Sunteți pe pagina 1din 32

4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.

sgm LaTeX2e(2002/01/18) P1: IKH


10.1146/annurev.matsci.33.022802.093856

Annu. Rev. Mater. Res. 2003. 33:581–610


doi: 10.1146/annurev.matsci.33.022802.093856
Copyright °c 2003 by Annual Reviews. All rights reserved
First published online as a Review in Advance on April 1, 2003

UNDERSTANDING MATERIALS COMPATIBILITY


Harumi Yokokawa
Energy Electronics Institute, National Institute of Advanced Industrial Science and
Technology, Tsukuba, Ibaraki 305-8565, Japan; email: h-yokokawa@aist.go.jp

Key Words solid oxide fuel cell, thermodynamic stability, chemical potential,
interface stability
■ Abstract The use of chemical potential diagrams to examine interface chemistry
is discussed in terms of the chemical reactions among oxides and associated interdif-
fusion across the interface. The driving force for both processes can be determined
from the chemical potential values. The geometrical features of the chemical potential
diagrams can be related to the valence stability of binary oxides and the stabilization en-
ergy of double oxides from the constituent oxides. The materials compatibility in solid
oxide fuel cell materials is discussed with a focus on a lanthanum manganite cathode
and a yttria-stabilized zirconia (YSZ) electrolyte. Emphasis is placed on the valence
numbers of manganese in the fluorite solid solution and the perovskite oxides, which
have been derived by thermodynamic analysis of the magnitude of the stabilization
energy/interaction parameters as a function of ionic size for respective valence num-
bers. The change of manganese valence on La2Zr2O7 formation and Mn dissolution in
YSZ are discussed in relation with the oxygen evolution/adsorption process. Oxygen
flow associated with electrochemical reactions exhibits markedly different features de-
pending on the direction of the polarization, which can lead to drastic changes in the
interface chemistry (precipitation or interdiffusion).

INTRODUCTION
Solid oxide fuel cells (SOFCs), which convert chemical energy (fuels) to elec-
tricity, consist of an electrolyte, two electrodes, interconnects, and other relevant
materials (1). SOFC stacks are fabricated around 1200–1400◦ C. Their desired
operating time is more that 50,000 h at ∼1000◦ C. This implies that the mate-
rial’s chemical behavior (2, 3) at such high temperatures is crucial for obtaining
good performance and high reliability, in addition to the materials’ electrochem-
ical properties that are required for the proper functioning of the electrochemical
cells. Chemical compatibility with other materials and the environment (fuels or
air) must be taken into account in designing SOFCs. For this purpose, thermo-
dynamic considerations should be made properly. Advances have been made in
constructing thermodynamic databases (4–6, 6a,b) and developing computer soft-
ware for calculating complicated chemical equilibria (7) in order to understand the
complex chemical processes at high temperatures. Although chemical equilibria
0084-6600/03/0801-0581$14.00 581
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

582 YOKOKAWA

calculations are powerful, they do not provide a good basis for examining the in-
terface stability or interdiffusion across the interfaces among the cell components.
This requires a thermodynamic analysis of equilibrium at the interfaces by using
a generalized chemical potential diagram (8).
Stack development of SOFCs with oxide interconnects can be categorized as
the first generation of SOFCs (3, 9). In this development, thermodynamic consid-
erations played important roles in obtaining an optimized solution to surmount
materials problems. A similar approach can be applied to other high-temperature
processes such as alkali metal thermo-electric converters (AMTEC) (10) or molten
carbonate fuel cells (11) and also to some low-temperature devices such as lithium
batteries (12).
In this article, we review the recent advance in understanding materials com-
patibility in some high-temperature SOFCs. The thermodynamic approach to
understand materials compatibility is discussed mainly in terms of the chemi-
cal potentials as a key concept in considering the chemical features of interfaces.
Among the SOFC materials and their interfaces, air electrode/electrolyte interfaces
are selected to examine in detail, using chemically useful quantities such as valence
stability, stabilization energy in complex oxides, and related acid/base properties.
Other topics associated with intermediate SOFCs, which can be categorized as
second generation SOFCs, are also discussed briefly.

HOW TO UNDERSTAND MATERIALS COMPATIBILITY

Chemical Potential and Interfacial Chemistry


In interface systems, many of the most important chemical phenomena that occur
involve precipitation of new phases at the interfaces and interdiffusion across the
interfaces (13).
Diffusion is usually discussed in terms of a concentration gradient. Strictly
speaking, however, diffusion should be treated in terms of chemical potential be-
cause the chemical potential gradient provides the true driving forces for diffusion.
Chemical potential is defined as
µ ¶
∂G
µi = , 1.
∂n i j6=i

where G is the Gibbs energy of a system consisting of several species, and ni is the
mole number of species i. Here, chemical potential is defined as the differentiation
of the Gibbs energy with respect to the mole number of species i, without changing
the mole number of other species. In an equilibrium state, chemical potential
species are defined as constant throughout the system. To properly treat diffusion,
a local equilibrium approximation is adopted; the system is divided into small areas
in which the thermodynamic quantities are defined as constant. From Equation 1,
it is easily understood that when a small amount of species i is taken out of one
phase, where its chemical potential is high, and placed on its adjacent point having
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 583

the lower value of chemical potential, the total Gibbs energy becomes lower as a
result of two processes. This explains why the chemical potential gradient gives
the true driving force for diffusion.
Interdiffusion across the interface cannot be treated in terms of concentration
in any case. Again, the chemical potential provides a basis for examining mass
transfer across the interface or the precipitation of new phases. Under the local
equilibrium approximation, chemical potential can be defined separately for two
adjacent phases. Even when the chemical potential has the same value for two
phases, interdiffusion can have taken place when there are chemical potential
gradients in the same direction across the interface.
New phase precipitation is usually discussed in terms of the Gibbs energy
change for the proposed precipitation reactions. All possible chemical reactions
should be counted and examined by using available thermodynamic data on the
Gibbs energy change for formation. For example, the following possible reactions
can be written for the interface between ZrO2 and LaCoO3;

LaCoO3 + ZrO2 = 0.5 La2 Zr2 O7 + CoO + 0.25 O2 (g), 2.


LaCoO3 + ZrO2 = 0.5 La2 Zr2 O7 + 1/3 Co3 O4 + 1/12 O2 (g). 3.

When some of those reactions have a negative Gibbs energy change for reaction,
new phases can be precipitated. However, when there are many thermodynami-
cally possible reactions, it is difficult to examine, from the Gibbs energy alone,
which kinds of reactions do take place and how they proceed. Note that a similar
examination can be made in terms of the chemical potential values in a more rea-
sonable manner. When the chemical potential values of La, Co, Zr, and O at the
LaCoO3/ZrO2 interface are determined as µ(La)∗ , µ(Co)∗ , µ(Zr)∗ , and µ(O)∗ , the
precipitation reaction of La2Zr2O7 can be determined in the following quantities:

1G = µ◦ (La2 Zr2 O7 ) − [2µ(La)∗ + 2µ(Zr)∗ + 7µ(O)∗ ]


= 1f G ◦ (La2 Zr2 O7 ) − {2[µ(La)∗ − µ(La)◦ ]
+ 2[µ(Zr)∗ − µ(Zr)◦ ] + 7[µ(O)∗ − µ(O)◦ ]}. 4.

Because the chemical potentials are key properties in this analysis, they are
easily extended to combine with diffusion. When several phases can be precipi-
tated, chemical potentials provide the basis for examining how those phases are
arranged across the interfaces (8).
The next question then is, how are the chemical potentials determined at the
interfaces? In alloy systems, it is not difficult to imagine that the elements have
their respective chemical potential values. In a system containing stoichiometric
compounds, however, there remains some ambiguity in determining the chemical
potentials. This problem is discussed further below.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

584 YOKOKAWA

Construction of Chemical Potential Diagram


The equilibrium state can be derived by several methods; one involves the chemical
equilibrium calculations (14) based on the Gibbs energy minimization technique.
When the mole numbers of elements are given at a constant temperature and
pressure, the equilibrium state can be defined as the state in which the total Gibbs
energy of the system becomes the minimum. Many computation programs for
calculating complicated chemical equilibria have been developed in the last two
decades on the basis of this technique (7).
The same information as provided by the Gibbs energy minimization can be
derived in the chemical potential space by adopting the duality theorem (see
Appendix C in Reference 8). In chemical potential space, the equilibrium state,
for example in the La-Co-O ternary system, can be defined as follows; from a
computer-geometrical point of view, a stable compound, LaCoO3, in the La-Co-O
ternary system can be represented by the following equation for a plane in chemical
potential space:
µ(La) + µ(Co) + 3µ(O) = µ◦ (LaCoO3 ), 5.
where the slope of the plane is determined from the stoichiometric numbers, [1,1,3],
whereas the location of this plane relative to the origin of the space is determined
from µ◦ (LaCoO3), which is easily derived form the Gibbs energy change for
formation (see below).
All stable compounds form a polyhedron, as shown in Figure 1 for the La-Co-
O system at 1273 K. Different compounds have different stoichiometric numbers
and, therefore, different slopes on the faces of the polyhedron. Thus polygonal
facets correspond to stable compounds, whereas edges and corners correspond
to two-phase and three-phase boundaries, respectively. The dual theory indicates
that when mole numbers are given for respective elements in a ternary system, the
equilibrium state can be obtained from the condition that a plane having the slope
corresponding to the given molar ratio is maximized within the polyhedron (8).
When this plane is touched to a corner, an edge, or a facet, the equilibrium state
can be derived as three-phase, two-phase, and single-phase state.
Their features correspond well to those in the triangle compositional diagram as
shown in Figure 1. Borderlines of polygonal phases represent two-phase tie lines
in the triangle compositional diagram.
When compared with the conventional chemical potential diagrams, the poly-
hedron shown in Figure 1 and its derivative can be regarded as a generalized

−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−→
Figure 1 Theremodynamic phase relations for the La-Co-O system at 1273 K. (a)
Polyhedron in the three-dimensional chemical potential space. Respective planes are
represented by planes whose slopes and location are determined from the stoichiometric
numbers and the Gibbs energy changes for formation, respectively. (b) Phase relations
in a normal triangle composition diagram.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 585


4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

586 YOKOKAWA

chemical potential diagram in the following sense. (a) There is no need to specify
a redox element, which makes it possible to treat alloys and double oxides in a
proper manner, (b) and there is no limitation of the coordinate axes to the environ-
mental variables. Elemental chemical potentials are used as coordinate axes. This
implicitly indicates that the diagram coordinates give the driving force for mass
transfer inside condensed phases as well as through gas phases. Diffusion in con-
densed phases is driven by the chemical potential difference, which is represented
by two different points in the chemical potential polyhedron. In the dilute solution
approximation, the chemical potential of impurities such as La in Co can directly
correspond to the concentration of impurities.

µ(La) = µφ (La) + RT ln c(La in Co), 6.

where µφ (La) is the value for the Henrian reference state. The chemical potentials
of the major components of the stoichiometric compound can also change inside
the stability polygon, depending on the energetic state of their related defects.
In LaCoO3, the concentration of oxide ion vacancies can be given as a function
of oxygen potential. Cation defects are usually negligible in the B-site in the
perovskite lattice. This implies that the difference in the µ(La) and µ(Co) inside
the stability field of LaCoO3 corresponds to those in concentration of the A-site
vacancies. To describe these features quantitatively, modeling based on defect
chemistry is required. When concentrations of defects are significant beyond the
stoichiometric approximation, the phase should be treated as a mixture; in such a
case, the slope of the plane in the chemical potential space becomes curved.
In practical materials applications where sophisticated doping techniques are
frequently adopted, phase relations in multicomponent systems are required to
account for the chemical behavior of dopants. In chemical equilibrium calculations,
the complication caused from increasing the number of elements can be well treated
without any difficulty. On the other hand, the chemical potential diagram has some
weak points in displaying phase relations in multicomponent systems. Even so, the
generalized chemical potential diagram provides a better solution for constructing
diagrams for such systems. By setting some of chemical potentials at selected
values, appropriate displaying can be made to some extent (8).
Figure 2 gives an example for the phase relations in the La-Zr-Co-O system in
order to examine interface stability. When the oxygen potential is fixed at 1 atm, the
thermodynamic state of binary oxides is represented with the metallic chemical po-
tential. In the La-Zr-Co-O system, the three elemental chemical potentials, µ(La),
µ(Zr), andµ(Co), can be used to set up the three-dimensional chemical potential
diagram. Two such diagrams are shown at (a) 1273 and (c) 1073 K; note that the
Gibbs energy change for Equation 2 is negative at 1273 K and positive at 1073 K.
Geometrical features in Figure 2 exactly correspond to the Gibbs energy values.
At higher temperatures (1273 K), LaCoO3 reacts with ZrO2 to form La2Zr2O7 and
CoO. Reaction experiments (14a) confirmed that the products are arranged in a
layer structure: LaCoO3/CoO/La2Zr2O7/ZrO2. From this sequence, a possible reac-
tive diffusion path can be drawn (Figure 2). The same phase relation and associated
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 587

Figure 2 The three-dimensional chemical potential diagram (a,c) and corresponding


composition diagram (b,d) for the La-Zr-Co-O system at p(O2) = 1 atm. (a,b) 1273
K and (c,d) 1073 K. Reaction product arrangement, LaCoO3/CoO/La2Zr2O7/ZrO2, is
represented by a continuous line in (a) and (b).
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

588 YOKOKAWA

Figure 2 (Continued )
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 589

reactive diffusion path are also shown in the triangle composition diagram
(Figure 2b,d ). It is clear that the reactive diffusion path is properly represented in
the chemical potential diagram.

Which Thermodynamic Properties


Dominate the Chemical Potential?
The chemical potentials are practically derived from the Gibbs energy related func-
tions. For a stoichiometric compound, the chemical potentials have the following
relations with the Gibbs energy;

1f G ◦ (LaCoO3 ) = µ◦ (LaCoO3 ) − [µ◦ (La) + µ◦ (Co) + 1.5µ◦ (O2 )] 7.


G ◦ (LaCoO3 ) = µ◦ (LaCoO3 ) − (µ◦La + µ◦Co + 1.5µ◦O2 ). 8.

Here, µ◦ (La) and µ◦ La are the chemical potentials of element La in the refer-
ence state at a given temperature and at the standard temperature (298.15 K),
respectively. 1fG◦ (LaCoO3) is the standard Gibbs energy change for formation
and depends on how the reference state is selected at high temperatures for each
element. G◦ (LaCoO3) is the Gibbs energy relative to the reference sate of elements
at 298.15 K, which has been used by Barin (15). By combining Equations 7 and 8
with Equation 5, the basic equations for constructing chemical potential diagrams
can be correlated with Gibbs energies.
Through the Gibbs energy values, the chemical potentials are also closely related
with enthalpy and entropy. Enthalpy is correlated directly with the energetics
of the complex oxides, whereas entropy is related mainly with the redox of the
constituent oxides. Both functions are important when macroscopic behavior such
as chemical reactions or diffusion are discussed in terms of the physicochemical
properties.
Because there are large differences between gas and condensed phases in en-
tropy values, valence changes accompanied with the evolution or absorption of
oxygen gas usually exhibit a large entropy change. Note, however, that the entropy
change per evolved oxygen mole number is about the same regardless of the binary
oxide or complex oxides (16). In contrast, the enthalpy change on valence change
is strongly affected by the ionic configuration in complex oxides. These features
can be well characterized in terms of the valence stability for binary oxides, and
the stabilization energy of double oxides from constituent oxides can be defined
as

MO0.5 n + 0.25 (m − n)O2 (g) = MO0.5 m ,


1[Mn+ : Mm+ ] = 1f G ◦ (MO0.5 m ) − [1f G ◦ (MO0.5 n )
+ 0.25(m − n)1f G ◦ (O2 )], 9.

AO1.5 + BO1.5 = ABO3 ,


δ(ABO3 ) = 1f G ◦ (ABO3 ) − [1f G ◦ (AO1.5 ) + 1f G ◦ (BO1.5 )]. 10.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

590 YOKOKAWA

By using the valence stability and the stabilization energy, the Gibbs energy change
for Equation 2 can be rewritten as follows;

1G(Equation 2) = 0.51f G ◦ (La2 Zr2 O7 ) + 1f G ◦ (CoO) + 0.251f G ◦ (O2 )


−[1f G ◦ (LaCoO3 ) + 1f G ◦ (ZrO2 )]
= 0.5[1f G ◦ (La2 Zr2 O7 ) − 1f G ◦ (La2 O3 ) − 21f G ◦ (ZrO2 )]
−{1f G ◦ (LaCoO3 ) − 0.5[1f G ◦ (La2 O3 ) − 1f G ◦ (Co2 O3 )]}
−{0.51f G ◦ (Co2 O3 ) − [1f G ◦ (CoO) + 0.251f G ◦ (O2 )]}
= 0.5 δ(La2 Zr2 O7 ) − δ(LaCoO3 ) − 1[Co3+ : Co2+ ]. 11.

Figure 1 shows that the cobalt trivalent ions are strongly stabilized in the
perovskite phase, which widens the stability field of LaCoO3 down to the lower
oxygen potential. In the presence of ZrO2, however, La2Zr2O7 phase is stabi-
lized similarily so that 0.5 δ(La2Zr2O7) is cancelled out with δ(LaCoO3). As a
result, the cobalt trivalent ions no longer are stable and CoO is formed
instead.
The features of stabilization energy and valence stability can be also correlated
with geometrical features in the chemical potential diagram shown in Figure 2.
For example, the magnitude of δ(La2Zr2O7) is represented by the location of a
La2Zr2O7 plane relative to those for La2O3 and ZrO2. Because δ(La2Zr2O7) does
not show strong temperature dependence, the relative location of the La2Zr2O7
plane does not change between 1273 and 1073 K. In other words, the width of
the stability polygon of La2Zr2O7 between La2O3 and ZrO2 is about the same.
However, the width of LaCoO3 between La2O3 and CoO becomes narrower at
1273 K than at 1073 K, which is the result of valence stability, 1[Co3+ : Co2+]
because δ(LaCoO3) is essentially temperature independent.

Valence Stability: Key Role of Oxygen


In many cases, the valence stability for binary oxides behaves in a complicated
manner because the enthalpy term is determined as the energy difference among
metallic substances, ionic substances, and covalent molecules. Thus it is difficult
to find a good correlation between the enthalpy change for formation and fun-
damental properties such as ionic radius. Fortunately, valence stability has been
experimentally determined and is available in various databases (6, 17). On the
other hand, the stabilization energy of double oxide from the constituent oxides
is not large compared with valence stability and exhibits good correlation with
ionic radii and other ionic properties (18–20) because the stabilization energy of
double oxides originates only from the energy difference between several ionic
substances and can be well characterized in terms of the ionic packing features for
respective crystals.
Characterization of reactions in terms of valence stability and stabilization
energy makes it easy to discuss reactivity in terms of the valence, the ionic size,
and the coordination number or other lattice environments.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 591

Whenever valence is changed in the oxide systems, oxygen has to be involved


in the chemical process and thus is one of the most important reactants or products.
From the equilibrium point of view, this suggests that phase relations should be
investigated as a function of oxygen potential. The redox equilibria are shifted
depending on the energetic states in double or complex oxides. Furthermore, to
properly understand such processes kinetically, it is essential to consider the oxygen
diffusion path through gaseous channels or inside materials. In electrochemical
cells with the oxide-ion conductor as electrolyte, oxygen movement is particularly
important for understanding the reactivity of the involved materials.

Acid-Base Properties
Another approach for understanding reactivity is based on the acid-base concept
(21–24). Although there are many definitions for determining acidity in oxide
systems, the chemical nature for salt formation reactions is commonly considered;
that is, the acidic oxide can easily react with the basic oxide. A typical example is
given as

(Zr0.8 Y0.2 )O2 + 0.2 V2 O5 = 0.8 ZrO2 (monoclinic phase) + YVO4 . 12.

Here, V2O5 is acidic and Y2O3 is basic so that the YVO4 formation can easily
proceed, although Y2O3 is well stabilized in the cubic fluorite structure. From a
chemical point of view, this is a convenient means of understanding solid-state
reactions.
This tendency can be correlated with the stabilization energy in double ox-
ides, and these stabilization energies can be interpreted in terms of ionic size,
valence, and other crystal chemical features as described above. For example,
V2O5 has a small valence stability because the pentavalent V5+ requires a large
number of coordinate oxide ions, which makes it difficult for oxide ions to be
kept apart in order to maintain low electrostatic repulsive energy. When V2O5
forms a double oxide with those oxides having large ions with lower valency
states, it possible for V5+ to have a large number of oxide ion coordinates, i.e.,
the oxide ions are separated by large cation neighbors. This explains why the
stabilization energy of YVO4 is large and therefore Equation 12 can be easily
processed.

Compound Formation Versus Solid Solution


in Multicomponent Systems
As shown above, significant stabilization can be expected when several cations
with different valence and ionic size form double or complex oxides. By increas-
ing the component number, however, these effects will be weakened. Thus it
becomes difficult to obtain greater stabilization upon further addition of oxide
components.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

592 YOKOKAWA

SOLID OXIDE FUEL CELL MATERIALS


Materials Issues in Solid Oxide Fuel Cells
To realize electrochemical energy conversion, SOFCs consist mainly of an oxide-
ion conductive electrolyte, two (air and fuel) electrodes, and interconnect materials
(2). Because those materials are made of solid components, some technological
difficulty of fabricating SOFC stacks is encountered compared with fabricating
other types of fuel cells that contain liquid substances (3). To establish a good
electrical contact at a solid-solid interface, high-temperature heat treatments or
physically activated processes are needed. On the other hand, those stacks have to
experience changes in temperature and ambients during fabrication, operation, and
maintenance. When materials are tightly bonded, differences in volume expansion
give rise to thermal stress so that thermal expansion matching among the cell
components becomes another important requirement.
The interface stability among the cell components is crucial. When interface
reaction products have poor electrical properties, a degradation in electrochemical
performance can occur. When interface reaction products have different thermal
expansion coefficients, micro-crack formation at the interface may lead to me-
chanical instability (e.g., interface delamination).
In the following, interface stability among the cell components is discussed from
the thermodynamic point of view with emphasis on fluorite oxides and perovskite
oxides.

Thermodynamic Representation of
Solid Oxide Fuel Cells Materials
In order to understand interdiffusion, as well as interface reactions, it is essential to
thermodynamically represent materials including dopants or possible impurities.

FLUORITE SOLID SOLUTION The fluorite-type oxides, YSZ and doped ceria used
as an electrolyte, are well represented by a simple sub-regular solution model (20,
26). The Gibbs energy is given as follows

G total = xA G A + xB G B
= xA (G ◦A + RT ln xA ) + xB (G ◦B + RT ln xB )
+ xA xB [A0 + A1 (xA − xB )], 13.

where G◦ A and G◦ B are the lattice stability of the fluorite phase, xA and xB are the
concentration of A and B, and A0 and A1 are the first and the second interaction
parameters in the subregular solution. Figure 3 shows the calculated phase diagram
and associated chemical potential behavior of a ZrO2-YO1.5 system. In pure ZrO2,
the cubic fluorite phase is stable only above 2662 K. At low temperatures, the
monoclinic structure, in which the Zr ions have seven coordinates, is stable. When
the YO1.5 component is added to the cubic ZrO2 phase, the large stabilization is
obtained as shown in Figure 3. This stabilization is expressed as the large negative
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 593

Figure 3 (a) Calculated phase diagram for the ZrO2-YO1.5 system and (b) Gibbs
energy and chemical potentials of ZrO2 and YO1.5 at 1273 K.

value of A0 in Figure 4. Because the oxide ion vacancies are formed by YO1.5
doping, the cubic structure is stabilized.
Recent investigations on the vacancy configuration by atomistic calculation (27,
28) and by EXAFS analyses (29) indicate that the oxide ion vacancies are formed
selectively around the Zr ions, leading to the ionic configuration in which the Zr4+
ions have about seven coordinate oxide ions even in the cubic structure. With this
in mind and with the aid of atomistic calculations, the macroscopic property can
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

594 YOKOKAWA

be well correlated with features of the ionic configuration. Successful attempts


have been made to correlate the chemical potential of the YO1.5 component with
the defects properties such as proton solubilities and hole conductivities (30–32).
Although the phase diagrams of the MO2-M’On (M = Zr, Ce; M’ = dopant)
systems exhibit a variety of behaviors, the interaction parameters for respective
dopant valences show excellent regularity with the dopant radius (20). Figure 4
shows the interaction parameters for the ZrO2-MO and the ZrO2-MO1.5 systems.
These parameters provide analyses for following chemical behaviors.

REACTIVITY OF THE DOPANT COMPONENTS Chemical reactivity of doped fluo-


rite oxides can be separately discussed for host materials and dopants. Although
the chemical potential of host materials is not sensitive to the types of dopants,

Figure 4 Interaction parameters for the fluorite solid solution between the cubic
zirconium oxide and metal oxides with divalent (upper) and trivalent (lower) ions,
respectively. The coefficient a represents the lattice stability of MO (or MO1.5) relative
to the most stable lattice structure. The parameters A0 and A1, are the first and second
Redlich-Kister equations. Dopant radius for the 8 coordinates is used for each ion.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 595

their chemical potential does depend on composition of the dopants and the hosts.
For example, the YO1.5 component chemical potential values are quite different
between the ZrO2-YO1.5 and the CeO2-YO1.5 systems (30). This gives rise to
different reactivities with other substances; e.g., for Al2O3, the YO1.5 component
in YSZ does not react, whereas that in the doped ceria does react to form YAlO3.
Similarly, the basic oxide YO1.5 tends to react with the acidic oxides such as H2O,
CO2, or SiO2. The reactivity with such substances also depends strongly on the
chemical potential of YO1.5 in host materials (32).

Figure 5 Calculated solubility of the transition metal oxides and the associated equi-
librated phase at 1273 K as a function of oxygen potential. The dash and dotted lines
separate the contributions from the different valences.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

596 YOKOKAWA

SOLUBILITY OF THE TRANSITION METAL OXIDES It is difficult to find correlations


among solubilities of the transition metal oxides in YSZ. Even so, by using the
present interaction parameters, the difference in solubilities can be well interpreted
in terms of the valence number and the ionic size. Figure 5 shows the calculated
solubility of transition metal oxides as a function of oxygen potential. The solubility
of Mn oxide is larger than that of Fe oxide. This is because Mn2+ ion size is larger
and closer to the critical dopant size, as shown in Figure 3, and also because of the
valence stability of binary oxides; the Fe solubility decreases rapidly with lowering
oxygen potential in the Fe metal stable region, whereas the stability region of MnO
is large, which makes the solubility of Mn oxide constant over a wide range of
oxygen potential.
The valence state of transition metal ions in YSZ has been experimentally deter-
mined by Sasaki et al. (33). Figure 6 compares the calculated and observed values
of transition metal ion concentration in YSZ. Except for V3+/V4+ ions, essentially
the same behavior is obtained, which confirms that the valence number and the
ionic size effects are important in determining the thermodynamic state of dopants.

PEROVSKITE OXIDE The perovskite type oxides, ABO3, consist of two cations and
oxide ions (34, 35). The B cations can have valences from +3 to +5 with 6 oxygen
coordinates, whereas the A cations have the valences from +3 to +1 with 12
oxygen coordinates. The stabilization energy depends strongly on how the ionic
configuration can be fitted to the perovskite structure; this is measured in terms
of the Goldshmidt tolerance factor, t, which is determined from ionic radii, rA, rB,
and rO, as follow:
(rA + rO )
t =√ . 14.
2(rB + rO )
The most ideally packed ionic configuration can be obtained when t = 1; cor-
respondingly, the stabilization energy shows an excellent dependence of t with the
largest value around t = 1 as shown in Figure 7 (36). The large stabilization en-
ergy of LaMO3(M = transition metal) stabilizes these perovskite phases against
reduction. Figure 8 shows that the reductive decomposition limit of LaMO3 is
shifted to the more reductive side when compared with the corresponding M2O3
phase. Although LaCoO3 has a large stabilization energy, its decomposition oxy-
gen potential is high compared with that of other perovskites because the valence
stability of Co3+ is originally low. When comparison is made between LaMnO3
and LaFeO3, interesting differences can be seen in decomposition products and
oxygen potential despite the fact that the magnitude of stabilization is about the
same. Such differences originate from differences in the valence stability.
The chemical features of perovskite oxides can be summarized as follows
1. A large oxygen nonstoichiometry; although the interstitial positions are not
available for oxygen atoms in the perovskite structure, oxide ion vacancies
can be formed to a large extent. Nondoped perovskites can have oxide ion
vacancies when the valence of the B-site cation is changed.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 597

Figure 6 Comparison between calculated and observed valence states of


transition metal ions in YSZ as a function of oxygen potential.

2. The A-site ions can be substituted with cations with different valences. To
compensate for the charge, the oxide ion vacancy formation or the hole
(electron) formation will be accompanied by such a substitution.
3. An A-site deficiency can occur when the B-site cations have high valence
states. Because the B-site ions change their valency, it is possible that oxide
ion vacancies will form in addition to A-site vacancies.

Figure 7 The stabilization energy of perovskite oxides, AIIIBIIIO3 as a


function of tolerance factor determined from the ionic size alone.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

598 YOKOKAWA

Figure 8 The two-dimensional chemical potential diagram for the


La-Mn-O, La-Fe-O, and La-Co-O systems at 1273 K.

In many cases, perovskite oxides exhibit a large mutual solubility despite the
difference in the ionic size among respective A-site and B-site ions. This behavior
suggests an ideal association model in which the constituents are assumed to have
a perovskite-related structure. To investigate the Sr substitution on the La site,
a mixture of LaCrO3 and SrCrO3 can be considered. When hypothetical (unsta-
ble) constituents such as CaCrO3 are considered, the good correlation between
the stabilization energy and the ionic size allows for estimating the thermody-
namic properties. To represent oxide ion vacancies, another type of hypothetical
compound, LaMnO2.5, can be considered. For the A-site deficiency, La2/3MnO3
or La2/3MnO2.5 can be included as possible constituents (36). By adopting these
constituents for LaMnO3 perovskites, attempts have been made to represent the
thermogravimetric results on the A-site-deficient lanthanum manganite, which has
been observed by Mizusaki and coworkers (37). As shown in Figure 9a, the re-
producibility of the thermogravimetric results on LaMnO3-d, La0.95MnO3-d, and
La0.9MnO3 is satisfactory.
From the defect chemistry point of view, several investigators have taken into
account the thermogravimetric results. Among them, Roosmalen & Cordfunke
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 599

(38–42) assumed that the disproportionation of manganese trivalence ions into


divalent and tetravalences may occur; furthermore, they assumed that the tetrava-
lence will exist at a selected concentration throughout the oxygen-deficient region.
These assumptions have been adopted by Mizusaki et al. (43, 43a) to account for
the electron conductivity and related defect properties of lanthanum manganite. In
Figure 9c, the change of respective valence concentration is shown as a function
of average Mn valence. The main point is that the tetravalent Mn4+ concentration
is assumed to be constant over a wide range of Mn valence number; this is made
to reproduce the thermogravimetric behavior. On the other hand, Figure 9b clearly
indicates that without assuming the tetravalent manganese ions in the oxygen de-
ficient region, the thermogravimetric results can be well taken into account within
the ideal association model.

Valence Stability and Cathode/Electrolyte Interaction


The transition metal elements play important roles in SOFC materials (44–46).
In particular, air electrode and interconnect materials are made of manganese
and chromium compounds, respectively. More recently, iron has attracted much
attention as an intermediate SOFC cathode (47). The validity of such materials
selection can be examined in terms of the valence stability and the stabilization
energy. In what follows, cathode and electrolyte interaction are examined for the
viewpoint of valence stability.

Figure 9 The nonstoichiometry of LaMnO3 and its relation with valence numbers.
(a) Experimental thermogravimetric results (37) (square, circle and diamond) and
model fitting results (solid lines), (b) calculated valence numbers as a function of aver-
age manganese valence, (c) proposed valence numbers by Mizusaki et al. (43, 43a).
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

600 YOKOKAWA

Figure 9 (Continued)
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 601

The LaCoO3-based cathode was not selected for use with the YSZ electrolyte
because the thermal expansion coefficient is too high and the reactivity with YSZ
is severe. However, the LaMnO3-based perovskite has attracted much interest.
Although the electrode performance of LaMnO3-based cathodes are relatively
good, some reactions with YSZ have occurred. Lau & Singhal (48) investigated
the interface stability between (La,Sr)MnO3 (LSM) and YSZ. They found that
interdiffusion across the interface, particularly manganese dissolution into YSZ,
is significant and also that La2Zr2O7 is formed at the interface.
The thermodynamic considerations on reactivity with YSZ were made to op-
timize the LaMnO3-based cathode (49–59). As described above, the manganese
ions can be dissolved into YSZ as Mn3+ and Mn2+ in the cathode atmosphere, and
the concentration of Mn2+ will increase with decreasing oxygen potential. From
the electrical measurements, some anomaly was observed in the oxygen potential
where manganese valence changed from +3 to +2 (60).
Reactions of LaMnO3-based perovskites with YSZ are not destructive so that
the electrochemical activity does not degrade significantly compared with that of
LaCoO3-based perovskite. As shown in Figure 10, the Gibbs energy change for
the following reactions are essentially positive in the fabrication temperature and
the operation temperature:

LaMnO3 + ZrO2 = 0.5 La2 Zr2 O7 + MnO + 0.25 O2 , 15.


LaMnO3 + ZrO2 = 0.5 La2 Zr2 O7 + MnO1.5 , 16.

Figure 10 Temperature dependence of the possible reactions between


LaMnO3 and ZrO2.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

602 YOKOKAWA

LaMnO3 + ZrO2 + 0.25 O2 = 0.5 La2 Zr2 O7 + MnO2 . 17.


Since Lau & Singhal (48) actually observed the La2Zr2O7 formation at inter-
faces, other type of reactions should also be considered.
To properly account for the reaction behavior of lanthanum manganite, it is
necessary to consider the valence changes in the perovskite phase, which is re-
lated to the A-site deficiency as described above. The consideration of the A-site
deficiency on the reactivity can lead to the following reaction:
LaMnO3 + xZrO2 + 0.25 O2 (g) = 0.5xLa2 Zr2 O7 + La1−x MnO3 . 18.
Here, the average manganese valence changes from 3 to 3(1+x). When the initial
A-site deficiency increases in La1-xMnO3, the driving force of reaction decreases
and eventually the three-phase coexistence field, La1-x0 MnO3, ZrO2, and La2Zr2O7,
appears. Below this critical A-site deficiency, x0 , La1-xMnO3 no longer reacts with
YSZ. Note that the above relation is oxidative.
In the A-site-deficient LaMnO3, the manganese dissolution into YSZ becomes
significant because the MnOn activity in the A-site-deficient LaMnO3 becomes
high. This Mn dissolution is reductive.
Figure 11 shows that the chemical potential diagram for the La-Zr-Mn-O system
at a constant oxygen potential. The solubility of Mn and La oxides in the cubic
YSZ phase are included.

Chemical Reactions During Cell Fabrication


and Cell Operation
On La2Zr2O7 formation or Mn dissolution, the manganese valence changes are
accompanied by evolving or adsorbing oxygen gas changes. Under electrochem-
ical polarization, oxygen potential distribution is developed inside the cathode
layer. This gives rise to an interesting situation in which the chemical reactivity at
the LaMnO3/YSZ interfaces depends on the direction and extent of polarization.
This reaction was experimentally confirmed by Tricker & Stobbs using TEM (61,
62). They observed the La2Zr2O7 formation along the LaMnO3/YSZ interface after
high-temperature heat treatment. Because this reaction needs oxygen as a reactant,
it is reasonable for La2Zr2O7 to extend along the interface, as shown in Figure 12a,
because only the LSM/YSZ interface is an effective oxygen diffusion path. Note
that the oxygen permeation rate through YSZ and LaMnO3 is very small be-
cause in the oxide-ion conductive YSZ, the electron conductivity is small and the
oxide-ion conductivity is also small in the electron-conductive LaMnO3. Quite
different features were observed after 24-h cell operation. When SOFC operation
is performed, the oxygen flow and related oxygen potential distribution drastically
change depending on the overpotential or the current density. Recent investigations
on cathode reaction mechanism have revealed that the electrochemical active sites
are distributed along the electrolyte/electrode/gas three-phase boundaries (TPBs).
Oxygen gas is diffused on the surface of the LSM or through gaseous channels.
Thus the oxygen potential in the gas phase or the LSM surface becomes minimal at
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 603

Figure 11 Chemical potential diagram for the La-Zr-Mn-O system at


p(O2) = 1 atm. [Reproduced from (56) with permission from The Electro-
chemical Society]

the TPBs. Because oxide ions are formed at TPBs and are transported to the other
side of the electrolyte, the inside LSM/YSZ interface becomes more reductive than
the surface along TPBs, which provides the driving force for La2Zr2O7 to move to
the oxidative side and disappear from the interface region as shown in Figure 12b
(58). This phenomenon was first observed by Tricker & Stobbs (61) and later by
Weber et al. (63).
When the A-site LaMnO3 is adopted, the high-temperature heat treatment with
YSZ produces two distinct interface areas (64, 65). Because the manganese dis-
solution is a reductive and therefore oxygen-evolving reaction, the manganese-
dissolved area can be well distinguished from the region where Zr or Y dissolved
into perovskite phases (see Figure 12c). Under the cell operation, the Mn disso-
lution is enhanced at the entire interface, and the LSM/YSZ interface becomes
homogeneously flat as shown in Figure 12d.
On anodic polarization of the LSM/YSZ interface, the oxygen potential of the
interface shifts to the more oxidative side. In the middle of the LSM interface
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

604 YOKOKAWA

Figure 12 (a) La2Zr2O7formation at high temperature; (b) La2Zr2O7 disappearing


from the interface under a cell operation; (c) formation of spot-like Mn-rich region at
convex parts of YSZ; (d) flattening of a YSZ/LaMnO3 interface under a cell operation.

islands, where the driving force of the La2Zr2O7 formation becomes large, the
oxygen potential is the highest. Once the La2Zr2O7 phase is formed, this blocks the
oxygen flow, which results in the increase of oxygen potential. Increased internal
gas pressure leads to the detachment of LSM from YSZ. In view of this, it is
essential in a water electrolyzer to avoid the composition region where the La2Zr2O7
formation can be expected (66).

Further Improvement of YSZ/(La,M)MnO3 Interface


With the knowledge that the A-site-deficient lanthanum manganite does not react
with YSZ, Suzuki et al. (67) fabricated the LSM/YSZ interface by electrochemical
vapor deposition (EVD). The resulting cathode exhibits excellent performance as
illustrated in Figure 13. Their investigation clearly indicated that (a) when TPBs
are long enough, the activation overpotential is extremely small. For this purpose,
the EVD process is excellent. (b) When air was used instead of pure oxygen, some
overpotential appears. This is apparently due to the gas diffusion inside pores in a
cathode layer.
In practical stack fabrication, therefore, a key issue has been made clear: how to
fabricate the cathode/electrolyte interface using inexpensive sintering procedures.
Suzuki et al. (67) have revealed that the cathode overpotential is 200 times higher
when the sintering method is adopted by using the same materials.
Although the sintering behavior of LSM is not significant when sintering is
made on LSM alone, the microstructure of LSM is usually damaged when LSM
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 605

Figure 13 (a) The cathode performance reported by Suzuki et al. (67) and
(b) the temperature dependence of the reaction resistivity.

is sintered with YSZ (68). This is probably because the dissolution of Zr and/or
Y into perovskite phase enhances the sinterability. In particular, this degradation
due to the change in microstructure occurs after the heat treatment at temperatures
higher than 1473 K. Several attempts have been made to overcome this issue:
■ The main efforts have involved the use the LSM/YSZ composite cathodes.
■ Other efforts have sought alternative cathodes.
■ Further doping to LSM has been tried: Mori et al. (69) found that the A-site-
deficient lanthanum manganite showed some degradation after heat treatment
above 1473 K and attempted to improve this behavior by doping the chromium
component with lanthanum manganite with the expectation that this would
give rise to poor sinterability (69). Although some improvement was obtained,
the overall results were not good. The most important demerit of Cr doping
is the weakening of the cathode layer adhesion.
■ Herbstritt and coworkers (70), using a Nobel technique, attempted to prepare a
metal organic deposition thin (80 nm) layer of LSM on YSZ by spin coating
process. Interestingly, this thin layer of LSM became porous during cell
operation, providing a fine microstructure with long TPBs.

Composite Electrode
Mizusaki et al. (68) also realized that the degradation of cathode can be derived
from the microstructure change due to sintering. Their cathode, (La0.6Ca0.4)MnO3,
did not react with YSZ so that the degradation is not due to the reactions among
the components but rather to the morphological change, leading to the decrease in
the TPBs. They found that addition of YSZ to cathode improved the degradation.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

606 YOKOKAWA

Risø National Laboratory have made systematic investigations on the com-


posite cathodes (71–75), whereas Virkar and coworkers (76, 77) and Bernett and
coworkers (78–80) made efforts in obtaining excellent performance with compos-
ite electrodes at lower temperatures.
From the thermodynamic and related points of view, the success of the LSM/
YSZ composite cathodes can be interpreted in terms of the following features:
■ During high-temperature heat treatment, La2Zr2O7 is formed if equilibrium
is not achieved. Thus it is possible to avoid the La2Zr2O7 formation between
LSM in the composite cathode and the YSZ electrolyte plate, which is the
most important interface in view of the fact that it is here that the electro-
chemical active sites are distributed around this interface.
■ The La2Zr2O7 phase, if formed inside the composite cathode, can inhibit
sintering of LSM. Without La2Zr2O7, YSZ can also inhibit sintering.
■ Recent attempts of using doped ceria (80, 81) instead of YSZ imply that
doped ceria can also inhibit sintering of LSM.
■ The LSM/YSZ composites exhibit higher oxide ion diffusivity and surface
reaction rate. This assists the oxygen flow to the TPBs on the YSZ electrolyte
plate.

Alternative Cathodes
Many attempts have been made to clarify the interface stability of LSM on YSZ and
to seek alternative cathodes that are chemically stable against YSZ (82–86). Initial
reaction experiments were made around 1000 ◦ C, and essentially no new cathodes
that stable against reactions with YSZ were proposed. Recently, attempts have
been made to obtain cathodes for the intermediate temperature SOFCs. Among
perovskite oxides, LaFeO3-based perovskites (47) have attracted much interest
because reactions with YSZ are found to be insignificant. Reactivity and its relation
with the A-site deficiency in those LaFeO3-based cathodes can be examined in
terms of the valence stability and related properties in a manner similar to those
of the LaMnO3-based cathodes.

SUMMARY
A thermodynamic method of investigating interface stability has been discussed
in terms of chemical potentials. Within the chemical potential space, phase rela-
tions associated with chemical reactions are represented using the stoichiometric
numbers and the Gibbs energy for their respective compounds. Furthermore, dif-
fusion through compounds can be well characterized in terms of the chemical
potential values, which are given within the stable regions. Chemical potentials
can also be correlated with valence stability and the stabilization energy through
Gibbs energy because the stabilization energy is well explained in terms of the va-
lence number, ionic size, and ionic-packing features. These considerations make it
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 607

possible to investigate interface chemistry by using chemical means such as va-


lence, ionic size, or ionic configuration. For the LaMnO3-based cathode/YSZ elec-
trolyte interface, this approach made it possible to correlate those fundamental
findings with the results of industrial efforts. This approach can be applied to other
materials problems of SOFCs (87, 88), molten carbonate fuel cells (11), and other
electrochemical devices (10, 12).

The Annual Review of Materials Research is online at


http://matsci.annualreviews.org

LITERATURE CITED
1. Minh NQ, Takahashi T. 1995. Science and 11. Yokokawa H, Sakai N, Kawada T, Dokiya
Technology of Ceramic Fuel Cells. Ams- M, Ota K. 1993. J. Electrochem. Soc. 140:
terdam: Elsevier. 3565–77
2. Kawada T, Yokokawa H. 1997. In Elec- 12. Yokokawa H, Sakai N, Yamaji K, Horita
trical Properties of Ionic Solids, ed. J T, Ishikawa M. 1998. Solid State Ionics
Nowotny, CC Sorrell, pp. 187–248. 113–115:1–9
Zurich:Trans Tech Publ. 13. Kirkaldy JS, Young DJ. 1987. Diffusion in
3. Yokokawa H. 2003. History of High Tem- the Condensed State. London: Institute of
perature Fuel Cell Development: Hand- Metals
book of Fuel Cells, ed. W Vielstich, H 14. van Zeggeren F, Storey SH. 1970. The
Gasteiger, A Lamm, pp 219–66. London: Computation of Chemical Equilibria,
Wiley & Sons London: Cambridge Univ. Press
4. Bale CW, Chartrand P, Degterov SA, 14a. Tagawa H, Mizusaki J, Katou M, Hirano
Eriksson G, Hack K, et al. 2002. Calphad K, Sawata A, Tsuneyoshi K. 1991. Proc.
26(2):189–228 2nd Int. Symp. Solid Oxide Fuel Cells, ed.
5. Anderson JO, Helander T, Höglund L, Shi F Grosz, P Zegers, SC Singhal, O Ya-
P, Sundman B. 2002. Calphad 26(2):273– mamoto, pp. 681–88. Brussels: Commi-
312 sion Eur. Communities
6. Yokokawa H, Yamauchi S, Matsumoto T. 15. Barin I. 1993. Thermochemical Data of
2002. Calphad 26(2):155–66 Pure Substances, Weinheim: VCH.
6a. Yokokawa H, Yamaji K, Horita T, Sakai 16. Spear KE. 1976. In Treatise on Solid
N. 2000: Calphad 24(4):435–48 State Chemistry, ed. NB Hannay, 4:115–
6b. Yokokawa H, Yamauchi S, Matsumoto T. 92. New York: Plenum
1994. Thermochim. Acta 245:45–55 17. Wagman DD, Evans WH, Parker VB,
7. Eriksson G, Hack K. 1990. Metal. Trans Schumm RH, Halow I, et al. 1982. J. Phys.
B 21B:1013–23 Chem Ref. Data 11:Suppl. 2
7a. Bale CW, Eriksson G. 1990. Can. Metal. 18. Yokokawa H, Kawada T, Dokiya M. 1989.
Q. 29:105–32 J. Am. Ceram. Soc. 72:C152–53
8. Yokokawa H. 1999. J. Phase Equilibria 19. Yokokawa H, Sakai N, Kawada T, Dokiya
20(3):258–87 M. 1991. J. Solid State Chem. 94:106–
9. Singhal SC. 2000. Solid State Ionics 135: 20
305–13 20. Yokokawa H. 1998. In Zirconia Engi-
10. Yokokawa H, Horita T, Sakai N, Kawada neering Ceramics: Old Challenges–New
T, Dokiya M. 1995. Solid State Ionics Ideas, ed. E Kisi, pp. 37–74. Zurich: Trans
78:203–10 Tech Pub.
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

608 YOKOKAWA

21. Flood H, Førland T. 1947. Acta Chemica M. 1996. Solid State Ionics, 86–88:1161–
Scand. 1:592 65
22. Maier J. 2001. Chem. Eur. J. 7:4762–70 37. Mizusaki J, Tagawa H, Yonemura Y, Mi-
23. Yamaguchi S, Nakamura K, Higuchi T. namiue H, Nambu H. 1994. Proc. 2nd
Shin S, Iguchi Y. 2000. Solid State Ionics Ionic and Mixed Conducting Ceramics,
136–137:191–95 ed. TA Ramanarayanan, WL Worrell, HL
24. Ingram MD, Janz GJ. 1965. Electrochim. Tuller, pp. 402–11. Pennington, NJ: Elec-
Acta 10:783–92 trochem. Soc.
25. Deleted in proof 38. Van Roosmalen JAM, Cordfunke EHP.
26. Yokokawa H, Sakai N, Kawada T, Dokiya 1993. J. Solid State Chem. 93:212–23
M. 1993. In Science and Technology of 39. Van Roosmalen JAM, Cordfunke EHP,
Zirconia V, ed. SPS Badwal, MJ Bannis- Helmholdt RB. 1994. J. Solid State Chem.
ter, RHJ Hannink, pp. 59–68. Lancaster, 110:100–5
Australia: Technomic Pub. 40. Van Roosmalen JAM, Cordfunke EHP.
27. Minervini LM, Zacate MO, Grimes 1994. J. Solid State Chem. 110:106–8
RW. 2000. Solid State Ionics 116:339– 41. Van Roosmalen JAM, Cordfunke EHP.
49 1994. J. Solid State Chem. 110:109–12
28. Zacate MO, Minervini L, Bradfield DJ, 42. Van Roosmalen JAM, Cordfunke EHP.
Grimes RW, Sickafus KE. 2000. Solid 1994. J. Solid State Chem. 110:113–17
State Ionics 128:243–54 43. Mizusaki J, Mori N, Takai H, Yonemura
29. Yoshida H, Deguchi H, Inagaki T, Miura Y, Minamiue H, et al. 2000. Solid State
K, Horiuchi M. 2001. Solid Oxide Fuel Ionics 129:163–77
Cells VII, ed. H Yokokawa, SC Sing- 43a. Mizusaki J, Yonemura Y, Kamata H,
hal, pp. 384–92. Pennington, NJ: Elec- Ohyama K, Mori N, et al. 2000. Solid State
trochem. Soc. Ionics 132:167–80
30. Yokokawa H, Sakai N, Horita T, Yamaji 44. Yokokawa H, Horita T, Sakai N, Dokiya
K, Xiong YP, et al. 2001. J. Phase Equi- M. 1998. Ceramic Interfaces: Proper-
libria 22(3):331–38 ties and Applications, ed. RSC Smart,
31. Yokokawa H, Sakai N, Horita T, Yamaji J Nowotny, pp. 171–201. London: IOM
K, Xiong YP, et al. 2001. In Solid Oxide Commun.
Fuel Cells VII, ed. H Yokokawa, SC Sing- 45. Sakai N, Tsunoda T, Kojima I, Yamaji K,
hal, pp. 339–48. Pennington, NJ: Elec- Horita T, et al. 2001. Ceramic Interfaces
trochem. Soc. 2, ed. HI Yoo, SJL Kang, pp. 135–56. Lon-
32. Yokokawa H, Sakai N, Horita T, Yamaji K, don: IOM Commun.
Xiong YP. 2002. High Temperature Ma- 46. Sakai N, Tsunoda T, Fukumoto N, Kojima
terials, pp. 26–37. Pennington, NJ: Elec- I, Yamaji K, et al. 1999. J. Electroceram-
trochem. Soc. ics 4:119–26. Suppl. 1
33. Sasaki K, Claus J, Maier J. 1999. Solid 47. Ralph JM, Vanghey JT, Krumpelt M.
State Ionics 121:51–60 2001. Solid Oxide Fuel Cells VII, ed. H
34. Goodenough JB, Longo JM. 1970. Yokokawa, SC Singhal, pp. 466–75. Pen-
Landolt-Bornstein, Group IV, Magnetic nington, NJ: Electrochem. Soc.
Oxides and Related Oxides, Vol. 4a, Chpt. 48. Lau SK, Singhal SC. 1985. Corrosion 85,
3. Berlin: Springer-Verlag Pap. 345:1–9
35. Nomura S. 1978. Landolt-Bornstein, 49. Yokokawa H, Sakai N, Kawada T, Dokiya
Group IV, Magnetic Oxides and Related M. 1989. Denki Kagaku 57:821–28
Oxides, Vol 12a. Chpt. 3. Berlin: Springer- 50. Yokokawa H, Sakai N, Kawada T, Dokiya
Verlag M. 1989. Denki Kagaku 57:829–36
36. Yokokawa H, Horita T, Sakai N, Dokiya 51. Yokokawa H, Sakai N, Kawada T, Dokiya
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

MATERIALS COMPATIBILITY 609

M. 1990. Proc. Int. Conf. SOFC Nagoya, 64. Horita T, Yamaji K, Negishi H, Sakai N,
pp. 118–34. Tokyo: Science House Yokokawa H, Kato T. 2000. Solid State
52. Yokokawa H, Sakai N, Kawada T, Dokiya Ionics 137:897–904
M. 1990. Denki Kagaku 58:161–71 65. Mitterdorfer A, Gauckler LJ. 1998. Solid
53. Yokokawa H, Sakai N, Kawada T, Dokiya State Ionics 111:185–18
M. 1990. Denki Kagaku 58:489–97 66. Kaiser A, Monreal E, Koch A, Stolten D.
54. Yokokawa H, Sakai N, Kawada T, Dokiya 1996. Ionics 2:184–89
M. 1990. Solid State Ionics 40–41:398– 67. Suzuki M, Sasaki H, Otoshi S, Kajimura
401 A, Sugiura N, Ippommatsu M. 1994. J.
55. Yokokawa H, Sakai N, Kawada T, Dokiya Electrochem. Soc. 141(7):1928–31
M. 1991. Proc. 2nd. Int. Symp. Solid Ox- 68. Mizusaki J, Tagawa H, Tsuneyoshi K,
ide Fuel Cells, ed. F Grosz, P Zegers, SC Sawata A, Katou M, Hirano K. 1990.
Singhal, O Yamamoto, pp. 663–70. Brus- Denki Kagaku 58(6):520–27
sels: Commission Eur. Communities 69. Mori M, Sakai N, Kawada T, Yokokawa H,
56. Yokokawa H, Sakai N, Kawada T, Dokiya Dokiya M. 1990: Denki Kagaku 58:528–
M. 1991. J. Electrochem. Soc. 138:2719– 32
27 70. Herbstritt D, Warga C, Weber A, Ivers-
57. Yokokawa H, Sakai N, Kawada T, Dokiya Tiffée E. 2001: Solid Oxide Fuel Cells VII,
M. 1991. Solid State Ionics 52:43–56 ed. H Yokokawa, SC. Singhal, pp. 349–57.
58. Yokokawa H, Horita T, Sakai N, Kawada Pennington, NJ: Electrochem. Soc.
T, Dokiya M. 1994. Proc. First Eur. Solid 71. Juhl M, Primdahl S, Manon C, Mogensen
Oxide Fuel Cell Forum, ed. U. Bossel, M. 1996. Solid State Ionics 61:173–
pp. 425–34. Oberrohrdorf: Eur. Fuel Cell 81
Forum 72. Østergård MJL, Mogensen M. 1993. Elec-
59. Yokokawa H, Horita T, Sakai N, Kawada trochim. Acta 38:2015–20
T, Dokiya M, et al. 1995. Solid Oxide Fuel 73. Østergård MJL, Clausen C, Bagger C,
Cells IV, ed. M Dokiya et al., pp. 975–84. Mogensen M. 1995. Electrochim. Acta
Pennington, NJ: Electrochem. Soc. 40(2):1971–81
60. Kawada T, Sakai N, Yokokawa H, Dokiya 74. Jørgensen MJ, Mogensen M. 2001. J.
M. 1992. Solid State Ionics 53–56: 418– Electrochem. Soc. 148:A433–42.
25 75. Jørgensen MJ, Bonanos N, Mogensen M.
61. Tricker DM, Stobbs WM. 1993. In High 2001: Solid Oxide Fuel Cells VII, ed. H
Temperature Electrochemical Behavior of Yokokawa, SC. Singhal, pp. 547–54. Pen-
Fast Ions and Mixed Conductors, ed. FW nington, NJ: Electrochem. Soc.
Poulsen, JJ Bentzen, T Jacobsen, E Skou, 76. Virkar AV, Chen J, Tanner CW, Kim JW.
MJL Østergård, pp. 453–60. Roskilde: 2000. Solid State Ionics 131:189–98
RISØ National Laboratory 77. Tanner CW, Fung KZ, Virkar AV. 1997. J.
62. Tricker DM. 1993. The microstructure Electrochem. Soc. 144:21–30
of solid oxide fuel cells and related 78. Tsai T, Barnett SA. 1997. Solid State Ion-
metal/oxide interfaces. PhD thesis. Cam- ics 93:207–17
bridge, UK: Univ. Cambridge. 205 pp. 79. Murray EP, Tsai T, Barnett SA. 1998.
63. Weber A, Männer R, Jobst B, Schiele M, Solid State Ionics 110:235–43
Cerva H, et al. 1996. High Temperature 80. Murray EP, Barnett SA. 2001. Solid State
Electrochemistry: Ceramics and Metals, Ionics 143:265–73
ed. FW Poulsen, N Bonanos, S Linderoth, 81. Hiwatashi K, Ueno A, Aizawa M. 2002.
M Mogensen, B Zachar-Christianse, 5th. Eur. SOFC Forum, ed. J Huijsmans,
pp. 473–78. Roskilde: RISØ National pp. 1141–21. Oberrohrdorf: Eur. Fuel Cell
Laboratory Forum
4 Jun 2003 12:51 AR AR189-MR33-20.tex AR189-MR33-20.sgm LaTeX2e(2002/01/18) P1: IKH

610 YOKOKAWA

82. Sakaki Y, Takeda Y, Kato A, Imanishi N, R, Nickel H, Hilpert K. 1997. J. Am. Ce-
Yamamoto O et al. 1999. Solid State Ion- ram. Soc. 80:909–14
ics 118:187–94 87. Yokokawa H, Horita T, Sakai N, Kawada
83. Mori M, Abe T, Itoh H, Yamamoto O, T, Dokiya M. 1993. In High Tempera-
Shen GQ, et al. 1999. Solid State Ionics ture Electrochemical Behavior of Fast Ion
123:113–19 and Mixed Conductors, ed. FW Poulsen,
84. Kindermann L, Das D, Nickel H, Hilpert JJ Bentzen, T Jacobsen, E Skou, MTL
K, Appel CC, Poulsen FW. 1997. J. Elec- Østergård, pp. 474–78. Roskilde: Risø
trochem. Soc. 144:717–20. National Laboratory
85. Kindermann L, Das D, Bahadur D, Nickel 88. Horita T, Choi JS, Lee YK, Sakai N,
H, Hilpert K. 1998. 106:165–72 Kawada T, et al. 1995. J. Am. Ceram. Soc.
86. Kindermann L, Das D, Bahadur D, Weiss 78(7):1729–36

S-ar putea să vă placă și