10
Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells B Part I. Variation of composition Andreas Mai, Vincent A.C. Haanappel, Sven Uhlenbruck, Frank Tietz * , Detlev Sto ¨ver Institute for Materials and Processes in Energy Systems, Forschungszentrum Ju ¨lich, IWV-1, D-52425 Ju ¨ lich, Germany Received 16 August 2004; received in revised form 9 March 2005; accepted 10 March 2005 Abstract The properties and the applicability of iron- and cobalt-containing perovskites were evaluated as cathodes for solid oxide fuel cells (SOFCs) in comparison to state-of-the-art manganite-based perovskites. The materials examined were La 1 x y Sr x Co 0.2 Fe 0.8 O 3 d (x = 0.2 and 0.4; y =0 0.05), La 0.8 Sr 0.2 FeO 3 d , La 0.7 Ba 0.3 Co 0.2 Fe 0.8 O 3 d and Ce 0.05 Sr 0.95 Co 0.2 Fe 0.8 O 3 d . The main emphasis was placed on the electrochemical properties of the materials, which were investigated on planar anode-supported SOFCs with 8 mol% yttria-stabilised zirconia (8YSZ) electrolytes. An interlayer of the composition Ce 0.8 Gd 0.2 O 2 d was placed between the electrolyte and the cathode to prevent undesired chemical reactions between the materials. The sintering temperatures of the cathodes were adapted for each of the materials to obtain similar microstructures. In comparison to the SOFCs with state-of-the-art manganite-based cathodes, SOFCs with La 1 x y Sr x Co 0.2 Fe 0.8 O 3 d cathodes achieved much higher current densities. Small A-site deficiency and high strontium content had a particularly positive effect on cell performance. The measured current densities of cells with these cathodes were as high as 1.76 A/cm 2 at 800 -C and 0.7 V, which is about twice the current density of cells with LSM/YSZ cathodes. SOFCs with La 0.58 Sr 0.4 Co 0.2 Fe 0.8 O 3 d cathodes have been operated for more than 5000 h in endurance tests with a degradation of 1.0 – 1.5% per 1000 h. D 2005 Elsevier B.V. All rights reserved. PACS: 84.60.Dn (Electrochemical conversion and storage); 81.05.Je (Ceramics and refractories) Keywords: Solid oxide fuel cells; SOFC; Cathode; Perovskite; LSCF; (La,Sr)(Co,Fe)O 3 ; Lanthanum ferrite 1. Introduction Reducing the costs for solid oxide fuel cell (SOFC) systems is one of the main current issues of this technology [1,2]. Therefore, efforts are being made to lower the operating temperature below the commonly used 800 – 1000 -C and to increase the power density of the SOFCs. Operating temperatures of 750 -C or below make it possible to use cheaper interconnect materials and a wider range of sealing materials [4]. It also reduces the degradation of stack materials and therefore leads to improved reliability and long-term stability. The decrease of the power densities at lower temper- atures is mainly due to overpotentials at the cathode [3,5,6]. This makes it necessary to develop new cathode materials with a higher electrocatalytic activity than those of the state- of-the-art La 1 x y Sr x MnO 3 d (LSM) perovskites. In addition to a high electronic conductivity, these materials should have high oxygen ion conductivity to enlarge the area where oxygen reduction can take place as well as high oxygen surface exchange coefficients for faster kinetics at the gas/cathode interface. Furthermore, no chemical reac- tions with the surrounding materials should occur, and the thermal expansion coefficient (TEC) should be close to that of the electrolyte to avoid mechanical stresses. 0167-2738/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.ssi.2005.03.009 i In memoriam Professor B.C.H. Steele. * Corresponding author. Tel.: +49 2461 61 5007; fax: +49 2461 61 2455. E-mail address: [email protected] (F. Tietz). Solid State Ionics 176 (2005) 1341 – 1350 www.elsevier.com/locate/ssi

Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

Embed Size (px)

Citation preview

Page 1: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

www.elsevier.com/locate/ssi

Solid State Ionics 176 (

Ferrite-based perovskites as cathode materials for anode-supported

solid oxide fuel cellsB

Part I. Variation of composition

Andreas Mai, Vincent A.C. Haanappel, Sven Uhlenbruck, Frank Tietz*, Detlev Stover

Institute for Materials and Processes in Energy Systems, Forschungszentrum Julich, IWV-1, D-52425 Julich, Germany

Received 16 August 2004; received in revised form 9 March 2005; accepted 10 March 2005

Abstract

The properties and the applicability of iron- and cobalt-containing perovskites were evaluated as cathodes for solid oxide fuel cells

(SOFCs) in comparison to state-of-the-art manganite-based perovskites. The materials examined were La1�x�ySrxCo0.2Fe0.8O3� d (x =0.2

and 0.4; y =0�0.05), La0.8Sr0.2FeO3� d, La0.7Ba0.3Co0.2Fe0.8O3� d and Ce0.05Sr0.95Co0.2Fe0.8O3� d. The main emphasis was placed on the

electrochemical properties of the materials, which were investigated on planar anode-supported SOFCs with 8 mol% yttria-stabilised zirconia

(8YSZ) electrolytes. An interlayer of the composition Ce0.8Gd0.2O2� d was placed between the electrolyte and the cathode to prevent

undesired chemical reactions between the materials. The sintering temperatures of the cathodes were adapted for each of the materials to

obtain similar microstructures. In comparison to the SOFCs with state-of-the-art manganite-based cathodes, SOFCs with

La1� x� ySrxCo0.2Fe0.8O3� d cathodes achieved much higher current densities. Small A-site deficiency and high strontium content had a

particularly positive effect on cell performance. The measured current densities of cells with these cathodes were as high as 1.76 A/cm2 at

800 -C and 0.7 V, which is about twice the current density of cells with LSM/YSZ cathodes. SOFCs with La0.58Sr0.4Co0.2Fe0.8O3� d cathodes

have been operated for more than 5000 h in endurance tests with a degradation of 1.0–1.5% per 1000 h.

D 2005 Elsevier B.V. All rights reserved.

PACS: 84.60.Dn (Electrochemical conversion and storage); 81.05.Je (Ceramics and refractories)

Keywords: Solid oxide fuel cells; SOFC; Cathode; Perovskite; LSCF; (La,Sr)(Co,Fe)O3; Lanthanum ferrite

1. Introduction

Reducing the costs for solid oxide fuel cell (SOFC)

systems is one of the main current issues of this technology

[1,2]. Therefore, efforts are being made to lower the

operating temperature below the commonly used 800–

1000 -C and to increase the power density of the SOFCs.

Operating temperatures of 750 -C or below make it possible

to use cheaper interconnect materials and a wider range of

sealing materials [4]. It also reduces the degradation of stack

0167-2738/$ - see front matter D 2005 Elsevier B.V. All rights reserved.

doi:10.1016/j.ssi.2005.03.009

i In memoriam Professor B.C.H. Steele.

* Corresponding author. Tel.: +49 2461 61 5007; fax: +49 2461 61 2455.

E-mail address: [email protected] (F. Tietz).

materials and therefore leads to improved reliability and

long-term stability.

The decrease of the power densities at lower temper-

atures is mainly due to overpotentials at the cathode [3,5,6].

This makes it necessary to develop new cathode materials

with a higher electrocatalytic activity than those of the state-

of-the-art La1� x � ySrxMnO3� d (LSM) perovskites. In

addition to a high electronic conductivity, these materials

should have high oxygen ion conductivity to enlarge the

area where oxygen reduction can take place as well as high

oxygen surface exchange coefficients for faster kinetics at

the gas/cathode interface. Furthermore, no chemical reac-

tions with the surrounding materials should occur, and the

thermal expansion coefficient (TEC) should be close to that

of the electrolyte to avoid mechanical stresses.

2005) 1341 – 1350

Page 2: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

A. Mai et al. / Solid State Ionics 176 (2005) 1341–13501342

A group of materials that fulfil some of these require-

ments are iron- and cobalt-containing perovskites, for

example La1� xSrxCo1� yFeyO3� d (LSCF), that have

already been known for their high oxygen permeability [7]

and high electrocatalytic activity (see, e.g., Refs. [8–10]).

However, these materials have to be selected carefully

because they have a significantly higher thermal expansion

coefficient than the commonly used 8 mol% yttria-stabilised

(8YSZ) electrolyte [11,12]. Furthermore, these type of

perovskites form SrZrO3 with the 8YSZ electrolyte at high

temperatures [10,11]. To overcome these problems, an

interlayer consisting of Ce0.8Gd0.2O2� d (CGO) between

cathode and electrolyte can be used [13,14].

In this first part on ferrite-based cathodes materials, the

applicability of several mixed-conducting perovskites is

evaluated as cathode materials for SOFCs with an 8YSZ

electrolyte. Their properties are compared with the state-of-

the-art LSM/YSZ composite cathodes. The processing of

the CGO interlayer and its influence on performance will be

the subject of a forthcoming paper, the second part in this

series.

2. Experimental

2.1. Powder preparation

Most of the cathode materials used were synthesised by

spray-drying as described by Kontouros et al. [15]. One of

the powders (La0.8Sr0.2Co0.2Fe0.8O3� d) was also synthes-

ised by a citrate complexation (Pechini) route [16], while

La0.7Ba0.3Co0.2Fe0.8O3� d was only synthesised by citrate

complexation. In addition to these powders, La0.58Sr0.4Co0.2Fe0.8O3� d was also provided by ECN (Energy

Research Centre of the Netherlands, Petten) while

Ce0.05Sr0.95Co0.2Fe0.8O3� d was provided by the Dresden

University of Technology (Germany). All cathode powders

were calcined at 700–900 -C in order to develop the

perovskite phase. The abbreviations used hereafter for the

different compositions are listed in Table 1.

Table 1

Data of cathode materials used in this paper

Abbreviation Composition TEC (�10�6 K�1)

(30–1000 -C)

LSM La0.65Sr0.3MnO3� d 12.3 [19]

L55SCF La0.55Sr0.4Co0.2Fe0.8O3� d 17.1

L58SCF La0.58Sr0.4Co0.2Fe0.8O3� d 17.4

L60SCF La0.6Sr0.4Co0.2Fe0.8O3� d 17.5 [12]

L78SCF La0.78Sr0.2Co0.2Fe0.8O3� d 13.8

L80SCF La0.8Sr0.2Co0.2Fe0.8O3� d 14.8 [12]

L70BCF La0.7Ba0.3Co0.2Fe0.8O3� d 16.8

CSCF Ce0.05Sr0.95Co0.2Fe0.8O3� d 23.7

L80SF La0.8Sr0.2FeO3� d 11.9

* P=perovskite, (S)= traces of spinel, (L)= traces of La2O3.

The stoichiometry of the powders was controlled by

optical emission spectroscopy (ICP-OES), while the phase

composition was evaluated by X-ray diffraction using a

Siemens D500 equipped with a monochromated Cu Ka

radiation source. The XRD patterns showed traces of La2O3,

La(OH)3 and of a spinel-type formation (Co,Fe)3O4 for

some of the powders that were calcined at 900 -C or below.

After calcination at 1100 -C, traces of La2O3 were only

found in L80SF, while all the other powders did not show

any La2O3 or La(OH)3. For the La0.55Sr0.4Co0.2Fe0.8O3� d

with 5% A-site deficiency, traces of the spinel were visible

in the XRD pattern after calcination at 1100 -C. These tracesof (Co,Fe)3O4 were not considered to be detrimental for use

as an SOFC cathode. After calcination of the powders, they

were ground by ball milling for several hours until a mean

particle size (d50) of approx. 0.8 Am was achieved.

Ce0.8Gd0.2O2� d (CGO) was obtained from Treibacher

Auermet (Austria) and was also ground by ball milling.

2.2. Sample preparation

All the cathode materials were tested on anode-supported

single cells consisting of an anode substrate (Ni/8YSZ) with

an average thickness of about 1.5 mm. The substrates with a

size of 50�50 mm2 were produced by warm pressing using

a so-called Coat-Mix material [17]. An electrochemically

active anode functional layer (Ni/8YSZ, thickness approx.

10 Am) and an electrolyte (8YSZ, thickness approx. 10 Am)

were both deposited by vacuum slip casting on de-bindered

substrates and co-fired at 1400 -C. More details about the

manufacturing process can be found elsewhere [2,18].

The subsequent layers were screen-printed using pastes

based on the above-mentioned ceramic powders, an ethyl

cellulose binder and a terpineol-based solvent. The area

of the cathode layers was 40�40 mm2. The cells used

for a comparison to the reference cathode material

La0.65Sr0.3MnO3� d (LSM) had a LSM/YSZ cathode func-

tional layer and a LSM current collector as described in Ref.

[18]. For all cells with ferrite-based cathodes, a CGO

interlayer preventing chemical reactions between the cath-

Sintering temperature Crystalline phases* formed afte

calcination at 1100 -C

1100 -C P

1080 -C P, (S)

1080 -C P

1200 -C P

1060 -C P

Spray-dried: 1150 -C P

Pechini: 1080 -C P

1120 -C P

1000 -C P

1150 -C P, (L)

r

Page 3: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

A. Mai et al. / Solid State Ionics 176 (2005) 1341–1350 1343

ode and the 8YSZ electrolyte was applied on the electrolyte

and sintered at 1300 -C. Some of the cells were prepared

with an interlayer having a thickness of 5 Am and consisting

of coarser CGO powder (d50=0.89 Am) (hereafter denoted

as interlayer type I), while the others had a thickness of 7

Am and finer CGO powder (d50=0.2 Am) (interlayer type

II). The cathode was then printed and sintered on the

interlayer resulting in a thickness of 45 Am. For the sintering

temperatures of the cathodes, see Table 1 and Section 3.2.

2.3. Characterisation of the structural and chemical

properties

Characterisation of the microstructures was performed

using scanning electron microscopes (Jeol T300 and LEO

1530 (Gemini)) equipped with energy dispersive X-ray

(EDX) analysis systems.

Thermal expansion measurements on sintered dense

bodies of the cathode materials were carried out as described

in Ref. [19].

The element distribution was examined by EDX and

SIMS (secondary ion mass spectroscopy, ION-TOF Tof-

SIMS) on polished cross sections of the samples.

2.4. Electrochemical characterisation

Electrochemical measurements of single cells were

performed in an alumina test housing placed inside a

furnace. In order to obtain sufficient electronic contact

between the cell and the electronic devices a Ni mesh and a

Pt mesh were used at the anode side and the cathode side,

respectively. Air tightness of the gas compartment was

obtained by a gold sealant. During the start-up of the tests an

argon flow was injected at the anode side and an air flow at

the cathode side. The temperature was then slowly increased

to the temperature for anode reduction. After reaching this

temperature, the anodes of the single cells were reduced by

gradually replacing the argon with hydrogen. Water vapour

(3 vol.%) was added by saturating the hydrogen gas through

a water bubbler and condenser (supersaturation and con-

densation) at the desired dew point of 24 -C. The total gas

flows of hydrogen and air were both set at 1000 ml/min

(standard temperature and pressure: STP) using mass flow

Fig. 1. L58SCF cathodes sintered at 1040 (left), 1080 (middle) and 1120 -C (righ

type I and the electrolyte YSZ.

controllers. The electrochemical performance was measured

between 650 and 800 -C. All electrochemical data were

obtained by DC methods using a current-control power

supply type Gossen 62N-SSP500-40 (Gossen-Metrawatt

GmbH, Germany), and a computer-controlled data acquis-

ition system including a data logger type NetDAQ 2640A

(Fluke, The Netherlands). The current–voltage character-

istics were measured with increasing current load by a

sequential step change of 0.0625 A/cm2 starting from zero

until either the voltage dropped below 0.7 V or the

maximum current load of 1.25 A/cm2 was reached. A

comparison of the electrochemical performance regarding

the different types of single cells was made by comparing

the current densities at 0.7 V. Calculations of the current

density at 700 mV and 650, 700, 750, 800 -C are based on

inter- or extrapolation. Calculations of the area-specific

resistance are based on linear regression of the current–

voltage curves at 0.7 V. For all different types of single cells,

at least two nominally identical cells were measured. The

data given in the tables is the average value of the

measurements, while the error margins show the differences

between these identically prepared cells.

3. Results and discussion

3.1. Thermal expansion coefficients

Iron- and cobalt-based perovskites are known to have

thermal expansion coefficients (TEC) that are well above

the TEC of the 8YSZ electrolyte (aYSZ=10.8�10�6 K�1

(30–1000 -C), [11]). The TECs of the materials used in the

present work as measured by dilatometry are shown in Table

1. For most of the powders, the slope of the TEC curves

increases with temperature, which can be explained by a

loss of oxygen at elevated temperatures [20,21]. The A-site

deficiency has only a small effect on the TEC, while a

higher Sr content results in higher TECs due to higher

oxygen vacancy concentrations, as has already been out-

lined in Ref. [21].

The cathodes made of CSCF and L60SCF showed cracks

and were partially spalled off from the CGO interlayer when

sintered on the 50�50 mm2 cells. This was considered to

t). All micrographs show the cathode on top followed by a CGO interlayer

Page 4: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

Fig. 3. Microstructure of an LSM/YSZ composite cathode on a YSZ

electrolyte.

A. Mai et al. / Solid State Ionics 176 (2005) 1341–13501344

result from thermal stresses due to the high TEC of the

materials, which was aggravated by the high sintering

temperature in the case of the L60SCF. For the other types

of cathodes only microcracks were visible on the surface.

For these cathodes, the porous CGO interlayer with a

slightly higher TEC (12.8�10�6 K�1 (20–1000 -C), [19])than 8YSZ might reduce the mechanical stresses on the

cathode preventing spallation of the cathode layer.

3.2. Sintering behaviour of the different materials

It has already been shown on LSM cathodes that the

microstructure has a strong influence on the performance

of the cathode [5] and that it can be tuned by varying the

sintering temperature. A coarser structure improves factors

like ionic and electronic conductivity and gas permeability

of the cathode, while a finer structure leads to a higher

specific surface area of the cathode and therefore to a

greater number of reaction sites. Often, when comparing

different cathode materials, the sintering temperature was

optimised beforehand for one of the materials under

investigation [10,22,23] and all of the various composi-

tions were subsequently sintered at that temperature. As a

consequence, the materials’ influence on the electrochem-

ical performance of the SOFC could not be easily

identified.

Fig. 1 shows three cathode layers consisting of L58SCF

sintered at 1040, 1080 and 1120 -C, for 3 h each. The three

pictures show a coarsening of the microstructure with

increasing temperature due to enhanced sintering of the

powder particles.

The sintering activity of the materials is, however,

already very different when the stoichiometry is slightly

changed. In Fig. 2, the effect of the A-site deficiency of the

cathode material on the sintering activity is shown. While

the L60SCF powder has hardly formed any sintering necks

at 1080 -C, the L55SCF particles are in good contact to each

other, the latter appearing similar to those of the L58SCF

cathode in Fig. 1b. Sintering the L60SFC cathode at 1200

-C leads to a microstructure that is similar to the above-

mentioned L58SFC and L55SCF cathodes sintered at 1080

-C as is shown in Fig. 2c. In Fig. 3, the microstructure of an

LSM/YSZ composite cathode is shown for comparison.

Fig. 2. Cathodes consisting of L55SCF and L60SCF sintered at 1080 -C (left and m

micrographs show the cathode on top followed by a CGO interlayer type I and t

Unless noted otherwise, an appropriate sintering temper-

ature was chosen so that the cathode of all tested cells had a

microstructure similar to the L58SCF cathode sintered at

1080 -C (Fig. 1b). The sintering temperatures for all

cathodes investigated here are listed in Table 1. For the

two L80SCF powders which were differently synthesised

(citrate complexation and spray-drying), the temperature

was also adjusted to the different sintering activities.

3.3. Chemical reaction with 8YSZ

As has already been previously reported, LSCF reacts

with 8YSZ at high temperatures forming the insulating

SrZrO3 [10,24]. This can be partially prevented by an

interlayer of an additional material, for example CGO,

between cathode and electrolyte. Due to the low sintering

activity of CGO when compared with 8YSZ, the interlayer

is porous. This allows the Sr ions to diffuse through the

interlayer and form an insulating SrZrO3 barrier between

CGO and 8YSZ. The formation of this barrier is increased

by two parameters: high sintering temperatures of the

cathode and high strontium content of the cathode material.

In general, higher Sr content leads to a less stable perovskite

[25], resulting in an increased SrZrO3 formation. As can be

seen in the EDX mappings in Fig. 4, the cell with L80SCF

cathode showed no detectable SrZrO3 formation, while for

the L58SCF cell, a small zone of a Sr-rich phase can be

detected close to the electrolyte. The slightly elevated Sr-

signal in the YSZ-electrolyte is an artefact due to an overlap

of the La-series in the characteristic X-ray spectra of

iddle, respectively) and an L60SCF cathode sintered at 1200 -C (right). All

he electrolyte YSZ.

Page 5: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

0.0 0.2 0.4 0.6 0.8 1.0 1.20.6

0.7

0.8

0.9

1.0

1.1

750 °C

Cel

l vol

tage

(V

)

Current density (A/cm2)

TS= 1040 °C

TS= 1080 °C

TS= 1120 °C

Fig. 5. Current–voltage curves at 750 -C of single cells with L58SCF-

based cathodes sintered at various temperatures including a CGO interlayer

type I.

A. Mai et al. / Solid State Ionics 176 (2005) 1341–1350 1345

strontium and zirconium. Because of this superposition, the

polished cross sections were additionally examined by

SIMS. In accordance with the EDX mappings, there was

no SrZrO3 detectable for the perovskites with 20% Sr on the

A-site (L80SCF, L78SCF, L80SF), while there was a small

zone of the Sr-rich phase detectable for all of the materials

with 40% Sr on the A-site (L55SCF, L58SCF, L60SCF).

The A-site deficiency did not result in any definite effect on

SrZrO3 formation. This is in contrast to findings on

manganite-based perovskites [26], where an A-site defi-

ciency led to a higher stability with respect to strontium

depletion. There was also an inhomogeneous distribution of

cobalt visible for the L58SCF and L55SCF cathodes. This

was considered to result from cobalt oxide in the cathode (as

in Ref. [27]) and was not thought to be harmful to the

cathode’s performance.

3.4. Electrochemical performance of the materials

3.4.1. Influence of sintering temperature

As already mentioned, the sintering temperature has a

strong influence on the microstructure of the cathodes and

therefore influences the electrochemical performance of the

cells [5]. Consequently, the electrochemical performance of

cells with an L58SCF cathode sintered at 1040, 1080 and

1120 -C was tested to find the optimum in electrochemical

performance. These cells contain a CGO interlayer of type I

(mean particle size 0.89 Am, thickness 5 Am). As can be seen

in Fig. 5 and in Table 2, the electrochemical performance is

significantly lower after sintering at 1120 -C than after 1080

-C, which can be explained by the smaller intrinsic surface

area of the cathode due to particle growth. The higher

sintering temperature also increases the above-mentioned

SrZrO3 formation, which lowers the electrochemical per-

formance of the cathodes. A lower sintering temperature

(1040 -C) leads to a slightly lower current density at 800 -Cand a slightly higher one at 650 -C. This can be explained bya poorer adhesion between cathode and interlayer or among

the cathode particles themselves, which compensates the

effect of a higher surface area for the samples sintered at a

lower temperature. However, the difference between these

two is too small to draw final conclusions.

Fig. 4. Formation of a Sr-rich phase at the electrolyte: EDX line scans for stronti

electrolytes, CGO interlayers, and (a) L80SCF, (b) L58SCF cathode. The baselin

3.4.2. Influence of A-site deficiency and synthesis route

A-site deficiency influences the sintering activity and

also the stability with regard to Sr depletion. As mentioned

above, the sintering temperature was adapted to the different

sintering activities to exclude an influence of the micro-

structure. In Table 3 and Fig. 6, the results of the

electrochemical measurements of cells with L55SCF,

L58SCF, L60SCF (all with x =0.4), and L78SCF and

L80SCF (with x =0.2) cathodes are shown. The difference

of about 10% between the values for the L58SCF cells listed

in Tables 2 and 3 is due to the different CGO layers (see

Section 2.2).

The performance is significantly higher for the A-site

deficient than for the non-deficient materials. The main

reason for this might be that the oxygen defect concentration

is higher for the A-site deficient than for the non-deficient

powders [28]. For L60SCF, the cell performance also

decreased due to partial spallation of the cathode during

the measurement, which is thought to result from higher

mechanical stresses due to the sintering temperature of 1200

-C. This is in contrast to Simner et al. [23], who found that

La0.8Sr0.2FeO3� d has a higher performance than their A-site

or B-site deficient equivalents. However, here the same

sintering temperature was used for all materials, so that the

um, cerium and zirconium taken along the black lines of SOFCs with YSZ

e of the Zr and Ce lines are shifted for better readability.

Page 6: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

Table 2

Current density (A/cm2, at 700 mV) and area-specific resistance (mV cm2) of L58SCF-based single cells as a function of the sintering temperature TS of the

cathode

Temperature (-C) Current density (A/cm2, 700 mV) Area-specific resistance (mV cm2)

TS=1040 -C 1080 -C 1120 -C TS=1040 -C 1080 -C 1120 -C

800 1.51T0.12 1.60T0.08 1.00T0.10 215T2 195T3 307T2

750 1.26T0.10 1.31T0.04 0.79T0.09 272T3 239T3 409T4700 0.93T0.04 0.92T0.01 0.52T0.05 360T7 326T4 638T7

650 0.60T0.01 0.55T0.01 0.29T0.01 511T7 531T16 1275T37

A. Mai et al. / Solid State Ionics 176 (2005) 1341–13501346

cells with deficient cation stoichiometry were probably

over-sintered or poorly adhered.

Performance of the materials with the same A-site

deficiency was generally higher with higher Sr content. A

higher amount of Sr atoms instead of the trivalent lanthanum

on the A-site is known to increase the ionic and electronic

conductivity and the surface exchange of oxygen, which can

be explained by the larger number of oxygen vacancies and

electronic holes [21,29,30]. However, the thermal expansion

coefficient also increases with higher Sr content [12,21],

which can cause mechanical problems. The electrochemical

data in Table 3 and Fig. 6 show that the materials with 40%

Sr on the A-site (L55SCF, L58SCF, L60SCF) perform much

better than the corresponding perovskites with 20% Sr

substitution (L78SCF, L80SCF). Of the compositions

investigated, the cells with a L58SCF cathode achieved the

highest current densities of 1.76 A/cm2, at 800 -C and 0.7 V.

For L80SCF, cathodes made of powders synthesised by

two routes were tested. In addition to the spray-dried

powder, another powder was synthesised by the Pechini

route. The citrate complexation (Pechini) method leads to a

powder with very fine particles and high specific surface

area. Also, a calcination temperature of 700 -C was

sufficient to produce a phase-pure powder. This resulted in

a higher sintering activity, and therefore a lower sintering

temperature than for the spray-dried powder had to be

chosen (1080 instead of 1150 -C). When comparing the

Table 3

Current density (A/cm2, at 700 mV) and area-specific resistance (mV cm2) of LSCF

route

Temperature (-C) Synthesis by spray-drying

CGO layer type II

L55SCF L58SCF L60SCF L

Current density (A/cm2, 700 mV)

800 1.23T0.04 1.76T0.08 0.90T0.04 1

750 1.04T0.04 1.43T0.08 0.77T0.02 0

700 0.74T0.03 0.99T0.06 0.54T0.02 0

650 0.46T0.02 0.58T0.03 0.34T0.01 0

Area-specific resistance (mX cm2)

800 253T5 179T2 325T3 3

750 302T2 219T3 437T3 4

700 431T9 297T4 650T6 5

650 694T12 522T10 1076T29 9

+ Results from only one cell.

electrochemical performance, however, the difference of the

two materials is negligible. Nevertheless, the lower sintering

temperature for the Pechini powder could lower SrZrO3

formation and therefore the Pechini method might be an

option for other compositions, e.g. L58SCF. In addition to

the L58SCF synthesised by spray drying, we tested L58SCF

provided by ECN (Petten, Netherlands). The performance of

cells with cathodes made of this L58SCF was lower than

using the spray-dried L58SCF powder (Table 3).

3.4.3. Influence of perovskite composition

Additionally to the LSCF materials, other perovskites

were tested such as L70BCF with barium instead of

strontium, L80SF without cobalt on the B-site, CSCF with

strontium and cerium on the A-site as well as the state-of-

the-art material LSM. For all of the cells except the LSM-

based cells, a CGO interlayer type I was used.

For L70BCF a high contact resistance was measured, and

therefore an additional LSM current collector was applied

on top of the L70BCF layer. With respect to the electro-

chemical data (Table 4, Fig. 7), its performance is

approximately the same as the corresponding L80SCF

material. Thus, substitution of the alkaline atom seems to

have little influence on performance. Together with the

findings of Ahmad-Khanlou [31] and a more recent paper

by Qiu et al. [32] reporting that replacing lanthanum by

other lanthanides has little influence on electrochemical

-based single cells with different A-site deficiency, Sr content and synthesis

CGO layer type I

78SCF+ L80SCF L58SCF powder

from ECN

L80SCF citrate

complexation

.20 0.68T0.01 1.28T0.07 0.74T0.04

.96 0.56T0.01 1.10T0.03 0.59T0.04

.63 0.42T0.01 0.75T0.01 0.41T0.03

.36 0.24T0.01 0.49T0.01 0.25T0.01

32 493T4 239T2 399T6

09 546T4 283T3 540T9

57 744T16 444T7 777T10

76 1228T37 680T12 1251T17

Page 7: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

Fig. 6. Current–voltage curves at 750 -C of single cells with LSCF cathodes containing 40% (left) and 20% Sr (right) on the A-site and different levels of

deficiency. The cathode powders were synthesised by spray-drying except L80SCF-CC, which was synthesised by citrate complexation.

A. Mai et al. / Solid State Ionics 176 (2005) 1341–1350 1347

behaviour, this points to a low influence of the A-site atoms

in general.

The type of B-site atom is generally known to have a

stronger influence on the electrochemical behaviour.

Therefore, L80SF was tested in comparison to L80SCF,

with the latter having 20% cobalt on the B-site instead of

iron. With L80SF lower current densities were achieved

over the full range of temperatures. This can be explained

by the beneficial effects of cobalt for the catalytic activity

for oxygen reduction resulting in a higher performance of

L80SCF. In addition to the lanthanum-based perovskites

the highly oxygen-deficient CSCF was tested. The oxygen

ion mobility and surface exchange coefficients for these

kinds of materials are much higher than for the lanthanum-

based materials and they were proposed as a high-

performance cathode material by Trofimenko and Ullmann

[33], and by Colomer et al. [34]. The drawbacks of these

materials are, however, quite obvious when using them as a

cathode in an SOFC: although a rather low sintering

temperature could be chosen (1000 -C) to achieve the

desired microstructure, the mechanical stresses due to the

high TEC mismatch lead to partial detachment of the

cathode before and after the electrochemical character-

isation. As a consequence, the electrochemical performance

Table 4

Current density (A/cm2, at 700 mV) and area-specific resistance (mV cm2)

of single cells with different perovskite-based cathode materials

Temperature (-C) L70BCF L80SF CSCF LSM/YSZ

Current density (A/cm2, 700 mV)

800 0.65T0.05 0.51T0.08 0.64T0.11 0.92T0.05750 0.48T0.03 0.47T0.07 0.49T0.09 0.55T0.11

700 0.30T0.01 0.36T0.05 0.34T0.06 0.36T0.08

650 0.17T0.01 0.24T0.04 0.20T0.03 0.22T0.03

Area-specific resistance (mX cm2)

800 476T3 586T6 520T3 343T64

750 590T18 714T11 644T8 517T174

700 821T19 936T7 938T28 767T248650 1444T44 1408T19 1704T74 1327T312

is much lower (see Table 4 and Fig. 7) than one would

expect from the properties of this material.

For reasons of comparison, the performance of a state-of-

the-art LSM/YSZ composite cathode, which is used as the

standard material for stack tests at Forschungszentrum

Julich, is shown in Fig. 7. The current densities of the

better performing LSCFs (L55SCF, L58SCF, L78SCF) are

up to two times higher than for the LSM cathode, while the

performance of the other perovskites is about the same or

lower than that of LSM.

3.5. Behaviour of L58SCF cathodes under various

measurement conditions

Before using a new cathode material (i.e. replacing

LSM by LSCF) in SOFC stacks, it has to be ensured that

the material does not fail under stack operating conditions.

Single cells with a cathode made of L58SCF from ECN

and a CGO interlayer type II were therefore tested under

various experimental conditions: long-term performance,

with methane and various water vapour contents in the

fuel gas.

0.0 0.2 0.4 0.6 0.8 1.0 1.20.6

0.7

0.8

0.9

1.0

1.1 L80SCF L80SF L70BCF CSCF LSM/YSZ

750 °C

Cel

l vol

tage

(V

)

Current density (A/cm2)

Fig. 7. Current–voltage curves of single cells at 750 -C with different

perovskite compositions used as cathode material.

Page 8: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

0.0 0.2 0.4 0.6 0.8 1.0 1.20.6

0.7

0.8

0.9

1.0

1.1 3 vol.% H2O

25 vol.% H2O

50 vol.% H2O

Cel

l vol

tage

(V

)

Current density (A/cm2)

750 °C

Fig. 8. Current–voltage curves of an SOFC with an L58SCF-based cathode

as a function of the water vapour concentration (fuel gas: H2=1000 ml/min,

oxidant: air=1000 ml/min).

0.0 0.2 0.4 0.6 0.8 1.0 1.20.6

0.7

0.8

0.9

1.0

1.1 33 vol.% CH4 + 67 vol.% H

2O

Cel

l vol

tage

(V

)

Current density (A/cm2)

797 °C 757 °C 698 °C 643 °C

Fig. 9. Current–voltage curves of an SOFC with an L58SCF-based cathode

as a function of temperature with methane as fuel gas: CH4=330 ml/min,

H2O=670 ml/min; oxidant: air=1000 ml/min.

A. Mai et al. / Solid State Ionics 176 (2005) 1341–13501348

3.5.1. Effect of H2O concentration with constant hydrogen

flow

The effect of H2O concentration on the cell voltage and

power output at 750 -C with a constant hydrogen flow

(1000 ml/min) is shown in Fig. 8. With increasing water

concentration the OCV decreased, which is obvious due to a

lower Nernst potential, as the oxygen partial pressure

increases with higher H2O content in the fuel gas. The

average current density also slightly decreased with higher

amounts of water (see Table 5). The area-specific resistance

did not significantly change with different water concen-

trations. A similar behaviour of the electrochemical data was

obtained with other test temperatures.

No obvious differences were found with different H2O

concentrations including a constant total gas flow of 1000

ml/min (STP).

3.5.2. Effect of the presence of methane

In addition to H2–H2O gas mixtures, experiments were

performed with methane. Methane is the major compound

of natural gas, and can be used as a fuel gas for SOFCs. For

the methane reformation process, water vapour was added to

the fuel gas. The study was performed with three different

gas mixtures consisting of methane, hydrogen, water and

argon (in ml/min.): 330–0–670–0, 280–125–560–60, and

235–235–470–60. These mixtures correspond to the

situation of no external reformation of methane and two

different pre-reformation situations. Experiments were

performed between 650 and 800 -C.

Table 5

Current density (A/cm2, at 700 mV) and area-specific resistance (mV cm2) of L58S

Temperature (-C) Current density (A/cm2, 700 mV) H2–H2O flow (ml

1000–30 1000–333 1000–1

800 1.27T0.10 1.24T0.17 1.23T0.

750 1.11T0.09 0.92T0.03 0.83T0.700 0.73T0.06 0.66T0.01 0.60T0.

650 0.47T0.04 0.41T0.04 0.35T0.

* Powder from ECN.

Fig. 9 shows as an example the current–voltage curves

as a function of temperature with 33 vol.% CH4 and 67

vol.% H2O. With decreasing temperature, the current

density at 700 mV was significantly reduced; from a

calculated average value of 0.88T0.05 A/cm2 at 800 -C to

0.29T0.03 A/cm2 at 650 -C (see also Table 6). A similar

behaviour was found for the other two methane-containing

gas compositions. These lower values of the current

densities and open circuit voltages (OCVs) can be explained

by the high amount of water resulting in a higher oxygen

partial pressure.

Table 6 also shows the average current densities and

area-specific resistances as a function of the gas composi-

tion. It can be concluded that with increasing H2 concen-

tration, and thus decreasing CH4 and H2O concentration, the

electrochemical performance was only slightly improved.

For example, the average current density at 700 mV and

calculated for 750 -C increased from 0.76T0.10 A/cm2 with

a 330–0–670–0 CH4–H2–H2O–Ar gas mixture to

0.86T0.07 A/cm2 with a 235–235–470–60 gas mixture,

whereas the area-specific resistance decreased from 412T59to 314T17 mV cm2.

It is worth noting that the oxygen partial pressure of a

fuel gas including methane with 67 vol.% H2O corresponds

to that of a gas mixture of hydrogen with about 33 vol.%

H2O. Due to this relatively high oxygen partial pressure a

lower OCV was reached: for example, the OCV at 800 -Cfor the methane–water mixture was about 970 mV. In the

case of hydrogen–water mixtures, the OCV for H2-25

CF-based single cells* as a function of the temperature and gas composition

/min) Area-specific resistance (mV cm2) H2–H2O flow (ml/min)

000 1000–30 1000–333 1000–1000

20 239T2 236T3 237T4

03 283T3 302T4 288T401 444T7 432T7 405T6

02 680T12 700T18 721T14

Page 9: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

Table 6

Current density (A/cm2, at 700 mV) and area-specific resistance (mV cm2) of L58SCF-based single cells* as a function of the temperature and gas composition

Temp. (-C) Current density (A/cm2, 700 mV)

CH4–H2–H2O–Ar flow (ml/min)

Area-specific resistance (mV cm2)

CH4–H2–H2O–Ar flow (ml/min)

330–0–670–0 280–125–560–60 235–235–470–60 330–0–670–0 280–125–560–60 235–235–470–60

800 0.88T0.05 1.01T0.08 1.05T0.10 283T5 287T14 278T3

750 0.76T0.10 0.81T0.09 0.86T0.07 412T59 337T29 314T17700 0.56T0.02 0.57T0.03 0.59T0.03 597T107 499T24 467T36

650 0.29T0.03 0.30T0.03 0.32T0.04 967T55 853T79 826T51

* Powder from ECN.

A. Mai et al. / Solid State Ionics 176 (2005) 1341–1350 1349

vol.% H2O and H2-50 vol.% H2O was 990 and 910 mV,

respectively. From these data it can be concluded that the

OCV of the methane–water mixture corresponds more or

less to that of hydrogen including a relatively high

concentration of water.

3.5.3. Long-term behaviour

Endurance tests regarding the long-term electrochemical

behaviour of two SOFCs with L58SCF-type cathodes (one

based on the powder from ECN (The Netherlands), the other

one on spray-dried powder) were made for a period of 5200

and 2000 h, respectively (Fig. 10). Both single cells were

tested at 750 -C under a constant load of 0.5 A/cm2 with

1000 ml/min H2 (3 vol.% H2O) as the fuel gas. For the first

cell an average degradation rate of about 1.5% per 1000 h

was measured in the first 2200 h, which decreased to 1.0%

for the last 2600 h. For the second cell the degradation was

about 0.9% per 1000 h. This is rather high for single cell

measurements. Possible reasons for the degradation include

further SrZrO3 formation and coarsening of the micro-

structure due to sintering effects. Further work is in progress

to determine the reason for the loss in performance.

4. Conclusions

The applicability of mixed ionic-electronic conductive

perovskites as SOFC cathodes was compared with state-of-

0 1000 2000 3000 4000 50000.50

0.75

1.00

1.25

Cel

l vol

tage

(V

)

Time (h)

power blackout

Cel

l vol

tage

(V

)

Fig. 10. Long-term behaviour of two L58SCF-based cells (at 750 -C and a load

spray-dried L58SCF.

the-art LSM/YSZ cathodes. Although there are problems

with regard to chemical stability and thermal expansion,

some of the materials proved to be superior in electro-

chemical performance.

For a meaningful comparison of the electrochemical

performance of the cathode materials, the microstructure of

the samples has to be similar. This can be achieved by

applying different sintering temperatures adjusted for each of

the materials. Care has to be taken, however, that the TEC is

not exceptionally high (>18�10�6 K�1). Furthermore, the

materials should not show a high tendency to Sr depletion.

The performance of some types of perovskites was

significantly higher than that of Ftraditional_ LSM/YSZ

cathodes. Especially the LSCFs with 40% Sr on the A-site

and an A-site deficiency (L55SCF, L58SCF) showed high

power densities, in particular at low temperatures (at 0.7 V:

1.23 W/cm2 at 800 -C, 1.0 W/cm2 at 750 -C and 0.7 W/cm2

at 700 -C). This is nearly twice the power density of LSM/

YSZ cells.

It was shown that for L58SCF cathodes long-term

stability is still an issue to be considered, while the reasons

have to be investigated further.

Acknowledgements

The authors thank the staff of the IWV department at

Forschungszentrum Julich for processing the anode sub-

0 1000 20000.50

0.75

1.00

1.25

Time (h)

of 0.5 A/cm2). Left: cell based on powder from ECN. Right: cell based on

Page 10: Ferrite-based perovskites as cathode materials for anode-supported solid oxide fuel cells: Part I. Variation of composition

A. Mai et al. / Solid State Ionics 176 (2005) 1341–13501350

strates and electrolyte layers, namely Mr. Blaß, Mr. F. J.

Dias, Ms. M.-T. Gerhards, Mr. M. Kampel, and Ms. G.

Klein, and for synthesis of some of the cathode materials in

particular Mr. W. Jungen, Mr. W. Herzhof and Ms. K.

Portulidou, for SEM investigations Dr. D. Sebold, for the

SIMS analyses Dr. U. Breuer and Ms. A. Scholl, and for

performing the electrochemical measurements, in particular

Ms. C. Tropartz, Ms. B. Rowekamp, and Mr. H. Wese-

meyer. In addition, thanks go to Dr. G. Rietveld from ECN,

Petten, The Netherlands, for providing La0.58Sr0.4Fe0.8Co0.2O3� d powder and to Prof. H. Ullmann (TU Dresden)

and Dr. N.E. Trofimenko (FhG-IKTS, Dresden) for provid-

ing the CSCF powder.

The work was carried out in the networking project

‘‘Renewable Energies’’ under contract no. 01SF0039 and

financial support from the German Federal Ministry of

Science and Education (BMBF) is gratefully acknowledged.

References

[1] T. Lipman, D. Sperling, W. Vielstich, A. Lamm, H.A. Gasteiger

(Eds.), Handbook of Fuel Cells, vol. 4, John Wiley and Sons Ltd.,

Chichester, UK, 2003, pp. 1318–1328.

[2] D. Stover, H.P. Buchkremer, J.P.P. Huijsmans, in: W. Vielstich, A.

Lamm, H.A. Gasteiger (Eds.), Handbook of Fuel Cells, Volume 4:

Fuel Cell Technology and Applications Part 2, John Wiley and Sons

Ltd., Chichester, UK, 2003, pp. 1013–1031.

[3] M. Lang, T. Franco, R. Henne, S. Schaper, G. Schiller, in: A.J.

McEvoy (Ed.), Proceedings of the 4th European Solid Oxide Fuel Cell

Forum, European Fuel Cell Forum, Oberrohrdorf, Switzerland, 2000,

p. 231.

[4] B.C.H. Steele, Solid State Ionics 134 (2000) 3–20.

[5] A. Ahmad-Khanlou, F. Tietz, I.C. Vinke, D. Stover, in: H. Yokokawa,

S.C. Singhal (Eds.), Procceedings of the 7th Int. Symposium on Solid

Oxide Fuel Cells (SOFC VII), The Electrochemical Society Proceed-

ings, Pennington, NJ, 2001, p. 476 (PV 2001-16).

[6] M. Sahibzada, B.C.H. Steele, K. Zheng, R.A. Rudkin, I.S. Metcalfe,

Catal. Today 38 (1997) 459–466.

[7] Y. Teraoka, H.M. Zhang, S. Furukawa, N. Yamazoe, Chem. Lett.

(1985) 1743–1746.

[8] B.C.H. Steele, S. Carter, J. Kajda, I. Kontoulis, J.A. Kilner, in: F.

Grosz, P. Zegers, S.C. Singhal, O. Yamamoto (Eds.), Proceedings of

the 2nd Int. Symposium on Solid Oxide Fuel Cells, Commission of the

European Communities, Luxembourg, 1991, pp. 517–525.

[9] V.V. Kharton, A.A. Yaremchenko, E.N. Naumovich, J. Solid State

Electrochem. 3 (1999) 303.

[10] H.Y. Tu, Y. Takeda, N. Imanishi, O. Yamamoto, Solid State Ionics 117

(1999) 277.

[11] G. Stochniol, A. Gupta, A. Naoumidis, D. Stover, in: U. Stimming,

S.C. Singhal, H. Tagawa, W. Lehnert (Eds.), Proceedings of the 5th

Int. Symposium on Solid Oxide Fuel Cells (SOFC-V), The Electro-

chemical Society Proceedings, Pennington, NJ, 1997, pp. 888–896

(PV 97-40).

[12] A. Petric, P. Huang, F. Tietz, Solid State Ionics 135 (2000) 537.

[13] H. Uchida, S. Arisaka, M. Watanabe, Electrochem. Solid-State Lett. 2

(1999) 428–430.

[14] A. Tsoga, A. Gupta, A. Naoumidis, P. Nikolopoulos, Acta Mater. 48

(2000) 4709.

[15] P. Kontouros, R. Forthmann, A. Naoumidis, G. Stochiniol, E.

Syskakis, Ionics 1 (1995) 40.

[16] M.P. Pechini, United States Patent Office, Patent no. 3,330,697

(1967).

[17] D. Simwonis, A. Naoumidis, F.J. Dias, J. Linke, A. Moropoulou,

J. Mater. Res. 12 (1997) 1508–1518.

[18] D. Stover, H.P. Buchkremer, F. Tietz, N.H. Menzler, in: J. Huijsmans

(Ed.), Proceedings of the 5th European Solid Oxide Fuel Cell Forum,

European Fuel Cell Forum, Oberrohrdorf, Switzerland, 2002, pp. 1–9.

[19] F. Tietz, Ionics 5 (1999) 129–139.

[20] L.W. Tai, M.M. Nasrallah, H.U. Anderson, D.M. Sparlin, S.R. Sehlin,

Solid State Ionics 76 (1995) 259–271.

[21] H. Ullmann, N. Trofimenko, F. Tietz, D. Stover, A. Ahmad-Khanlou,

Solid State Ionics 138 (2000) 79–90.

[22] T. Ishihara, T. Kudo, H. Matsuda, Y. Takita, J. Electrochem. Soc. 142

(1995) 1519–1524.

[23] S.P. Simner, J.F. Bonnett, N.L. Canfield, K.D. Meinhardt, J.P. Shelton,

V.L. Sprenkle, J.W. Stevenson, J. Power Sources 113 (2002) 1–10.

[24] L. Kindermann, D. Das, H. Nickel, K. Hilpert, Solid State Ionics 89

(1996) 215–220.

[25] H.U. Anderson, Solid State Ionics 52 (1992) 33–41.

[26] G. Stochniol, E. Syskakis, A. Naoumidis, J. Am. Ceram. Soc. 78

(1995) 929–932.

[27] F. Riza, Ch. Ftikos, F. Tietz, W. Fischer, J. Eur. Ceram. Soc. 21 (2001)

1769–1773.

[28] A. Mai, ‘‘Katalytische und elektochemische Eigenschaften von eisen-

und kobalthaltigen Perowskiten als Kathoden fur die oxidkeramische

Brennstoffzelle (SOFC)’’, PhD thesis, Ruhr-Universitat Bochum

(2004).

[29] J. Misuzaki, I. Yasuda, J. Shimayoma, S. Yamauchi, K. Fueki,

J. Electrochem. Soc. 140 (1993) 467–471.

[30] J.A. Kilner, R.A. De Souza, I.C. Fullarton, Solid State Ionics 86-88

(1996) 703–709.

[31] A. Ahmad-Khanlou, ‘‘Alternative Werkstoffe fur Komponenten der

Hochtemperatur-Brennstoffzelle (SOFC) zur Herabsetzung der

Betriebstemperatur’’, PhD thesis, Ruhr-Universitat Bochum (2000).

[32] L. Qiu, T. Ichikawa, A. Hirano, N. Imanishi, Y. Takeda, Solid State

Ionics 158 (2003) 55–65.

[33] N.E. Trofimenko, H. Ullmann, J. Eur. Chem. Soc. 20 (2000)

1241–1250.

[34] M.T. Colomer, B.C.H. Steele, J.A. Kilner, Solid State Ionics 147

(2002) 41–48.