Vous êtes sur la page 1sur 17

Various synthesis methods of aliovalent-doped ceria and

their electrical properties for intermediate temperature solid


oxide electrolytes
Gihyun Kim
a
, Naesung Lee
a
, Ki-Beum Kim
a
, Byung-Kook Kim
b
, Hyejung Chang
b
,
Song-Ju Song
c
, Jun-Young Park
a,
*
a
HMC & INAME, Green Energy Research Institute, Faculty of Nanotechnology and Advanced Materials Engineering, Sejong University,
Seoul 143-747, Republic of Korea
b
Korea Institute of Science and Technology, Seoul 136-791, Republic of Korea
c
Department of Materials Science and Engineering, Chonnam National University, Gwangju 500-757, Republic of Korea
a r t i c l e i n f o
Article history:
Received 16 August 2012
Received in revised form
6 November 2012
Accepted 8 November 2012
Available online 14 December 2012
Keywords:
Intermediate temperature-solid
oxide electrolytes
Ceria-based materials
Composite electrolytes
Conductivity
Ceramic processing
a b s t r a c t
This article investigates the relationship between ionic conductivity and various process-
ing methods for aliovalent-doped, ceria solid solution particles, as an intermediate
temperature-solid oxide electrolyte to explain the wide range of conductivity values that
have been reported. The effects of doping material and content on the ionic conductivity
are investigated comprehensively in the intermediate temperature range. The chemical
routes such as coprecipitation, combustion, and hydrothermal methods are chosen for the
synthesis of ceria-based nanopowders, including the conventional solid-state method. The
ionic conductivity for the ceria-based electrolytes depends strongly on the lattice param-
eter (by dopant type and content), processing parameters (particle size, sintering temper-
ature and microstructure), and operating temperature (defect formation and transport).
Among other doped-ceria systems, the Nd
0.2
Ce
0.8
O
2d
electrolyte synthesized by the
combustion method exhibits the highest ionic conductivity at 600

C. Further, a novel
composite Nd
0.2
Ce
0.8
O
2d
electrolyte consisting of a combination of powders (50:50)
synthesized by coprecipitation and combustion is designed. This electrolyte demonstrates
an ionic conductivity two to four times higher than that of any singly processed
electrolytes.
Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.
1. Introduction
Solid oxide fuel cells (SOFCs) are increasingly recognized as
a next generation clean technology for electrical energy
conversion due to their high energy conversion efciency and
various fuel capabilities. SOFCs can convert hydrocarbon-
based resources including fossil fuels, potentially, biomass
and municipal solid waste to electricity and the electrical
efciency of SOFCs is 45e60% based on the lower heating
value of the fuel [1,2]. Over the past decade, many studies have
attempted to reduce the operational temperatures of SOFCs in
order to increase the durability and electrical efciency, and to
reduce the fabrication cost through wider choices of constit-
uent materials for the interconnector and sealant. In
* Corresponding author. Tel.: 82 2 3408 3848; fax: 82 2 3408 4342.
E-mail address: jyoung@sejong.ac.kr (J.-Y. Park).
Available online at www.sciencedirect.com
j ournal homepage: www. el sevi er. com/ l ocat e/ he
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7
0360-3199/$ e see front matter Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2012.11.044
particular, aliovalent-doped ceria (such as Y
2
O
3
, Gd
2
O
3
, Sm
2
O
3
and Nd
2
O
3
) has received great attention as an intermediate
temperature-solid oxide electrolyte (IT-SOE) due to its supe-
rior ionic conductivity at lower temperatures (500e800

C)
compared to conventional yttria-stabilized zirconia (YSZ)
SOEs under air atmosphere [3e5].
Yahiro et al. [6] examined the total ionic conductivity of the
series (CeO
2
)
0.8
(LnO
1.5
)
0.2
(Ln = La, Nd, Sm, Eu, Gd, Y, Ho, Tm
and Yb) at 800

C. The samaria-doped samples had the highest


conductivity (0.0945 S cm
1
) and lanthania the lowest
(0.0416 S cm
1
). Steele [7] demonstrated that Gd
0.1
Ce
0.9
O
2d
(GDC) exhibited the highest ionic conductivity among the
other doped-ceria systems in the IT range, with an ionic
conductivity of 0.0253 S cm
1
at 600

C in air, in contradiction
with the above results. This low conductivity is related to the
deleterious effects of SiO
2
impurities, which are responsible
for the high grain boundary resistivities that obscure the
intrinsic lattice ionic conductivities for large dopant concen-
trations [8e10]. Mogenson et al. [11] also attributed the
apparent disagreement to the differences in the methods of
sample fabrication, which is in turn dependent on the grain
boundary resistivity.
Eguchi et al. [12] showed that the lattice constant of ceria-
based materials is linearly changed with increasing dopant
content until its solubility limit and that the ionic conductivity
of (CeO
2
)
0.8
(SmO
1.5
)
0.2
is higher than with other rare-earth
Fig. 1 e XRD patterns of aliovalent-doped ceria powders synthesized by (a) solid-state, (b) combustion, (c) coprecipitation,
and (d) hydrothermal methods and (e) comparisons of all synthesis methods of Nd
0.2
Ce
0.8
C
2Ld
powders.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1572
oxide-doped ceria. Further, Kim [13], Hong and Virkar [14] and
Kilner and Brook [15] suggested that the minimal elastic strain
(in case of dopant addition) results in the maximum ionic
conductivity in uorites based on an empirical relationship
between room temperature lattice parameter and dopant ion
radius (and concentration), and that Gd
3
exhibits the highest
ionic conductivity as the ionic radius of Gd
3
lies close to
a critical ionic radius. Recently, Omar et al. [16] proposed and
tested a different dopant strategy based on a density func-
tional theory (DFT) [17] of the interaction energy between
oxygen vacancy and dopant cation. According to the DFT
theory, Sm
3
and Nd
3
exhibited a grain ionic conductivity
14% higher than that of GDC at 550

C in air. Furthermore,
Omar et al. [18] showed that Nd
0.1
Ce
0.9
O
2d
, among doped
ceria, further enhanced the ionic conductivity and that the
grain ionic conductivity was around 17% higher than that of
GDC at 500

C, in air. They also reported that a structureeionic
conductivity relationship based on minimal elastic strain in
the crystal lattice was not sufcient to explain the ionic
conductivity behavior in doped ceria when the processing
variables were kept constant. In addition, in the dilute regime,
the ionic conductivity of aliovalent-doped ceria increases with
increasing dopant concentration. Experimentally, the grain
ionic conductivity reaches a maximum at a dopant concen-
tration of about 10e20 mol% due to defect associations by
electrostatic attraction between the oppositely charged
defects [19,20]. Nonetheless, a wide range of conductivity
values has been reported in various publications for ceria-
based IT-SOEs depending on the type of dopant (e.g. Nd
2
O
3
,
Sm
2
O
3
and Gd
2
O
3
) and its concentration (e.g., 10 vs. 20 wt%)
[19e23]. This deviation is principally due to the processing
variables and impurities within the SOE [7,10,18]. Hence, to
improve the ionic conductivity further, ceria-based SOEs
should be studied systemically and comprehensively by
a variety of synthesis techniques with various dopants and
concentrations, and by dening the dominant causes of these
differences (e.g. contamination and microstructure).
Another scientic approach for enhancing the electrical
conductivity of ceria-based electrolyte is nanostructured
electroceramics for IT-SOEs, known as Nanoionics, due to
their enhanced properties compared with bulk materials
[24e27]. The reduction to the nanoscale has been quite
benecial because the increase in mobile ionic defects in
space charge regions increases the conductivity. An under-
standing of nanostructured materials is dependent on
knowledge of the microstructure (e.g., grain boundaries in
SOEs) of the materials, even though some controversies such
as induced electronic contribution and validities of various
transport models in nanocrystalline doped-ceria SOEs remain
[28e33]. Nanostructured aliovalent-doped ceria solid solu-
tions can be synthesized by various wet chemical processes
such as the combustion [35e38], coprecipitation [39,40] and
hydrothermal [41e43] methods. These processes are capable
of synthesizing homogeneous, high purity and ultrane
powders at lower temperature. Moreover, the nanosized
Fig. 2 e Lattice parameter of Re
0.1
Ce
0.9
C
2Ld
and
Re
0.2
Ce
0.8
C
2Ld
powders synthesized by various ceramic
synthesis methods at 600

C as a function of dopant type
and concentration.
Table 1 e Lattice parameter and crystallite size of
aliovalent-doped ceria powders by XRD analysis.
Materials Processing
method
Lattice
parameter (

A)
Crystallite
size (

A)
10GDC Solid-state 5.402 132.0
Combustion 5.435 54.8
Coprecipitation 5.411 25.3
10SDC Solid-state 5.416 84.1
Combustion 5.452 59.0
Coprecipitation 5.425 47.5
Hydrothermal 5.421 48.2
10NDC Solid-state 5.434 104.7
Combustion 5.469 39.4
Coprecipitation 5.472 25.4
Hydrothermal 5.461 43.6
20GDC Solid-state 5.438 129.5
Combustion 5.455 63.0
Coprecipitation 5.428 25.4
20SDC Solid-state 5.456 86.5
Combustion 5.475 65.5
Coprecipitation 5.444 36.2
Hydrothermal 5.435 79.1
20NDC Solid-state 5.473 123.6
Combustion 5.486 78.9
Coprecipitation 5.483 35.8
Hydrothermal 5.478 45.2
Table 2 e Characterization data of aliovalent-doped ceria
powders by FE-SEM, HR-TEM and BET measurements.
Processing
method
Surface
area (BET, m
2
/g)
Particle
size
(SEM/TEM, nm)
Crystallite
size
(XRD,

A)
Solid-state 2.3e2.9 90e400 90 54.8
Combustion 29.8e36.1 5e8 54 10.3
Coprecipitation 88.4e122.7 3e5 32 9.0
Hydrothermal 42.5e57.1 4e8 54 16.8
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1573
powder is essential to obtain a dense electrolyte at a sintering
temperature which is lower than that where the change in
valency state of cerium occurs.
The primary objective of the present work is to understand
the change of electrical conductivity of different doped-ceria
materials with various nanosized powder processing tech-
niques. Further, through a clear understanding of the
conductivity trend as a function of dopant kind and concen-
tration, and of microstructure and morphology by processing
variables, a novel strategy is suggested to improve the ionic
conductivity of doped-ceria systems in the IT range.
2. Experimental procedure
2.1. Preparation of powder and pellet
Pure powders of Re
0.1
Ce
0.9
O
2d
and Re
0.2
Ce
0.8
O
2d
(where
Re = Gd, Sm and Nd), hereby referred as 10RDC and 20RDC
(where 10 and 20 = dopant concentration and R = G, S and N),
respectively, were synthesized by conventional solid-state
reaction, hydrothermal, coprecipitation, and hydrothermal
methods. Dopant (Re) concentrations of 10 and 20 mol% were
chosen in these systems to investigate the effect of the change
in the lattice parameter between the host and dopant cations
on the ionic conductivity. For the conventional solid-state
synthesis method, CeO
2
and Re
2
O
3
(both with 99.9% purity
from LTS Research Lab, USA) powders were mixed in stoi-
chiometric proportions and ball-milled (zirconia balls) with
ethanol solvent for 24 h. After milling phase, the ceramic
slurries were dried in an oven at 80

C for 10 h and then
calcined at 1200

C for 10 h in air. The agglomerated powders
were pulverized using a mortar and pestle, and subsequently
sieved through a mesh (53 mm) to improve the sinterability of
the powders. The powders were uniaxially pressed at 10 MPa
in a cylindrical mold, and then cold-isostatically pressed at
200 MPa (less than 25 mmin diameter and 2 mmin thickness).
Each pressed pellet was nally sintered at 1650

C for 10 h in
air. The densities of the prepared 10/20GDC, 10/20SDC, and 10/
20NDC pellets were measured by Archimedes technique with
a density meter (MD300S, Alfa Mirage, Japan), using water as
the liquid medium and the direct measurements of geometric
volume with the corresponding mass.
Ceria powders of Re
0.1
Ce
0.9
O
2d
and Re
0.2
Ce
0.8
O
2d
were
also prepared by coprecipitation, combustion and hydro-
thermal methods using Ce(NO
3
)
3
$6H
2
O (99.99% purity, Kanto
Chemical, Japan) and Re(NO
3
)
3
$6H
2
O (Re = Gd, Sm and Nd,
99.9%, Alfa Aesar, USA) as starting precursors. In the
Fig. 3 e HR-TEM and FE-SEM micrographs of Nd
0.2
Ce
0.8
C
2Ld
powders synthesized by (a) solid-state, (b) combustion, (c)
coprecipitation and (d) hydrothermal methods.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1574
coprecipitation route, each nitrate was dissolved in deion-
ized (DI) water individually and then the solutions (0.2 M)
were mixed. Ammonium hydroxide solution (1.0 N Stan-
dardized Solution, Alfa Aesar, USA) was added dropwise to
the cation solution under vigorous stirring until a pH of 9
was attained, because the acidity (pH) in the reaction
medium plays a key role in producing nanoparticles with
good crystallinity and morphology [42,43]. The obtained
solution was left at rest for 2 h and then the precipitate
was separated by centrifugation and washed with DI water.
The precipitate was then calcined at 400

C for 2 h in air. In
the hydrothermal route, on the other hand, the nitrate
solutions (pH 9) were stirred with ammonium hydroxide
solution in DI water. The low viscosity gel was washed with
distilled water and the precipitated material was placed in
an autoclave reactor (ECO solution, Korea). The system was
kept at 180

C for 4 h (ramp rate of 1

C min
1
), after which
the precipitated powder was washed with distilled water.
Finally, the powder was calcined at 400

C for 2 h. In the
combustion process, glycine (NH
2
CH
2
COOH, 99.5% purity,
Alfa Aesar, USA) was added to the mixture of cerium and
metal (Re) nitrate hexahydrate in DI water. The molar ratio
Fig. 4 e Raman spectra of aliovalent-doped ceria powders synthesized by (a) solid-state, (b) combustion, (c) coprecipitation,
and (d) hydrothermal methods and (e) comparisons of all synthesis methods of Nd
0.2
Ce
0.8
C
2Ld
powders.
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1575
of fuel to metal nitrates was 1. The solutions were gradu-
ally heated on a hot plate, evaporated to a pink gel, and
nally changed to whiteeyellow ash with ame. The
products were calcined at 400

C for 2 h to form the uorite
structure. The powders were uniaxially pressed at 10 MPa
in a cylindrical mold and then isostatically pressed at
200 MPa. Each pressed pellet was sintered at 1450, 1550 and
1650

C for 4 h in air. The sintering temperature was
increased until each pellet reached above 95% of the rela-
tive density, as measured by the density meter.
2.2. Powder characterization and electrical
measurements
X-ray diffraction(XRD, D/MAX2500, Rigaku, USA) was used for
phase identication, crystal size, and lattice parameter
calculation of the calcined powders. The XRD analysis was
carried out using Cu-Ka over a 2q range of 20

e100

. Raman
spectroscopy (inVia Raman microscope, RENISHAW, UK) was
used to identify the structural characterization of each doped-
ceria powder. The BrunauereEmmetteTeller (BET) specic
surface area and particle size of the powders were measured
using a Zetatrac particle size analyzer (Microtrac, USA) with
nitrogen at about 77 K. Powder morphology and size, and the
morphological features of the sintered pellets were observed
by high resolution-transmission electron microscopy (HR-
TEM, Titan, FEI, USA) and eld emission-scanning electron
microscopy (FE-SEM, S-4700, Hitachi High tech, Japan). TEM
specimens were prepared by a focused ion beam instrument
(FIB; Helios NanoLab DualBeam, FEI, USA) with a spatial
accuracy of within tenths of a nanometer to obtain the elec-
tron transparency necessary to measure the grain boundary
thickness of the pellet samples and identify the presence of
minor and impurity phases quantitatively with energy
dispersive spectroscopy (EDX).
The linear thermal expansions of the samples were
measured in the temperature range from 100 to 1000

C at
a heating rate of 3

C min
1
with an air purge at a ow rate of
50 cc min
1
. A parallelepiped specimen of dimensions
1 5 10 mm
3
was cut out of the sintered samples by a dia-
mond saw. The one-dimensional displacement in the direc-
tion of the specimen length was measured using an
isothermal dilatometer (Netzch L75 PT1600, Germany). The
detecting rod was pushed against the sample with a load of
0.02 N.
For electrical measurements, Pt paste (Heraeus, USA) was
coated on both sides of the ceria-based electrolyte for use as
electrodes. After painting, the Pt was dried at 125

C for 30 min
followed by sintering at 850

C for 1 h. After painting two
times, Au wires were attached to the electrodes using Pt paint,
dried and sintered under the same conditions as described
above. Electrochemical impedance spectroscopy (EIS)
measurements were monitored with an SP-240 Potentiostat/
Galvannostat/EIS instrument (SP240, BioLogic, France) from
10 mHz to 7 MHz in the temperature range from250

Cto 800

C
over 50

C intervals in air. An equivalent circuit model was
used to separate out the components of the conductivity using
EC-Lab software. The ow rate of dry air was set at 200 sccm.
Conductivity activation energies were determined from
Arrhenius plots of sT vs. 1/T.
3. Results and discussion
3.1. Powder characterizations
The doping effect of the acceptor dopant cations in the ceria
crystal lattice was investigated using XRD. Fig. 1 shows the
XRD results of aliovalent-doped ceria powders synthesized
from the solid-state, combustion, coprecipitation and hydro-
thermal methods. The XRD patterns conrmed the reaction
crystallization into the uorite structure after calcinations.
The XRD prole did not exhibit any other extra peaks of other
phases. Further, the broad peaks of the combustion, copreci-
pitation and hydrothermal powders indicated the nano-
crystalline structure of each synthesized powder (Fig. 1(e)). By
contrast, the powder synthesized by conventional solid-state
method exhibited the development of relatively sharp peaks,
indicating a larger crystallite size (Fig. 1(e)). In addition, the
XRD patterns revealed the negligible inuence of the dopant
type (Gd
2
O
3
, Sm
2
O
3
and Nd
2
O
3
) and concentration (10 and
20 mol%), as shown in Fig. 1(a)e(d).
Fig. 5 e Relative densities of Re
0.1
Ce
0.9
C
2Ld
and
Re
0.2
Ce
0.8
C
2Ld
pellets prepared by various ceramic
synthesis methods as a function of (a) sintering conditions
(Nd
0.2
Ce
0.8
C
2Ld
only) and (b) dopant type and
concentration.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1576
Fig. 6 e Microstructural analysis by HR-TEM and FE-SEM of Nd
0.2
Ce
0.8
C
2Ld
pellets prepared by (a) solid-state (sintered at
1650

C, 10 h), (b) combustion (sintered at 1450

C, 4 h), (c) coprecipitation (sintered at 1650

C, 10 h) and (d) hydrothermal
methods (sintered at 1650

C, 10 h).
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1577
In order to provide further insight into the inuence of
dopants onthe ceria structure, trends in the lattice constant of
the primary uorite ceria phase for samples were analyzed
from the XRD data. Lattice constant uncertainties were esti-
mated from the measurements step size uncertainty of
0.00465

, without considering the uncertainty from instru-


mental systematic error. The unit cell parameter a increased
with increasing dopant content in the investigated substitu-
tion range of 10e20 mol% and atomic radius
(r
3
Gd
= 1:053

A; r
3
Sm
= 1:079

A and r
3
Nd
= 1:109

A), as shown in
Fig. 2 and Table 1. This was attributed to the crystal lattice
expansion of the material with increasing dopant content and
atomic radius, in agreement with the previous results [11].
Further, the nanocrystalline powders synthesized by the
combustion, coprecipitation, and hydrothermal methods
showed a larger lattice parameter than those by the solid-
state technique, except for the 20GDC and 20SDC samples.
The volume change in the lattice of nanoparticles is associ-
ated with the interior atomic structure, surface atoms and
surroundings, dangling bonds and oxygen concentration on
the surface [44,45]. The expansion of the parameter may have
been due to softening of the lattice vibration caused by the
large surface area to volume ratio in the nanoparticles. This
trend has also been reported in nanoscale compound particles
such as MgO and ZrO
2
[46,47]. In addition, the grain size-
induced nonstoichiometry may have induced the larger
lattice parameter in nanosized powder processing, because
the lattice parameter of ceria is strongly sensitive to changes
in the oxygen nonstoichiometry. Further, the relationship
between lattice parameter and conductivity is explained in
the next section.
Table 2 shows the surface area and size distribution of the
ceria-based particles obtained with the different preparation
procedures. The average diameters of 3e5 nm for coprecipi-
tation, 4e8 nm for hydrothermal and 5e8 nm for combustion
methods were accompanied by relatively narrow particle size
distributions. In contrast, the solid-state powder had an
average diameter of 0.09e0.4 mm with a relatively large
particle size distribution, which agrees very well with the
trend evident in the XRD patterns. The order of the particle
size for the doped ceria was coprecipitation < combu
stion = hydrothermal solid-state methods, and this trend
was maintained comparatively regardless of the dopant type
and content. The SEM and TEM image analyses further sup-
ported these particle dimensions, as shown in the particle size
and powder morphology observations presented in Fig. 3. The
TEM images of samples obtained by coprecipitation and
hydrothermal methods show a homogeneous distribution of
spherical nanoparticles, even if aggregates of tenths of
a nanometer are observed, whereas the combustion powders
present a porous fractal consisting of tightly agglomerated
nanocrystallites, typical of the morphology of the
combustion-synthesized ceria-based powders [34e38]. For the
case of combustion route, radical aming ignition produces
irregular morphologic particles, whereas the precipitation
process for the coprecipitation and hydrothermal methods
produces ne and spherical particles. There was, however, no
clear trend of particle size with the dopant concentration and
type due to the wide distribution of particle size. This result is
in good agreement with the XRDresults shown in Table 1. The
Fig. 7 e (a) Thermal expansion curves and (b) thermal expansion coefcients (TECs) of ceria-based SOEs synthesized by the
combustion method in the temperature range of 100e1000

C in air.
Table 3 e Average thermal expansion coefcients (TECs)
derived from thermal expansion data of SOFC materials.
Component Material TEC Reference
Electrolyte Zr
0.92
Y
0.08
O
2d
10 58
CeO
2
11.7 59
Ce
0.9
Gd
0.1
O
2d
(10GDC) 12.3 58
Ce
0.8
Sm
0.2
O
2d
(20SDC) 13 57
Ce
0.9
Gd
0.1
O
2d
(10GDC) 12.6 This work
Ce
0.9
Sm
0.1
O
2d
(10SDC) 12.8 This work
Ce
0.9
Nd
0.1
O
2d
(10NDC) 12.5 This work
Ce
0.8
Nd
0.2
O
2d
(20NDC) 12.6 This work
Cathode La
0.8
Sr
0.2
MnO
3
12.4 60
La
0.6
Sr
0.4
Co
0.2
Fe
0.8
O
3d
12.5 61
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1578
particle surface area of aliovalent-doped ceria powder
synthesized by coprecipitation method was 88.4e122.7 m
2
g
1
,
which was the largest among the prepared powders. All the
microscopy data indicated that the coprecipitation-
synthesized powder has the nest nanoparticles.
Raman spectroscopy was used to study the microstructure
of acceptor dopant cation-based ceria powders. The Raman
spectra of the 20NDC powders synthesized by various
processes are shown in Fig. 4. Aliovalent-doped CeO
2
showed
a strong band at 461 cm
1
, which is typical for the cubic CeO
2
phase [48,49]. There was little evident difference between any
of the samples according to the dopant type and content.
However, the Raman band at 461 cm
1
broadened with
decreasing particle size (powders prepared by
coprecipitation _ combustion < hydrothermal solid-state
methods) in Fig. 5(e). Zhang [50] and Askrabic [51] attributed
this change in the Raman peak to the increasing concentra-
tion of the point defect (oxygen vacancies) due to the lattice
expansion caused by the decreasing particle size. Further, one
additional peak was observed at an intensity of 550e600 cm
1
,
which was attributed to the presence of intrinsic and extrinsic
oxygen vacancies caused by the nonstoichiometry of the ceria
nanopowders and the substitution of Ce
4
ions, respectively
[52]. Gouadec and Colomban [53] showed that oxygen atoms
Fig. 8 e (a) Representative impedance plots of the 20NDC samples synthesized by solid-state method at temperatures of
300

C and 700

C, and, (b) Arrhenius plots of a set of aliovalent-doped ceria electrolytes from EIS measurements in the IT
range. Solid, com, co-pre, and hydro stand for solid-state, combustion, coprecipitation and hydrothermal methods,
respectively. For the detailed explanations, some of the data in Fig. 8(c) are reused in Figs. 9(a) and 11(a).
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1579
associated with this vibration mode are strongly related to the
oxygen sublattice disorder by processes leading to the grain
size-induced nonstoichiometry.
3.2. Pellet characterizations
Fig. 5 shows the measured relative densities of ceria-based
samples according to the sintering condition. The densities
of the pellets were estimated by Archimedes technique with
a density meter, as described in the experimental section.
After sintering at 1450

C for 4 h, the pellet pressed from the
combustion 20NDC powder exhibited a high relative density
(97.3% of the theoretical density), whereas the pellets pressed
from the hydrothermal and coprecipitation 20NDC powders
showed lower relative densities of 77.5% and 63.6%, respec-
tively, despite the uniform shape and size of their nanosized
particles. Moreover, the 20NDC pellets pressed from the
hydrothermal and coprecipitation powders showed a low
relative density of 90.0% and 77.5%, respectively, even after
sintering at 1650

C for 10 h, whereas the solid-state 20NDC
pellet exhibited a high relative density of 96.4% (Fig. 5(b)). The
densities of the prepared 10/20GDC, 10/20SDC, and 10/20NDC
pellets from the combustion (sintered at 1450

C for 4 h) and
solid-state (sintered at 1650

C for 10 h) methods were greater
than 96% of the theoretical density by Archimedes technique.
The lower densities of the pellets made by sintered
coprecipitation followed by the hydrothermal method could
not yet be explained, compared to those of the pellets made by
combustion with similar particle sizes. However, very
different particle morphology may lead to different packing
density in the sample fabrication. Powders obtained by
coprecipitation and hydrothermal methods show a homoge-
neous distribution of spherical nanoparticles, whereas the
combustion powders present a porous spongy, thin ake type
nanocrystallites. This morphology of combustion powders
may have a comparative advantage in producing high packing
density pellets. In addition, nanosized powders synthesized
by coprecipitation and hydrothermal methods are likely to be
agglomerated with higher friction force between particles due
to the high ratio of surface area to volume, compared to that of
microsized particles. This property may decrease the packing
density due to the reduced mobility of the interparticles
during the formation of the green body [54,55]. Further, this
phenomenon produces a number of intergranular pores
during the signicant volume shrinkage of the green compact
by the sintering process. In addition, the capillary effect
during the pressing can prevent evacuation of gas from the
green body since the time required to evacuate the gas
increases with decreasing particle size [56]. This increases the
probability of pore formation within the pellet since the gas is
captured between particles during the sintering of nanosized
particles. The relationship between density and conductivity
of ceria-based samples is explained in the next section.
The average grain sizes and grain boundary thicknesses of
the sintered 20NDC pellets were determined from the FE-SEM
and HR-TEM micrographs, as shown in Fig. 6. The grain sizes
were 9.375 mm(solid-state, sintered at 1650

C for 10 h), 1.55 mm


(coprecipitation, sintered at 1650

C for 10 h), 1.7 mm (hydro-
thermal, sintered at 1650

C for 10 h) and 0.82 mm (combustion,
sintered at 1450

C for 4 h). In addition, the grains in the
solid-state and combustion 20NDC pellets were connected to
eachother witharelativedensityabove96%. Ontheother hand,
the pellets of the coprecipitation and hydrothermal methods
showed intergranular pores with a low relative density after
being sintered at extremely high temperature conditions.
Another important characteristic to be considered for IT-
SOEs is the attainment of a thermal expansion coefcient
Fig. 9 e Arrhenius plots of aliovalent-doped ceria
electrolytes (prepared by combustion method) as functions
of dopant type and concentration: (a) total, (b) grain and (c)
grain boundary conductivity.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1580
(TEC) similar to the SOFC components, since such materials
are operated at high temperatures. Otherwise, the compo-
nents of SOFC can cause cracks and delamination during
heating (operating) cycles due to the thermal stress of ceramic
materials [57]. The TEC values of ceria-based pellets synthe-
sized by the combustion route were determined by dilatom-
etry in the temperature range from 100 to 1000

C and were
compared with the values previously reported in the litera-
ture. The linear thermal expansion of rare earth-doped ceria
(10GDC, 10SDC, 10NDC and 20NDC) in air is shown in Fig. 7.
The expansions of the ceria-based specimens were constant
(0.1%) in the isothermal regions. The normalized thermal
expansion (DL/L
o
) of all samples, where L
o
is the sample length
at room temperature and DL is the difference of the length
between L
T
and L
o
, increased with increasing temperature.
The DL/L
o
for the ceria-based samples was 0.935% at 800

C, as
shown in Fig. 7(a). TEC was obtained from the slope of DL/L
o
,
since the slope was considered to be nearly linear. The
average TEC values of the 10GDC, 10SDC, 10NDC and 20NDC
samples were about 12.6, 12.8, 12.5 and 12.6 10
6
K
1
from
100 to 1000

C, respectively, in agreement with previous
results [57e59]. Currently, high performance La
0.8
Sr
0.2
MnO
3
and La
0.6
Sr
0.4
Co
0.2
Fe
0.8
O
3d
cathodes with TEC values of
12.4 10
6
K
1
[60] and 12.5 10
6
K
1
[61], respectively, are
compatible with ceria-based electrolytes in terms of their TEC
values. The obtained results are summarized in Table 3 and
compared with those of the materials of the SOFC compo-
nents. However, more detailed work is necessary to investi-
gate the relationship between these TEC values and various
processing methods for aliovalent-doped ceria solid solutions.
3.3. Ionic conductivity
The oxygen ionic conductivity for ceria electrolytes was
determined by EIS. The impedance spectra were tted to the
conventional equivalent electronic circuit in series so that
semicircles were generated on the Nyquist plots. The repre-
sentative impedance plots of the 20NDC samples synthesized
by solid-state method at 300

C and 700

C are presented in
Fig. 8(a). In the impedance plots of 300

C, bulk and grain
boundary arcs are distinct, along with rising electrode arcs. In
the order of decreasing frequency, arcs corresponding to the
response of the grain bulk, the grain boundaries, and the
electrodes were recorded and the high and middle frequency
intercepts of the semicircle with the real impedance axis was
taken as the total resistance. In the impedance plots of 700

C,
however, two arcs were observed and the high frequency
intercept of the arc was taken as the total resistance. Fig. 8(b)
shows the Arrhenius plots of total conductivity of the
aliovalent-doped ceria electrolytes measured at 250e800

C.
Fig. 11 e Arrhenius plots of Nd
0.2
Ce
0.8
C
2Ld
electrolytes as
a function of powder synthesis method: (a) total, (b) grain
and (c) grain boundary conductivity.
Fig. 10 e Conductivity values of Re
0.1
Ce
0.9
C
2Ld
and
Re
0.2
Ce
0.8
C
2Ld
electrolytes at 600

C as a function of the
lattice parameter.
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1581
Table 4 e Activation energies and pre-exponential factors of a set of aliovalent-doped ceria SOEs measured by EIS in the IT
range.
Materials Processing Conductivity Activation energy (eV) Pre-exponential (SK/cm)
10GDC Solid-state Total 1.04498 1.814 10
6
Bulk 1.14826 2.668 10
6
Grain boundary 1.08645 1.091 10
2
Combustion Total 0.85381 1.821 10
6
Bulk 1.20202 8.048 10
4
Grain boundary 1.31139 7.720 10
5
Coprecipitation Total 0.78677 5.884 10
5
Bulk 0.96567 9.241 10
7
Grain boundary 0.96479 1.594 10
3
10SDC Solid-state Total e e
Bulk
Grain boundary
Combustion Total 1.08165 3.540 10
6
Bulk e e
Grain boundary e e
Coprecipitation Total 1.01307 2.351 10
6
Bulk 1.16900 8.259 10
8
Grain boundary 2.04390 1.196 10
5
Hydrothermal Total 0.86492 3.914 10
6
Bulk 1.74136 1.800 10
6
Grain boundary 0.96479 1.594 10
4
10NDC Solid-state Total 1.20927 5.266 10
7
Bulk e e
Grain boundary e e
Combustion Total 1.10060 2.691 10
7
Bulk 1.20202 1.306 10
8
Grain boundary 0.90579 7.494 10
1
Coprecipitation Total 1.22593 1.116 10
8
Bulk 1.05714 8.632 10
8
Grain boundary 1.29089 2.172 10
4
Hydrothermal Total 0.9792 4.154 10
6
Bulk 1.13975 2.873 10
8
Grain boundary 0.82225 1.800 10
2
20GDC Solid-state Total 0.91843 4.964 10
5
Bulk e e
Grain boundary e e
Combustion Total 1.05265 2.842 10
7
Bulk 0.91796 7.866 10
6
Grain boundary 1.06992 3.433 10
3
Coprecipitation Total 0.87383 1.652 10
6
Bulk 1.43549 4.538 10
9
Grain boundary 0.56076 3.418 10
4
20SDC Solid-state Total 1.06133 2.254 10
7
Bulk 1.09611 1.815 10
8
Grain boundary 0.90579 7.904 10
2
Combustion Total e e
Bulk
Grain boundary
Coprecipitation Total 1.11941 5.159 10
7
Bulk 1.10623 6.881 10
8
Grain boundary 1.07033 6.127 10
4
Hydrothermal Total 1.00499 1.141 10
7
Bulk 0.76076 3.144 10
5
Grain boundary 1.02940 1.996 10
4
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1582
The activation energy E
a
of total conductivity was determined
fromthe slope in the ln(sT ) vs. 1/T Arrhenius plot, as shownin
Table 3. Firstly, according to the doping content and type,
20 mol% Nd
2
O
3
in ceria solid solutions (synthesized by
combustion powders) was determined to be the best dopant
cation and concentration to maximize the ionic conductivity
at 600

C (Figs. 8(b), 9 and 10). In addition, among the other
aliovalent-doped ceria electrolytes, the 10NDC (combustion
powders) and 10GDC (coprecipitation powders) SOEs showed
higher ionic conductivities at below 550

C, whereas the high
doping concentrations of 20GDC, 20NDC, and 20NDC exhibi-
ted higher conductivities at more than 600

C (Figs. 8(b) and 9).
Fig. 10 compares the ionic conductivity values (at 600

C) of
Re
0.1
Ce
0.9
O
2d
and Re
0.2
Ce
0.8
O
2d
materials (Re = Gd
3
, Sm
3
,
and Nd
3
) synthesized using the same combustion procedure.
The ionic conductivity increased with increasing dopant ionic
radius (r
3
Gd
= 1:053

A; r
3
Sm
= 1:079

A and r
3
Nd
= 1:109

A) and
lattice parameter, which is consistent with the previously
reported results [11,16e18,55]. Andersson et al. [17] predicted
this effect based on the DFT calculations by showing that the
oxygen vacancy preference to reside in the nearest neigh-
boring site of Pm
3
is nearly the same and that Nd
3
lies next
to Pm
3
, which should result in higher ionic conductivity for
ceria. Further, based on the oxygen diffusion coefcient of
Nd
0.23
Ce
0.77
O
2d
reported by Kamiya et al. [62], doping with
Nd
3
is much more effective due to the smaller association of
point defects in Nd
2
O
3
-doped-ceria system, compared to other
aliovalent doping materials.
In order to further explain the cause of the variation in
conductivity behavior according to the doping material and
concentration under identical fabrication method (combus-
tion only), specic analysis of the ionic conductivity must be
undertaken. It is possible to resolve the bulk and grain
boundary contributions individually using a geometrical
method, as specied by previous works [63e65]. In particular,
to evaluate the contribution of grain boundaries to the
conduction, grain boundary conductivity obtained using
sample thickness only cannot provide any insight because the
grain boundary area parallel to the current ow is much
smaller than that of the grain. Hence, the grain boundary
conductivity should be calculated in other ways. Recently,
research on the grain boundary conductivity and micro-
structure relations has been extensively reviewed by Hong
et al. [64], Tian and Chan [65] and Kidner et al. [66] using the
brick layer model or the nano-grain composite model. The
specic grain boundary conductivity determined by geomet-
rical factor is dened as
s
gb
=
t
R
gb
A

d
S
(1)
where R
gb
, A, S, and d are the resistance of the grain boundary,
the electrode area, the grains size, and the grain boundary
layer thickness, respectively. Arrhenius plots of the grain and
grain boundary conductivities of the acceptor-doped ceria
electrolytes in air are shown in Fig. 9(b) and (c). As shown in
those gures, the grain and grain boundary conductivities
increase with increasing dopant ionic radius and content
(10 mol% < 20 mol%), except for Nd
2
O
3
-doped ceria SOEs. In
particular, this effect was more pronounced at the grain
conductivities of acceptor-doped ceria materials, compared to
those of grain boundary conductivities. That is, the effect of
the acceptor dopant cations physical properties, induced by
the change in the doping material and concentration, on the
ionic conductivity is more closely associated with the grain
property of ceria-based SOEs at below 500

C. However, the
inuence of the doping material on the ionic conductivity was
small at high temperatures, as shown in Fig. 9(a). This was
attributed to the temperature dependence of the activation
energy for oxygen diffusion in SOEs. The activation energy for
oxygen diffusion is comprised of the migration enthalpy for
the oxygenion and the association enthalpy of the local defect
structures. At low temperature, with the acceptor dopant-
induced increase in the oxygen vacancy concentration, the
resulting increase in defect structure formation raises the
ionic conductivity, since the migration enthalpy for the
oxygen ion is independent of the dopant concentration.
Table 4 e (continued)
Materials Processing Conductivity Activation energy (eV) Pre-exponential (SK/cm)
20NDC Solid-state Total 1.17971 1.455 10
8
Bulk e e
Grain boundary e e
Combustion Total 0.90132 1.040 10
6
Bulk e e
Grain boundary e e
Coprecipitation Total 1.19631 1.184 10
8
Bulk 0.96439 2.846 10
7
Grain boundary 1.10468 1.473 10
3
Hydrothermal Total 0.97193 8.697 10
6
Bulk 1.06435 5.055 10
8
Grain boundary 0.82920 4.015 10
2
20NDC
(combination,
1450

C, 4 h)
50% Coprecipitation
50% combustion
Total 1.03736 1.227 10
7
Bulk 0.87996 2.601 10
6
Grain boundary 0.88761 4.584 10
2
20NDC (combination,
1550

C, 4 h)
50% Coprecipitation
50% combustion
Total 1.05589 6.374 10
6
Bulk 0.73224 7.888 10
3
Grain boundary 1.29089 2.172 10
4
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1583
Further investigations were conducted only with the
20NDC electrolytes for simplicity, with the highest ionic
conductivity being attained at 600

C. In terms of the ceramic
processing, 20NDC pellets fabricated by the combustion
method showed the highest ionic conductivity in the IT range,
as shown in Fig. 11(a), among the other synthesis methods for
acceptor-doped ceria systems under identical dopant
concentration. In order to explain the dependence of the
conductivity behavior on the synthesis methods in ceria
powder, the ionic conductivity of ceria-based SOEs was
analyzed by the grain and grain boundary contributions
individually. Arrhenius plots of the grain and grain boundary
conductivities of acceptor-doped ceria electrolytes in air are
plotted in Fig. 11(b) and (c). The bulk conductivities of the
20NDC electrolytes were signicantly affected by the pro-
cessing method. The low densication was one of the major
causes of the low grain conductivity of 20NDC pellets fabri-
cated by the hydrothermal and coprecipitation methods, as
shown in Fig. 11(b) [67,68]. The poor sinterability resulted in
a large, poorly conductive grain region and hence with a low
total ion conductivity (see Fig. 6(c) and (d)).
The grain boundary conductivities also showed signicant
differences in the IT range even with the same material. This
result may be because bulk transport properties depend on
the level of densication and composition homogeneity,
whereas grain boundary transports are dependent on the
microstructures (geometry and grain size), impurities, and
space charge phenomena [8e19,69e71]. The grain boundary
conductivity of the combustion 20NDC pellet was four-orders
of magnitude higher than those of the solid-state and hydro-
thermal pellets. The lower grain boundary conductivity of the
solid-state pellet has previously been attributed to the high
temperature sintering condition [27,38,64,65,72]. The parallel
model explains the sintering condition effect on the low
temperature conductivity. As the sintering temperature
increases with time, more impurities (mainly silica phases)
diffuse into the ceria electrolyte at the grain boundary region.
Furthermore, the nite amount of silica phases hinders the
formation of a continuous and uniform phase along the grain
boundary region at lower sintering temperature. The effect of
the SiO
2
content is the main cause of the variation in the
synthesis methods and compositions reported in the litera-
ture for different reported values of maximum total conduc-
tivity [8e11,16e19,24]. Furthermore, charged point defects
accumulated at the grain boundary regionduring the sintering
process and the excess charge induced space charge layers
with segregation of the dopant-substituted defect in the grain
interface at higher sintering temperature [27,69e71]. The
double Schottky barrier formed in the grain boundary regions
is a result of the depletion of oxygen vacancies, thereby
resulting in the degradation of the grain boundary ionic
conductivity [7,24].
In the case of the coprecipitation 20NDC pellet, the low
densication was one of the major causes of low grain
conductivity, as shown in Fig. 11(b) and (c). The poor sinter-
ability resulted in low total ion conductivity due to poorly
connected grain regions (see Fig. 6(c)), as mentioned earlier.
However, the grain boundary conductivity of the coprecipita-
tion electrolyte was higher than that of the hydrothermal and
solid-state electrolytes in the IT range (Fig. 11(c)). The high
grain boundary conductivity of the coprecipitation 20NDC
electrolyte was attributed to the more uniform distribution of
submicron-sized grains through the nanostructured homoge-
neous particles than other processed electrolytes, as shown in
Figs. 3 and 6. Hence, this process was chosen as an advanced
processing method to improve the ionic conductivity of ceria-
based electrolytes and is discussed in more detail in the next
Fig. 12 e Arrhenius plots of composite and singly
processed Nd
0.2
Ce
0.8
C
2Ld
electrolytes: (a) total conductivity,
(b) bulk conductivity and (c) grain boundary conductivity.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1584
section. The activation energies and pre-exponential factors
for ceria-based SOEs are listed in Table 4.
3.4. Advanced electrolytes
As previously indicated, the bulk conductivity of the 20NDC
electrolytepreparedbycoprecipitationmethodwashigher than
that of the hydrothermal and solid-state electrolytes in the IT
range in Fig. 11. The high bulk conductivity of the coprecipita-
tion 20NDC may have been due to a more homogeneous
dopants distribution in the electrolyte through the nano-
structured uniformspherical particles (Figs. 3 and 6). Hence, an
advanced electrolyte consisting of a combination of powders
(50:50) synthesizedbycoprecipitationandcombustionmethods
was fabricated to improve the ionic conductivity of the elec-
trolyte. One strategy to increase the conductivity is to improve
the densication of the coprecipitation pellet having homoge-
neous and uniform distributions of submicron-sized grains by
using combustion powders consisting of porous fractal and
tightly agglomerated nanocrystallites. Two combination pro-
cessed20NDCpellets weresinteredat 1450

Cand1550

Cfor 4h
to compare the ionic conductivity of combustion processed
electrolytes. The relative densities of two combination pro-
cessed20NDCpellets sinteredat 1450

Cand1550

Cwere 93.7%
and 97.1%, respectively.
The conductivity data of the 20NDC composite pellet are
presented in Fig. 12 and compared with the singly processed
samples. The composite pellets showed higher total conduc-
tivities than the singly processed electrolytes at all tempera-
tures. In particular, the grain boundary conductivities of the
20NDC composite electrolytes were several orders of magni-
tude higher than those of the singly processed materials. This
result was attributed to the easy transport of the oxygen ions
in the ceria electrolyte due to the homogeneous and uniform
distributions of submicron-sized grains prepared by copreci-
pitation method. Furthermore, compared to the 20NDC SOE
prepared by the coprecipitation method, the grain boundary
of conductivity of the 20NDCcomposite SOEs was signicantly
improved up to the level of the combustion SOE. This result
was related with the enhanced density of the coprecipitation
pellets. Fig. 13 shows the microstructure of the composite
20NDC pellets sintered at 1450

C and 1550

C. The composite
20NDC electrolyte showed high levels of relative density
(_97%) after sintering at 1550

C for 4 h, thereby resulting in
high ionic conductivity in the IT range. In conclusion, the total
ionic conductivity of the composite electrolyte was
0.0736 0.022 S cm
1
at 600

C, which was 2.5e4 times more
than the values of less than 0.03 S cm
1
for the singly pro-
cessed electrolytes in the IT range, as shown in Fig. 14.
4. Conclusions
Aliovalent-doped ceria electrolytes were prepared by solid-
state, combustion, coprecipitation and hydrothermal
synthesis methods to improve the ionic conductivity of IT-SOEs
by elucidating the correlations between electrical properties
and processing variables. Asystematic set of ionic conductivity
data for 24 doped-ceria systems was analyzed using EIS and
various physicochemical tools such as XRD, FE-SEM and HR-
TEM. The doped-ceria SOEs showed a wide range of conduc-
tivity in the IT range depending on the dopant type, concen-
tration, powder synthesis method and pellet sintering
condition. Among these doped-ceria systems, Nd
0.2
Ce
0.9
O
2d
Fig. 13 e Microstructure of Nd
0.2
Ce
0.8
C
2Ld
composite pellets sintered at 1550

C for 4 h.
Fig. 14 e Comparisons of conductivity values of a set of
aliovalent-doped ceria electrolytes at 600

C measured by
EIS.
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1585
and Nd
0.1
Ce
0.9
O
2d
SOEs synthesized by combustion method
and sintered at 1450

C for 4 h exhibited the highest ionic
conductivities in air at 600

C and at below 550

C, respectively.
The ionic conductivity increased with increasing lattice
parameter in the ceria lattice, and with increasing dopant ionic
radius (r
3
Gd
< r
3
Sm
< r
3
Nd
) and content (10 mol% < 20 mol%) at
600

C. In addition, a lower sintering temperature raised the
ionic conductivity in the IT range due to the deleterious effects
of SiO
2
impurities and space charge layers which were respon-
sible for grain boundary resistivity. The grain boundary
conductivity of the combustion20NDCpellet sinteredat 1450

C
for 4 hwas one- andtwo-orders of magnitude higher thanthose
of the coprecipitation and solid-state pellets sintered at 1650

C
for 10 h, respectively. This result demonstrated that the
combination of nanoparticles and low temperature sintering
can effectively decrease the grain boundary resistance.
Further investigations were carried out on an advanced
electrolyte fabricated with a combination of powders (50:50)
synthesized by coprecipitation and combustion methods in
order to improve the electrical conductivity of the electrolyte.
The total ionic conductivity of the composite electrolyte was
2.5e4timeshigher thanthat of anysinglyprocessedelectrolytes
in the IT range. In particular, the grain boundary conductivities
of the 20NDC composite electrolytes were several orders of
magnitude higher than those of the single processed materials.
This was attributed to the homogeneous and uniform distri-
butions of submicron-sized grains prepared by coprecipitation
powders, as well as the enhanced grain boundary densication
achieved by using combustion powders consisting of porous
fractal and tightly agglomerated nanocrystallites.
Acknowledgments
This research was supported by the Fusion Research Program
for Green Technologies through the National Research Foun-
dation of Korea (NRF) funded by the Ministry of Education,
Science and Technology (#2011-0004428).
r e f e r e n c e s
[1] Steele BCH, Heinzel A. Materials for fuel-cell technologies.
Nature 2001;414:345e52.
[2] Wachsman ED, Lee KT. Lowering the temperature of solid
oxide fuel cells. Science 2011;334:935e9.
[3] Inoue T, Setoguchi T, Eguchi K, Arai H. Study of a solid oxide
fuel cell with a ceria-based solid electrolyte. Solid State
Ionics 1989;35:285e91.
[4] Inaba H, Tagawa H. Ceria-based solid electrolytes. Solid State
Ionics 1996;83:1e16.
[5] Huang J, Xie F, Wang C, Mao Z. Development of solid oxide
fuel cell materials for intermediate-to-low temperature
operation. Int J Hydrogen Energy 2012;37:877e83.
[6] Yahiro H, Eguchi Y, Eguchi K, Arai H. Oxygen ion conductivity
of the ceriaesamarium oxide system with uorite structure. J
Appl Electrochem 1988;18:527e31.
[7] Steele BCH. Appraisal of Ce
1y
Gd
y
O
2y/2
electrolytes for IT-
SOFC operation at 500

C. Solid State Ionics 2000;129:95e110.
[8] Zaja c W, Molenda J. Properties of doped ceria solid electrolytes
in reducing atmospheres. Solid State Ionics 2011;192:163e7.
[9] Zhang TS, Ma J, Cheng H, Chan SH. Ionic conductivity of
high-purity Gd-doped ceria solid solutions. Mater Res Bull
2006;41:563e8.
[10] Fuentesa RO, Baker RT. Synthesis and properties of
gadolinium-doped ceria solid solutions for IT-SOFC
electrolytes. Int J Hydrogen Energy 2008;33:3480e4.
[11] Mogensen M, Sammes NM, Tompsett GA. Physical, chemical
and electrochemical properties of pure and doped ceria.
Solid State Ionics 2000;129:63e94.
[12] Eguchi K, Setoguchi T, Inoue T, Arai H. Electrical properties of
ceria-based oxides and their application to solid oxide fuel
cells. Solid State Ionics 1992;52:165e72.
[13] Kim DJ. Lattice-parameters, ionic conductivities, and
solubility limits in uorite-structure MO
2
oxide (M = Hf
4
,
Zr
4
, Ce
4
, Th
4
, U
4
) solid-solutions. J Am Ceram Soc 1989;
72:1415e21.
[14] Hong SJ, Virkar AV. Lattice-parameters and densities of rare-
earth-oxide doped ceria electrolytes. J Am Ceram Soc 1995;
78:433e9.
[15] Kilner JA, Brook RJ. Astudy of oxygenionconductivity indoped
nonstoichiometric oxides. Solid State Ionics 1982;6:237e52.
[16] Omar S, Wachsman ED, Nino JC. A co-doping approach
towards enhanced ionic conductivity in uorite-based
electrolytes. Solid State Ionics 2006;177:3199e203.
[17] Andersson DA, Simak SI, Skorodumova NV, Abrikosov IA,
Johansson B. Optimization of ionic conductivity in doped
ceria. Proc Natl Acad Sci USA 2006;103:3518e21.
[18] Omar S, Wachsman ED, Jones JL, Nino JC. Crystal
structureeionic conductivity relationships in doped ceria
systems. J Am Ceram Soc 2009;92:2674e81.
[19] Fu Y, Chen S. Preparation and characterization of
neodymium-doped ceria electrolyte materials for solid oxide
fuel cells. Ceram Int 2010;36:483e90.
[20] Fang P, Li SP, Lua JQ, Pu ZY, Cen SQ, Luo MF. Effect of phase
structure on electrical conductivity of Ce
x
Gd
1x
O
2d
solid
electrolytes. Mater Sci Eng B 2009;164:101e5.
[21] Zha S, Xia C, Meng G. Effect of Gd (Sm) doping on properties
of ceria electrolyte for solid oxide fuel cells. J Power Sources
2003;115:44e8.
[22] Stephens IEL, Kilner JA. Ionic conductivity of Ce
1x
NdO
2x/2
.
Solid State Ionics 2006;177:669e76.
[23] Fu Y-P, Chen S-H, Huang J-J. Preparation and
characterization of Ce
0.8
M
0.2
O
2d
(M= Y, Gd, Sm, Nd, La) solid
electrolyte materials for solid oxide fuel cells. Int J Hydrogen
Energy 2010;35:745e52.
[24] Yan D, Liu X, Bai X, Pei L, Zheng M, Zhu C, et al. Electrical
properties of grain boundaries and size effects in samarium-
doped ceria. J Power Sources 2010;195:6486e90.
[25] Pe rez-Coll D, Nu n ez P, Frade JR. Improved conductivity of
Ce
1x
SmO
2d
ceramics with submicrometer grain sizes. J
Electrochem Soc 2006;153:A478e83.
[26] Christie GM, van Berkel FPF. Microstructure-ionic
conductivity relationships in ceria-gadolinia electrolytes.
Solid State Ionics 1996;83:17e27.
[27] Bellino MG, Lamas DG, de Reca NEW. Enhanced ionic
conductivity in nanostructured, heavily doped ceria
ceramics. Adv Funct Mater 2006;16:107e13.
[28] Karaca T, Altnc ekic TG, O

ksu zo mer MF. Synthesis of


nanocrystalline samarium-doped CeO
2
(SDC) powders as
a solid electrolyte by using a simple solvothermal route.
Ceram Int 2010;36:1101e7.
[29] Kosinski MR, Baker RT. Preparation and
propertyeperformance relationships in samarium-doped
ceria nanopowders for solid oxide fuel cell electrolytes. J
Power Sources 2011;196:2498e512.
[30] Guo X, Sigle W, Maier J. Blocking grain boundaries in yttria-
doped and undoped ceria ceramics of high purity. J Am
Ceram Soc 2003;86:77e87.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1586
[31] Reis SL, Souza ECC, Muccillo ENS. Solid solution formation,
densication and ionic conductivity of Gd- and Sm-doped
ceria. Solid State Ionics 2011;192:172e5.
[32] Souza ECC, Chueh WC, Jung W, Muccillo ENS, Haile SM. Ionic
and electronic conductivity of nanostructured, samaria-
doped ceria. J Electrochem Soc 2012;159:K127e35.
[33] Kim S, Maier J. On the conductivity mechanism of
nanocrystalline ceria. J Electrochem Soc 2002;149:J73e83.
[34] Lenka RK, Mahata TK, Tyagi AK, Sinha PK. Inuence of
grain size on the bulk and grain boundary ion conduction
behavior in gadolinia-doped ceria. Solid State Ionics 2010;
181:262e7.
[35] Lenka RK, Mahata T, Sinha PK, Tyagi AK. Combustion
synthesis of gadolinia-doped ceria using glycine and urea
fuels. J Alloy Compd 2008;466:326e9.
[36] Mukherjee ST, Bedekar V, Patra A, Sastry PU, Tyagi AK. Study
of agglomeration behavior of combustion-synthesized nano-
crystalline ceria using new fuels. J Alloy Compd 2008;466:
493e7.
[37] Tian RF, Zhao F, Chen FL, Xia CR. Sintering of samarium-
doped ceria powders prepared by a glycine-nitrate process.
Solid State Ionics 2011;192:580e3.
[38] Chang HY, Wang YM, Lin CH, Cheng SY. Effects of rapid
process on the conductivity of multiple elements doped
ceria-based electrolyte. J Power Sources 2011;196:1704e11.
[39] Prasad DH, Kim HR, Park JS, Son JW, Kim BK, Lee HW, et al.
Superior sinterability of nano-crystalline gadolinium doped
ceria powders synthesized by co-precipitation method. J
Alloy Compd 2010;495:238e41.
[40] Fu Y, Wen S, Lu C. Preparation and characterization of
samaria-doped ceria electrolyte materials for solid oxide fuel
cells. J Am Ceram Soc 2008;91:127e31.
[41] Dikmen S, Shuk P, Greenblatt M, Gocmez H. Hydro-thermal
synthesis and properties of Ce
1x
GdO
2d
solid solutions.
Solid State Sci 2002;4:585e90.
[42] Wu N-C, Shi E-W, Zheng Y-Q, Li W-J. Effect of pH of medium
on hydrothermal synthesis of nanocrystalline cerium(IV)
oxide powders. J Am Ceram Soc 2002;85:2462e8.
[43] Dos Santos ML, Lima RC, Riccardi CS, Tranquilin RL,
Bueno PR, Varela JA, et al. Preparation and characterization
of ceria nanospheres by microwave-hydrothermal method.
Mater Lett 2008;62:4509e11.
[44] George KC, Kurien S, Mathew J. Lattice strain and lattice
expansion of nanoparticles of MgAl
2
O
4
as a function of
particle size. J Nanosci Nanotechnol 2007;7:2016e9.
[45] Fukuhara M. Lattice expansion of nanoscale compound
particles. Phys Lett A 2003;313:427e30.
[46] Cimino A, Porta P, Valigi M. Dependence of the lattice
parameter of magnesium oxide on crystallite size. J Am
Ceram Soc 1996;49:152e6.
[47] Schmid HK. Quantitative analysis of polymorphic mixes of
zirconia by X-ray diffraction. J Am Ceram Soc 1987;70:367e76.
[48] Khakpour Z, Youzbashi AA, Maghsoudipour A, Ahmadi K.
Synthesis of nanosized gadolinium doped ceria solid solution
by high energy ball milling. Powder Tech 2011;214:117e21.
[49] Chandradass J, Nam B, Kim KH. Fine tuning of gadolinium
doped ceria electrolyte nanoparticles via reverse
microemulsion process. Colloids Surf A Physicochem Eng
Aspects 2009;348:130e6.
[50] Zhang F, Chan SW, Spanier JE, Apak E, Jin Q, Robinson RD,
et al. Cerium oxide nanoparticles: size-selective formation
and structure analysis. Appl Phys Lett 2002;80:127e9.
[51] Askrabic S, Dohcevic-Mitrovic ZD, Radovic M, Scepanovic M,
Popovic ZV. Phononephonon interactions in Ce
0.85
Gd
0.15
O
2d
nanocrystals studied by Raman spectroscopy. J Raman
Spectrosc 2009;40:650e5.
[52] McBride JR, Hass KC, Poindexter BD, Weber WH. Raman and
X-ray studies of Ce
1x
RE
x
O
2y
, where RE = La, Pr, Nd, Eu, Gd
and Tb. J Appl Phys 1994;76:2435e41.
[53] Gouadec G, Colomban Ph. Raman spectroscopy of
nanomaterials: how spectra relate to disorder, particle size
and mechanical properties. Prog Cryst Growth Charact Mater
2007;53:1e56.
[54] Chen P-L, Chen I-W. Sintering of ne oxide powders: I,
microstructural evolution. J Am Ceram Soc 1996;79:3129e41.
[55] Chen P-L, Chen I-W. Sintering of ne oxide powders: II,
sintering mechanisms. J Am Ceram Soc 1997;80:637e45.
[56] Kingery WD, Francois B. In: Kuczynski GC, Hooton NA,
Gibbon CF, editors. Sintering and related phenomena; 1967.
p. 471e96. New York.
[57] de Larramendi IR, Ortiz-Vitoriano N, Acebedo B, de
Aberasturi DJ, de Muro IG, Arango A, et al. Pr-doped ceria
nanoparticles as intermediate temperature ionic conductors.
Int J Hydrogen Energy 2011;36:10981e90.
[58] Minh NQ, Takahashi T. Science and technology of ceramic
fuel cells. Amsterdam: Elsevier; 1995.
[59] Dikmen S. Effect of co-doping with Sm
3
, Bi
3
, and Nd
3
on
the electrochemical properties of hydrothermally prepared
gadolinium-doped ceria ceramics. J Alloys Compd 2010;491:
106e12.
[60] Tagawa H. SOFC and environmental of earth. Tokyo, Japan:
Agune Shohu Publishing; 1998.
[61] Steele BCH, Zheg K, Rudkin RA, Kiratzis N, Christie M. In:
Dokiya M, Yamamoto O, Tagawa H, Singhal SC, editors.
Proceedings of the fourth international symposium on solid
oxide fuel cells (SOFC-IV). New Jersey: The Electrochemical
Society; 1995. p. 1028e38.
[62] Kamiya M, Shimada E, Ikuma Y, Komatsu M, Haneda H,
Sameshima S, et al. Oxygen self-diffusion in cerium oxide
doped with Nd. J Mater Res 2001;16:179e84.
[63] Park J-Y, Yoon H, Wachsman ED. Fabrication and
characterization of high-conductivity bilayer electrolytes for
intermediate-temperature solid oxide fuel cells. J Am Ceram
Soc 2005;88:2402e8.
[64] Hong SJ, Mehta K, Virkar AV. Effect of microstructure and
composition on ionic conductivity of rare-earth oxide-doped
ceria. J Electrochem Soc 1998;145:638e47.
[65] Tian C, Chan S. Ionic conductivities, sintering temperatures
and microstructures of bulk ceramic CeO
2
doped with Y
2
O
3
.
Solid State Ionics 2000;134:89e102.
[66] Kidner NJ, Perry NH, Mason TO. The brick layer model
revisited: introducing the nano-grain composite model. J Am
Ceram Soc 2008;91:1733e46.
[67] Gibson IR, Dranseld GP, Irvine JTS. Sinterability of
commercial 8 mol% yttria-stabilized zirconia powders and
the effect of sintered density on the ionic conductivity. J
Mater Sci 1998;33:4297e305.
[68] Pe rez-Coll D, Sa nchez-Lo pez E, Mather GC. Inuence of
porosity on the bulk and grain-boundary electrical properties
of Gd-doped ceria. Solid State Ionics 2010;181:1033e42.
[69] Zhao F, Virkar AV. Effect of morphology and space charge on
conduction through porous doped ceria. J Power Sources
2010;195:6268e79.
[70] Tschope A, Bauerle C, Birringer R. Numerical analysis of space
charge layers andelectrical conductivity inmesoscopic cerium
oxide crystals. J App Phys 2004;95:1203e10.
[71] Lubomirsky I, Fleig J, Maier J. Modeling of space-charge
effects in nanocrystalline ceramics: the inuence of
geometry. J App Phys 2002;92:6819e27.
[72] Singh V, Babu S, Karakoti1 AS, Agarwal A, Seal S. Effect of
submicron grains on ionic conductivity of nanocrystalline
doped ceria. J Nanosci Nanotechnol 2010;10:6495e503.
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 1 5 7 1 e1 5 8 7 1587

Vous aimerez peut-être aussi