Você está na página 1de 11

Solid State Ionics 125 (1999) 111

www.elsevier.com / locate / ssi

Solid-state protonic conductors: principles, properties, progress and


prospects
Truls Norby*
Department of Chemistry, University of Oslo, Centre for Materials Science, Gaustadalleen 21, N-0349 Oslo, Norway

Abstract
A brief overview is given of the types and principles of solid state protonic conductors. Their properties are summarised in
terms of protonic conductivity and operation temperature. A few trends and highlights of the current scientific advances are
given for some of the main classes of protonic conductors. Despite our progress, a crucial point still seems to be the ability
and computer power required to understand and model the complex quantum mechanical proton transfer process. Many of
the commercial prospects are closely connected with the need for cleaner energy technologies and the use of hydrogen as an
energy carrier. In particular, polymer-based proton conductors are quickly approaching the market for small fuel cells and
electrolysers: from bench-scale demonstration units for hydrogen energy to automobiles. There is also emerging interest in
mixed protonicelectronic conductors for use as hydrogen permeable membranes in hydrogen separation technologies. This
will also be helpful in the much needed development of electrodes for high- and intermediate-temperature proton conductors.
1999 Elsevier Science B.V. All rights reserved.
Keywords: Oxides; Perovskites; Oxyhydroxides; Hydroxides; Polymers; Mixed conduction; Hydrogen permeable membranes; Hydrogen
separation

1. Introduction
Materials with high and pure protonic conductivity
are candidates for electrolytes in sensors, batteries,
fuel cells, electrolysers, etc. The typical solid-state
protonic conductors developed a couple of decades
ago were mainly acidic or hydrous inorganic compounds, as reflected in the focii of the early international conferences on Solid State Protonic Conductors (SSPC). Later, entirely different classes of
materials gained increasing interest as proton con*Tel.: 147-229-58722; fax: 147-229-58749.
E-mail address: truls.norby@fys.uio.no (T. Norby)

ductors: polymers, oxide ceramics, intercalation


compounds, etc. Geoscientists have a long tradition
of analysing water and hydroxide in minerals, and
proton transfer and migration is studied in detail in
biological systems and processes, e.g. cell walls and
photosynthesis. In the development of solid-state
proton conductors we may thus seek contribution
from studies of protons irrespective of the medium or
the magnitude or dominance of the protonic conductivity, and thus include organics, biological systems, minerals, ceramics, and aqueous / liquid systems. The present paper, serving as an introduction
to the proceedings of SSPC9 (Bled, Slovenia, August
1998) [1], will briefly introduce the field and some

0167-2738 / 99 / $ see front matter 1999 Elsevier Science B.V. All rights reserved.
PII: S0167-2738( 99 )00152-6

T. Norby / Solid State Ionics 125 (1999) 1 11

central classes of materials, and point out current


advances and future developments.

2. Principles of protonic conduction classes


of proton conductors
Protonic conductors are often considered to be
electrolytes in which hydrogen is transported towards
and evolved at the cathode during electrolysis. This
would include electrolytes carrying positively
charged protonic species such as H 1 and H 3 O 1 and
even NH 41 . On the other hand, OH 2 falls outside this
definition, but is still of relevance and interest to us.
The uniqueness of the bare proton nucleus is lost
when we consider hydrogen in other oxidation states
(atoms, molecules, hydride ions). These are therefore
considered outside our scope, except that they may
share the strong and interesting isotope effects of
protons.
Thus, for our purposes, protonic transport includes
transport of protons (H 1 ) and any assembly that
2
carries protons (OH 2 , H 2 O, H 3 O 1 , NH 1
4 , HS ,
1
etc.). The transport of protons (H ) between relatively stationary host anions is termed the Grotthuss
or free-proton mechanism. Transport by any of the
other species is termed a vehicle mechanism.
Vehicle mechanisms are most frequently encountered in aqueous solution and other liquids / melts.
Thus, both aqueous and molten NaOH are OH 2 ion
conductors. H 2 SO 4 (l) and H 2 SO 4 (aq) are H 1 and
H 3 O 1 conductors depending on the water content. In
solids, however, vehicle mechanisms are usually
restricted to materials with open structures (channels,
layers) to allow passage of the large ions and
molecules. For instance, SiO 2 and many silicates
allow diffusion of H 2 O molecules, and b-Al 2 O 3
allows migration of H 3 O 1 and NH 1
4 ions. Although
OH 2 transport is also possible in open structures
(interstitial mechanism) it is interesting to note that
there is hardly any unambiguous evidence for OH 2
transport in a crystalline solid. For instance, NaOH(s)
is a proton conductor by a free-proton mechanism
[2].
Above we have to some extent classified conduction according to the transported species. Other
classifications may involve type of synthesis, low vs.
high operating temperature, organic vs. inorganic

compounds, etc. In the present treatment we will


classify materials according to their content and state
of water and protons.
Many solid protonic conductors have liquid or
liquid-like regions or layers of water, and thus
substantial water contents. Such materials may be on
the borderline between solid and liquid states, or
simply be two-phase systems. Examples include the
so-called particle hydrates where adsorbed hydrous
layers on each grain provide the proton conductive
medium, polymer membranes which must be hydrated in order to provide a liquid-like H 2 OH 3 O 1
system for protonic transport, and phosphonates in
which phosphate layers may be separated by spacer
groups allowing water to become absorbed. The
protons are often introduced by ion exchange of
alkali cations, and their positive charge often compensates the negative charge of anions fixed to the
backbone of the structure. High protonic conduction
in such water-rich liquid-like systems relies on the
dynamics of water molecules as protonic vehicles.
The next type of water-containing material has
crystallographically fixed protons. This group comprises organic and inorganic hydrates, i.e. materials
where there are oxygen ions with two (or three)
protons. We may also have a single proton per anion
and, depending on the relative strength of the XH
bond, they are denoted acids or bases ( hydroxides).
Examples of acids include CsHSO 4 and other acidic
salts and solid acids, while bases comprise a range of
hydroxides, oxyhydroxides, hydroxycarbonates, etc.
In crystalline hydrates, acids, and hydroxides, conduction requires the formation of defects such as
proton vacancies or interstitials. These can, in principle, be formed by thermal intrinsic disorder, nonstoichiometry, or aliovalent doping. The latter is
rarely pursued, and instead the best conductors here
have a finite number of protons distributed over a
larger number of equivalent sites, typically occurring
in high-temperature disordered phases, e.g. in
CsHSO 4 .
The above classes of water-containing materials
are vulnerable to decomposition (loss of water) at
high temperature. The tolerance for high-temperature
increases as we reduce the water content. The liquidlike systems often lose too much water even below
1008C at ambient humidity, and most crystalline
hydrates are dehydrated below 2008C. Some solid

T. Norby / Solid State Ionics 125 (1999) 1 11

acids are stable above 2008C, and some hydroxides


and oxyhydroxides are stable to several hundred
centigrade.
Finally, we have materials in which protons are
not part of the nominal structure or stoichiometry,
but are present as foreign species (defects) remaining
from synthesis or in equilibrium with ambient hydrogen or water vapour. These materials thus withstand
high-temperature treatments and dehydration as their
structure is not affected. The introduction of protonic
defects has become an important extension of studies
of defect structures of oxides, and at present also
phosphates, sulfates, halides, sulfides, etc. are receiving increased attention in this respect. In these
materials the protons dissolve as interstitial foreign
species, and their mobility is then not dependent on
native defects. However, acceptor doping may increase the concentration of protons, in competition
with positive native defects like oxygen vacancies.
Many investigations thus focus on the equilibrium
between water vapour, oxygen vacancies and
protons:
H 2 O 1 O xO 1 V ??O 5 2OH ?O ,

of protons (water) from oxides. Thus, proton conductors are generally functional over relatively narrow
temperature ranges.
This becomes evident from Fig. 1, which shows an
overview of proton conductivities for selected materials as a function of temperature in an Arrheniustype plot. The data are extracted from reviews by
Kreuer [35] and other references to be given below.
First, note the high proton conductivity of hydrated
Nafion , almost comparable to aqueous HCl and
liquid H 3 PO 4 , but only utilisable up to around
1008C, where it starts to dehydrate. Fig. 1 further
includes the reported conductivity for an imidazoleintercalated sulfonated polyaromatic polymer (IIS-

(1)

which, in general, is exothermic so that protons rule


at low temperatures and oxygen vacancies at high
temperatures.
Biological as well as geological structures and
processes can exhibit and depend on protonic transport covering all the above classes: water-mediated
transport, crystallographic proton transfer via proton
acceptor and donor sites (or vacancies and interstitials), and, in the case of minerals, also as defect
protons in nominally proton-free compounds.

3. Properties
Many classes of protonic conductors are represented by members which at some temperature
exhibit protonic conductivities of up to 10 23 10 22
S / cm. This typically represents a maximum as a
function of temperature: at higher temperatures the
protonic conduction decreases because of reversible
or irreversible loss of vehicle water (e.g., polymers),
because of decomposition or melting of hydrates,
hydroxides or acid salts, or because of reversible loss

Fig. 1. Selected literature data for proton conductivity as a


function of inverse temperature. IISPAP, imidazole-intercalated
sulfonated polyaromatic polymer; BYSO, Ba 2 YSnO 5.5 . The curve
for Y:BaCeO 3 is a calculated estimate. The interpretation of the
conductivity of Gd:BaPrO 3 as protonic is somewhat ambiguous at
present. For further explanation, discussion and references, see
text.

T. Norby / Solid State Ionics 125 (1999) 1 11

PAP), showing that the temperature tolerance can be


increased, but thus far at the expense of a lower
conductivity. As a representative of a system with
crystallographic protons, CsHSO 4 goes through a
disorder transformation and becomes a fast proton
conductor around 1408C, but melts only somewhat
above 2008C. Finally, we have doped oxides and
phosphates, some of which are decent proton conductors in wet atmospheres from 5 to 6008C and a
couple of hundred degrees higher. At higher temperatures the proton conduction in these materials
usually goes through a maximum and decreases (as
the concentration of proton defects starts to fall) or
electronic or oxygen ion conductivities become too
high. In order to demonstrate the maximum in proton
conductivity, Fig. 1 includes a curve for Y-doped (10
mol%) BaCeO 3 , which was calculated using reported
and estimated values of entropy and enthalpy of
hydration (Eq. (1)) and proton mobility [6,7]. Data
for Ba 2 YSnO 5.5 (BYSO) [4] and Y-doped BaZrO 3
[5] show that these compete in terms of proton
conduction with BaCeO 3 at low temperatures, but
the high-temperature behaviour was not investigated
in any detail. Among the many studies of
Ba 3 Ca 1.18 Nb 1.82 O 8.73 (BCN18), the figure refers to
results where the maximum in proton conductivity is
evident [8].
Results for another perovskite, Ba-doped LaErO 3
[9], and a phosphate, Sr-doped LaPO 4 [10], provide
examples of materials with their maxima at higher
temperatures but without a gain in the proton conductivity. The conductivity for Gd-doped BaPrO 3
was extracted from a study in which there are
indications that the conductivity is protonic [11], but
this needs to be established more unambiguously and
is under investigation.
We may conclude that, at present, solid proton
conductors do not parallel the best oxygen ion
conductors (with conductivities .1 S / cm). However, proton conductors, in general, work at substantially lower temperatures and may offer the highest
conductivities at intermediate and low temperature.
But there are no solid proton conductors working
satisfactorily in the gap between, say, 200 and
5008C, as shown in Fig. 1. While the gap may seem
small in an Arrhenius plot, it covers a most important and desirable range of operating temperatures
for both chemical processes and energy conversion

processes. Narrowing this gap is of prime interest in


the development of proton conductors for practical
applications.

4. Progress
Progress can be measured in many units, but with
reference to the gap in Fig. 1 it is natural to refer
mainly to proton conductivity and operating temperature.

4.1. Water-containing systems: PEMs


For water-containing solid-state proton conductors
for fuel cells it is generally desirable to increase the
thermal stability and thus the working temperature:
this may increase the achievable proton conductivity
and the quality of the waste heat. It may also reduce
the need for costly noble metal catalyst loads.
Proton exchange membranes (PEM) is a commonly used term for polymer-based proton conductors.
The commercially available membrane materials
such as Nafion 117, which have formed the basis for
the success of the PEMs, are perfluorinated copolymers with sulfonated side-branches. These are
expensive to produce and have limited temperature
stability due to dehydration and desulfonation. There
is thus much interest in more stable proton donor
acceptor systems. Equally important are efforts to
produce the polymers at a more competitive price.
Thus, other polymers and synthesis routes are being
investigated. Mainly to bring the unfamiliar reader in
touch with the relevant terminology, we describe a
typical example: poly(vinylidene fluoride) (PVDF)
can be grafted by radiation with gamma-rays, electrons, or heavy ions. In this process the chains are
broken, leaving active sites (radicals). These can
react and branch with polystyrene (PS) which has
been, or can be, sulfonated (i.e. reacted with concentrated sulfuric acid) to add SO 3 H groups which act
as proton donors for water molecules. With sufficient
water (hydration) these protons become mobile. Such
a membrane is typically denoted irradiation-grafted
PVDFPSsulfonic acid, PVDF-g-pssa, etc. [12,13].
While giving the possibility of cheaper production,
the new materials have difficulties in matching the
performance of the commercial PEMs. In a recent

T. Norby / Solid State Ionics 125 (1999) 1 11

review, Kreuer [4] gives a useful introduction to the


principles of PEMs and the problems with the newer,
cheaper ones: the sturdiness of the perfluorinated
polymer backbone is lost in more aliphatic sections
of other polymers. He suggests the use of aromatic
polymers, and further to use, for example, imidazole,
as an alternative to water, intercalated into the
polymers. While having proton donor / acceptor
abilities similar to those of water, the imidazole has a
higher melting point and gives better high-temperature tolerance.

4.2. Low-temperature inorganic proton conductors


Also for crystalline proton-containing materials
progress is measured in terms of operating temperature range. This is to some extent a matter of
substituting cation or anion groups to manipulate
phase stability and in some cases produce supercooled disorder. For instance, a number of phases
reminiscent of CsHSO 4 but with varying CsH
ratios (e.g., Cs 5 H 3 (SO 4 ) 4 ?xH 2 O) have been studied
in the last few years [14], and new results are also
presented in this proceedings volume [1]. Others
have found that the sulfur in SO 22
can be partially
4
and randomly replaced by aliovalent phosphorous,
leading to partially occupied proton sites [15]. The
various materials exhibit what is often termed superprotonic conduction to lower temperature than
CsHSO 4 , but a breakthrough in terms of a practical
electrolyte seems not to have emerged.
For new results on well-known classes of what we
here choose to call low-temperature protonic conductors, such as b0-alumina, 12-tungstophosphoric acid
hydrates, and various zirconium phosphate hydrates,
the reader is referred to contributions in the present
proceedings [1] or in the preceding SSPC proceedings, e.g. a recommended paper is Ref. [16] in the
latter. The present proceedings volume also contains
reports on entirely new materials in terms of proton
conduction.
I use this opportunity to mention that we have
briefly investigated the conductivity of SrHPO 4 ,
Sr(H 2 PO 4 ) 2 and Ba(H 2 PO 4 ) 2 to look for possible
disorder transitions giving increased proton conductivity. However, for SrHPO 4 the conductivity was
below 10 27 S / cm in the investigated range up to
3208C, and for the di-hydrogen phosphates the

conductivities were also low, except after exposure


to a too high temperature, which produced partial
decomposition into hygroscopic and highly conductive phosphoric acid on the grain boundaries and
surfaces.

4.3. High-temperature proton conductors


For high-temperature proton conductors, progress
is a matter of combining high mobility with a high
concentration of protons, but also of avoiding accompanying contributions from oxygen ions and
electrons. Focus has mainly been on the use of
complex perovskites in which charge deficit and
compensating protons are achieved by cation nonstoichiometry rather than normal doping, supposedly
alleviating proton trapping at dopants. Accordingly,
Ba 3 Ca 1.18 Nb 1.82 O 8.73 (BCN18) has been a favourite
subject of study among the high-temperature proton
conductors for the last few years. BCN18 can take
up 0.18 protons per ABO 3 perovskite unit, which is
of the same order of magnitude as typical acceptordoped perovskites. The last couple of years have
seen reports of higher proton contents by hydration
of more oxygen-deficient perovskites, comprising
Ba 2 YSnO 5.5 [17] and Ba 2 InSnO 5.5 [18], which may
take up 0.5 protons per ABO 3 unit, and of Ba 2 In 2 O 5
[19,20], which may take up one proton per ABO 3
unit. These cases may be regarded as inherently
oxygen-deficient structures (perovskite-derived structures with ordered, structural oxygen vacancies) in
which oxygen ions from water may dissolve on the
vacant sites, with charge compensation by protons.
The protons may be disordered, whereby we retain a
protonated perovskite structure, or they may order,
whereby we obtain an oxyhydroxide. Further search
for proton conductors among oxides will surely lead
to the investigation of many new and old oxyhydroxides. This necessitates the development and use of
theory and high-temperature in situ structural and
defect-structural studies of orderdisorder processes,
more or less correlated on the cation, oxygen, and
proton sublattices. These matters are briefly introduced elsewhere [21] and touched upon also in an
experimental study of the complex perovskite
Sr 4 (Sr 2 Nb 2 )O 11 reported in these proceedings [22].
Much remains in the ability to determine experimentally, model, and predict equilibrium proton

T. Norby / Solid State Ionics 125 (1999) 1 11

concentrations in oxides and the thermodynamics


behind them. It will be necessary to control the
microstructure and composition of samples at a
higher level than hitherto and to take into account
different kinds of defectdefect interactions. BaCeO 3
has come to play an important role here as a model
material, and numerous papers have been and are
being published on this oxide, for example considering stability and the effects of composition and
cation non-stoichiometry [4,2325] and modelling of
defect chemistry [26,27]. The latter studies indicate
that the higher basicity of the lower-valent dopants
(such as Gd or Y) normally substituted for the
tetravalent B-site cation (here Ce) increases the
exothermic hydration enthalpy, giving a relatively
higher proton content at high temperatures.
It is interesting to note that a lattice simulation
study of LaGaO 3 indicated that the hydration of this
perovskite is very unfavorable, in fact reporting a
large positive enthalpy for the reaction in Eq. (1)
[28]. In accordance with this, not even traces of
proton transport have been found in acceptor-doped
LaGaO 3 , neither by changes in total conductivity as
a function of water vapour pressure [29] nor by
water vapour concentrationcell type transport number measurements [30]. In contrast, the acceptordoped lanthanum rare-earth perovskites LaScO 3
[31,32] and LaErO 3 [9] are readily hydrated and
exhibit protonic conduction.
Ba 3 Ca 1.18 Nb 1.82 O 8.73 (BCN18) has also become a
much studied material, and a recent report [8]
indicates that this and probably other complex
perovskites contain protons in numbers exceeding
the number of lowest-energy sites such that a second
type of site may have to be used, resulting in a
two-stage hydration process. Otherwise, the tendency
of forming ordered domains of such complex perovskites may give rise to large domain border impedances, and this needs to be addressed carefully in
future studies if these materials are to be used, for
example, as electrolytes in fuel cells.
It was recently reported that Gd-doped BaPrO 3
exhibits very high and presumably protonic conductivity [11] and, all in all, there is seemingly still
room for discoveries among old and new oxide
materials.
Proton conduction at high temperature is also
reported in meta- and polyphosphates [33,34] and in

acceptor-doped LaPO 4 [10,35], but the defect


chemistry of phosphates is an open issue. For
instance, it is uncertain whether or not the introduction of acceptor substituents is accompanied by
phosphate group vacancies during synthesis. If not,
what is then the charge compensation mechanism?
Protons represent one possibility, pyrophosphate
groups (P2 O 742 ) another. Normal oxygen vacancies
are less probable, and the crucial point is that a
possible phosphate deficiency is permanent after
synthesis, unlike an oxygen deficiency of an oxide. It
is possible that a missing phosphate group tends to
be filled with four OH 2 groups. This would be
similar to the four protons believed to compensate
for a missing Si 41 in quartz / silicates and Mn 41 in
MnO 2 , except that in the phosphate the four OH 2
groups would be left with one negative effective
charge.
Muons are short-lived elementary particles of one
positive charge, but with only a fraction of the
protons mass. They expectedly have much in common with protons, and their decay can be used to
monitor their behaviour in the structure, as recently
demonstrated in a study of muons in a perovskite
[36].

4.4. Theory and modelling of proton transport


Nowick and Vaysleyb [37] recently reviewed
isotope effects and concluded that the classical
harmonic oscillator approach of proton transfer in the
OHO system may not rule the ground alone: a
non-classical zero-level effect on the activation
enthalpy is generally observed, and there are details
of the proton transfer by which the heavier isotopic
species may obtain a higher pre-exponential of
diffusion. Furthermore, the number of successful
jumps may be an order of magnitude lower than
predicted classically, much in agreement with experimental results. In order to make advances in the
understanding of proton transfer in many of the
materials we investigate, it appears crucial that the
predictions from isotope effect considerations can be
held against experimental data of sufficient reliability
and accuracy.
There is nevertheless continuous progress in our
ability to model protons and simulate their behaviour
by use of computers, to confer with experimental

T. Norby / Solid State Ionics 125 (1999) 1 11

results, and predict proton conductivities in new


materials. Recently, we have seen quantum mechanical molecular dynamics simulations of short-range
proton motion in CsHSO 4 and in perovskite oxides
[38]. The simulations can now reproduce a small
number of longer jumps (to neighbouring oxygen
ions) at extremely high temperatures, but still not to
the extent that the thermodynamical parameters of
long-range diffusion can be predicted [5]. Modelling
has recently led Kreuer to conclude that doped
BaZrO 3 should be a better protonic conductor than
reported. Suspecting that experimental problems had
obscured this in earlier data, he prepared samples
with carefully controlled composition and microstructure and reports that doped BaZrO 3 exhibits the
highest proton conductivity attained so far among
oxides at low temperatures [5]. These and other
aspects of proton transport in perovskites are treated
in more detail in keynote contributions to the present
proceedings volume [5,39].

5. Prospects

5.1. Systems with proton-conducting electrolytes


hydrogen as an energy carrier
By most measures, the leading commercially
exploited solid proton conductors are the polymerbased proton exchange materials (PEMs). The price
of these and of the Pt catalysts needed has been
almost prohibitive; it appears to have taken a force as
powerful as the automobile market to overcome this
barrier, with promises of mass production deregulating the price. In the last couple of years we have
seen prototype fuel cells taking the step from buses,
via trucks and vans, to medium and small cars. It is
likely that many of us will drive such cars within a
decade, and that also other energy needs in transportation and otherwise will be covered to an
increasing extent by hydrogen and hydrogen-driven
fuel cells [40].
For educational and promotional purposes you can
now buy commercially a complete small-scale hydrogen-power unit [41] which demonstrates solar
cells, water electrolysis by a solid polymer protonconducting membrane, gas storage at atmospheric
pressure or in a metal hydride, and a fuel cell, again

with a solid polymer electrolyte, driving, for example, an electric fan. The unit runs entirely with
distilled water and at room temperature (as it does
not have a very high power density demanding high
conductivity or fast electrode kinetics).
A decade ago, polymers were considered a longterm alternative, but they already lead the way in
terms of solid electrolytes. They will probably
remain competitive in the automobile and other
small-scale segments of fuel cells and electrolysers,
but for larger facilities it is likely that we will need
to employ high-temperature ceramic conductors.
These may be proton conductors, but the lesson to be
learned from the development of zirconia-based solid
oxide fuel cells (SOFC) is that electrode development and materials compatibility must be addressed
as soon as possible.

5.2. Hydrogen separation membranes: mixed


protonicelectronic conductors
5.2.1. Introduction applications
The growing importance of hydrogen will increase
the interest for hydrogen separation membranes. This
comprises the possibility of pumping hydrogen out
of synthesis gas (e.g., from reforming of methane,
see Fig. 2) or possibly in the direct dehydrogenation
of methane or other hydrocarbons or organics.
Finally, such membranes may be used in hydrogen
purification systems. Palladium is an established
material in hydrogen permeable membranes, but has
some limitations. Ceramic hydrogen permeable
membranes with mixed protonicelectronic conduction provide an interesting alternative which
academia has barely shown interest in [42]. It should
be noted that studies of mixed protonelectron
conductors and mixed conductors in general are most
relevant also for the development of electrodes for
proton-conducting electrolytes [43].
5.2.2. Special considerations
The concept of mixed protonicelectronic conductors has of course many aspects in common with the
mixed oxygen ionelectronic conductors, but there
are also important differences. First, consider the
large gradient in chemical potential which we have,
for instance, when we use a mixed oxygenelectron
conductor to separate oxygen from air through a

T. Norby / Solid State Ionics 125 (1999) 1 11

be found among materials with oxidation states not


normally considered stable under near-atmospheric
conditions.
Mixed protonelectron conductors can be developed along different strategies:

Fig. 2. Schematic of the use of a mixed oxygen ionelectronic


conductor for oxygen separation with direct reforming of methane,
followed by the use of a mixed protonicelectronic conductor for
hydrogen extraction. The products are thus pure hydrogen and
synthesis gas with reduced hydrogen content, the latter suitable
for, for example, FisherTropsch synthesis of methanol.

membrane directly exposed to methane. This is not


available in hydrogen separation processes, unless
hydrogen can be reacted through the protonelectron
conductor with a reactant giving a stable and valuable product. Most processes will instead involve
hydrostatic pumping or pressing hydrogen through,
with one or at best a few atmospheres in pressure
difference as driving force.
Perhaps more important is the difference in the
conditions the membrane will work under. The
oxygen permeable membrane will be exposed to
either oxidising vs. reducing conditions (e.g., air vs.
methane) or to conditions where both atmospheres
are relatively oxidising (in the hydrostatic pumping
mode). The hydrogen permeable membrane will
most likely find use with reducing conditions on both
sides. When, in addition, we may have to use mixed
valency elements to obtain sufficiently high electronic conduction, candidate materials are likely to

5.2.3. High-temperature, acceptor-doped systems


with mixed protonicp-type or n-type electronic
conduction
Traditionally, high-temperature proton conductors
are acceptor-doped oxides exhibiting mixed conduction from protons, oxygen vacancies and electron
holes. Rare-earth-doped BaCeO 3 is a good example,
exhibiting large contributions of all three conductivities under high humidities, high oxygen partial
pressures, and at high temperature [24]. The p-type
electronic conductivity can be of different magnitude
in different oxides, depending on the valence states
of the cations involved, but usually it decreases
rapidly with decreasing oxygen activity. Thus, most
of these materials will not be suitable as mixed
conductors under reducing conditions. However,
some of them exhibit an electronic n-type conduction
under sufficiently reducing conditions, despite the
acceptor doping which suppresses the concentration
of defect electrons. The oxides with remaining ntype conduction will be those which contain a readily
reducible cation. Thus, the n-type conduction is
relatively high in BaCeO 3 and other cerates, while
almost negligible in the zirconates [44]. There is
currently some interest in alkaline earth
praseodymates and terbates (e.g., BaPrO 3 and
BaTbO 3 ) and these may be interesting in terms of
n-type conduction because of their reducibility.
Lately, Shimura et al. [45] have found proton
conduction in Sr 2 TiO 4 , a material of potential interest as a mixed conductor due to its high basicity and
the reducibility of Ti 41 . In the present proceedings
an attempt to increase mixed conduction in this
material by addition of mixed valency cations such
as Fe or Co is reported [46].
If the p-type conduction of acceptor-doped oxides
is to be utilised under reducing conditions, one may
have to employ compounds with a cation in a
relatively stable, yet oxidisable, valency under reducing conditions. By way of illustration, the wellknown Ca- or Sr-doped LaCrO 3 is a mixed p-type
and oxygen ion conductor even under reducing

T. Norby / Solid State Ionics 125 (1999) 1 11

conditions due to the relatively stable 31 valency of


chromium. However, proton conduction is not found
in this material (by our own emf measurements at
90010008C in wet hydrogen). Potential materials of
interest for future studies could be the relatively
ambivalent rare earth oxides Pr 2 O 3 and Tb 2 O 3 or
31
compounds including these. Pr
is found as an
A-site cation in perovskites, but electronic conduction in perovskites is generally attributed to OB
OB pathways, and a mixed valency A-cation may
be inefficient. Tb 31 could possibly enter as a B-site
cation (although, for example, LaTbO 3 might not
take on the perovskite structure). Many other possibilities exist, such as Ti 31 , Ce 31 , and reduced states
of V, Mo, Mn, etc., especially if compounds with
other structures than the perovskite are considered.
Another problem with the acceptor-doped oxides
is the tendency of high oxygen ion conduction
accompanying the high proton conductivity. The
alkali earth zirconates and a few rare earth sesquioxides have been shown to have ionic conduction
which is relatively purely protonic, but the protonic
conductivities are not sufficient for high throughput
applications in these materials. Most of the materials
investigated for high-temperature proton conduction
are perovskites or related structures with a tendency
to exhibit high oxygen vacancy mobilities. This will
be detrimental in many applications of hydrogen
permeable membranes, and it is of interest to consider other structure types. While many close-packed
types such as spinels may possess mixed valency and
electronic conduction, there are no indications of
high protonic transport in these, probably due to the
relatively rigid oxygen ion host sublattice. So what is
needed may be a more dynamic oxygen sublattice,
but without the possibility of long-range oxygen
transport. This is what we have in, for instance,
CsHSO 4 with its rotating SO 4 tetrahedra passing on
protons without exchanging oxygen ions. Can this be
transposed to high-temperature materials? LaPO 4 has
PO 4 octahedra where each oxygen has three nearest
neighbours (within the tetrahedron). Oxygen is covalently strongly bonded to the acidic phosphorous
central atom and is not expected to move easily
outside the tetrahedron (and oxygen is accordingly
not considered very mobile in phosphates). In agreement with this it appears that Sr-doped LaPO 4
possesses protonic conduction and some p-type

electronic conduction, but no oxygen ion transport


[10,35]. It might be of interest to investigate other
materials with related structures, i.e. acidic central
atoms with tetrahedrally coordinated oxygen. At
ambient oxygen activities this would, for instance,
comprise alkaline earth tungstates (AWO 4 ) and rare
earth vanadates, niobates, and tantalates (LnBO 4 ).
During a brief survey we did indeed find indications
that the conductivity of acceptor-doped BaWO 4 is
protonic, although not high [47]. Under reducing
conditions, reduced forms of the central atoms may
have to be considered, e.g. Mo 51 , but as the options
here are quite few, the use of tetrahedra-based
structures may still prove more useful for the development of pure proton-conducting electrolytes for
use as sensors under oxidising conditions.

5.2.4. Reduced materials with mixed protonic and


n-type electronic conduction
Some materials dissolve protons with charge
compensation by defect electrons, given a sufficiently high hydrogen activity; we may regard the hydrogen as donor dopants. (Other donor dopants will not
be beneficial here since they will decrease the proton
content.) In principle, these materials will be mixed
protonelectron conductors useful under reducing
conditions. A well-known example of this defect
situation is found in ZnO [48], but the protonic
mobility is probably not sufficient in this case. Other
candidates are reducible oxides without a too strong
tendency of forming oxygen vacancies, comprising
TiO 2 and corundum-type sesquioxides such as
Cr 2 O 3 , Fe 2 O 3 and Mn 2 O 3 and more complex oxides
with these and other cations.
5.2.5. Materials with high intrinsic or metallic
electronic disorder
In the above cases we have considered materials,
mainly oxides, where we know or can anticipate a
certain proton conductivity, and have discussed how
this can be combined with an electronic conductivity.
On the other hand, many oxides have intrinsically
high or even metallic electronic conductivities. A
few of these have high oxygen conductivities (e.g.,
LaCoO 3 - or SrFeO 3-x -based materials) and are thus
of little interest as mixed proton conductors. Others,
however, are known or assumed to have negligible
ionic conductivities, and it would be of interest to

10

T. Norby / Solid State Ionics 125 (1999) 1 11

map protonic transport in some of these. Even with


high proton transport, the protonic transport number
may be small in a metallic conductor, and electrical
methods may then not be applicable. Instead, diffusivity measurements may have to be used, for
example SIMS analyses of tracer-annealed samples
or steady-state measurements of the hydrogen flux
through a membrane. Incoherent neutron scattering
(INS) appears to be useful for characterising proton
dynamics in materials with high protonic and electronic disorder [49].
The experimental difficulty of directly measuring
proton conductivity calls for the modelling of proton
solubility and transport in highly electronically conducting materials. This has been attempted for
LaMnO 3 [50]: the values obtained predict that
acceptor-doped LaMnO 3 may dissolve significant
amounts of protons in wet atmospheres at temperatures which are fairly low, but where equilibrium
may still be achieved in practice. We are trying to do
this by thermogravimetry under reducing, wet conditions, but without success so far.

5.2.6. Materials with water or crystallographic


protons
Materials with crystallographic protons may be
mixed protonicelectronic conductors by many
combinations of defect disorder, but all require a
tendency towards mixed valency and proton defects
(interstitials, vacancies, or disordered intrinsic deficiencies, as mentioned before). As an example, an
oxyhydroxide may be acceptor-doped to give interstitial protons (random, extra hydroxide groups on an
oxygen ion sublattice) and electron holes, or donordoped to give proton vacancies and electrons. Alternatively, the material may be reduced to give interstitial protons and electrons, or oxidised to give
proton vacancies and electron holes. There are vast
possibilities, but the Ni(OH) 2 NiOOH system may
provide a simple illustration [51]: the electrochemical oxidation of Ni(OH) 2 can, for instance, be
written
Ni(OH) 2 1 OH 2 5 NiOOH 1 H 2 O 1 e 2 ,

(2)

and is a reaction utilised at the Ni electrodes of


certain accumulators. Chemically, the oxidation
could well be written

Ni(OH) 2 1 1 / 4O 2 5 NiOOH 1 1 / 2H 2 O,

(3)

but for our purposes we only want to indicate that


oxidation of Ni(OH) 2 creates defects reminiscent of
NiOOH, while reduction of NiOOH creates defects
reminiscent of Ni(OH) 2 . Depending on the conditions and defect levels (nonstoichiometries) allowed, each of these phases may thus, in principle,
be a mixed protonicelectronic conductor.

5.2.7. Hydrogen insertion compounds


Hydrogen and protonic species may be intercalated into a number of materials. Many of these may
be of present interest as insertion or storage electrodes for secondary batteries. The above example of
the Ni hydroxides can be considered such a case,
with the hydrogen (or proton) as the intercalated
species. Of immense interest for Li-ion batteries is
the reductive Li intercalation into MnO 2 to form a
LiMn 2 O 4 spinel. However, the Li ions can be
exchanged with protons [52] and the transport of
protonic species is apparently fast enough at ambient
temperatures that the prospects of hydrogen permeable membranes of this type of materials are interesting. Other examples comprise the so-called hydrogen
bronzes obtained when hydrogen is inserted into
WO 3 , MoO 3 , and other transition metal oxides
[53,54]. As mentioned above, incoherent neutron
scattering (INS) studies [49] are particularly useful
for characterising the proton dynamics of such
intercalation compounds and other compounds with
high proton and electronic disorder and indicate a
number of materials of potential interest for hydrogen storage and / or permeable membranes.

6. Concluding remarks
The growing importance of cleaner energy introduces hydrogen as a carrier and increases the need
for materials with higher proton conductivity for fuel
cells, electrolysers, and electrochemical reactors, for
materials with purer proton conductivity for hydrogen sensors, and for materials with high mixed
protonicelectronic conduction for hydrogen separation membranes. An improved understanding of
protons and proton transport in general is important
and increasingly acknowledged also for a number of

T. Norby / Solid State Ionics 125 (1999) 1 11

other applications (hydrogen and water barriers,


protection of metals and alloys, etc.) and scientific
fields (geosciences, biochemistry, etc.).

References
[1] G. Lahajnar, R. Blinc (Eds.), Proceedings of the International
Conference on Solid State Protonic Conductors, Bled,
Slovenia,1998 (special issue), Solid State Ionics (this volume).
[2] M. Spaeth, K.D. Kreuer, Th. Dippel, J. Maier, Solid State
Ionics 97 (1997) 291.
[3] K.-D. Kreuer, Chem. Mater. 8 (1996) 610.
[4] K.-D. Kreuer, Solid State Ionics 97 (1997) 1.
[5] K.-D. Kreuer, Solid State Ionics (this volume).
[6] K.-D. Kreuer, Th. Dippel, Yu.M. Baikov, J. Maier, Solid
State Ionics 8688 (1996) 613.
[7] T. Norby, Y. Larring, Curr. Opin. Solid State Mater. Sci. 2
(1997) 593.
[8] B. Gross, St. Marion, R. Hempelmann, D. Grambole, F.
Herrmann, Solid State Ionics 109 (1998) 13.
[9] Y. Larring, T. Norby, Solid State Ionics 70 / 71 (1994) 305.
[10] T. Norby, N. Christiansen, Solid State Ionics 77 (1995) 240.
[11] T. Fukui, S. Ohara, S. Kawatsu, J. Power Sources 71 (1998)
164.
[12] S.D. Flint, R.C.T. Slade, Solid State Ionics 97 (1997) 299.
[13] D.I. Ostrovski, L.M. Torell, M. Paronen, S. Hietala, F.
Sundholm, Solid State Ionics 97 (1997) 315.
[14] A.I. Baranov, V.V. Sinitsyn, V.Yu. Vinnichenko, D.J. Jones, B.
Bonnet, Solid State Ionics 97 (1997) 153.
[15] S.M. Haile, P.M. Calkins, D. Boysen, Solid State Ionics 97
(1997) 145.
[16] Ph. Colomban, J. Tomkinson, Solid State Ionics 97 (1997)
123.
[17] P. Murugaraj, K.-D. Kreuer, T. He, T. Schober, J. Maier,
Solid State Ionics 98 (1997) 1.
[18] T. Schober, Solid State Ionics 109 (1998) 1.
[19] T. Schober, J. Friedrich, F. Krug, Solid State Ionics 99
(1997) 9.
[20] W. Fischer, G. Reck, T. Schober, Solid State Ionics 116
(1999) 211.
[21] T. Norby, Korean J. Ceram. 4 (1998) 128.

[22] R. Glockner,
A. Neiman, Y. Larring, T. Norby, Solid State
Ionics (this volume).
[23] D. Shima, S.M. Haile, Solid State Ionics 97 (1997) 443.
[24] J. Guan, S.E. Dorris, U. Balachandran, M. Liu, J. Electrochem. Soc. 145 (1998) 1780.
[25] Z. Wu, M. Liu, J. Electrochem. Soc. 144 (1997) 2170.

[26] K.D. Kreuer, W. Munch,


M. Ise, T. He, A. Fuchs, U. Traub,
J. Maier, Ber. Bunsenges. Phys. Chem. 101 (1997) 1344.

[27] R. Glockner,
M.S. Islam, T. Norby, Solid State Ionics 122
(1999) 145.

11

[28] S. Khan, M.S. Islam, D.R. Bates, J. Phys. Chem. B 102


(1998) 3099.
[29] M. Feng, J.B. Goodenough, Eur. J. Solid State Inorg. Chem.
31 (1994) 663.
[30] E.M. Ottesen, Cand. Scient. Thesis, Department of Chemistry, University of Oslo, 1998.
[31] H. Fujii, Y. Katayama, T. Shimura, H. Iwahara, J. Electroceram. 2 (1998) 119.
[32] D. Lybye, N. Bonanos, Solid State Ionics (this volume).
[33] M. Greenblatt, P.P. Tsai, T. Kodama, S. Tanase, Solid State
Ionics 40 / 41 (1990) 444.
[34] M. Cappadonia, S. Haufe, O. Niemzig, U. Stimming, Solid
State Ionics (this volume).
[35] K. Amezawa, S. Kjelstrup, T. Norby, Y. Ito, J. Electrochem.
Soc. 145 (1998) 3313.

[36] R. Hempelmann, M. Soetratmo, O. Hartmann, R. Wappling,


Solid State Ionics 107 (1997) 269.
[37] A.S. Nowick, A.V. Vaysleyb, Solid State Ionics 97 (1997) 17.

[38] W. Munch,
G. Seifert, K.-D. Kreuer, J. Maier, Solid State
Ionics 97 (1997) 39.
[39] A.S. Nowick, Yang Du, K.C. Liang, Solid State Ionics (this
volume).
[40] T.O. Saetre, in: Proceedings of HYPOTHESIS II, Grimstad,
Norway, 1997, Hydrogen Power: Theoretical and Engineering Solutions, Kluwer, Dordrecht, 1998.
[41] H-TEC Wasserstoff-Energie-Systeme GmbH, Lindenstrasse

48a, D-23558 Lubeck.


[42] J. Guan, S.E. Dorris, U. Balachandran, M. Liu, Solid State
Ionics 100 (1997) 45.
[43] W.L. Rauch, M. Liu, J. Electrochem. Soc. 144 (1997) 4049.
[44] J.B. Smith, Cand. Scient. Thesis, Department of Chemistry,
University of Oslo, 1998.
[45] T. Shimura, K. Suzuki, H. Iwahara, Solid State Ionics 104
(1997) 17.
[46] T. Shimura, K. Suzuki, H. Iwahara, Solid State Ionics (this
volume).
[47] T. Norby, unpublished results.
[48] D.G. Thomas, J.J. Lander, J. Chem. Phys. 25 (1956) 1136.
[49] F. Fillaux, Solid State Ionics (this volume).
[50] M. Cherry, M.S. Islam, J.D. Gale, C.R.A. Catlow, J. Phys.
Chem. 99 (1995) 14614.

[51] R.A. Huggins, H. Prinz, M. Wohlfahrt-Mehrens, L. Jorissen,


W. Witschel, Solid State Ionics 70 / 71 (1994) 417.
[52] B. Ammundsen, P.B. Aitchison, G.R. Burns, D.J. Jones, J.
Solid State Ionics 97 (1997) 269.
Roziere,
[53] P.G. Dickens, A.M. Chippindale, in: Ph. Colomban (Ed.),
Proton Conductors. Solids, Membranes and Gels Materials and Devices, Chemistry of Solid State Materials, Vol. 2,
Cambridge University Press, Cambridge, 1992, Chapt. 7.
[54] R.C.T. Slade, in: Ph. Colomban (Ed.), Proton Conductors.
Solids, Membranes and Gels Materials and Devices,
Chemistry of Solid State Materials, Vol. 2, Cambridge
University Press, Cambridge, 1992, Chapt. 29.

Você também pode gostar