Sunteți pe pagina 1din 41

Accepted Manuscript

Studies on structural, morphological and electrical properties of Ce0.8 Ln


3+
, Gd3+, Sm3+, Nd3+ and La3+) solid solutions prepared by citrate
0.2O2− δ , (Ln = Y

complexation method

K.C. Anjaneya, G.P. Nayaka, J. Manjanna, G. Govindaraj, K.N. Ganesha

PII: S0925-8388(13)02264-0
DOI: http://dx.doi.org/10.1016/j.jallcom.2013.09.101
Reference: JALCOM 29455

To appear in:

Received Date: 26 July 2013


Revised Date: 13 September 2013
Accepted Date: 16 September 2013

Please cite this article as: K.C. Anjaneya, G.P. Nayaka, J. Manjanna, G. Govindaraj, K.N. Ganesha, Studies on
structural, morphological and electrical properties of Ce0.8 Ln 0.2O2− δ , (Ln = Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid
solutions prepared by citrate complexation method, (2013), doi: http://dx.doi.org/10.1016/j.jallcom.2013.09.101

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Studies on structural, morphological and electrical properties of Ce0.8Ln0.2O2−δ, (Ln = Y3+,

Gd3+, Sm3+, Nd3+ and La3+) solid solutions prepared by citrate complexation method

K.C. Anjaneyaa, G.P. Nayakaa, J. Manjannaa*, G. Govindarajb, K.N. Ganeshab

a
Department of Industrial Chemistry, Kuvempu University, Shankarghatta 577 451, India.
b
Department of Physics, Pondicherry University, Pondicherry, 605 001, India.

Corresponding author (J. Manjanna)

E-mail: jmanjanna@rediffmail.com, Tel: +91 8282−256 228; Fax: +91 8282−256 255

Keywords: Fuel cells, Ceramics, Chemical synthesis, Vacancy formation, Electrochemical

impedance spectroscopy
Abstract

A modified sol-gel method which is also known as citrate complexation is used here to prepare

Ce0.8Ln0.2O2−δ (Ln= Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid solutions for their proposed

application in intermediate-temperature solid oxide fuel cells (IT-SOFCs) as electrolytes. All the

samples exhibit the fluorite type crystalline structure without any phase segregation,

characteristics of CeO2, as revealed in XRD pattern. The lattice parameters and densities

calculated based on the oxygen vacancy model agreed well with the experimental (Archimedes

method) values. The formation of solid solution is also confirmed by Raman spectroscopy. The

microstructural features of the samples are recorded by using FE-SEM/TEM. The ionic

conductivities of various Ln-doped samples at 623 K as obtained from electrochemical

impedance measurement are as follows: Ce0.8Sm0.2O2−δ > Ce0.8Gd0.2O2−δ > Ce0.8Nd0.2O2−δ >

Ce0.8La0.2O2−δ > Ce0.8Y0.2O2−δ. The activation energy for total conductivity are found to be 0.86,

0.87, 0.89, 0.96 and 1.02 eV for Sm, Gd, Nd, La and Y doped ceria, respectively. The

relationship between the dopant radius, chemical composition, lattice parameters, morphology

and electric properties are discussed here.


 

1. Introduction

Solid oxide fuel cells (SOFCs) are regarded as one of the clean sources of energy with high

conversion efficiency of chemical energy to electric energy without any pollution [1−2]. The

high working temperature of SOFC has been a major hurdle in the realization of this important

green technology in real life [3]. When compared to other types of fuel cells, SOFCs have

several advantages like fuel flexibility i.e., they are able to work with hydrogen, hydrocarbon

reformate and in some conditions directly with hydrocarbon fuels [4]. However, the widespread

commercialization of SOFCs has not been realized due to its cost associated with cell

fabrication, electrode/ electrolyte materials and its maintenance. To increase its

commercialization, lowering the operation temperature to intermediate temperature (IT) ca. 873

K−1073 K is one of the main goals of the current SOFCs research. In the IT range, less

expensive material such as steel could be used in the auxiliary components [5]. Among the

SOFC components, the electrolyte (ionic conductor) plays an important role, serve as heart of the

device, and influence on the performance of the cell. IT-SOFCs can operate in dual mode: in fuel

cell mode, stored hydrogen can be used to generate electricity and heat, while in electrolysis

mode it can produce hydrogen from steam. IT solid oxide electrolysis cells (SOECs) uses waste

heat from power stations and other industries such as steel making industry to produce hydrogen.

Due to the requirement similarity, much of the materials used in IT-SOFCs can also be used in

IT-SOECs. Unlike SOFC, SOEC has not been studied in detail. The efficiency of IT-SOFC/ IT-

SOEC can be improved by developing novel electrolyte materials with high oxygen ion

conductivity [6,7]. A typical high temperature SOFCs, which operates at 1073 K−1273 K uses

yttria stabilized zirconia (YSZ) as the electrolyte. But this high temperature causes several
problems like high fabrication cost, reaction between the cell components and also thermal

expansion mismatch [8].

To overcome these problems and to reduce the operating temperature, rare earth (RE) doped

ceria is considered as the potential electrolyte materials for IT−SOFCs. This is due to its inherent

high oxide ion conductivity at IT when compared to YSZ. RE doped ceria is also preferred due

to its low cost when compared to lanthanum gallate based electrolytes [9]. Pure cerium oxide is

fundamentally a poor ionic conductor having a fluorite structure with space group Fm3m over

the temperature range from room temperature to its melting point and oxygen vacancies as

the predominant ionic defects. Partial substitution of tetravalent cerium by trivalent RE dopants

(RE = Y3+, Gd3+, Sm3+, Nd3+ and La3+) results in the formation of and thus increases the

oxygen ion conductivity. The defect reactions can be written in Kroeger−Vink notation as:

where, the symbols denote the usual meanings of the Kroger−Vink notations i.e. each pair of

RE3+ cations creates one defect for the charge compensation. The and point defects

tend to cluster due to static electrical attraction and the tendency increases at higher doping level,

leading to the formation of larger (RE − ) pairs of defect aggregation [10−12]. Ionic

conductivity of ceria can be increased by increasing the oxygen vacancies. However, at higher

dopant levels defect association occurs between cation and vacancies. Once an associate is

formed, the oxygen vacancy is not free to participate in the conduction process and thus free

vacancy concentration is lowered [13,14]. Depending on the characteristics introduced by the

dopant, doped ceria is also used for variety of applications such as, oxygen storage materials,

oxygen sensors, catalysts etc at high temperature [15,16].


Design of the ceria based electrolytes for IT-SOFCs mainly depends on dopant ion, dopant

concentration, oxygen vacancy concentration, defect association energy and local defect

structure. All of these strongly influence on the homogeneity, stability and electrical conductivity

of the solid solutions [17]. However, ionic conductivity can be optimized when the dopant ionic

radius is close to critical radius (rc) i.e., when the ionic radius of the dopant causes neither

expansion nor contraction of the fluorite lattice. For trivalent cation in ceria, rc 1.038 Å [18].

Below this critical value, the ionic conductivity increases with dopant radius and above which

the ionic conductivity decreases with dopant radius i.e. ionic radius of dopant strongly influence

the ionic conductivity of ceria [19]. The doping of lanthanide group ions viz., Y3+, Gd3+, Sm3+,

Nd3+, and La3+ in ceria are known to enhances the oxygen ion conductivity significantly.

One of the major problems associated with the application of RE doped ceria as SOFC

electrolytes is the ceramic grain boundary (GB) resistance, which is usually larger than or in the

same range as the bulk resistance, although the GB areas are expected to be thin [20]. This

implies that GBs constitute barriers for ionic charge transport and they indeed often act as

Schottky barriers [21]. Therefore, the reduction of the GB resistance is of uppermost importance.

According to Bauerle model [22], the GB resistance is closely associated with the impurity

(SiO2) and RE cations segregating at GB, which are mainly influenced by the synthesis routes.

To reduce the GB resistance several attempts have been made among them preparation method

of solid solution is also one, which influence not only electrical properties, but also on the

structural and micro structural properties.

Ceria based solid solutions can be synthesized by variety of techniques such as solid-state

reaction [23,24], precipitation using ammonia [25] or ammonium carbonate [26−28],

hydrothermal [20], combustion [29] sol−gel [30] and hexamethylenetetramine−based


homogeneous precipitation [31]. It is well established that wet chemical routes are advantages

over solid−state route. In the present work, we have synthesized Ce0.8Ln0.2O2-δ, (Ln= Y3+, Gd3+,

Sm3+, Nd3+, and La3+) by variant sol-gel technique known as citrate - complexation as proposed

by Muccillo et al [32]. This method was advantages such as low-cost, relative simplicity and it

emits less carbon residues than the other similar techniques. Although RE-doped ceria have been

investigated widely since past three decades, no detailed reports are available on the effect of

dopant radius on the total conductivity of Ce0.8Ln0.2O2−δ solid solutions prepared by citrate –

complexation method. Thus, here we have focused on the influence of the different dopant ion

radius and concentration on the lattice parameter, density, micro structural characteristics and

electric properties of Ce0.8Ln0.2O2−δ solid solutions.

2. Experimental

2.1. Sample preparation

Solid solutions of Ce0.8Ln0.2O2−δ, (Ln = Y3+, Gd3+, Sm3+, Nd3+ and La3+) were synthesized by

citrate – complexation method. Stoichiometric amounts of cerium and RE nitrates i.e.,

(Ce(NO3)3·6H2O & Ln(NO3)3·6H2O were dissolved in distilled water to get a homogenized

mixture of metal ions. To this solution, citric acid was added so that the total metal ion to citrate

ratio fixed to 1: 2. On heating this mixture to 353 K under stirring for 6 h, the excess water and

the nitrate gases were removed to form a transparent and viscous gel. The gel was transferred to

a crucible and heated to 573 K for 4 h in a muffle, which led to decomposition of precursors. The

residual powder was then calcined at 873 K for 4 h to get final products, designated hereon as

YDC20, GDC20, SDC20, NDC20 and LDC20 respectively.


For pelletizing (10 mm 2 mm), the required powder samples were uniaxially pressed to 5 Mpa

and sintered at 1473 K in air for 4 h with programmed heating rate of 5 K per minute and cooling

rate of 3 K per minute.

2.2 Characterization

X−ray powder diffractometer (XRD; Philips PW 3710 diffractometer) with Cu Kα radiation (λ =

0.15418 nm) was used to identify the crystalline phase and lattice parameters of calcined

Ce0.8Ln0.2O2−δ samples. The relative density of the sintered samples were obtained using the

equation:

where, dm is the density of samples measured by Archimedes method and dth is the theoretical

density, which is given by:

where, is the dopant ion content, a is the lattice constant of Ce0.8Ln0.2O2−δ, NA is the Avogadro

number and MCe is the atomic weight of cerium, MLn is the atomic weight of corresponding

lanthanide ion and Mo is the atomic weight of oxygen respectively [33].

The vibrational measurement of the samples were carried out using Perkin-Elmer IR spectra in

the frequency range 4000−400 cm−1 and Raman spectroscopy (HORIBA−JOBIN YVON and λ =

633 nm) in the frequency range 200−1000 cm−1. The morphology of the calcined powder and

sintered pellets were recorded by transmission electron microscope (TEM, JEM-3010, JEOL,
JAPAN) and field-emission scanning electron microscope (FE-SEM Carl Zeiss, Supra 40 VP)

with energy-dispersive X-ray spectroscopy (EDX) respectively. AC electrical properties were

measured by electrochemical impedance spectroscopy at different temperature (423 K−623 K)

using Material Mates (M2) – 7260 impedance analyzer in the frequency range 1 Hz to 10 MHz

with an AC signal of 10 mV. For ionic conductivity measurement, silver paste was applied on

both sides of pellets followed by baking at 673 K for 1 h. Data were fitted to the corresponding

equivalent circuits using Z−view software. Conductivity (σ) was calculated using the following

equation:

where, = thickness of the pellet, A = area of cross section and R = resistance.

3. Results and discussion

Fig. 1 is the XRD pattern of Ce0.8Ln0.2O2−δ calcined at 873 K. These patterns matches with the

standard cubic fluorite structure (JCPDS File No 34−0394; Fm3m) and there was no impurity

phase. Also, a small shift in the ceria peaks (as shown in Fig. 2) depending on the radius of RE-

dopant ion indicates the formation of single-phasic solid solutions with a marginal increase in

lattice parameter, a (= 5.411 Å for pure CeO2 [34]), which is calculated as:

where, d is the inter planar spacing and θ is the diffraction angle. Theoretically, the lattice

parameter of the RE doped ceria is given by Hong and Virkar [35] as:
where, x is the concentration of dopant and rLn, rCe, rO and are the radii of dopant cations

(Y3+= 1.03 Å, Gd3+ = 1.05 Å, Sm3+ = 1.08 Å, Nd3+ = 1.11 Å and La3+ = 1.15 Å), cerium ion

(Ce4+ = 0.97 Å), the oxygen ion (O2− = 1.38 Å) and the oxygen vacancy ( = 1.164 Å)

respectively. The measured and calculated lattice parameters of Ce0.8Ln0.2O2−δ as a function of

dopant radius are shown in Fig. 3. The lattice parameter increases with increasing atomic radius

of the dopant, which is attributed to the lattice expansion with increasing ionic radius. As shown

in Fig. 3, the measured lattice parameters are in accordance with calculated lattice parameters.

The crystallite size D was calculated from the Scherrer’s equation:

where, λ is the wavelength of the X rays (1.5418Å), θ is the angle of main reflection (111) and β

is full width at half-maximum intensity (in radians). Fig. 3 is the plot of crystallite size versus

dopant ionic radius for the samples calcined at 873 K. It clearly indicates that crystallite size

increase with increasing the radius of the dopant from Y3+ to Sm3+ and then decrease with further

increase in the size of the dopant ion. Therefore, in Ce0.8Ln0.2O2−δ, crystallite growth strongly

depends on the dopant size. The lattice parameters and crystallite sizes of Ce0.8Ln0.2O2−δ as a

function of dopant radius are given in Table 1.

During the formation of rare earth doped ceria solid solutions, several defect reactions can be

created. Among them, oxygen vacancy model and interstitial model are very important, which

are described as follows:


Oxygen vacancy model: MO1.5 M Ce 0.5V O 1.5 O O (8)

Interstitial model: MO1.5 0.75M Ce 0.25M i 1.5O O (9)

From the above reactions, density for oxygen vacancy and interstitial models are given as

follows:

Here Dvac is the density of the oxygen vacancy model, Dint is the density of the interstitial model,

Md is the atomic weight of dopant cation, MCe is the atomic weight of cerium, Mo is the atomic

weight of oxygen, NA is the Avogadro number and a is the lattice parameter obtained from Eq.

(6). Fig. 4 shows measured density, oxygen vacancy model and interstitial model density for

Ce0.8Ln0.2O2−δ. Lattice parameters and densities based on oxygen vacancy model agreed well

with the experimental results and hence support the oxygen vacancy model for RE doped ceria

[36]. The measured, calculated, and relative densities of solid solutions as a function of dopant

ion radius are given in Table 2.

Fig. 5 represents the FT-IR spectrum of calcined (at 573 K) Ce0.8Ln0.2O2−δ. In all the samples,

bands were observed at 500−800, 1100−1600 and 3100− 3450 cm−1 ranges. The observed bands

are assigned as follows: The bands around 500−800 cm−1 range are ascribed to the characteristic

stretching Ce−O vibrations [37]. The band at 1300 cm-1 is most probably due to the presence of
small amount remnant nitrates. The band at 1600 is due to CO2. Less intense peak at 1100 cm-1

might be due residual organics present in the samples [38]. The C−H stretching mode is observed

in the range 3100−3300 cm-1 [39]. Lastly, the intense band observed at 3430 cm−1 corresponds to

γ(O−H) stretching frequency of water molecule, which is due to the presence of adsorbed

moisture.

A Raman spectrum is a very good technique to know about both M−O bond arrangement and

lattice defects. Thus, it acts as potential tool to obtain additional structural information apart from

XRD data. Raman spectra of the sintered Ce0.8Ln0.2O2−δ is shown in Fig. 6. It is well-known that,

CeO2 is having fluorite structure with single allowed Raman mode at 465 cm-1, which may be

attributed to the symmetric F2g breathing mode of oxygen atoms around each cerium ion

[40−42]. As shown in Fig 6, the shift in the F2g mode attributed to change in M−O vibration

frequency after incorporation with RE cations. The shift in frequency depends on the ionic radius

of the dopant, larger the radius of the dopant more is the shift. The reduction of frequency

follows the order: Y3+(467 cm-1) > Gd3+ (463 cm-1) > Sm3+(461 cm-1) > Nd3+(460 cm-1) > La3+

(458 cm-1). Furthermore, one broad additional peak was observed between 550 – 600 cm−1,

which may be attributed to the presence of intrinsic and extrinsic oxygen vacancies caused by the

nonstoichiometry of the ceria and substitution of Ce4+ by Ln3+ ions respectively [43−45]. The

intensity of above peaks depends on the number of vacancies. Since the concentration is fixed in

all investigated samples, it depends on size of the vacancies, which increases with increase in the

radius of the dopant [46]. It is important to note that there are no vibrational mode corresponding

to any free rare earth sesquioxides (Y2O3, Gd2O3, Sm2O3, Nd2O3 and La2O3) phase expected at

360 cm-1 [47], which reveals that solid solution formation was complete at 1473 K. During heat

treatment, heat energy supplied to the system is used for the formation of solid solutions. This
observation confirms the formation of Ce0.8Ln0.2O2−δ solid solutions in corroboration with XRD

data.

Fig. 7 (a-b) shows the TEM image and selected area electron diffraction (SAED) patterns of

calcined SDC20 powder. The TEM micrograph clearly shows the presence of nanoparticles in

the sample. The nanocrystallinity is also confirmed by the presence of rings in the SAED as

shown in Fig. 7(b). SAED ring pattern corresponds to the fluorite structure of RE doped ceria.

Microstructural features of Ce0.8Ln0.2O2−δ (Ln= Y3+, Gd3+, Sm3+, Nd3+ and La3+) ceramics

sintered at 1473 K are shown in Fig 8 (a-e). All the sintered samples show good densification

with few closed pores at the GB or at the triple junction point and some intragranular pores.

These closed pores can be removed from the sintered ceramics via a higher sintering temperature

or prolonging the soaking time. Moreover, the grain size seems to be independent of dopant ion

radius and the grain sizes are distributed in the range of 100−150 nm for Ce0.8Ln0.2O2−δ solid

solutions. As electrical conductivity depends on the sintered density, here all samples exhibits

well-sintered density, hence enhanced total conductivity and low activation energy. Even though

EDX is taken for all the sintered samples (not shown here), here we are given EDX only for

SDC20 (Fig 8 (f)), which confirms the desired stoichiometry of all the constituents elements.

Impedance spectroscopy is a well-developed tool to study the electrical properties of solid

electrolytes. Bauerle [48] first introduced the use of this spectroscopy to determine oxygen ion

conductivity. In impedance spectroscopy addition to the total ionic conductivity, it is possible to

get the information about electrode process and separate contribution of grain and the GB

resistance. The impedance spectrum is typically represented as negative of imaginary component

of impedance ( Z'') versus real component of impedance (Z') which is referred as Nyquist plot.

Typical impedance spectra of electrolyte material consist of three semicircles, each semicircle
represents distinct process. The semicircles at high frequency and low frequency arc ascribed to

grain and electrode process while semicircle at intermediate frequency represents the GB

contribution. At high temperature, the semi circle due to the grain and GB disappears. This is due

to time constant associated with grain and GB impedances are much smaller when compare to

electrode process [49,50]. In the present study, the equivalent circuit model had adopted for the

data analysis, as this is most frequently used technique for solid ionic materials and the use of

simple capacitor is not sufficient here to model the electric properties of the materials. Hence, for

this purpose a constant phase element (CPE) was used to fit the results. Using the equation ωRC

= 1 where ω is the angular frequency ( = 2πf, f is the frequency in Hz) and R is the arc

magnitude, the grain and GB capacitances (C) can be calculated. Grains have a capacitance of

pF, whereas GB have a capacitance of nF.

Fig 9 (a−c) represents the Nyquist plots for YDC20, GDC20, SDC20, NDC20 and LDC20 at 573

K, 598 K and 623 K temperature respectively. The impedance spectra of all the rare earth doped

samples measured at different temperature shows that decrease in resistance with increasing

temperature and thus the conductivity of all samples increases with increase in temperature. Here

the semicircles corresponding to grain and GB conductivities are highly overlapped, which might

be due to similar relaxation time of charge carriers both inside the grain and GB and it was

difficult to distinguish between them. Therefore, only total conductivity is calculated here for all

RE doped ceria solid solutions. The total conductivity can be calculated by using the Eq. (4).

Some samples at 623 K exhibits of an arc beside the overlapped grain and GB arc. The arc is due

to the electrode processes, which shows the interface conduction between the electrolyte and

electrode in these samples.


The Arrhenius plot for total conductivity of all the RE doped ceria samples is shown in Fig. 10. It

can be clearly seen that, experimental data is well fitted with straight line throughout the

measured temperature. This indicates that the conductivity can be expressed in the form of

Arrhenius relation:

where Ea = activation energy, k = Boltzmann constant, σ0 = pre-exponential factor. Ceria doped

with RE elements bestow higher oxide ion conductivity than those doped with other elements

[51]. Here the ionic conductivity was highest for SDC20 among the RE doped ceria solid

solution. The total ionic conductivities of Ce0.8Ln0.2O2−δ at 623 K are ranked as follows: SDC20

> GDC20 > NDC20 > LDC20 > YDC20. The size of the dopant plays an important role in the

determination of total conductivity. As shown in Fig. 11, the conductivity of RE doped ceria

generally increases as the dopant radius increases and reaches a maximum for samarium. After

that, the conductivity decreases with increase in dopant radius. Since the dopant concentration

was fixed the mobility of the oxygen vacancy depend on the radius of the dopant ion. The

striking observation here is that our samples exhibits the enhanced conductivity even at low

temperature compared to the data available in the literature. Yamashita et al [52] reported a

conductivity of only 4×10-6/ 8×10-7 Scm-1 at 773 K for SDC20/ YDC20 compared to 9.73×10-4/

4.65×10-5 Scm-1 observed in the present study at 623 K respectively. B.S. Prakash et al [53]

measures the ionic conductivities of glycine-nitrate synthesized GDC20 and reported a

maximum conductivity of 8.1×10-5 Scm-1at 673 K, compared to 4.41×10-4 Scm-1 at 623 K

observed in the present study.


To explain this enhanced oxygen ion conductivity observed in our samples, we tried to correlate

a crystallographic index term called “effective index, (Ei)” with the conductivity. Mori [54,55] et

al., originally introduced the term Ei for designing new oxide ion conducting materials, taking

into the account of the ionic radii mismatch between the dopant ion and host cation and the

amount of oxygen vacancies that are expected to produce by the dopant substitution in an

idealized crystallographic sites. The Ei is given by the formula

where, avg.rc, rd and rh are the average ionic radius of the cation, dopant and host ion (Ce4+)

respectively. The eff ro is the effective oxygen ion radius, which is given by the formula

where, 1.4 Å is the ionic radius of O2- in oxides and δ is the level of oxygen vacancies in CeO2

based oxides.

As per the Eq. (13), the system will have an idealized non-distorted fluorite structure when the

value of Ei becomes unity or close to unity. Thus, in general for high oxygen conductivity, Ei of

that system must be close to unity [56]. The Ei, ionic radius ratio rd/rh, σt and Ea at 623 K for

Ce0.8Ln0.2O2−δ are given in Table 3. Although it has been suggested that the conductivity is high

when Ei of the system was close to unity, it does not seem applicable in this case. For example,

even though Ei for LDC20 is close to unity than that of SDC20, the oxide ion conductivity of

LDC20 is lower than that of SDC20. According to the critical ionic (rc) radius explanation, the

conductivity is high for those dopants whose radius was to close to 1.038 Å [18] and Y3+ is
closer to this value than that of Sm3+. This is not in accordance with tendency of the conductivity

observed in the present study. Hence, from these reasons we explain it in terms of ionic radius

ratio (rd/rh) as explained by Kilner [57]. Accordingly, the conductivity is high for those dopants,

which are having minimum association enthalpy between the dopant cation and the anion

vacancy i.e., rd/rh value close to unity. When rd/rh approaches to unity, the lattice strain is

expected to be minimum. Minimum lattice strain could lead to the increased anion mobility and

hence higher conductivity. As shown in Table.3, rd/rh close to unity for SDC20 and therefore it

exhibits highest conductivity among the Ce0.8Ln0.2O2−δ, (Ln= Y3+, Gd3+, Sm3+, Nd3+ and La3+)

solid solutions. However, the global lattice strain is not always a major factor for the high oxide

ion conductivity, local structure may be important. Hence, the conductivity also depends on

sample preparation method, amount of SiO2 contamination and RE cations segregating at GB

and micro structural feature of sintered pellets etc. In the present citrate method, the amount of

SiO2 contamination and RE cations segregating at GB is low in comparison to other methods

such as solid-state reaction etc. Both the amount of SiO2 contamination at GBs and the RE

cations segregating to GBs may have a significant effect on the total conductivities. Here for

SDC20, SiO2 contamination and Sm cations segregating at GBs might be lowest and also SDC20

contain more continuous GB phases (as shown in Fig 8 (C)), than those of GDC20, NDC20,

LDC20 and YDC20. Hence, SDC20 exhibits highest total conductivity among the RE doped

ceria solid solutions, which is in corroboration with Kilner theory. The activation energy (Ea) is

calculated from the slope of the straight line plot of log(σT) vs. 1/T. The Ea for total conduction

was found out to be 0.86, 0.87, 0.89, 0.96 and 1.02 eV for Sm, Gd, Nd, La and Y doped ceria

respectively. The Ea for total conduction shows that all the solid solutions are oxide ion

conductors.
Based on the electrical measurement in the temperature range of 473−623 K, we believe that all

Ce0.8Ln0.2O2−δ (Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid solutions can be used as a potential

electrolytes for IT−SOFCs/ IT−SOECs. Nevertheless, detailed electrical measurements (ca. open

circuit voltage, ionic transference number etc.) and mechanical properties are necessary.

Conclusion

In this study, the lattice parameters, densities, microstructure, ionic conductivity and activation

energy of Ce0.8Ln0.2O2−δ (Ln = Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid solutions prepared by

citrate complexation method are determined. The lattice parameters and densities of Ln-doped

ceria were examined by introducing the concept of oxygen vacancy model. All the calcined

samples exhibit cubic fluorite-type phase formation. Unit cell parameters increased with

increasing the dopant radius. The solid solution formation was also confirmed by Raman

spectroscopy. The microstructural features of the calcined powder and sintered pellets were

recorded by using TEM/FE-SEM respectively. Sm doped ceria exhibits highest total conductivity

among the RE doped ceria samples. The activation energy for total conductivity was found to be

0.86, 0.87, 0.89, 0.96 and 1.02 eV for Sm, Gd, Nd, La and Y doped ceria respectively. Hence, we

envisage that these RE doped ceria materials can be used as a potential electrolyte material for

IT-SOFCs/ IT-SOECs applications.

Reference
1. T. Suzuki, B. Liang, T. Yamaguchi, H. Sumi, K. Hamamoto, Y. Fujishiro, J. Power
Sources 199 (2012) 170−173.
2. Z. Li, T. Mori, P. Yan, Y. Wu, Z. Li, Mat. Sci. Eng. B 177 (2012) 1538−1541.

3. H.L. Tuller, A.S. Nowick, J. Electrochem. Soc. 122 (1975) 255−259.


4. M.G. Chourashiya, L.D. Jadhav, Int. J. Hydrogen Energy 36 (2011) 14984−14995.

5. M.R. Kosinski, R.T. Baker, J. Power Sources 196 (2011) 2498−2512.

6. N. Chaubey, B.N. Wani, S.R. Bharadwaj, M.C. Chattopadhyaya, Solid State Sciences 20
(2013) 135−141.

7. S. Kim, J. Yu, D. Seo, I. Han, S. Woo, Int. J. Hydrogen Energy 37 (2012) 78−83.

8. Y. Zhang, S. He, L. Ge, M. Zhou, H. Chan, L. Guo, Int. J. Hydrogen Energy 36 (2011)
5128−5135.

9. V.V. Kharton, F.M. Figueiredo, L. Navarro, E.N. Naumovich, A.V. Kovalevsky, A.A.
Yaremchenko, A.P. Viskup, A. Carneiro, F.M. B. Marques, J.R. Frade, J. Mater. Sci. 36
(2001) 1105−1117.

10. L.J. Guang, I. Takayasu, M. Toshiyuki, Acta Mater. 52 (2004) 2221−2228.

11. M. Mogensen, N.M. Sammes, G.A. Tompsett, Solid State Ionics 129 (2000) 63−94.

12. B.C. H. Steele, Solid State Ionics 134 (2000) 3−11.

13. C. Min, H.K. Bok, X. Qing, K.A. Byeong, J.K. Woo, P.H. Duan, Ceram. Int. 35 (2009)
1335−1343.

14. R.V. Mangalaraja, S. Ananthakumar, A. Schachtsiek, M. Lopez, C.P. Camurri, R.E.


Avila, Mater. Sci. Eng. A. 527 (2010) 3645−3650.

15. M. Balaguer, C. Soils, J. M. Serra, J. Phys. Chem. C 116 (2012) 7975−7982.

16. R.O. Fuentes, R.T. Baker, Int. J. Hydrogen Energy 33 (2008) 3480−3484.

17. W. Chang, Y. Wang, C. Lin, S. Cheng, J. Power Sources 196 (2011) 1704−1711.

18. S. Ramesh, V.P. Kumar, P. Kistaiah, C.V. Reddy, Solid State Ionics 181 (2010) 86−91.

19. W. Chen, A. Navrotsky, J. Mater. Res. 21 (2006) 3242−3251.

20. J.P. Gonjal, R. Schmidt, J.E. Canuto, P.R. Alvarez, E. Moran, J. Power Sources 209

(2012) 163−171.
21. X. Guo, R. Waser, Prog. Mater. Sci. 51 (2006) 151−157.

22. T.S. Zhang, J. Ma, Y.Z. Chen, L.H. Luo, L.B. Kong, S.H. Chan, Solid State Ionics 177

(2006) 1227−1235.

23. E.Y. Pikalova, A.A. Murashkina, V.I. Maragou, A.K. Demin, V.N. Strekalovsky, P.E.

Tsiakaras, Int. J. Hydrogen Energy 36 (2011) 6175−6183.

24. K. Eguchi, T. Setoguchi, T. Inoue, H. Arai, Solid State Ionics 52 (1992) 165−171.

25. J.V. Herle, T. Horita, T. Kawada, N. Sakai, H. Yokokawa, M. Dokiya, Solid State Ionics

86 (1996) 1255−1258.

26. J.V. Herle, T. Horita, T. Kawada, N. Sakai, H. Yokokaya, M. Dokiya, Ceram. Int. 24

(1998) 229−241.

27. K.C. Anjaneya, G.P. Nayaka, J. Manjanna, G. Govindaraj, K.N. Ganesha, J. Alloys

Comp. 578 (2013) 53−59.

28. Y. Fu, S. Chen, J. Huang, Int. J. Hydrogen Energy 35 (2010) 745−752.

29. D.H. Prasad, S.Y. Park, H.I. Ji, H.R. Kim, J.W. Son, B.K. Kim, H.W. Lee, J.H. Lee, J.

Phys. Chem. C 116 (2012) 3467−3476.

30. M. Guo, J. Lu, Y. Wu, Y. Wang, M. Luo, Langmuir 27 (2011) 3872−3877.

31. G.B. Jung, T.J. Huang, M.H. Huang, C.L. Chang, J. Mater. Sci. 36 (2001) 5839−5844.

32. E.N.S. Muccillo, R.A. Rocha, R. Muccillo, Mate. Lett. 53 (2002) 353−358.

33. T.S. Zhang, J. Ma, H.T. Huang, P. Hing, Z.T. Xia, S.H. Chan, J.A. Kilner, Solid State

Sci. 5 (2003) 1505−1516.

34. W.H. Weber, K.C. Hass, J.R. McBride, Phys. Rev. B 48 (1993) 178−184.

35. S.J. Hong, A.V. Virkar, J. Am. Ceram. Soc. 78 (1995) 433−442.

36. T.H. Etsell, S.N. Flengas, Chem. Rev. 70 (1970) 339−341.


37. T.H. Hsieh, D.T. Ray, Y.P. Fu, Ceram. Int. 39 (2013) 7967−7973.

38. L.D. Jadhav, M.G. Chourashiya, A.P. Jamale, A.V. Chavan, S.P. Patil, J. Alloys Comp.

506 (2010) 739−744.

39. R. Suresh, V. Ponnuswamy, R. Mariappan, Appl. Surf. Sci. 273 (2013) 457−464.

40. M.J. Godinho, R.F. Gonclaves, L.P.S. Santos, J.A. Varela, E. Longo, E.R. Leite, Mater.

Lett. 61 (2007) 1904−1907.

41. B. Choudhury, A. Choudhury, Curr. Appl. Phys. 13 (2013) 217−223.

42. E.R. Trejo, J. Phys. Chem. Solids 74 (2013) 605−610.

43. G.R. Li, D.L. Qu, L. Arurault, Y.X. Tong, J. Phys. Chem. C 113 (2009) 1235−1241.

44. A. Askrabic, Z.D. Mitrovic, M. Radovic, M. Scepanovic, Z.V. Popovic, J. Raman

Spectrosco. 40 (2009) 650−655.

45. A. Nakajima, A. Yoshihara, M. Ishigame, Phys. Rev. B 50 (1994) 13297−13307.

46. J.R. Mcbridge, K.S. Hass, B.D. Poindexter, W.H. Weber, J. Appl Phys 76 1994

2435−2441.

47. A. G. Murillo, C.L. Luyer, C. Garapon, C. Dujardin, E. Bernstein, C. Pedrini, J. Mugnier,

Opt. Mater. 19 (2002) 161−168.

48. J.E. Bauerele, J. Phys. Chem. Solids. 30 (1969) 2657−2670.

49. G.B. Balazs, R.S. Glass, Solid State Ionics 76 (1995) 155−162.

50. M.G. Chourashiya, J.Y. Patil, S.H. Pawar, L.D. Jadhav, Mater. Chem. Phys. 109 (2008)

39−44.

51. H. Inaba, H. Tagawa, Solid State Ionics 83 (1996) 1−9.

52. K. Yamashita, K.V. Ramanujachary, M. Greenblatt, Solid State Ionics 81 (1995) 53−60.

53. B.S. Prakash, V.K.W. Grips, S.T. Aruna, J. Power Sources 214 (2012) 358−364.
54. T. Mori, T. Ikegami, H. Yamamura, J. Electrochem. Soc. 146 (1999) 4380−4387.

55. T. Mori, T. Ikegami, H. Yamamura, T. Atake, J. Therm. Anal. Calorim. 57 (1992)

831−840.

56. S. Banerjee, P.S. Devi, D. Topwal, S. Mandal, K. Menon, Adv. Funct. Mater. 17 (2007)

2847−2854.

57. J.A. Kilner, Solid State Ionics 8 (1983) 201−207.

 
Figure Captions

Fig. 1. XRD patterns of Ce0.8Ln0.2O2−δ (Ln = Y3+, Gd3+, Sm3+, Nd3+ and La3+) powders calcined
at 873 K for 4 h.

Fig. 2. Shift in (220) XRD peak position of Ce0.8Ln0.2O2−δ with different Ln cations.

Fig. 3. Measured and calculated lattice parameters and crystallite sizes of Ce0.8Ln0.2O2−δ as a
function of dopant cation radius.

Fig. 4. Measured density, oxygen vacancy and interstitial model density of Ce0.8Ln0.2O2−δ (Y3+,
Gd3+, Sm3+, Nd3+ and La3+) solid solutions as a function of dopant cation.

Fig. 5. FT−IR spectra of calcined Ce0.8Ln0.2O2−δ (Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid solutions.

Fig. 6. Raman spectra of Ce0.8Ln0.2O2−δ solid solutions sintered at 1473 K.

Fig. 7. (a) TEM image of SDC20 calcined at 873 K for 4 h, (b) SAED pattern of calcined
SDC20.

Fig. 8 (a−e). FE−SEM images of Ce0.8Ln0.2O2−δ (Y3+, Gd3+, Sm3+, Nd3+ and La3+) pellets sintered
1473 K. Fig. 8 (f) EDX spectra of sintered SDC20

Fig. 9 (a−c). The Nyquist plots for Ce0.8Ln0.2O2−δ (Y, Gd, Sm, Nd and La) pellets sintered at
1473 K and measured at 573, 598 and 623 K respectively.

Fig. 10. Arrhenius plot for Ce0.8Ln0.2O2−δ (Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid solutions
sintered at 1473 K.

Fig. 11. Total ionic conductivity of Ce0.8Ln0.2O2−δ at 623 K as a function of radius of dopant ion.
Ce0.8Ln0.2O2­δ

(111)

(220)

(311)
(200)

(331)

(422)
(400)
(122)

(420)
YDC20
Intensity (a.u.)
GDC20

SDC20

NDC20

LDC20

CeO2

20 30 40 50 60 70 80 90
2θ / deg.
 

Fig. 1.
(220)
Ce0.8Ln0.2O2­δ

YDC20
Intensity (a.u.)

GDC20

SDC20

NDC20

LDC20

CeO2

45 46 47 48 49 50
2θ / deg.
 

Fig. 2.
Lattice Parameter (angstrom)

5.48
La 24

Crystallite Size (nm)


5.46 Nd
22
5.44
Gd 20
Y Sm
5.42
Crystallite size
18
5.40 Measured
Calculated 16
5.38
1.02 1.04 1.06 1.08 1.10 1.12 1.14 1.16
Dopant Cation radius (angstrom)

Fig. 3

 
7.8
measured
vacancy
7.6 interstitial
Density (g/cm3)

7.4

7.2

7.0

6.8

6.6
Y Gd Sm Nd La
Dopant Cation
 

Fig. 4.
LDC20

NDC20
% Transmittance (a.u.)

SDC20

GDC20

YDC20

500 1000 1500 2000 2500 3000 3500 4000


Wavelength (nm)
 

Fig. 5.
LDC20

NDC20
Intensity (a.u)

SDC20

GDC20

YDC20

300 400 500 600 700


Raman Shift (cm­1)
 

Fig. 6.
Fig. 7 (a-b)

(a) 

20nm 

Fig. 7 (a)

(b) 

Fig. 7 (b)
Fig. 8 (a-f)

(a)  YDC20  (b)  GDC20 

   

(c)  SDC20  (d)  NDC20 

   

(e)  LDC20  (f)  SDC20 

 
Fig. 9 (a-c)

10000 6000
@ 573 K
YDC20
5000 LDC20
NDC20
SDC20
4000
GDC20
Z'' (Ohm)

8000 3000

2000

1000 YDC20
­ Z'' (Ohm)

6000 LDC20
0
0 1000 2000 3000 4000 5000 6000
Z' (Ohm) NDC20
f SDC20
4000 64 KHz GDC20
86 KHz

2000 @ 573 K
205 KHz

274 KHz

0
0 2000 4000 6000 8000 10000
Z' (Ohm)
 

Fig. 9 (a)
6000 3000
YDC20 @ 598 K @ 598 K YDC20
 LDC20
2500
 NDC20  LDC20
5000 2000
 GDC20
 SDC20  NDC20
Z'' (Ohm)

1500  GDC20
 SDC20
4000 1000
­ Z'' (Ohm)

500

3000 0
0 500 1000 1500 2000 2500 3000
Z' (Ohm)

154 KHz f
2000
178 KHz
1000
422 KHz
649 KHz
0
0 1000 2000 3000 4000 5000 6000
Z' (Ohm)
 

Fig. 9 (b)
8000 3000
 YDC20
  LDC20
@ 623 K
@ 623 K
7000 2500   NDC20
 YDC20
  GDC20

2000
  SDC20
  LDC20
6000

­ Z'' (Ohm)
f   NDC20
1500
  GDC20
5000 1000   SDC20
­ Z'' (Ohm)

4000 500

0
3000 0 500 1000 1500
Z' (Ohm)
2000 2500 3000

205 KHz
2000 f
237 KHz
487KHz 562 KHz
1000

0
0 1000 2000 3000 4000 5000 6000 7000 8000
Z' (Ohm)
 

Fig. 9 (c)
0
LDC20
NDC20
-1 GDC20
SDC20
log(σ T) [Scm K]

YDC20
-1

-2

-3

-4
Total conductivity
-5

-6
1.6 1.8 2.0 2.2 2.4
-1
1000/T [K ]
 

Fig. 10
-0.9 3+
Sm
-1.0
3+ 3+
Gd Nd
-1.1
log(σ *T) [Scm K]
-1

-1.2

-1.3
3+
-1.4 La

-1.5 Y
3+

-1.6
0.90 0.95 1.00 1.05 1.10 1.15 1.20 1.25
Dopant Ionic Radius (Angstrom)
 

Fig. 11
Table.1. Measured, calculated lattice parameters and crystallite sizes of Ce0.8Ln0.2O2−δ (Y3+,

d
Composition Sample ID Dopant Lattice parameter Crystallite
3
radius (Å) Size
+ (Å) Measured Calculated nm

,
Ce0.8Y0.2O2-δ YDC20 1.03 5.4103(3) 5.4140 21

Ce0.8Gd0.2O2-δ GDC20 1.05 5.4199(2) 5.4234 22


S
Ce0.8Sm0.2O2-δ SDC20 1.08 5.4325(1) 5.4370 24
m
Ce0.8
3 Nd0.2O2-δ NDC20 1.11 5.4533(5) 5.4510 19

Ce0.8
+ La0.2O2-δ LDC20 1.15 5.4721(4) 5.4695 16

, Nd3+ and La3+) as a function of dopant radius.


Table 2. Measured, calculated and relative densities of Ce0.8Ln0.2O2−δ as a function of

dopant radius.

Dopant Density (g/cm3) Relative


Sample ID radius Measured Calculated Density
(Å) (g/cm3)
YDC20 1.03 6.637(3) 6.732 98.58

GDC20 1.05 7.123(5) 7.216 98.78

SDC20 1.08 7.067(2) 7.125 99.22

NDC20 1.11 6.943(4) 7.027 98.80

LDC20 1.15 6.799(1) 6.884 98.76


Table 3. Radius ratio (rd/rh), effective index (Ei), conductivity (σt) and activation energy

(Ea) of Ce0.8Ln0.2O2−δ (Y3+, Gd3+, Sm3+, Nd3+ and La3+) solid solutions.

Radius Effective At 623 K


Sample ID ratio Index Conductivity Activation energy
rd / rh Ei σt Ea (eV)
YDC20 1.11 0.7833 4.65×10-5 1.02±0.01

GDC20 1.08 0.8006 4.41×10-4 0.87±0.01


SDC20 1.06 0.8279 9.73×10-4
0.86±0.01
NDC20 1.14 0.8584 2.30×10-4
0.89±0.01
-5
LDC20 1.18 0.8963 6.39×10
0.98±0.01
Studies on structural, morphological and electrical properties of Ce0.8Ln0.2O2−δ, (Ln= Y3+,

Gd3+, Sm3+, Nd3+ and La3+) solid solutions prepared by citrate complexation method

K.C. Anjaneyaa, G.P. Nayakaa, J. Manjannaa*, G. Govindarajb, K.N. Ganeshab

a
Department of Industrial Chemistry, Kuvempu University, Shankarghatta 577 451, India.
b
Department of Physics, Pondicherry University, Pondicherry, 605 001, India.

Graphical Abstract:

0 -0.9
3+
Sm
LDC20
-1.0
-1 NDC20 3+
Gd 3+
GDC20 Nd
-1.1
log(σ T) [Scm K]

SDC20 log(σ *T) [Scm K]


-1

-2
-1
YDC20
-1.2
-3
-1.3
3+
-4 La
-1.4
Total conductivity
-5 -1.5 3+
Y

-6 -1.6
0.90 0.95 1.00 1.05 1.10 1.15 1.20 1.25
1.6 1.8 2.0 2.2 2.4
-1 Dopant Ionic Radius (Angstrom)
1000/T [K ]
   
Highlights

► Ln doped CeO2 was successfully synthesized by citrate-complexation method.

► Unit cell parameter increases with increase in dopant radius.

► Densities based on oxygen vacancy model agreed with the experimental results.

► Raman spectroscopy confirms the formation of Ce0.8Ln0.2O2−δ solid solutions.

► Ce0.8Sm0.2O2−δ exhibits the highest oxygen ionic conductivity among the RE doped ceria.

► Total conductivity of the systems studied follows the theory of ionic radius ratio rd/rh.

S-ar putea să vă placă și