Colossal Dielectric Behavior and Relaxation in Nd-Doped BaTiO3 at Low

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Ceramics International 44 (2018) 7251–7258

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Colossal dielectric behavior and relaxation in Nd-doped BaTiO3 at low T


temperature
⁎ ⁎
Qiaoli Liua,b, Junwei Liub, Dayong Lub, , Weitao Zhenga,
a
School of Materials Science and Engineering, Key Laboratory of Mobile Materials, MOE, and State Key Laboratory of Superhard Materials, Jilin University, Changchun
130012, China
b
Research Center for Materials Science and Engineering, Jilin Institute of Chemical Technology, Jilin 132022, China

A R T I C L E I N F O A B S T R A C T

Keywords: Nd-doped BaTiO3 ceramics with Ti vacancy compensation were prepared by a solid-state reaction method, and
BaTiO3 the lattice structures were identified by XRD and Raman spectra. Samples with light doping possess a tetragonal
Nd-doping structure at room temperature, while the cubic phase appeared as the doping content increased. Dielectric
Dielectric relaxation properties at low temperature (−100–0 °C) were characterized by the dielectric constant, complex impedance
Defect complex
and electric modulus. Colossal dielectric constant and dielectric relaxation processes appeared in all samples.
Three electrical responses were observed in the low doping content ceramics, while four appeared in the high
doping content ceramics by a combined plot of Z″ and M″. These responses were attributed to the microstructure
inhomogeneity, which in turn can be ascribed to charge segregation when the doping content is high.
Localization of charge carriers was also observed in all of the ceramics, and the confined carriers preferred to
exist in the grain when the doping content was low, while localization would be more significant in grain
boundaries with increased dopant. The localization could be ascribed to the formation of defect complexes,
which on one hand maintained a high resistivity to the ceramics, and on the other hand, resulted in polarization
of the dipoles. Both Maxwell-Wagner polarization and dipole polarization were found to exist in the Nd-doped
BaTiO3 ceramics and were responsible for the colossal dielectric constant and relaxations. The space charge,
defect complexes, and microstructure are assumed to be responsible for the change in the dielectric constant and
the relaxation mechanism.

1. Introduction materials. Hopping of localized carriers produces a dipolar-type polar-


ization and is assumed to be responsible for CDC behavior and dielectric
Since CaCu3Ti4O12 (CCTO) was found to have a colossal dielectric relaxation [9–11].
constant (CDC) by Subramanian et al. in 2000 [1], potential techno- Apart from CCTO, giant permittivity has been observed in other
logical applications and the origin of CDC materials have attracted compounds, for example, the double perovskite structure [12,13], TiO2
significant attention. Unfortunately, the origin of the giant permittivity [14–16], and (Ba, Sr)TiO3 [17–20]. For BaTiO3 (BTO) ceramics, CDC
remains controversial, and the various models used to interpret CDC behavior is usually induced by a complex multistage process, involving
behavior can be divided into two categories: extrinsic effect and in- high temperature, reduced atmospheres, and limited diffusion of
trinsic mechanism. Although certain problems remain unresolved, the oxygen and dopant ions along grain boundaries [4]. Nano-crystalline
extrinsic nature, including spatial inhomogeneity, contact effect, and BTO ceramics prepared by the spark plasma sintering (SPS) technique
internal barrier layer capacitor (IBLC) are widely accepted as depicting have also been found to have a colossal constant [21–23]. Although
CDC behavior [2–5]. According to the IBLC model, the high dielectric CDC behavior and dielectric relaxation for BTO ceramics prepared by
constant can be attributed to the Maxwell-Wagner type interfacial po- conventional solid-state reaction method have been observed in recent
larization at the grain boundaries (GBs) [4]. The internal domains years, most studies have focused on CDC behavior and dielectric re-
within the grain have also been found to play an important role in CDC laxation at high temperatures [24–26]. Oxygen vacancies and related
materials [6–8]. On the other hand, the intrinsic mechanism is a pop- defects were assumed to be responsible for CDC and relaxation above
ular explanation for CDC. Polarons caused by oxygen deficiency or di- Curie temperature [25–28]. However, only a few studies exist on CDC
poles caused by a mixed-valent structure are convincing in many CDC and dielectric relaxation of BTO at low temperature, and the origins of


Corresponding authors.
E-mail addresses: [email protected] (D. Lu), [email protected] (W. Zheng).

https://doi.org/10.1016/j.ceramint.2018.01.181
Received 2 November 2017; Received in revised form 17 January 2018; Accepted 22 January 2018
Available online 01 February 2018
0272-8842/ © 2018 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

CDC and relaxation are not yet clear. Yu et al. observed the CDC phe-
nomenon and dielectric relaxation in a hexagonal BTO single crystal
below room temperature, where the high dielectric constant was as-
cribed to the Maxwell-Wagner effect [17]. Han et al. prepared BTO
ceramics using the SPS method and studied the origin of colossal per-
mittivity below room temperature, suggesting that the dielectric re-
laxation and giant permittivity originated from a combined effect: in-
terfacial and hopping polarization [22,23], whereas a single Maxwell-
Wagner effect was found to be dominant in La and Cr co-doped BTO
[29]. The published results suggest that the dielectric responses of BTO
may be more complicated at low temperature, and the defect types as
well as the movement of charge carriers need to be considered.
In order to obtain an improved understanding of CDC behavior and
dielectric relaxation of BTO at low temperature, in this work, we pre-
pared Nd-doped BTO by means of a solid-state reaction, and titanium
vacancy compensation was adopted to maintain the charge balance: Fig. 1. Room temperature XRD patterns of pure BaTiO3, Ba1-xNdxTi1-x/4O3 (a) and (002)/
(200) peaks indicating the structural evolution (b).
′ ′′′
4Ba×Ba + Ti×Ti ⇒ 4Nd•Ba + VTi

Rare-earth-doped BTO with titanium vacancy compensation has such as dielectric constant (ε), ac conductivity (σac ), impedance (Z) and
been studied by various authors [30–36]. Phase transition temperature modulus (M) of the ceramics were recorded with a broadband dielectric
was found to decrease with an increase in dopant, and the Curie peak spectrometer (Concept 41, Novocontrol Technologies, Germany) within
decreased at a rate of −24 °C/at% La [30,33], which contributed to the a range of temperature (−100–0 °C) and frequency (1 Hz–15 MHz).
distortion and tilting of nearby TiO6 octahedra caused by defect clusters
[36]. A high dielectric constant of 25,000 was obtained at La = 0.06 at 3. Results and discussion
curie temperature [33], and the diffuse phase transition also occurred
with heavy doping [33,34], which is usually related to the in- 3.1. Structural analysis by XRD and Raman spectra
homogeneous structure [37]. In this study, we focus on CDC behavior
and dielectric relaxation observed at low temperature. Four dielectric The room temperature XRD patterns of the pure and Nd-doped
responses are observed with an increase in doping content, implying BaTiO3 are illustrated in Fig. 1. Single-phase perovskite structure of
that the electrically heterogeneous structure is significantly different BaTiO3 was prepared, and no impurity phases could be detected within
from published results. The localization of charge carriers was also the XRD instrument limits. In fact, the solid solution of rare earth ions
confirmed, and its origin and effect on polarization and relaxation are in BaTiO3 depends on the compensation mechanisms: for Ba-site doping
discussed. with electronic compensation, Ti-rich phases such as Ba6Ti17O40 usually
exist [38,39], whereas the Ti-vacancy ionic compensation mechanism
2. Experimental results in a larger solid solubility (0–0.25 for La3+) [39]. Because Nd3+
(1.27 Å) has a smaller radius than La3+ (1.36 Å), it is reasonable that
Nd3+ ions completely entered the BaTiO3 lattice in our results. The
2.1. Preparation of Nd-doped BaTiO3 ceramics
change in the (002)/(200) peak files at 2θ ≈ 45° indicates the evolution
of the unit cell symmetry with the increase in Nd content. For pure
The ceramics were prepared by means of the conventional solid-
BaTiO3, the obvious splitting of the peak at 45° confirms the tetragonal
state reaction method. Starting materials (BaCO3, TiO2 and Nd2O3) with
structure. The intensity ratio of (002)/(200) decreases, and these two
high purity (≥ 99.9%) were weighted according to the formula Ba1-
peaks gradually collapse into one single (200) peak when x = 0.06,
xNdxTi1-x/4O3 (x = 0.02, 0.04 and 0.06). Precursor oxides were mixed
suggesting that the tetragonality decreases with the increase in Nd
in an agate mortar and calcined at 1100 °C for 5 h, following which the
content, and the structure transforms into a completely cubic phase for
calcined powder was milled again, mixed with a 10% polyvinyl alcohol
x = 0.06.
(PVA) solution, and uniaxially pressed into disk-shaped pellets (12 mm
The room temperature Raman spectra of Ba1-xNdxTi1-x/4O3 ceramics
diameter and approximately 2 mm thickness) at 200 MPa. Finally, these
from 100 to 1200 cm−1 are illustrated in Fig. 2. For pure BaTiO3, the
pellets were sintered at 1400 °C for 12 h in air and then cooled at a rate
tetragonal phase is usually characterized by sharp bands at ~ 170 and
of 200 °C/h. An undoped BaTiO3 sample was also prepared for com-
306 cm−1, with asymmetric broader bands at ~ 270, 520 and
parison.

2.2. Characterization

Powder X-ray diffraction (XRD) (DX-2700, Dandong, China) using


Cu Kα radiation was performed in order to confirm the crystal structure
of the sintered ceramics, with a step of 0.02° and count time of 3 s.
Room temperature Raman spectra of the pellets were recorded using a
LabRAM XploRA Raman spectrometer (Horiba Jobin Yvon), while
spectral excitation was provided by a Nd: YAG laser, using the 532 nm
line. The pellet microstructures were examined by means of scanning
electron microscopy (SEM) (EVO MA10, Zeiss, Germany) and energy-
dispersive X-ray spectrometry (EDS) (Aztec 2.3, Oxford, Britain). In
order to measure the dielectric properties, the ceramics were firstly
sputtered by gold layer, and then painted with silver paste on both sides
of the polished pellet as electrodes to ensure ohmic contact; the painted
electrodes were fired at 300 ̊C for 15 min. The dielectric characteristics, Fig. 2. Room temperature Raman spectra for BaTiO3 and Ba1-xNdxTi1-x/4O3 ceramics.

7252
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

720 cm−1 [40]. The latter three bands, which relate to the disorder of illustrated in Fig. 3(d). The EDS reveals that the cation ratios are almost
Ti displacements in the octahedron [41,42], are significantly broader the same as the designed stoichiometry within the error range. The Nd
and more symmetrical in the cubic paraelectric phase. From our results contents in the grain and grain boundary are found to be equal (1.9%)
in Fig. 2, bands of 242, 305, 515, 721, and 840 cm−1 can be observed. for x = 0.02, while the Nd content is slightly higher than 6% in the
The sharp 305 cm−1 peak related to the tetragonal phase exists in pure grain boundary for x = 0.06, indicating that it is an Nd-rich region.
BaTiO3, x = 0.02 and 0.04 samples, but becomes weaker and broader Although EDS is a rough tool for estimating element contents, parti-
with an increased Nd content, following which this peak disappears in x cularly when the proportion is low, the space charge segregation has
= 0.06. The 721 cm−1 band, which is a further characteristic of the been observed by numerous authors [43,44]. When donor or acceptor
tetragonal phase, is also broadened in x = 0.06. These results confirm segregation occurs in the grain boundary, a compositional and elec-
the structural transformation from tetragonal to cubic, and are con- trical inhomogeneity will be formed; thus, the total sample resistivity
sistent with the XRD patterns. A new symmetrical bands is observed at can be represented by the resistance of the grain, potential barrier, and
~ 840 cm−1, and its intensity increases with Nd content. The 840 cm−1 the dopant-segregation-induced insulating region [45]. The electrical
band is attributed to an A1g octahedral breathing mode, and is sensitive inhomogeneity can more easily be characterized by impedance spec-
to the perovskite structure octahedra [42], suggesting distortion of the troscopy [32], and is discussed in the following section.
TiO6 octahedra. In A-site doped BaTiO3 with rare earth, three charge
compensation mechanisms may exist: ionic compensation with Ba and/
3.3. Dielectric properties
or Ti vacancies, electronic compensation with free electrons, and only
Ti vacancies may have a direct and significant influence on the TiO6
The frequency responses of the dielectric constant (εˊ) and dielectric
octahedra; thus, the 840 cm−1 band is direct proof of the dominant Ti
loss (tan δ) for Ba1-xNdxTi1-x/4O3 ceramics at selected temperatures are
vacancy compensation mechanism in our results.
displayed in Fig. 4(a)–(c). It is clearly indicated that all ceramics exhibit
a considerably high dielectric constant (105–106), which is almost fre-
3.2. Microstructure and element analysis by SEM and EDS quency independent in the low-frequency range. The dielectric constant
decreases rapidly to a lower plateau (about 103) at a high frequency,
Fig. 3(a)–(c) display the surface morphologies of Ba1-xNdxTi1-x/4O3 accompanied by a relaxation peak (peak 1 in Fig. 4) in dielectric loss. A
ceramics, where the microstructure reveals the well-sintered nature new relaxation peak (peak 2 in Fig. 4(b) and (c)) is observed at high
with different grain sizes. Sample x = 0.02 exhibits a uniform grain size frequency as the Nd content increases, while the peak is almost sub-
of ~ 1 µm, while larger grains (2–4 µm) and higher density are observed merged in peak 1 at x = 0.04 and shifts to a higher frequency (105–107
in x = 0.04, and the grain size decreases to 1–2 µm when x = 0.06. It is Hz) at x = 0.06. These two relaxation peak sets separate from one
worth noting that the grain boundary layer becomes continuous in x = another completely at x =0.06. A comparison of the dielectric constant
0.04, and can hardly be distinguished from the grain for x = 0.06, for pure BaTiO3 and Ba1-xNdxTi1-x/4O3 at −100 °C is also presented in
indicating that a liquid phase may exist in the grain boundary. This Fig. 4(d). The colossal dielectric behavior only emerges in Nd-doped
result suggests that the eutectic temperature is lowered by heavy Nd samples, and the pure BaTiO3 exhibits only an intrinsic polarization
doping. However, no secondary phases were detected by the XRD and with a dielectric constant of 2000–2500 depending on the frequency,
back scattered electron imaging (BSE, not shown) results. As the Max- suggesting that Nd ions play a decisive role in the colossal dielectric
well-Wagner polarization is closely related to the grain boundary constant. The dielectric plateaus are enhanced with the temperature
structure, the liquid phase is expected to have a significant influence on increasing for all Nd-doped ceramics, which has also been observed in
the dielectric properties. EDS was also used to examine the cation perovskite-type oxides [46] and transition metal oxides [47]. As de-
content in the Ba1-xNdxTi1-x/4O3 ceramics, and the result for x = 0.06 is monstrated in the following discussion, the colossal dielectric plateaus

Fig. 3. SEM images for Ba1-xNdxTi1-x/4O3 (a−c) and


EDS of x = 0.06 (d).

7253
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

Fig. 5. Arrhenius plots of relaxation peaks for Ba1-xNdxTi1-x/4O3 ceramics.

fmax = f0 exp(−Ea/ κB T ),

where Ea is the activation energy required for the thermally activated


process, f0 represents the pre-exponential factor, κB is the Boltzmann
constant, and T is the absolute temperature. The activation energy can
be calculated from the plot slopes of lnfmax versus 1000/T, as illustrated
in Fig. 5. The activation energy values of relaxation 1 are 0.17 eV,
0.13 eV and 0.06 eV for x = 0.02, 0.04 and 0.06, respectively, while the
values for relaxation 2 are 0.13 eV and 0.08 eV for x = 0.04 and 0.06,
respectively. The activation energy decreases with the doping level,
which may be attributed to the high concentration and mobility of
charge carriers in heavy doping samples. It is worth noting that the
values of fmax in peak 1 of x = 0.04 and 0.06 deviates from the Ar-
rhenius law in a higher temperature range (> −40 °C), which implies
that more complicated relaxation mechanisms in these two samples.
Four mechanisms are usually responsible for polarization in di-
electric materials: electronic, ionic, orientational and space charge or
interfacial polarization. The electronic and ionic mechanisms are re-
lated to the elastic displacement of the valence electron clouds and
displacement of the relative ion positions, the former occurs at the
optical frequency and the latter in the infrared region. For BaTiO3, the
dielectric constant corresponding to electronic and ionic polarization is
significantly lower than the observed colossal dielectric constant
(> 105) at a low frequency, and can be neglected in our discussion.
However, the dielectric properties in the low-frequency regime offer a
great deal of information concerning the impurity, defect, and space-
charge carrier transport processes [25]. The interfacial or space charge
polarization is also known as Maxwell-Wagner polarization. According
to the widely accepted IBLC model, which is based on Maxwell-Wagner
polarization, the CDC phenomenon arises from an electric in-
homogeneous structure of the composition. The giant dielectric con-
stant at low frequencies is attributed to the grain boundary barrier layer
with high resistance, while the low permittivity at high frequencies is
associated with low-resistance grain [4]. The inhomogeneous structure
can be effectively depicted by the impedance complex plane plot, which
offers a powerful tool for determining the dielectric contributions from
Fig. 4. Frequency dependence of dielectric constant and loss of Ba1-xNdxTi1-x/4O3 at se- the grain interiors, grain boundaries and sample-electrode contacts
lected temperatures (a−c) and comparison of pure and doped BaTiO3 at −100 °C (d).
[48]. In the complex impedance plane plots of Z′ versus Z″ (where Z′
and Z″ are the real and imaginary parts of the complex impedance Z*)
can be attributed to Maxwell-Wagner and dipole polarization. At low at a certain temperature, a semicircle represents an electric homo-
temperature, the space charge and dipoles may freeze and thus exhibit geneous structure. The high frequency arc is usually attributed to bulk
lower activity than that at a high temperature, while the polarization conduction, while the lower frequency arc is related to the grain-
rate formed is rapid and thus relaxation occurs in a high frequency boundary conduction, and a third arc at an even lower frequency will
when the temperature is high [47]. The relaxation peaks move towards be observed if the sample-electrode contact effect cannot be neglected
higher frequencies with increased temperature in all samples, in- [49]. Additional arc(s) will exist at mid and/or high frequencies if the
dicating a thermally activated dielectric relaxation process. The ther- grain is inhomogeneous or the domain boundary exists [32,50]. The
mally activated relaxation peak positions (fmax) of tan δ are found to impedance complex plane plots for Ba1-xNdxTi1-x/4O3 at selected tem-
obey the Arrhenius law: peratures are presented in Fig. 6(a)–(c). Two obvious arcs exist at x =
0.02 and x = 0.04: a well-developed semicircle at high frequencies and

7254
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

Fig. 7. Equivalent circuit of electrical properties of grain and grain boundary effects.

Z * = Z ′ − Z ′′
Rg Rgb
Z′ = +
1 + (ωRg Cg )2 1 + (ωRgb Cgb)2

and

ωRg Cg ωRgb Cgb


Z ′′ = Rg ⎡ ⎤+R ⎡
gb ⎢
⎤,
⎢ 1 + (ωRg Cg )2 ⎥ 1 + (ωRgb Cgb)2 ⎥
⎣ ⎦ ⎣ ⎦
where Rg, Rgb and Cg, Cgb are the resistances and capacitances of the
grain and grain boundary, respectively.
We fitted our experimental data using the ZView impedance soft-
ware at different temperatures. Because the low-frequency resistances
are too high for our instrument to obtain a well-developed semicircle, it
is difficult to achieve effective fitting of the low-frequency arcs. In fact,
the semicircles of grain boundary are more clearly visible at higher
temperatures, but often strongly overlap with a non-ohmic electrode
interface relaxation [51,52]. Moreover, the Ba1-xNdxTi1-x/4O3 possesses
a cubic structure at high temperatures, which differs from the orthor-
hombic structure at low temperatures. These drawbacks prevent us
from extracting reliable GB conductivity values. Fortunately, the
smaller semicircles at high frequencies enable us to extract accurate
resistances values. The high frequency arcs for x = 0.02 and 0.04 were
fitted by using an equivalent circuit containing a resistor and capacitor,
while x = 0.06 was fitted with two RC elements for the two semicircles
at high frequency. The resistances for the ceramics at various tem-
peratures can be fitted effectively by the Arrhenius law, the results of
which are illustrated in Fig. 8. The activation energy values are 0.19 eV
for x = 0.02, 0.12 eV for x = 0.04, and 0.12 eV (semicircle 1, inset of
Fig. 6(c)) and 0.08 eV (semicircle 2, inset of Fig. 6(c)) for x = 0.06.
These values are similar to the fitting results of relaxation peaks in
Fig. 5. As addressed by Liu et al., the activation energy obtained from
Fig. 6. Complex impedance plane (Z*) plots for x = 0.02 (a), x = 0.04 (b), and x = 0.06
relaxation peaks is close to that of grains [53]; therefore, the high-
(c). The inset of (c) illustrates the high-frequency semicircles of x = 0.06. frequency arcs represent the contribution of grain, while the multi-arcs
at high frequency may be related to an inhomogeneous grain. These
impedance results suggest a more complicated inhomogeneous struc-
a large arc with a nonzero intercept at low frequencies, indicating the
ture in heavy doped ceramics.
inhomogeneous structure in the samples. The high frequency semi-
circles are fully developed with a small diameter (approximately 105
and 7.5 × 104 for x = 0.02 and 0.04 at −100 °C), while the low-
frequency arcs are significantly larger (> 107 diameter for x = 0.02
and 0.04, > 106 for x = 0.06). With an increase in temperature, the
arcs become progressively smaller. The results of x = 0.02 and 0.04
suggest that both semiconductive bulk (grain) and an insulating inter-
face are contained in the ceramics. In sample x = 0.06, three obvious
arcs are observed, as shown in Fig. 6(c), two small well-developed
semicircles at high frequencies (semicircles of 1 and 2 in inset), and a
large arc at low frequencies, which means that a new dielectric re-
sponse appears as the doping content increases.
The impedance data of inhomogeneous structures are often modeled
by an equivalent circuit consisting of two or more parallel RC elements
connected in serial. If the dielectric response is contributed by the grain
and grain boundary, the equivalent circuit will contain two RC ele-
ments, as illustrated in Fig. 7. The complex impedance in this condition
Fig. 8. Arrhenius plot of resistances obtained from the high-frequency semicircles of
may be expressed by the following equations: complex impedance plane plots.

7255
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

boundary and is responsible for the high resistivity and giant permit-
tivity at low frequencies, because the domain boundary area per vo-
lume is larger than that of the grain boundary [8]. However, from the
results of Shao et al., this domain contribution appears only at a high
temperature, and the corresponding relaxation was not observed at low
temperatures. Based on the above, we attribute the electric response at
low frequencies to the grain boundary contribution. The M″ data of x =
0.02 indicate two low-resistance components at mid frequencies and
high frequencies: one at 5 × 103 Hz and the other at > 107 Hz
(Fig. 9(a)). The combined Z″ and M″ plot indicates that three responses
exist for x = 0.02, which at first glance is inconsistent with the result of
Fig. 6(a). In fact, when further enlarging the high frequency region, we
could identify a small nonzero intercept on Z′ (approximately several
ohms, inset of Fig. 6(a)), indicating that a highly conductive bulk region
exists in the complex impedance plane plot. West et al. pointed out that,
in La-doped BaTiO3 ceramics, the grain exhibits an inhomogeneous
structure containing two different parts: the outer and interior grain
region. The high frequency response in M″ can be attributed to a
semiconducting grain core, while the mid frequency response is related
to the outer grain region [32,33]. In our results, the small resistance at a
high frequency confirms the existence of a grain core. From the fre-
quency-dependence plot of the dielectric constant in Fig. 4(a), only two
contributions can be observed at low and high frequencies, respec-
tively, suggesting that the grain core effect can be ignored and the at-
tribution is mainly dominated by the outer region. If we ignore the
existence of the small grain core, we can assign the outer region as the
entire grain and x = 0.02 exhibits a homogeneous grain structure.
Furthermore, we ignore the small core effects of x = 0.04 and 0.06 in
the following discussion. For x = 0.04, a new electric response ap-
peared in M″ at 105 Hz, accompanied by a decreasing of the peak at
103–104 Hz. The combined M″ and Z″ plot of x = 0.04 suggests that
four inhomogeneous structures exist in Ba1-xNdxTi1-x/4O3 ceramics.
Moreover, the low- and high-frequency responses originate from the
grain boundary and highly conductive core, while the other two re-
sponses at mid frequencies imply that inhomogeneity appears in the
grain compared with x = 002: an outer or surface region with larger
resistance and an inner region with smaller resistance may exist in high
doping ceramics. Thus, from the combined M″ and Z″ plot, the in-
homogeneous structures of Ba1-xNdxTi1-x/4O3 ceramics could be inter-
Fig. 9. Combined Z″ and M″ plot of x = 0.02 (a), x = 0.04 (b), and x = 0.06 (c) at
preted as follows: the grain core has the smallest resistance of several
−100 °C. The inset of (c) shows an expanded view of the plots at 102–105 Hz. The dashed ohms and could hardly be distinguished in the complex impedance
lines are guidelines for the eye of the relaxation peaks of Z″ and M″. plane plot, while the grain core has little effect on polarization and
relaxation and can thus be ignored. The grain exhibits a nearly homo-
In order to gain an improved understanding of the inhomogeneous geneous structure at a low doping content and is separated into two
structure and the dielectric response contributions, we applied the different components for high doping samples, the outer region pos-
combined impedance and electric modulus (M*, extracted from the sesses a larger resistivity and a corresponding electric response is ob-
permittivity data using the relation: M* = M′ + iM″ = 1 / ε*.) analysis served at lower frequency, while the inner grain response appears at
in order to clarify the relaxation origins. The combined plot provides a higher frequency. The inhomogeneity implies that the Nd ions inside
useful tool for analyzing the impedance data of electroceramics. Each the grain are not uniform, which is difficult to distinguish by means of
RC element results in a Debye peak in the Z″ and M″ spectroscopic plots XRD and Raman spectra.
[48]. According to the IBLC model, Z″ is dominant according to the From Fig. 9, we can also observe the mismatch between Z″ and M″
most resistive element, namely, the grain-boundary component Rgb, peaks. This insignificant mismatch indicates that the charge carrier
whereas M″ is dominated by the element with a small capacitance, movements are not completely long range, and short-range migration
namely, the bulk element Cb. The Z″ and M″ peaks should be coincident also exists [56,58,59]. In order to clarify the nature of the short range
on the frequency scale if the migration of charge carriers is long range, movement, the universal dielectric response (UDR) model is adopted to
otherwise, the short range movement of the charge carriers will be study the dielectric response caused by localized charge carriers [60].
dominant [54–56]. Because the relaxation peak related to the smallest According to this model, ε′ can be calculated as
capacitance grain in Z″ and the peak of most resistive grain boundary in ε′ = tan(sπ /2) σ0 f s − 1 / ε0,
M″ are too weak to be distinguished, the combined Z″ and M″ plot is
powerful in providing a comprehensive depiction of the dielectric re- where σ0 and s are temperature-dependent constants, ε0 is the electric
sponse. Fig. 9 displays the combined Z″ and M″ plot of the ceramics at permittivity of free space, and f denotes the experimental frequency (f
−100 °C. The plot indicates that Z″ is dominated by a low frequency = ω / 2π). The above equation can be rewritten as
incline with an associated high resistivity, which is attributed to the fε′ = A (T ) f s ,
insulating grain boundary region by Morrison et al. [57]. Shao et al.
claimed that the domain boundary has a larger resistance than the grain where A(T) = tan(sπ / 2)σ0 / ε0 is a temperature-dependent constant.
Therefore, a straight line with a slope of s should be obtained if

7256
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

tendency to exist in grain, while the localization is more obvious at


grain boundaries for x = 0.06. When the doping level is low, most Nd
ions can enter the BaTiO3 lattice and are localized due to the com-
pensation of Ti vacancies, while a large amount of carriers will gather at
the outer grain and grain boundary for heavily doped samples. When
different defect types exist, various defect associations occur, as con-
firmed by other researchers, for example Mn′Ti′ −V ··O [61] V ··O−2Ti3Ti+, [25],
Mn′Ti′ −2La·Ba [62], Ti′Ti−La·Ba [63], and Mn′Ti′ −2La·Ba [64]. As proven from
our Raman results, Ti vacancy compensation, rather than other com-
pensation mechanisms, exists in Ba1-xNdxTi1-x/4O3 ceramics; therefore,
Nd·Ba and VTi ′ ′′′ may form the charge complex 4Nd·Ba−VTi ′ ′′′, and these
complexes may lead to localization of carriers, while also resulting in
dipoles. The dipoles rotate along the applied electric field direction and
are responsible for polarization of the Ba1-xNdxTi1-x/4O3 ceramics. We
can interpret the dielectric plateau of the samples at low frequencies. A
Maxwell-Wagner polarization in the grain boundary dominates the
plateau of x=0.02 at low temperatures with the Nd doping content
increasing, and Nd ions and Ti vacancies will associate with one an-
other to produce 4Nd·Ba−VTi ′ ′′′ defect complexes. These defect complexes,
which are the source of dipole polarization, partly reduce the space
charge contributing to the Maxwell-Wagner polarization, resulting in a
reduction in the dielectric constant in x = 0.04. Moreover, the anom-
alous large grain in x = 0.04 (Fig. 3) will decrease the amount of grain
boundaries, which may be another reason for the lower dielectric
constant. For x = 0.06, a high concentration of space charge exists in
the grain boundaries because of charge segregation, leading to a higher
dielectric constant plateau and Maxwell-Wagner relaxation. The new
dielectric constant plateau and relaxation in x = 0.04 and x = 0.06 at
higher frequencies may be related to the defect complex dipoles. These
results imply that both space charge and dipole polarization exist in
Ba1-xNdxTi1-x/4O3 ceramics, and this combined effect is consistent with
the published results [22].

4. Conclusions

Colossal dielectric constant behavior and dielectric relaxations were


observed in Nd-doped BaTiO3 ceramics. The combined plot of Z″ and
M″ provides a comprehensive depiction of electric responses, from
which the grain inhomogeneity could be confirmed. The highly con-
ductive grain core was determined, and its contribution to polarization
and relaxation could be ignored. The outer grain exhibits a nearly
uniform structure in the low doping sample and separates into two
components: surface area and inner region. The slight mismatch be-
tween Z″ and M″ peaks suggests that the carriers are partly localized,
Fig. 10. Log10(ε′f) vs. log10f plot for x = 0.02 (a), x = 0.04 (b), and x = 0.06 (c) at ′ ′′′
and this localization may be a result of the formation of the 4Nd·Ba−VTi
selected temperatures. defect complex. The defect complexes as well as microstructure in the
Nd-doped BaTiO3 ceramics, play an important role in the formation of
log10(fε′) is plotted as a function of log10f at a given temperature. dipoles and distribution of space charges. Both Maxwell-Wagner and
Fig. 10 represents the log10(fε′) versus log10f plot for the ceramics at dipole polarization were observed in our samples.
different temperatures. Straight lines appear at high temperatures and
low frequencies, the data deviate from the linear behavior at higher Acknowledgments
frequencies because of the relaxation process, and a straight line ap-
pears again at low temperatures and high frequencies, with a similar The authors wish to acknowledge financial support from the
value of slope s. The fitted values of s at low and high frequencies are National Natural Science Foundation of China (21271084) and of Jilin
0.94 and 0.93 for x = 0.02, 0.92 and 0.91 for x = 0.04, and 0.97 for x Province (20160101290JC), and Changbai Mountain Scholar
= 0.06, respectively. It is also determined that the low-frequency linear Distinguished Professor (2015047).
behavior is only significant at a high temperature for x = 0.02, while
the straight lines can be obtained in a wider temperature and frequency References
range for x = 0.06, which indicates that the charge carriers of x = 0.06
are more localized than those of x = 0.02 at low temperatures and [1] M.A. Subramanian, D. Li, N. Duan, B.A. Reisner, A.W. Sleight, High dielectric
constant in ACu3Ti4O12 and ACu3Ti3FeO12 phases, J. Solid State Chem. 151 (2000)
frequencies. This situation is contrary to that in high frequencies, where 323–325.
x = 0.02 exhibits a more localized state. Because the low frequency [2] M.H. Cohen, J.B. Neaton, L. He, D. Vanderbilt, Extrinsic models for the dielectric
dielectric response represents the effects of the grain boundaries and response of CaCu3Ti4O12, J. Appl. Phys. 94 (2003) 3299–3306.
[3] P. Lunkenheimer, V. Bobnar, A.V. Pronin, A.I. Ritus, A.A. Volkov, A. Loidl, Origin of
high frequency responses are attributed to the grain, the results of apparent colossal dielectric constants, Phys. Rev. B 66 (2002) 052105.
Fig. 10 imply that the localized carriers of x = 0.02 demonstrate a [4] D.C. Sinclair, T.B. Adams, F.D. Morrison, A.R. West, CaCu3Ti4O12: one-step internal

7257
Q. Liu et al. Ceramics International 44 (2018) 7251–7258

barrier layer capacitor, Appl. Phys. Lett. 80 (2002) 2153. 509 (2011) 10040–10049.
[5] T.B. Adams, D.C. Sinclair, A.R. West, Giant barrier layer capacitance effects in [35] C.L. Freeman, J.A. Dawson, H.-R. Chen, L. Ben, J.H. Harding, F.D. Morrison,
CaCu3Ti4O12 ceramics, Adv. Mater. 14 (2002) 1321–1323. D.C. Sinclair, A.R. West, Energetics of donor-doping, metal vacancies, and oxygen-
[6] T.-T. Fang, C.P. Liu, Evidence of the internal domains for inducing the anomalously Loss in A-site rare-earth-doped BaTiO3, Adv. Funct. Mater. 23 (2013) 3925–3928.
high dielectric constant of CaCu3Ti4O12, Chem. Mater. 17 (2005) 5167–5171. [36] C.L. Freeman, J.A. Dawson, J.H. Harding, L.-B. Ben, D.C. Sinclair, The influence of
[7] T. Fang, L. Mei, H. Ho, Effects of Cu stoichiometry on the microstructures, barrier- A-site rare earth ion size in controlling the curie temperature of Ba1−xRExTi1−x/
layer structures, electrical conduction, dielectric responses, and stability of 4O3, Adv. Funct. Mater. 23 (2013) 491–495.
CaCu3Ti4O12, Acta Mater. 54 (2006) 2867–2875. [37] I. Fujii, M. Ugorek, S. Trolier-McKinstry, Grain size effect on the dielectric non-
[8] S.F. Shao, J.L. Zhang, P. Zheng, W.L. Zhong, C.L. Wang, Microstructure and elec- linearity of BaTiO3 ceramics, J. Appl. Phys. 107 (2010) 104116.
trical properties of CaCu3Ti4O12 ceramics, J. Appl. Phys. 99 (2006) 084106. [38] D.M. Smyth, The defect chemistry of donor-doped BaTiO3: a rebuttal, J.
[9] L. Ni, X.M. Chen, Enhancement of giant dielectric response in CaCu3Ti4O12 ceramics Electroceram. 9 (2002) 179–186.
by Zn substitution, J. Am. Ceram. Soc. 93 (2010) 184–189. [39] F.D. Morrison, D.C. Sinclair, A.R. West, Electrical and structural characteristics of
[10] L. Zhang, Z.-J. Tang, Polaron relaxation and variable-range-hopping conductivity in lanthanum-doped barium titanate ceramics, J. Appl. Phys. 86 (1999) 6355–6366.
the giant-dielectric-constant material CaCu3Ti4O12, Phys. Rev. B 70 (2004). [40] U.D. Venkateswaran, V.M. Naik, R. Naik, High-pressure Raman studies of poly-
[11] C.C. Wang, L.W. Zhang, Polaron relaxation related to localized charge carriers in crystalline BaTiO3, Phys. Rev. B 58 (1998) 14256–14260.
CaCu3Ti4O12, Appl. Phys. Lett. 90 (2007) 142905. [41] M. DiDomenico, S.H. Wemple, S.P.S. Porto, R.P. Bauman, Raman spectrum of
[12] G. Wang, C. Wang, S. Huang, C. Lei, X. Sun, T. Li, L. Liu, J. Jones, Origin of colossal single-domain BaTiO3, Phys. Rev. 174 (1968) 522–530.
dielectric behavior in double perovskite Ba2CoNbO6, J. Am. Ceram. Soc. 96 (2013) [42] J. Pokorný, U.M. Pasha, L. Ben, O.P. Thakur, D.C. Sinclair, I.M. Reaney, Use of
2203–2210. Raman spectroscopy to determine the site occupancy of dopants in BaTiO3, J. Appl.
[13] W.Z. Yang, M.S. Fu, X.Q. Liu, H.Y. Zhu, X.M. Chen, Giant dielectric response and Phys. 109 (2011) 114110.
mixed-valent structure in the layered-ordered double-perovskite ceramics, Ceram. [43] S.H. Yoon, H. Kim, Space charge segregation during the cooling process and its
Int. 37 (2011) 2747–2753. effect on the grain boundary impedance in Nb-doped BaTiO3, J. Eur. Ceram. Soc. 22
[14] W. Hu, Y. Liu, R.L. Withers, T.J. Frankcombe, L. Noren, A. Snashall, M. Kitchin, (2002) 689–696.
P. Smith, B. Gong, H. Chen, J. Schiemer, F. Brink, J. Wong-Leung, Electron-pinned [44] N. Wilcox, V. Ravikumar, R.P. Rodrigues, V.P. Dravid, M. Vollmann, R. Waser,
defect-dipoles for high-performance colossal permittivity materials, Nat. Mater. 12 K.K. Soni, A.G. Adriaens, Investigation of grain boundary segregation in acceptor
(2013) 821–826. and donor doped strontium titanate, Solid State Ion. 75 (1995) 127–136.
[15] H. Han, P. Dufour, S. Mhin, J.H. Ryu, C. Tenailleau, S. Guillemet-Fritsch, Quasi- [45] S.B. Desu, D.A. Payne, Interfacial segregation in perovskites III, microstructure and
intrinsic colossal permittivity in Nb and In co-doped rutile TiO2 nanoceramics electrical properties, J. Am. Ceram. Soc. 73 (1990) 3407–3415.
synthesized through a oxalate chemical-solution route combined with spark plasma [46] W. Somphan, N. Sangwong, T. Yamwong, P. Thongbai, Giant dielectric and elec-
sintering, Phys. Chem. Chem. Phys. 17 (2015) 16864–16875. trical properties of sodium yttrium copper titanate: Na1/2Y1/2Cu3Ti4O12, J. Mater.
[16] W. Hu, K. Lau, Y. Liu, R.L. Withers, H. Chen, L. Fu, B. Gong, W. Hutchison, Colossal Sci.: Mater. Electron. 23 (2011) 1229–1234.
dielectric permittivity in (Nb+Al) codoped rutile TiO2 ceramics: compositional [47] Y. Lin, L. Jiang, R. Zhao, C.-W. Nan, High-permittivity core/shell structured NiO-
gradient and local structure, Chem. Mater. 27 (2015) 4934–4942. based ceramics and their dielectric response mechanism, Phys. Rev. B 72 (2005)
[17] J. Yu, P.-F. Paradis, T. Ishikawa, S. Yoda, Maxwell–Wagner effect in hexagonal 014103.
BaTiO3 single crystals grown by containerless processing, Appl. Phys. Lett. 85 [48] M. Li, A. Feteira, D.C. Sinclair, Origin of the high permittivity in (La0.4Ba0.4Ca0.2)
(2004) 2899–2901. (Mn0.4Ti0.6)O3 ceramics, J. Appl. Phys. 98 (2005) 084101.
[18] Q. Sun, Q. Gu, K. Zhu, R. Jin, J. Liu, J. Wang, J. Qiu, Crystalline structure, defect [49] K. Meeporn, T. Yamwong, S. Pinitsoontorn, V. Amornkitbamrung, P. Thongbai,
chemistry and room temperature colossal permittivity of Nd-doped barium titanate, Grain size independence of giant dielectric permittivity of CaCu3Ti4−xScxO12
Sci. Rep. 7 (2017) 42274. ceramics, Ceram. Int. 40 (2014) 15897–15906.
[19] Z. Wang, M. Cao, Z. Yao, Q. Zhang, Z. Song, W. Hu, Q. Xu, H. Hao, H. Liu, Z. Yu, [50] W. Li, R.W. Schwartz, Ac conductivity relaxation processes in CaCu3Ti4O12 cera-
Giant permittivity and low dielectric loss of SrTiO3 ceramics sintered in nitrogen mics: grain boundary and domain boundary effects, Appl. Phys. Lett. 89 (2006)
atmosphere, J. Eur. Ceram. Soc. 34 (2014) 1755–1760. 242906.
[20] S. Yan, C. Mao, G. Wang, C. Yao, F. Cao, X. Dong, High temperature dielectric [51] R. Schmidt, M.C. Stennett, N.C. Hyatt, J. Pokorny, J. Prado-Gonjal, M. Li,
relaxation anomaly of Y3+ and Mn2+ doped barium strontium titanate ceramics, J. D.C. Sinclair, Effects of sintering temperature on the internal barrier layer capacitor
Appl. Phys. 116 (2014) 144103. (IBLC) structure in CaCu3Ti4O12 (CCTO) ceramics, J. Eur. Ceram. Soc. 32 (2012)
[21] S. Guillemet-Fritsch, Z. Valdez-Nava, C. Tenailleau, T. Lebey, B. Durand, 3313–3323.
J.Y. Chane-Ching, Colossal permittivity in ultrafine grain size BaTiO3–x and [52] A.R. West, D.C. Sinclair, N. Hirose, Characterization of electrical materials, espe-
Ba0.95La0.05TiO3–x materials, Adv. Mater. 20 (2008) 551–555. cially ferroelectrics, by impedance spectroscopy, J. Electroceram. 1 (1997) 65–71.
[22] H. Han, C. Voisin, S. Guillemet-Fritsch, P. Dufour, C. Tenailleau, C. Turner, [53] J. Liu, C. Duan, W. Yin, W.N. Mei, R.W. Smith, J.R. Hardy, Large dielectric constant
J.C. Nino, Origin of colossal permittivity in BaTiO3 via broadband dielectric spec- and Maxwell-Wagner relaxation in Bi2∕3Cu3Ti4O12, Phys. Rev. B 70 (2004).
troscopy, J. Appl. Phys. 113 (2013) 024102. [54] D.C. Sinclair, A.R. West, Impedance and modulus spectroscopy of semiconducting
[23] H. Han, C. Davis, J.C. Nino, Variable range hopping conduction in BaTiO3 ceramics BaTiO3 showing positive temperature coefficient of resistance, J. Appl. Phys. 66
exhibiting colossal permittivity, J. Phys. Chem. C 118 (2014) 9137–9142. (1989) 3850–3856.
[24] J. Tao, Z. Yi, Y. Liu, M. Zhang, J. Zhai, X. Tan, Dielectric tunability, dielectric re- [55] M.A.L. Nobre, S. Lanfredi, Grain boundary electric characterization of Zn7Sb2O12
laxation, and impedance spectroscopic studies on (Ba0.85Ca0.15)(Ti0.9Zr0.1)O3 lead- semiconducting ceramic: a negative temperature coefficient thermistor, J. Appl.
free ceramics, J. Am. Ceram. Soc. 96 (2013) 1847–1851. Phys. 93 (2003) 5576–5582.
[25] S.J. Lee, K.Y. Kang, S.K. Han, Low-frequency dielectric relaxation of BaTiO3 thin- [56] A.K. Dubey, P. Singh, S. Singh, D. Kumar, O. Parkash, Charge compensation,
film capacitors, Appl. Phys. Lett. 75 (1999) 1784–1786. electrical and dielectric behavior of lanthanum doped CaCu3Ti4O12, J. Alloy.
[26] S.H. Cha, Y.H. Han, Effects of Mn doping on dielectric properties of Mg-doped Compd. 509 (2011) 3899–3906.
BaTiO3, J. Appl. Phys. 100 (2006) 104102. [57] F.D. Morrison, D.C. Sinclair, A.R. West, An alternative explanation for the origin of
[27] C. Ang, Z. Yu, High capacitance-temperature sensitivity and “giant” dielectric the resistivity anomaly in La-doped BaTiO3, J. Am. Ceram. Soc. 84 (2001) 474–476.
constant in SrTiO3, Appl. Phys. Lett. 90 (2007) 202903. [58] D.K. Mahato, T.P. Sinha, Electrical conductivity and dielectric relaxation in
[28] C.E. Ciomaga, M.T. Buscaglia, V. Buscaglia, L. Mitoseriu, Oxygen deficiency and Pr2CoZrO6 double perovskite, J. Alloy. Compd. 634 (2015) 246–252.
grain boundary-related giant relaxation in Ba(Zr,Ti)O3 ceramics, J. Appl. Phys. 110 [59] C.-C. Wang, M.-N. Zhang, K.-B. Xu, G.-J. Wang, Origin of high-temperature relaxor-
(2011) 114110. like behavior in CaCu3Ti4O12, J. Appl. Phys. 112 (2012) 034109.
[29] M. Fukunaga, Y. Uesu, G. Li, CaCu3Ti4O12-type colossal dielectric relaxation in [60] K.J. Andrew, Dielectric relaxation in solids, J. Phys. D: Appl. Phys. 32 (1999) R57.
complex perovskite (Ba1-xLax)(Ti1-xCrx)O3, Ferroelectrics 354 (2007) 106–114. [61] R. Merkle, J. Maier, Defect association in acceptor-doped SrTiO3: case study for
[30] F.D. Morrison, D.C. Sinclair, J.M.S. Skakle, A.R. West, Novel doping mechanism for Fe'TiVO and Mn''TiVO, Phys. Chem. Chem. Phys. 5 (2003) 2297–2303.
very-high-permittivity barium titanate ceramics, J. Am. Ceram. Soc. 81 (1998) [62] G. Er, S. Ishida, N. Takeuchi, Investigations of the electrical property, diffuse re-
1957–1960. flectance and ESR spectra of the La-(Fe,Mn)-codoped PTCR BaTiO3 annealed in
[31] F.D. Morrison, A.M. Coats, D.C. Sinclair, A.R. West, Charge compensation me- reducing atmosphere, J. Mater. Sci. 34 (1999) 4265–4270.
chanisms in La-doped BaTiO3, J. Electroceram. 6 (2001) 219–232. [63] M.D. Glinchuk, I.P. Bykov, S.M. Kornienko, V.V. Laguta, A.M. Slipenyuk,
[32] F.D. Morrison, D.C. Sinclair, A.R. West, Characterization of lanthanum-doped A.G. Bilous, O.I. V'Yunov, O.Z. Yanchevskii, Influence of impurities on the prop-
barium titanate ceramics using impedance spectroscopy, J. Am. Ceram. Soc. 84 erties of rare-earth-doped barium-titanate ceramics, J. Mater. Chem. 10 (2000)
(2001) 531–538. 941–947.
[33] A.R. West, T.B. Adams, F.D. Morrison, D.C. Sinclair, Novel high capacitance ma- [64] H. Kishi, N. Kohzu, Y. Iguchi, J. Sugino, M. Kato, H. Ohsato, T. Okuda, Occupational
terials: – BaTiO3:La and CaCu3Ti4O12, J. Eur. Ceram. Soc. 24 (2004) 1439–1448. sites and dielectric properties of rare-earth and Mn substituted BaTiO3, J. Eur.
[34] A. Ianculescu, Z.V. Mocanu, L.P. Curecheriu, L. Mitoseriu, L. Padurariu, R. Truşcă, Ceram. Soc. 21 (2001) 1643–1647.
Dielectric and tunability properties of La-doped BaTiO3 ceramics, J. Alloy. Compd.

7258

You might also like