarify the role that defects play in the transport processes
changes in the water self-diffusion coefficient at low temper-
in these types of materials.
atures that were attributed to freezing events.47,48 These
The PFG-NMR technique rigorously measures the water self- membranes were probably more hydrated than the samples
diffusion coefficient (Dself), which is related to the Fickian studied by Ma et al., which lead to some portion of the water
diffusion coefficient (DFick) through: in the membrane freezing. These studies of the water mobil-
ity at sub-zero temperatures demonstrate that the water–
 
@lnðaw Þ polymer associations observed in PEMs show a wide range
DFick ¼ Dself ; (3) of behaviors and have a significant influence on the trans-
@lnðkÞ
port properties, even under conditions where bulk water is
where aw is the activity of water, so qln(aw)/qln(k) describes frozen.
the hydration of the membrane with relative humidity (RH). Dielectric spectroscopy has been used to explore the motion
Fickian diffusion coefficients for water in these types of poly- of water in hydrated PEMs. Paddison et al.49 showed that
the dielectric constant of NAFIONV increased with hydration
R
mers can be derived from dynamic water sorption experi-
ments or concentration gradient diffusion experiments. As number and decreased with increasing frequency, similar to
the water motion in sorption or concentration cell experi- that of liquid water. In sulfonated poly(aromatic) samples,
ments is due to a concentration gradient, these measure- the measured dielectric constant was a weaker function of
ments probe the Fickian diffusion coefficient and cannot be hydration number than what was observed for the PFSA
compared directly to PFG-NMR self-diffusion coefficients samples.50 Kreuer11 showed that the dielectric constants
unless the water uptake properties are known. measured in each case are a reflection of the ionic nano-
phase morphology of the material. Additionally, the dielectric
Hallinan and Elabd42 used attenuated total reflection-FTIR constant as a function of water volume fraction extrapolated
sorption measurements to quantify the water uptake proper- to 64 in NAFIONV at a water volume fraction of unity where
R

ties of NAFIONV during small and large step-changes in RH.


R

a value of 20 was observed for the sulfonated poly(aromatic)


These authors found that a Fickian diffusion model described sample. The dielectric response is reflective of the lower
most small RH step change experiments, for example, from water mobility observed in sulfonated poly(aromatic) sam-
43% to 56% RH, but for dry membranes or large steps in RH, ples compared to PFSAs. Lu et al.51 resolved different popu-
for example, from 0% to 22% RH or 0% to 100% RH, the lations of water in NAFIONV membranes using dielectric
R

uptake curves were non-Fickian. The authors were able to spectroscopy. Gigahertz relaxations were observed reflecting
decrease the mass transfer resistance at the vapor–membrane water with bulk-like properties in addition to a process that
interface by increasing the velocity of the gas phase surround- had a lower dielectric relaxation frequency response. The
ing the membrane, which is critical for recovering the correct dielectric strength of these two processes showed an
value of the diffusion coefficients. Water permeation experi- increase with hydration number indicating that the popula-
ments under steady-state concentration gradients can also tions of water molecules giving rise to each process
give insight into the water mobility of membranes as they are increased with hydration number. A process with kilohertz
exposed to different hydration conditions.43,44 However, per- relaxations was also assigned to the rotational motions of
forming these types of experiments for small gradients in the hydrated sulfonate groups and their first hydration
water activity is important to maintain consistent hydration spheres.51,52 Shifts in the relaxation frequencies to higher

WWW.MATERIALSVIEWS.COM JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20 13
REVIEW WWW.POLYMERPHYSICS.ORG

ferential scanning calorimetry experiments, which were cor-


related to increased water mobility. The activation energies
for transport also declined with increased heats of fusion of
water leading the researchers to conclude that more rapidly
diffusing water could be linked to faster molecular diffusion
of the solutes. The ionic domain size and connectivity was
not sufficient to explain these differences in the transport
properties of the membranes and the measurement of a sig-
nature of water binding seemed to bridge the understanding
of transport across chemically dissimilar materials. However,
it must be stressed in this study that the ionic domains
were studied in their dry state. TEM or scattering of dry
samples is often performed out of convenience, but techni-
ques to measure the size and connectivity of the hydrated
domains for a range of polymers are becoming more com-
FIGURE 4 Thermal signature of water in the membrane corre-
monplace.53–56
lates to proton conductivity (*) and methanol permeability The concept of electrochemical selectivity57 or the perme-
(l). Data taken from Ref. 4. ability-normalized conductivity is useful for quantifying the
tradeoff between the different transport properties of inter-
values reflected a larger population of faster moving water est for a given application. However, this type of figure of
as the hydration numbers were increased. merit can have shortcomings, especially for materials that
show high selectivity, but may have very low absolute values
There are a variety of methods for understanding the dy-
of conductivity.58 Thus, a minimum conductivity must be
namics of water and its distribution of environments when
first defined for a given application, and then candidate
absorbed in ion-containing membranes. PFG-NMR has the
materials can be selected once the primary goal of conduc-
advantage of a direct measurement of the water self-diffu-
tivity is met. Many sulfonated aromatic polymers show
sion coefficient, but it is difficult to extract distribution infor-
increased relative selectivity for proton conduction versus
mation from this type of measurement. Additionally, the
methanol permeation compared to PFSAs. This increase in
length scale over which the measurement is probing the dy-
selectivity can be interpreted both by considering the rela-
namics of water must be considered. Vibrational and dielec-
tive size of the ionic domains in each system and by visualiz-
tric spectroscopy can extract information on the different
ing how the water is distributed throughout these different
environments that water experiences in the material, but the
sized phases (Fig. 5). It has been shown that the permittivity
length-scale dependence of the measurements, the exchange
within an ionic pore changes radially.59–61 The high dielectric
of water between different microenvironments, and the
constant in the center of the pore gives rise to high rates of
meaningful parameters for proton transport are still being
water and solute transport. Therefore, larger pores contain
studied for a range of different PEMs. Comprehensive meas-
water that is less associated with the polymer, and high
urements tying conductivity values and measurements of
transport rates result. This situation not only increases pro-
water motion will further the community’s understanding of
ton conductivity but also can significantly increase methanol
proton transport in these types of ion-containing polymers
and water transport.
and this basic knowledge can be extended to study ion and
molecular transport beyond proton transport. Xu et al.62 were able to increase the selectivity of PEMs
through tuning the ion content of crosslinked polymers. In

TRANSPORT OF WATER AND SMALL MOLECULES

The water motion in fuel cells membranes is often measured


to interpret the proton conductivity of the materials, but the
water dynamics also have a pronounced impact on the trans-
port of water and small molecules in fuel cell membranes.
Hickner et al.4 correlated the thermal signatures of water
absorbed in a series of sulfonated poly(phenylene)s to the
transport of methanol (Fig. 4) and glucose through the mem-
branes. For highly sulfonated poly(phenylene)s, the diffusion
of methanol and glucose was greater than for NAFIONV
R

membranes. The ionic domains as observed by transmission


electron microscopy (TEM) were not as well organized in
the sulfonated aromatic polymers as they appeared in
NAFIONV; however, the absorbed water in the samples with
R
FIGURE 5 Down-pore view of ionic domains in PFSAs and sul-
high transport rates showed increased heats of fusion in dif- fonated poly(aromatic) PEMs.

14 JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20


WWW.POLYMERPHYSICS.ORG REVIEW

this work, the authors were able to connect the increased steady-state water content (except possibly at very low hydra-
electrochemical selectivity of their materials with a lower tion), but during changes in water uptake, polymer relaxation
thermal signature of water by calorimetry experiments. For a plays a significant role in determining the rate of water swel-
PEM with an ion exchange capacity (IEC) of 0.97 mequiv g1 ling or deswelling. Various models have been proposed for
and a water uptake of 19 wt %, a relative selectivity of 22 how polymer structure at the polymer/vapor interface controls
was obtained while maintaining a conductivity of 0.048 S water transport into a membrane.65 The difference in water
cm1. The low water thermal signature and high selectivity uptake of a membrane from the vapor or liquid phase with
was the result of the small ionic domains in these materials unit activity (saturated vapor or liquid water at the same tem-
and perhaps some added polarity due to the presence of the perature) has been ascribed to a hydrophobic surface layer
urethane bonds at the crosslink junctions. that develops in PFSAs when exposed to dry conditions or low
activity water vapor. This Schroeder’s paradox has been
Hickner and Pivovar63 correlated the water self-diffusion coef-
invoked to explain the interfacial resistance to slow water dif-
ficient, electro-osmotic drag, relative selectivity, and backbone
fusion into the membrane from the vapor phase. Onishi
stiffness for a variety of PEMs. The observation of electro-
et al.66 provided evidence that observations of Schroeder’s par-
osmotic drag being correlated with the relative selectivity is
adox are primarily governed by the equilibration time of the
interesting because this connection emphasizes the linkage
sample in the vapor phase and thus emphasized that polymer
between water motion and the other transport properties of
dynamics, even at extremely long time scales, control, in part,
the membrane. Electro-osmotic drag is a function of the num-
the water uptake of these types of materials.
ber of water molecules transported with each proton across
the membrane. For membranes with significantly confined Relaxation dynamics of the polymers are evident during
water, the electro-osmotic drag values will be low. However, dynamic water uptake experiments. Quantitative expressions
for membranes with less water–polymer associations, the vis- for the relaxation dynamics have been developed by Hallinan
and Elabd67 using a diffusion-relaxation model for NAFIONV
R
cous drag of the proton as it moves through the ionic domains
results in more water transport in concert with the proton adsorption. In this work, the authors demonstrated that by
and thus a higher electro-osmotic drag coefficient. Hickner minimizing the vapor–polymer boundary layer thickness dur-
and Pivovar’s correlation demonstrates that more mobile ing the water uptake experiments, the sorption curves could
water is tied with lower electrochemical selectivity. be described by a coupled Fickian diffusion process and a poly-
Interpreting the transport properties of PEMs through mer relaxation expression that likely occur on similar time-
insights into the mobility and binding of water absorbed in scales. For various models, the calculated diffusion coefficients
the polymer is a useful method for understanding the trade- were similar and spanned values of 2  107 to 9  107
off in properties that are prevalent in these types of materi- cm2 s1 over the full range of water activities. The self-diffu-
als. In many cases, if an increase in conductivity is desired, sion coefficient of water in ion-containing polymers increases
the diffusion of methanol and water will increase as well, systematically with hydration number (see Fig. 1). The Fickian
which is not always favorable for a given application. The diffusion coefficients, however, display a maximum at interme-
permeability of methanol can be suppressed in many PEMs diate hydration.5,68,69 This difference in the shape of the self-
but often at the expense of ionic conductivity. Even though diffusion and Fickian diffusion coefficients is due to the shape
ion conducting and molecular diffusion occurs through the of the sorption relationship with water activity (eq 3).
aqueous ionic domains in the material, the transport prop-
Satterfield et al.70 reported creep rates of dry and hydrated
erty tradeoffs can be optimized because the migration of
NAFIONV membranes over 1000 min. Hydrated samples
R

ions depends on the ionic concentration in the sample but


were observed to creep quickly at short times due to their
the molecular diffusion does not. For instance, Xu et al.64
plasticization by water, and dry samples have slower creep
were able to maintain a high degree of mobile water in
rates, but showed long-time relaxations. Interestingly, the
NAFIONV/sulfonated poly(silsesquioxane) composite samples
R

by increasing the IEC; however, the selectivity was improved swelling pressures measured during in situ hydration of the
even as the water uptake of the materials was increased. membrane showed relaxation on the timescale of the
This unusual trend in properties was due to maintaining mechanical creep experiments. Results for the swelling prop-
high ion exchange capacities and hence conductivities in the erties of the membranes in this work were discussed in the
materials due to the inclusion of the sulfonated filler, but the context of the competing energies of solvation of the ionic
filler also provided some blockage of the ionic domains thus groups and the energy required to swell the polymer matrix.
preventing methanol permeability through the material. This Interesting effects have been observed for the creep and me-
chanical properties of NAFIONV with low water contents.71–73
R
type of strategy will be useful going forward in optimizing
the transport property tradeoffs in other types of mem- Under specific conditions of temperature and water activity,
NAFIONV can undergo antiplasticization, where the addition
R
branes for various applications.
of water increases the tensile modulus of the membrane. The
POLYMER RELAXATION DURING UPTAKE AND mechanism of this antiplasticization is not entirely clear, but it
MECHANICAL PROPERTIES
may be due to microstructural changes in the polymer or ioni-
The dynamic motion of the polymer molecules do not make a zation/bonding interactions of the sulfonic acid groups on
primary contribution to proton motion under conditions of small additions of water.

WWW.MATERIALSVIEWS.COM JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20 15
REVIEW WWW.POLYMERPHYSICS.ORG

Harrison et al.74 reported that the modulus of a disulfonated deficiencies of aromatic polymers in mechanical properties
poly(arylene ether sulfone) membrane was one order of are a major concern for long lifetime devices. The synthesis of
magnitude higher than that of NAFIONV but declined precipi-
R
block copolymer PEMs with soft blocks may be one route to
tously with hydration. Similar observations have been made materials with greater toughness and elongation properties,
by Fujimoto et al.75 for sulfonated poly(phenylene)s where but synthesizing these types of materials with chemical moi-
the Young’s modulus of NAFIONV declined by a factor of four
R
eties that can withstand the electrochemical stresses and reac-
for hydrated samples, but the modulus of the sulfonated tive species in an operating device is a technical challenge.
poly(phenylene) dropped by a greater percentage due to a
The mechanical stabilization of the membrane is important to
larger plasticization effect. Under hydrated conditions, the
preserve the morphology of the water-filled ionic domains that
modulus of sulfonated aromatic polymers is generally
contribute to high transport rates in these types of membranes.
greater than that of PFSAs. The difference in mechanical
Additionally, too much water uptake by the polymer can dilute
properties between materials narrows under hydrated condi- the concentration of ionic groups and lower the resulting con-
tions due to the greater extent of plasticization of aromatic ductivity.79 Kim et al.80 measured the disruption of the ionic do-
polymers by water. The aromatic polymers are more plasti- main morphology of NAFIONV and disulfonated poly(arylene
R

cized because of the greater extent of water–polymer inter- ether sulfone) proton conducting membranes and correlated
actions in non-perfluorianted samples as compared to PFSAs. these microscopic changes to the overall membrane swelling
Kim et al.76 showed that the Kelley–Bueche equation (a vari- and proton conductivity. The authors proposed an ‘‘upper limit
ant of the Fox equation) described the decrease in the Tg of use temperature’’ that describes the maximum temperature at
NAFIONV and disulfonated poly(arylene ether sulfone)s which the polymer will not over-swell when exposed to liquid
R

membranes up to a certain water content. For NAFIONV the water. This concept can be extended to polymers that swell
R

Tg strongly decreased with increasing water content and fol- highly at room temperature, an upper swelling parameter, and
lowed the Kelley–Bueche equation up to a bulk water uptake provides guidelines for the ideal concentration of ions in the
of 5 wt %. The drop in Tg was more severe for the disulfo- membrane in the swollen state. In other work, Kim et al.81
nated poly(arylene ether sulfone) sample and the Tg depres- showed that even though the hydrated ion concentration can
sion followed the Kelley–Bueche equation up to nearly 20 wt decline for samples that uptake significant amounts of water,
% water uptake. At higher water contents than these critical the proton conductivity can continue to increase up to a point
values, the Tg of each material departed from the prediction even though the ion concentration in the sample decreases.
of plasticization and was only weakly influenced by the fur- This increase in conductivity, even under declining proton con-
ther addition of water. The authors hypothesized that the centration, is due to faster proton transport through the aque-
critical value of water uptake for each material was con- ous phase that displays more rapid water rotation and diffusion
trolled by the amount of water–polymer interactions for a with the increased water uptake. These observations lead to the
given sample. As the disulfonated poly(sulfone) material had conclusion that there is no one descriptor of high conductivity
small domains and a polarizable backbone, it had more con- in PEMs, but the ionic domain morphology, water–polymer
tact between the polymer and water which increased the interactions, and ion concentration must all be weighed when
critical value of water uptake. NAFIONV with its perfluori-
R designing a material for a specific use.
nated backbone and well-defined ion domain structure was Dynamic water uptake has been investigated in submicron
NAFIONV films. Kongkanand82 ascribed the slow water uptake
R
not as plasticized by water and instead the water was col-
lected in the ionic nanophase of the material. of nm-thick films to the hydrophobic interface layer that com-
monly features in Schroeder’s paradox arguments to explain
An important aspect of a membrane is its ability to withstand
the water uptake of thick (>20 lm) films. However, relaxation
mechanical deformation during device assembly or operating
may also make a significant contribution to the slow uptake
stresses.77 Many ion-containing aromatic polymers have
dynamics, especially for these thin films with very short diffu-
higher moduli than what is observed for PFSAs, but the elon-
sion lengths. In the absence of structural data to confirm and
gation properties of aromatic polymer rarely exceed 30–50%
quantify the structure and composition of the interface layer,
in the wet state, where NAFIONV’s elongation to break is
R

it is difficult to separate interfacial transport resistance and


about 170% and does not change significantly from wet to
slow uptake dynamics due to polymer relaxation.
dry conditions.75 The poor elongation properties of many
PEMs is demonstrated by their inability to withstand wet–dry Understanding how nanometer-scale thick films of ion-con-
cycling. In this type of test, the membrane is mechanically taining polymers absorb water and probing the water self-
constrained, and the RH surrounding the membrane is cycled diffusion, conductivity, and ionic domain morphology in
on the timescale of minutes between wet (usually 80% RH or these thin films are important in understanding how these
higher) and dry (usually 20% RH or lower) conditions. Vari- polymers behave in high surface area porous electrodes. In
ous PFSAs and reinforced PFSAs can approach or exceed these electrode structures, the ionomer is generally distrib-
thousands or tens of thousands of cycles in this type of test. uted as thin films throughout the electrode layer as a com-
However, few aromatic polymers have demonstrated good posite with the catalyst particles. The function of many elec-
performance under these conditions. The low oxygen perme- trochemical devices relies on a delicate balance of ion
ability of sulfonated aromatic polymers increases their resist- conductivity, electrical conductivity, reactant and product
ance to chemical attack by reactive oxygen species,78 but the molecular transport, and exposure of the reactive sites.

16 JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20


WWW.POLYMERPHYSICS.ORG REVIEW

Currently in fuel cells, the electrode is a complex mixture of


components without the appearance of regular order,83 with
perhaps just one notable exception that has demonstrated
excellent performance but still has shortcomings in terms of
its ability to operate under freezing conditions.84 Extending
studies of the structure–function relationships of bulk ion-
containing membranes to nanometer-scale thin films85,86 will
promote more rational design of polymer composite electro-
des and promote new tools and concepts to understand their
performance.

CONCLUSIONS AND DIRECTIONS

As the use of ion-containing polymers are expanded to new


uses and new technologies, specific focus on the role of
water in determining these polymers’ transport, permselec-
tivity, and mechanical properties will help to accelerate their FIGURE 6 Water self-diffusion coefficients of NAFIONV
R (~),

development. Specifically, development of materials for alka- sulfonate (n) and quaternary ammonium (l) functionalized
line fuel cells that require hydroxide (OH) polymeric con- poly(aromatic) membranes. Data taken from Ref. 3.
ductors and redox flow batteries where high proton conduc-
tivity and low vanadium ion crossover is desired can benefit
from the body of knowledge surrounding proton conducting in the presence of anions versus cations was invoked to
membranes for fuel cells. explain the differences in the two systems. Similar physics
may be at play in quaternary ammonium-containing AEMs in
Alkaline membrane fuel cells are important because they comparison to sulfonated PEMs, which could drive different
have the potential to free fuel cell technology from the cost relationships in their water uptake and ion conductivity
constraints of precious metal catalysts. Catalyst corrosion properties.
may not be as severe in fuel cells with a high internal pH as
compared to the more well-known acidic PEMFCs, and the It is likely that other studies of AEMs have uncovered similar
membrane-based architecture affords facile device and sys- trends in the water uptake.90–92 However, more focus on the
tem operation. Yan and Hickner87 have outlined the critical basic structure–property relationships of these materials is
relationship between bicarbonate conductivity and water needed if their conductivity and water uptake properties are
uptake for a series of poly(sulfone)-based anion exchange to be further optimized. There also exists the interesting
membranes (AEMs). A strong correlation was observed potential to study the influence of polymer architecture93,94
between the bulk water uptake of a series of AEMs and their and cation type95,96 in these types of materials, which will
bicarbonate conductivity, regardless of IEC. The methanol expand the fundamental knowledge surrounding the proper-
permeability also scaled proportionately with the bulk water ties of ion-containing polymer membranes with tethered cat-
uptake and the activation energies of ion conduction and ions. The hydration/conductivity tradeoff is critical to opti-
methanol permeability were inversely related to the amount mize in AEMs, but hydration may be important for avoiding
of water in the membrane. It is possible that the hopping degradation of the cation as demonstrated by Chempath
properties of hydroxide will reveal different trends as has et al.97 Sulfonated polymers are known to degrade when
been observed with bicarbonate (which does not undergo they are stored in anhydrous conditions or at high tempera-
hopping transport), but rigorous measurements of hydroxide tures for significant amounts of time in the acidic form.
conductivity remain difficult due to rapid adsorption of CO2 There is much less degradation observed for sulfonated
in AEMs. membranes in the sodium or tetrabutyl ammonium neutral-
Interestingly, the water self-diffusion in AEMs appears to ized form. The quaternary ammonium functionality of AEMs
scale differently with ion content compared to the relation- is less stable than sulfonate groups, so hydration strategies
ships observed for PEMs (Fig. 6).3,88 NAFIONV has a high
R to preserve the integrity of the cation for membranes in hy-
water self-diffusion coefficient because of its large ionic droxide or bicarbonate (that becomes hydroxide at higher
domains. The difference in water association with cationic temperatures under low pCO2 conditions) forms are a consid-
polymers compared to anion polymers may be similar to eration when designing membranes for a specific use. Stor-
that measured in cationic and anionic reverse micelles. age of these materials in their chloride form appears to miti-
Dokter et al.89 have shown that water interacts less strongly gate any degradation due to the weak nucleophilicity of the
with ions in CTAB (cetyltrimethylammonium bromide, a cati- chloride anion.
onic quaternary ammonium-based surfactant) reverse Redox flow batteries are a viable technology for grid-scale
micelles compared to water in AOT (sodium bis(2-ethy- energy storage and can serve as buffering capacity for wide-
thexyl) sulfosuccinate, an anionic sulfonate-based surfactant) spread renewable energy deployment. In vanadium redox
reverse micelles. The difference in water hydrogen bonding flow batteries (VRFBs), the redox reactions that store and

WWW.MATERIALSVIEWS.COM JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20 17
REVIEW WWW.POLYMERPHYSICS.ORG

discharge electrical energy are balanced by the flow of pro- (CBET-0932740) for support of his group’s research in the syn-
tons across a membrane separating the anolyte and catholye thesis, properties, and applications of ion-containing polymers.
containing V2þ/V3þ and V4þ/V5þ reversible redox species.
High proton conductivity of the membrane lowers the resis-
REFERENCES AND NOTES
tive contributions to the efficiency loss in these types of
devices, but vanadium ion diffusion through the membrane 1 Peckham, T. J.; Schmeisser, J.; Rodgers, M.; Holdcroft, S.
must be prevented. Transport of the vanadium redox species J. Mater. Chem. 2007, 17, 3255–3268.
across the membrane causes capacity fade during reversible 2 Kreuer, K. D. Solid State Ionics 1997, 94, 55–62.
cycling of the battery. The redox activity of the electrolytes 3 Hibbs, M. R.; Hickner, M. A.; Alam, T. M.; McIntyre, S. K.; Fuji-
can be regenerated, but maintenance of the electrolyte moto, C. H.; Cornelius, C. J. Chem. Mater. 2008, 20, 2566–2573.
causes downtime of the energy storage device and adds 4 Hickner, M. A.; Fujimoto, C. H.; Cornelius, C. J. Polymer 2006,
maintenance costs to the overall system. As grid-scale energy 47, 4238–4244.
storage devices are multi MW-class systems, any additional 5 Lee, D. K.; Saito, T.; Benesi, A. J.; Hickner, M. A.; Allcock, H.
R. J Phys Chem B 2011, 115, 776–783.
fixed or operating expenses must be kept to an absolute
minimum. 6 Mangiagli, P. M.; Ewing, C. S.; Xu, K.; Wang, Q.; Hickner, M.
A. Fuel Cells 2009, 9, 432–438.
In one study by Kim et al.,98 NAFIONV-based cells experienced
R

7 Mauritz, K. A.; Moore, R. B. Chem. Rev. 2004, 104, 4535–4585.


a capacity loss of 13 mA h per cycle over 40 cycles compared 8 Gebel, G. Polymer 2000, 41, 5829–5838.
to a 6.4 mA h per cycle decline over an equivalent number of 9 Rubatat, L.; Rollet, A. L.; Gebel, G.; Diat, O. Macromolecules
cycles for VRFBs with sulfonated poly(sulfone) RADELV mem-
R
2002, 35, 4050–4055.
branes. The sulfonated poly(sulfone) membrane displayed six 10 Schmidt-Rohr, K.; Chen, Q. Nat. Mater. 2008, 7, 75–83.
times less V4þ ion transport than NAFIONV which was the ori-
R

11 Kreuer, K. D. J. Membr. Sci. 2001, 185, 29–39.


gin of its low capacity fade. So far, there are no studies of 12 Ratner, M. A., in: Polymer Electrolytes Reviews-l, eds. Mac-
how the water binding in ion-conducting membranes influen- Callum, J. R. and Vincent, C. A. Elsevier Applied Science, Lon-
ces the vanadium crossover in VRFBs or transport processes don, 1987, p. 173.
in different types of flow batteries, such as iron-chromium or 13 Dou, S. C.; Zhang, S. H.; Klein, R. J.; Runt, J.; Colby, R. H.
zinc-bromine.99 However, the trends in the comparison of va- Chem. Mater. 2006, 18, 4288–4295.
nadium ion transport in NAFIONV membranes and sulfonated
R
14 Fragiadakis, D.; Dou, S. C.; Colby, R. H.; Runt, J. Macromole-
aromatic PEMs appears to be similar to what has been cules 2008, 41, 5723–5728.
observed for methanol transport (Fig. 5). The biggest caveat 15 Chen, Y. B.; Thorn, M.; Christensen, S.; Versek, C.; Poe, A.;
thus far in applying membranes other than PFSAs to VRFBs Hayward, R. C.; Tuominen, M. T.; Thayumanavan, S. Nat.
Chem. 2010, 2, 503–508.
appears to be the increased degradation observed in the case
16 Schuster, M. F. H.; Meyer, W. H.; Schuster, M.; Kreuer, K. D.
of sulfonate poly(sulfone) membranes.100 New, more intrinsi-
Chem. Mater. 2004, 16, 329–337.
cally stable polymers and membrane engineering strategies101
17 Tokuda, H.; Hayamizu, K.; Ishii, K.; Susan, M.; Watanabe, M.
will help to alleviate the lifetime problems of non-PFSA mate- J. Phys. Chem. B 2005, 109, 6103–6110.
rials, but the demand for tens of thousands of hours of cycle
18 Stickel, F.; Fischer, E. W.; Richert, R. J. Chem. Phys. 1996,
life will remain a challenge. 104, 2043–2055.
To conclude, there are many facets to the design of new ion- 19 Xu, W.; Cooper, E. I.; Angell, C. A. J. Phys. Chem. B 2003,
containing membranes for electrochemical and water treat- 107, 6170–6178.
ment applications.102 This review article has emphasized the 20 Zawodzinski, T. A.; Neeman, M.; Sillerud, L. O.; Gottesfeld,
role that water–polymer associations play in controlling the S. J. Phys. Chem. 1991, 95, 6040–6044.
properties of these materials, but the effect of ionic domain 21 Kreuer, K. D. Solid State Ionics 1997, 97, 1–15.
morphology and other factors such as charge concentration 22 Zawodzinski, T. A.; Derouin, C.; Radzinski, S.; Sherman, R.
in these systems influence the water behavior and ultimately J.; Smith, V. T.; Springer, T. E.; Gottesfeld, S. J. Electrochem.
Soc. 1993, 140, 1041–1047.
couple to the transport properties. There is no single, simple
predictor of conductivity and permeability in ion-conductiv- 23 Dippel, T.; Kreuer, K. D.; Lassegues, J. C.; Rodriguez, D.
Solid State Ionics 1993, 61, 41–46.
ity membranes, but investigations of the water behavior
have opened up areas of fundamental interest and technolog- 24 Subbaraman, R.; Ghassemi, H.; Zawodzinski, T. A. J. Am.
Chem. Soc. 2007, 129, 2238–2239.
ical importance and provide one set of design criteria for
25 Zhou, Z.; Li, S. W.; Zhang, Y. L.; Liu, M. L.; Li, W. J. Am.
optimizing these materials for a specific purpose.
Chem. Soc. 2005, 127, 10824–10825.
26 Paddison, S. J.; Zawodzinski, T. A. Solid State Ionics 1998,
ACKNOWLEDGMENTS
113, 333–340.
M.A.H. thank the Penn State Materials Research Institute, the 27 Wang, Y. Q.; Kawano, Y.; Aubuchon, S. R.; Palmer, R. A.
Penn State Institutes of Energy and the Environment, the U.S. Macromolecules 2003, 36, 1138–1146.
Army Research Office (W911NF-08-1-0282 and W911NF-11-1- 28 Laporta, M.; Pegoraro, M.; Zanderighi, L. Phys. Chem.
0411), the U.S. Office of Naval Research (N00014-08-1-0730 and Chem. Phys. 1999, 1, 4619–4628.
N00014-10-1-0875), and the U.S. National Science Foundation 29 Falk, M. Can. J. Chem. 1980, 58, 1495–1501.

18 JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20


WWW.POLYMERPHYSICS.ORG REVIEW

30 Moilanen, D. E.; Piletic, I. R.; Fayer, M. D. J. Phys. Chem. A 59 Yang, Y.; Pintauro, P. N. Ind. Eng. Chem. Res. 2004, 43,
2006, 110, 9084–9088. 2957–2965.
31 Spry, D. B.; Fayer, M. D. J. Phys. Chem. B 2009, 113, 60 Yang, Y.; Walz, J.; Pintauro, P. J. Chem. Soc.-Faraday Trans.
10210–10221. 1997, 93, 603–611.
32 Moilanen, D. E.; Spry, D. B.; Fayer, M. D. Langmuir 2008, 24, 61 Paul, R.; Paddison, S. J. J. Phys. Chem. B 2004, 108,
3690–3698. 13231–13241.
33 Spry, D. B.; Goun, A.; Glusac, K.; Moilanen, D. E.; Fayer, M. 62 Xu, K.; Chanthad, C.; Hickner, M. A.; Wang, Q. J. Mater.
D. J. Am. Chem. Soc. 2007, 129, 8122–8130. Chem. 2010, 20, 6291–6298.
34 Moilanen, D. E.; Piletic, I. R.; Fayer, M. D. J. Phys. Chem. C 63 Hickner, M. A.; Pivovar, B. S. Fuel Cells 2005, 5, 213–229.
2007, 111, 8884–8891. 64 Xu, K.; Chanthad, C.; Gadinski, M. R.; Hickner, M. A.; Wang,
35 Lang, E. W.; Ludemann, H. D.; Piculell, L. J. Chem. Phys. Q. ACS Appl. Mater. Interfaces 2009, 1, 2573–2579.
1984, 81, 3820–3827. 65 Freger, V. J. Phys. Chem. B 2009, 113, 24–36.
36 Geil, B.; Fujara, F.; Sillescu, H. J. Magn. Reson. 1998, 130, 66 Onishi, L. M.; Prausnitz, J. M.; Newman, J. J. Phys. Chem. B
18–26. 2007, 111, 10166–10173.
37 Woessner, D. E. In Encyclopedia of Nuclear Magnetic Reso-
67 Hallinan, D. T.; De Angelis, M. G.; Baschetti, M. G.; Sarti, G.
nance; Grant, D. M.; Harris, R. K.; Eds.; Wiley: Chichester, UK,
C.; Elabd, Y. A. Macromolecules 2010, 43, 4667–4678.
1996, pp 1068–1084.
68 Morris, D. R.; Sun, X. D. J. Appl. Polym. Sci. 1993, 50,
38 Roy, A.; Hickner, M. A.; Yu, X.; Li, Y. X.; Glass, T. E.;
1445–1452.
McGrath, J. E. J. Polym. Sci. Part B-Polym. Phys. 2006, 44,
2226–2239. 69 Legras, M.; Hirata, Y.; Nguyen, Q. T.; Langevin, D.; Metayer,
M. Desalination 2002, 147, 351–357.
39 Roy, A.; Yu, X.; Dunn, S.; McGrath, J. E. J. Membr. Sci.
2009, 327, 118–124. 70 Satterfield, M. B.; Majsztrik, P. W.; Ota, H.; Benziger, J. B.;
Bocarsly, A. B. J. Polym. Sci. Part B-Polym. Phys. 2006, 44,
40 Li, J.; Wilmsmeyer, K. G.; Madsen, L. A. Macromolecules
2327–2345.
2008, 41, 4555–4557.
71 Majsztrik, P. W.; Bocarsly, A. B.; Benziger, J. B. Macromole-
41 Hou, J. B.; Li, J.; Madsen, L. A. Macromolecules 2010, 43,
cules 2008, 41, 9849–9862.
347–353.
72 Satterfield, M. B.; Benziger, J. B. J. Polym. Sci. Part
42 Hallinan, D. T.; Elabd, Y. A. J. Phys. Chem. B 2009, 113,
B-Polym. Phys. 2009, 47, 11–24.
4257–4266.
73 Bauer, F.; Denneler, S.; Willert-Porada, M. J. Polym. Sci.
43 Adachi, M.; Navessin, T.; Xie, Z.; Li, F. H.; Tanaka, S.; Hold-
Part B-Polym. Phys. 2005, 43, 786–795.
croft, S. J. Membr. Sci. 2010, 364, 183–193.
44 Romero, T.; Merida, W. J. Membr. Sci. 2009, 338, 135–144. 74 Harrison, W. L.; Hickner, M. A.; Kim, Y. S.; McGrath, J. E.
Fuel Cells 2005, 5, 201–212.
45 Siu, A.; Schmeisser, J.; Holdcroft, S. J. Phys. Chem. B 2006,
110, 6072–6080. 75 Fujimoto, C. H.; Hickner, M. A.; Cornelius, C. J.; Loy, D. A.
Macromolecules 2005, 38, 5010–5016.
46 Ma, Z. R.; Jiang, R. C.; Myers, M. E.; Thompson, E. L.; Gittle-
man, C. S. J. Mater. Chem. 2011, 21, 9302–9311. 76 Kim, Y. S.; Dong, L. M.; Hickner, M. A.; Glass, T. E.; Webb,
V.; McGrath, J. E. Macromolecules 2003, 36, 6281–6285.
47 Saito, M.; Hayamizu, K.; Okada, T. J. Phys. Chem. B 2005,
109, 3112–3119. 77 Lai, Y. H.; Mittelsteadt, C. K.; Gittleman, C. S.; Dillard, D. A.
J. Fuel Cell Sci. Technol. 2009, 6.
48 Nicotera, I.; Coppola, L.; Rossi, C. O.; Youssry, M.; Ranieri,
G. A. J. Phys. Chem. B 2009, 113, 13935–13941. 78 Sethuraman, V. A.; Weidner, J. W.; Haug, A. T.; Protsailo, L.
V. J. Electrochem. Soc. 2008, 155, B119–B124.
49 Paddison, S. J.; Reagor, D. W.; Zawodzinski, T. A. J. Elec-
troanal. Chem. 1998, 459, 91–97. 79 Gottesfeld, S.; Zawodzinski, T. A. In Advances in Electro-
chemical Science and Engineering; Alkire, R. C.; Gerischer, H.;
50 Paddison, S. J.; Bender, G.; Kreuer, K. D.; Nicoloso, N.;
Kolb, D. M.; Tobias, C. W., Eds.; Wiley-VCH, New York, 1997,
Zawodzinski, T. A. J. New Mater. Electrochem. Syst. 2000, 3,
pp 195–301.
291–300.
80 Kim, Y. S.; Wang, F.; Hickner, M.; McCartney, S.; Hong, Y.
51 Lu, Z. J.; Polizos, G.; Macdonald, D. D.; Manias, E. J. Elec-
T.; Harrison, W.; Zawodzinski, T. A.; McGrath, J. E. J. Polym.
trochem. Soc. 2008, 155, B163–B171.
Sci. Part B-Polym. Phys. 2003, 41, 2816–2828.
52 Tsonos, C.; Apekis, L.; Pissis, P. J. Mater. Sci. 2000, 35,
5957–5965. 81 Kim, Y. S.; Einsla, B.; Sankir, M.; Harrison, W.; Pivovar, B. S.
Polymer 2006, 47, 4026–4035.
53 Park, M. J.; Downing, K. H.; Jackson, A.; Gomez, E. D.;
Minor, A. M.; Cookson, D.; Weber, A. Z.; Balsara, N. P. Nano 82 Kongkanand, A. J. Phys. Chem. C 2011, 115, 11318–11325.
Letters 2007, 7, 3547–3552. 83 Eikerling, M. J. Electrochem. Soc. 2006, 153, E58–E70.
54 Gebel, G.; Diat, O. Fuel Cells 2005, 5, 261–276. 84 Debe, M. K.; Schmoeckel, A. K.; Vernstrorn, G. D.; Atana-
55 Gebel, G.; Lambard, J. Macromolecules 1997, 30, soski, R. J. Power Sources 2006, 161, 1002–1011.
7914–7920. 85 Siroma, Z.; Kakitsubo, R.; Fujiwara, N.; Ioroi, T.; Yamazaki,
56 Aieta, N. V.; Stanis, R. J.; Horan, J. L.; Yandrasits, M. A.; S.-i.; Yasuda, K. J. Power Sources 2009, 189, 994–998.
Cookson, D. J.; Ingham, B.; Toney, M. F.; Hamrock, S. J.; Her- 86 Paul, D. K.; Fraser, A.; Karan, K. Electrochem. Commun.
ring, A. M. Macromolecules 2009, 42, 5774–5780. 2011, 13, 774–777.
57 Pivovar, B. S.; Wang, Y. X.; Cussler, E. L. J. Membr. Sci. 87 Yan, J. L.; Hickner, M. A. Macromolecules 2010, 43,
1999, 154, 155–162. 2349–2356.
58 Siu, A.; Pivovar, B.; Horsfall, J.; Lovell, K. V.; Holdcroft, S. J. 88 Hickner, M. A.; Tudryn, G. J.; Alam, T. M.; Hibbs, M. R.; Fuji-
Polym. Sci. Part B-Polym. Phys. 2006, 44, 2240–2252. moto, C. H. Macromol. Symp. 2009, 279, 59–62.

WWW.MATERIALSVIEWS.COM JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20 19
REVIEW WWW.POLYMERPHYSICS.ORG

89 Dokter, A. M.; Woutersen, S.; Bakker, H. J. J. Chem. Phys. 96 Wang, J.; Li, S.; Zhang, S. Macromolecules 2010, 43,
2007, 126. 3890–3896.
90 Wang, G. G.; Weng, Y. M.; Chu, D.; Chen, R. R.; Xie, D. J. 97 Chempath, S.; Einsla, B. R.; Pratt, L. R.; Macomber, C. S.;
Membr. Sci. 2009, 332, 63–68. Boncella, J. M.; Rau, J. A.; Pivovar, B. S. J. Phys. Chem. C
91 Slade, R. C. T.; Varcoe, J. R. Solid State Ionics 2005, 176, 2008, 112, 3179–3182.
585–597. 98 Kim, S.; Yan, J.; Schwenzer, B.; Zhang, J.; Li, L.; Liu, J.;
92 Robertson, N. J.; Kostalik, H. A.; Clark, T. J.; Mutolo, P. F.; Yang, Z.; Hickner, M. A. Electrochem. Commun. 2010, 12,
Abruna, H. D.; Coates, G. W. J. Am. Chem. Soc. 2010, 132, 1650–1653.
3400–3404.
99 Yang, Z.; Zhang, J.; Kintner-Meyer, M. C. W.; Lu, X.; Choi,
93 Faraj, M.; Elia, E.; Boccia, M.; Filpi, A.; Pucci, A.; Ciardelli, F. D.; Lemmon, J. P.; Liu, J. Chem. Rev. 2011, 111, 3577–3613.
J. Polym. Sci. Part A-Polym. Chem. 2011, 49, 3437–3447.
100 Kim, S.; Tighe, T.; Schwenzer, B.; Yan, J.; Zhang, J.; Liu,
94 Tanaka, M.; Fukasawa, K.; Nishino, E.; Yamaguchi, S.;
J.; Yang, Z.; Hickner, M. J. Appl. Electrochem. 2011, 41, 1–13.
Yamada, K.; Tanaka, H.; Bae, B.; Miyatake, K.; Watanabe, M.
J. Am. Chem. Soc. 2011, 133, 10646–10654. 101 Luo, Q.; Zhang, H.; Chen, J.; You, D.; Sun, C.; Zhang, Y.
95 Gu, S.; Cai, R.; Luo, T.; Chen, Z.; Sun, M.; Liu, Y.; He, G.; J. Membr. Sci. 2008, 325, 553–558.
Yan, Y. Angew. Chem.-Int. Ed. 2009, 48, 6499–6502. 102 Hickner, M. A. Mater. Today 2010, 13, 34–41.

20 JOURNAL OF POLYMER SCIENCE PART B: POLYMER PHYSICS 2012, 50, 9–20

You might also like