Você está na página 1de 178

PAUL INGE DAHL

Synthesis and characterization of ionic


conductors based on ZrO2, BaZrO3 and SrCeO3
and
Preparation of LaFeO3 and LaCoO3 thin films

DEPARTMENT OF MATERIALS SCIENCE


AND ENGINEERING

NORWEGIAN UNIVERSITY OF SCIENCE


AND TECHNOLOGY

NTNU

IMT-Report 2006:86
IUK-Thesis 121

SEPTEMBER 2006
ii
This thesis has been submitted to

Department of Materials Science and Engineering

Norwegian University of Science and Technology

in partial fulfillment of the requirements for

the academic degree

Doktor Ingeniør

September 2006

iii
iv
PREFACE
First and foremost, I wish to express my gratitude to my supervisors
Professor Mari-Ann Einarsrud and Professor Tor Grande, for the invaluable
guidance in the process of completing this thesis. Their inspiration through
the experimental work and efficient feedbacks during the writing of the
thesis is highly appreciated.

Many additional people need to be honored for their contribution to the


work presented in this thesis:

- Professor Mats Nygren, Dr. Zhe Zhao and Mats Johnsson for
operating the Spark plasma sintering apparatus.

- Dr. Reidar Haugsrud, Professor Truls Norby and Christian Kjølseth


at the University of Oslo, for electrical conductivity measurements
on BaZrO3 and SrCeO3 materials and for assisting the preparation of
manuscript to the corresponding parts of the papers.

- MSc. Øystein Andersen operating the spray pyrolysis, administrated


by Associate professor Kjell Wiik.

- Dr. Ingeborg Kaus for assisting the work on YSZ.

- Dr. Hilde Lea Lein for experimental assistance on the preparation of


BaZrO3 and SrCeO3 materials.

- Dr. Hasan Okuyucu and Laura Bertolo for experimental work on


LaCoO3 films and Dr. Mohan Menon for his contribution to this
topic.

- The technical staff at the Department of Materials Science and


Technology, in particular Elin Nilsen for assisting the XRD
measurements.

- Dr. Julian Tolchard for crystallographic illustrations and discussions


and for feeding me while my wife was gone.

- Dr. Tommy Mokkelbost and Dr. Johann Mastin for sharing both
brilliant ideas and frustrations in the process of completing this
thesis. I hope we never meet again – at 2 am in the office!

- All other co-workers in Chemistry building II, friends and family for
making this time enjoyable.

v
Financial support from Norwegian University of Science and Technology
and the Research Council of Norway, Grant No.1585171431, is appreciated.

Last but not least, I would like to give a special thank my dear wife Melinda
and my children Isobel and Jonas, for reminding me that there is more to life
than science. Without your patience and encouragement this would not be
possible.

vi
TABLE OF CONTENT
1 Summary………….…………………………………………..………….1
2 Background and motivation…………………………………...…………5
2.1 Ionic conducting oxides and their applications…………….……...5
2.1.1 Oxygen ion conductors………………………….……….…...5
2.1.2 Proton conductors…...………………………….……….……5
2.1.3 Mixed ionic/electronic conductors……...………..……….….7
2.1.4 Solid oxide fule cells.....………………………..……….……8
2.1.5 Sensors…………………………………………….…….….10
2.2 Thin film technology…………….………………………….……10
2.3 Aim of work………..………………………………………….…11
3 Review of existing literature………………………………...………….13
3.1 Oxygen ion conductors – Zirconia………………………………..13
3.1.1 Crystal structure………………………………………….…..13
3.1.2 Stabilized zirconia……………………………………….…..13
3.1.3 Ionic conductivity in YSZ………………………………...…14
3.1.4 Powder synthesis…………………………………………….18
3.1.5 Densification and microstructure……………………………19
3.2 Proton conductors – Barium zirconate and strontium cerate………20
3.2.1 Crystal structure………………………………………….…..20
3.2.2 Electrical properties………………………………………….20
3.2.3 Powder synthesis and densification………………………….22
3.2.4 Mechanical properties and chemical stability……..…………23
3.3 Thin films of perovskite-type oxides………………………………26
3.3.1 Lanthanum ferrite thin films…………………………………26
3.3.2 Lanthanum cobaltite thin films………………………………27

References………………………………………………………………….28

vii
Scientific papers

PAPER I: Synthesis and characterization of nanocrystalline


YSZ powder by combustion synthesis……………………. 39

PAPER II: Densification and properties of zirconia prepared


by three different sintering techniques……………………. 50

PAPER III: Synthesis, densification and electrical properties


of strontium cerate ceramics………………………………. 71

PAPER IV: Preparation and characterization of


barium zirconate ceramics………………………………...101

PAPER V: Oriented LaFeO3 thin films grown on


NdGaO3 by spin-coating………………………………….131

PAPER VI: Crystallization and surface morphology of oriented

LaCoO3 films prepared by three different sol-gel routes....147

viii
LIST OF ACRONYMS

AFM Atomic Force Microscopy


BZ Barium Zirconate
CS Conventional Sintering
FESEM Field Emission Scanning Electron Microscopy
G/N Glycine to Nitrate (ratio)
HP Hot Pressing
HTXRD High Temperature X-Ray Diffraction
IR Infrared Spectroscopy
LC Lanthanum Cobaltite
LCC 20% Calcium substituted Lanthanum Cobaltite
p(O2) Partial Pressure of Oxygen
p(H2O) Partial Pressure of water vapor
p(CO2) Partial Pressure of carbon dioxide
p(H2) Partial Pressure of Hydrogen
SC Strontium Cerate
SC5Yb 5% Ytterbium substituted Strontium Cerate
SEM Scanning Electron Microscopy
SOFC Solid Oxide Fuel Cell
SPS Spark Plasma Sintering
TEC Thermal Expansion Coefficient
TEM Transmission Electron Microscopy
TGA Thermo Gravimetrical Analysis
XRD X-Ray Diffraction
YBZ 10% Yttrium substituted Barium Zirconate
YSZ Yttria Stabilized Zirconia

ix
x
1 SUMMARY
Ceramic electrolytes that conduct oxygen ions, protons or exhibit mixed
ionic and electronic conductivity at intermediate to high temperatures are
important materials for use in electrochemical devices like solid oxide fuel
cells (SOFCs), gas separation membranes and sensors. Special interest in
SOFC technology, where further optimization is still needed, is due to better
and more environmental friendly utilization of fossil fuel or hydrogen as an
alternative energy carrier. One of the main topics regarding such oxide
materials has been to improve in particular the ionic conductivity at suitable
operation temperatures. While the research on SOFCs mainly involves bulk
materials, oxide thin films are applicable for e.g. gas sensing devices and
catalysts.

The electrical and chemical properties in a specific oxide system are


dependent on the amount of impurities in the system and, in the case of
cation substitution e.g. the valance and size of the substitutes compared to
the original ions. Effects of microstructure must be considered as the
electrical and chemical properties may differ with varying grain size. This
leads to the overall aim of the present work, which has been to develop
complete synthesis routes to bulk materials and thin films of ceramic oxides
with designed microstructure. Yttria stabilized zirconia (YSZ), being the
oxygen ion conductor traditionally used in oxygen sensors and SOFCs, has
been prepared and characterized along with two perovskite-type oxide
materials showing high proton conductivity; strontium cerate and barium
zirconate. Additionally thin films of two perovskite systems, lanthanum
cobaltite and lanthanum ferrite, applicable for e.g. gas sensors, catalysts as
well as SOFCs, have been prepared.

The two first scientific papers presented here deals with yttria stabilized
zirconia. In Paper I the synthesis of nanocrystalline YSZ powders (ZrO2
with 8 mol% Y2O3) using smoldering combustion synthesis with glycine as
fuel and nitrate as oxidizer is reported. The influence of glycine to nitrate
ratio (G/N) was investigated along with the dependence of calcination
temperature with regards to the crystallite size, surface area and content of
residual carbonates in the prepared powders. A G/N ratio of 0.23 and
calcination in the range above 600ºC in oxygen flow gave high quality
powders with a crystallite size less than 10 nm. The removal of residual
carbonates by calcination above 600°C led to increased sinterability.

In Paper II, three different sintering techniques have been used to prepare
dense YSZ materials from high-quality powder described in Paper I; spark

1
plasma sintering, hot pressing and conventional sintering. The spark plasma
sintering technique was shown to be superior to the other methods giving
dense materials (≥ 96%) with uniform morphology at lower temperatures
(1100°C) and shorter sintering time (5 min). The lowest obtained grain size
of dense specimens was 0.21, 0.37 and 12 µm using spark plasma sintering,
hot pressing and conventional sintering, respectively. The total electrical
conductivity of the YSZ materials, measured by the van der Pauw
technique, showed no clear correlation with the grain size. The same for the
hardness, measured by the Vickers indentation method. The only effect of
grain size was found for the fracture toughness, where a small decrease with
increasing grain size was observed.

Preparation and characterization of SrCeO3 and BaZrO3 proton conducting


oxide materials are reported in Paper III and IV. In Paper III the preparation
of pure and 5% ytterbium substituted strontium cerate (SrCeO3 /
SrCe0.95Yb0.05O3-δ) by spray pyrolysis of nitrate salt solutions is presented.
Secondary phases detected in the as-synthesized powders were removed
after calcination in nitrogen atmosphere at 1100°C (SrCeO3) and 1200°C
(SrCe0.95Yb0.05O3-δ). Ball milling of calcined powders in iso-propanol
resulted in particle size down to 0.06 µm. Dense SrCeO3 and
SrCe0.95Yb0.05O3-δ materials with homogenous microstructure were obtained
by sintering at 1350 - 1400°C in air. The grain size of the sintered
specimens was in the range 6 - 10 µm for SrCeO3 and 1 - 2 µm for
SrCe0.95Yb0.05O3-δ. The electrical properties of SrCe0.95Yb0.05O3-δ were in
good agreement with reported data, showing mixed ionic-electronic
conduction. The ionic contribution was dominated by protons below
1000°C and the proton conductivity reached a maximum of ~0.005 S/cm
above 900°C. In oxidizing atmosphere, the p-type electronic conduction
was dominating above ~700°C, while the contribution from n-type
electronic conduction in reducing atmosphere only was significant above
~1000°C.

Preparation of pure and 10% yttrium substituted barium zirconate (BaZrO3 /


BaZr0.9Y0.1O2.95) powders using the same procedure as for strontium cerate
is presented in Paper IV. The crystalline powders were calcined at 1000°C
to remove secondary phases and agglomerates were effectively broken
down by ball milling giving particle size in the range 0.09 – 0.17 µm.
Despite of similar characteristics for the two powders, the densification
properties were poorer for the yttrium substituted material. Pressure less
sintering of BaZrO3 at 1600°C resulted in severe grain growth (up to 18 µm)
and low density (< 92%). High density (~98%) was obtained by hot
pressing of both BaZrO3 and BaZr0.9Y0.1O2.95 materials. This sintering
technique was shown to reduce the mobility of grain boundaries, giving

2
homogenous microstructures with average grain size down to 0.42 µm.
Excess of ZrO2 was observed in some hot pressed specimens and may be
due to high sintering temperatures (>1500°C) and reducing atmosphere
causing evaporation of BaO(g) and Ba(g). The electrical properties of the
BaZrO3 and BaZr0.9Y0.1O2.95 materials were in agreement with the literature,
showing high grain boundary resistance for both materials. Slightly lower
conductivity was observed for BaZrO3 in wet compared to dry atmosphere.
The activation energy for bulk conductivity was higher for BaZrO3 (~100
kJ/mol) than for the acceptor substituted material (~30 kJ/mol) confirming
charge compensation of the protons due to substitution.

Preparation and characterization of thin films of LaFeO3 and LaCoO3 are


reported in Paper V and VI. Paper V shows how LaFeO3 precursor solution
with good spinning properties was prepared using nitrate salts dissolved in
methanol added acetic acid with acetyl acetone as chelating agent. Thin
films were prepared by spin coating on (100)- and (110)-oriented single
crystalline NdGaO3 substrates as well as (0001)-oriented single crystal
Al2O3 substrates. By controlled heat treatment up to 400°C with heating
rate of 0.1°C/min, homogenous continuous amorphous films were obtained
on all substrates. Annealing for 1 h at 500 – 1000°C caused the formation
of oriented polycrystalline thin films on NdGaO3 with both (100)- and
(110)-orientation. Films prepared on sapphire were polycrystalline,
randomly oriented and inhomogeneous after annealing at 700°C. The
LaFeO3 films grown on the NdGaO3 substrates crystallized between 400
and 500°C and the average grain size increased from 40 to 250 nm as the
temperature was increased from 500 to 700°C.

Paper VI presents the synthesis of precursor solutions for preparation of


LaCoO3 and La0.8Ca0.2CoO3 films by three different sol-gel routes.
Precursor solutions were synthesized from acetates, alkoxides or nitrate salts
of the respective cations (La, Co and Ca), using methanol or 2-
methoxyethanol as solvents and suitable chelating agents. Films were
prepared on (100)-oriented single crystal yttria stabilized zirconia (YSZ)
and (0001)-oriented single crystal Al2O3 substrates by dip-coating or spin-
coating. Oriented polycrystalline films of LaCoO3 and La0.8Ca0.2CoO3 were
grown on the cubic (100)-oriented YSZ substrates. These films were (104)-
oriented when indexing according to the hexagonal structure, corresponding
to (110)-orientation in the cubic structure. The orientation was independent
of both type of precursor solution and deposition technique. Crystallization
of the films started between 800 and 900°C. Controlled heating with
heating rate of 0.1°C/min to 400°C resulted in smooth homogenous surface
morphology of the prepared films. Smooth surfaces were maintained by
heating of these films up to 900°C. Heating to 1000°C was assisted with a

3
severe anisotropic grain growth, giving increased surface roughness. Films
prepared on hexagonal (0001)-oriented sapphire substrates were
polycrystalline and randomly oriented.

4
2 BACKGROUND AND MOTIVATION
In the following introduction to the present work the main focus has been
put on oxide ceramics with characteristic electrical properties. To give the
reader an idea of the relevance of this work for technological applications a
brief overview of the different types of conducting oxide materials is
presented, along with their potential application fields. As a part of the
work has dealt with thin films of mixed ionic and electronic conducting
oxides, an introduction to thin film technology, and relevant applications is
also given. This chapter concludes with the aim of the present work.

2.1 Ionic conducting oxides and their applications


Among the many types of ionic conductors the ones included in the present
work are; oxide materials exhibiting high oxygen ion conductivity, proton
conductivity or mixed electronic and ionic conduction, where the ionic
contribution is from oxygen ions and/or protons.

2.1.1 Oxygen ion conductors


Solid oxide materials that exhibit oxygen ion conductivity are mainly used
as oxygen conducting electrodes for solid oxide fuel cells (SOFCs),
electrolysers, oxygen pumps and amperometric oxygen monitors [1]. Solid
electrolyte materials in such high temperature electrochemical systems are
required to exhibit relatively high oxygen ion conductivities. The most well
characterized oxygen ion conductors are those based on fluorite structure [2],
e.g. zirconia (ZrO2) and ceria (CeO2) properly substituted with aliovalent
cations such as yttrium (Y3+) and ytterbium (Yb3+), in order to introduce
oxygen vacancies on which the ionic conductivity depends. Other fluorite-
related materials such as pyrochlores (A2B2O7) and materials derived from
bismuth oxide (δ-Bi2O3) have also been reported to exhibit high oxygen ion
conductivity [2]. Oxygen ion conduction is also found in perovskite-type
oxides (ABO3), however, such materials are more often associated with
mixed ionic and electronic conductivity, when containing mixed valent
cations of transition metals [3]. Fig. 1 presents the conductivity in some
fluorite-type oxide systems with predominantly oxygen ion conductivity.
Although other oxide systems than yttria stabilized zirconia (YSZ, ZrO2
with 8 – 10% Y2O3 incorporated) have been reported to exhibit higher
conductivity, YSZ is still commercially used for both oxygen sensors and
as oxide electrolyte material in SOFCs. This is due to the high mechanical

5
strength, chemical and thermal stability of YSZ materials compared to other
candidate oxide materials, such as acceptor-substituted ceria (CeO2 with 5 –
10% substitution of Ce with e.g. Yb, Sm or Gd).

Fig. 1 Total conductivity (σ) with varying temperature for a selection


of fluorite-type oxide materials with predominantly oxygen ion
conductivity, as presented by Steele [3].

Development of materials with dimensions in the nanoscale range has


become an important research field due to the often different properties of
such materials compared to large scaled bulk materials. This research is
relevant also for oxygen ion conductors, as the electrical properties may be

6
related to the microstructure. The properties of the grain boundaries
compared to bulk are determining for the ionic conductivity. Higher
concentration and mobility of defects in the grain boundaries may lead to
enhanced conductivity while impurities such as SiO2 or subsistent
segregation on the grain boundaries, will increase the grain boundary
resistance compared to bulk [4-6]. Additionally, the ionic conductivity at
grain boundaries is related to the effect of space charge regions in the grains
adjacent to the grain boundaries [7]. If bulk defects with high mobility, e.g.
oxygen vacancies, are accumulated in the space charge region, the grain
boundary conductivity should increase. However, oxygen vacancy
depletion due to space charge in the vicinity of grain boundaries may
increase the grain boundary resistance. Considering these effects, reducing
the grain size, and consequently introducing a larger amount of grain
boundaries, may increase or decrease the conductivity of a material,
depending on which effects are dominating.

2.1.2 Proton conductors


Although polymeric compounds such as hydrated Nafion® exhibit high
proton conductivity, comparable to aqueous HCl and liquid H3PO4, these
materials are not utilisable at temperatures above ~200°C. For use at
intermediate to high temperatures (>500°C) substituted phosphates and
oxides are applicable [8]. Fig. 2 demonstrates how the bulk proton
conductivity in these materials goes through a maximum at higher
temperatures.

Materials with high and pure proton conductivity, including perovskite-type


oxides (ABO3) based on barium or strontium zirconates and cerates (A =
Ba, Sr, B = Zr, Ce), are candidates for electrolytes in sensors, batteries, fuel
cells and electrolysers. The advantage of using proton conducting oxide
materials as electrolyte material in SOFCs (see chapter 2.1.4) has been a
motivation for the investigation of the proton conductors in the present
work.

2.1.3 Mixed ionic/electronic conductors


The group of materials exhibiting high conductivity of both ions and
electrons are also applicable to fuel cells, sensors, gas pumps, batteries and
electrochemical reactors [9]. Oxides with mixed electronic and oxygen ion
conductivity are e.g. perovskite-type materials (ABO3) based on lanthanum
(A-site) and manganese, chromium, iron, cobalt (B-site). Additional
substitution (on both A- and B-site) alters the electrical properties, and

7
custom designed materials, e.g. for SOFC cathodes, oxygen separation
membranes and sensors can be prepared.

Fig. 2 Bulk proton conductivities of various oxide materials as


presented by Kreuer [10] based on data from Norby & Larring
[11]
.

2.1.4 Solid oxide fuel cells


Solid oxide fuel cells (SOFCs) are energy conversion devices that produce
electricity directly from a gaseous fuel by electrochemical combination of
the fuel with an oxidant [12]. The advantages of SOFCs compared to
conventional electric power generation systems are high conversion
efficiency and environmental compatibility, which is important in a society
striving to reduce the pollution. In the traditional SOFC, oxygen (from air)

8
is reduced by a porous electrode (cathode) producing oxide ions which
migrate through a solid electrolyte to the porous fuel electrode (anode) and
react with the fuel (H2, CO, CH4) forming H2O and/or CO2. In turn a proton
conducting solid electrolyte can be used where H2 is oxidized to produce
protons that subsequently react with oxygen to form water. The operation
principle in such a proton conductor based fuel cell is schematically shown
in Fig. 3. The main advantage of such SOFCs based on proton conductors
is that water forms and leaves the system on the side exposed to air, as
opposed to the traditional SOFCs based on oxygen ion conductors where
water forms on the fuel side, and consequently dilutes the fuel.

Fig. 3 Illustration of how a proton conductor based SOFC operates, by


Norby et al. [13].

9
2.1.5 Sensors
The most common type of oxygen sensors based on oxide materials utilizes
the oxygen transport through a solid electrolyte (commercially used YSZ).
This principle used in the so-called lambda sensors in gasoline engines,
illustrated in Fig. 4 a). The inner and outer platinum electrode, pasted onto
the YSZ electrolyte, are connected to atmospheric air (or reference gas) and
the exhaust gas from the engine, respectively. The different partial
pressures of oxygen create a potential which gives a direct indication of the
oxygen content in the exhaust gas. Oxygen partial pressure may also be
monitored by amperometric devices, as illustrated by Fig. 4 b). Replacing
the solid electrolyte (YSZ) in such a device with a pure proton conducting
material also allows the monitoring of hydrogen gas.

a) b)

Fig. 4 Schematic illustrations of oxygen sensors based on a solid


electrolyte like YSZ (CCM Technologies [14]).

2.2 Thin film technology


Thin film technology has become an important research field, which
supplements the research on bulk materials. Application of oxide thin films
(typical thickness ~100 nm or less) as protective layers on surfaces
exploited to rough conditions or improving the substrate properties are two
examples on use of such thin films, however, often the film itself is the
functional material [15]. The properties may differ from, and in some cases
be improved compared to bulk materials. Typical fields of interest for
applying oxide thin films include electronics, optics and energy storage. In

10
additionally, and more importantly, for the perovskite-type oxide films
included in the present work, is the use as thin electrolyte layers (with
improved electrochemical properties) in SOFCs, for gas sensors and as
catalysts.

For preparation of thin oxide films several physical techniques may be


applied, such as; pulsed laser deposition (PLD), atomic layer epitaxial
(ALE), electrochemical oxidation, ion beam sputtering and spray pyrolysis
in inductively coupled plasma (spray-ICP). Chemical techniques, like
chemical vapour deposition (CVD) and chemical solution deposition (CSD)
may also be applied. Using these chemical techniques imply the advantage
of good ability to control the chemical composition of thin films of e.g.
mixed and substituted metal oxides. CSD techniques, including sol-gel
synthesis, chelate and metal-organic decomposition have successfully been
applied to prepare perovskite-type thin films with good properties [16].
Relatively simple procedures make CSD techniques more cost efficient and
for this reason, more applicable for commercial use. This has been the
motivation for using sol-gel techniques and simple film deposition methods
(dip-coating and spin-coating) in the present study of preparation of oxide
films.

2.3 Aim of work


The overall aim of the work presented in this thesis has been to develop
complete synthesis routes to bulk materials and thin films of ceramic oxides
with designed microstructure and nanostructure, and to perform a
characterization of selected properties in these materials. The work can be
divided in three main subjects; synthesis and characterization of oxygen ion
conductors, proton conductors and oxide thin film preparation.

In the first part based on oxygen conductors the focus has been put on yttria
stabilized zirconia (YSZ) materials, commonly used in oxygen sensing
devices and solid oxide fuel cells. The most relevant literature reviewed in
the next chapter forms the basis from which the aim of presented work on
YSZ origins. Though there are many reports on powder synthesis,
densification and characterization of YSZ materials, not many of them
present complete studies with all parts included. The aim of the present
work on YSZ has been to obtain routes for synthesis of nanocrystalline
powder and densification of fine bulk materials with well defined micro-
and nanostructure. In addition to thorough microstructure studies of
prepared powders and bulk materials, it has been in our interest to study the

11
electrical and mechanical properties of YSZ and relate these to the micro-
/nanostructure.

The second part on proton conductors focuses on two perovskite-type oxide


systems, barium ziconate (BaZrO3) and strontium cerate (SrCeO3)
substituted with yttrium (Y) and ytterbium (Yb) respectively. Although
well studied by others, the synthesis and densification of these materials are
reported to be challenging. The main aim of this work has been to obtain
control over the microstructure development in the materials during
densification. By use of synthesis routes and densification techniques still
not reported for these materials, we have aimed to present a study of both
BaZrO3 and SrCeO3 materials, which should be of importance for others
working on these systems. Furthermore, it has been in our interest to study
the electrical properties of both systems, focusing on proton conductivity.
The electrical properties of prepared proton conductors were investigated at
the University of Oslo. Results from these studies are presented here,
however, more elaborate discussions are left to forthcoming publications.

The last part of this work has been somewhat different as it deals with films
rather than bulk materials. However, as for the work on proton conductors,
two perovskite-type oxides known to exhibit mixed ionic and electronic
conductivity have been studied; lanthanum cobaltite (LaCoO3) and
lanthanum ferrite (LaFeO3). The overall aim of this part of the presented
work has been to obtain fundamental understanding and experience in the
preparation of oxide thin films based on these materials. The reported
literature on thin films from these systems is limited, in particular on films
prepared by chemical routes. With this in mind, our goal has been to
contribute to this research field by presenting simple chemical synthesis
routes and methods for preparing homogenous films with designed
microstructure/ surface morphology. Due to the superior properties of
epitaxially grown films compared to randomly oriented polycrystalline
films, it has been of special interest to study preferential growth orientation
on various substrate materials.

12
3 REVIEW OF EXISTING LITERATURE
The disposition of this chapter based on the materials investigated in the
present work, includes three main parts; oxygen ion conductors represented
by YSZ, proton conductors involving SrCeO3 and BaZrO3 materials and,
finally, thin films where relevant work on LaFeO3 and LaCoO3 films is
reviewed.

3.1 Oxygen ion conductors - Zirconia


Intense investigations of the science and technology of zirconia (ZrO2) over
the last half century have propelled it into an outstanding, versatile material.
Zirconia is a remarkable material and a case study in materials science,
since structure-property correlations have been extensively examined.
Atomic structure and microstructure, defects, phase transformations, and
processing on one hand and properties (thermal, mechanical, electrical and
optical) on the other, are intimately connected. [17,18]

3.1.1 Crystal structure


Zirconia, having monoclinic crystal structure at ambient temperature,
exhibits a phase transition to tetragonal at ~1170°C, further to cubic at
~2370°C [19]. The cubic phase, which is stable from ~2370°C to the melting
point (2680 ±15°C), has a fluorite-type crystal structure, illustrated in Fig. 5,
in which each zirconium ion is coordinated by eight equidistant oxygen ions
and each oxygen is tetrahedrally coordinated by four zirconium.

3.1.2 Stabilized zirconia


Proper substitution with certain aliovalent oxides stabilizes the cubic fluorite
structure of ZrO2 from ambient temperature to the melting point [20]. The
cubic phase can exist in a wide range of compositions and temperatures.
Fig. 6 presents the phase relations in the ZrO2 – Y2O3 system as a function
of temperature. In substituted zirconia the aliovalent guest cations reside on
the sites for host cations (Zr4+), generating oxygen vacancies for charge
neutrality, as demonstrated by the defect equation (Eq. 1) for substitution
with Y2O3.
ZrO
′ + v •O• + 3O O
Y2 O 3 2 → 2YZr x
(1)

13
The oxygen vacancies lead to substantial ionic conductivity over an
extended oxygen partial pressure range, where the electronic conductivity is
negligible [21].

Fig. 5 The cubic fluorite structure of zirconia.

3.1.3 Ionic conductivity in YSZ


The concentration of vacancies is of importance for the ionic conductivity,
hence the degree of substitution is as well. For YSZ the maximum level of
ionic conductivity has been found for 8 to 10% Y2O3 incorporated into the
ZrO2 lattice. Beyond this optimal level the conductivity begins to decrease,
as indicated in Fig. 7. At the optimal level the conductivity may be
increased by raising the operating temperature to enhance the mobility of
the defects. This is why most SOFCs become efficient at operating
temperatures around 1000°C. Additionally the microstructure / grain size is
determining for the conductivity in the material.

14
Fig. 6 Phase diagram for the ZrO2 – Y2O3 system as presented by
Stubican et al. [20].

The bulk resistivity of polycrystalline ZrO2 is equivalent to that of a single


crystal and it is usually not affected by sintering temperature, atmosphere
and heat treatment. Efforts to reduce the bulk resistivity by changing
composition (adding aliovalent oxides) are now nearly exhausted and the
maximum stabilizer concentrations (e.g. 8 - 10 mol% for substitution with
Y2O3) have been established [22]. In order to increase the overall ionic
conductivity it has therefore been natural to look for ways to decrease the
grain boundary resistivity which is influenced by the impurity level,
sintering temperature, atmosphere, heat treatment, etc.

15
Fig. 7 Dependence of conductivity on composition for ZrO2-based
solid solutions at 800°C [23].

A brief review of conductivity data for YSZ bulk materials presented in the
literature in addition to density and grain size (when available), is given in
Table 1, along with comparable data for YSZ thin films. As YSZ shows
mainly ionic conductivity [24], oxygen ion conductivity has been evaluated
equal to reported total conductivities. Differences in bulk and grain
boundary conductivity (listed in Table 1) have not been considered in the
following and some of the listed conductivity values have been extrapolated
from given data, assuming Arrhenius behavior with the total conductivity, σ,
given by Eq. 2.

σT = A exp 
− Ea
kT 
(2)

A is a constant, Ea is the activation energy and k is the Boltzmann constant.

16
Table 1 Reported total conductivity, σ, at 1000°C and activation energy,
Ea, in YSZ bulk materials prepared by different sintering
techniques (CS: conventional, MS: microwave and SPS: spark-
plasma sintering). D notes average grain size. Data for YSZ
thin films (TF) are added for comparison.
Mol% Technique / Density D σ Ea Reference
Y2O3 T (°C) (%) (µm) (mS/cm) (kJ/mol) #
8 SPS / 1400 99 10 160 [25]
8 SPS / 1300 97 <1 190 92 [26]
8 SPS / 1200 95 0.15 183 109 [27]
8 SPS / 1100 96 0.21 172 102 [28]
8 SPS / 1000 91 120 [25]
8 CS / 1600 99 >5 180 92 [26]
8 CS / 1500 98 12 137 94 [28]
8 CS / 1500 99 12 156 90 b [29]
8.5 CS / 1500 97 9.5 149 107 b [29]
7.7 CS / 1500 97 8.5 160 104 b [29]
8 CS / 1500 99 169 [30]
b b
9 CS / 1400 93 <1 181 113 [31]
8 CS / 1350 98 3.5 107 [32]
8 CS / 1150 87 1.2 85 [32]
8 MS / 1500 99 172 [30]
8 MS / 1200 97 158 [30]
8.7 TF / 1000 - 0.02 334 90 [35]
b: bulk conductivity / activation energy

From the reviewed data in Table 1, no clear conclusions can be made for the
influence of grain size on the conductivity in bulk materials. Factors such
as sintering technique and temperature must also be considered. There are
reports of correlation between grain size and grain boundary conductivity in
zirconia bulk materials, as clearly demonstrated for calcia-substituted
zirconia in Fig. 8 [33].

Many studies have been carried out on conductivity of YSZ thin films,
where nanoscaled grains are more easily obtained compared to bulk
materials. There is strong support for the notion that the energetics for
defect formation may be substantially reduced in nanocrystalline oxides [34].
Enhanced grain boundary conductivity with decreasing grain size is reported
for YSZ films [35,36], similar to the correlation between grain size and

17
conductivity shown in Fig. 8. This effect would be due to greater grain
boundary diffusion than corresponding “bulk” diffusion because of high
defect densities and high mobilities. An additional explanation for
enhanced conduction at grain boundaries is related to the formation of space
charge regions in the grain adjacent to the boundaries [37]. Enhancement of
ionic conductivity in nanocrystalline oxides remains unresolved due to
conflicting reports and inadequate efforts to isolate the ionic from the total
conductivity. No clear lines can be drawn between the grain size effects
observed in thin films and bulk materials, possibly because reduced grain
size in bulk materials is more challenging than for thin films, where the
sintering temperature can be lowered.

Fig. 8 Grain boundary conductivity as a function of grain size for ZrO2


stabilized with 15 mol% CaO [33].

3.1.4 Powder synthesis


Nanosized powders are essential as a starting point for making
nanocrystalline ceramics. According to conventional sintering theory, the
sinterability increases with decreasing crystallite size in the starting powder
[38]
. There are various methods for synthesis of high quality oxide powders;
among them are precipitation techniques [39,40], combustion techniques [41-43],

18
sol-gel techniques [44,45], and hydrothermal techniques [46]. All these
different techniques are based on a solution type chemistry, where
precursors of the various cations are dissolved in a solvent, commonly
water, and then mixed in appropriate proportions.

Oxide ceramic powders prepared by a glycine-nitrate combustion technique,


as described by Chick et al. [47], are superior to powders prepared by e.g.
sol-gel synthesis, showing greater compositional uniformity, lower residual
carbon level and smaller particle size. In this synthesis technique glycine
complexes the cations and may do so on both the carboxylic acid end or by
the amino group depending on the size and charge of the cation as well as
the pH in the solution. The amount of glycine added must be sufficient to
retain the cations completely complexed as the solution is being evaporated.
In addition to being a complexing agent, the glycine works as a fuel and
will, in the presence of nitrates from the metal nitrate precursor, working as
an oxidizer, cause a spontaneous ignition reaction upon evaporation. The
temperature of the ignition reaction depends on the fuel-to-oxidizer ratio
(glycine:nitrate (G/N) ratio) and which metal cations are introduced into the
solution.

Mukasyan et al. [48] have defined three combustion modes depending on the
G/N ratio for La0.8Sr0.2CrO3; Smoldering combustion synthesis (SCS): G/N
< 0.39, T < 600ºC; Volume combustion synthesis (VCS): 0.39 < G/N <
0.66, 1150ºC < T < 1350ºC; Self-propagating high-temperature synthesis
(SHS): 0.66 < G/N < 0.88, 800ºC < T < 1100ºC. According to Chick et al.
[47]
, the best powder quality is achieved when the combustion temperature is
the highest. For La0.8Sr0.2CrO3 the highest combustion temperature was
observed for a G/N ratio in the range of 0.4 – 0.7 however, Kaus et al. has
indicated that the maximum crystallite size and combustion temperature in
the synthesis of nanocrystalline zirconia occurred for a G/N ratio of 0.30 [49].
This report also states that the best powders were obtained for a lower G/N
ration (0.23). As the reaction temperature affects the crystallite size in the
final powder it is desirable to keep this at a minimum in order to prepare
finer powders.

3.1.5 Densification and microstructure


Densification of zirconia materials is widely investigated [26,28-32] and it is
possible to obtain dense (> 95%) specimens by conventional sintering at
temperatures in the range 1300 – 1600°C. Such sintering with long
sintering time (hrs), in particular in the higher end of this temperature range
will result in rather large grained materials. Larger grains may be
undesirable, for example with respect to mechanical and electrical

19
properties. Other sintering techniques such as spark plasma sintering (SPS)
have become more popular for controlling the grain growth during sintering.
This is also indicated by the density/grain size data reviewed in Table 1,
showing average grain size for YSZ materials prepared by SPS in the sub-
micron range. In an SPS experiment both sintering time and temperature
needed to obtain dense materials can be reduced significantly compared to
conventional sintering. This can be ascribed to the effective heat transfer, as
the pressing die (and green body if conducting) works as the heating
element. The applied pulsed electrical field may create spark discharges
during the initial part of the sintering, and clean the particle surface,
facilitate grain boundary diffusion and possibly contribute to increased
densification rate [50].

3.2 Proton conductors - Barium zirconate and


strontium cerate
The research on proton conducting ceramics accelerated after Iwaharas
discovery of relatively high proton conductivity in acceptor doped strontium
cerate (SrCeO3) at high temperatures, in the presence of water vapor or
hydrogen [51]. SrCeO3 materials, along with barium zirconates (BaZrO3) are
reviewed here. As these materials are often compared, due to competing
properties, it has been advantageous to present their properties together in
the following.

3.2.1 Crystal structure


Both BaZrO3 and SrCeO3 exhibit perovskite-type (ABO3 structures) crystal
structures. BaZrO3 exists in the cubic perovskite-structure, illustrated by
Fig. 8 a), while SrCeO3 has a distorted version of this structure, making the
material orthorhombic, as shown in Fig. 8 b).

3.2.2 Electrical properties


The available conductivity data for acceptor substituted SrCeO3 [51-65,69],
generally proves mixed ionic-electronic conduction. No major deviations
from the results of the extensive work on SrCe2.95Yb0.05O3-δ, performed by
Iwahara et al. [51-53,55,60-64,69], have been reported, and the total conductivity
at 1000°C for this material ranges from 0.05 to 0.005 S/cm, depending on
atmosphere. Kosacki and Tuller reported that the electrical properties of
SrCe2.95Yb0.05O3-δ between 800 and 1000°C are dominated by holes and
electrons at high p(O2) (≥ 1 atm) and low p(O2) (< 10-20 atm), respectively

20
[57]
. The proton conductivity in humid air reaches a maximum of ~0.004
S/cm around 900°C [53]. The proton transport number under such conditions
exceeds 0.5 below 800°C, however increases with decreasing p(O2). At
700°C and p(O2) ~5·10-4 atm, the ionic contribution to the conductivity is
predominated by protons [60].

Fig. 8 a) Cubic and b) orthorhombic perovskite structure. Central


atom represent A-site cation, while the B-site cations take
octahedral coordination with oxygen.

The electrical properties of yttrium substituted BaZrO3 are well documented


in the literature [10,11,66-73], and some central data are listed in Table 2. Total
bulk (ionic + electronic) conductivity up to 0.01 S/cm at 800°C in wet O2, is
reported by Bohn and Schober [71]. While hole conduction suppresses the
transport number for protons in oxidizing atmosphere, BaZr0.9Y0.1O2.95 is
close to a perfect proton conductor under reducing conditions [71]. The main
issue regarding ionic conductivity in BaZrO3 materials is the high grain
boundary resistance, inhibiting proton transport, however this is not yet
fully understood. For 10% yttrium substituted BaCeO3, the total
conductivity is about one order of magnitude higher than the corresponding
BaZrO3 material; σtot = 0.7 and 0.06 S/cm, respectively, at 1000°C in wet
hydrogen atmosphere [75]. The idea of combining the improved electrical
properties of BaCeO3 with the mechanical properties and chemical stability
of BaZrO3 has been studied in mixed systems (BaCe0.9-xZrx Y0.1O3-δ). For
these systems increasing chemical stability against CO2 was reported with
increasing Zr content, however, this was accompanied by decreased proton
conductivity [74,75]. Snijkers et al. has reported improved conductivity for
BaZr0.9Y0.1O2.95 specimens prepared with excess of BaO, in the range 200 -

21
300°C, when extrapolating the data to ~500°C the data equals those of
specimens prepared without excess of BaO (~8·10-4 S/cm) [72].

Table 2 Reported conductivity data for BaZr0.9Y0.1O2.95 for 500°C (bulk)


and 600°C (total) under various conditions. Water partial
pressure, p(H2O), is indicated when available.

Sintering σtotal at σbulk at


Density
temperature 600°C 500°C Atmosphere Ref. #
(%)
(°C) (S/cm) (S/cm)

Wet synthetic
1715°C 97 2·10-3 [68]
air

Wet air
1800°C 95 4·10-3 p(H2O) = [66]
1.7·103 Pa
Wet H2
1800°C 95 1.5·10-3 p(H2O) = [66]
1.7·103 Pa

1700°C 95 ~10-3 * 7·10-4 * Ambient air [72]

Wet Ar +
1715°C 97 7·10-4 [68]
4%H2

Wet H2 and
1715°C 97 4·10-3 [71]
wet Ar/O2

Wet
1700°C 91 5·10-2 ** p(H2O) = [10,73]
2.3·103 Pa
* Extrapolated values
** BaZr0.8Y0.2O2.9

3.2.3 Powder synthesis and densification


Most reports concerning preparation of ceramic BaZrO3 and SrCeO3
powders have been by solid state ceramic method at temperatures in the
range 1000 – 1450°C [10,11,55,57,66-73,76-83,94,97]. This method generally
produces coarse powders not suitable for preparation of bulk materials with
custom designed fine grained microstructures. Fine BaZrO3 powders with

22
more defined morphology have been prepared by co-precipitation [84,85], sol-
gel techniques [86,87] and hydrothermal synthesis [87]. Other reported
synthesis routes for preparation of SrCeO3 powders include complexation
with EDTA [88] or citric acid [89], as well as with glycine in a combustion
synthesis route [90]. Powders prepared by complexation with EDTA appear
phase pure after calcination at 1200°C in inert atmosphere (He).

No thorough sintering studies of BaZrO3 and SrCeO3 ceramics seem to be


available. Dense (> 95%) BaZrO3 from powders made by solid state
method can only be obtained by conventional sintering above 1700°C
[67,68,71,72,75]
. Using spark plasma sintering, dense (>95%) ceramics of pure
and Y-substituted BaZrO3 can be obtained at 1500 and 1600°C respectively
[91]
. Reported densities of SrCeO3 and SrCe2.95Yb0.05O3-δ materials prepared
from powders by the solid state ceramic method sintered in air at 1350°C for
96 hrs, are >85 and >75%, respectively [90]. Generally, sintering
temperature in the range of 1500 to 1650°C [57,79,83] in air is needed to obtain
materials with >95% density when prepared from such powders. Higher
densities (>97%) have been obtained from commercial powders by sintering
at 1450°C in air for 2 hrs [92] and from powder prepared by a complexation
route by sintering at 1300°C for 12 hrs in nitrogen atmosphere [88],
confirming the importance of starting powder properties.

3.2.4 Mechanical properties and chemical stability


Chemical, mechanical and electrical properties are all of importance when
considering proton conducting materials for commercial use. Among the
perovskite-type zirconates and cerates of barium and strontium, there has
been a continuous battle arising from higher conductivity in the cerates
versus improved mechanical properties and chemical stability in the
zirconates [93,94]. Young’s modulus (E) and Vickers hardness (HV) for
SrCeO3 (substituted with 5% Yb) are reported to be 145 GPa and 5.5 GPa
[95]
. The mechanical properties of BaZrO3 are improved compared to those
of SrCeO3 with reported values of E = 181 GPa and HV = 11.1 GPa [96]. It
could be mentioned that Hassan reported lower bending strength, hardness
and fracture toughness for SrZrO3 compared to SrCeO3 [95], in contrast to the
trend with improved mechanical properties of zirconates compared to
cerates. The density (crucial for the mechanical properties) for the
presented SrZrO3 material appears to be lower than for SrCeO3.

The chemical instability of SrCeO3 and BaCeO3 materials when exposed to


CO2 atmosphere is well documented by Scholten et al. [94]. In an extensive
review, Kreuer has compared the equilibrium reactions for barium and

23
strontium cerates and zirconates in both CO2 and H2O atmosphere, as
demonstrated by Eq. 3 and 4, respectively.

ABO3 (s) + CO 2 (g) = ACO3 (s) + BO 2 (s) (3)

ABO3 (s) + H 2O(g) = A(OH) 2 (s) + BO 2 (s) (4)

Additionally the decomposition of the resulting carbonates (ACO3) and


hydroxides (A(OH)2) according to Eq. 5 and 6 have been considered.

ACO3 (s) = AO(s) + CO 2 (g) (5)

A(OH)2 (s) = AO(s) + H 2O(g) (6)

The equilibrium constants for these reactions (Eq. 3 – 6), as a function of


partial pressures of CO2 and H2O are presented for SrCeO3 and BaZrO3 in
Fig. 9 a) and b).

In the case of SrCeO3 the equilibrium constants for Eq. 3 and 5, as a


function of p(CO2), are close to identical and both equal unity around 700°C
in air. Hence, heat treatment of SrCeO3 above 700°C in air is more likely to
result in the formation of strontium oxide (SrO) rather than SrCO3. The
onset temperature of decomposition of SrCO3 (according to Eq. 5) obtained
by thermogravimetric analysis have been reported to 845°C in 1 atm N2, and
1220 – 1275°C in 1 atm CO2. The corresponding onset temperatures for
synthesis of SrCeO3 from a mixture of SrCO3 and CeO2, are 800°C and
1168 - 1190°C, in 1 atm N2 and CO2, respectively [94,97]. At lower
temperatures, carbonate formation may be critical, with respect to phase
purity and mechanical properties of SrCeO3 materials. Reactions with water
should also be considered when operating in atmospheres with high p(H2O)
[93]
. BaZrO3 is far more stable in CO2 atmosphere, compared to SrCeO3,
and reaction in Eq. 4 is not significant for water vapour pressure < 10 bar.

For BaZrO3 the high sintering temperatures (>1500°C) needed for obtaining
dense materials is more crucial than the stability against CO2 / H2O, as the
material tend to decompose, according to Eq. 7, as documented by
thermodynamic studies [98,99].

BaZrO3 (s) = ZrO 2 (s) + BaO(g) (7)

24
a)

b)

Fig. 9 Stability diagrams for SrCeO3 and BaZrO3 in CO2 or H2O


atmosphere, as presented by Kreuer [93].

25
Heat treatment of Y-stabilized BaZrO3 bulk materials above 1500°C have
resulted in reduced lattice parameter and yttria/zirconia excess, indicating
evaporation of BaO(g) above this temperature [86]. On the surface of these
BaZrO3 specimens, similar effects were observed at lower temperatures (T >
1250°C) [86]. In reducing atmosphere in contact with graphite during hot
pressing or spark plasma sintering, evaporation of Ba(g), according to Eq. 8,
must also be considered.

BaZrO3 (s) + C(s) = ZrO 2 (s) + CO(g) + Ba(g) (8)

3.3 Thin films of perovskite-type oxides

3.3.1 Lanthanum ferrite thin films


Lanthanum ferrite (LaFeO3) has an orthorhombic distortion of the cubic
perovskite structure, comparable to that of SrCeO3, as shown in Fig. 8 b).
LaFeO3 is an antiferromagnetic insulator at room temperature, making the
material feasible for use in magnetic sensors and as read heads in computer
hard drives [100]. At elevated temperatures LaFeO3 exhibits mixed ionic-
electronic conductivity [101] and a linear response for log σ versus log p(O2)
have been reported for LaFeO3 thin films (at 1000°C), making the material
promising for oxygen sensor applications [102]. Furthermore, LaFeO3 have
been reported to show excellent sensitivity towards non-flammable gases
such as CO [103] and NOx [103-107] and exhibits selective sensitivity towards
flammable gases such as ethanol [108], methane [103] and volatile sulfides
such as CH3SH [109]. LaFeO3 thick films are reported to be promising for
detection of γ-radiation [110]. The gas sensitivity has been reported to
increase with decreasing grain size [111] making the morphology an
important feature in gas sensing films. In the literature, the preparation of
LaFeO3 films is reported by use of sputtering techniques [102,104,112-115],
screen printing of slurries from nanosized powder [105,106,110,116] and
electrochemical reduction [117]. Only sputtering techniques have resulted in
growth of epitaxial LaFeO3 thin films, and rhombohedral (001)-oriented
LaAlO3 substrates [102,112] and cubic (001)-oriented MgO substrates [113] have
been used for this purpose. This observation is not unexpected as sputtering
techniques allows good control of deposition of thin film layers, making for
epitaxial orientation, compared to thicker films (> 0.5 µm) prepared by
screen printing and electrochemical reduction. Randomly oriented LaFeO3
films have also been prepared on quartz substrates by CSD, using a colloidal
sol of hydrous LaFeO3 particles with an average particle size of 7 nm and

26
dip-coating technique [118]. These films were shown to be fully crystalline
and single phase after annealing at 650°C.

3.3.2 Lanthanum cobaltite thin films


Lanthanum cobaltite, LaCoO3, has a rhombohedral distorted version of the
perovskite structure (ABO3). At elevated temperatures (above ~1400°C) the
material undergoes a phase transition from rhombohedral to cubic [119], and
interesting ferroelastic behavior is associated with this phase transition [120-
122]
. LaCoO3 materials exhibit characteristic changes in magnetic and
electrical properties with varying temperature [123]. The properties of
LaCoO3 may be adjusted by substitution with e.g. calcium, strontium or
barium on the A-site [120-122,124], which as an example will lower the
rhombohedral to cubic phase transition temperature (materials become cubic
at ambient temperature for ~50 mol% substitution [119]. Mixed
ionic/electronic conductivity at elevated temperatures, makes LaCoO3-
materials applicable in membrane technology and as cathode materials in
solid oxide fuel cells (SOFC’s) [125]. LaCoO3 materials are also investigated
with respect to catalytic properties [126] and for use in gas sensors [127,128].

The literature reveals thin films of LaCoO3 (pure and Ca-/Sr-substituted)


prepared by different techniques, such as pulsed laser deposition (PLD) [128],
chemical vapour deposition (CVD) [129], atomic layer epitaxy (ALE) [130],
electrochemical oxidation [131], ion beam sputtering [132] and by spray
pyrolysis in inductively coupled plasma (spray-ICP) [133]. LaCoO3 films
have also been prepared from various sol-gel precursor solutions by spin- or
dip-coating [134-143]. Sol-gel precursor solutions have been synthesized by
traditionally accepted routes using metal alkoxides dissolved in alcohol with
proper chelating agents, such as 2-ethylacetoacetate or polyethylene glycol
(PEG) [134-136]. Less expensive variants using nitrate salts or acetates as
starting materials for alcohol-based precursor solutions, with e.g.
butylacetate or polyvinyl alcohol (PVA) as chelating agents, have also been
reported [137-141]. LaCoO3 films have also been prepared from water based
routes using ethylendiaminetetraacetic acid (EDTA) or
[141-144]
diethylenetriamineoentaacetic acid (DTPA) as complexing agents .
None of these so-called chemical solution deposition (CSD) techniques have
resulted in epitaxial growth of the films. Due to the often superior
properties of crystalline epitaxial films, these are more attractive for
applications and fundamental studies, compared to randomly oriented
polycrystalline films [145].

27
References
[1] B.C.H. Steel, Oxygen ion conductors and their technological
applications, Mater. Science and Engineering, B13, 79-87 (1992)
[2] J. A. Kilner, Fast oxygen transport in acceptor doped oxides, Solid
State Ionics , 129 [1-4] , 13-23 (2000)
[3] B.C.H. Steele, Oxygen ion conductors, in High Conductivity Solid
Ionic Conductors, Recent Trends and Applications (T. Takahashi,
ed.), World Scientific, Singapore, 402-446 (1989)
[4] J.E. Bauerle, Study of solid electrolyte polarization by a complex
admittance method, J. Phys. Chem. Solids, 30, 2657-2670 (1969)
[5] R. Gerhardt and A. S. Nowick, Grain-boundary effect in ceria doped
with trivalent cations. 1. Electrical measurements, J. Am. Ceram.
Soc., 69 [9], 641-646 (1986)
[6] D. Y. Wang and A. S. Nowick, The grain-boundary effect in doped
ceria solid electrolytes, J. Solid State Chem., 35 [3] (1980) 325-333.
[7] J. Maier, Ionic-conduction in-space charge regions, Prog. Sol. State
Chem., 23 [3], 171-263 (1995)
[8] T. Norby, Solid-state protonic conductors: principles, properties,
progress and prospects, Solid State Ionics, 125, 1-10 (1999)
[9] I. Riess, D.S. Tannhauser, Mixed Ionic Electronic Conductors, in
High Conductivity Solid Ionic Conductors, Recent Trends and
Applications (T. Takahashi, ed.), World Scientific, Singapore, 402-
446 (1989)
[10] K.D. Kreuer, Proton-conducting oxides, Annu. Rev. Mater. Res., 33,
333-359 (2003)
[11] T. Norby, Y. Larring, Concentration and transport of protons in
oxides, Curr. Opin. Solid State & Mater. Sci., 2, 593-599 (1997)
[12] O. Yamamoto, Solid oxide fuel cells: fundamental aspects and
prospects, Electrochimica Acta, 45 [15-16], 2423-2435 (2000)
[13] T. Norby, R. Haugsrud, N. Vajeeston, New Proton Conducting
Materials for Fuel Cells and Hydrogen Separation Membranes,
Proc. 9th Int. Conf. Inorg. Membr., Lillehammer – Norway, 260-267
(2006)
[14] http://rtreport.ksc.nasa.gov/techreports/2003report/100/106.html
[15] M. Gelfi, E. Bontempi, R. Roberti, L. Armelao, L.E. Depero,
Residual stress analysis of thin films and coatings through XRD2
experiments, Thin Solid Films, 450, 143-147 (2004)
[16] R.W. Schwartz, Chemical solution deposition of perovskite thin
films, Chem. Mater., 9, 2325-2340 (1997)
[17] E. C. Subbarao, Solid electrolytes and their applications, ed. by E.C.
Subbarao, Plenum Press (1980)

28
[18] D.W. Richerson, Modern Ceramic Engineering, Marcel Dekker Inc.,
2nd ed. (1992)
[19] E.C. Subbarao, Zirconia - an overview, Science and Technology of
Zirconia, 3, 1–23 (1981)
[20] V.S. Stubican, R.C. Hink, S.P. Ray, Phase equilibria and ordering in
the system ZrO2 – Y2O3, J. Am. Ceram. Soc., 61 [1-2], 17-21 (1978)
[21] N. Q. Minh and T. Takahashi, Science and Technology of Ceramic
Fuel Cells, Elsevier Science B. V., Amsterdam, 70-92 (1995)
[22] X. Guo, R.-Z. Yuan, On the grain boundaries of ZrO2-based solid
electrolyte, Solid State Ionics, 80, 159-166 (1995)
[23] T. Takahashi, Physics of Electrolytes, Vol. 2, (J. Hladic ed.),
Academic Press, London, 980-1049 (1972)
[24] A. Weyl, D. Janke, High-temperature ionic conduction in
multicomponent solid oxide solutions based on zirconia, J. Am.
Ceram. Soc., 80 [4], 861-873 (1997)
[25] X.J. Chen, K.A. Khor, S.H. Chan, L.G. Yu, Preparation yttria-
stabilized zirconia electrolyte by spark-plasma sintering, Mater. Sci.
Eng., A341, 43-48 (2003)
[26] T. Takeuchi, I. Kondoh, N. Tamari, N. Balakrishnan, K. Nomura, H.
Kageyama, Y. Takeda, Improvement of mechanical strength of 8
mol% yttria-stabilized zirconia ceramics by spark-plasma sintering,
J. Electrochem. Soc., 149 [4], A455-A461 (2002)
[27] U. Anselmi-Tamburini, Spark plasma sintering and characterization
of bulk nanostructured fully stabilized zirconia: Part II.
Characterization studies, J.E. Garay, Z.A. Munir, J. Mater. Res., 19
[11], 3263-69 (2004)
[28] P.I. Dahl, I. Kaus, K. Wiik, Z. Zhao, M. Johnsson, M. Nygren, T.
Grande, M.-A. Einarsrud, Densification and properties of zirconia
prepared by three different sintering techniques, Ceram. Int. (2006) -
Accepted
[29] I.R. Gibson, G.P. Dransfield, J.T. Irvine, Influence of yttria
concentration upon electrical properties and susceptibility to ageing
of yttria-stabilised zirconias, J. Europ. Ceram. Soc., 18, 661-667
(1998)
[30] F.T. Ciacchi, S.A. Nightingale, S.P.S. Badwal, Microwave sintering
of zirconia-yttria electrolytes and measurement of their ionic
conductivity, Solid State Ionics, 86-88, 1167-1172 (1996)
[31] R. Ramamoorthy, D. Sundararaman, S. Ramasamy, Ionic
conductivity of ultrafine-grained yttria stabilized zirconia
polymorphs, Solid State Ionics, 123, 271-278 (1999)
[32] X.J. Chen, K.A. Khor, S.H. Chan, L.G. Yu, Influence of
microstructure on the ionic conductivity of yttria-stabilized zirconia
electrolyte, Mater. Sci. Eng., A335, 246-252 (2002)

29
[33] M. Aoki, Y.-M. Chiang, I. Kosacki, J.-R. Lee, H.L. Tuller, Y.J. Liu,
Solute segregation and grain-boundary impedance in high-purity
stabilized zirconia, J. Am. Ceram. Soc., 79, 1169-1180 (1996)
[34] H.L. Tuller, Ionic conduction in nanocrystalline materials, Solid
State Ionics, 131, 143-157 (2000)
[35] I. Kosacki, H.U. Anderson, Microstructure – property relationships
in nanocrystalline oxide thin films, Ionics, 6, 294-311 (2000)
[36] I. Kosacki, T. Suzuki, V. Petrovsky, H.U. Anderson, Electrical
conductivity of nanocrystalline ceria and zirconia thin films, Solid
State Ionics, 136-137, 1225-1233 (2000)
[37] J. Maier, Space-charge regions in solid 2-phase systems and their
conduction contribution. 3. Defect chemistry and ionic-conductivity
in thin films, Solid state ionics, 23, 59-67 (1987)
[38] R.M. German, Sintering Theory and Practice, John Wiley & Sons,
Inc., New York, 104-110 (1996)
[39] S.K. Tadokoro, E. N. S. Muccillo, Physical characteristics and
sintering behavior of ultrafine zirconia-ceria powders, J. Europ.
Ceram. Soc., 22, 1723-1728 (2002)
[40] S.-G. Chen, Y.-S. Yin, D.-P. Wang , J. Li, Reduced activation
energy and crystalline size for yttria-stabilized zirconia nano-
crystals: an experimental and theoretical study, J. Cryst. Growth,
267 [1-2], 100-109 (2004)
[41] S. Jiang, W. A. Schulze, V. R. W. Amarakoon, G. C. Stangle,
Electrical properties of ultrafine-grained yttria-stabilized zirconia
ceramics, J. Mater. Res., 12 [9], 2374-2380 (1997)
[42] Y. Wu, A. Bandyopadhyay, S. Bose, Processing of alumina and
zirconia nano-powders and compacts, Mater. Sci. Eng. A, 380, 349-
355 (2004)
[43] K.C. Patil, S.T. Aruna and T. Mimani, Combustion synthesis: an
update, Curr. Opin. Solid State Mat. Sci., 6, 507-512 (2002)
[44] C. Laberty-Robert, F. Ansart, C. Deloget, M. Gaudon, A. Rousset,
Powder synthesis of nanocrystalline ZrO2-8%Y2O3 via a
polymerization route, Mater. Res. Bull., 36, 2083-2101 (2001)
[45] C. Laberty-Robert, F. Ansart, C. Deloget, M. Gaudon, A. Rousset,
Dense yttria stabilized zirconia: sintering and microstructure,
Ceram. Int., 29, 151-158 (2003)
[46] Y. B. Khollam, A. S. Deshpande, A. J. Patil, H. S. Potdar, S. B.
Deshpande, S. K. Date, Synthesis of yttria stabilized cubic zirconia
(YSZ) powders by microwave-hydrothermal route, Mater. Chem.
Phys., 71 [3], 235-241 (2001)
[47] L. A. Chick, L. R. Pederson, G. D. Maupin, J. L. Bates, L. E.
Thomas and G. J. Exarhos, Glycine-nitrate combustion synthesis of
oxide ceramic powders, Mater. Lett., 10 [1-2], 6 (1990)

30
[48] A. S. Mukasyan, C. Costello, K. P. Sherlock, D. Lafarga, A. Varma,
Perovskite membranes by aqueous combustion synthesis: synthesis
and properties, Sep. Pur. Tech., 25, 117-126 (2001)
[49] I. Kaus, P.I. Dahl, J. Mastin, T. Grande, M.-A. Einarsrud, Synthesis
and characterization of nanycrystalline YSZ powder by smoldering
combustion synthesis, J. Nanomater., 2006, 1-7 (2006)
[50] Z. Shen, M. Johnsson, Z. Zhao, M. Nygren, Spark plasma sintering
of alumina, J. Am. Ceram. Soc., 85 [8], 1921-27 (2002)
[51] H. Iwahara, T. Esaka, H. Uchida and N. Maeda, Proton conduction
in sintered oxides and its application to steam electrolysis for
hydrogen-production, Solid State Ionics, 3-4, 359-363 (1981)
[52] H. Iwahara, Technological challenges in the application of proton
conducting ceramics, Solid State Ionics, 77, 289-298 (1995)
[53] T. Yajima, H. Iwahara, Studies on proton behavior in doped
perovskite-type oxides: (II) Dependence of equilibrium hydrogen
concentration and mobility on dopant content in Yb-doped SrCeO3,
Solid State Ionics, 53-56, 983-988 (1992)
[54] T. Scherban, A.S. Nowick, Bulk protonic conduction in Yb-doped
SrCeO3, Solid State Ionics, 35, 189-194 (1989)
[55] N. Matsunami, T. Yajima, H. Iwahara, Permeation of implanted
deuterium through SrCeO3 (5% Yb), Nucl. Instr. Meth. Phys. Res.,
B65, 278-281 (1992)
[56] I. Kosacki, J. Schoonman, M. Balanski, Raman scattering and ionic
transport in SrCe1-xYbxO3, Solid State Ionics, 57, 345-351 (1992)
[57] I. Kosacki, H.L. Tuller, Mixed conductivity in SrCe0.95Yb0.05O3
protonic conductors, Solid State Ionics, 80, 223-229 (1995)
[58] U. Reichel, R.R. Arons, W. Schilling, Investigation of n-type
electronic defects in the protonic conductor SrCe1-xYxO3-α,, Solid
State Ionics, 86-88, 639-645 (1996)
[59] T. Matzeke, M. Cappadonia, Proton conductive perovskite solid
solutions with enhanced mechanical stability, Solid State Ionics, 86-
88, 659-663 (1996)
[60] H. Uchida, N. Maeda, H. Iwahara, Relation between proton and hole
conduction in SrCeO3-based solid electrolytes under water-
containing atmospheres at high temperatures, Solid State Ionics, 11,
117-124 (1983)
[61] H. Uchida, H. Yoshikawa, T Eseka, S. Ohtsu, H. Iwahara, Formation
of protons in SrCeO3-based proton conducting oxides. Part II.
Evaluation of proton concentration and mobility in Yb-doped
SrCeO3, Solid State Ionics, 36, 89-95 (1989)
[62] T. Yajima, H. Iwahara, H. Uchida, K. Koide, Relation between
proton conduction and concentration of oxide ion vacancy in

31
SrCeO3 based sintered oxides, Solid State Ionics, 40/41, 914-917
(1990)
[63] T. Hibino, K. Mizutani, T. Yajima, H. Iwahara, Evaluation of proton
conductivity in SrCeO3, BaCeO3, CaZrO3 and SrZrO3 by
temperature programmed desorption method, Solid State Ionics, 57,
303-306 (1992)
[64] H. Iwahara, Proton conducting ceramics and their applications,
Solid State Ionics, 86-88, 9-15 (1996)
[65] T. Schober, F. Krug, W. Schilling, Criteria for the application of
high temperature proton conductors in SOFCs, Solid State Ionics,
97, 369-373 (1997)
[66] R.C.T. Slade, S.D. Flint, N. Singh, Investigation of protonic
conduction in Yb- and Y-doped barium zirconates, Solid State Ionics,
82, 135-141 (1995)
[67] K.D. Kreuer, St. Adams, W. Münch, A. Fuchs, U. Klock, J. Maier,
Proton conducting alkaline eart zirconates and titanates for high
drain electrochemical applications, Solid State Ionics, 145, 295-306
(2001)
[68] T. Schober, H.G. Bohn, Water vapor solubility and electrochemical
characterization of the high temperature proton conductor
BaZr0.9Y0.1O2.95, Solid State Ionics, 127, 351-360 (2000)
[69] H. Iwahara, T. Yajima, T. Hibino, K. Ozaki, H. Suzuki, Protonic
conduction in calcium, strontium and barium zirconates, Solid State
Ionics, 61, 65-69 (1993)
[70] W. Wang, A.V. Virkar, Ionic and electron-hole conduction in
BaZr0.93Y0.07O3-δ by 4-probe dc measurements, J. Power Sources,
142, 1-9 (2005)
[71] H.G. Bohn, T. Schober, Electrical conductivity of the high-
temperature proton conductor BaZr0.9Y0.1O2.95, J. Am. Ceram. Soc.,
83 [4], 768-772 (2000)
[72] F.M.M. Snijkers, A. Nuekenhoudt, J. Cooymans, J.J. Luyten, Proton
conductivity and phase composition in BaZr0.9Y0.1O3-δ, Scripta
Materialia, 50, 655-659 (2004)
[73] K.D. Kreuer, Aspects of the formation and mobility of protonic
charge carriers and the stability of perovskite-type oxides, Solid
State Ionics, 125 [1-4], 285-302 (1999)
[74] K.H. Ryu, S.M. Haile, Chemical stability and proton conductivity of
doped BaCeO3 – BaZrO3 solid solutions, Solid State Ionics, 125,
355-367 (1999)
[75] K. Katahira, Y. Kohchi, T. Shimura, H. Iwahara, Protonic
conduction in Zr-substituted BaCeO3, Solid State Ionics, 138, 91-98
(2000)

32
[76] H. Iwahara, T. Esaka, H. Uchida, N. Maeda, Proton conduction in
sintered oxides and it’s application to steam electrolysis for
hydrogen production, Solid State Ionics, 3/4, 359-363 (1981)
[77] T. Scherban, A.S. Nowick, Bulk protonic conduction in Yb-doped
SrCeO3, Solid State Ionics, 35, 189-194 (1989)
[78] I. Kosacki, J. Schoonman, M. Balanski, Raman scattering and ionic
transport in SrCe1-xYbxO3, Solid State Ionics, 57, 345-351 (1992)
[79] H. Iwahara, T. Yajima, T. Hibino, K. Ozaki, H. Suzuki, Protonic
conduction in calcium, strontium and barium zirconates, Solid State
Ionics, 61, 65-69 (1993)
[80] U. Reichel, R.R. Arons, W. Schilling, Investigation of n-type
electronic defects in the protonic conductor SrCe1-xYxO3-α,, Solid
State Ionics, 86-88, 639-645 (1996)
[81] T. Matzeke, M. Cappadonia, Proton conductive perovskite solid
solutions with enhanced mechanical stability, Solid State Ionics, 86-
88, 659-663 (1996)
[82] S.V. Chavan, A.K. Tyagi, Sub-solidus phase equilibria in CeO2-SrO
system, Thermochimica Acta, 390, 79-82 (2002)
[83] H. Matsumoto, T. Shimura, H. Iwahara, T. Higuchi, K. Yashiro, A.
Kaimai, T. Kawada, J. Mizusaki, Hydrogen separation using proton-
conducting perovskites, J. All. Comp., 408-412, 456-462 (2006)
[84] J. Brzezinska-Miecznik. K. Haberko, M.M. Bucko, Barium zirconate
ceramic powder synthesis by the coprecipitation – calcination
technique, Mater. Lett., 56, 273-278 (2002)
[85] F. Boschini, B. Robertz, A. Rulmont, C. Cloots, Preparation of
nanosized barium zirconate powder by thermal decomposition of
urea in an aqueous solution containing barium and zirconium, and
by calcination of the precipitate, J. Eur. Ceram. Soc., 23, 3035-3042
(2003)
[86] A. Magrez, T. Schober, Preparation, sintering and water
incorporation of proton conducting BaZr0.9Y0.1O3-δ: comparison
between three different synthesis techniques, Solid State Ionics, 175,
585-588 (2004)
[87] P.P. Phule, D.C. Grundy, Pathways for the low temperature
synthesis of nano-sized crystalline barium zirconate, Mater. Sci.
Eng., B23, 29-35 (1994)
[88] K.J. de Vires, Electrical and mechanical properties of proton
conducting SrCe0.95Yb0.05O3-α, Solid State Ionics, 100, 193-200
(1997)
[89] S. Cheng, V.K. Gupta, J.Y.S. Lin, Synthesis and hydrogen
permeation properties of asymmetric proton-conducting ceramic
membranes, Solid State Ionics, 176, 2653-2662 (2005)

33
[90] S.V. Chavan, A.K. Tyagi, Preparation of Sr0.09Ce0.91O1.91, SrCeO3,
and Sr2CeO4 by glycine-nitrate combustion: Crucial role of oxidant-
to-fuel ratio, J. Mater. Res., 19 [11], 3181-3188 (2004)
[91] U. Anselmi-Tamburini, M.T. Buscaglia, M. Viviani, M. Bassoli, C.
Bottini, V. Buscaglia, P. Nanni, Z.A. Munis, Solid-state synthesis
and spark plasma sintering of submicron BaYxZr1-xO3-x/2 (x = 0, 0.08
and 0.16) ceramics, J. Eur. Ceram. Soc., 26, 2313-2318 (2006)
[92] H. Taherparvar, J.A. Kilner, R. Baker, M. Sahibzada, Solid State
Ionics, 162-163, 297-303 (2003)
[93] K.D. Kreuer, On the development of proton conducting materials for
technological applications, Solid State Ionics, 97, 1-15 (1997)
[94] M.J. Scholten, J. Schoonman, J.C. van Miltenburg, H.A.J. Oonk,
Synthesis of strontium and barium cerate and their reaction with
carbon dioxide, Solid State Ionics, 61, 83-91 (1993)
[95] D. Hassan, S. Janes, R. Clasen, Proton-conducting ceramics as
electrode/electrolyte materials for SOFC's - part 1: preparation,
mechanical and thermal properties of sintered bodies, J. Europ.
Ceram. Soc., 23, 221-228 (2003)
[96] K.J. de Vires, Electrical and mechanical properties of proton
conducting SrCe0.95Yb0.05O3-α, Solid State Ionics, 100, 193-200
(1997)
[97] A.N. Shirsat, K.N.G. Kaimal, S.R. Bharadwaj, D. Das,
Thermodynamic stability of SrCeO3, J. Solid State Chem., 177,
2007-2013 (2004)
[98] T. Tsuneo, S. Stølen, H. Yokoi, Thermodynamic study of barium
zirconates by mass-spectrometry, J. Nucl. Mater., 209, 174-179
(1994)
[99] T. Tsuneo, Thermodynamic properties of ternary barium oxides,
Thermochimica Acta, 253, 155-165 (1995)
[100] J.B. Kortright, D.D. Awschalom, J. Stöhr, S.D. Bader, Y.U, Idzerda,
S.S.P. Parkin, I.K. Schuller, H.-C. Siegmann, Research frontiers in
magnetic materials at soft X-ray synchrotron radiation facilities, J.
Magnetism Magn. Mater., 207, 7-44 (1999)
[101] O. Yamamoto, Y. Takeda, R. Kanno, M. Noda, Perovskite-type
oxides as oxygen electrodes for high temperature oxide fuel cells,
Solid State Ionics, 23, 241-246 (1987)
[102] I. Hole, T. Tybell, J.K. Grepstad, I. Wærnhus, T. Grande, K. Wiik,
High temperature transport kinetics in heteroepitaxial LaFeO3 thin
films, Solid State Electronics, 47, 2279-2282 (2003)
[103] N.N. Toan, S. Saukko, V. Lantto, Gas sensing with semiconducting
perovskite oxide LaFeO3, Physica B: Condensed Matter, 327, 279-
282 (2003)

34
[104] E. Traversa, S. Matsushima, G. Okada, Y. Sadaoka, Y. Sakai, K.
Watanabe, NO2 sensitive LaFe3 thin films prepared by r.f. sputtering,
Sensors and Actuators B, 25, 661-664 (1995)
[105] M. Carotta, M. Butturi, G. Martinelli, Y. Sadaoka, P. Nunziante, E.
Traversa, Microstructural evolution of nanosized LaFeO3 powders
from the thermal decomposition of a cyano-complex for thick film
gas sensors, Sensors and Actuators B, 44, 590-594 (1997)
[106] J. Yoon, M. Grilli, E. Di Bartolomeo, R. Polini, E. Traversa, The
NO2 response of solid electrolyte sensors made using nano-sized
LaFeO3 electrodes, Sensors and Actuators B, 76, 483-488 (2001)
[107] H. Aono, E. Traversa, M. Sakamoto, Y. Sadaoka, Crystallographic
characterization and NO2 gas sensing property of LnFeO3 prepared
by thermal decomposition of Ln---Fe hexacyanocomplexes,
Ln[Fe(CN)6]·nH2O, Ln = La, Nd, Sm, Gd, and Dy, Sensors and
Actuators B, 94, 132-139 (2003)
[108] S. Zhao, J. Sin, B. Xu, M. Zhao, Z. Peng, H. Chai, A high
performance ethanol sensor based on field-effect transistor using a
LaFeO3 nano-crystalline thin-film as a gate electrode, Sensors and
Actuators B, 64, 83-87 (2000)
[109] C. Xiangfeng, P. Siciliano, CH3SH-sensing characteristics of
LaFeO3 thick-film prepared by co-precipitation method, Sensors and
Actuators B, 94, 197-200 (2003)
[110] K. Arshak, O. Korostynska, S. Clifford, Screen printed thick films of
NiO and LaFeO3 as gamma radiation sensors, Sensors and
Actuators A, 110, 354-360 (2004)
[111] C. Xu, J. Tamaki, N. Miura, N. Yamazoe, Grain size effects on gas
sensitivity of porous SnO2-based elements, Sensors and Actuators B,
3, 147-155 (1991)
[112] J.K. Grepstad, Y. Takamura, A. Scholl, I. Hole, Y. Suzuki, T.
Tybell, Effects of thermal annealing in oxygen on the
antiferromagnetic order and domain structure of epitaxial LaFeO3
thin films, Thin Solid Films, 486, 108-112 (2005)
[113] Y.-H. Lee, J.-M. Wu, Epitaxial growth of LaFeO3 thin films by RF
magnetron sputtering, J. Crystal Growth, 263, 436-441 (2004)
[114] J.P. Locquet, J. Perret, J. Fompeyrine, E. Mächler, J.W. Seo, G. Van
Tendeloo, Doubling the critical temperature of La1.9Sr0.1CuO4
using epitaxial strain, Nature, 394, 453-456 (1998)
[115] A. Scholl, J. Stöhr, J. Lüning, J.W. Seo, J. Fompeyrine,H. Siegwart,
J.P. Locquet, F. Nolting, S. Anders, E.E. Fullerton, M.R. Scheinfein,
H.A. Padmore, Observation of antiferromagnetic domains in
epitaxial thin films, Science, 287, 1014-1016 (2000)
[116] E. Traversa, Y. Sadaoka, M. Carotta, G. Martinelli, Environmental
monitoring field tests using screen-printed thick-film sensors based

35
on semiconducting oxides, Sensors and Actuators B, 65, 181-185
(2000)
[117] Y. Mastumoto, J. Hombo, Preparation of LaFeO3 perovskite film
using electrochemical reduction, J. Electroanal. Chem., 348, 441-
445 (1993)
[118] M. Rajendran, M.G. Krishna, A.K. Bhattacharya, Fabrication and
characterization of aqueous sol-gel-derived LaFeO3 thin films,
Modern Physics Lett. B, 14 (22-23), 801-808 (2000)
[119] J. Mastin, M.-A. Einarsrud, T. Grande, Crystal structure and thermal
properties of La1-xCaxCoO3-delta (0 <= x <= 0.4), Chem. Mater.,
18, 1680-1687 (2006)
[120] K. Kleveland, N. Orlovskaya, T. Grande, A.M.M. Moe, M.-A.
Einarsrud, Ferroelastic behaviour of LaCoO3-based ceramics, J.
Am. Ceram. Soc., 84, 2029-2033 (2001)
[121] S. Faaland, P.E. Wullum, R. Holmestrand, T. Grande, M.-A.
Einarsrud, Stress-strain behaviour during compression of
polycrystalline La1-xCaxCoO3-based ceramics, J. Am. Ceram. Soc.,
88 [3], 726-730 (2005)
[122] J. Mastin, H.L. Lein, T. Grande, M.-A. Einarsrud, Mechanical
properties of LaCoO3-based materials, to be published
[123] M.A. Señarís-Rodríguez, J.B. Goodenough, Magnetic and transport
properties of the system La1-xSrxCoO3-δ (0 < x ≤ 0.50), J. Solid State
Chem., 118, 323-336 (1995)
[124] M.A. Señarís-Rodríguez, M.P. Breijo, S. Castro, C. Rey, M.
Sanchez, R.D. Sanchez, J. Mira, A. Fondado, J. Rivas, Peculiarities
in the electrical and magnetic properties of cobalt perovskites Ln1-
3+ 3+ 2+ 2+
xMxCoO3 (Ln : La , M : Ca , Sr2+ , Ba2+ ; Ln3+ : Nd3+, M2+ :
2+
Sr ), Int. J. Inorg. Mater., 1, 281-287 (1999)
[125] V.V. Kharton, E.N. Naumovich, A.V. Kovalevsky, A.P. Viskup,
F.M. Figueiredo, I.A. Bashmakov, F.M.B Marques, Mixed electronic
and ionic conductivity of LaCo(M)O3 (M=Ga, Cr, Fe or Ni), Solid
State Ionics, 138, 135-148 (2000)
[126] R.N. Singh, B. Lal, High surface area lanthanum cobaltate and its A
and B sites substituted derivatives for electrocatalysis of O2
evolution in alkaline solution, Int. J. Hydrogen Energy, 27, 45-55
(2002)
[127] E. Brosha, R. Mukundan, D.R. Brown, F. H. Garzon, J.H. Visser, M.
Zanini, Z. Zhou, E.M. Logothetis, CO/HC sensors based on thin
films of LaCoO3 and La0.8Sr0.2CoO3-8 metal oxides, Sensors and
Actuators B, 69, 171-182 (2000)
[128] D.T.V. Anh, W. Olthuis, P. Bergveld, Sensing properties of
perovskite oxide La0.5Sr0.5CoO3-8 obtained by using pulsed laser
deposition, Sensors and Actuators B, 103, 165-168 (2004)

36
[129] M. Losurdo, A. Sacchetti, P. Capezzuto, G. Bruno, L. Armelao, D.
Batteca, G. Bottaro, A. Gasparotto, C. Maragno, E. Tondello,
Optical and electrical properties of nanostructure LaCoO3 thin
films, App. Phys. Lett., 87, 0601909 (2005)
[130] H. Seim, M. Nieminen, L. Niinistö, H. Fjellvåg, L.-S. Johansson,
Growth of LaCoO3 thin films from β-diketonate precursors, Appl.
Surface Sci., 112, 243-250 (1997)
[131] Y. Matsumoto, T. Sesaki, J. Hombo, A new preparation method of
LaCoO3 perovskite using electronic oxidation, Inorg. Chem., 31,
738-741 (1992)
[132] T. Hattori, T.Matsui, H. Tsuda, H. Mabuchi, K. Morii, Fabrication
and electric properties of LaCoO3 thin films by ion-beam sputtering,
Thin Solid Films, 388, 183-188 (2001)
[133] H. Ichinose, H. Katsuki, M. Nagano, Deposition of LaMO3(M=Co,
Cr, Al) films by spray pyrolysis in inductively coupled plasma, J.
Crystal Growth, 144, 59-64 (1994)
[134] H.J. Hwang, J. Moon, M. Awano, K. Maeda, Sol-Gel Route to
porous lanthanum cobaltite (LaCoO3) thin films, J. Am. Ceram.
Soc., 83, 2852-2854 (2000)
[135] H.J. Hwang, M. Awano, Preparation of LaCoO3 catalytic thin film
by the sol-gel process and its NO decomposition characteristics, J.
Europ. Ceram. Soc., 21, 2103-2107 (2001)
[136] H.J. Hwang, A. Towata, M. Awano, K. Maeda, Sol-Gel route to
perovskite-type Sr-substituted LaCoO3 thin films and effects of
polyethylene glycol on microstructure evolution, Scripta Mater., 44,
2173-2177 (2001)
[137] B. Trummer, O. Fruhwirth, K. Reichmann, M. Holzinger, W. Sitte,
P. Pölt, Preparation and characterisation of LaNixCo1-xO3, J. Europ.
Ceram. Soc., 19, 827-829 (1999)
[138] S. Javorič, G. Dražič, M. Kosec, J. Europ. Ceram. Soc., 21, 1543-
1546 (2001)
[139] E. Bontempi, L. Armelao, D. Barreca, L. Bertolo, G. Bottaro, E.
Pierangelo, L.E. Depero, Structural characterized of sol-gel
lanthanum, cobaltite thin films, Crystal Eng., 5, 291-298 (2002)
[140] M. Gelfi, E. Bontempi, R. Roberti, L. Armelao, L.E. Depero,
Residual stress analysis of thin films and coatings through XRD2
experiments, Thin Solid Films, 450, 143-147 (2004)
[141] Y. Zhang, Y. Zhu, R. Tan, W. Yao, L. Cao, Influence of PEG
additive and precursor concentration on the preparation of LaCoO3
film with peroskite structure, Thin Solid Films, 388, 160-164 (2001)
[142] Y. Zhu, R. Tan, J. Feng, S. Ji, L. Cao, The reaction and poisoning
mechanism of SO2 and perovskite LaCoO3 film model catalysts,
Appl. Catalysis A: General, 209, 71-77 (2001)

37
[143] L. Hong, X. Chen, Z. Cao, Preparation of a perovskite
La0.2Sr0.8CoO3-x membrane on a porous MgO substrate, J. Europ.
Ceram. Soc., 21, 2207-2215 (2001)
[144] Y. Zhu, R. Tan, T. Yi, S. Ji, X. Ye, L. Cao, Preparation of nanosized
LaCoO3 perovskite oxide using amorphous heteronuclear complex
as a precursor at low temperature, J. Mater. Sci., 35, 5415-5420
(2000)
[145] D. P. Norton, Synthesis and properties of epitaxial electronic oxide
thin-film materials, Mater. Sci. Eng. R, 43, 139-247 (2004)

38
PAPER I
40
41
42
43
44
45
46
47
48
PAPER II
50
Densification and properties of zirconia prepared
by three different sintering techniques

Paul Inge Dahl, Ingeborg Kaus, Kjell Wiik,


Tor Grande, Mari-Ann Einarsrud †

Department of Materials Science and Engineering


Norwegian University of Science and Technology
N-7491 Trondheim, Norway

Zhen Zhao, Mats Johnsson, Mats Nygren

Inorganic Chemistry, Arrhenius Laboratory


Stockholm University
SE-106 91 Stockholm, Sweden

Abstract
Densification of nanocrystalline yttria stabilized zirconia (YSZ) powder
with 8 mol% Y2O3, prepared by a glycine/nitrate smoldering combustion
method, was investigated by spark plasma sintering, hot pressing and
conventional sintering. The spark plasma sintering technique was shown to
be superior to the other methods giving dense materials (≥ 96%) with
uniform morphology at lower temperatures and shorter sintering time. The
grain size of the materials was 0.21, 0.37 and 12 µm after spark plasma
sintering, hot pressing and conventional sintering, respectively. Total
electrical conductivity of the materials showed no clear correlation with the
grain size, but the activation energy for spark plasma sintered materials was
slightly higher than for materials prepared by the two other densification
methods. The hardness, measured by the Vickers indentation method, was
found to be independent on grain size while fracture toughness, derived by
the indentation method, was slightly decreasing with increasing grain size.

Keywords: YSZ, densification, grain size, electrical conductivity, hardness,


fracture toughness.

Corresponding author. E-mail: Mari-Ann.Einarsrud@material.ntnu.no

51
Introduction
Yttria stabilized zirconia (YSZ) with its high ionic and low electronic
conductivity is generally the material of choice as an electrolyte in solid
oxide fuel cells (SOFC). In order to improve the conductivity and thereby
lower the operating temperature alternative electrolyte materials (i.e. doped
CeO2) have been suggested but also often rejected due to poor mechanical
properties. Improved conductivity in the traditional YSZ materials is
therefore desired. Increasing conductivity with decreasing grain size has
been reported for YSZ thin films [1,2] and nanocrystalline YSZ bulk
materials may therefore give the desired improvement of the ionic
conductivity. It is not given, however, that the properties observed for thin
films can be directly converted into nanostructured bulk material properties.
The reported changes in ionic conductivity for bulk YSZ with different
grain size are small (or not existing) [3-8]. Factors such as density, sintering
technique and effects due to impurities must be considered before
concluding with an effect of grain size.

Thin films with grain size down to ~10 nm have been prepared [2]. The
preparation of nanocrystalline bulk materials is more challenging since
higher temperatures and longer sintering times compared to those for films
are usually needed. The grain growth normally occurs during the final stage
of the sintering process were the temperature is highest. In hot pressing
(HP) and spark plasma sintering (SPS) uniaxial pressure is applied during
sintering allowing densification at reduced temperature and time and
thereby suppressing the grain growth. The SPS technique, where a pulsed
direct current is passed through an electrical conducting pressing die
working as the heating element gives more rapid densification rate due to
the use of pressure and rapid heating rate. The presence of a pulsed
electrical field that during the initial part of the sintering might create sparks
that clean the particles surface and thus facilitate grain boundary diffusion
and electrical field enhanced diffusion processes might also contribute to
increased densification rate [9].

The objective of the present work is to evaluate different densification


methods for the preparation of dense fine grained bulk YSZ materials.
Moreover, possible effects of the grain size on the total conductivity and the
mechanical properties of these materials are also explored.

52
Experimental
Sample preparation
Nanocrystalline YSZ powder (8 mol% Y2O3) was prepared as previously
described using a glycine/nitrate smoldering combustion method [10].
Powder synthesized with the optimal glycine/nitrate ratio of 0.23 was
calcined in oxygen flow at 650°C for 24 hrs, ball milled for 12 hrs and dried
at 400°C for 12 hrs. This preparation method resulted in single phase
powder with crystallite size less than 10 nm and a particle size less than 50
nm calculated from surface area [10]. A small amount of residual carbonate
species was present in the powder after calcination [10]. The theoretical
density calculated the X-ray diffraction data was 5.96 g/cm3 [10]. The
powders were sieved (250 µm) before compacting and sintering.

Green bodies for conventional sintering were prepared by uniaxial pressing


at 64 MPa followed by cold isostatic pressing (CIP) at 200 MPa giving
green body density around 43% of theoretical density.

Densification
Dilatometry (Netzsch, DIL 402 C) on green body cylinders was performed
in ambient air with a heating rate of 120°C/h up to 1450°C. For
densification studies, three different sintering techniques were used; spark
plasma sintering (SPS), hot pressing (HP) and conventional sintering. The
SPS (Dr. Sinter 2050, Sumitomo Coal Mining Co. Ltd., Japan) was
performed in vacuum using cylindrical graphite dies also working as the
heating element. A uniaxial pressure varying from 50 MPa to 110 MPa was
applied at room temperature and released at the end of the holding time at
the sintering temperature. The sintering temperature varied from 1100°C to
1300°C and the holding time varied from 0 to 10 min. A heating rate of
100°C/min between 600°C and 1300°C and a cooling rate of >350°C/min
down to 1000°C was used. Previous studies have shown that this is a
reasonable cooling rate [11].

HP was performed under nitrogen flow in a clam furnace (Entech, VSTF


40/15) using cylindrical graphite dies with an inner diameter of 15 mm. No
compaction of the starting powders was made prior to the sintering. After
heating to 600°C with a rate of 600°C/h, a uniaxial pressure of 25 MPa was
applied before heating to sintering temperature (1150°C - 1300°C) at a rate
of 300°C/h. The pressure was released after desired sintering time (varying
from 0 to 6 hrs) at this temperature followed by a cooling to room
temperature at 600°C/h.

53
Conventional sintering was performed in air in a muffle furnace (Entech,
SF-4/17). Green bodies were sintered at temperatures from (i) 1150°C to
1300°C for 1 h with a heating rate of 300°C/h from 600°C (cooling rate
600°C/h) or (ii) 1500°C for 12 hrs with a heating/cooling rate of 200°C/h.
Final density was measured by the Archimedean method (ISO 5017) in
isopropanol.

Microstructure
The microstructure of the dense specimens was studied by field emission
scanning electron microscopy (FE-SEM) (Hitachi S-4300SE) on polished
and thermally etched surfaces. Thermal etching for 12 min at a temperature
50°C below sintering temperature was used. Average grain size was
estimated from the FE-SEM micrographs using the linear intercept method
[12]
. For the specimens with average grain size above 1 µm the
measurements were done over a minimum of 50 grains while a minimum of
100 grains were used for determining grain sizes below 1 µm.

Electrical conductivity
Electrical conductivity measurements were performed using the van der
Pauw technique [13] on sintered, polished (parallel planar) specimens. With
four platinum electrodes attached to the specimen surface close to the
circumference, the samples were slowly heated (1°C/min) to 1000°C in a
vertical tube furnace. At constant temperatures, from ~1000°C to ~500°C
(and back up to ~1000°C), the specific electrical conductivity was
calculated from perpendicular sets of currents (I1, I2) and voltages (U1, U2).

Mechanical properties
Hardness was calculated using the Vickers indentation method (Matsuzawa
DVK-1S) [14] on polished surfaces. For each sample, 10-12 indents were
made with an applied load of 2.9 N and measured by optical microscopy
(Reickert MeF3 A) using a digital camera (Sony DXC-950P colour video
camera). For fracture toughness (KIC) calculations [14], 10 indents with an
applied load of 49 N, were made for each sample. KIC was derived from
average crack length, measured by SEM (Hitachi S-3500N), as well as
experimental Vickers hardness values and a modulus of elasticity (E = 218
GPa) previously reported by Donzel and Roberts [15].

54
Results
Densification
The curves in Fig. 1, derived from linear shrinkage, show how the density
varied with time during SPS and conventional sintering. The sintering rate
was higher for SPS as close to full density (> 96%) was achieved within
minutes whereas hours were needed in order to obtain dense materials by
conventional sintering. Dense materials were achieved at significantly
lower temperature by SPS (1150 – 1200°C) compared to conventional
sintering (1450°C). Maximum sintering rate (derived from the slope of the
curves) was found between 1100°C and 1150°C for SPS and about 1250°C
for the conventional sintering. As seen from Fig. 1 the densification started
slightly earlier when the applied pressure during SPS was increased from 50
to 100 MPa however, the sintering rate was not affected.

100

90 1150oC 1450oC
1200oC

80
Density (%)

70

60

50
SPS 50 MPa
40 SPS 100 MPa
Conventional
30
0 2 4 6 8 200 300 400 500 600 700

Time (min)

Fig. 1 Change in density of YSZ during SPS and conventional


sintering as a function of sintering time. The marks (x)
indicate time for which the isothermal sintering temperatures
were obtained. Densities for SPS specimens are derived
directly from linear shrinkage during sintering and the final
density measured by the Archimedean method. Densities for
conventional sintering are calculated using both green and
final density with the linear shrinkage (dL/L0) as a scaling
factor.

55
The SPS and HP specimens were grey or black depending on sintering time.
The colour changed to white upon heating to temperatures above ~1000°C
in ambient air. Sintering conditions as well as final density of the sintered
specimens are listed in Table 1. Each specimen is given a numerical ID
indicating temperature-pressure-time. Final densities achieved by
isothermal SPS, HP and conventional sintering are displayed in Fig. 2 a).
SPS is superior as a relative density of 96.3% was obtained at 1150°C (SPS-
1150-70-5). Conventional sintering at equivalent temperatures gives
significantly lower density than SPS and HP.

Table 1 Sintering parameters and densities of YSZ specimens obtained


from different sintering techniques. CS, HP and SPS notes
conventional sintering, hot pressing and spark plasma sintering,
respectively. Number code indicates temperature-pressure-time.
Temperature Applied pressure Density
Specimen ID Time
(°°C) (MPa) (%)
CS-1500-0-12 1500 - 12 h 97.4
HP-1200-25-1 1200 25 1h 90.8
HP-1200-25-6 1200 25 6h 93.6
HP-1250-25-0.5 1250 25 0.5 h 93.9
HP-1250-25-1 1250 25 1h 96.9
HP-1250-25-3 1250 25 3h 97.0
HP-1300-25-0 1300 25 0h 92.3
HP-1300-25-0.5 1300 25 0.5 h 97.5
HP-1300-25-1 1300 25 1h 99.3
SPS-1150-70-5 1150 70 5 min 96.3
SPS-1200-70-5 1200 70 5 min 97.3
SPS-1250-70-5 1250 70 5 min 98.4
SPS-1300-70-5 1300 70 5 min 98.5
SPS-1200-50-0 1200 50 0 min 85.8
SPS-1200-50-1 1200 50 1 min 94.0
SPS-1200-50-5 1200 50 5 min 98.8
SPS-1200-50-10 1200 50 10 min 99.5
SPS-1100-110-8 1100 110 8 min 96.0
SPS-1150-100-5 1150 100 5 min 96.3
SPS-1150-100-3* 1150 100* 3 min 98.5
SPS-1150-100-5* 1150 100* 5 min 99.3

* Pressure applied at sintering temperature

56
100

90

Density (%)
80

a) 70

60 SPS, 70 MPa, 5 min


HP, 25 MPa, 1 h
Conventional, 1 h

50
1150 1200 1250 1300

Temperature (oC)

SPS time (min)


0 2 4 6 8 10
100

95
Density (%)

90

b)
85

80
o
SPS, 1200 C, 50 MPa
HP, 1250oC, 25 MPa
75 HP, 1200oC, 25 MPa

0 1 2 3 4 5 6

HP time (hrs)

Fig. 2 Densities of YSZ specimens sintered by different techniques at;


a) different temperatures and b) different time.

57
Fig. 2 b) shows how the density varies with sintering time at constant
temperature. SPS shows the highest degree of densification. The density of
hot pressed specimen HP-1250-25-1 was measured to 96.9%. In
comparison the density of SPS-1200-50-5 was 98.8 showing that both
temperature and in particular sintering time can be reduced significantly for
the SPS technique compared to HP. An attempt to lower the SPS
temperature to 1100°C by applying a pressure of 110 MPa (SPS-1100-110-
8) resulted in a final density of 96.0%. By applying a pressure at the
sintering temperature, a final density of 98.5% and 99.3% was obtained for
SPS-1150-100-3* and SPS-1150-100-5*, respectively.

Microstructure and grain growth


FE-SEM micrographs of a selection of sintered YSZ materials are shown in
Fig. 3. As seen in Fig. 3 a) through d), SPS gives homogenous
microstructures with narrow grain size distribution. The grain growth
during SPS, with increasing isothermal time at 1200°C (Fig. 3 a) and b)), is
moderate compared to the grain growth with increasing sintering
temperature (Fig. 3 c) and d)). The smallest grain size (0.21 µm) for spark
plasma sintered YSZ was obtained for the SPS-1100-110-8 specimen (Fig. 3
c)). In contrast severe grain growth was observed when the SPS sintering
temperature was increased from 1100°C to 1300°C (SPS-1300-70-5 shown
in Fig. 3 d)). Grain growth during HP with increasing isothermal time at
1250°C (Fig. 3 e) and f)) is moderate, comparable to SPS (Fig. 3 a) and b)).
The time scale however, is different and specimen HP-1250-25-3 (Fig. 3 f))
showed slightly smaller grains than SPS-1200-50-10 (Fig. 3 b)). The
conventionally sintered specimen (CS-1500-12) in Fig. 3 g) has larger
grains and with enclosed pores.

The average grain size development for YSZ materials, sintered at different
temperatures and times, is displayed in Fig. 4 a) and b), respectively. The
results indicate that HP is a good technique for densification of YSZ with
suppressed grain growth. For relatively dense YSZ (97%), an average grain
size down to 0.37 µm was obtained (HP-1250-25-1). The SPS technique
show a more severe grain growth with increasing temperature (Fig 4 a))
compared to HP, but as a function of time (Fig. 4 b)) the growth is moderate
for SPS as well. The average grain size of the conventionally sintered
specimen (CS-1500-0-12) was 12 µm.

58
a) SPS-1200-50-0 500 nm b) SPS-1200-50-10 500 nm

c) SPS-1100-110-8 500 nm d) SPS-1300-70-5 3µm

e) HP-1250-25-0 500 nm f) HP-1250-25-3 500 nm

g) CS-1500-0-12 3 µm

Fig. 3 FESEM micrographs of sintered YSZ specimens after oxidation


and thermal etching 50°C below sintering temperature.

59
3.5
SPS, 70 MPa, 5 min
3.0 HP, 25 MPa, 1 h

2.5
Grain size (µm)
2.0

a) 1.5

1.0

0.5

0.0
1150 1200 1250 1300

Temperature (oC)

SPS time (min)


0 2 4 6 8 10
0.55

0.50

0.45
Grain size (µm)

0.40
b)
0.35

0.30
o
SPS, 1200 C, 50 MPa
0.25 o
HP, 1250 C, 25 MPa

0.20
0.0 0.5 1.0 1.5 2.0 2.5 3.0

HP time (hrs)

Fig. 4 Average grain size as a function of sintering temperature (a)


and time (b) for YSZ sintered by hot pressing and spark plasma
sintering.

60
Electrical conductivity
The electrical conductivity of selected specimens is presented in Fig. 5 in
the form of plots of log(σT) versus 1000/T. Table 2 summarizes the
electrical conductivity at 900°C and the associated activation energy for
specimens sintered by the different techniques. For YSZ specimens sintered
from synthesized powder no clear trend of variation with grain size (varying
from 0.21 µm to 12 µm) was observed for neither conductivity (σ) nor
activation energy (EA). It can however be remarked that EA for the SPS
specimens are in the slightly higher (99 – 102 kJ/mol) compared to that of
the HP specimen (96 kJ/mol) and conventional sintered specimen (94
kJ/mol).

2.5

2.0
σ T) / (Scm K)]

1.5
-1

1.0
log[(σ

0.5 CS-1500-0-12
HP-1250-25-1
SPS-1300-70-5
0.0 SPS-1150-70-5
SPS-1100-110-8

0.8 0.9 1.0 1.1 1.2 1.3


-1
1000/T (K )

Fig. 5 Logarithmic plot of the electrical conductivity for YSZ materials


measured by the van der Pauw technique.

61
Table 2 Properties of YSZ specimens obtained from different sintering
techniques.
Average Conductivity Activation
Density
Specimen ID grain size at 900°C energy
(%)
(µm) (mS/cm) (kJ/mol)
CS-1500-0-12 97.5 12 70 94
SPS-1300-70-5 98.5 3.3 83 101
HP-1250-25-1 96.9 0.37 82 96
SPS-1150-70-5 96.3 0.24 75 99
SPS-1100-110-8 96.0 0.21 82 102

Mechanical properties
The SEM micrograph in Fig. 6 a) show an intergranular crack obtained from
a 49 N indent on the SPS-1300-70-5 specimen. Intergranular cracks were
observed for all samples except the larger grained conventionally sintered
(1500°C) YSZ specimen which showed intragranular crack behaviour (Fig.
6 b)). Vickers hardness (HV) and fracture toughness (KIC) of YSZ materials
with varying grain size are shown in Fig. 7 a) and b), respectively. As seen
there is no significant change in HV for specimen with grain size from 0.2 to
12 µm, while the fracture toughness is slightly increasing with decreasing
grain size.

a)a) SPS-1300-70-5 4 µm b) CS-1500-0-12


b) 4µm

Fig. 6 a) SEM micrograph of a intergranular crack (from 49 N indent)


in YSZ sintered by SPS at 1300°C. b) SEM micrographs of the
intragranular crack (from 49 N indent) in YSZ sintered
conventionally at 1500°C.

62
15
a)
14

HV (GPa)
13

12

1.7 b)
KIC (MPa m1/2)

1.5

1.3

1.1

0.1 1 10

Grain size (µm)

Fig. 7 Vickers hardness (a) and fracture toughness (b) for YSZ
materials with different grain size. The uncertainty bars indicate
the standard deviations.

Discussion
Densification
SPS is shown to be a highly efficient technique for densification of YSZ at
temperatures 100-200°C lower than that needed by HP. For HP with an
applied pressure of 25 MPa, a temperature of 1300°C is needed to obtain
fully dense (>99% of theoretical value) YSZ materials. In comparison,
1150°C for 5 min (100 MPa) was sufficient when using the SPS technique.
The higher pressure used in SPS can partly explain the lower sintering
temperature. The main reason for both the short sintering time and reduced
temperature, however, is the efficient heat transfer as well as the self-
heating from spark discharge between the particles. Under these conditions

63
residual carbonates in the powder will be efficiently removed as long as the
system is not closed. By not applying the load until the isothermal SPS
temperature is reached, CO2 is allowed to escape the system before the main
densification starts. Residual carbonates in the powder have been shown to
inhibit the sintering [10]. A more efficient removal of carbonates during SPS
compared to HP and conventional sintering can therefore also explain the
higher sintering rate and lower sintering temperature. The densification rate
by SPS has a maximum 50 – 100°C lower than reported by Anselmi-
Tamburini et al. for SPS of commercial powder (TZ-8Y) [16].

Microstructure and grain growth


HP and SPS have been proved to produce dense YSZ materials with small
grains compared to conventional sintering where severe grain growth is
inevitable in the final stage of densification. The kinetics of grain growth
during HP and SPS was investigated by calculating the growth order, n, and
(for HP) the activation energy, EA, for grain growth. The growth order was
calculated from the slope of the linear regression plots shown in Fig. 8 a)
assuming the growth obeys a power-law relationship [17]. According to the
LSW theory [18, 19] the exponent, n, should be equal to 3 when the grain
growth is controlled by volume diffusion. Furthermore, n = 2 can be
assigned to growth controlled by energy difference across curved interfaces
while n = 4 is associated with coarsening controlled by grain boundary
diffusion. Hence, grain growth for both SPS and HP, with calculated
exponents n = 3.3 and n = 2.6 respectively, seem to be dominated by volume
diffusion. Even though the exponent n for SPS is higher, the relative grain
growth D - D0 is lower for HP. It is shown in Fig. 4 a) how the grain growth
for SPS with increasing sintering temperature is far more severe than for
HP. More effective decomposition of carbonate species during SPS, giving
grain boundaries with higher purity and therefore higher mobility, can
explain the higher grain growth rate compared to that of HP.

It should be noted that plotting ln(D) instead of ln(D - D0) as a function of


ln(t) led to unreasonably high growth order values. Simplified plots like this
are often reported [20-22] and will for severe grain growth (several orders of
magnitude) give satisfactory estimates of n. From our observations we can
state that this fails for smaller grain size changes with sintering time. The
grain growth observed for SPS follow the same trend as recently reported
work on YSZ [16]. The activation energy for grain growth during HP,
calculated from the slope in Fig. 8 b), was 226 kJ/mol. This is comparable
to reported value of 289 kJ/mol for YSZ (8% Y2O3) [23].

64
-1.0

-1.5 SPS
ln[(D - D0) / µm] n = 3.3

-2.0
a)
HP
-2.5
n = 2.6

-3.0
-4 -3 -2 -1 0 1

ln[t / hrs]

-1.2

-1.6
ln[(D-D0)2.6 / t]

-2.0
EA = 226 kJ/mol
-2.4
b)
-2.8

-3.2

-3.6
0.76 0.78 0.80 0.82 0.84

1000/T (K-1)

Fig. 8 a) Grain growth exponents, n, for HP at 1250°C and SPS at


1200°C, calculated from the slopes of the linear regression plots
of ln(D – D0) versus ln(t). b) Activation energy for grain growth
during HP, calculated from the slope of the linear regression
plot of ln((D – D0)/t) versus 1000/T.

65
Electrical conductivity
It has been suggested that the difference in conductivity for the
polycrystalline specimens from different starting powders could be due to a
different level of impurities in the powders. The powder contained a small
amount of residual carbonates. However, it is unlikely that carbonates are
still present after sintering, particularly by using the SPS technique. At
900°C, σ varies from 75 to 83 mS/cm and in comparison a total ionic
conductivity in the range of 40 – 60 mS/cm at 900°C is reported for YSZ
(8% Y2O3) by Chen [5]. For YSZ (8.7% Y2O3) bulk materials with grain
size varying from 6.5 down to 1.3 µm, a small variation in electrical
conductivity (3 – 6 mS/cm at 600°C) has been reported by Tuller [6]. This is
comparable to the measured electrical conductivity of 3 – 5 mS/cm at
600°C, however in this case no systematic variation with grain size is
observed. Calculated activation energies varies from 94 to 102 kJ/mol
which is somewhat lower than what is reported by Tuller (111 kJ/mol) [6]
and Anselmi-Tamburini (109 kJ/mol) [24]. The colouring of SPS and HP
specimens after sintering is an indication of reduced material as previously
reported [24]. Increased electronic conductivity would be expected in YSZ
specimen containing reduced species acting as charge carriers. Such an
effect is not observed as the specimens are oxidized by heating to ~1000°C
prior to the electrical conductivity measurement. A possible explanation for
the higher EA values for SPS and HP specimens could be carbon impurities
from the graphite pressing dies used during sintering. It should be pointed
out that no significant weight gain or loss due to oxidation of reduced
species or removal of carbon was observed upon heating.

Mechanical properties
Measured Vickers hardness are in agreement with reported values of 13-15
GPa [15,17,25] and no significant change is observed for different grain sizes.
The fracture toughness is somewhat decreasing with increasing grain size.
As intergranular fracture is observed, an increase in KIC with increasing
grain size should be expected [26]. The observed decrease might be due to
the segregation of pores along the grain boundaries, as seen for the SPS
specimen in Fig. 6 a). Intergranular fracture in conventionally sintered YSZ
(Fig. 6 b)) can be explained by the trapped pores inside individual grains.
The intergranular fracture results in lower KIC value compared to the SPS
sample sintered at 1100°C. The obtained KIC values are in agreement with
literature values [15, 23].

66
Conclusions
In a comparative study of the sintering techniques SPS, HP and
conventional sintering, dense YSZ materials (> 96% of theoretical), with
average grain size of 210 nm, 370 nm and 12 µm, respectively, were
obtained by all methods. SPS enables preparation of dense materials with
limited grain growth provided appropriate sintering conditions are applied,
if not very fast grain growth might occur. No significant differences in
electrical conductivity were observed for YSZ materials with grain size
varying from 210 nm to 12 µm, however the activation energy for spark
plasma sintered material (99-102 kJ/mol) was slightly higher than for HP
(96 kJ/mol) and conventional sintering (94 kJ/mol). No clear variation in
Vickers hardness was found with varying grain size. The fracture toughness
showed a small increase with decreasing grain size.

References
[1] I. Kosacki, T. Suzuki, V. Petrovsky, H.U. Anderson, Electrical
conductivity of nanocrystalline ceria and zirconia thin films, Solid
State Ionics, 136-137, 1225-1233 (2000)
[2] I. Kosacki, H.U. Anderson, Microstructure – Property relations in
nanocrystalline oxide thin films, Ionics, 6, 294-311 (2000)
[3] M.C. Martin, M.L. Mecartney, Grain boundary ionic conductivity of
yttrium stabilized zirconia as a function of silica content and grain
size, Solid State Ionics, 161, 67-79 (2003)
[4] X. Guo, Z. Zhang, Grain size dependent grain boundary defect
structure: case of doped zirconia , Acta Materialia, 51, 2539-2547
(2003)
[5] X.J. Chen, K.A. Khor, S.H. Chan, L.G. Yu, Influence of
microstructure on the ionic conductivity of yttria-stabilized zirconia
electrolyte, Mater. Sci. Eng., A335, 246-252 (2002)
[6] H.L. Tuller, Ionic conduction in nanocrystalline materials, Solid
State Ionics, 131, 143-157 (2000)
[7] P. Mondal, A. Klein, W. Jaegermann, H. Hahn, Enhanced specific
grain boundary conductivity in nanocrystalline Y2O3-stabilized
zirconia, Solid State Ionics, 118, 331-339 (1999)
[8] C. Petot, M. Filal, A.D. Rizea, K.H. Westmacott, J.Y. Laval, C.
Lacour, R Ollitrault, Microstructure and ionic conductivity of freeze-
dried yttria-doped zirconia, J. Eur. Ceram. Soc., 18, 1419-28 (1998)
[9] Z. Shen, M. Johnsson, Z. Zhao, M. Nygren, Spark plasma sintering
of alumina, J. Am. Ceram. Soc., 85 (8), 1921-27 (2002)

67
[10] I. Kaus, P. Dahl, J. Mastin, T. Grande, M.-A. Einarsrud, Synthesis
and characterization of nanocrystalline YSZ powder by smoldering
combustion synthesis, Journal of Nanomaterials, 2006, 1-7 (2006)
[11] Z. Shen, Z. Zhao, H. Peng, M. Nygren, Formation of tough
interlocking microstructures in silicon nitride ceramics by dynamic
ripening, Nature, 417, 266-269 (2002)
[12] M.I. Mendelson, Average grain size in polycrystalline ceramics, J.
Am. Ceram. Soc., 52 (8), 443-446 (1969)
[13] L.J. Van der Pauw, A method for measuring specific resistivity and
hall effect of discs of arbitrary shape, Philips Res. Repts., 13, 1-9
(1958)
[14] G.R. Anstis, P. Chantikul, B.R. Lawn, D.B. Marshall, A critical-
evaluation of indentation techniques for measuring fracture
toughness: 1. Direct crack measurements, J. Am. Ceram. Soc., 64,
533-538 (1981)
[15] L. Donzel, S.G. Roberts, Microstructure and mechanical properties
of cubic zirconia (8YSZ)/SiC nanocomposites, J. Eur. Ceram. Soc.,
20, 2457-2462 (2000)
[16] U. Anselmi-Tamburini, J.E. Garay, Z.A. Munir, Spark plasma
sintering and characterization of bulk nanostructured fully stabilized
zirconia: Part I. Densification studies, J. Mater. Res., 19 (11), 3255-
3262 (2004)
[17] H. Jiang, Y. Lu, W. Huang, X Li, M. Li, Microstructural evolution
and mechanical properties of the semisolid Al-4Cu-Mg alloy, Mater.
Char., 51, 1-10 (2003)
[18] I.M. Lifshitz, V.V. Slyozov, The kinetics of precipitations from
supersaturated solid solutions, J. Phys. Chem. Solids, 19 (1/2), 35-
40 (1961)
[19] C. Wagner, Theorie der alterung von niederschlagen durch umlosen
(Ostwald-Reifung), Zeitschrift für Elektrochemie, 65, 581-591
(1961)
[20] S. Tekeli, Influence of alumina addition on grain growth and room
temperature mechanical properties of 8YSCZ/Al2O3 composites,
Comp. Sci. Tech., 65, 967-972 (2005)
[21] L. Yang, J.S. Wu, L.T. Zhang, Microstructure evolvements of a rare-
earth filled skutterudite compound during annealing and spark
plasma sintering, Materials and Design, 25, 97-102 (2004)
[22] T.S. Zhang, J. Ma, Y.J. Leng, Z.M. He, Sintering, microstructure
and grain growth of Fe-doped Ce0.9Gd0.1O2−δ ceramics derived from
oxalate coprecipitation, J. Crystal Growth, 274, 603-611 (2005)
[23] S. Tekili, Fracture toughness (K-IC), hardness, sintering and grain
growth behaviour of 8YSCZ/Al2O3 composites produced by colloidal
processing, J. All. Comp., 391, 217-224 (2005)

68
[24] U. Anselmi-Tamburini, Spark plasma sintering and characterization
of bulk nanostructured fully stabilized zirconia: Part II.
Characterization studies, J.E. Garay, Z.A. Munir, J. Mater. Res., 19
(11), 3263-69 (2004)
[25] G.A. Gogotsi, S.N. Dub, E.E. Lomonova, B.I. Ozersky, Vickers and
Knoop indentation behaviour of cubic and partially-stabilized
zirconia crystals, J. Eur. Ceram. Soc., 15, 405-413 (1995)
[26] P.F. Becher, Microstructural design of toughened ceramics, J. Am.
Ceram. Soc., 74 (2), 255-269 (1991)

69
70
PAPER III
72
Synthesis, densification and electrical properties
of Strontium Cerate ceramics

Paul Inge Dahl, Hilde Lea Lein, Tor Grande, Mari-Ann Einarsrud †

Department of Materials Science and Engineering


Norwegian University of Science and Technology
NO-7491 Trondheim, Norway

Truls Norby, Reidar Haugsrud

Centre for Materials Science and Nanotechnology


Department of Chemistry, University of Oslo
NO-0349 Oslo, Norway

Abstract
Powders of pure and 5% ytterbium substituted strontium cerate (SrCeO3 /
SrCe0.95Yb0.05O3-δ) were prepared by spray pyrolysis of nitrate salt solutions.
The powders were single phase after calcination in nitrogen atmosphere at
1100°C (SrCeO3) and 1200°C (SrCe0.95Yb0.05O3-δ). Hollow spherical
particles formed during spray pyrolysis were effectively broken down by
ball milling giving particle size down to 0.06 µm. Dense SrCeO3 and
SrCe0.95Yb0.05O3-δ materials were obtained by sintering at 1350 - 1400°C in
air, but heat treatment at 850 and 1000°C, respectively, was necessary prior
to sintering. The dense materials had homogenous microstructures with
grain size in the range 6 - 10 µm for SrCeO3 and 1 - 2 µm for
SrCe0.95Yb0.05O3-δ, after sintering up to 1400°C. The electrical conductivity
of SrCe0.95Yb0.05O3-δ was in good agreement with reported data, showing
mixed ionic-electronic conduction. The ionic contribution was dominated
by protons in below 1000°C and the proton conductivity reached a
maximum of ~0.005 S/cm above 900°C. In oxidizing atmosphere the p-type
electronic conduction was dominating above ~700°C, while the contribution
from n-type electronic conduction only was significant above ~1000°C in
reducing atmosphere.

Corresponding author. E-mail: mari-ann.einarsrud@material.ntnu.no

73
Introduction
Perovskite-type oxide materials (ABO3 structures), such as orthorhombic
strontium cerate, SrCeO3, have been widely studied due to the electrical
properties of these materials. When properly substituted with rare earth
oxides on the cerium site, SrCeO3, exhibit high proton conductivity as first
demonstrated by Iwahara et al. [1]. This property makes SrCeO3-based
ceramics suitable for electrolyte materials in solid oxide fuel cells (SOFCs),
hydrogen pumps and sensors [2-4]. In particular SrCeO3 substituted with 5%
Yb (SrCe2.95Yb0.05O3-δ) has been classified as one of the perovskite-type
oxides with the highest proton conductivity (~0.004 S/cm at 900°C) [5].
However, SrCeO3 based materials have often been rejected for use as proton
conductors in commercial devices due to poor mechanical properties and
chemical stability [2].

High stability of strontium carbonate (SrCO3) makes preparation of single


phase SrCeO3 powders and bulk materials challenging. This is described in
the literature [6,7], by the reaction between SrCeO3 and CO2 according to Eq.
1, forming SrCO3 which may decompose to strontium oxide (SrO) as
described by Eq. 2.

SrCeO3(s) + CO2(g) = SrCO3(s) + CeO2(s) (1)

SrCO3(s) = SrO(s) + CO2(g) (2)

The onset temperature for synthesis of SrCeO3 from a mixture of SrCO3 and
CeO2 (according to Eq. 1) has been reported to 800°C and between 1168
and 1190°C, for reaction in N2 and CO2 atmosphere, respectively. The
corresponding onset temperature for decomposition of SrCO3 (according to
Eq. 2) in N2 and CO2 atmosphere have been reported to 845°C and between
1220 and 1275°C, respectively [6,8]. Hence, at lower temperatures (<800°C),
carbonate formation may be critical, with respect to phase purity and
mechanical properties of SrCeO3 materials. Reactions with water,
equivalent to Eq. 1 and 2, should also be considered when operating in
atmospheres with high p(H2O) [7].

Preparation of SrCeO3 powders have been reported by methods such as


complexation with EDTA [9] or citric acid [10] and combustion methods [11],
but the dominating synthesis route for these materials is by the solid state
ceramic method [1,6,8,12-20]. The main objectives of work on SrCeO3
ceramics have been to investigate the electrical proton conductivity. Hence,
little focus has been put into densification and microstructure of sintered
specimens. Generally sintering temperatures in the range 1500 - 1650°C (in
air) are needed for preparing dense (>95%) SrCeO3 and SrCe2.95Yb0.05O3-δ

74
materials by the solid state ceramic method [13,15,16,20]. Higher densities
(>97%) have been obtained from commercial powders by sintering at
1450°C in air for 2 hrs [21], and from powder prepared by a complexation
route by sintering at 1300°C for 12 hrs in nitrogen atmosphere [9].

The limited literature data on synthesis and densification of SrCeO3 based


materials leads to the motivation for the present work, where the overall aim
have been to give descriptive routes for synthesis of SrCeO3 powders and
preparation of dense materials with homogenous microstructures.
Preparation routes for pure and 5% Yb-substituted SrCeO3 powders will be
presented and the influence of carbonates (and other secondary phases) will
be discussed. The sinterability of the materials has been studied with focus
on how the microstructure in these materials evolves during densification.
Furthermore, characterization of the electrical conductivity will be presented
and thoroughly discussed, and thermodynamic data extracted from these.

Experimental
Powder synthesis and sample preparation
Powders of SrCeO3 and SrCe0.95Yb0.05O3-δ (noted SC and SC5Yb in the
following) were synthesized by spray pyrolysis of nitrate salt solutions.
Precursor solutions of cerium nitrate (Ce(NO3)2·xH2O, Alfa Aesar, 99.5%)
and ytterbium nitrate (Yb(NO3)3·5H2O, Aldrich, 99.9%) dissolved in
distilled water were thermogravimetrically standardized and mixed in
stoichiometric proportions with dried (250°C, 24 hrs) strontium nitrate
(Sr(NO3)2, Merck, >99%), giving a total cation concentration of 0.9 M. The
mixed solutions were atomized into a furnace at a rate of 1 L/h. The
temperature by the nozzle was 840-845°C and the output temperature was
540-550°C. The as-synthesized SC and SC5Yb powders were calcined at
900°C in ambient air for 48 hrs. Additional calcination at 1000 - 1200°C
for 6 hrs in nitrogen flow ( pO2 < 10-4 atm, pCO2 < 2·10-7 atm) were performed
in order to remove carbonates and obtain phase pure powders. The calcined
powders were ball milled with yttria stabilized zirconia (YSZ) balls in iso-
propanol for 6 - 24 hrs. The milled powders were dried at 400°C for 12 hrs
in ambient air and sieved (150 µm).

Green bodies were made by uniaxial pressing at 64 MPa. Generally


powders calcined at higher temperatures (>1000°C) needed addition of 2
wt% binder (ethyl cellulose, Sigma) in order to increase the green strength
and green density (~55%). No binder was needed for compaction of
powders calcined at 900°C giving green bodies with green density of 50-

75
53%. Green bodies were sintered in ambient air at temperatures varying
from 1250°C to 1450°C in a muffle furnace (Entech, SF-4/17). In order to
remove carbonate species from the specimens prior to sintering, heat
treatment at 850°C for 24 hrs and 1000°C for 1 h was used for sintering of
SC and SC5Yb, respectively.

Characterization of sintered materials


The prepared powders were analyzed and crystal structure confirmed by X-
ray diffraction (XRD) (Cu Kα, Phillips PW1730/10). Powders from crushed
samples sintered at 1400°C and added 30 weight% Si standard were used to
obtain XRD data from which the lattice parameters and theoretical densities
were calculated by the Rietveld method. Infrared spectroscopy (IR) was
performed on potassium bromide (Merck, KBr for IR spectroscopy
Uvasol®) discs containing 1 wt% powder using a Bruker IFS 66v
spectrophotometer. Thermogravimetric analysis (TGA) (Netzsch STA 449
C Jupiter) on powders was carried out in air with a heating rate of 3°C/min
to 1450°C. Nitrogen adsorption on de-gassed powders (24 hrs, 250°C) was
measured (Micromeritics ASAP 2000) using the 5 point BET equation to
obtain specific surface area, from which the particle size was calculated
assuming spherical particles.

Dilatometry (Netzch, DIL 402 C) on green body cylinders was performed in


air with a heating rate of 3°C/min up to 1450°C. Density of sintered
specimens was measured by the Archimedean method (ISO 5017) in iso-
propanol. The prepared powders as well as sintered specimens were studied
by scanning electron microscopy (SEM) (Hitachi S-3500N).

Conductivity measurements
A circular Pt electrode (~10 mm diameter) was attached to each side of
sintered SC5Yb specimens, before mounting on top of the support tube in a
ProboStat measurement cell for electrochemical characterization (two-point
measurements) [22]. When the purpose was to measure the open-circuit
voltage (OCV) from gradients in the pressures of oxygen, hydrogen and
water vapor, a gold gasket was placed between the alumina support tube and
the specimen. Consequently, the specimen served as a membrane between
the inner and the outer gas compartment in the cell. Electrode leads (current
supplies and voltage probes) from the cell base were contacted to the
electrodes on the specimen. Relatively strong spring loads with alumina

76
parts in the hot zone held the assembly together; to maintain contact
between the Pt leads and the electrodes and to facilitate the gold sealing.

The total conductivity was measured in the temperature range 300 to


1050°C by means of impedance spectroscopy and at a constant frequency of
10 kHz (Solartron 1260 FRA, oscillation voltages between 0.1 and 1 V) as a
function of the oxygen partial pressure; in wet (0.025 atm H2O) O2-Ar
mixtures for oxidizing and wet (0.025 atm H2O) H2-Ar mixtures for
reducing conditions, as well as a function of the water vapor partial
pressure. The partial conductivities, measured under wet conditions, were
calculated based on transport numbers from EMF measurements (Solartron
7150+ high impedance voltmeter) as a function of temperature. Details
regarding the EMF-method and the set-up of the gas mixer are described in
the literature [23,24].

Impedance spectroscopy in the frequency range 1 MHz to 0.1 Hz was


applied to determine the different contributions of bulk (b), grain boundaries
(gb) and electrodes (e) to the overall impedance. Different circuits were
assigned to fit the impedance data depending on the conditions (T, p(O2)
and p(H2O)). Up to 650°C it was possible to separate the impedance of bulk
and grain boundaries by fitting the impedance data to circuits of parallel
resistors (R) and constant phase capacitive elements (Q) in series:
(RbQb)(RgbQgb)(ReQe) at temperatures below 450°C, and Rb(RgbQgb)(ReQe)
from 450 to 650°C. Above 650°C the bulk and grain boundary impedance
could not be separated and the volume resistance, Rvol = Rb + Rgb, was
determined. Under conditions where electronic conduction plays a
significant role, it has been assumed that Relectronic, representing the
electronic conductance, connects effectively directly between electrodes,
yielding circuits which in a simplified version may look like
Relectronic[Rvol(ReQe)]. At temperatures >500°C the impedance spectra were
corrected for an inductive element resulting from different types of parasitic
contribution. The (ReQe)-elements represent complex processes at the
electrodes and can be further separated into several different circuit
elements.

77
Results
Characterization of powders
The morphology of prepared powders is presented by SEM micrographs in
Fig. 1. During spray pyrolysis characteristic agglomerates in the form of
hollow spheres were formed as seen for the SC powder calcined at 900°C.
The spherical shape was maintained upon heating to 1100°C, however,
some coarsening of the powder was observed. The agglomerates in the SC
and SC5Yb powders calcined at 900°C were effectively broken down to
smaller particles by ball milling in iso-propanol. Due to coarsening of the
powders during heat treatment at 1100°C, the particle size after milling
seemed somewhat larger in the powders calcined at 1100°C. Ball milling in
water resulted in decomposition of the calcined powders, confirming the
chemical instability of SrCeO3.

Surface area and calculated particle size for a selection of powders is


presented in Table 1. Ball milling of SrCeO3 powder calcined at 900°C
resulted in an increase in surface area, from 1.7 to 5.4 m2/g. Calcination at
1100°C gave a slight reduction in surface area to 3.9 m2/g (after ball
milling). The surface area of ball milled SC5Yb powder calcined at 900°C
was 17.8 m2/g, corresponding to particles 3 - 8 times smaller compared to
that found for the SC powders.

Table 1 Surface area (SA) of milled SrCeO3 and SrCe0.95Yb0.05O3-δ


powders. Spherical particles were assumed for calculation of
the particle size (D).
Tcalcination tcalcination
Compound SA (m2/g) D (µm)
(°C) (hrs)
SrCeO3* 900 48 1.7 0.62
SrCeO3 900 48 5.4 0.19
SrCeO3 1100 6 3.9 0.27
SrCe0.95Yb0.05O3-δ 900 48 17.8 0.06
SrCe0.95Yb0.05O3-δ 1100 6 9.2 0.11
* Not milled

78
Fig. 1 SEM micrographs of not milled SrCeO3 and SrCe0.95Yb0.05O3-δ
powders calcined at 900 and 1100°C before and after milling in
iso-propanol.

79
The orthorhombic phase was confirmed for SC and SC5Yb by XRD. The
calculated lattice parameters and theoretical densities for both compositions
are listed in Table 2. X-ray diffractograms of SC and SC5Yb powders are
shown in Fig. 2 a) and b), respectively. Secondary phases (Sr2CeO4 and
CeO2) were present in the SC powder after calcination at 900°C. After
calcination at 1000°C in nitrogen flow these phases were still detected by
XRD, however, calcination under these conditions at 1100°C resulted in
phase pure SC powder. Larger amounts of secondary phases were detected
in the SC5Yb powder calcined at 900°C in air. In addition to the secondary
phases found in the SC powders, SrCO3 was detected in SC5Yb powder
calcined at 900°C. SrCO3 was not detected in the SC5Yb powder calcined
at 1000°C in nitrogen flow, but both Sr2CeO4 and CeO2 were present.
Traces of Sr2CeO4 were also detected after calcination at 1100°C but the
powder was apparently phase pure after calcination at 1200°C. XRD on the
surface of sintered specimen exposed to ambient atmosphere for 2 weeks
gave no indications of formation of SrCO3.

Table 2 Lattice parameters and theoretical density (TD) for SrCeO3 and
SrCe0.95Yb0.05O3-δ powders.

Compound a (Å) b (Å) c (Å) TD (g/cm3)


SrCeO3 6.16 8.60 6.01 5.76
SrCe0.95Yb0.05O3-δ 12.28 8.58 6.00 5.82

80
a)  SrCeO3
 Sr2CeO4
 CeO2

Relative intensity 1100°C

1000°C 


 

  
  

900°C

20 25 30 35 40

2 θ (°)

b)  SrCe0.95Yb0.05O2.975
 SrCO3
  Sr2CeO4

 CeO2
Relative intensity

1200°C

1100°C

1000°C 


 
 
 
   

900°C

20 25 30 35 40

2 θ (°)

Fig. 2 X-ray diffractograms of a) SrCeO3 powders and b)


SrCe0.95Yb0.05O3-δ powders calcined in ambient air at 900°C and
in nitrogen flow from 1000 to 1200°C.

81
Fig. 3 shows TG data for SC and SC5Yb powders. Significant weight loss
was observed for both SC and SC5Yb powder calcined at 900°C and ball
milled in iso-propanol. The observed weight loss below 400°C (mainly
seen for SC powder calcined at 900°C) is most likely due to adsorption of
CO2 and/or water from the atmosphere as all powders were dried at 400°C
after milling. As seen from Fig. 3, the main weight loss of milled powders
was found between 700 and 800°C for SC and around 900°C for SC5Yb.
This weight loss is associated with decomposition of SrCO3. The higher
loss (4.0%) for SC5Yb compared to SC (1.6%) is in agreement with a larger
amount of SrCO3 in the SC5Yb powder, as confirmed by XRD. A total
weight loss of ~0.3% was observed for SC powder calcined at 1100°C. This
indicates little or no presence of strontium carbonate after calcination at
1100°C.

100
SC - 1100°C, not milled

99
Weight change (%)

SC - 900°C, milled

98

97

SC5Yb - 900°C, milled


96

100 300 500 700 900 1100 1300 1500

Temperature (°C)

Fig. 3 TG curves obtained from milled SrCeO3 (SC) and


SrCe0.95Yb0.05O3-δ (SC5Yb) powders calcined at 900°C as well as
SC powder calcined at 1100°C before milling.

82
IR spectra of SC / SC5Yb powders calcined at different temperatures are
presented in Fig. 4. A reference spectrum of SrCO3 is also included. The
first broad band in the range of 500 to 800 cm-1, observed for all prepared
powders, is due to the stretching of the metal-oxygen bonds [24]. The
characteristic frequencies for SrCO3 are also found in the spectra of SC and
SC5Yb powders calcined at 900°C, respectively. Higher intensity of the
carbonate bands, indicate larger amount of SrCO3 in the SC5Yb powder
compared to SC. SC powders calcined at 1000°C and 1100°C were
carbonate free according to the IR spectra. SC5Yb powders calcined at
higher temperatures (1000 - 1100°C) show similar behaviour as observed
for SC. IR spectra of ball milled powders indicated that SrCO3 were
reintroduced by CO2 adsorption from the air or from reaction with residual
milling liquid in the powder.
Relative transmittance

1100°C
1000°C

900°C
900°C

SrCO3
SrCeO3
SrCe0.95Yb0.05O2..975

400 1000 1600 2200 2800 3400 4000


-1
Wave number (cm )

Fig. 4 IR spectra of SrCO3 powder, SrCe0.95Yb0.05O3-δ and SrCeO3


powders calcined 900°C in air as well as SrCeO3 powders
calcined at 1000°C and 1100°C in nitrogen atmosphere.

83
Densification
The linear shrinkage versus temperature presented in Fig. 5 a) show an onset
for densification at ~800°C for SC and ~1000°C for SC5Yb, both calcined
at 900°C. The presintering feature above 900°C for SC5Yb is most likely
due to decomposition of residual SrCO3, resulting in evolving CO2 gas that
could cause a small expansion in the material. Calcination at 1100°C lead
to a small shift in the onset temperature for densification. Though the onset
temperature for densification is lower for SC, the sintering rate for this
compound is lower than for SC5Yb, as displayed in Fig. 5 b). The highest
sintering rate (Fig. 5 b)) for SC calcined at 900°C was found at ~1150°C
and as high as ~1375°C for SC calcined at 1100°C. In comparison the
highest sintering rate for SC5Yb calcined at 900°C and 1100°C was
~1330°C and ~1370°C, respectively.

-5
dL/L0 (%)

-10 a)
SC - 900°C
-15 SC - 1100°C
SC5Yb - 900°C
-20 SC5Yb - 1100°C

0.0
d(dL/L0)/dT (%/°C)

-0.1
b)

-0.2

-0.3
200 400 600 800 1000 1200 1400
Temperature (°C)

Fig. 5 a) Linear shrinkage and b) sintering rate from associated


derivative curves for SrCeO3 (SC) and SrCe0.95Yb0.05O3-δ
(SC5Yb) after calcination at 900°C and 1100°C.

84
Fig. 6 displays the variation in density and estimated grain size as a function
of temperature during isothermal sintering of SC and SC5Yb. The
presintering heat treatment at 850°C for 24 hrs for SC and 1000°C for 1 h
for SC5Yb resulted in 1 - 4% reduction of closed porosity. As seen in Fig. 6
a), the density of SC and SC5Yb increases when the temperature is
increased from 1250 to 1350°C, however further increase in temperature
gave little change in density. After sintering at 1350°C the density was 96.6
and 97.6% of theoretical density for SC and SC5Yb, respectively.

100

95
Density (%)

90 a)

SC
85 SC5Yb

15
Grain size (µm)

10
b)

0
1250 1300 1350 1400 1450

Temperature (°C)

Fig. 6 a) density and b) estimated grain size of SrCeO3 (SC) and


SrCe0.95Yb0.05O3-δ (SC5Yb) specimens sintered at indicated
temperatures for 2 hrs in ambient air.

85
Microstructure and grain growth
Estimated grain size in sintered SC and SC5Yb specimens are presented in
Fig. 6 b). The grain size in SC increased from ~6 to ~14 µm when the
sintering temperature was increased from 1250 to 1450°C. For SC5Yb the
grain growth was less severe giving average grain size in the range of 1 to 2
µm after sintering at 1250 - 1400°C. Sintering at 1450°C, however, resulted
in grain size of ~8 µm for this composition as well. The homogenous
microstructures of sintered SC and SC5Yb specimens are displayed by SEM
micrographs in Fig. 7. The microstructure evolution is shown in the left and
right row of micrographs for SC and SC5Yb, respectively, with increasing
sintering temperature from 1250 to 1450°C downwards in the figure. While
the grain size of the SC specimens were rather large, even after sintering at
1250°C, SC5Yb specimen sintered at this temperature exhibit grains of ~1
µm, however this specimen appears not fully dense. The inevitable amount
of closed porosity observed in SC specimens, due to higher mobility of
grain boundaries compared to pores, was not observed in the SC5Yb
specimens sintered at temperatures up to 1400°C. Sintering at 1450°C
produced larger grains resulting in some enclosed pores for SC5Yb as well.

The severe grain growth in SC resulted in microcracking reducing the


mechanical properties of this material. This may be due to the anisotropic
properties of the orthorhombic SrCeO3. Microcracking was not observed in
the SC5Yb specimens, which may indicate critical grain size for
microcracking in in the range 2 - 6 µm, explaining why this was observed
for the larger grained SC materials.

86
Fig. 7 SEM micrographs of polished and etched (1200°C, 12 min)
surfaces of SrCeO3 and SrCe0.95Yb0.05O3-δ specimens sintered for
2 hrs 1250°C - 1450°C.

87
Electrical properties
The partial conductivities as a function of the inverse temperature for
SC5Yb in wet oxygen (0.025 atm H2O + 0.975 atm O2) and wet hydrogen
(0.025 atm H2O + 0.975 atm H2) are presented in Fig. 8 a) and b),
respectively. These figures include the total AC (10 kHz) conductivity
measured in continuous temperature ramps (12°C/h), along with partial
conductivities deconvoluted from impedance spectroscopy at selected
temperatures in wet hydrogen (Fig. 8 b)). The proton and oxygen ion
conductivities are similar under oxidizing and reducing conditions, whereas
the electronic contribution under oxidizing conditions (p-type) is higher than
under reducing conditions (n-type). One should note that the OCV of
concentration cells has only been measured in the temperature region where
the proton transport number is less than 1 and below the melting point of
gold (700 to 1050°C). The total conductivities measured at temperatures
lower than ~700°C are predominantly protonic, although this is not
indicated in Fig. 8, since it has not been directly measured. The grain
boundary resistance in this material was low and, as mentioned, the bulk
and grain boundary impedance could not be separated above 650°C.

The water vapor partial pressure dependence of the total conductivity from
500 to 1000°C, as measured in hydrogen, is shown in Fig. 9 a). The total
conductivity increases as a function of increasing water vapor partial
pressure. The typical conductivity behavior observed for a mixed ionic-
electronic conductor upon large variations in the oxygen partial pressure, is
shown in Fig. 9 b), for temperatures in the range 800 - 1000°C. The total
conductivity increases with decreasing oxygen partial pressures under
reducing conditions, and increases with increasing oxygen partial pressures
under oxidizing conditions. These two behaviors reflect an increasing n-
type and p-type contribution to the total conductivity, respectively.
Between these two regimes the total conductivity is essentially independent
of the oxygen partial pressure reflecting ionic conduction. Below 800°C
under reducing conditions, the conductivity is predominantly ionic showing
no significant n-type electronic conduction. One should note that the total
conductivity degrades with time, and that this seems to depend particularly
on the time measured under wet conditions and at low temperature.

88
a) Temperature (°C)
1000 800 600 400

σTotal(10 kHz)
10-2 σElectron
σProton

Conductivity (S/cm)
σOxygen ion
10-3

10-4

10-5

0.8 1.0 1.2 1.4 1.6 1.8


-1
1000/T (K )

b) Temperature (°C)
1000 800 600 400

σTotal(10 kHz)
σTotal(IS)
10-2 σElectron
Conductivity (S/cm)

σProton
σOxygen ion
10-3

10-4

10-5

0.8 1.0 1.2 1.4 1.6 1.8


-1
1000/T (K )

Fig. 8 Total conductivity and partial conductivities as a function of the


inverse absolute temperature for SrCe0.95Yb0.05O3-δ in a) wet
oxygen and b) wet hydrogen. σTotal(10 kHz) and σTotal(IS) are
obtained from a continuous ramp at 10 kHz and impedance
spectroscopy with 50°C steps, respectively.

89
a) 10-2
1000°C

Total conductivity (S/cm)


900°C

800°C

10-3 700°C

600°C

500°C

10-4
10-3 10-2

Water partial pressure (atm)

b) 0.050
1000°C
900°C
Total conductivity (S/cm)

800°C

0.010

0.005

10-20 10-15 10-10 10-5 100

Oxygen partial pressure (atm)

Fig. 9 Total conductivity of SrCe0.95Yb0.05O3-δ specimens with a)


varying water vapor partial pressure, measured in hydrogen at
500 - 1000°C, and b) varying oxygen partial pressure in the
range of 800 - 1000°C.

90
Discussion
Powder characteristics
As demonstrated in this work, phase pure SrCeO3 powders can be prepared
by spray pyrolysis followed by calcination at 1100 - 1200°C in nitrogen
atmosphere. Fine powders with particle size down to ~60 nm have been
obtained by ball milling of calcined powders, however some coarsening was
observed with increasing calcination temperature. The TG analysis (Fig. 3)
of SC powders in air indicates that SrCO3 introduced during ball milling is
removed below 800°C, in agreement with reported onset temperature for
synthesis of SC from a mix of SrCO3 and CeO2 in nitrogen [6]. The IR
spectrum of SC calcined at 900°C (Fig. 4), however, indicate that small
amount of SrCO3 is present in the powder even before milling. A larger
amount of SrCO3 in the SC5Yb powder calcined at 900°C, confirmed by
XRD (Fig. 2 b)) and IR (Fig. 4)) and assisted by the larger weight loss from
TG analysis (Fig. 3), indicate a stabilization of this secondary phase by
substitution with Yb. Strontium rich secondary phase (Sr2CeO4) as well as
ceria (CeO2) were also more dominating in the synthesized SC5Yb powder,
the former being detected by XRD in the SC5Yb powder after calcination at
1100°C, but not in the SC powder after calcination at 1000°C. These
features may be explained by the more basic nature of Yb-substituted
material, as reported for ceria [25], which in turn e.g. will make the material
more susceptible to CO2.

Densification and microstructure


By introducing a heat treatment at 850 and 1000°C prior to sintering, dense
SC and SC5Yb materials, with well defined microstructures, were obtained
by sintering for 2 hrs in air at 1350 - 1400°C. SC was found to sinter at a
lower temperature compared to SC5Yb and onset temperatures for sintering
being around ~200°C higher for the latter. The higher sintering temperature
in the Yb-substituted material can not be assigned to any ionic size effects,
as the ionic radius of Yb3+ equals that of Ce4+ [27]. The assisted introduction
of vacancies in the material is also unlikely to increase the sintering
temperature. Increased mobility of oxygen ions would rather aid the
sintering, however, the cations are more likely the rate limiting species. The
shift in onset temperature for sintering may be explained by the presence of
secondary phase, inhibiting the sintering in the SC5Yb material. Sr2CeO4
was detected in the SC5Yb powder, even after calcination at 1100°C. All
secondary phases seem to vanish at higher temperatures (1200°C), however,
at such temperatures coarsening of the powder will reduce the sinterability
as well. Despite lower onset temperature, the sintering rate was higher in

91
the SC5Yb material compared to SC, as demonstrated in Fig. 5. This effect
may be explained by larger surface area in starting powders compared to
that of SC. Densities of 96-98% in sintered SC and SC5Yb materials
sintered at 1350 - 1450°C in air are comparable to reported densities >98%
obtained by solid state ceramic method and sintering at 1550 - 1600°C in air
for 10 - 12 hrs [16,17], and from powders prepared by complexation route, by
sintering at 1300°C for 12 hrs in nitrogen flow [9].

Homogenous, fine grained microstructures were obtained for the SC5Yb


specimens sintered up to 1400°C. The more severe grain growth in the SC
materials, with resulting trapped pores indicates a high mobility of grain
boundaries compared to pores. The reduced growth in SC5Yb may, as the
sinterability, be explained by the presence of secondary phases in the
material. Segregation of secondary phases on the grain boundaries will pin
these and inhibit the growth. At elevated sintering temperatures the
secondary phases will disappear (presuming stoichiometric cation ratio in
the starting powder), alleviating the grain boundary pinning. This may be
the effect observed in SC5Yb when increasing the sintering temperature
from 1400 to 1450°C, resulting in remarkable grain growth.

Due to the anisotropic properties of SrCeO3, microcracking was observed in


large grained SC materials (> 6 µm), while the finer grained (1 - 2 µm)
SC5Yb materials were resistant towards microcracking. Based on these
observations the critical grain size may be estimated between 2 and 6 µm.
In a comparable work, grain size of ~7 µm have been reported for dense
SC5Yb materials [9], however studies of microstructure and grain size in
these materials are rare. The commonly used solid state ceramic route
produces coarse powders, hence, high sintering temperatures (>1500°C) are
needed to obtain dense materials. Consequently, the sintered materials will
consist of rather large grains (10 - 20 µm) [13]. This exceeds the suggested
critical size for microcracking, which in turn may help explain the poor
mechanical properties observed for SrCeO3 materials [7]. SrCeO3 based
materials with improved mechanical properties have been prepared by
introducing SrZrO3 (SrCe1-xZrxO3), however, the lower conductivity of
SrZrO3 based materials compared to SrCeO3 affects the final composite as
well [18]. Observations made in the present work, points out the importance
of high quality starting powders and controlled densification, with respect to
the mechanical properties of the sintered SrCeO3-based materials.

92
Conductivity
The conductivity results from this investigation generally reflect the
behaviour reported in the literature on acceptor substituted SrCeO3 [1,2,5,12-
18,28-33]
. SrCe0.95Yb0.05O3-δ exhibit mixed ionic-electronic conduction where
the ionic contribution is predominated by protons below ~1000°C. The
proton conductivity reaches a maximum of ~0.005 S/cm above 900°C.
Oxygen ion conductivity becomes the major ionic charge carrier above
1000°C. The p-type conductivity is considerably higher under oxidizing
conditions than the corresponding n-type conductivity under reducing
conditions. This conductivity behavior can be interpreted with basis in a
few elementary point defect reactions. Substitution of trivalent Yb for
fourvalent Ce yields an effectively negative defect (acceptor) that must be
charge-compensated either by consumption of other effectively negative or
formation of effectively positive defects. In SrCeO3 charge-compensation
occurs by formation of oxygen vacancies according to Eq. 3.

Yb 2 O 3 + 2SrO = 2Yb 'Ce + 2SrSrx + v •O• + 5O O


x
(3)

Under wet conditions oxygen vacancies may be hydrated through


interaction with water vapor. The protonic defect that forms is assumed to
be associated with a structural oxygen ion through the reaction in Eq. 4.

H 2 O(g ) + v •O• + O O
x
= 2OH •O (4)

Native defects are furthermore in equilibrium with the surrounding


atmospheres, as shown in Eq. 5, for oxygen vacancies in equilibrium with
electron holes predominating under high-temperature oxidizing conditions.

x 1
OO + 2h • = v •O• + O 2 (g ) (5)
2

By taking the intrinsic ionization between electron holes and electrons into
account all the point defects necessary to model the behavior encountered
during this investigation are in place. However, in order to obtain a full
mathematical description of the defect structure, and to determine
physicochemical parameters reflecting the conductivity characteristics, the
site balance and the electroneutrality must be included [34]. Within the
experimental window of the present investigation, acceptor doping in
SrCeO3 has, generally, been concluded to be charge compensated
predominantly by oxygen vacancies and protons. Hence, the
electroneutrality condition may be expressed by Eq. 6.

93
[Yb ] = [2v ]+ [OH ]
'
Ce
••
O

O (6)

By combining this electroneutrality and the site balance with the


equilibrium expression in Eq. 4, the concentration of protons and oxygen
vacancies may be resolved. Moreover, the conductivity of the different
charge carriers (indexed i) is proportional to their concentration (ci) and
mobility (µi), σi = zieciµi. On these bases the dependencies of the partial and
total conductivities with variations in the conditions can be modeled.
Thermodynamic parameters extracted from this modeling include the
standard entropy and enthalpy of the hydration reaction in Eq. 4, ∆S0 = -125
± 5 J/molK, ∆H0 = -145 ± 10 kJ/mol and, moreover, the preexponential
mobility and the activation enthalpy of defect mobility for the protons and
oxygen vacancies: µ0(H+) = 20 ± 5 cm2K/Vs, ∆Hmob(H+) = 55 ± 5 kJ/mol,
µ 0( v•O• ) = 20 ± 10 cm2K/Vs, ∆Hmob( v•O• ) = 60 ± 5 kJ/mol. Since it is not
possible to resolve independent expressions for the concentration of
electrons and electron holes, only apparent activation energies for p- and n-
type conductivity can be listed: EA(h·) = 65 kJ/mol and EA(e) = 350 kJ/mol.
As a first approximation in the modeling the entropy of the hydration
reaction was assumed to be -120 J/mol and, furthermore, the concentrations
of protons and oxygen vacancies were assumed to be small relative to the
concentration of oxygen sites [34-36].

Fig. 10 a) demonstrates, by way of example, the fit of the parameter-set to


the water vapor partial pressure dependence in hydrogen. Moreover, Fig. 10
b) illustrates how the different partial conductivities individually influence
the overall functional water vapor partial pressure dependence at 500 and
1000°C. One may note here how protons predominate at the low
temperature as a consequence of the exothermic nature (negative enthalpy)
of the hydration of oxygen vacancies. Oxygen ion conductivity barely
influences the total conductivity even under dry conditions at 500°C, and n-
type electronic conduction is orders of magnitude lower than the ionic
conductivity. By increasing the temperature the influence of protons
decreases, as indicated by reduced proton concentration relative to the
oxygen vacancy concentration. At high temperature, under these rather
reducing conditions, also n-type electronic conduction comes into play,
particularly at high water vapor partial pressures. The shift in the relative
influence of the three charge-carriers within the 500°C temperature interval
explains the observed change in curvature of the functional water vapor
partial pressure dependence (Fig. 10 a) and b)).

94
a) 10-2

1000°C
Total conductivity (S/cm)
900°C

800°C
10-3

700°C

600°C

10-4
500°C

10-5 10-4 10-3 10-2 10-1 100

Water partial pressure (atm)

b)
10-2 σTotal(1000°C)
σProton(1000°C)
10-3 σOxygen ion(1000°C)
Conductivity (S/cm)

σElectron(1000°C)
10-4

10-5

10-6 σTotal(500°C)
σProton(500°C)
10-7
σOxygen ion(500°C)
σElectron(500°C)
10-8

10-5 10-4 10-3 10-2 10-1 100

Water partial pressure (atm)

Fig. 10 a) Total conductivity curves fit to the experimental data for


SrCe0.95Yb0.05O3-δ, for varying water vapor partial pressure, in
hydrogen from 500 to 1000°C. b) Total and partial
conductivities with varying partial pressure of water vapor at
500°C (bold curves) and 1000°C (slim curves). Circles and
lines experimental results data and modeled conductivity curves,
respectively.

95
Despite the extensive literature on the functional properties of the acceptor
substituted SrCeO3, only a few investigations have so far derived similar
physicochemical properties as reported here. The thermodynamics of the
dissolution of water according to the reaction in Eq. 4 have been measured
by thermogravimetry and also estimated from conductivity measurements
[17,37,38]
. Hydration enthalpies are reported in the range from -115 to -160
kJ/mol, which correspond nicely with the values obtained here.

The degradation in the total conductivity encountered during these


experiments may have several reasons. The basic high-temperature proton
conducting perovskites, in particular BaCeO3 and SrCeO3, are renowned for
their reactivity towards CO2. However, since gas mixtures with relatively
low levels of CO2 have been applied during this investigation, this is not
believed to cause the decline observed in the total conductivity with time.
More likely this is a result of SrCeO3 decomposeing to Sr(OH)2 and CeO2
[7]
, due to long-term exposure in wet atmospheres at relatively low
temperature (slow temperature ramps down to 300°C). It should also be
recognized that SrCeO3 seems to react with Pt and that this may change both
the properties of the material and affect the processes at the electrode.

Conclusions
Spray pyrolysis has been proven a good method for preparing powders of
orthorombic SrCeO3 and SrCe0.95Yb0.05O3-δ. Effective breaking of
agglomerate spheres formed during spray pyrolysis was done by ball milling
in iso-propanol resulting in particle size down to 0.06 µm. All secondary
phases present in the as-synthesized SrCeO3 and SrCe0.95Yb0.05O3-δ powders
can be removed by calcination in nitrogen atmosphere at 1100 and 1200°C,
respectively. Dense (>96%) materials with homogenous microstructures
have successfully been prepared. Due to lower mobility of grain boundaries
in the material the grain size in the SrCe0.95Yb0.05O3-δ (1 - 2 µm) materials
were somewhat lower than those found in SrCeO3 (>6 µm). Some closed
porosity and microcracking was observed in the SrCeO3 materials, while the
fully dense (~98%) SrCe0.95Yb0.05O3-δ materials seemed were free of cracks
and closed pores. The electrical properties of the Yb-substituted materials
were in very good agreement with literature data, showing p-type
conductivity in oxidizing and n-type conductivity in reducing atmosphere.
The ionic contribution to the total conductivity is predominately from
protons below 1000°C and the protonic conductivity reaches a maximum of
0.005 S/cm at 1000°C.

96
Acknowledgement
Work was supported by the Research Council of Norway, Grant No.
1585171431 (NANOMAT). Øystein Andersen is acknowledged for
operating the spray pyrolysis.

References
[1] H. Iwahara, T. Esaka, H. Uchida, N. Maeda, Proton conduction in
sintered oxides and it’s application to steam electrolysis for
hydrogen production, Solid State Ionics, 3/4, 359-363 (1981)
[2] H. Iwahara, Technological challenges in the application of proton
conducting ceramics, Solid State Ionics, 77, 289-298 (1995)
[3] T. Norby, Solid-state protonic conductors: principles, properties,
progress and prospects, Solid State Ionics, 125, 1-10 (1999)
[4] T. Schober, Applications of oxidic high-temperature proton
conductors, Solid State Ionics, 162-163, 277-281 (2003)
[5] T. Yajima, H. Iwahara, Studies on proton behavior in doped
perovskite-type oxides: (II) Dependence of equilibrium hydrogen
concentration and mobility on dopant content in Yb-doped SrCeO3,
Solid State Ionics, 53-56, 983-988 (1992)
[6] M.J. Scholten, J. Schoonman, Synthesis of strontium and barium
cerate and their reaction with carbon dioxide, Solid State Ionics, 61,
83-91 (1993)
[7] K.-D. Kreuer, On the development of proton conducting materials
for technological applications, Solid State Ionics, 97, 1-15 (1997)
[8] A.N. Shirsat, K.N.G. Kaimal, S.R. Bharadwaj, D. Das,
Thermodynamic stability of SrCeO3, J. Solid State Chemistry, 177,
2007-2013 (2004)
[9] K.J. de Vires, Electrical and mechanical properties of proton
conducting SrCe0.95Yb0.05O3-α, Solid State Ionics, 100, 193-200
(1997)
[10] S. Cheng, V.K. Gupta, J.Y.S. Lin, Synthesis and hydrogen
permeation properties of asymmetric proton-conducting ceramic
membranes, Solid State Ionics, 176, 2653-2662 (2005)
[11] S.V. Chavan, A.K. Tyagi, Preparation of Sr0.09Ce0.91O1.91, SrCeO3,
and Sr2CeO4 by glycine-nitrate combustion: Crucial role of oxidant-
to-fuel ratio, J. Mater. Res., 19 (11), 3181-3188 (2004)
[12] T. Scherban, A.S. Nowick, Bulk protonic conduction in Yb-doped
SrCeO3, Solid State Ionics, 35, 189-194 (1989)

97
[13] N. Matsunami, T. Yajima, H. Iwahara, Permeation of implanted
deuterium through SrCeO3 (5% Yb), Nuclear Instruments and
Methods in Physics Research, B65, 278-281 (1992)
[14] I. Kosacki, J. Schoonman, M. Balanski, Raman scattering and ionic
transport in SrCe1-xYbxO3, Solid State Ionics, 57, 345-351 (1992)
[15] H. Iwahara, T. Yajima, T. Hibino, K. Ozaki, H. Suzuki, Protonic
conduction in calcium, strontium and barium zirconates, Solid State
Ionics, 61, 65-69 (1993)
[16] I. Kosacki, H.L. Tuller, Mixed conductivity in SrCe0.95Yb0.05O3
protonic conductors, Solid State Ionics, 80, 223-229 (1995)
[17] U. Reichel, R.R. Arons, W. Schilling, Investigation of n-type
electronic defects in the protonic conductor SrCe1-xYxO3-α,, Solid
State Ionics, 86-88, 639-645 (1996)
[18] T. Matzeke, M. Cappadonia, Proton conductive perovskite solid
solutions with enhanced mechanical stability, Solid State Ionics, 86-
88, 659-663 (1996)
[19] S.V. Chavan, A.K. Tyagi, Sub-solidus phase equilibria in CeO2-SrO
system, Thermochimica Acta, 390, 79-82 (2002)
[20] H. Matsumoto, T. Shimura, H. Iwahara, T. Higuchi, K. Yashiro, A.
Kaimai, T. Kawada, J. Mizusaki, Hydrogen separation using proton-
conducting perovskites, Journal of Alloys and Compounds, 408-412,
456-462 (2006)
[21] H. Taherparvar, J.A. Kilner, R. Baker, M. Sahibzada, Effect of
humidification at anode and cathode in proton-conducting SOFCs,
Solid State Ionics, 162-163, 297-303 (2003)
[22] T. Norby, "Electronic Book", Norwegian Electro Ceramics AS,
www.norecs.com
[23] T. Norby, EMF method determination of conductivity contributions
from protons and other ions in oxides, Solid State Ionics, 28-30,
1586-1591 (1988)
[24] D.P. Sutija, T. Norby, P. Björnbom, Transport number
determination by the concentration-cell / open-circiut voltage
method for oxides with mixed electronic, ionic and protonic
conductivity, Solid State Ionics, 77, 167-174 (1995)
[25] K. Nakamoto, Infrared and Raman Spectra of Inorganic and
Coordination Compounds, 2nd ed., John Wiley & Sons Inc., New
York (1997)
[26] T. Mokkelbost, I. Kaus, T. Grande, M.-A. Einarsrud, Combustion
synthesis and characterization of nanocrystalline CeO2-based
powders, Chem. Mater., 16, 5489-5494 (2004)
[27] R.D. Shannon, Revised effective ionic radii and systematic studies of
interatomic distances in halides and chalcogenides, Acta Cryst.,
A32, 751-767 (1976)

98
[28] H. Uchida, N. Maeda, H. Iwahara, Relation between proton and hole
conduction in SrCeO3-based solid electrolytes under water-
containing atmospheres at high temperatures, Solid State Ionics, 11,
117-124 (1983)
[29] H. Uchida, H. Yoshikawa, T Eseka, S. Ohtsu, H. Iwahara, Formation
of protons in SrCeO3-based proton conducting oxides. Part II.
Evaluation of proton concentration and mobility in Yb-doped
SrCeO3, Solid State Ionics, 36, 89-95 (1989)
[30] T. Yajima, H. Iwahara, H. Uchida, K. Koide, Relation between
proton conduction and concentration of oxide ion vacancy in SrCeO3
based sintered oxides, Solid State Ionics, 40/41, 914-917 (1990)
[31] T. Hibino, K. Mizutani, T. Yajima, H. Iwahara, Evaluation of proton
conductivity in SrCeO3, BaCeO3, CaZrO3 and SrZrO3 by
temperature programmed desorption method, Solid State Ionics, 57,
303-306 (1992)
[32] H. Iwahara, Proton conducting ceramics and their applications,
Solid State Ionics, 86-88, 9-15 (1996)
[33] T. Schober, F. Krug, W. Schilling, Criteria for the application of
high temperature proton conductors in SOFCs, Solid State Ionics,
97, 369-373 (1997)
[34] R. Haugsrud, T. Norby, Proton conduction in rare-earth ortho-
niobates and ortho-tantalates, Nature Materials, 5, 193196 (2006)
[35] K.D. Kreuer, Aspects of the formation and mobility of protonic
charge carriers and the stability of perovskite-type oxides, Solid
State Ionics, 125, 285-302 (1999)
[36] Y. Larring, Protons and oxygen vacancies in acceptor-substituted
rare earth oxides, PhD. Thesis, University of Oslo (1998)
[37] T. Yajima, H. Iwahara, Studies on behavior and mobility of protons
in doped perovskite-type oxides (I) In situ measurement of hydrogen
concentration in SrCe0.95Yb0.05O3-α at high temperatures, Solid State
Ionics, 50, 281-286 (1992)
[38] F. Krug, T. Schober, T. Springer, In situ measurements of the water
uptake in Yb doped SrCeO3, Solid State Ionics, 81, 111-118 (1995)

99
100
PAPER IV
102
Preparation and characterization of Barium
Zirconate ceramics

Paul Inge Dahl, Hilde Lea Lein, Tor Grande, Mari-Ann Einarsrud †

Department of Materials Science and Engineering


Norwegian University of Science and Technology
NO-7491 Trondheim, Norway

Christian Kjølseth, Truls Norby, Reidar Haugsrud

Centre for Materials Science and Nanotechnology


Department of Chemistry, University of Oslo
NO-0349 Oslo, Norway

Abstract
Powders of pure and 10% yttrium substituted barium zirconate (BaZrO3 /
BaZr0.9Y0.1O2.95) were prepared by spray pyrolysis of nitrate salt solutions.
The crystalline powders were calcined at 1000°C to remove secondary
phases and agglomerates were effectively broken down by ball milling
giving particle size in the range 0.09 – 0.17 µm. Despite of similar
characteristics of the powders, the densification properties were poorer for
the yttrium substituted material. Severe grain growth was observed during
conventional sintering (1600°C) of BaZrO3 resulting in average grain size
up to 18 µm and only < 92% relative density. Dense BaZrO3 and
BaZr0.9Y0.1O2.95 materials (~98%) were prepared by hot pressing. The
mobility of the grain boundaries was efficiently suppressed by application
of uniaxial pressure during sintering, resulting in homogenous
microstructures and average grain size down to 0.42 µm. The electrical
properties of BaZr0.9Y0.1O2.95 are in agreement with the literature, and high
grain boundary resistance was observed for both materials. Slightly lower
conductivity was observed for BaZrO3 in wet compared to dry atmosphere.

Corresponding author. E-mail: mari-ann.einarsrud@material.ntnu.no

103
Introduction
Perovskite oxides (ABO3) such as barium and strontium based cerates and
zirconates (A = Ba, Sr and B = Ce, Zr) are well known for their electrical
properties. Substituted with rare earth oxides these materials exhibit high
proton conductivity. Potential application areas for such ceramic proton
conductors include gas sensors, fuel cells and membrane technology [1].
Barium zirconate (BaZrO3) based materials exhibit lower proton
conductivity compared to the isostructural barium cerates (BaCeO3). This is
due to grain boundaries being highly resistive to proton transport, and
consequently decreasing the conductivity by several orders of magnitude
which is detrimental to applications like electrolytes in an SOFC. The
reason for the high grain boundary resistance is not yet understood. Despite
higher conductivity reported for the cerate based ceramics, zirconate
materials seem superior with respect to chemical and mechanical properties
[2-5]
. The idea of combining the properties of BaCeO3 and BaZrO3 has been
studied in mixed systems (BaCe0.9-xZrx Y0.1O3-δ) where increasing chemical
stability against CO2 was reported with increasing Zr content. However, the
increasing Zr content resulted in decreased proton conductivity [6,7].
Consequently, BaZrO3 substituted with trivalent rare earths such as Y or Yb
has become state of the art materials among ceramic high temperature
proton conductors.

The conductivity of yttrium substituted BaZrO3 is well documented in the


literature [8-17]. For all these studies ceramic powders have been prepared by
solid state reaction of oxides, carbonates and/or acetates at temperatures in
the range 1200 – 1450°C. This method generally produces coarse powders.
Fine powders with more defined morphology have been prepared by co-
precipitation [18,19], sol-gel techniques [20,21] and hydrothermal synthesis [21].

A challenge in preparation and use of Ba/Sr-based zirconates (and cerates)


is possible reactions with CO2 as indicated for BaZrO3 in Eq. (1).

BaZrO3(s) + CO2(g) = BaCO3(s) + ZrO2(s) (1)

The presence of BaCO3 has not been reported in BaZrO3 powders or dense
specimens heat treated at high temperatures (>1000°C). However, as the
equilibrium constant for Eq. (1) equals 1 around 320°C in air [4], formation
of BaCO3 may occur during cooling. The stability of carbonates can narrow
the temperature region of use for these ceramic proton conductors. It is not
clear how the presence of BaCO3 affects the conductivity.

At high temperatures, BaZrO3 can decompose according to Eq. (2) as


documented by thermodynamic studies [22,23].

104
BaZrO3(s) = BaO(g) + ZrO2(s) (2)

Evaporation of BaO(g) has been observed on the surface of Y-substituted


BaZrO3 specimens heat treated above 1200°C and above 1500°C, bulk is
affected as well [20]. These observations indicate that preparation of dense
and phase pure BaZrO3 ceramics might be challenging.

No thorough sintering studies of BaZrO3 ceramics seem to be available.


Dense (> 95%) BaZrO3 from powders made by solid state method can only
be obtained by conventional sintering above 1700°C [7,11,12,15,16]. Using
spark plasma sintering, dense (>95%) ceramics of pure and Y-substituted
BaZrO3 can be obtained at 1500 and 1600°C, respectively [24]. Besides the
latter work there are no consistent densification routes for controlled
microstructure and grain size available for these materials unless
undesirable sintering aids such as ZnO is added [25].

Based on the above information the preset study aims to establish


reproducible procedures for preparation of dense BaZrO3 materials with
designed microstructure. The quality of the starting powders is important
for this matter. Hence, the first part of this deals with the preparation of
pure and 10% Y-substituted BaZrO3 powders with special focus on the
mentioned problem with formation of BaCO3. Two different sintering
techniques have been used for densification of BaZrO3 specimens (from
prepared powders) in order to design different microstructures. The
materials have been thoroughly characterized, and being the most central
property of the BaZrO3 materials, the electrical conductivity has been
investigated. The conductivity measurements reported herein are, as such,
preliminary results from a large matrix of measurements on Y-substituted
BaZrO3 where the overall purpose is to contribute to the understanding of
the grain boundary effects. The different electrical behaviour of the
nominally BaZrO3 compared to the Y-substituted material are discussed in
the present work, however, for a deeper discussion on the conductivity
characteristics of acceptor substituted BaZrO3 and in particular of the effects
of grain boundary resistance, forth-coming papers should be consulted [26].

Experimental
Powder synthesis
Powders of BaZrO3 (BZ) and BaZr0.90Y0.10O2.95 (two similar batches: YBZ-
1, YBZ-2) were synthesized by spray pyrolysis of nitrate salt solutions.
Aqueous solutions of zirconyl nitrate (ZrO(NO3)2·xH2O, Acros, 99.5%) and

105
yttrium nitrate (Y(NO3)3·6H2O, Merck, 99.9%) were standardized by
thermogravimetry, mixed in stoichiometric proportions with dried (250°C,
24 hrs) barium nitrate powder (Ba(NO3)2, Merck, 99%). Due to the limited
solubility of Ba(NO3)2 (10.1 g / 100 g H2O), the total cation concentration in
the mixed solution was 0.2 M. The spray pyrolysis was performed with a
feed rate of 2 L/h. The temperature by the nozzle was 840-845°C and the
output temperature was ~560°C. The as-synthesized powders were calcined
at 1000°C for 48 hrs in ambient air and 1100°C for 6 hrs in nitrogen flow.
Calcined powders were ball milled with yttria stabilized zirconia (YSZ)
balls in isopropanol for 24 hrs, dried at 400°C for 24 hrs in ambient air and
sieved (150 µm).

Sample preparation
Green bodies were made by uniaxial pressing at 64 – 128 MPa giving green
densities in the range 44 – 51% of theoretical value. Conventional sintering
of BZ and YBZ-1 was performed in a muffle furnace (Entech, SF-4/17) at
1600°C in ambient air with a presintering step at 1000°C for 24 hrs. For
BaZrO3, the sintering time was varied from 1 to 18 hrs while the sintering
time for BaZr0.9Y0.1O2.95 was 6 hrs. Two different hot presses were used,
depending on desired temperature region and applied pressure. Hot pressing
(HP) up to 1500°C was performed in a clam furnace (Entech, VSTF 40/15)
under flowing nitrogen gas, with an applied pressure of 25 MPa. A
commercial hot press (Thermal Technology Inc. HP50-7010G) was used for
hot pressing at 1450 – 1750°C in flowing argon gas, with an applied
pressure of 50 MPa. Cylindrical graphite pressing dies (15 – 25 mm) was
used. The uniaxial pressure was applied after heating to sintering
temperature with a rate of 600°C/h. The sintering time was varied from 1 to
6 hrs and the pressure was released before cooling to room temperature at
600°C/h. As the HP was performed in reducing atmosphere the specimens
were reoxidized at 1400°C in flowing O2 for 24 hrs after sintering.

Characterization
Thermogravimetric analysis (TGA) (Netzsch STA 449 C Jupiter) was
performed on as-synthesized powders (BZ and YBZ-1) by heating to
1500°C in air with a heating rate of 3°C/min. As-synthesized and calcined
powders as well as surface of sintered specimens were analyzed by X-ray
diffraction (XRD) (Cu Kα, Phillips PW1730/10). Powders from crushed
samples sintered at 1500°C for 6 hrs and added 30 weight% Si standard
were used to obtain XRD data from which the lattice parameters and

106
theoretical densities were calculated by the Rietveld method. Infrared
spectroscopy (IR) was performed on potassium bromide (Merck, KBr for IR
spectroscopy Uvasol®) pellets containing 1 wt% BZ or YBZ-1 powder using
a Bruker IFS 66v spectrophotometer. Nitrogen adsorption on de-gassed
powders (250°C, 24 hrs) was measured (Micromeritics Tristar) using the 5
point BET equation to give the specific surface area. The particle size was
calculated from these results assuming spherical particles.

Dilatometer studies (Netzch, DIL 402 C / 402 E) were performed in air


using a heating rate of 3°C/min up to 1500°C and 1600°C for BaZrO3 (BZ)
and BaZr0.9Y0.1O2.95 (YBZ-1 and YBZ-2), respectively. Densities of
sintered specimens were measured by the Archimedean method (ISO 5017)
in isopropanol. The microstructure of synthesized powders and dense
specimens (oxidized/etched at 1400°C) was studied by scanning electron
microscopy (SEM) (Hitachi S-3500N). Average grain size of sintered
specimens was calculated by the linear intercept method [27] over a
minimum of 50 grains.

Electrical properties
Electrical characterization was performed on (BZ), with > 95 % density, and
(YBZ-1), with > 88% density. Circular Pt electrodes (~10 mm diameter)
were attached to each side of the specimens and mounted in ProbostatTM
measurement cells [28]. The electrical characterization was carried out in wet
(p(H2O) ~ 0.025 atm) and dry (p(H2O) ~ 3·10-5 atm) oxygen by means of
impedance spectroscopy (Hewlett Packard (HP) 4192A) in the frequency
range 5 MHz to 5 Hz, with an oscillation voltage of 0.5 V. Dependencies of
the oxygen partial pressure at different temperatures were obtained by
varying the ratio of argon and oxygen at constant water partial pressure.
Impedance spectroscopy was performed every 50°C between 200 and
1000°C for BZ and between 150 and 600°C for YBZ-1. However, due to
instrument limitations it was only possible to get reasonable values from
deconvolution of the spectra in the temperature range 450-750°C (wet) and
450-600°C (dry) for BZ and 200-400°C for YBZ-1 (wet).

The impedance spectra were deconvoluted using the Equivalent Circuit for
Windows program [29] and fitted to equivalent circuits consisting of two
parallel RQ elements in series, where R is the impedance and Q is a constant
[
phase capacitive element. Q has the impedance Z Q = Y ( jω) n ]
−1
, where
j = − 1 , ω = frequency, Y and n are constants, and n ranges between 0 and
1. Experimentally, the exponent n characterizing the subcircuit element

107
(RQ) is rather close to 1, and thus these elements behave much like
(1 ) (1 −1)
capacitors. The capacitance is given by C = Y n R n . The equivalent
circuit (RbCb)(RgbCgb) was used in the deconvolution and had capacitances
for bulk and grain boundaries in the typical order of 10 pF and 10 nF,
respectively. From these data specific conductivities for both bulk and grain
boundaries were, furthermore, calculated applying the Brick-Layer model
[30]
. The presented specific conductivity data have not been corrected for
porosity. This can be done using the measured density ρ and the theoretical
density, ρth, following the empirical relationship: σ = σmeasured / (ρ/ρth)2 [31].

Results
Characterization of synthesized powders
The TG curves in Fig. 1 show weight loss observed for as-synthesized
powder of both BaZrO3 (BZ) and BaZr0.9Y0.1O2.95 (YBZ-1) upon heating to
1500°C. Though there was a continuous weight loss all the way to 1500°C,
less than 0.5wt% weight loss was observed above ~950°C.

100

99
Weight change (%)

98

97

96

95
BZ
94 YBZ-1

0 300 600 900 1200 1500

Temperature (°C)

Fig. 1 TG analysis of as-synthesized BaZrO3 (BZ) and BaZr0.9Y0.1O2.95


(YBZ-1) powders.

108
From XRD of the powder, the lattice parameter for cubic BaZrO3 was
calculated to a = 4.185 Å giving a theoretical density of 6.26 g/cm3. Lattice
parameter and theoretical density for BaZr0.9Y0.1O2.95 (YBZ-1) was
calculated to a = 4.193 Å and 6.21 g/cm3, respectively. Calculated lattice
parameters are slightly lower, but still in accordance with reported values
[11-12]
.

X-ray diffractograms of BaZrO3 and BaZr0.9Y0.1O2.95 powders are displayed


in Fig. 2 a) and b), respectively. For the as-synthesized BZ powder (Fig. 2
a)) presence of secondary phases (BaCO3 and BaO) was found, however,
these phases were apparently removed by calcination at 1000°C. During hot
pressing at 1500°C, a secondary phase of ZrO2, indicated by the additional
reflection near 2θ = 29°, was formed on the specimen surface. The
reflection, though weak, was also observed for bulk material. No traces of
BaCO3 could be detected by XRD in the hot pressed BZ specimens. The X-
ray diffractograms in Fig. 2 b) demonstrates that BaCO3 is present in the as-
synthesized YBZ-1 powder, however, a lower intensity of the BaCO3
reflections indicates a smaller amount compared to the as-synthesized
BaZrO3 powder (Fig. 2 a)). After calcination at 1000°C and hot pressing at
1500°C no secondary phases were detected by XRD for YBZ-1. For YBZ-
2, treated the same way as YBZ-1, BaCO3 was still observed by XRD after
calcination at 1000°C. As for BZ, a reflection around 2θ = 29° was detected
for the hot pressed YBZ-2 specimens, even in the bulk. The intensity of this
reflection was not significantly different for YBZ-2 specimens hot pressed
at 1450 and 1750°C, indicating that the composition of bulk is constant in
this temperature range.

109
a) BaZrO3: 
BaCO3: 
BaO: 
ZrO2: 
Relative intensity

BZ HP 1500°C
bulk

BZ HP 1500°C
surface


BZ 1000°C

 
    
 
   BZ as-synth.

20 30 40 50 60

2 θ (°)
()

b) BaZr0.9Y0.1O2.95: 
BaCO3: 
BaO:
ZrO2:
Relative intensity

YBZ-2 HP 1750°C

YBZ-2 HP 1450°C

YBZ-2 1000°C

YBZ-1 HP 1500°C

YBZ-1 1000°C

       
      YBZ-1 as-synth.

20 30 40 50 60

2 θ (°)
()

Fig. 2 a) XRD of as-synthesized and calcined (1000°C) BZ powder, as


well as surface and bulk of BZ specimen hot pressed (25 MPa,
1 h) at 1500°C.
b) XRD of as-synthesized and calcined (1000°C) YBZ-1 and YBZ-
2 powders and bulk of YBZ-1 specimen hot pressed (25 MPa, 1
h) at 1500°C and YBZ-2 specimens hot pressed (50 MPa, 1 h)
at 1450 and 1750°C.

110
Fig. 3 presents SEM micrographs of calcined BaZrO3 (BZ) and
BaZr0.9Y0.1O2.95 (YBZ-1 and YBZ-2) powders, before and after milling.
Spray pyrolysis produced homogenous powders with particle size <5 µm, as
seen for the BZ powder shown in Fig. 3 a). The particles are typically
shaped like distorted hollow spheres as displayed for BZ and YBZ-1 in Fig.
3 b) and c), respectively. The particle size was reduced significantly by ball
milling, as demonstrated by the SEM micrographs of ball milled BZ, YBZ-1
and YBZ-2 in Fig. 3 d), e) and f), respectively. Table 1 demonstrates how
the surface area and calculated particle size of BZ and YBZ-1 powders vary
with calcination temperature. Only minor decrease in surface area was
observed for BZ and YBZ-1 with increasing calcination temperature
showing that these powders are resistant to coarsening with increasing
calcination temperature. The YBZ-1 powder calcined at 1000°C for 48 hrs
had a higher surface area (10.3 m2/g) compared to the YBZ-2 powder
treated the same way (5.8 m2/g).

Table 1 Surface area of BaZrO3 (BZ) and BaZr0.9Y0.1O2.95 (YBZ-1 and


YBZ-2) powders calculated from nitrogen adsorption using the
BET equation. Spherical particles were assumed for estimation
of the particle size.
Tcalcination tcalcination Surface area Particle size
Compound
(°C) (hrs) (m2/g) (µm)
BZ 850* 0 9.0 0.11
BZ 1000 48 7.5 0.13
BZ 1100 6 5.8 0.16
YBZ-1 1000 48 10.3 0.09
YBZ-1 1100 6 7.0 0.14
YBZ-2 1000 48 5.8 0.17

* As-synthesized powder.

111
Fig. 3 SEM micrographs of BZ, YBZ-1 and YBZ-2 powders calcined at
1000°C. a) and b) BZ before ball milling. c)YBZ-1 before ball
milling. d) BZ after ball milling. e)YBZ-1 after ball milling and
f) YBZ-2 after ball milling.

112
IR spectra of BZ and YBZ-1 powders are shown in Fig. 4. The BaCO3
spectrum is included as a reference. The broad band in the range of 500 to
800 cm-1, observed for all prepared powders, is due to the stretching of the
metal-oxygen bonds [32]. The intensity of the BaCO3 bands in the YBZ-1
powders is reduced with increasing calcination temperature. No bands
assigned to BaCO3 were detected for YBZ-1 powder from specimen
sintered at 1600°C. In addition to the wide carbonate band in the range of
1200 to 1700 cm-1, a narrow sharp peak was detected for the as-synthesized
YBZ-1 powder, at frequency 1384 cm-1, probably due to physically
adsorbed CO2 [33]. The wide band in the frequency region from 3000 to
3600 cm-1 is assigned to O-H stretching of physiosorbed water or from
surface adsorbed hydroxyl groups. IR spectra of BZ powders calcined at
different temperatures show the same trends as the YBZ-1 powders.

1600°C
1100°C
Relative transmittance

1000°C

as-synthesized

as-synthesized

BaCO3
YBZ-1
BZ

400 1000 1600 2200 2800 3400 4000

Wave number (cm-1)

Fig. 4 IR spectra of BaCO3 powder, as-synthesized BZ and YBZ-1


powder, YBZ-1 powder calcined at 1000°C and 1100°C as well
as YBZ-1 powder from crushed YBZ-1 specimen sintered at
1600°C.

113
Densification
The linear shrinkage curve for BZ in Fig. 5 shows onset for sintering around
1200°C. The curve for YBZ-1 shows the same on-set temperature,
however, the curve is remarkably wider due to lower sintering rate. The
shrinkage for YBZ-2 starts around 950°C and flattens out in a plateau before
the main shrinkage appears above 1300°C. The presence of a secondary
phase (BaCO3) as confirmed by XRD (Fig. 2 b)), may explain the first
shrinkage step for YBZ-2. The density and porosity of BZ conventionally
sintered at 1600°C are shown in Fig. 6 a). An increase in sintering time
from 1 to 6 hrs resulted in an increase in density from 82.5% to 91.1%. No
significant change in density was observed with increased hold time beyond
6 hrs at 1600°C, and the porosity was mainly closed (7.5%). In comparison,
the density of YBZ-1 specimen sintered at 1600°C for 6 hrs was 74%.
Density and grain size of sintered BaZrO3 and BaZr0.9Y0.1O2.95 specimens
are presented in Table 2.

-2

-4
dL/Lo (%)

YBZ-2

-6

-8 YBZ-1

-10

BZ
-12

800 1000 1200 1400 1600

Temperature (°C)

Fig. 5 Linear shrinkage curves for BaZrO3 (BZ) and BaZr0.90Y0.10O2.95


(YBZ-1 and YBZ-2). All powders were calcined at 1000°C for 48
hrs, ball milled with YSZ balls in iso-propanol for 24 hrs and
dried for 24 hrs at 400°C. Green body densities were in the
range 49 – 51% of theoretical density.

114
a) 100 10

Closed porosity 8
95

Porosity (%)
Density (%)
Density 6
90
4

85
2
Open porosity

80 0
0 2 4 6 8 10 12 14 16 18

Time (hrs)

b) 100 10

Density
8
95

Porosity (%)
Density (%)

6
90
4
Closed porosity
85
2
Open porosity

80 0
0 1 2 3 4 5 6

Time (hrs)

c) 100 10
Density
8
95
Porosity (%)
Density (%)

6
90
4
Closed porosity
85
2
Open porosity
80 0
1450 1500 1550 1600 1650 1700 1750

Temperature (°C)

Fig. 6 Density, closed and open porosity for a) BaZrO3 conventionally


sintered at 1600°C for 1 – 18 hrs in ambient air, b) BaZrO3 hot
pressed at 1500°C (25 MPa) for 1 – 6 hrs and c) YBZ-2
specimens hot pressed for 1 h (50 MPa) at indicated
temperatures.

115
Density and porosity as a function of HP time at 1500°C (25 MPa) are
shown for BaZrO3 (BZ) in Fig. 6 b). As for the conventionally sintered BZ
(Fig. 6 a)) some closed porosity seems inevitable for HP at 1500°C,
however, the amount is lower (~3%). Density, closed and open porosity for
YBZ-2 specimens obtained by HP at different temperatures are shown in
Fig. 6 c). A significant increase in the density, from 85.6 to 96.7%, was
observed for YBZ-2 when the HP (1 h, 50 MPa) temperature was increased
from 1450 to 1550°C. 97.6% density was obtained for both YBZ-2 after HP
(1h, 50 MPa) at 1750°C and BZ after HP (1 h, 50 MPa) at 1650°C. The
results of HP are summarized in Table 2. The BZ and YBZ-1 specimens
appeared black due to reduction after hot pressing. The colour changed to
light brown for BaZrO3 and green/white for BaZr0.9Y0.1O2.95 after reheat
treatment at 1400°C for 24 hrs in flowing oxygen.

Table 2 Density and grain size of conventionally sintered and hot


pressed BaZrO3 (BZ) and BaZr0.90Y0.10O2.95 (YBZ-1 / YBZ-2)
specimens.
Tsintering tsintering Pressure Density Gran size
Compound
(°C) (hrs) (MPa) (%) (µm)
BZ 1600 1 - 82.5
BZ 1600 18 - 91.9 18
BZ 1400 1 25 87.5
BZ 1500 1 25 93.4 0.46
BZ 1500 3 25 95.8 0.77
BZ 1500 6 25 96.1 1.0
BZ 1650 1 50 97.6 0.71
YBZ-1 1600 6 - 74.2
YBZ-1 1500 1 25 88.5 0.4*
YBZ-2 1450 1 50 85.6 0.4*
YBZ-2 1550 1 50 96.7 0.46
YBZ-2 1650 1 50 96.3 0.49
YBZ-2 1750 1 50 97.8 0.60

*Porous material – exact value not attainable.

116
Microstructure and grain growth
As revealed by the optical micrograph in Fig. 7 a), BZ conventionally
sintered at 1600°C for 12 hrs contained large grains (17 µm). In Fig. 7 b)
the SEM micrograph of BZ conventionally sintered at 1600°C for 18 hrs
confirms the presence of pores enclosed inside grains. Only a minor
increase in average grain size (16 - 18 µm, as given in Table 2, was
observed for BZ when the sintering time at 1600°C was increased from 6 to
18 hrs. The SEM micrographs in Fig. 7 c), d) and e) demonstrates how the
grain size in BZ increases when the HP time is increased from 1 to 6 hrs at
1500°C. HP at 1650°C for 1 h, with an applied pressure of 50 MPa also
resulted in grains (Fig. 7 f)) similar to, or smaller than those obtained by HP
at 1500°C for 3 and 6 hrs (Fig. 7 d) and e)). Fig. 7 g) and h) reveals the
homogenous fine grained microstructures of YBZ-2 specimens HP (1 h, 50
MPa) at 1550 and 1750°C, respectively. Table 2 summarizes the average
grain size of hot pressed specimens of BZ and YBZ-2.

117
Fig. 7 a) Optical micrograph of BZ conventionally sintered at 1600°C
for 12 hrs. b) through h) SEM micrographs of: b) BZ
conventionally sintered at 1600°C for 18 hrs, c) BZ specimens
hot pressed at 1500°C (1 h, 25 MPa) for 1 h, d) BZ hot pressed

118
at 1500°C (3 hrs, 25 MPa), e) BZ hot pressed at 1500°C (6 hrs,
25 MPa), f) BZ hot pressed at 1650°C (1 h, 50 MPa), g) YBZ-2
hot pressed at 1550°C (1 h, 50 MPa) and h) YBZ-2 hot pressed
at 1750°C (1 h, 50 MPa)

Electrical properties
Specific bulk and grain boundary conductivities as a function of the inverse
absolute temperature in wet and dry oxygen, are shown for hot pressed BZ
(1500°C, 3h, 25 MPa), with >95 % density and average grain size ~0.8 µm,
in Fig. 8 a). The conductivity in BZ was slightly higher under dry as
compared to wet conditions and the bulk conductivity was considerably
higher than the specific grain boundary conductivity both under wet and dry
conditions. The grain boundary conductivity under wet conditions exhibits a
straight-line Arrhenius behavior with an apparent activation energy of 170
kJ/mol. The bulk data exhibits lower activation energy and the temperature
dependence decreases with increasing temperature from 600 to 800°C. Fig.
8 b) shows the p(O2)-dependecies for the BZ specimen under relatively
oxidizing conditions at 500, 600 and 700°C. The conductivity increases with
increasing oxygen partial pressure.

As demonstrated in Fig. 9 a), the impedance spectra of hot pressed YBZ-1


(1500°C, 1 h, 25 MPa), with >88% density and average grain size ~0.4 µm,
consisted of two distinguishable semicircles, with values of the capacitance
typical representing grain boundaries. Hence, two sets of grain boundary
conductivities have been reported, in Fig. 9 b), where the temperature
dependence (from 200 to 400°C) for bulk and grain boundary conductivities
of YBZ-1 is presented. The reason for this behaviour is not yet clear and
needs further investigation. The bulk semicircle was not observed in the
YBZ-1 spectra presumably due to instrument limitations. The spectra were,
consequently, fitted to the equivalent circuit Rb(Rgb1Qgb1)(Rgb2Qgb2) where
Rb corresponds to the apex of the first grain boundary semi-circle in the
Nyquist plot. The specific grain boundary conductivities in Fig. 9 b) have
been calculated as for the BZ sample using the Brick Layer Model.
However, in the absence of bulk semi-circles for the YBZ-1 spectra, we
have assumed an average empirical value for the bulk capacitance based on
data from the low temperature spectra of the YBZ-1 specimen where only
the bulk semi-circle is visible. This value was used as input to calculate
specific grain boundary conductivities for YBZ-1.

119
a) Temperature (°C)
800 700 600 500
10-3

10-4

Conductivity (S/cm)
10-5

10-6

10-7 σBulk(wet O2)


σGb(wet O2)
10-8
σBulk(dry O2)
σGb(dry O2)
10-9
1.0 1.1 1.2 1.3 1.4

-1
1000/T (K )

b)
500°C
600°C
700°C
Bulk conductivity (s/cm)

1e-4

1e-5

10-5 10-4 10-3 10-2 10-1 100

Oxygen partial pressure (atm)

Fig. 8 a) Conductivity as a function of inverse temperature for BZ in


wet and dry O2. The specific conductivity of the grain boundary
was calculated applying the Brick Layer Model.
b) pO2-dependence for BZ, bulk, at 500, 600 and 700°C.

120
a)
YBZ-1, 400°C, wet O2
80000

60000

-Z (Ohm) 40000

20000

0
0 20000 40000 60000 80000

R (Ohm)

b) Temperature (°C)
400 300 200

10-2
Apparent conductivity (S/cm)

10-3

10-4

10-5

10-6

10-7 σBulk
σGb-1
10-8
σGb-2
10-9
1.4 1.6 1.8 2.0 2.2
-1
1000/T (K )

Fig. 9 a) Impedance spectrum for YBZ-1 at 400°C in wet O2. The


spectrum shows two grain boundary semicircles.
b) Apparent conductivities as a function of inverse temperature
for YBZ-1 in wet O2. The apparent specific grain boundary
conductivities were calculated applying the Brick Layer Model.

121
Discussion
Powder characteristics and densification
The TG curves in Fig. 1 demonstrated that BaCO3 detected in as-
synthesized powders by XRD (Fig. 2) was hard to remove completely, as
weight loss was observed even above 1000°C. Calcination at 1000°C for 48
hrs resulted in apparently phase pure BZ (Fig. 2 a)) and YBZ-1 (Fig. 2 b))
powders, however, carbonates were detected by IR even after calcination at
1100°C in nitrogen (as seen for YBZ-1 in Fig. 4). For YBZ-2 powder,
BaCO3 along with ZrO2, was detected by XRD after calcination at 1000°C
for 48 hrs (Fig. 2 b)). The presintering step observed for YBZ-2 (Fig. 5) can
be explained by the presence and decomposition of BaCO3. YBZ-1 and
YBZ-2 showed considerably poorer sinterability compared to BaZrO3 (BZ).
This could indicate a slight deviation in the stoichiometric composition of
the BaZr0.9Y0.1O2.95 powders. This is the case for YBZ-2, where the
secondary phase of ZrO2 becomes more apparent after hot pressing. A
general explanation for the reduced sinterability can also be the substitution
of yttrium for zirconium. As Y3+ is larger than Zr4+ (0.90 versus 0.72 Å [34])
reduced sinterability will be observed if the Zr/Y-site cation diffusion is rate
limiting.

Conventional sintering of BaZrO3 at 1600°C resulted in density <92%


independent of sintering time. This is in agreement with the literature
[7,11,12,15,16]
stating that T >1700°C is needed in order to obtain >95% dense
BaZrO3 specimens. Sintering at 1600°C also involved severe grain growth
giving average grain size ≥16 µm. Due to the high mobility of grain
boundaries pores became entrapment in grains, as seen in Fig. 7 b).
Entrapped pores have also been reported for BaZrO3 prepared from fine
grained powders sintered at 1450°C [19]. Decomposition of traces of BaCO3
may also contribute to pore growth and increased porosity. In the presented
work, HP has been shown to effectively suppress the grain growth observed
for conventional sintering. The reduced sintering temperature in addition to
applied uniaxial pressure reduces the mobility of grain boundaries, and
dense (>96%) materials with uniform microstructures and average grain size
in the range of 0.42 to 1.0 µm have been obtained by HP. Close to 98%
density was obtained for both BaZrO3 and BaZr0.9Y0.1O2.95 after HP (1h, 50
MPa) at 1650 and 1750°C, respectively. Based on these results, HP is
clearly superior to conventional sintering when it comes to densification of
BaZrO3-based materials.

122
Chemical stability
It should be remarked that the high temperatures used in particular during
HP (up to 1750°C) have been reported to give evaporation of BaO(g), as
shown by Eq. 2. In the HP experiments the system is fairly closed and the
atmosphere highly reducing. The reaction between BaZrO3 and the carbon
pressing dies resulting in evaporation of Ba(g), according to Eq. 3, is
therefore more significant than the evaporation of BaO(g).

BaZrO3(s) + C(s) = ZrO2(s) + CO(g) + Ba(g) (3)

In Fig. 10 the equilibrium constants for the reactions in Eq. 2 and 3 are
plotted as a function of temperature. Fig. 9 clearly demonstrates how the
reaction described by Eq. 3 will dominate at temperatures above ~1250°C
when carbon is present. The equilibrium constant for the reaction in Eq. 3
equals 10-6 below 1600°C. This implies that the partial pressure of Ba(g)
may exceed 10-3 bar when the HP temperature is increased beyond 1600°C,
giving significant loss of Ba.

If Ba(g) has diffused out of the pressing die one would expect to observe
ZrO2(s) in the material, particularly close to the surface and at grain
boundaries. The latter is possibly what was observed for hot pressed BZ
and specially YBZ-2 specimens. As all specimens were polished after
densification one would expect possible ZrO2(s) on the surface to disappear.
In the case of YBZ-2 hot pressed at T > 1450°C, ZrO2(s) was detected even
in bulk. For YBZ-2 there are indications that the starting powder differ
slightly from optimal stoichiometry, resulting in ZrO2 rich materials after
hot pressing. No obvious observation of secondary phases were observed
by XRD of polished pellets of both BZ and YBZ-1 hot pressed at 1500°C,
however the surface of the BZ specimen contained ZrO2(s) according to
XRD. Density larger than 96% can be obtained for both BaZrO3 and
BaZr0.9Y0.1O2.95 by HP at 1500 and 1550°C, respectively and these
temperatures might be sufficiently low to obtain stoichiometric, phase pure
materials. For full densification (~98% density) higher temperatures are
needed (1650 -1750°C) and evaporation of BaO(g) and Ba(g) is expected to
take place. Loss of Ba(g) or BaO(g) during sintering may result in ZrO2 at
grain boundaries, which will explain the grain boundary resistance
discussed further below.

123
Temperature (°C)
900 1100 1300 1500 1700

-5

-10
log K

-15

-20
BaZrO3(s) + C(s) = ZrO2(s) + CO(g) + Ba(g)
BaZrO3(s) = ZrO2(s) + BaO(g)

1000 1200 1400 1600 1800 2000

Temperature (K)

Fig. 10 Equilibrium constants, K, for decomposition of BaZrO3 plotted


as a function of temperature. For the reaction in normal
atmosphere the equilibrium constant equals the BaO partial
pressure (K = p(BaO)) while for the reaction in reducing
atmosphere the constant is a product of the Ba- and the CO
partial pressure (K = p(Ba)·p(CO)).

Electrical properties
The oxygen partial pressure dependence of the bulk conductivity for the BZ
specimen can be interpreted as consisting of two different contributions. At
high temperatures and under oxidizing conditions the conductivity increases
with increasing oxygen partial pressure, generally concluded to reflect p-
type electronic conductivity. Towards lower oxygen partial pressures, the
conductivity becomes independent of the oxygen partial pressure which
either reflects that an aliovalent impurity determines the concentration of the
charge carrier or that intrinsic disorder dominates the defect structure. Since
the conductivity is higher under dry than under wet conditions and that the
difference increases with increasing temperature, one may conclude that
electron holes are the predominating charge carrier under high-temperature
oxidizing conditions. It is however not straight-forward to decide which
defect is responsible for the ionic contribution. On basis of the protonic
dominance reported in the literature for acceptor substituted BZ and the fact
that water decreases the p-type contribution, as shown in Fig 8 a), protons

124
clearly influence the defect structure. Whether these protons are charge-
compensated by an effectively negative impurity (acceptor), frozen in metal
vacancies or in thermal equilibrium with metal vacancies cannot be
established. One may note, though, that the apparent activation energy for
the bulk conductivity in the low temperature region is relatively high (~100
kJ/mol) as compared to the YBZ-1 (~30 kJ/mol), where the acceptor
substitution clearly is charge compensated by protons.

From the conductivity characteristics of the BZ material one should also


recognize that there is a large difference between the bulk and the grain
boundary conductivity and, furthermore, that essentially the same effects are
observed for the bulk conductivity under wet versus dry conditions. One
may speculate whether this indicates that the same defect situation prevails
in the grain boundaries as in the bulk. The apparent activation energy of the
grain boundary conductivity is, however, considerably higher than for bulk.

The conductivities of BZ and YBZ-1 are compared to the state-of-the-art


values reported for proton conductivity in BaZr0.9Y0.1O3-δ by Bohn and
Schober [15] and Kreuer [17] in Fig. 11. Acceptor substitution clearly has a
tremendous impact increasing the conductivity several orders of magnitude.
The values for the bulk conductivity of YBZ-1 are essentially the same as
those observed by Bohn and Schober [15]. Although it has not been directly
measured we may assume, based on the similar temperature dependence of
the present data and the literature, that the conductivity is predominantly
protonic in the present temperature region (200 to 400°C).

There are different hypotheses in the literature so as to explain the high


grain boundary resistance of acceptor substituted BaZrO3. It has been
suggested that a non-uniform distribution of charge carriers across grain
boundaries may result in high resistance towards the ionic transport [35].

Through-out the discussion above, one must bear in mind that interpretation
of nominally BaZrO3 materials always is associated with high uncertainties
due to the unknown level and type of impurities. In our case, this is of
particular importance since grain boundaries are sensitive in this respect.
Difference in the chemical composition at the grain boundaries due to loss
of Ba(g) / BaO(g) during sintering, may be of importance for understanding
the grain boundary resistance in BaZrO3 ceramics.

125
Temperature (°C)
800 600 400 200

10-2 BZ
YBZ-1
YBZ Bohn
YBZ Kreuer

Conductivity (S/cm)
10-3

10-4

10-5

10-6

1.0 1.5 2.0 2.5 3.0


-1
1000/T (K )

Fig. 11 Bulk conductivity as a function of inverse temperature for BZ


and YBZ-1 in wet O2, obtained in the present investigation,
compared with bulk data for YBZ in wet O2 presented by Bohn
and Schober [35] and data for single crystal YBZ by Kreuer [36].

Conclusions
High quality powders of BaZrO3 and BaZr0.9Y0.1O2.95 with particle size of
~0.1 µm were prepared by spray pyrolysis of nitrate salt solutions followed
by calcination at 1000°C and ball milling in iso-propanol. By conventional
sintering of BaZrO3, a maximum density of 92% was obtained and the
average grain size for these specimens was in the range of 16 to 18 µm.
Both for BaZrO3 and BaZr0.9Y0.1O2.95, a density >96% was obtained by hot
pressing at 1500 and 1550°C, respectively. The HP specimens showed
homogenous microstructures with average grain size down to 0.42 µm.
Density close to 98% was obtained for BaZrO3 and BaZr0.9Y0.1O2.95 after HP
(1h, 50 MPa) at 1650 and 1750°C, respectively. However, at these
temperatures evaporation of BaO(g) and Ba(g) is an issue. The prepared
specimens exhibit electrical properties comparable to what has been
reported in the literature. The conductivity in BaZrO3 is slightly higher in
dry atmosphere compared to wet, and for both compositions the proton
conductivity is clearly limited by high grain boundary resistance.

126
Acknowledgement
Work was supported by the Research Council of Norway, Grant No.
1585171431 (NANOMAT). Øystein Andersen is acknowledged for
operating the spray pyrolysis.

References
[1] H. Iwahara, Y. Asakura, K. Katahira, M. Tanaka, Prospect of
hydrogen technology using proton-conducting ceramics, Solid State
Ionics, 168, 299-310 (2004)
[2] N. Taniguchi, K. Hatoh, J. Niikura, T. Gamo, H. Iwahara, Proton
conductive properties of gadolinium doped barium cerates at high
temperatures, Solid State Ionics, 53-56, 998-1003 (1992)
[3] M.J. Scholten, J. Schoonman, J.C. van Miltenburg, H.A.J. Oonk,
Synthesis of strontium and barium cerate and their reaction with
carbon dioxide, Solid State Ionics, 61, 83-91 (1993)
[4] K.D. Kreuer, On the development of proton conducting materials for
technological applications, Solid State Ionics, 97, 1-15 (1997)
[5] T. Norby, Solid-state protonic conductors: principles, properties,
progress and prospects, Solid State Ionics, 125, 1-10 (1999)
[6] K.H. Ryu, S.M. Haile, Chemical stability and proton conductivity of
doped BaCeO3 – BaZrO3 solid solutions, Solid State Ionics, 125,
355-367 (1999)
[7] K. Katahira, Y. Kohchi, T. Shimura, H. Iwahara, Protonic
conduction in Zr-substituted BaCeO3, Solid State Ionics, 138, 91-98
(2000)
[8] K.D. Kreuer, Proton-conducting oxides, Annu. Rev. Mater. Res., 33,
333-359 (2003)
[9] T. Norby, Y. Larring, Concentration and transport of protons in
oxides, Current Opinion in Solid State & Materials Science, 2, 593-
599 (1997)
[10] R.C.T. Slade, S.D. Flint, N. Singh, Investigation of protonic
conduction in Yb- and Y-doped barium zirconates, Solid State Ionics,
82, 135-141 (1995)
[11] K.D. Kreuer, St. Adams, W. Münch, A. Fuchs, U. Klock, J. Maier,
Proton conducting alkaline eart zirconates and titanates for high
drain electrochemical applications, Solid State Ionics, 145, 295-306
(2001)
[12] T. Schober, H.G. Bohn, Water vapor solubility and electrochemical
characterization of the high temperature proton conductor
BaZr0.9Y0.1O2.95, Solid State Ionics, 127, 351-360 (2000)

127
[13] H. Iwahara, T. Yajima, T. Hibino, K. Ozaki, H. Suzuki, Protonic
conduction in calcium, strontium and barium zirconates, Solid State
Ionics, 61, 65-69 (1993)
[14] W. Wang, A.V. Virkar, Ionic and electron-hole conduction in
BaZr0.93Y0.07O3-δ by 4-probe dc measurements, J. Power Sources,
142, 1-9 (2005)
[15] H.G. Bohn, T. Schober, Electrical conductivity of the high-
temperature proton conductor BaZr0.9Y0.1O2.95, J. Am. Ceram. Soc.,
83 [4], 768-772 (2000)
[16] F.M.M. Snijkers, A. Nuekenhoudt, J. Cooymans, J.J. Luyten, Proton
conductivity and phase composition in BaZr0.9Y0.1O3-δ, Scripta
Materialia, 50, 655-659 (2004)
[17] K.D. Kreuer, Aspects of the formation and mobility of protonic
charge carriers and the stability of perovskite-type oxides, Solid
State Ionics, 125 (1-4), 285-302 (1999)
[18] J. Brzezinska-Miecznik. K. Haberko, M.M. Bucko, Barium zirconate
ceramic powder synthesis by the coprecipitation – calcination
technique, Materials Letters, 56, 273-278 (2002)
[19] F. Boschini, B. Robertz, A. Rulmont, C. Cloots, Preparation of
nanosized barium zirconate powder by thermal decomposition of
urea in an aqueous solution containing barium and zirconium, and
by calcination of the precipitate, J. Eur. Ceram. Soc., 23, 3035-3042
(2003)
[20] A. Magrez, T. Schober, Preparation, sintering and water
incorporation of proton conducting BaZr0.9Y0.1O3-δ: comparison
between three different synthesis techniques, Solid State Ionics, 175,
585-588 (2004)
[21] P.P. Phule, D.C. Grundy, Pathways for the low temperature
synthesis of nano-sized crystalline barium zirconate, Materials
Science and Engineering, B23, 29-35 (1994)
[22] T. Tsuneo, S. Stølen, H. Yokoi, Thermodynamic study of barium
zirconates by mass-spectrometry, J. Nuclear Materials, 209, 174-179
(1994)
[23] T. Tsuneo, Thermodynamic properties of ternary barium oxides,
Thermochimica Acta, 253, 155-165 (1995)
[24] U. Anselmi-Tamburini, M.T. Buscaglia, M. Viviani, M. Bassoli, C.
Bottini, V. Buscaglia, P. Nanni, Z.A. Munis, Solid-state synthesis
and spark plasma sintering of submicron BaYxZr1-xO3-x/2 (x = 0, 0.08
and 0.16) ceramics, J. Eur. Ceram. Soc., 26, 2313-2318 (2006)
[25] Babilo, S.M. Haile, Enhanced sintering of yttrium doped barium
zirconate by addition of ZnO, J. Am. Ceram. Soc., 88 [9], 2362-2368
(2005)

128
[26] C. Kjølseth, R. Haugsrud, T. Norby, Electrical characterization of
undoped and acceptor doped BaZrO3 - effects of grain boundaries,
To be published (2007)
[27] M.I. Mendelson, Average grain size in polycrystalline ceramics, J.
Am. Ceram. Soc., 52 (8), 443-446 (1969)
[28] T. Norby, "Electronic Book", Norwegian Electro Ceramics AS,
www.norecs.com
[29] B.A. Boukamp, Enschede. p. AC-immittance analysis system, EqC.
2003, University of Twente/WisseQ
[30] T. Norby, Electrical measurements, KJM-MEF 4010, Oslo:
University of Oslo (2005)
[31] S. Marion, A.I. Becerro, T. Norby, Ionic and electronic conductivity
in CaTi1-xFexO3-δ (x = 0.1-0.3), Ionics, 5 (5-6), 385-392 (1999)
[32] K. Nakamoto, Infrared and Raman Spectra of Inorganic and
Coordination Compounds, 2nd ed., John Wiley & Sons, Inc., New
York, (1997)
[33] T. Mokkelbost, I. Kaus, T. Grande, M.-A. Einarsrud, Combustion
synthesis and characterization of nanocrystalline CeO2-based
powders, Chem. Mater., 16, 5489-5494 (2004)
[34] R.D. Shannon, Revised effective ionic radii and systematic studies of
interatomic distances in halides and chalcogenides, Acta Cryst.,
A32, 751-767 (1976)
[35] X. Guo. R. Waser, Electrical properties of the grain boundaries of
oxygen ion conductors: Acceptor-doped zirconia and ceria, Progress
in Materials Science, 51 (2), 151-210 (2006)

129
130
PAPER V
132
Oriented LaFeO3 thin films grown on NdGaO3
by spin-coating

Paul Inge Dahl, Tor Grande, Mari-Ann Einarsrud †

Department of Materials Science and Engineering

Norwegian University of Science and Technology

N-7491 Trondheim, Norway

Abstract
LaFeO3 precursor solution was prepared using nitrate salts dissolved in
methanol. Addition of acetic acid and acetyl acetone as chelating agent
resulted in a solution with good spinning properties. Spin-coating was
performed at 2000 rpm for 1 min on (100)- and (110)-oriented single
crystalline NdGaO3 substrates as well as (0001)-oriented single crystal
Al2O3 substrates. Controlled heat treatment up to 400°C with heating rate of
0.1°C/min resulted in homogenous continuous amorphous films on all
substrates. Annealing for 1 h at 500 – 1000°C caused the formation of
oriented polycrystalline thin films on NdGaO3 with both (100)- and (110)-
orientation. Films prepared on sapphire were polycrystalline, randomly
oriented and inhomogenous after annealing at 700°C. The LaFeO3 films
grown on the NdGaO3 substrates crystallize between 400 and 500°C and the
average grain size increased from 40 to 250 nm as the temperature was
increased from 500 to 700°C.

Corresponding author. E-mail: Mari-Ann.Einarsrud@material.ntnu.no

133
Introduction
Many perovskite-type oxide materials, such as lanthanum ferrite (LaFeO3)
exhibit mixed ionic-electronic conductivity [1]. LaFeO3, with orthorhombic
structure, is an antiferromagnetic insulator at room temperature, making the
material feasible for use in magnetic sensors and as read heads in computer
hard drives [2]. At elevated temperatures the material shows significant
conductivity. At 1000°C a linear response for log σ versus log p(O2) have
been reported for LaFeO3 thin films, making the material promising for
oxygen sensor applications [3]. Furthermore, LaFeO3 show excellent
sensitivity towards non flammable gases such as CO [4] and NOx [4-8] and
exhibit selective sensitivity towards flammable gases such as ethanol [9],
methane [4] and volatile sulfides such as CH3SH [10]. LaFeO3 films have also
been investigated for γ-radiation dosemetry purposes [11]. To optimize
response time, thin film sensors are advantageous, and these sensors have
the benefit of needing no reference cell. Additionally the gas sensitivity has
been reported to increase with decreasing grain size [12], making the
morphology an important feature in gas sensing films. In the literature, the
preparation of LaFeO3 films is reported by use of sputtering techniques
[3,5,13-16]
, screen printing of slurries from nanosized powder [6,7,11,17], and
electrochemical reduction [18].

Among the above mentioned techniques, only sputtering has resulted in


epitaxial LaFeO3 thin films. (110)-oriented LaFeO3 films have been grown
on rhombohedral (001)-oriented LaAlO3 substrates [3,13], and (001)-oriented
LaFeO3 films have been prepared on cubic (001)-oriented MgO substrates
[14]
. The crystal structure and lattice parameters of LaFeO3 and a selection
of substrate materials are given in Table 1. NdGaO3 only give a minor
mismatch in the orthorhombic lattice parameters compared to LaFeO3.
Hence, NdGaO3 substrate is a suitable candidate for studying epitaxial
growth of LaFeO3 films. The aim of the present fundamental study has
been to develop a simple sol-gel preparation route for precursor solution of
LaFeO3, and to obtain uniform thin films with defined microstructure via
spin-coating. Further we have investigated the crystallization behaviour of
LaFeO3 films on single crystal NdGaO3 substrates ((100)- and (110)-
oriented). Sapphire substrates, with a noncompatible hexagonal crystal
structure, were chosen for comparable LaFeO3 film growth of random
orientation.

134
[19]
Table 1 Crystal structure and lattice parameters of LaFeO3 and
selected substrate materials [20].

Lattice parameters (Å)


Composition Crystal structure
a b c

LaFeO3 Orthorhombic 5.555 5.566 7.855


NdGaO3 Orthorhombic 5.43 5.50 7.71

Al2O3 (0001) Hexagonal 4.758 - 12.99

Experimental
LaFeO3 precursor solution was prepared by a sol-gel synthesis route
comparable to well accepted routes with metal alkoxides/acetates [21,22],
however modified with nitrate salts as starting agents. Stoichiometric
quantities of La and Fe nitrate salts (La(NO3)3·6H2O and Fe(NO3)2·xH2O,
Merck or Fluka, >99%) were dissolved in methanol (CH3OH, Normapur,
>99.8%) before addition of acetic acid (CH3COOH, Merck, 99.5%)
followed by stirring for 1 h. Finally, acetyl acetone (CH2(COCH3)2, Merck,
>99%) was added as a chelating agent. The addition of acetyl acetone
resulted in an immediate reaction, observed in form of a rapid colour change
in the solution from transparent red-brownish to dark opaque. A molar
cation : acetic acid : acetyl acetone ratio of 1 : 6 : 6 was used and the total
cation concentration in the solution was ∼0.20 M. The solution was
refluxed at 65°C for 1 h. The as-prepared precursor solution was stable for
1 – 2 months after which the wetting properties onto the substrates were
found to deteriorate.

The LaFeO3 precursor solution was deposited onto single crystal


neodymium gallate (NdGaO3, (100)-oriented, MTI Corp. and (110)-
oriented, Crystal & Coating Tech.) and sapphire (Al2O3, (0001)-oriented,
MTI Corp.) substrates by spin-coating (Laurell, WS-400A-6NPP/C-1). The
precursor solution was dispersed on the substrate surface and spun at 2000
rpm for 1 min. The films were dried at 100°C for 5 min or heat treated at
400°C for 1 h with heating and cooling rates of 0.1 and 1°C/min,
respectively. A total of 1 - 5 film layers were deposited onto the substrates,
after which they were annealed (Entech, MF 1/12) at 500 – 1000°C for 1 h
with heating/cooling rate of 1 – 3°C/min. Additionally, a part of the
precursor solution was evaporated forming a thick gel and after drying at
100°C (1 h) a hard gel was obtained, which was easily crushed down to a

135
black powder. This powder self ignited at 135°C in a volatile reaction, to
give a fluffy product which was calcined at 1000°C. The prepared LaFeO3
powder was single phase with orthorhombic crystal structure, indicating
stoichiometric La:Fe ratio in the precursor solution. The calculated lattice
parameters were a = 5.553 Å, b = 5.566 Å and c = 7.859 Å, in agreement
with literature [19].

The crystallisation of the films was studied by X-ray diffraction (XRD)


(Siemens D5000 and Bruker AXS, D8 Advance) with Cu Kα radiation. The
alignment of the substrate and film was performed using rocking curve
scans over the main reflections to record the offset angle. Crystallite size of
crystalline films was estimated by Scherrer’s equation using the full width at
half maximum (FWHM) of the dominating film reflections, and FWHM of
comparable reflections from LaB6 (NIST standard no. 660A [23]) to indicate
the broadening contribution from the instrument. The morphology of
prepared films was studied by atomic force microscopy (AFM) (Digital
Instruments, MMAFM-1) in tapping mode using doped silicon tips. The
average grain size of polycrystalline films with defined grain structure, was
calculated by use of the linear intercept method over at least 50 grains.

Results
X-ray diffractograms of LaFeO3 films grown on NdGaO3 substrates are
shown in Fig. 1, and compared to the diffractogram of powder obtained
from the same precursor solution as used for film preparation. LaFeO3 films
grown on (100)-oriented NdGaO3 (Fig. 1 a) were oriented in the substrate
direction, as only reflections corresponding to (200)/(112) and (400)/(224)
were detected. Assuming the d-values of film reflections correspond to the
(100)-plane distances (d200 and d400), the lattice parameter a, was calculated
to 5.543 Å for the film (annealed at 700°C) compared to a = 5.553 Å for
bulk (powder calcined at 1000°C).

Oriented LaFeO3 films were also grown on NdGaO3 with (110)-orientation


(Fig. 1 b)) and only reflections corresponding to LaFeO3 (110)/(002) and
(220)/(004) were detected. These LaFeO3 films (annealed at 700°C) issued
d-values from which the lattice parameter c = 7.825 Å was calculated,
assuming (hkl) to be (002) and (004) for this set of d-values. In comparison
the corresponding lattice parameter for bulk was c = 7.858 Å. Prepared
LaFeO3 film on (110)-oriented NdGaO3 was amorphous after annealing at
400°C. LaFeO3 film deposited on (100)-oriented NdGaO3 substrate showed
crystallinity after annealing at 500°C. The crystallization temperature can

136
therefore be estimated to be between 400 and 500°C, assuming it is
independent of substrate orientation.

(100)
NdGaO3 (200)
LaFeO3 (200)

(200)
NdGaO3 (400)
LaFeO3 (400)
Relative intensity

LaFeO3 film on (100)-oriented NdGaO3


(112)/(200)

(114)/(222)
(023)/(221)
(220)/(004)

(224)/(400)
(202)/(022)

(204)/(312)
(131)
LaFeO3 powder

25 30 35 40 45 50 55 60 65 70 75

a) 2 θ (°)

NdGaO3 (220)
LaFeO3 (220)
NdGaO3 (110)
LaFeO3 (110)
Relative intensity

LaFeO3 film on (110)-oriented NdGaO3


(202)/(022)
(110)/(002)

(112)/(200)

(220)/(004)

LaFeO3 powder

20 25 30 35 40 45 50

b) 2 θ (°)

Fig. 1 X-ray diffractograms of LaFeO3 films grown on NdGaO3


substrate with a) (100)-orientation and b) (110)-orientation.

137
Both films have been annealed at 700°C. Diffractogram of
powder calcined at 1000°C is shown for comparison.

The evolution of LaFeO3 film crystallization on (100)-oriented NdGaO3


substrates, with increasing annealing temperature, is demonstrated in Fig. 2.
Increased intensity of the reflection at 2θ = 32.3°, corresponding to LaFeO3
(200), was observed with increasing temperature up to 1000°C. The full
width at half maximum (FWHM) decreased from 0.234° to 0.178° over the
temperature range 500 to 1000°C, corresponding to a small increase in
crystallite size from 37 to 51 nm.

1000°C
Relative intensity

700°C

500°C

Substrate

32.0 32.2 32.4 32.6

2 θ (°)

Fig. 2 X-ray diffractograms of LaFeO3 films grown on (100)-oriented


NdGaO3 substrates annealed at various temperatures for 1 h.

The offset angles for the NdGaO3 substrates and corresponding films were
in the range of 0.3 – 0.5°. The difference in offset angle of LaFeO3 film and
NdGaO3 substrate was generally <0.1°, confirming epitaxial growth of the
film. No offset angle was recorded for films grown on sapphire, indicating
random orientation of these films. The random orientation of LaFeO3 films
deposited on sapphire (annealed at 700°C) was confirmed by XRD, as
demonstrated in Fig. 3. All high intensity reflections of LaFeO3 are present

138
in the diffractogram, however, the intensities are weak compared to those
obtained for the oriented films (Fig. 1). Characteristic features from the
sapphire substrate dominate the diffractogram.

LaFeO3 on sapphire 700°C

(220)/(004)
Sapphire substrate

(112)/(200)
Relative intensity

(202)/(022)
(111)

20 25 30 35 40 45 50

2 θ (°)

Fig. 3 X-ray diffratogram of LaFeO3 film on sapphire substrate, after


annealing at 700°C for 1 h. Diffraction profile for uncoated
substrate is added for comparison. The non-linear background
is due to the properties of the detector used.

AFM micrographs of LaFeO3 films grown on (100)-oriented NdGaO3, are


presented in Fig. 4, where Fig. 4 a) shows that fast heating (3°C/min) to
relatively high temperature (1000°C) resulted in films with a rather rough,
porous microstructure. Slow heating (0.1°C/min up to 400°C) resulted in
more continuous fine grained films after annealing at 500°C as shown by
the AFM micrograph in Fig. 4 b). As the slow heating (0.1°C) improved the
homogeneity, it was used for all LaFeO3 films grown on (110)-oriented
NdGaO3 substrates. The morphology of these films are presented in Fig. 5.
The homogenous film morphology obtained by slow heating to 400°C (Fig.
5 a)) was maintained upon further annealing at 700°C (1°C/min) for 1 h
(Fig. 5 b)). A fine grained LaFeO3 film structure (~40 nm) was obtained
after heating to 400°C, as seen in Fig. 5 c), although the film was

139
amorphous. Further annealing at 700°C resulted in a crystalline film with a
larger grain structure (~250 nm) (Fig. 5 d)).

Fig. 4 AFM micrographs of LaFeO3 films on (100)-oriented NdGaO3


substrates after annealing at a) 1000°C with a heating rate of
3°C/min and b) 500°C with a heating rate of 0.1°C/min up to
400°C, then 1°C/min.

The morphology of LaFeO3 films grown on sapphire substrate is presented


in Fig. 6. The film is homogenous after heat treatment at 400°C (Fig. 6 a)),
however, Fig. 6 b) clearly shows that this is not the case after 700°C. The
formation of islands, giving a rough noncontinuous film, confirms that
sapphire was not a suitable substrate for LaFeO3 films. Although the
surface roughness was high, the average grain size obtained for LaFeO3 film
on sapphire annealed at 700°C was the same (~250 nm) as for the film on
(110)-oriented NdGaO3, heat treated the same way. The crystallite size
obtained from XRD was lower than the grain size obtained from AFM. The
trend of increasing crystallite/grain size with increasing temperature was the
same.

140
Fig. 5 AFM micrographs of LaFeO3 film grown on (110)-oriented
NdGaO3 substrate after heat treatment with a heating rate of
0.1°C/min up to 400°C (a) and c)) followed by annealing at
700°C (1°C/min) (b) and d)). The scalebar on the bottom right
indicates height differences of the surfaces in Fig. c) and d).

141
Fig. 6 AFM micrographs of LaFeO3 film grown on (0001)-oriented
sapphire substrate after slow heat treatment up to 400°C
(0.1°C/min) (Fig. a)) and annealing at 700°C for 1 h (1°C/min)
(Fig. b)). Corresponding height scalebars are placed right of
micrographs.

Discussion
The importance of low heating rate, during first heat treatment of the film
layers have clearly been demonstrated by the AFM micrographs in Fig. 4.
The volatile self ignition reaction observed for the dried LaFeO3 sol at
135°C can explain why the fast heat treatment resulted in a rough,
nonhomogenous film surface (Fig. 4 a)). When using slow heating rate the
solvent (with boiling point at 65°C) will be completely evaporated by the
time the ignition temperature is reached. Hence, the reduction of nitrates
may be more controlled as the additional fuel (methanol) no longer is
present.

Epitaxial growth of LaFeO3 thin films, as demonstrated in the present work,


have to our knowledge not been reported by similar methods (only by
sputtering techniques). (110)-oriented LaFeO3 films have been grown on
rhombohedral (001)-oriented LaAlO3 substrates [3,13] and (001)-oriented
LaFeO3 films have been prepared on cubic (001)-oriented MgO substrates
[14]
. Both LaFeO3 and NdGaO3 exhibit an orthorhombic crystal structure
with < 2% mismatch in lattice parameters. LaFeO3 film on NdGaO3
substrate is therefore expected to orient in the same direction as the
substrate. The observed LaFeO3 film reflections can be assigned to two sets
of hkl values. As an example, distance between two corresponding (200)-
planes is close to identical to that for (112)-type planes, hence the d-values
for these two reflections (d200 and d112) will be the same. Although, not

142
evidenced, the LaFeO3 films on the NdGaO3 substrates are likely to orient in
the substrate direction

The lattice parameters for the NdGaO3 substrates are 1-2% smaller, giving
~5% lower unit cell volume, compared to LaFeO3 bulk material.
Consequently it is expected that a LaFeO3 film epitaxially grown on
NdGaO3 will be under compression, which was confirmed by the smaller
lattice parameters (a and c for films on (100)- and (110)-oriented substrates,
respectively) for LaFeO3 films on NdGaO3 substrates compared to those
obtained from bulk. The prepared films were polycrystalline, indicating
growth starting by nucleation at several sites on the substrate surface. The
compressive stress in the film may be alleviated at the grain boundaries in
the polycrystalline film while maintaining the film continuity. If the stress
gets too large, however, cracking of the film will occur. This was not the
case for LaFeO3 grown on NdGaO3 as the prepared polycrystalline films
were all crack free.

In order to be epitaxial, the growth must start at the interface between the
film and the substrate surface. If the growth starts at the film surface
moving down towards the substrate, the growth is unlikely to be epitaxial.
In addition, the surface structure of the substrate must be compatible with a
corresponding surface of the film. As demonstrated, the prepared LaFeO3
films on (0001)-oriented Al2O3 substrates were randomly oriented and
polycrystalline. As the crystal structure of substrate and film did not match
(hexagonal versus orthorhombic) this was expected, however not obvious.
Packing of the atoms of one crystal structure (film) on top of a different
structure (substrate) is possible if the interatomic distances of the
corresponding surfaces match. Furthermore, epitaxial growth is possible if
a repetitive order can be found between the interatomic distances of the film
and the substrate surface. This could not be obtained for LaFeO3 film on
(0001)-oriented sapphire, hence the random orientation of these films. A
more critical aspect of films prepared on sapphire was the formation of
islands, observed when heating from 400 to 700°C. The inhomogenous and
rough film surfaces may be a result of high strain effects in the films due to
non compatibility with the substrate surface. The strain will be introduced
when crystallization starts and this may explain why the formation of
islands was not observed in the amorphous films on sapphire heat treated at
400°C.

143
Conclusion
A simple preparation route for LaFeO3 precursor solution from nitrate salts
dissolved in methanol was obtained using acetyl acetone as a chelating
agent. Single crystal NdGaO3 was proved to be a suitable substrate for
growth of oriented polycrystalline LaFeO3 thin films by spin-coating of the
LaFeO3 precursor solution. By gentle heat treatment to 400°C (0.1°C/min
heating rate) amorphous films with well defined smooth microstructure
were obtained. Annealing of these films at up to 700°C resulted in
continuous, highly oriented polycrystalline films due to epitaxial growth and
average grain size up to 250 nm. Hexagonal sapphire was excluded as a
suitable substrate due to formation of LaFeO3 islands upon heating to
700°C.

Ackowledgement
Dr. Julian Tolchard is acknowledged for assisting the interpretation of XRD
data.

Literature
[1] O. Yamamoto, Y. Takeda, R. Kanno, M. Noda, Perovskite-type
oxides as oxygen electrodes for high temperature oxide fuel cells,
Solid State Ionics, 23, 241-246 (1987)
[2] J.B. Kortright, D.D. Awschalom, J. Stöhr, S.D. Bader, Y.U, Idzerda,
S.S.P. Parkin, I.K. Schuller, H.-C. Siegmann, Journal of Magnetism
and Magnetic Materials, 207, 7-44 (1999)
[3] I. Hole, T. Tybell, J.K. Grepstad, I. Wærnhus, T. Grande, K. Wiik,
High temperature transport kinetics in heteroepitaxial LaFeO3 thin
films, Solid State Electronics, 47, 2279-2282 (2003)
[4] N.N. Toan, S. Saukko, V. Lantto, Gas sensing with semiconducting
perovskite oxide LaFeO3, Physica B: Condensed Matter, 327, 279-
282 (2003)
[5] E. Traversa, S. Matsushima, G. Okada, Y. Sadaoka, Y. Sakai, K.
Watanabe, NO2 sensitive LaFe3 thin films prepared by r.f. sputtering,
Sensors and Actuators B, 25, 661-664 (1995)
[6] M. Carotta, M. Butturi, G. Martinelli, Y. Sadaoka, P. Nunziante, E.
Traversa, Microstructural evolution of nanosized LaFeO3 powders
from the thermal decomposition of a cyano-complex for thick film
gas sensors, Sensors and Actuators B, 44, 590-594 (1997)

144
[7] J. Yoon, M. Grilli, E. Di Bartolomeo, R. Polini, E. Traversa, The
NO2 response of solid electrolyte sensors made using nano-sized
LaFeO3 electrodes, Sensors and Actuators B, 76, 483-488 (2001)
[8] H. Aono, E. Traversa, M. Sakamoto, Y. Sadaoka, Crystallographic
characterization and NO2 gas sensing property of LnFeO3 prepared
by thermal decomposition of Ln---Fe hexacyanocomplexes,
Ln[Fe(CN)6]·nH2O, Ln = La, Nd, Sm, Gd, and Dy, Sensors and
Actuators B, 94, 132-139 (2003)
[9] S. Zhao, J. Sin, B. Xu, M. Zhao, Z. Peng, H. Chai, A high
performance ethanol sensor based on field-effect transistor using a
LaFeO3 nano-crystalline thin-film as a gate electrode, Sensors and
Actuators B, 64, 83-87 (2000)
[10] C. Xiangfeng, P. Siciliano, CH3SH-sensing characteristics of LaFeO3
thick-film prepared by co-precipitation method, Sensors and
Actuators B, 94, 197-200 (2003)
[11] K. Arshak, O. Korostynska, S. Clifford, Screen printed thick films of
NiO and LaFeO3 as gamma radiation sensors, Sensors and
Actuators A, 110, 354-360 (2004)
[12] C. Xu, J. Tamaki, N. Miura, N. Yamazoe, Grain size effects on gas
sensitivity of porous SnO2-based elements, Sensors and Actuators B,
3, 147-155 (1991)
[13] J.K. Grepstad, Y. Takamura, A. Scholl, I. Hole, Y. Suzuki, T.
Tybell, Effects of thermal annealing in oxygen on the
antiferromagnetic order and domain structure of epitaxial LaFeO3
thin films, Thin Solid Films, 486, 108-112 (2005)
[14] Y.-H. Lee, J.-M. Wu, Epitaxial growth of LaFeO3 thin films by RF
magnetron sputtering, J. Crystal Growth, 263, 436-441 (2004)
[15] J.P. Locquet, J. Perret, J. Fompeyrine, E. Mächler, J.W. Seo, G. Van
Tendeloo, Doubling the critical temperature of La1.9Sr0.1CuO4
using epitaxial strain, Nature, 394, 453-456 (1998)
[16] A. Scholl, J. Stöhr, J. Lüning, J.W. Seo, J. Fompeyrine,H. Siegwart,
J.P. Locquet, F. Nolting, S. Anders, E.E. Fullerton, M.R. Scheinfein,
H.A. Padmore, Observation of antiferromagnetic domains in
epitaxial thin films, Science, 287, 1014-1016 (2000)
[17] E. Traversa, Y. Sadaoka, M. Carotta, G. Martinelli, Environmental
monitoring field tests using screen-printed thick-film sensors based
on semiconducting oxides, Sensors and Actuators B, 65, 181-185
(2000)
[18] Y. Mastumoto, J. Hombo, Preparation of LaFeO3 perovskite film
using electrochemical reduction, J. Electroanal. Chem., 348, 441-
445 (1993)
[19] S.E Dann, D.B. Currie, M.T. Weller, M.F. Thomas, A.D. Al-
Rawwas, The Effect of Oxygen Stoichiometry on Phase Relations

145
and Structure in the System La1-xSrxFeO3-δ (0 ≤ x ≤ 1, 0 ≤ δ ≤ 0.5), J.
Solid State Chem., 109, 134-144 (1994)
[20] http://www.mticrystal.com/products_crystal.html
[21] N. Tohge, S. Takahashi, T. Minami, Preparation of PbZrO3-PbTiO3
ferroelectric thin-films by the sol-gel process, J. Am. Ceram. Soc.,
74 [1], 67-71 (1991)
[22] D.M. Tahan, A. Safari, L.C. Klein, Preparation and characterization
of BaxSr1-xTiO3 thin films by a sol-gel technique, J. Am. Ceram.
Soc., 79 [6], 1593-1598 (1996)
[23] http://ts.nist.gov/ts/htdocs/230/232/232.htm

146
PAPER VI
148
Crystallization and surface morphology of
oriented LaCoO3 films prepared by three
different sol-gel routes

Paul Inge Dahl, Hasan Okuyucu*, Julian Tolchard


Tor Grande, Mari-Ann Einarsrud †

Department of Materials Science and Engineering


Norwegian University of Science and Technology
N-7491 Trondheim, Norway

* Present address: Gazi University, Faculty of Technical Education


Besevler, 06500 Ankara, Turkey

Abstract
Precursor solutions for preparation of LaCoO3 and La0.8Ca0.2CoO3 films
were prepared by three different sol-gel routes. The precursor solutions
were synthesized from acetates, alcoxides or nitrate salts of the respective
cations (La, Co and Ca), using methanol or 2-methoxyethanol as solvent and
suitable chelating agents. The solutions were deposited onto single crystal
yttria stabilized zirconia (YSZ) and sapphire substrates by spin- or dip-
coating. Oriented polycrystalline films of both LaCoO3 and La0.8Ca0.2CoO3
were grown on the cubic (100)-oriented YSZ substrates. These films were
(104)-oriented when indexing according to the hexagonal structure,
corresponding to (110)-orientation in the cubic structure, and the orientation
was independent of both type of precursor solution and deposition
technique. Crystallization of the films started between 800 and 900°C. The
most critical influence on surface roughness was the heating rate.
Controlled heating, with heating rate of 0.1°C/min to 400°C resulted in
smooth homogenous surface morphology (Rq = 0.6 nm) of the prepared
films, and smooth surfaces were maintained by heating of these films up to
900°C. Heating to 1000°C was assisted with a severe anisotropic grain
growth giving increased surface roughness. Films prepared on hexagonal
(0001)-oriented sapphire were polycrystalline and randomly oriented.

Corresponding author. E-mail: Mari-Ann.Einarsrud@material.ntnu.no

149
Introduction
Lanthanum cobaltite, LaCoO3, with perovskite-type structure (ABO3),
exhibit interesting magnetic and electronic behaviour with change in
temperature. Depending on the spin-state of the trivalent Co-ions, LaCoO3
goes from being an insulator at low temperatures (T < 35 K), to showing
semiconducting behaviour at first, and then at elevated temperatures (T >
650 K), close to metallic conductivity [1]. The rhombohedral crystal
structure of LaCoO3 is described by hexagonal lattice parameters as listed in
Table 1. At elevated temperatures (above ~1400°C) the material undergoes
a phase transition from rhombohedral to cubic structure [2]. Orientation of
ferroelastic domains, arising from the cubic to rhombohedral phase
transition, gives the LaCoO3-based materials interesting ferroelastic features
[3-5]
. The mentioned properties may be adjusted by substitution with e.g.
calcium, strontium or barium on the A-site [3-6]. Such substitution will lower
the rhombohedral to cubic phase transition temperature, and the materials
become cubic at ambient temperature for ~50 mol% substitution [2].
Furthermore, mixed ionic/electronic conductivity at elevated temperatures,
makes LaCoO3-materials applicable in membrane technology and as
cathode materials in solid oxide fuel cells (SOFCs) [7]. LaCoO3 materials
are also investigated with respect to catalytic properties [8] and for use in gas
sensors [9,10].

The literature reveals thin films of LaCoO3 (pure and Ca-/Sr-substituted)


prepared by different techniques, such as pulsed laser deposition (PLD) [10],
chemical vapour deposition (CVD) [11], atomic layer epitaxy (ALE) [12],
electrochemical oxidation [13], ion beam sputtering [14] and by spray
pyrolysis in inductively coupled plasma (spray-ICP) [15]. LaCoO3 films
have also been prepared from various sol-gel precursor solutions by spin- or
dip-coating [16-25]. Sol-gel precursor solutions have been synthesized by
traditionally accepted routes using metal alkoxides dissolved in alcohol with
proper chelating agents, such as 2-ethylacetoacetate or polyethylene glycol
(PEG) [16-18]. Less expensive variants using nitrate salts or acetates as
starting materials for alcohol-based precursor solutions, with e.g.
butylacetate or polyvinyl alcohol (PVA) as chelating agents, have also been
reported [19-23]. LaCoO3 films have also been prepared from water based
routes using ethylendiaminetetraacetic acid (EDTA) or
diethylenetriamineoentaacetic acid (DTPA) as complexing agents [23-26].
None of these so-called chemical solution deposition (CSD) techniques have
resulted in epitaxial growth of the films. Due to the often superior
properties of crystalline epitaxial films, these are more attractive for
applications and fundamental studies, compared to randomly oriented
polycrystalline films [27].

150
Among the variety of film preparation techniques mentioned above, the
present work puts focus on CSD methods. The aim of the work has been to
obtain knowledge and experience on film preparation, by fundamental
studies of LaCoO3 films prepared by several CSD techniques. Different sol-
gel precursor routes have been investigated and films prepared from the
different routes, by both spin- and dip-coating, have been compared.
Furthermore, the heat treatment of prepared films has been studied, with the
goal of obtaining crystalline films with smooth, homogenous surfaces.

Experimental
Synthesis routes
Precursor solutions for preparation of pure and calcium substituted
lanthanum cobaltite (LaCoO3 and La0.8Ca0.2CoO3) films, noted LC and
LCC in the following, were prepared by three different sol-gel routes. Each
synthesis route presented in the following has been named based on the
starting precursors used: acetate route, alkoxide route and nitrate route.
Literature references to comparable sol-gel routes are given for each of the
presented three. However, all routes have been modified from those
reported in the literature.

Acetate route [21,22,28]

Stoichiometric amounts of acetates of La, Co and Ca [La(C2H2O2)3·1.5H2O,


Co(C2H2O2)2·4H2O and Ca(C2H2O2)2·xH2O, Acros or Alfa Aesar, >98%]
were dissolved in methanol [CH3OH, Normapur, >99.8%]. Trifloroacetic
acid [CF3COOH, Fluka, >98%] and triethanolamine [N(C2H5O)3, Fluka,
99%) were added as chelating agents. The molar ratio of trifluoroacetic
acid/triethanolamine/cations was 6.4/0.30/1.0 and the total cation
concentration was 0.07 M. The prepared solutions were stirred for 12 hrs
and appeared transparent with a bright violet color.

Alkoxide route [16-18]

Lanthanum isopropoxide [La(C3H7O)3, Gelest, >95%], cobalt acetate and


calcium acetate [Co(C2H3O2)2·4H2O, Ca(C2H3O2)2·xH2O, Acros, >99%]
were used as starting materials. Precursor solutions were prepared in N2
atmosphere in order to obtain control over hydrolysis/condensation
reactions of lanthanum isopropoxide. The solvent, 2-methoxyethanol
[C3H9O2, Acros, >99%), was dried over molecular sieves [4 Å, Merck] for
24 hrs and distilled (122 - 124°C) in N2 flow. The alkoxide was dissolved

151
in 2-methoxyethanol and the solution transferred to flasks containing
cobalt/calcium acetates, dried to constant weight at 100°C in vacuum. The
acetates immediately dissolved giving clear violet solutions. After heating
and stirring for 1 h at 60°C, ethylacetoacetate was added as a chelating
agent. The molar ratio of chelating agent/cations was 1 and the total cation
concentration in the prepared solution was ~0.1 M. The solution was
refluxed at 75°C for 2 hrs in nitrogen atmosphere giving a slight colour
change from violet to dark pink at first, and then, to intense dark red (LC-
sol) or brown (LCC-sol). To start a partial hydrolysis and condensation
reaction a small amount of water was added (water/alkoxide molar ratio
equal to 1).

Nitrate route [19-22]


Stoichiometric amounts of nitrate salts of La, Co and Ca [La(NO3)3·6H2O,
Co(NO3)2·6H2O, Ca(NO3)2·4H2O, Fluka, KeboLab or Merck, >99%] were
dissolved in methanol or 2-methoxyethanol. Acetic acid [CH3COOH,
Merck, 99.5%] was added and the solutions stirred for 1 h. Acetyl acetone
[CH2(COCH3)2, Merck, >99%] was added as a chelating agent. A molar
cation/acetic acid/acetyl acetone ratio of 1/6/6 was used and the total cation
concentrations in the solutions were 0.07 or 0.20 M. The solutions were
refluxed at 65°C or 124°C for 1 h.

Film processing
Films were prepared by spin- and dip-coating of the precursor solutions onto
single crystal substrates of (100)-oriented cubic YSZ [ZrO2 with 9.5 mol%
Y2O3, Crystal & Coating Inc.] and (0001)-oriented hexagonal sapphire
[Al2O3, MTI Corp.]. The crystal structure, lattice parameters and thermal
expansion coefficients (TEC) for substrate and film materials are compared
in Table 1. In the dip-coating method, ultrasonically cleaned substrates
were dipped into the precursor solution and manually pulled up into a
vertical furnace preheated to 550°C. In the spin-coating method, the
precursor solution was dispensed at substrates cleaned with methanol and
spun at 2000 rpm for 30 sec [Laurell, WS-400A-6NPP/C-1]. Each film
layer was dried at 100 – 150°C for 5 – 10 min or heat treated at 400°C for 1
h, with a heating rate of 0.1°C/min. A total of 1 – 5 film layers were
deposited. Prepared films were annealed [Entech, MF 1/12] at 800 –
1000°C in ambient air for 20 min – 3 hrs using heating/cooling rates from
0.1 to 3°C/min. Parts of the precursor solutions were evaporated forming
hard gels that could be crushed down to powders. The evaporated nitrate
precursor solution self ignited, in a volatile combustion reaction around

152
135°C, while decomposition of the two other precursor solutions were more
controlled. The synthesized powders were calcined at 1000°C in ambient
air for 1 – 12 hrs, giving single phase powders of rhombohedral crystal
structure (confirmed by X-ray diffraction).

Table 1 Crystal structure, lattice parameters and thermal expansion


coefficients (TEC) for film [2] and substrate materials [29,30].

Lattice TEC
Compound Structure
parameters (Å) (10-6 K-1)

YSZ Cubic a = 5.140 10.3

Al2O3 Hexagonal a = 4.758 7.50


c = 12.99

LaCoO3 Rhombohedral ah = 5.437 22.4


(hexagonal) ch = 13.09

La0.80Ca0.20CoO3 Rhombohedral ah = 5.427 19.3


(hexagonal) ch = 13.09

Characterization
The prepared films and powders were characterized by X-ray diffraction
(XRD) [Siemens D5000] with Cu Kα radiation through a primary
monochromator. The alignment of the substrate and film was performed
using rocking curve scans over the main reflections to record the offset
angle. Additional studies of the film crystallisation was performed in situ
by high temperature X-ray diffraction (HTXRD) [Siemens D5005] using
heating/cooling rates of 2°C/h and 1 h dwell time at each temperature. The
morphology of prepared films was studied by atomic force microscopy
(AFM) [Digital Instruments, MMAFM-1] in tapping mode using doped
silicon tips. The roughness of film surfaces was estimated by calculation of
root mean square roughness (Rms, Rq) from scanned 5 x 5 µm2 areas.
Average grain size of the crystalline films were estimated from the AFM
images using the linear intercept method over 30 – 60 grains.

153
Results
Dip-coated films from acetate solution
LC/LCC films completely covered the YSZ substrate surfaces after dip-
coating and heat treatment at 550°C. The morphology of dip-coated LC
films on YSZ is presented by AFM micrographs in Fig. 1. The as-prepared
films after heat treatment at 550°C showed uniform surfaces (Rq = 0.3 nm),
as indicated by Fig. 1 a). After annealing at 900°C for 20 min, with a
heating rate of 3°C/min (Fig. 1 b)), the surface was still smooth (Rq = 2.7
nm) and the formation of grains had taken place. Annealing at 1000°C for
20 min led to a significant increase in surface roughness (Rq = 22 nm), as
demonstrated by Fig. 1 c). The increase in annealing temperature from 900
to 1000°C was accompanied by grain growth, average grain size increasing
from ~300 nm to ~580 nm.

X-ray diffractograms of the dip-coated films prepared on YSZ substrates


from the acetate solutions are shown in Fig. 2 an X-ray diffractogram of
synthesized powder is included for comparison. Fig. 2 a) shows that both
LC and LCC films annealed at 1000°C are (104)-oriented, as only
reflections corresponding to the (104)- and (208)-planes in the diffractogram
for LC powder, can be observed. The HTXRD diffractograms in Fig. 1 b)
shows that crystallization of the LC film takes place between 850 and
900°C.

Spin-coated films from alkoxide solution


Homogenous films of LC and LCC were obtained by spin-coating of the
alkoxide solutions onto YSZ substrates. AFM micrographs of the films are
shown in Fig. 3. As indicated by Fig. 3 a), fast heating (3°C/min) to 800°C
(3 hrs) results in formation of islands on top of a finer grained underlying
film surface (Fig. 3 b)). Annealing at 1000°C for 3 hrs resulted in complete
domination of islands on the surface, hence the rough morphology shown in
Fig. 3 c). The surface roughness after fast heating to 800 and 1000°C, was
calculated to Rq = 4.4 and 28 nm, respectively. The corresponding increase
in grain/particle size was from ~75 to ~400 nm. Slow heating (0.1°C/min)
to 400°C produced smooth films (Rq = 0.6 nm) as demonstrated by the AFM
micrograph in Fig. 3 d). Annealing of such film at 800 for 1 h, with a
heating rate of 1°C, resulted in a homogenous morphology (Fig. 3 e)) and
surface roughness of 4.0 nm. Further annealing at 1000°C for 1 h, resulted
in island formation and a rather rough surface (Rq = 9.8 nm), as indicated by
Fig. 3 f).

154
Fig. 1 AFM micrographs of LC films prepared by dipcoating in acetate
precursor solution; a) as prepared film (heat treated at 550°C),
b) and c) films annealed at 900°C and 1000°C, respectively, for
20 min with heating rate of 3°C/min.

155
YSZ (200)
YSZ (100)
Relative intensity
La0.8Ca0.2CoO3-δ film

a) LaCoO3-δ film

(214)
(110)
(104)

(202)

(300)
(024)

(018)

(208)
(220)
(122)
(006)

(116)
LaCoO3-δ powder

25 30 35 40 45 50 55 60 65 70 75

2 θ (°)

LaCoO3-δ film
Relative intensity

b) 950°C
900°C
850°C

32.5 33.5 34.5

2 θ (°)

Fig. 2 a) X-ray diffractograms of LC and LCC films, prepared on YSZ


substrates by dipcoating with acetate-sol and annealed at
1000°C for 20 min. The diffractogram of powder obtained by
evaporation of sol and calcination at 1000°C for 12 hrs is added
for comparison. Indexing (hkl) based on hexagonal structure
[31]
. b) X-ray diffractograms of LC film prepared on YSZ
substrates by dipcoating with acetate-sol, obtained by in situ
HTXRD.

156
Fig. 3 AFM micrographs of LCC films obtained by spincoating of
alkoxide precursor solution onto YSZ substrates. a) and b) 5-
layered film Annealed at 800°C for 3 hrs with a heating rate of

157
3°C/min, c) 5-layered film annealed at 1000°C for 3 hrs with a
heating rate of 3°C/min, d) 1-layered film heat treated at 400°C
for 1 h with a heating rate of 0.1°C, e) and f) 1-layered film heat
treated as in d) followed annealing at 800 and 1000°C,
respectively, for 3 h.

The X-ray diffractograms of LC and LCC films in Fig. 4 demonstrate the


evolution of crystallization with increasing temperature from 800 to
1000°C. For both compositions, only one reflection, corresponding to the
(104)-reflection in the powder diffractogram, can be assigned to the films.
This shows that the films of both LC and LCC were textured. Rocking-
curve scans confirmed this by showing similar offset angles for the films
compared to the substrate. As an example the offset angle for a LCC film
annealed at 1000°C for 3 hrs was measured to 0.41°, compared to 0.44° for
the substrate. Some smaller reflections (that may be assigned to Co3O4)
could be detected, indicating a slight deviation from stoichiometric
composition of the precursor solutions.

Spin- and dip-coated films from nitrate solutions


Due to poor wetting properties, films on YSZ substrates where not obtained
by spin-coating of the nitrate-methanol precursor solution. This effect may
be due to the different acidity of the substrate materials (zirconia being
acidic, alumina having amphoteric properties). Solutions prepared without
acetic acid or with 2-methoxyethanol as solvent did not improve the wetting
on the YSZ substrates. Films were however successfully prepared on YSZ
substrates by dip-coating and rapid heat treatment. The AFM micrographs
in Fig. 5 a) and b) shows the surface morphology of dip-coated films from
nitrate precursor solution after annealing at 800 and 1000°C, respectively,
for 1 h with a heating rate of 3°C/min. The corresponding surface
roughness (Rq) actually decreased from 17 to 11 nm, however the
homogenous grain structure (~140 nm) obtained after annealing at 800°C
was lost when increasing the temperature to 1000°C. The X-ray
diffractograms of dip-coated film annealed at 800 and 1000°C, given in Fig.
5 c), are equivalent to those in Fig. 2 and 4, showing (112)-orientation of the
films.

Homogenous LC films were obtained by spin-coating of methanol based


precursor solution onto (0001) sapphire substrates, as demonstrated by the
AFM micrograph in Fig. 6 a). Surface roughness and average grain size for
such film, annealed at 900°C, was 2.8 nm and ~220 nm, respectively. XRD

158
of this film (Fig. 6 b)) reveals reflections assigned to several crystal planes
in LaCoO3, indicating random orientation. The intensity of these reflections
was low, compared to those obtained for textured films, which is reasonable
as all crystal planes are not oriented in the same direction.

Estimated d-values for (104)-reflections selected of LC / LCC powders and


films on (100) YSZ substrates, prepared by the three different techniques
and annealed at 1000°C, are listed in Table 2. As seen from Table 2, all
film d-values match well with those for powders, indicating no significant
lattice distortion perpendicular to the (104)-planes.

YSZ (100)

La0.8Ca0.2CoO3 films
1000°C
Relative intensity

800°C

LaCoO3 films
1000°C
900°C
800°C
(104)
(110)

(202)
(012)

(024)
(006)

1000°C
LaCoO3 powder

20 25 30 35 40 45 50

2 θ (°)

Fig. 4 X-ray diffractograms of 5-layered LC and LCC films, prepared


by spincoating of alkoxide-sol onto YSZ substrates and annealed
at indicated temperatures for 3 hrs. The diffractograms of
powder obtained by evaporation of sol and calcination at
1000°C for 12 hrs is added for comparison. Hexagonal
indexing (hkl) [31].

159
YSZ (200)
YSZ (100)
Relative intensity

Film 1000°C

c) Film 800°C
(110)
(104)

(214)
(202)

(024)

(300)
(018)

(208)
(006)

(220)
(122)
(116)

Powder 1000°C

25 30 35 40 45 50 55 60 65 70 75

2 θ (°)

Fig. 5 AFM micrographs of LC films on YSZ substrate dipcoated from


nitrate precursor solution and annealed for 1 h at a) 800 and b)
1000°C, with a heating rate of 3°C/min. c) Corresponding X-
ray diffractograms with hexagonal indexing (hkl) [31].

160
a)

b)
Relative logarithmic intensity

Substrate
(024)

(214)

LaCoO3 film
(202)
(104)
(110)
(012)

(006)

(300)
(116)
(122)

(018)

(220)

LaCoO3 powder

20 30 40 50 60 70

2 θ (°)

Fig. 6 a) AFM micrograph of 5-layered LCfilm on (0001)-oriented


sapphire substrate annealed at 900°C for 1 h. b) X-ray
diffractograms of (0001)-oriented sapphire substrate with and
without LC film annealed at 900°C for 1 h. Arrows indicate
reflections assigned to the film. Powder diffractograms, with
hexagonal indexing (hkl) [31], is added for comparison.

161
Table 2 Comparison of d-values obtained from XRD on LC/LCC
powders and films, all heat treated at 1000°C.
Precursor
Compound Technique d104 (Å) d208 (Å)
solution
LaCoO3 Acetate Powder 2.6868

LaCoO3 Acetate Dip coating 2.6849 1.3438

La0.80Ca0.20CoO3 Acetate Powder 2.6802

La0.80Ca0.20CoO3 Acetate Dip coating 2.6801 1.3409

LaCoO3 Alkoxide Powder 2.6862

LaCoO3 Alkoxide Spin coating 2.6892

La0.80Ca0.20CoO3 Alkoxide Powder 2.6804 1.3424

La0.80Ca0.20CoO3 Alkoxide Spin coating 2.6869 1.3447

LaCoO3 Nitrate Powder 2.6858 1.3466

LaCoO3 Nitrate Dip coating 2.6821 1.3453

Discussion
Film orientation
Films of LC and LCC grown on (100)-oriented YSZ showed preferential
growth, independent of precursor solution and deposition technique. The
observed (104)-orientation (hexagonal indexing, equivalent to rhombohedral
(112)) have also been reported for La0.8Sr0.2CoO3 films grown on (100)-
oriented YSZ by sputtering techniques [32], however no explanation for this
growth orientation was given.

The (104)-planes in the hexagonal structure of the LaCoO3 correspond to


the cubic (110)-planes, as obtained from the transformation matrices
presented in Eq. (1). Subscript indexes c and h notes cubic and (h k l)
values, respectively.

162
2 1 1
3 3 3
(hc kc l c ) = (hh kh l h ) 13 1 1 (1)
1 3 3
6 1 1
 6 6

Considering the correlation between the hexagonal (104) and the cubic
(110)-planes may assist the understanding of how LaCoO3 films may orient
in this direction on top of cubic (100)-oriented YSZ, despite the different
crystal structures. However, in order to explain such orientation of the film
the interatomic distances in the (104)-plane of LaCoO3 must be compared
with those of the YSZ (100)-plane. In Fig. 7, this is done by comparing the
(100)-surface of cubic YSZ with the (104)-surface of LaCoO3 (Fig. 7 a) and
b), respectively). The interatomic O – O / La – La distance in the LaCoO3
ccp in the horizontal direction of the illustrated (104)-plane is 3.87 Å (Fig. 7
b)). Compared to the interatomic O – O distance of 2.57 Å in YSZ (Fig. 7
a)), this makes for close to perfect repetitive stacking in this direction, as 3
O- or La-atoms in LaCoO3 fits on top of 4 octahedral holes on the YSZ
surface, with a mismatch of < 0.4%. In the vertical direction of the
illustrated crystal planes (Fig. 7) the interatomic O – La distances in
LaCoO3 are 2.46 and 3.04 Å, the latter giving a more significant mismatch
(~18%) with the O – O distance in YSZ. This mismatch may be reduced by
a local compression of the film, in this direction, at the interface between the
film and the substrate.

As the illustrated vertical interatomic distances in the film are larger than
those of the underlying substrate, an ordered the film is expected to be under
compressive stress. Some of the stress introduced from lattice mismatch
may have been relieved at the grain boundaries, and in the case of fast
heated films assisted by severe grain growth/island formation. Additionally,
the compressive stress from lattice mismatch may to some extent be
extinguished by tensile stress introduced during cooling of the films, due to
the higher TEC of LC/LCC compared to YSZ (Table 1). In Fig. 7 c), the
two planes, illustrated in Fig. 7 a) and b), are overlaid, illustrating a possible
way for oriented growth to occur. A similar match to the (0001)-surface in
Al2O3 was not feasible, hence orientated growth did not occur in films
grown on sapphire substrates.

163
Fig. 7 Surface structures of a) (001)-plane YSZ [29], b) (104)-plane
LaCoO3 [2] and c) the latter structure on top of the first one,
suggesting a possible way of oriented growth.

Surface morphology
The surface morphology of the films has been a central part in this study.
The prepared films were polycrystalline, indicating that growth has started
by nucleation at several sites on the substrate surface. Both the heating rate
and film preparation technique are important for obtaining uniform films
with a smooth surface (here defined by Rq < 5 nm). It was demonstrated in
Fig.1 how smooth films on YSZ were obtained from the acetate precursor
solution by dip-coating and annealing up to 900°C. The heating rate was
not critical for these films as the dip-coated films were directly pulled up

164
into a furnace preheated at 550°C. An increase in the annealing
temperature, from 900 to 1000°C (Fig. 1 b) and c), respectively) resulted in
a significant roughening of the surface (Rq = 2.7 and 23 nm, respectively).
A possible explanation for this phenomenon is anisotropic growth of the
grains in a preferred direction (perpendicular to the (104)-plane). The film
crystallization was shown to start between 850 and 900°C, so after
annealing at 900°C for only 20 min the film may not have been fully
crystalline. Dip-coated films made from the nitrate precursor solution
showed rougher surfaces compared the film obtained from acetate precursor
annealed at 900°C (Fig. 1 b), Rq = 2.7 nm), even after annealing at 800°C
(Fig. 5 a), Rq = 18 nm). As the temperature was 100°C lower, the longer
annealing time (1 h versus 20 min) can not explain this difference. More
likely the rough surface in the dip-coated films from nitrate precursor
solution is due to the vigorous combustion reaction observed at 135°C for
the precursor. By pulling the dip-coated films directly into a furnace at
550°C, this combustion reaction was allowed to happen uncontrolled, hence
causing a rough surface morphology before crystallization.

Controlled heating (0.1°C/min up to 400°C followed by 1°C/min to 900°C)


of a film prepared from the nitrate sol by spin-coating onto sapphire,
resulted in a homogenous film with a smooth surface (Fig. 6 a)). For fast
heating (3°C/min) of films spin-coated from the alkoxide precursor solution,
the roughening of the surfaces were found to start around 800°C, and further
increase of annealing temperature gave a severe anisotropic (104)-oriented
grain growth. Again, controlled heating (0.1°C/min) gave smooth films
after heating up to 400°C (Fig. 3 d)), followed by annealing for 1 h at
800°C, with a heating rate of 1°C/min (Fig. 3 e)). As for the dip-coated
films, increase in the annealing temperature to 1000°C resulted in a
significant grain growth and roughening of the surface.

Based on these results we conclude that controlled heating up to 400°C


improved the film quality. This temperature range corresponds to where the
organics (solvent, chelating agents) and nitrates decompose, hence the
controlled heating in this range is critical. Uncontrolled decomposition of
these species may give rough surface morphology.

Conclusion
Highly oriented LaCoO3-based films can be grown on (100)-oriented YSZ
substrates by both spin- and dip-coating, independent of preparation route
for precursor solution. The observed (104)-orientation of these films
includes some lattice mismatch, resulting in compressive stress on the

165
interface between film and substrate. As the films are polycrystalline some
of this stress may be relieved on the grain boundaries, and possibly to some
extent be extinguished by tensile stress introduced during cooling of the
LaCoO3 films, due to the higher TEC values compared to that of the YSZ
substrates. Controlled decomposition of organics and nitrates in the
precursor solutions, performed by slow heating (0.1°C/min) up to 400°C,
followed by annealing at 800 – 900°C, seemed to be the key to obtaining
homogenous films with smooth surfaces (Rq < 5 nm). These annealing
temperatures should not be exceeded, as the surface roughness increased
quite dramatically when the annealing temperature was increased to 1000°C
(Rq up to 28 nm).

Acknowledgement
Laura Bertolo is acknowledged for experimental work done on LC films
prepared by the alkoxide route and Dr. Mohan Menon for his contribution
on heat treatment of these.

Literature
[1] M.A. Señarís-Rodríguez, J.B. Goodenough, Magnetic and transport
properties of the system La1-xSrxCoO3-δ (0 < x ≤ 0.50), J. Solid State
Chem., 118, 323-336 (1995)
[2] J. Mastin, M.-A. Einarsrud, T. Grande, Crystal structure and thermal
properties of La1-xCaxCoO3-delta (0 <= x <= 0.4), Chem. Mater.,
18, 1680-1687 (2006)
[3] K. Kleveland, N. Orlovskaya, T. Grande, A.M.M. Moe, M.-A.
Einarsrud, Ferroelastic behaviour of LaCoO3-based ceramics, J.
Am. Ceram. Soc., 84, 2029-2033 (2001)
[4] S. Faaland, P.E. Wullum, R. Holmestrand, T. Grande, M.-A.
Einarsrud, Stress-strain behaviour during compression of
polycrystalline La1-xCaxCoO3-based ceramics, J. Am. Ceram. Soc.,
88 [3], 726-730 (2005)
[5] J. Mastin, H.L. Lein, T. Grande, M.-A. Einarsrud, Mechanical
properties of LaCoO3-based materials, to be published
[6] M.A. Señarís-Rodríguez, M.P. Breijo, S. Castro, C. Rey, M.
Sanchez, R.D. Sanchez, J. Mira, A. Fondado, J. Rivas, Peculiarities
in the electrical and magnetic properties of cobalt perovskites Ln1-

166
xMxCoO3 (Ln3+ : La3+, M2+ : Ca2+ , Sr2+ , Ba2+ ; Ln3+ : Nd3+, M2+ :
2+
Sr ), Int. J. Inorg. Mater., 1, 281-287 (1999)
[7] V.V. Kharton, E.N. Naumovich, A.V. Kovalevsky, A.P. Viskup,
F.M. Figueiredo, I.A. Bashmakov, F.M.B Marques, Mixed electronic
and ionic conductivity of LaCo(M)O3 (M=Ga, Cr, Fe or Ni), Solid
State Ionics, 138, 135-148 (2000)
[8] R.N. Singh, B. Lal, High surface area lanthanum cobaltate and its A
and B sites substituted derivatives for electrocatalysis of O2
evolution in alkaline solution, Int. J. Hydrogen Energy, 27, 45-55
(2002)
[9] E. Brosha, R. Mukundan, D.R. Brown, F. H. Garzon, J.H. Visser, M.
Zanini, Z. Zhou, E.M. Logothetis, CO/HC sensors based on thin
films of LaCoO3 and La0.8Sr0.2CoO3-8 metal oxides, Sensors and
Actuators B, 69, 171-182 (2000)
[10] D.T.V. Anh, W. Olthuis, P. Bergveld, Sensing properties of
perovskite oxide La0.5Sr0.5CoO3-8 obtained by using pulsed laser
deposition, Sensors and Actuators B, 103, 165-168 (2004)
[11] M. Losurdo, A. Sacchetti, P. Capezzuto, G. Bruno, L. Armelao, D.
Batteca, G. Bottaro, A. Gasparotto, C. Maragno, E. Tondello,
Optical and electrical properties of nanostructure LaCoO3 thin
films, Applied Physics Letters, 87, 0601909 (2005)
[12] H. Seim, M. Nieminen, L. Niinistö, H. Fjellvåg, L.-S. Johansson,
Growth of LaCoO3 thin films from β-diketonate precursors, Applied
Surface Science, 112, 243-250 (1997)
[13] Y. Matsumoto, T. Sesaki, J. Hombo, A new preparation method of
LaCoO3 perovskite using electronic oxidation, Inorg. Chem., 31,
738-741 (1992)
[14] T. Hattori, T.Matsui, H. Tsuda, H. Mabuchi, K. Morii, Fabrication
and electric properties of LaCoO3 thin films by ion-beam sputtering,
Thin Solid Films, 388, 183-188 (2001)
[15] H. Ichinose, H. Katsuki, M. Nagano, Deposition of LaMO3(M=Co,
Cr, Al) films by spray pyrolysis in inductively coupled plasma, J.
Crystal Growth, 144, 59-64 (1994)
[16] H.J. Hwang, J. Moon, M. Awano, K. Maeda, Sol-Gel Route to
porous lanthanum cobaltite (LaCoO3) thin films, J. Am. Ceram.
Soc., 83, 2852-2854 (2000)
[17] H.J. Hwang, M. Awano, Preparation of LaCoO3 catalytic thin film
by the sol-gel process and its NO decomposition characteristics, J.
Europ. Ceram. Soc., 21, 2103-2107 (2001)
[18] H.J. Hwang, A. Towata, M. Awano, K. Maeda, Sol-Gel route to
perovskite-type Sr-substituted LaCoO3 thin films and effects of
polyethylene glycol on microstructure evolution, Scripta Mater., 44,
2173-2177 (2001)

167
[19] B. Trummer, O. Fruhwirth, K. Reichmann, M. Holzinger, W. Sitte,
P. Pölt, Preparation and characterisation of LaNixCo1-xO3, J. Europ.
Ceram. Soc., 19, 827-829 (1999)
[20] S. Javorič, G. Dražič, M. Kosec, J. Europ. Ceram. Soc., 21, 1543-
1546 (2001)
[21] E. Bontempi, L. Armelao, D. Barreca, L. Bertolo, G. Bottaro, E.
Pierangelo, L.E. Depero, Structural characterized of sol-gel
lanthanum, cobaltite thin films, Crystal Engineering, 5, 291-298
(2002)
[22] M. Gelfi, E. Bontempi, R. Roberti, L. Armelao, L.E. Depero,
Residual stress analysis of thin films and coatings through XRD2
experiments, Thin Solid Films, 450, 143-147 (2004)
[23] Y. Zhang, Y. Zhu, R. Tan, W. Yao, L. Cao, Influence of PEG
additive and precursor concentration on the preparation of LaCoO3
film with peroskite structure, Thin Solid Films, 388, 160-164 (2001)
[24] Y. Zhu, R. Tan, J. Feng, S. Ji, L. Cao, The reaction and poisoning
mechanism of SO2 and perovskite LaCoO3 film model catalysts,
Applied Catalysis A: General, 209, 71-77 (2001)
[25] L. Hong, X. Chen, Z. Cao, Preparation of a perovskite
La0.2Sr0.8CoO3-x membrane on a porous MgO substrate, J. Europ.
Ceram. Soc., 21, 2207-2215 (2001)
[26] Y. Zhu, R. Tan, T. Yi, S. Ji, X. Ye, L. Cao, Preparation of nanosized
LaCoO3 perovskite oxide using amorphous heteronuclear complex
as a precursor at low temperature, J. Mater. Sci., 35, 5415-5420
(2000)
[27] D.P. Norton, Synthesis and properties of epitaxial electronic oxide
thin-film materials, Mater. Sci. Engin. R, 43, 139-247 (2004)
[28] T. Schuller, M.A. Aegerter, Optical, electrical and structural
properties of sol gel ZnO:Al coatings, Thin Solid Films, 351, 125-
131 (1999)
[29] http://www.coatingandcrystal.com/html/super.html
[30] http://www.mticrystal.com/products_crystal.html
[31] V.V. Sikolenko, E.V. Pomjakushina, S.Y. Istomin, Neutron
diffraction studies of La1-xSrxCoO3 magnetic structure at x = 0.15
and x = 0.30, J. Magnetism and Magnetic Materials, 258-259, 300-
301 (2003)
[32] E.L. Brosha, B.W. Chung, F.H. Garzon, I.D. Raistrick, R.J. Houlton,
M.E. Hawley, In Situ Growth and Characterization of La0.8Sr0.2CoO3
Perovskite Mixed Conductor Films, J. Electrochem. Soc., 142 (5),
1702-1705

168

Você também pode gostar