Вы находитесь на странице: 1из 27

J. Am. Ceram. Soc.

, 95 [1] 1–26 (2012)


DOI: 10.1111/j.1551-2916.2011.04952.x
© 2011 The American Ceramic Society

Journal
Lead-Free Relaxor Ferroelectrics
Vladimir V. Shvartsman† and Doru C. Lupascu*
Institute for Materials Science, University of Duisburg-Essen, 45141, Essen, Germany

Feature size is a natural determinant of material properties. Its effects, namely the competition between different ordered
design offers the technological perspectives for material states or phases and related changes on the nano- and meso-
improvement. Grain size, crystallite size, domain width, and scale domain structures.5 Control over formation of complex
structural defects of different nature constitute the classical nanoscale structures opens the way to exploring the existing
design elements. Common ferroelectric ceramics contain micro- and even achieving new functionalities. Of particular interest
meter grain sizes and submicrometer domain widths. Domain are functionalities arising from the coupling between different
wall mobility is a major contribution to their macroscopic mate- degrees of freedom, for example, magnetoresistance, magne-
rial properties providing approximately half of the macroscopic toelectric coupling in multiferroics, and electromechanical
output in optimized materials. The extension into the dynamic coupling in relaxors.
nanoworld is provided by relaxor ferroelectrics. Ionic and nano- Since their discovery by Smolensky and co-workers in
scale field disorders form the base to a state with natural nano- 1954,6,7 relaxors have attracted continued interest due to
meter-size polar structures even in bulk materials. These polar their unusual properties. A generic feature of relaxors is a
structures are highly mobile and can dynamically change over broad maximum in temperature dependence of the dielectric
several orders of magnitude in time and space being extremely permittivity, whose position, Tm, is shifted to lower tempera-
sensitive to external stimuli. This yields among the largest tures as the frequency of the probing field decreases
dielectric and piezoelectric constants known. In this feature (Fig. 1).7 In contrast to conventional ferroelectrics, this maxi-
article, we want to outline how lead-free relaxors will offer a mum does not correspond to a phase transition from a para-
route to an environmentally safer option in this outstanding electric to a long-range-ordered ferroelectric state with
material class. Properties of uniaxial, planar, and volumetric polarization being homogeneous inside macroscopic domains
relaxor compositions will be discussed. They provide a broader [Fig. 2(a)].8,9 Instead, in relaxors, the polarization is corre-
and more interesting scope of physical properties and features lated on a local scale resulting in the appearance of polar
than the classical lead-containing relaxor compositions. nanometer-size regions [PNRs, Figs. 2(b)–(d)]. Large dielec-
tric permittivity, e ~ 104–105 [Ref. 10] observed in a broad
temperature range and high piezoelectric coefficients d33 of
I. Introduction
2500 pC/N, accompanied by an almost hysteresis-free actua-
tion strain of more than 0.6%11 have made relaxors a mate-
P ERFORMANCE of functional materials in some cases relies
on symmetry and purity of the material like in the fore-
front silicon semiconductor world. In other cases, it may
rial of choice for high-end industrial applications converting
mechanical into electrical forms of energy and vice versa.12
mostly depend upon imperfections in order or structure. Several reviews have been devoted to relaxors, most of
Colossal effects in complex oxides as an example are criti- them to the lead-containing systems.8,9,13,14 Existing surveys
cally related to inherent heterogeneity. Among well-known on lead-free relaxors15,16 focused on particular systems and
effects are colossal magnetoresistance1 and magnetocapaci- are ~10 years old. In the present feature article, we intent to
tance2 in doped manganites, high-temperature superconduc- overview the state of the art for the most interesting lead-free
tivity in doped cuprates,3 and giant dielectric permittivity relaxor systems, stressing features which are particular to
and electromechanical coupling in relaxor ferroelectrics these materials as compared to the “canonical” relaxors, like
which we will call “relaxors” in the remainder of the docu- PbMg1/3Nb2/3O3 (PMN).
ment.4 In all these quite unlike materials, giant responses to In what follows, we will briefly characterize relaxors in
relatively weak external stimuli are caused by proximity general and then delineate the properties, similarities, and
differences for lead-free relaxors with respect to the more
commonly discussed lead-containing systems. Volumetric,
D. J. Green—contributing editor
planar, and uniaxial relaxor systems will be discussed.

II. Characteristics of Relaxors


Manuscript No. 30125. Received August 05, 2011; approved October 10, 2011. (1) The Polar Nanoregions and their Dynamics
*Member, The American Ceramic Society. The properties of relaxors are closely related to their unique

Author to whom correspondence should be addressed. e-mail: vladimir.shvartsman@
uni-due.de polar structure, namely to the existence of polar nanometer-size

Feature
2 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

regions (PNRs) and their response to external stimuli. These


regions of locally correlated polarizations appear at the
“Burns” temperature, TD, which is typically far above the
maximum point of permittivity, Tm.17 While the spatial sym-
metry is broken inside the PNRs, relaxors retain the higher
symmetry of the parent paraelectric phase macroscopically.
In the absence of an external electrical field, the dipole
moments of the PNRs, Pl, are randomly distributed. There-
fore,
Pno2 measurable remanent polarization exists. However,
as Pl 6¼ 0, the existence of PNRs becomes apparent in
properties depending on P2, for example, in the deviation of
temperature dependences of the refractive index17,18 or the
thermal expansion coefficient13 from the behavior predicted
for the paraelectric state (Fig. 3). For normal ferroelectrics,
the refractive index n decreases linearly with decreasing T,
starting from the high-temperature paraelectric phase down
to the Curie temperature, below which only n(T) deviates
from linearity.19 This deviation is proportional to the square
of polarization. Burns and Dacol found that in the case of
relaxors, like PMN, this deviation already starts at
TD = 620 K, which is about 350 K above Tm (at
f =10 kHz).17 Highly dynamic polar nanoregions have been
made responsible for this effect. The existence of PNRs was
later confirmed by elastic diffuse neutron scattering experi-
ments20,21. In PMN single crystals, a significant diffuse scat-
tering appears below TD with increasing intensity for
Fig. 1. The temperature dependence of the real part of the decreasing temperature.21 This diffuse intensity has been
dielectric permittivity measured at different frequencies for a single associated with the PNRs. Various neutron and X-ray scat-
crystal of the canonical relaxor compound Pb(Mg1/3Nb2/3)O3 (from
Ref. 9).
tering experiments have been carried out to investigate how
PNRs are formed, and what sizes and polarization values

(a) (b)

(c) (d)

Fig. 2. (a) 180- and 90-degree domains in a ferroelectric PbTiO3 single crystal at RT (the sample has been provided by Prof. J. Dec, University
of Silesia, Katowice, Poland); (b) polar structures in a relaxor Pb0.9125La0.0975(Zr0.65Ti0.35)0.976O3 ceramic below the transition temperature [Ref.
286]; (c) polar nanoregions in a relaxor Sr0.61Ba0.49Nb2O6 single crystal above the transition temperature [Ref. 240]; (d) polar nanoregions in a
BaTi0.85Sn0.15O3 ceramic above the transition temperature [Ref. 83]. Polar structures were observed by piezoresponse force microscopy. Blue and
red colors correspond to polarization oriented up and down, respectively. Yellow contrast marks regions with a negligible response.
January 2012 Lead-Free Relaxors 3

are conjugated to the order parameter (the electric polariza-


tion in case of ferroelectrics and relaxors) will destroy the
transition to a long-range-ordered state.30 In the case of re-
laxors with the perovskite structure (“cubic” relaxors), like
PMN, the local polarization has eight allowed directions tak-
ing into account the local rhombohedral symmetry, and
hence, can be considered quasi-continuous. An equilibrium
phase transition into a long-range-ordered ferroelectric state
is therefore excluded. Indeed, the low-temperature state in
PMN is a short-range-ordered glassy-like state.
At the same time, at temperatures below TD, the RFs pro-
mote nucleation of PNRs with polarity controlled by the fluc-
tuations of the RFs.28 In the vicinity of the Burns
temperature, the PNRs are considered to be dynamical entities
with their dipole moments thermally fluctuating between
equivalent polarization directions.31 The dipole moments of
the PNRs are weakly correlated and are free to reorient. After
an external excitation, the system returns to the state with
lowest free energy. It is always the same state regardless of the
initial conditions. To distinguish this state from the paraelec-
tric PNR-state, it has been termed the ergodic relaxor (ER)
Fig. 3. Temperature dependences of the linear thermal expansion,
Dl/l, (a) and the refractive index, n, (b) for Pb(Mg1/3Nb2/3)O3 phase.9 The thermally activated reorientation of dipole
showing deviations from linear behavior starting at a temperature moments of the PNRs in the ER phase yields the major con-
(TD) much higher than the peak (Tm) of the dielectric permittivity. tribution to the dielectric permittivity of relaxors.31 On cool-
After G. A. Samara,13 reprinted with permission by Institute of ing, the interaction between the PNRs grows stronger
Physics Publishing Ltd. resulting in a slowing down of their dynamics. A broad distri-
bution of PNR sizes and randomness of interactions between
develop at different temperatures.14,22 The correlation length them yield a broad distribution of relaxation times giving rise
of the atomic displacements contributing to diffuse scattering, to a broad peak of dielectric permittivity versus temperature.
which is a direct measure of the PNRs’ size, is inversely pro- Finally, in systems like PMN, the divergence of the longest
portional to the width of the diffuse peak. According to relaxation time at a finite temperature results in the freezing
high-resolution neutron diffuse scattering performed on of PNR dynamics and in the transition to a glassy-like
PMN single crystals, the PNRs grow from about 1.5 nm at state.32,33
TD to ~10 nm at 10 K upon cooling. The most significant The freezing temperature, Tf, may be derived from fitting
change of their size was observed around the freezing tem- the frequency dependence of the temperature of the dielectric
perature Tf ~ 210 K. This temperature coincides with those permittivity maximum, Tm, to the phenomenological Vogel–
estimated from the analysis of the dynamic dielectric Fulcher law:
response of relaxors and corresponds to the slowing-down of
 
the relaxation related to the reorientation of PNRs.9 The Ea
number of PNRs estimated from the integrated intensity of f ¼ f0 exp ; (1)
kðTm  Tf Þ
scattering shows a monotonic increase upon cooling, starting
from TD, and then, an abrupt decrease at about Tf due to
the merging of smaller PNRs into larger ones.23 Qualita- where Ea is an activation energy, f0 is an attempt frequency,
tively, similar results have been obtained for other relaxor and k is the Boltzmann constant.31 In this state, the polar
compounds.14,24,25 order remains short-range. Polarization is correlated on the
The microscopic origins of PNRs are closely related to the nanometer scale inside the PNRs [Figs. 2(b)–(d)], but the sys-
inherent structural and charge inhomogeneities typical of tem remains nonpolar on a macroscopic scale. As relaxors
relaxors. In PMN, these inhomogeneities are due to the coex- display various characteristics of nonergodic behavior below
istence of two B-site cations (Mg2+ and Nb5+) with different Tf, viz. an anomalously wide relaxation time spectrum,34
charges and chemical bonding characteristics in the ABO3 aging,35 and dependence of the state on the thermal and field
perovskite structure. High-resolution electron microscopy has history of the sample,18,33,36 this state is often called the
revealed an inhomogeneous distribution of B-site cations: nonergodic relaxor (NR) state.9
nanosized chemically ordered regions (CORs, ~2–5 nm) exist A long-range-ordered ferroelectric-like state may be
in a disordered surrounding.26,27 Based on X-ray dispersive induced in relaxor compounds when poling by an electric
spectroscopy measurements, Davies and Akbas suggested that field larger than a critical strength,37 by applying a mechani-
one of the B-sublattices inside the CORs is occupied exclu- cal strain,38 or even by thermal annealing stimulating cation
sively by B5+ ions, while the other one contains a random ordering.39 A crossover from the relaxor to the ferroelectric
distribution of B2+ and B5+ ions in a 2:1 ratio.27 Both, ran- behavior may also be provided by chemical substitution,
domly distributed B-site cations and CORs are believed to be which is, for example, the case in (1x)PbMg1/3Nb2/3O3–x
the sources of quenched random electric fields (RFs).18,28,29 PbTiO3 (PMN-PT) with large Ti4+ content (x > 0.35)40 or
This means that, additional to the external electric field and SrxBa1xNb2O6 (SBN) with large Ba2+ content (x  0.5).41
the fields locally generated by polarization mismatch at grain Compounds from an intermediate concentration range com-
boundaries arising in all polycrystalline ferroelectrics to some bine both relaxor and ferroelectric properties. On the one
extent, relaxors contain sources of local electric fields within hand, they show strong frequency dispersion of the dielectric
the individual grains. These are randomly ordered according permittivity around Tm and evidence of the existence of
to the disorder induced by the differently hetero-charged sub- PNRs far above Tm.42 On the other hand, their low-tempera-
stitutional ions in the lattice. Different from pure acceptor- or ture state exhibits typical ferroelectric properties, such as fer-
donor-type defects, these are overall compensated within the roelectric hysteresis, pyro-, and piezoelectricity, etc.9
structure and only exhibit local charge mismatch, yielding the Nevertheless, there are some differences between this state
local random electric fields. and a “normal” ferroelectric state. The PNRs appear to per-
Imry and Ma have argued that in systems with continuous sist even in the low-temperature ferroelectric state.43,44 Some
symmetry of the order parameter, the quenched RFs which compounds exhibit specific macroscopic phase inhomogene-
4 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

ity. Namely, depth-resolved X-ray diffraction experiments


performed on single crystal (1x)PbZn1/3Nb2/3O3–xPbTiO3
(PZN-PT)45 and PMN-PT46 revealed that the structural
phase transition to the rhombohedral ferroelectric phase only
occurs in an outer layer (thickness several tens of micro-
meters), while the interior part remains cubic, similar to the
NR state of PMN.

(2) Lead-Free and Lead-Containing Systems


While the relaxor behavior was initially discovered in Ba-
TiO3–BaSnO3 solid solutions,6 most of the subsequent
research activity has focused on lead-containing systems,
such as PMN-, PZN-based solid solutions, (Pb,La)(Zr,Ti)O3
ceramics, etc., which is due to their larger material output
coefficients. The formation as well as the evolution of the
relaxor state have been thoroughly investigated for these
materials using a broad spectrum of experimental methods.
Several theoretical models have been developed.9,28,47,48
During the last few years, interest in the lead-free relaxor
compounds has grown considerably. According to the ISI Fig. 4. The crystallographic structure of BaTiO3.
Web of Knowledge, the portion of articles devoted to lead-
free relaxor systems among all articles devoted to relaxors
has continuously increased from ~25% in 2000 to around When solid solutions with other perovskite type com-
45% in 2010. The growth is stimulated by the increasingly pounds are formed, BaTiO3 yields relaxor compositions
severe restrictions on the use of lead in industry as a hazard- with a number of interesting features, some occurring in
ous substance. According to the directive for the Restriction certain compositions, others in several, and again some con-
of the use of certain Hazardous Substances in electrical and currently:
electronic equipment (RoHS)49 adopted by the European
Parliament in the year 2006, the maximum allowed concen- 1. Relaxor behavior in compositions without nominal
tration of lead is established to be 0.1 wt% in homogeneous charge disorder.
materials for electrical and electronic equipment used in 2. Diffuse phase transitions as an intermediate state
households as well as industry. Some specific applications are between relaxor and ferroelectric states.
exempt from the directive, if their elimination is technically 3. Existence of ferroelectric domains above the transition
or scientifically impracticable, lead-containing ferroelectrics temperature.
being one of them. They will be prohibited once a practicable 4. Coexistence of static and dynamic PNRs in a broad
substitution is available. Similar regulations have been estab- temperature range.
lished in a number of other countries. 5. Re-entrant behavior.
Lead-free relaxors also exhibit a broader variety of phe- In the following, we will distinguish between these differ-
nomena in comparison with the “canonical” lead-containing ent properties and outline which compositions offer which
compositions. Some of them transform to relaxor behavior peculiarities.
by isovalent substitution, i.e. without a nominal charge disor- A crossover to the relaxor state was observed in BaTiO3
der, some show re-entrant relaxor behavior, and others show by both heterovalent and isovalent ionic substitutions. In the
antiferroelectric (AFE) relaxor states, just to mention a few. first case, the relaxor behavior is induced by substitutions
Properties like large dielectric tunability,50 high dielectric per- either on A- or B- or on both A- and B-sites of the perov-
mittivity at elevated temperatures,51,52 large electrically skite lattice ABO3, for example, in Ba1x(Sm0.5Na0.5)xTiO3,57
induced mechanical strain,53 a strong photorefractive effect,54 Ba1x/2 x/2(Ti1xNbx)O3 ( denotes a vacant site),58 and
and other properties render the lead-free relaxor systems BaTiO3–BiScO359–61 or BaTiO3–La(Mg0.5Ti0.5)O3,62,63 respec-
attractive for a number of applications. tively. In the second case, the relaxor behavior appears in
compounds with substituted B-site cations, such as BaTiO3–
III. Lead-Free Relaxor Systems BaSnO36 or BaTiO3–BaZrO3.64
In most cases, the dielectric properties of BT transform
(1) Lead-Free Relaxors with Perovskite Structure according to the following scenario:
(A) BaTiO3-Based Relaxors: The BaTiO3–BaSnO3 First, the temperatures of the transition between different
was the first system in which relaxor properties were discov- ferroelectric phases increase and come close to each other.
ered.6 A number of relaxor compounds based on BaTiO3 Simultaneously, TC goes down [Fig. 6(a)].
(BT) have been found since. The BT is among the most stud- Second, at a certain concentration of substituting ele-
ied and most widely used ferroelectrics. The discovery of fer- ments, all these transitions merge, a so-called pinched transi-
roelectricity in this compound in the 1940s55,56 has triggered tion occurs.64 The system then directly transforms from the
a new era in the investigation of ferroelectrics. The BT crys- cubic paraelectric to the rhombohedral ferroelectric state.
tallizes in the cubic perovskite structure (point group m3m) The corresponding peak of e′(T) is usually much broader
above its Curie temperature, TC = 393 K (Fig. 4). Below TC, than that in BaTiO3, however, without the frequency shift of
it successively transforms to three ferroelectric phases: first to Tm typical of relaxors. Such behavior is occasionally denoted
the tetragonal (4mm), then to orthorhombic (mm) at about as a diffuse phase transition (DPT) to distinguish it from the
278 K, and finally, to a rhombohedral (3m) phase at 183 K. relaxor one. Another hallmark of the DPT is a deviation of
The spontaneous polarization in the three ferroelectric phases the dependence of the real part of the dielectric permittivity e
lies along [001], [011], and [111] direction of the parent cubic ′(T) from the Curie–Weiss law at T > Tm [Fig. 6(b)]. This
structure, respectively [Fig. 5(a)]. Above TC, the dielectric deviation usually starts at temperatures significantly larger
permittivity follows a Curie–Weiss law e = C/(Th) with the (up to hundreds of degrees) than Tm.
Curie–Weiss temperature, h < TC, indicating a first-order Several phenomenological approaches have been devel-
phase transition. oped to describe the DPT behavior. One is based on the
January 2012 Lead-Free Relaxors 5

Fig. 5. (a) Phase transitions in BaTiO3 according to the displacive scenario with Ti ions shifted from the center. (b) Phase transitions in BaTiO3
according to the order–disorder scenario with Ti ions exhibiting jumps between eight off-center positions. Shades of gray from bright to dark
illustrate occupancy probabilities from most to less probable (Refs. 117,119).

where e′m is the maximum of the dielectric permittivity at a


given frequency, Tm is the corresponding temperature, and
the parameter d determines the degree of diffuseness of the
transition. However, this quadratic law barely fits experimen-
tal e(T) dependences of many compounds showing DPT
behavior. Another empirical description was later proposed
by Santos and Eiras65

e0 m
e0 ¼ TT c ; (3)
1þ d
m

Here, Tm and d have similar meanings as in Eq. (2). The


parameter c gives information on the character of the phase
transition: c = 1 corresponds to a “normal” ferroelectric
transition, c = 2 indicates so-called complete diffuse phase
transition, and c 2 (1, 2) represents an intermediate state, a
so-called incomplete DPT with a certain degree of diffuseness
of its phase transition. Obviously, the Curie–Weiss law and
Eq. (2) are limiting cases of Eq. (3). This equation is more
general and better describes a variety of intermediate states
in between purely ferroelectric and relaxor behavior. In
particular, the frequency dispersion region around Tm can be
fitted maintaining the parameters d and c frequency-indepen-
dent.65 Instead of Eq. (3), another function, 1=e0  1=e0m ¼
ðT  Tm Þc =C, proposed by Uchino and Nomura,66 is often
Fig. 6. (a) Temperature dependences of the real part of the used to describe the DPT. While it is very similar to the Eq.
dielectric permittivity measured at 102–106 Hz for heterovalent- (3) and yields the same values of the parameter c, the dimen-
substituted (1x)BaTiO3–xLa(Mg1/2Ti1/2)O3 compounds. (b) sionality of the parameter C depends on the value of c,
Temperature dependences of the reciprocal permittivity for same which makes this notation physically difficult to interpret.
compositions fitted by the Curie–Weiss law. Reprinted with At higher substitution, Tm starts to show frequency disper-
permission from Ref. 63. Copyright Taylor & Francis Group, http:// sion matching the Vogel–Fulcher law [Eq. (1)]. This is often
www.informaworld.com.
used as a criterion of a completed crossover to relaxor
model proposed by Kirillov and Isupov.47 In this model, the behavior. Sometimes, especially when data of the frequency
DPT is considered to be due to local composition fluctua- dispersion of e(T) are missing, the relaxor state is associated
tions, resulting in microregions with different local Curie with the condition c = 2 [Eq. (3)]. For example, for the
points distributed in a Gaussian fashion around a mean (1x)BaTiO3–xLa(Mg0.5Ti0.5)O3 system, the parameter c
Curie temperature. The following equation was obtained to changes from 1.1 to 1.9 when x increases from 0 to 0.075,
describe the temperature dependence of the dielectric permit- respectively (Table I).63 Correspondingly, compounds with
tivity in the vicinity of Tm: x  0.075 show a distinct Tm(f) dispersion [Fig. 6(a)].
The degree of the crossover from the ferroelectric to the
e0 m relaxor state depends on the substitution type. Ravez and
e0 ¼ e01 þ 2 ; (2)
1 þ ðTT mÞ Simon systematized the effects of various substitutions on the
2d2 properties of BaTiO3.15,67 Most experimental data agree well
6 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

Table I. Summary of the Dielectric Characteristics for (1x)BaTiO3–xLa(Mg0.5Ti0.5)O3 Ceramics


x 5
em (10 Hz) Tm, K (105 Hz) DTm, K (102–106 Hz) c Tf , K Ea, eV f0, Hz

0.025 5300 331 0.5 1.34


0.05 5500 257 1.5 1.72
0.075 6400 181 3.5 1.91
0.10 4800 101 9 1.91 87 0.02 2.3∙1010
Here, em is the peak value of the dielectric permittivity at 100 kHz, DTm is the difference between positions of the maximum in the e(T) dependences, Tm, mea-
sured at 106 and 102 Hz, the parameter c reflects a degree of diffuseness of the e(T) peaks according to Eq. (3). Tf, Ea, and f0 are the Vogel–Fulcher fitting parame-
ters, according Eq. (1) [Refs. 62,63].

with their classification. In particular, the relaxor state is formations of parent BaTiO3 (pinched phase transition)
induced by a relatively low doping level in the case of hetero- occurs at x ~ 0.1079 and x ~ 0.0880 for (1x)BaTiO3–xBaZrO3
valent substitutions on both A- and B- sites of the perovskite (BTZr) and (1x)BaTiO3–xBaSnO3 (BTSn), respectively.
structure [e. g., x = 0.055 for (1x)BaTiO3–xLa(Mg0.5Ti0.5) Compositions in the concentration ranges
O3,63 x = 0.075 for (1x)BaTiO3–xNaNbO368, x = 0.10 for 0.10  x  0.18 (BTSn) and 0.15  x  0. 20 (BTZr)
(1x)BaTiO3–xBiScO359] or on both B- and oxygen-sites show DPT behavior (Fig. 7). The nature of this phase transi-
[x = 0.04 for Ba(Ti1xLix)O33xF3x69] (Table II). Heterova- tion is still a matter of discussion. Some researchers attribute
lent B-site substitution is exemplified by Nb-doped BaTiO3, it to the relaxor state with the distinction between a fre-
where the relaxor behavior has been reported for Nb concen- quency-independent DPT and a frequency-dependent relax-
trations higher than 7.5 at.%.58 However, in this case, A-site or-like behavior due to different sizes of the PNRs yielding
vacancies are required for charge compensation and the exact different relaxation times.81 When these relaxation times are
stoichiometry becomes Ba1x/2x/2(Ti1xNbx)O3.58 Thus, close to the experimental observation rates, relaxor behavior
these compounds can be also considered as both A- and B- is observed; when they differ much, only DPT behavior is
site hetero-substituted. For heterovalent substitution on the expected. However, there is an evidence that the DPT corre-
A-site relaxor behavior is induced at moderate doping levels sponds to the ferroelectric phase transition. Ferroelectric
(e.g., x = 0.10 for Ba1xBi2x/3x/3TiO370 or x = 0.20 for domains have been observed in compounds with DPT behav-
Ba1xKx/2Lax/2TiO371). In solid solutions with isovalent- ior in the vicinity of Tm both in BTSn82,83 and BTZr74
substituted Ti4+, like BaTi1xZrxO372–74 or BaTi1xSnxO36,75–78, ceramics. Broad-band dielectric spectroscopy reveals a contri-
the relaxor state is only observed at a fairly large substitution bution to the dielectric permittivity related to the relaxation
levels x > 0.25–0.30. of pinned ferroelectric domain walls (Fig. 8).74,83 This relaxa-
The latter systems are of particular interest. First, the
more gradual crossover from the ferroelectric to the relaxor
behavior allows investigating this process in more detail than
for compounds with heterovalent substitutions. Second,
appearance of the relaxor behavior in systems without nomi-
nal charge disorder is unique and has not been well under-
stood so far.
(B) Crossover to the Relaxor State in BaTiO3–BaMO3
(M = Zr, Sn) Solid Solutions: The isovalently substituted
BaTiO3–BaMO3 (M = Zr, Sn) solid solutions show a more
gradual crossover to the relaxor state than heterovalent-
substituted compositions. The merger of all polymorph trans-

Table II. Critical Concentrations Corresponding to the


Ferroelectric-Relaxor Crossover for Some BaTiO3-Derived
Systems
Type of
substitution Solid solution x0 Ref. Fig. 7. Temperature dependences of the real part of the dielectric
permittivity measured at different frequencies for isovalent-
Isovalent: BaTi1xZrxO3 0.30 74 substituted (1x)BaTiO3–xBaSnO3 compounds [Ref. 76].
B-site BaTi1xSnxO3 0.25 76
BaTi1xCexO3 0.20 279
Heterovalent: Ba1xBi2x/3x/3TiO3 0.10 70
A-site Ba1xKx/2Lax/2TiO3 0.25 71
Ba1x(Sm0.5Na0.5)xTiO3 * 0.20 57
Heterovalent: Ba1x/2x/2(Ti1-xNbx)O3 0.075 58
A- and Ba1xLax[Ti1x(Mg0.5Ti0.5)x]O3 0.055 63
B-site Ba1xNax[Ti1xNbx]O3 0.075 68
Ba1xBix[Ti1xScx]O3 0.10 59
Ba1xBix[Ti1xAlx]O3 0.10 280
Ba1xLax[Ti1xFex]O3 0.10 281
Heterovalent: Ba[Ti1xLix]O33xF3x 0.04 69
B- and
anion-site
 denotes vacant site.
*
While BaTiO3 doped by Nb5+ nominally corresponds to B-site substitu-
tion, A-site vacancies are needed for charge compensation in this case. Thus, Fig. 8. Frequency spectra of the imaginary part of the dielectric
one has actually the situation corresponding to both A- and B-site heterova- permittivity for (1x)BaTiO3-xBaSnO3 ceramics with DPT (15% of
lent substitution. Sn) and relaxor (30% of Sn) behavior (Ref. 76).
January 2012 Lead-Free Relaxors 7

Table III. Summary of the Dielectric and Relaxor Characteristics for (1x)BaTiO3–xBaZrO3 Ceramics.
Parameters are the Same as for Table I [Ref. 74]
x em (105 Hz) Tm, K (105 Hz) DTm, K (102–106 Hz) c Tf , K Ea, eV f0, Hz

0.15 31900 344.5 0.5 1.29


0.20 30200 317 0.6 1.46
0.25 13500 279.5 3 1.93 278 5 9 104 8.9 9 107
0.30 11900 233 6 1.86 228 0.001 1.2 9 107
0.35 9960 198 25 1.96 165 0.02 108
0.40 4600 161 42 1.9 70 0.13 1.5 9 1012

tion already appears above Tm, which correlates with the increases beyond 25%, a critical size and distribution density
observation of ferroelectric domains in the same temperature of PNRs was reached where the interaction between them
range for BTSn ceramics with x = 0.10, 0.15.83 The existence becomes essential. Correspondingly, BTZr compounds start
of ferroelectric domains in some temperature interval above to show relaxor behavior with finite freezing temperature
Tm has also been reported for the heterovalent-substituted gradually increasing from 4 to 122 K when x decreases from
solid solution (BaTiO3)0.975–(LaMg1/2Ti1/2O3)0.025 also show- 0.75 to 0.35.
ing DPT behavior.84 Figure 9 shows a phase diagram for (1x)BaTiO3–xBaZrO3
In compositions with larger content of Sn (or Zr), a shift solid solutions based on experimental data by several
of Tm with frequency following the Vogel–Fulcher law is authors. After a pinched transition at x ~ 0.10, the low-tem-
observed, which is typical of the relaxor behavior perature state below Tm is ferroelectric (with rhombohedral
(Table III).74,83 The dielectric spectra change qualitatively as symmetry) in compositions with 0.10 < x  0.20. For
compared to compositions with a diffuse phase transition: 0.25  x  0.35, a percolation transition (one assumes
the domain wall-related relaxation disappears, while another ferroelectric order inside the percolating clusters) occurs at
contribution in the high-frequency range becomes dominant Tf, but regions with dynamic PNRs simultaneously exist
(Fig. 8 shows the dielectric spectra for the BTSn ceramics below the transition temperature.74 In the range
with x = 0.30). This new component corresponds to the low- 0.35 < x < 0.80, canonical behavior is found transiting into
f wing of the Debye-like relaxation process related to the the NR state with macroscopic cubic symmetry below Tf.
reorientation of PNRs with characteristic frequencies Finally, compounds with 0.80  x  0.95 remain in the
between 106 and 109 Hz.85 A slowing down of the PNR ER state (or in a superparaelectric (SPE) state) down to the
dynamics upon cooling is reflected in a broadening of the lowest temperatures investigated (near 0 K). For some BTZr
relaxation time distribution and in its shift toward longer compositions, the Burns temperature was estimated from
relaxation times. The shape of e′′ (f) changes accordingly. In thermal expansion measurements.64 At high temperatures,
both BTSn (x  0.20) and BTZr (x  0.25) ceramics, a the strain follows a linear temperature dependence normal
shift of the PNR-related relaxation down to lower frequen- for the paraelectric state. Below TD ~ 440 K, a deviation
cies was observed for decreasing temperature.74,76,77 The from linear behavior was observed indicating an electrostric-
spectra broaden strongly and become almost flat at tempera- tive strain due to formation of the local spontaneous polari-
tures below 110–130 K. This is attributed to a transition to a zation. Thus, TD corresponds to the appearance of PNRs, i.
NR state accompanied by a freezing of the relaxation e. to a transition from the paraelectric state to the ER state.9
dynamics. The transition takes place approximately at the Approximately, the same temperature for the paraelectric–
same temperature for BTSn 0.20  x  0.30 and BTZr ergodic relaxor transition has been reported for Raman data
0.25  x  0.35 ceramics irrespective of composition, while BTSn with x = 0.30.88
the freezing temperatures estimated from the Vogel–Fulcher There is still some controversy regarding the nature of the
fits of Tm(f) differ for these compositions.74,76 For BTSn with low-temperature relaxor state in the isovalent-substituted
20% and 25% of Sn4+, for example, Tf is substantially BaTiO3 compounds. Some authors argue that it is qualita-
higher than the temperature corresponding to complete freez- tively different from canonical PMN-type relaxors64,89 based,
ing of PNRs relaxation. An explanation is that Tf estimated for example, on results of heat capacity measurements,89
from the Vogel–Fulcher fit corresponds to the appearance of
an infinite percolated cluster with frozen polarization, while
some re-orientable PNRs continue to exist below Tf. Bokov
et al. similarly suggested that in BTZr (x = 0.35), two inter-
penetrating sets of PNRs exist below Tf: static (large) PNRs
and dynamic (small) ones.73 They point out that in some
cases, the fulfillment of the VF law does not imply the freez-
ing of the relaxation.86 A maximum in the temperature
dependence of static permittivity and a gradual broadening
of the dielectric spectrum upon cooling necessary for such a
kind of behavior are satisfied for BTZr (x = 0.35) in the tem-
perature range around Tm.73 In compositions with relatively
high substitution rate, x  0.30 (BTSn), x  0.40 (BTZr)
Tf approximately matches complete freezing of the PNR
dynamics found from temperature-dependent dielectric per-
mittivity.
Maiti et al. have investigated the BTZr solid solutions on Fig. 9. Phase diagram of the (1x)BaTiO3–xBaZrO3 system.
the Zr-rich side.64,87 In compounds with Zr content Transition temperatures were determined based on the dielectric
permittivity measurements: 0.00  x < 0.10 (Ref. 79), 0.10  x 
0.80  x  0.95, the relaxation peak in the dielectric per- 0.40 (Ref. 74), 0.40  x  0.95 (Refs. 64,87): solid and open
mittivity, Tm, follows the Arrhenius law, f = f0exp(Ea/kTm). squares correspond to Tm (at 100 kHz) and Tf, respectively. The
The authors attributed this relaxation to PNRs appearing Burns temperature (stars) was estimated based on data of the
around the ferroelectrically active Ti4+ cations introduced acoustic emission for x = 0.00 (Ref. 125) and thermal expansion
into the nonpolar matrix of BaZrO3. As the Ti content measurements for 0.40  x  0.70 (Ref. 64).
8 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

Raman spectroscopy,90 or X-ray diffraction.91 However, data


reported by different groups are often contradictory. Naga-
sawa et al. have not observed an anomaly in specific heat
measurements of relaxor BTZr (x = 0.35) ceramics,89 which
would be typical of PMN.92 On the contrary, Gorev et al.93
have reported anomalous specific heat contributions for this
composition in the temperature ranges 250–350 and 150–
200 K, relating them to the appearance of PNRs and their
freezing, correspondingly. In PMN, long-range order can be
induced by field cooling from the paraelectric state.94 Sciau
et al. have not observed any electric field-induced long-range
polar order in BaZr0.35Ti0.65O3 by X-ray diffraction measure-
ments under electric field.91 Bokov et al. have argued that
the presence of nanometer-size nonpolar Zr-rich regions90,95
restricts the growth of PNRs in the BTZr relaxors and pre-
vents their transformation to macroscopic ferroelectric
domains even under a strong enough external electric field.73
Pyroelectric experiments on the other hand reveal that appli-
cation of suitable electric fields can induce a ferroelectric
order below Tf in BTZr relaxor compounds with
0.35  x  0.45.64 The zero-field cooling/field heating
(ZFC/FH) experiments reveal a decay of this electrically
induced state at depolarization temperatures ~81 (x = 0.35),
~46 (x = 0.40), and ~33 K (x = 0.45). These temperatures lie
below the corresponding freezing temperatures estimated
from Tm(f). The authors attributed this depolarization to a
transition from the field-induced ferroelectric state back to a
NR state.
(C) BaTiO3–BiScO3, A Re-Entrant Relaxor: A pecu-
liar behavior has been observed in (1x)BaTiO3–xBiScO3
relaxors. The BiScO3 is a paraelectric with a monoclinic dis-
torted perovskite structure characterized by both oxygen Fig. 10. (a) The peak (Tm) and freezing temperature (Tf) as a
octahedra tilting and antiparallel displacement of B-cations.96 function of BiScO3 concentration for (1x)BaTiO3–xBiScO3
Solid solution (1x)BaTiO3–xBiScO3 shows a fast crossover ceramics (from Ref. 59). (b) Remanent polarization as a function of
from the ferroelectric to the relaxor behavior typical of temperature in (1x)BaTiO3–xBiScO3 ceramics showing re-entrant
heterovalent substitutions.59,61 At x  0.10, the samples relaxor behavior. Reprinted with permission form Ref. 61.
show VF-type behavior with an activation energy Ea ~ 0.1– Copyright 2011, The American Physical Society.
0.2 eV, a value about one order of magnitude larger than in
PMN.59 Such values of Ea imply a weak interaction between
PNRs classifying this system as a “weakly coupled relaxor,” preted as a long-range-ordered state with a remanent polar-
where the PNRs are isolated and frustrated in a matrix that ization developing below Tm between the ER and the NR
provides only weak coupling between neighboring polar states in (1x)BaTiO3–xBiScO3 (0.10  x  0.45). Bharad-
regions.59 waja et al. assumed that re-entrant features in this system are
Contrary to most BT-derived relaxors, an increase of both due to a competitive dipole interaction between local ferro-
Tm and Tf was observed in compounds with higher BiScO3 electric and AFE order and/or local stresses.61 Guo et al.
content starting from x = 0.10 [Fig. 10(a)].59 Such enhance- observed by piezoresponse force microscopy that the disor-
ment of the transition temperature at relatively small amount dered state with presence of PNRs grows within the normally
of the second component is unusual in BT-based solid solu- ordered ferroelectric phase.102 They concluded that polar
tions. It was reported only when Ba was partially substituted clusters develop from a ferroelectric state and re-enter the
either by Pb97 or by Bi98. It was assumed that the Bi–O frozen disordered phase, instead of emerging from a para-
bonding due to its partially covalent nature99 induces a dif- electric state as in most other relaxor ferroelectrics.102
ferent component of polarization in the structure pushing the Similar re-entrant relaxor behavior has also been reported
transition temperature to higher values.98 A similar type of for some other lead-free systems, like 0.99BaTiO3–0.01
Tm(x) dependence with a minimum has been reported for AgNbO3103 and Ba2Nd1xPrxFeNb4O15104 [Fig. 11(b)].
some other systems, for example, (1x)BT–xKNbO3100 and (D) BaTiO3-Based Relaxors: Microscopic Properties: There
(1x)BT–xNaNbO3101 solid solutions. However, in these are many evidences that at the microscopic scale, BaTiO3-
cases, the increase of Tm starts at rather high concentrations based relaxors behave similar to pure barium titanate. The
of the second component, x = 0.7 and 0.5, respectively, Raman spectra of BT-derived relaxors, for example, look
reflecting a dilution of the high TC KNbO3 and NaNbO3 very much like those recorded in the paraelectric phase of
compounds by BaTiO3. BaTiO3 with broad bands centered at ~285 and 520 cm1
Another marked phenomenon reported for the BaTiO3– (Fig. 12).58,88,90,105,106 Even though a perfect cubic symmetry
BiScO3 system is re-entrant relaxor behavior called so analo- forbids any Raman modes, the Raman spectrum in the para-
gously to re-entrant magnetic spin glass systems.61,102 A more electric phase of pure BaTiO3 has been reported at tempera-
ordered state first develops at higher temperatures before tures up to 570 K. It is considered as an indication of a
going back to a more disordered state at low temperature. strong local disorder due to random displacement of Ti4+
Taking the remanent polarization, for example, it first ions at high temperatures.107,108 Presence of these bands in
increases as the temperature is lowered, but drops again on BT-derived relaxors implies that the local structure around
further cooling essentially vanishing at low T [Fig. 10(b)].61 the BO6 octahedra deviates from the perfect cubic symmetry.
Another manifestation of re-entrant behavior is a relaxation It was conjectured that the local structure of BTZr relaxors
peak related to the freezing of PNRs occurring below is rhombohedral.105 A high-frequency mode at ~780 cm1
another frequency-independent transition attributed to the (BTZr), not existing in pure BT is often considered as a sig-
ferroelectric phase transition [Fig. 11(a)].102 This was inter- nature of the relaxor phase.88,105,106 In contrast to ferroelec-
January 2012 Lead-Free Relaxors 9

(for the 520 cm1 line) shows saturation below the freezing
temperature, indicating a saturation of the correlation func-
tion of polarization,58,88,105 whereas for ferroelectric com-
pounds, a continuous increase of the Raman-integrated
intensity of most lines was observed reflecting the onset of
long-range order.58
Investigation of the local structure of BTZr relaxors by
neutron pair distribution function analysis also indicates that
the Ti4+ displacements present in the ferroelectric BaTiO3
persist in BTZr relaxors.110 Moreover, the local dipole
moments associated with the Ti4+ are similar in BTZr and
BT. This similarity shows that the difference in the dielectric
properties of ferroelectric BT and relaxor BTZr only lies in
different correlations between the cation displacements.
Using first principle calculations, the same group showed
that contrary to BT, where the Ti displacements are all the
same in amplitude and direction along a <111> direction; in
BTZr relaxors, other types of displacement also appear:
along <110>, <100> directions, and even nearly canceled.111
The type of Ti polar displacement is entirely determined by
the Ti/Zr distribution in the adjacent unit cells. The underly-
ing mechanism involves local strain effects that ensue from
the large size mismatch between the Ti4+ and Zr4+ cations.
Random strains lead to a reorganization of the covalent
bonds between the Ti atoms and their oxygen neighbors in
BTZr resulting in a reorientation of the Ti shifts.111 A ran-
dom distribution of the Ti displacement symmetries among
the Ti–O6 octahedra impedes a perfect alignment of all the
Ti displacements as it exists in ferroelectric BT at low
temperatures. Nevertheless, the local distribution of the
substituted cations does not completely determine the Ti
off-centering: <111> displacements can be directed in eight
possible directions, four possible directions for <110> dis-
Fig. 11. Temperature dependence of the real part of the dielectric placements, and two possible directions for <100> displace-
permittivity measured between 102 and 106 Hz for (a) (1x)BaTiO3–x ments. It was argued that this degree of freedom allows
BiScO3, x = 0.05 (Reprinted with permission from Ref. 102. Copyright
partial correlations of the polar displacements, and thus the
2008, American Institute of Physics) and (b) Ba2PrxNd1xFeNb4O15,
x = 0.6 ceramics each showing re-entrant relaxor behavior. Reprinted formation of polar nanoregions.111
from Ref. 104 with permission by Institute of Physics Publishing, Ltd. Off-centering of Ti4+ ions in BTZr solid solutions was
confirmed by the analysis of neutron total scattering mea-
surements by a reverse Monte Carlo approach,112 while the
Zr4+ ions remain in centers of the oxygen octahedra. The
trics, BT-derived relaxors exhibit a continuous variation of correlation of the Ti4+ displacements results in polar nanore-
the overall intensity of the Raman spectra with temperature gions, whose polarization direction evolves on a few nanome-
without any significant change in peak positions or band- ter length scale. Comparison between BTZr with different Zr
width.58,105,106,109 That is, from the Raman point of view, content shows that this correlation is stronger and extends
they do not show any structural phase transition. On the further for compositions with ferroelectric behavior
other hand, for relaxor BTZr, the integrated Raman intensity (x = 0.20) in comparison with relaxor BTZr (x = 0.35).112

Fig. 12. Temperature evolution of Raman spectra for (1x)BaTiO3–xBaZrO3 ceramics with x = 0.00 and 0.35. Reprinted with permission from
Ref. 109. Copyright 2009, American Institute of Physics.
10 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

Electron diffraction experiments on both isovalent and such cluster can reach maximal size. In this situation, the sys-
heterovalent-substituted BaTiO3 relaxors reveal diffuse scat- tem maintains ferroelectric behavior. When the concentration
tering, which implies existence of highly correlated aniso- of the impurities approaches and exceeds the threshold, the
tropic chain dipoles along the <001> directions.113,114 These concentration of clusters increases, but they cannot reach
chains were attributed to one-dimensional PNRs. The simi- their maximal size. The relaxation of these small clusters
larly correlated anisotropic <001> chain dipoles are also results in the dispersion of the dielectric permittivity and re-
known to be the characteristic of the paraelectric phase of laxor behavior of the system. The interaction between them
BaTiO3 itself,115,116 being induced by covalent hybridization finally yields blocking of their relaxation at Tf. The model of
between Ti 3d and O 2p orbitals along the <001> directions. Simon et al. correlates PNRs with the clusters appearing
Liu et al. suggested that development of the long-range fer- around substituting ions. However, it does not explain the
roelectric order in BaTiO3 is due to transverse correlations polar nature of the PNRs, i.e. how nonpolar inclusions like,
between <001> chain dipoles.114 In the relaxor compositions, for example, Sn4+ can create a polar cluster. To account for
the dopant ions set up random local electric and strain fields, the different degrees of crossover to the relaxor state in cases
suppressing transverse correlations of these <001> chain of isovalent and heterovalent substitutions, the authors have
dipoles and frustrating the development of long-range ferro- considered a perturbation strength exerted by the introduced
electric order.114 This interpretation of the diffuse electron ions. This phenomenological parameter is not related to any
scattering has been criticized by Miao et al.,109 who point physical or chemical property of the substituting ions,
out that diffuse scattering is a general phenomenon observed although.
in a broad number of ferroelectric/relaxor compositions and Considering appearance of the relaxor behavior in isova-
it is not necessarily related to the correlation length of polar lent-substituted BaTiO3, Shvartsman et al. proposed that the
order. In other words, the appearance of diffuse scattering crossover from the ferroelectric to the relaxor behavior corre-
does not directly relate to the presence of PNR. Indeed, these sponds to a transformation from a predominately broken
authors found that the electron diffraction data obtained bond-type (at a lower degree of substitution) to a predomi-
from a broad range of BTZr compositions, 0.05  x  nately random-field-type (for higher substitution) disorder.83
0.35, in which the polar order varies from the ferroelectric to Usually, the B-site substituting cations do not go off-center,
the relaxor state, show a self-consistent set of diffuse streaks giving rise to random breaking of Ti–O bonds. On the one
with the same diffraction conditions and qualitatively the hand, spatial fluctuations of the defected bonds result in
same intensity.109 fluctuations of the polar correlations, which in turn are
Summarizing, results of the different experiments probing responsible for the formation of precursor clusters, seen, for
properties of materials at the microscopic to atomic scale example, in BTSn ceramics by PFM.83 On the other hand,
indicate that the polar behavior of both pure BaTiO3 and the RFs were suggested to arise not only in compounds with
BaTiO3-based relaxors (at least in systems where the second heterovalent (due to the inherent charge disorder), but also
component is paraelectric) are substantially related to with isovalent substitutions. In the latter case, the RFs are
displacements of Ti4+ ions from the center position in the mainly related to the existence of nanoscale regions with
oxygen octahedra. enriched concentration of substituting cations.95,105 In these
(E) Origin of the Relaxor Behavior in BaTiO3-Based regions with accumulated concentration of broken bonds, the
Systems: The origins of the crossover to the relaxor state polar correlations are strongly diminished and polar domains
in BT-based compositions are now discussed. The relaxor are less likely to nucleate. However, due to the distortions
behavior requires two main “ingredients”: the existence of arising around the substituting ions, a redistribution of the
quenched RFs and of randomly interacting PNRs at temper- charges and a local formation of charged centers results.
atures much higher than Tm.29,48 Random interactions are at These are sources of local random fields. Clearly, this kind
the origin of the nonergodic glass-like behavior, while RFs of RFs is much weaker than that stemming from heterova-
prevent an equilibrium phase transition to a FE state, at lent cation substitution. Hence, the crossover to relaxor
least when the local symmetry of the order parameter is behavior requires relatively high doping levels. It was sug-
rhombohedral, that is close to be continuous.29 These two gested that ferroelectric or relaxor behavior is reflected in the
ingredients are linked mutually: nucleation of PNRs at interrelation between the three characteristic temperatures:
higher temperatures is assumed to be promoted by fluc- TD, TC, and Tf. If the strength of the random-field fluctua-
tuations of RFs.28 tions is weak, it might be insufficient to correlate the fluctuat-
It is natural to assume that the development of the relaxor ing dipole moments of the individual unit cells at high
state in hetero-substituted BaTiO3 systems is closely related temperature. In that case, the nominal Burns temperature,
to the quenched RFs arising due to the introduced charge TD, will be relatively low. If it is below the Curie tempera-
disorder. This is similar to lead-containing relaxors, like ture, no PNRs are expected and the conventional ferroelec-
PMN or (Pb,La)(Zr,Ti)O3. However, for the isovalent-substi- tric transition occurs. Even when the dynamic PNRs appear,
tuted BT-based relaxors, the role of random fields, or even their freezing temperature Tf, may lie below TC given by the
their existence, is not obvious. random bond fluctuations. In this case again, a transforma-
To explain the appearance of relaxor properties in B-site tion into the ferroelectric long-range-ordered state will take
(both isovalent and heterovalent)-substituted barium titanate, place. This scenario is assumed to be realized in composi-
Simon et al.72 recalled that the ferroelectric polarization in tions with DPT behavior, which shows both domain-like and
pure BaTiO3 results mainly from cooperative off-center shifts PNR relaxation in the dielectric spectra.73,74,83 Only if the
of the Ti4+ (B-site) cations in the oxygen cages. This is con- nominal TC lies below Tf, a transition to a NR state should
trary to the lead-containing ferroelectrics, where off-centrality take place.
of Pb2+ cations on the A-site plays a major role. In BaTiO3, The aforementioned models consider that the PNRs
substitution of Ti4+ breaks the correlations in displacement appear in the relaxor state originating from fluctuations of
of B-site cations and related long-range polar order. Simon the random fields or polarized regions around substitutional
et al. proposed a model, where each substituted atom is defects. However, there is experimental evidence that a
assumed to perturb the surrounding BaTiO3 host lattice in a dynamic polar disorder is inherent in the paraelectric state of
finite volume creating a polarized cluster.72 The size of these pure barium titanate. In particular, diffuse X-ray scatter-
perturbed clusters depends on the correlation length of the ing115,117 and electron paramagnetic resonance118 data indi-
host matrix. They grow at decreasing temperature and reach cate that the Ti4+ ions randomly occupy one of the eight
their maximal size at the Curie temperature, TC. When the equivalent off-center sites along the [111] cubic directions
concentration of substituted cations is below a critical thresh- [Fig. 5(b)]. Nuclear magnetic resonance experiments119 and
old, the number of the perturbed clusters is small, but each terahertz dielectric spectroscopy120 point out a coexistence of
January 2012 Lead-Free Relaxors 11

displacive and order–disorder features for the ferroelectric


phase transition of BaTiO3. Brillouin scattering,121 wide band
dielectric spectroscopy,122 and photon correlation spectros-
copy123 give evidence that “jumps” of Ti4+ ions in between
different positions are correlated in neighboring unit cells
resulting in polar clusters with a size of a few nanometers far
from TC and up to 200 nm123 in the vicinity of the transition
temperature. These polar clusters resemble PNRs in the
relaxors and manifest themselves in similar ways, for exam-
ple, in a deviation of the refractive index from the linear
temperature dependence,124 or in acoustic emission peaks at
temperatures attributed to their appearance (Burns tempera-
ture) and emergence of some static PNRs.125
It is still an open question, how the dynamic polar regions
observed in barium titanate above TC are related to PNRs in
BaTiO3-related relaxors. A high-frequency (f ~ 108–109 Hz) Fig. 13. Temperature dependences of the dielectric permittivity
dielectric relaxation similar to the one in pure BaTiO3 has measured at frequencies from 10 kHz to 1 MHz for (1x)KNN–x
been reported for BTSn composition relaxors with DPT SrTiO3 ceramics, x = 0.01 (1), 0.03 (2), 0.06 (3), and 0.10 (4).
behavior.76 One can assume that polar clusters in pure Reprinted with permission from Ref. 135. Copyright 2004, Elsevier.
BaTiO3 are precursors of dynamic PNRs in the derived
relaxor compounds. In this case, the crossover to the relaxor
state occurs as soon as the ferroelectric transition related to The properties of these different systems have some com-
correlated displacements of Ti4+ is instable against the ran- mon features. As the concentration of the second component
dom electric fields.28,30 Again, for isovalent substitution, the increases, TC (or Tm) shifts down to lower temperatures and
RFs are relatively weak and such a transformation requires a the peak of the dielectric permittivity becomes much broader
relatively large concentration of dopants. It has been argued than in pure KNN (Fig. 13). The samples with relatively
that the random strains resulting from chemical substitution high x show a deviation from Curie–Weiss behavior above
also affect the Ti off-center symmetries among the Ti–O6 Tm and the parameter c describing the degree of diffuseness
octahedra and impede a perfect alignment of all Ti displace- of the transition [Eq. (3)] increases. The Tt–o shifts to lower
ments as in ferroelectric BaTiO3.111 One can also assume values and the anomaly in the temperature dependence of
that the effect of isovalent substitution is mainly related to a the dielectric permittivity becomes strongly diffuse with pro-
gradual reduction of the Curie temperature. In that case, the nounced frequency dispersion (Fig. 13). Above a certain con-
transformation to the relaxor state occurs in compounds centration of the second component. the low-temperature
whose nominal Curie temperature falls below the tempera- transition cannot be revealed from the dielectric data. A
ture corresponding to the freezing of PNR relaxation which change of the crystallographic symmetry from orthorhombic
then occurs earlier. In some cases of A-site substitutions, for to tetragonal has been established by X-ray diffraction, for
example, (1x)BT–x(Na0.5Bi0.5)TiO3, TC increases in compo- example, for KNN–BaTiO3 (x = 0.06)134 and KNN–SrTiO3
sitions with larger x and no relaxor behavior occurs at least (x = 0.04)135. At a high enough degree of substitution, for
up to x = 0.30.98 example, x  0.15 and x  0.10 for KNN–SrTiO3135 and
(F) KNN-Based Relaxors: Among the most promising KNN–Ba0.5Sr0.5TiO3140, respectively, the peak position, Tm,
lead-free piezoelectrics are materials based on (K0.5Na0.5) shows relaxor-like frequency dispersion (Fig. 13). The macro-
NbO3 (KNN) and (Bi0.5Na0.5)TiO3 (BNT).126–128 Search for scopic symmetry below Tm remains cubic. The polarization
new compounds with improved electromechanical response hysteresis loops become slimmer with increasing addition of
has resulted in the discovery of a number of systems with the second component.
relaxor behavior. The KNN-based relaxors exhibit the Like in BT-based relaxors, heterovalent substitution
following features: induces quenched RFs owing to the local charge imbalance
and the local elastic fields in KNN-derived systems. The
1. Vogel–Fulcher behavior of Tm(f) typical of relaxors. fields hinder long-range ordering and give rise to PNRs. The
2. Diffuse phase transitions. degree of transformation into relaxor behavior varies from
3. An extremely broad maximum in e(T). x ~ 0.04 in KNN–BiScO3 to x ~ 0.15 in KNN–SrTiO3. Some
We will now discuss some remarkable properties. compounds, for example, 0.96KNN–0.04BiScO3 and
The advantages of solid solutions (K,Na)NbO3 between 0.85KNN–0.15SrZrO3 show an extremely broad plateau-like
ferroelectric KNbO3 and antiferroelectric NaNbO3 are their maximum in e(T), for which, the large dielectric permittivity
relatively high Curie temperatures, several polymorphic phase (~5000) remains almost constant in intervals of 100–150 K
transitions (PPT), and the existence of morphotropic phase width at temperatures above 300–400 K, respectively.138,139
boundaries (MPBs).126–130 The compositions around 50% of Based on the similarity between this behavior and that of
NaNbO3 lie in the vicinity of a MPB between two ortho- dipolar glasses,141 Bobnar et al. have argued that the compo-
rhombic phases and show the best piezoelectric properties, a sitional mixing for these compounds is so effective that they
longitudinal piezoelectric coefficient, d33 ~ 160 pC/N, and an can be described as a glass not only on a mesoscopic level
electro-mechanical coupling factor kp up to 40%, as well as a like other relaxors, but also on a microscopic scale.138 The
high Curie temperature (693 K).131–133 Besides the paraelec- ability of these compounds, especially KNN–BiScO3, to
tric (cubic) to ferroelectric (tetragonal) phase transition at maintain large dielectric permittivity in a broad temperature
TC, other transitions between ferroelectric phases with tetrag- range makes them quite attractive for high-temperature
onal to orthorhombic and orthorhombic to rhombohedral capacitors.139
symmetry change occur at Tt–o = 473 K and To–r = 80 K, (G) Bi0.5Na0.5TiO3—“Antiferroelectric” Relaxors: Sim-
respectively.129,130 Due to the high TC, KNN is a promising ilar to KNN, bismuth sodium titanate Bi0.5Na0.5TiO3 (BNT)
candidate for high-temperature applications. Relaxor behav- is considered as a promising lead-free piezoelectric mate-
ior has been observed in systems, like (1x)KNN–xBT rial.126–128 Of particular interest are solid solutions (1x)
(x > 0.10),134 (1x)KNN–xSrTiO3 (x > 0.15),135–137 (1x) BNT–xBT, which form a morphotropic phase boundary at
KNN–xSrZrO3 (x > 0.15),138 (1x)KNN–xBiScO3 x = 0.06 separating a rhombohedral (R3c) and a tetragonal
(x > 0.04),139 etc. (P4mm) ferroelectric phase isostructural with pure BNT and
12 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

BT, respectively.142 Compositions in the MPB region show


enhanced piezoelectric properties as compared to the end
members: d33 ~ 155 pC/N, and kp ~ 37% qualifying them
for piezoelectric actuators and high-frequency ultrasonic
applications.142,143 Both BNT and BNT-based solid solutions
show such features typical of relaxor ferroelectrics52,144,145
like:
1. A broad peak of the dielectric permittivity.
2. Absence of a structural transformation at Tm.
3. Frequency-dependent anomaly of the dielectric permit-
tivity.
4. Polar nanoregions.
as well as some peculiarities as:
1. local tetragonal symmetry
2. rather AFE than ferroelectric character of PNRs
The temperature evolution of the structure of undoped
BNT is still a matter of discussions.146,147 At room tempera-
ture, it is in the ferroelectric state with rhombohedral (R3c)
symmetry.148,149 Temperature-dependent X-ray and neutron
diffraction studies reveal a change from the rhombohedral to
the tetragonal phase in a broad temperature range between
528 and 593 K,126,149 or according to other reports between
493 and 673 K,146,150 where these phases coexist, and finally,
a transition from a tetragonal to a cubic phase at 813 K.149
The temperature dependence of the dielectric permittivity
shows several anomalies [Fig. 14(a)]: a broad maximum at
Tm ~ 593 K corresponds to a transition from a paraelectric
(according some reports ferroelastic147,150) to a presumed
AFE phase,148,151 and a frequency-dependent hump at
~473 K attributed to a transition to a ferroelectric
state.152,153 As the low-temperature anomaly is accompanied
by a decay of spontaneous polarization, it is often designated
as a depolarization temperature, Td.151,152 The nature of the
state in the temperature range between 473 and 593 K is still
debated. A double polarization hysteresis loop suggests AFE
ordering between Td and Tm.142,143 Neutron diffraction
studies suggest that the AFE phase should take the P4bm
symmetry with antiparallel displacements between A-site and
B-site cations.149 On the other hand, structural investigations Fig. 14. Temperature dependences of the dielectric permittivity
measured at frequencies from 102 to 106 Hz for (a) Bi0.5Na0.5TiO3.
do not reveal any structural transition at Tm,150,154 but indi- (Reprinted with permission from Ref. 153. Copyright 1998), with
cate coexistence of rhombohedral and tetragonal phases in permission from Elsevier), (b) 0.5Bi0.5Na0.5TiO3–0.5BaTiO3, and (c)
this temperature range.150 Weak pyroelectricity and a finite 0.3Bi0.5Na0.5TiO3–0.7BaTiO3 ceramics (Reprinted from Ref. 52.
bottleneck of the double hysteresis loops indicate the exis- Copyright 2004, Elsevier).
tence of polar regions at least at the mesoscale.155 Recent
transmission electron microscopy studies suggested that the
transformation from the ferroelectric rhombohedral into the in compositions with x = 0.05 and then only weakly depends
orthorhombic AFE phase proceeds via an intermediate AFE on the BT content. The temperature corresponding to the
modulated phase consisting of orthorhombic sheets in a transition to the AFE tetragonal state, T2, rapidly drops
rhombohedral matrix in the range between 473 and down and merges with the depolarization temperature, Td, at
573 K.146 The orthorhombic phase is then formed at 573 K x = 0.02-0.03. At x  0.06, Td(x) starts to rise, while T2(x)
and immediately turns into the tetragonal one.146 The inter- continues to decrease, determining the morphotropic phase
mediate state is characterized by polar disorder which boundary (Fig. 16). The transition between paraelectric
becomes apparent in relaxor-like behavior of the low-temper- phases with tetragonal and cubic symmetry also shifts to
ature dielectric anomaly.146,150 The NBT is considered to be lower temperatures approaching Tm at x ~ 0.05.145 It
a relaxor due to a substantial frequency dispersion at Td, remains unclear whether this transition merges with the
absence of a structural transformation at Tm, and indications Tm(x) line at x > 0.05 or dissolves into an average cubic
of phase coexistence at the nanoscale.150,151,156–158 The relax- phase with local tetragonal distortion.145 The nature of the
or behavior was attributed to the presence of ferroelectric region of the phase diagram, defined by 0.05 < x < 0.11 and
rhombohedral clusters embedded in a nonpolar tetragonal Td < T < Tm is not clear yet. Based on the analysis of the
matrix150 or polar Na+TiO3 and Bi3+TiO3 clusters in an elastic compliance, Cordero et al. suggested that for x > 0.05
AFE matrix.156 and T < Td, the phase is pseudocubic with local tetragonal
The structural complexity is intimately related to cation distortions.145 The compositions from this range show relaxor-
disorder in the Na/Bi sublattice.146 It is enhanced in solid like frequency dispersion near Td, which is substantially stron-
solutions with other perovskites, for example, in BNT–BT. ger than that in the parent BNT. Sheets et al. considered the
Indeed, a strong frequency dispersion of the dielectric permit- state between Tm and Td, as a “region of frustration where
tivity around Td has been reported for (1x)BNT–xBT solid multiple domain states are simultaneously available.”159 By
solutions with 0.06  x  0.09 (Fig. 15).144 Figure 16 analogy with other relaxor systems, the frequency dispersion
shows a phase diagram of the BNT–BT system.52,142,145 On of the dielectric permittivity might be related to the existence
the BNT side, Tm corresponding to the transition to a para- of the polar nanoregions, which probably form much above
electric phase decreases from 593 K in pure BNT to ~550 K Tm.150,151 Indeed, room temperature TEM studies reveal that
January 2012 Lead-Free Relaxors 13

Fig. 17. Schematic diagram of the nanodomains responsible for the


observed relaxor AFE behavior in BNT–BT. The AFE nanodomains
with P4bm symmetry are embedded in an undistorted cubic matrix
(the white background in the diagram). The antiparallel arrow pairs
denote the cation displacements which fluctuate among the six
equivalent <001>c directions. Reprinted with permission from Ref.
144. Copyright 2010, American Institute of Physics.

being confined within individual nanodomains, which are


embedded in a matrix of an undistorted cubic lattice
(Fig. 17). It was also suggested that the frustrated relaxor
state may be induced by the competition between long-range
AFE and FE orders,160 which arise in this case due to adding
ferroelectric BaTiO3 to antiferroelectric Bi0.5Na0.5TiO3.144
For compositions 0.07 < x < 0.09, the frustration due to the
competing orders was assumed to reach a maximum, leading
Fig. 15. Temperature dependences of the dielectric characteristics
for (1x)Bi0.5Na0.5TiO3–xBaTiO3 ceramics, 0.04  x  0.11.
to the relaxor behavior persisting to low temperatures. The
Reprinted with permission from Ref. 144. Copyright 2010, American charge disorder in the A-sites being randomly occupied by
Institute of Physics. Bi3+, Na+, and Ba2+ results in the random electric fields,
which may affect the local balance between the off-center
ferroelectric rhombohedral domains with sizes around 100 nm shift of the B-site Ti4+ cation and the displacement of the
forming colonies of complex morphology seen in the composi- A-site cation in the opposite direction. The RFs will destroy
tion with x = 0.04 transform into nanodomains observed for the long-range order and promote local dipole correlations.28
samples with x = 0.07 and 0.09, which manifest local tetrago- The resulting misbalance provides weak local dipole
nal P4bm symmetry.144 Further increase in BaTiO3 content to moments dynamically fluctuating along the six equivalent
x = 0.11 leads to the appearance of large lamellar ferroelectric <001> directions. These fluctuations are assumed to be an
domains corresponding to the stabilization of the tetragonal origin of the observed frequency dispersion of dielectric per-
P4mm phase at room temperature.144 mittivity in relaxor BNT–BT compounds.144
Ma et al. proposed that the relaxor state in compositions Considering this model, Ma et al. assumed that in BNT–
with x > 0.04 has rather an AFE than ferroelectric nature.144 BT, the MPB separates rather a ferroelectric and a relaxor
They suggested that the displacements of A-site cations (Bi3+, AFE phase than two ferroelectric phases.144 In this case, the
Na+, Ba2+) are antiparallel to the displacements of the maximum of piezoelectric properties and large electric field-
B-site cation (Ti4+) within each individual nanodomain. This induced strain in BNT–BT around MPB seems to profit from
is supported by the fact the local P4bm symmetry is identical both relaxor-to-ferroelectric and the antiferroelectric-to-ferro-
to that of the AFE phase of BNT with antiparallel displace- electric electric field-induced transitions.
ments of Bi3+/Na+ and Ti4+ along the [001] direction.149 The phase diagram for large BT (x > 0.20) content has
However, the AFE ordering in BNT–BT is only short-ranged been studied less. In compositions with x  0.50, Tm is con-
tinuously shifting from ~520 K (x = 0.50) to 393 K for pure
barium titanate (Fig. 16).52 A strong frequency dispersion
was observed around Tm for compositions with 0.30 
x  0.70 indicating relaxor behavior [Figs. 14(b) and
(c)].52,161,162 Similar to the canonical relaxor, this peak does
not correspond to any change of macroscopic symmetry, but
is attributed to the relaxation of PNRs with local tetragonal
symmetry embedded in a nonpolar matrix.162 Contrary to
the pure NBT and NBT–BT with small concentration of bar-
ium titanate, the low-frequency anomaly at Td is nearly fre-
quency-independent. It corresponds to the transition to a
ferroelectric state with tetragonal symmetry.161,162 Com-
pounds with BNT content between 15% and 30% show a
DPT behavior characterized by a negative difference between
Tm and the Curie–Weiss temperatures as compared to a posi-
Fig. 16. Phase diagram of (1x)Bi0.5Na0.5TiO3–xBaTiO3 system tive one for compositions closer to the BaTiO3 side.98
plotted based on results of dielectric, ferroelectric, and anelastic The properties of the BNT–BT system are unusual in com-
measurements. The data were taken from Refs. 142,144,145 for the parison with other BT-derived relaxors. The Curie tempera-
concentration range 0.00  x  0.20 and for the concentration ture increases when coming from the BT side, while most
range 0.50  x  1.00 from Refs. 52,98,161,162. other systems show the opposite trend. In spite of the hetero-
14 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

valent substitution, the relaxor behavior appears only at a tures.181 Upon heating, it transits to an AFE orthorhombic
fairly high concentration of BNT. This may be related to a (Pbma) state at ~230 K. This first-order phase transition
weaker effect of doping only on the A-site on the correlated exhibits a large thermal hysteresis, ~100 K.182 Several transi-
shifts of the Ti4+ ions.15 Low-temperature X-ray diffraction tions between AFE states with different orthorhombic sym-
studies indicate that two PPT, typical of BT, are suppressed metry follow. At ~753 K, a transition to the paraelectric
already in compounds with 5% of BNT. This is probably orthorhombic (Pnmm) phase takes place, which transform
due to the effect of the local strains generated by doping with further to orthorhombic (Ccmm), tetragonal (P4/mbm), and
cations of different sizes.98 The tetragonal symmetry of the finally, cubic (Pm3m) at 793, 848, and 913 K, respec-
ferroelectric state is stabilized by BNT doping contrary to tively.181,183 The temperature dependence of the dielectric
the rhombohedral one for other BT-based relaxors. permittivity shows a rather sharp maximum at Tm ~ 625–
The relaxor behavior has been reported also for other 645 K associated with a first-order phase transition between
binary systems containing BNT, for example, in (1x)BNT– two AFE phases (Pbma and Pmnm).182,184,185 Neutron dif-
xSrTiO3 (x > 0.25),52,163,164 xBNT–(1x)NaNbO3 (x  fraction186 and Raman scattering187 data indicate the coexis-
0.175),165 (1x)BNT–xBiFeO3 (x < 0.50),166 (1x)BNT–xBi0.5 tence of the ferroelectric and AFE phases in a broad
Li0.5TiO3 (x < 0.10),167 (1x)BNT–xBa(Al1/2Nb1/2)O3 (x > temperature range from 70 to 250 K. Competing ferroelec-
0.045),168 and (1x)BNT–xBi0.5K0.5TiO3 (x > 0.06).169 Prop- tric and AFE interactions were suggested to result in the
erties of the BNT-based binary solid solution were systemized frustration of local dipole ordering and in the relaxor-like
by Hiruma et al.170 Three types of phase diagrams were con- behavior.160 Lanfredi et al. reported on smeared dielectric
sidered. For both type A (e.g., BNT–NaNbO3) and type B relaxation peaks (much smaller than the main maximum at
(e.g., BNT–KNbO3, BNT–SrTiO3) solid solutions, the range ~630 K) showing frequency dispersion over the temperature
of existence of the relaxor state broadens significantly for range of 80–250 K.188 Theoretical calculations suggest that
increasing concentration of the second component. The BNT– the free energies of the ferroelectric and AFE phases over
KNbO3 and BNT–SrTiO3 show relaxor behavior at room this temperature range are quite close. The small energy dif-
temperature in compositions with x > 0.08 and x > 0.28, ference between the two phases makes it possible to easily
respectively. The range where the rhombohedral and tetrago- switch from the AFE to the ferroelectric state at moderate
nal phases coexist also broadens for the type A solid solu- electric fields.186 The NN is such an example where also
tions. However, it disappears in solid solutions of the type B small amounts of dopant ions, like Li+ or K+, will induce a
and C. For the latter (e.g., BNT–BT, BNT–PbTiO3, BNT– transition.180 There is evidence that local fluctuations of Na
(Bi0.5K0.5)TiO3), the relaxor state is stabilized at higher tem- content or oxygen vacancies are responsible for the coexis-
perature, while the low-temperature state is ferroelectric with tence of the AFE and ferroelectric phases in NaNbO3 single
the rhombohedral (R3c) and tetragonal (P4mm) symmetries crystals at room temperature and even above.189
for compositions with small and large x, respectively. The cor- The solid solutions based on NaNbO3 can be divided into
responding MPBs lie at x ~ 0.06 (BNT–BT), x ~ 0.20 (BNT– two groups.180 In the first, the high-temperature ferroelectric
BKT), and x ~ 0.14 (BNT–PT). state is already stabilized at a few mol% of substitution, the
Relaxor properties have also been reported for ternary sys- temperature dependences of the dielectric permittivity are
tems, like Bi0.5Na0.5TiO3–K0.5Bi0.5TiO3–BaTiO3 (Ref. 171), sharp, and the transition temperature smoothly depends on
Bi0.5Na0.5TiO3–NaNbO3–BaTiO3 (Ref. 172), Bi0.5Na0.5TiO3– the degree of substitution. This occurs when the second com-
K0.5Bi0.5TiO3–NaNbO3 (Ref. 173), Na0.5Bi0.5TiO3–Bi0.5K0.5 ponent is either a ferroelectric or paraelectric with tolerance
TiO3–BiCoO3 (Ref. 174), Bi0.5Na0.5TiO3–BaTiO3–K0.5Na0.5 (Goldschmidt) factor t > 1, t = (RA + RO)/√2(RB + RO),
NbO3 (Ref. 175). Compounds belonging to the latter system where RA, RB, and RO are the radii of A-, B-site, and oxygen
exhibit relatively large and virtually temperature-insensitive ions, respectively.190 Examples are (Na,Li)NbO3 and (Na,K)
dielectric permittivity (De/e425 K < 10% from 316 to 592 K, NbO3.180,191 In the second group, the AFE state remains sta-
e425 K ~ 2000)175 and giant unipolar strain (0.45%).53 These ble up to a comparatively high degree of substitution. The
make them promising for applications in high-capacitance peak position Tm decreases smoothly up to a certain thresh-
capacitors operating at high temperatures and electrome- old concentration x = x0. For x > x0, an abrupt drop of Tm
chanical devices. The superior properties originate from the along with a drastic broadening of the e(T) maximum is
relaxor nature of the BNT–BT–KNN compounds. The large observed [Fig. 18(b)]. Compositions with x > x0 exhibit a fre-
unipolar strain appears to be a consequence of the presence quency dispersion of Tm typical of relaxors.179,180 Such
of nonpolar regions at zero electric field destabilizing and behavior is observed when the second component is either an
randomizing the electrically induced ferroelectric order. AFE or a paraelectric with t < 1.190
These nonpolar regions drive the system back into its unp- Both x0 and the degree of the frequency shift of Tm
oled state once the applied electric field is removed.176 depend on the type of substitution. For the isovalent-substi-
(H) NaNbO3-Based Relaxors: Sodium niobate, NaN- tuted solid solutions, like, for example, (1x)NaNbO3–xNa-
bO3 (NN), is another lead-free base for relaxor compounds, TaO3, x0 is about 0.45, and the dielectric peak position only
such as heterovalent-substituted (1x)NaNbO3–xBaSnO3 shifts by a few K for two frequency decades indicating rather
(x > 0.1),177 (1x)NaNbO3–xLiNbO3 (0.01  x  0.04),178 a DPT than relaxor behavior.179 On the contrary, heterova-
(1x)NaNbO3–xBi0.5Na0.5TiO3 (x  0.175),165 (1x)NaNb lent cation substitutions, like, for example, in (1x)NaNbO3–x
O3–xSrCu1/3Nb2/3O3 (x  0.125),179 (1x)NaNbO3–xGd1/3 SrCu1/3Nb2/3O3 (Ref. 179) and (1x)NaNbO3–xNa0.5Bi0.5
NbO3 (x > 0.15),180 etc. TiO3 (Ref. 165), or heterovalent substitution accompanied by
Basic features of these systems are: formation of cation vacancies like in (1x)NaNbO3–xGd1/
3NbO3 (Ref. 180) result in a substantial frequency dispersion
1. A frequency dispersion of Tm above certain concentra-
at x > x0 (e.g., DTm ~ 40 K for 0.7NaNbO3–0.3Na0.5Bi0.5
tion threshold typical of relaxors.
TiO3)179 Similar to the canonical relaxor, Tm(f) dependences
2. A DPT in isovalent-substituted solid solutions.
can be well described by the Vogel–Fulcher law. The devia-
3. The extrapolated Curie–Weiss temperature lies below
tion from the Curie–Weiss behavior is observed in a broad
Tm.
temperature range above Tm for all these solid solutions.
The parent composition NaNbO3 exhibits a complex In the case of NaNbO3-derived relaxors, the extrapolated
sequence of phase transformations, for which, the number of Curie–Weiss temperature, h, is much lower than Tm. This
transitions as well as the symmetry of the individual phases is in contrast to the behavior of other relaxor systems
are still a matter of discussion. According to the most (PMN- or BaTiO3-based relaxors), where h is usually larger
accepted scheme, NaNbO3 is in a ferroelectric state with than Tm. It was suggested that this difference is a conse-
rhombohedral symmetry (R3c) at liquid helium tempera- quence of the AFE character of pure NaNbO3.180
January 2012 Lead-Free Relaxors 15

Fig. 18. (a) Temperature dependences of the dielectric permittivity


of (1x)NaNbO3–xSrCu1/3Nb2/3O3 (b) Concentration dependences
of the Tm for different binary NaNbO3-based solid solutions showing
a crossover to relaxor behavior. The second components of the solid
solutions are: (1) SrCu1/3Nb2/3O3, (2) Gd1/3NbO3, (3) Bi0.5Na0.5TiO3,
(4) Sr0.5NbO3, and (5) NaTaO3, respectively. Reprinted with
permission from Ref. 179. Copyright Taylor & Francis Group,
http://www.informaworld.com.

(2) Lead-Free Relaxors with Two-Dimensional Order


Parameter: Compounds with the Aurivillius Structure Fig. 19. Aurivillius crystal structure, BaBi2Nb2O9, some Ba2+
Bismuth layer structured ferroelectrics (BLSF) were synthe- substitute Bi3+ in [Bi2O2]2+ layers.
sized and described for the first time by Aurivillius.192 They
have a very anisotropic structure consisting of oxygen octa- different valence occupy the same positions in the perovskite
hedral (perovskite-like) blocks interleaved with (Bi2O2)2+ blocks of the Aurivillius structure, like K0.5La0.5Bi2Nb2O9,206
layers (Fig. 19). The c parameter, along the normal to the and K0.5La0.5Bi2Ta2O9,207 form the second group. Compounds
fluorite-like (Bi2O2)2+ layers, is much larger than a and b where the relaxor behavior is induced by substitutions for Bi3+
parameters of the orthorhombic or tetragonal unit cell. The in the (Bi2O2)2+-layers, for example, SrBi1.6Nd0.4Nb2O9,208
general chemical formula of the BLSF compounds is SrBi1.65La0.35Nb2O9,209 Bi2LaTiNb2O9,210 SrBi3LaTi4O15,211
(Bi2O2)2+(Am1BmO3m+1)2-, where m is an integer between 1 and Bi2.75La1.25Ti3O12,211 may be put into the third group.
and 5 indicating the number of the perovskite-like blocks All these compositions again show typical relaxors behav-
between two adjacent (Bi2O2)2+ layers. 12-coordinated A- ior (frequency dispersion of the dielectric permittivity, Tm(f)
sites are usually occupied by Pb2+, Bi3+, alkaline, or alka- dependence following the Vogel–Fulcher law, and deviation
line earth cations, whereas in 6-coordinated B-sites cations of of the e(T) dependences from the Curie–Weiss law above
transition metals, such as Ti4+, Nb5+, or Ta5+, are found. Tm). Some characteristic parameters for various Bi-layered
The structural anisotropy of the Aurivillius compounds relaxors are listed in Table IV. Contrary to canonical perov-
results in highly anisotropic electric properties.193,194 Namely, skite relaxors like PMN, it is difficult to induce long-range
the spontaneous polarization lies predominately in the (a,b)- order in Bi-layered relaxors with m = 2 even when applying
plane being quasi two-dimensional.194 The interest to investi- strong electric fields.199 Typically, no polarization nonlinear-
gate BLSF compounds is motivated by ity and apparent ferroelectric hysteresis are found for these
materials below Tm.199,212,213 However, BaBi4Ti4O15 (m = 4)
1. their fairly high Curie temperature, for example,
and Ba2Bi4Ti5O18 (m = 5) exhibit remanent polarization val-
875 K for Bi4Ti3O12 and 928 K for Na0.5Bi4.5Ti4O15
ues ~15 and 12 lC/cm2, respectively.214 It was argued that
(Ref. 195),
the remanent polarization is large in BLSFs having high
2. relatively low dielectric permittivity, and
Curie temperatures. Indeed, Tm is much larger for BaBi4-
3. large anisotropy in the electromechanical coupling fac-
Ti4O15 and Ba2Bi4Ti5O18 than for the relaxors with m = 2.
tor opening possibilities to use them for high-tempera-
The coercive electric field is smaller in compounds with a lar-
ture piezoelectric applications.128 Moreover, their
ger value for m supposedly due to the decrease in the strain
4 excellent stability against repetitive polarization switching
energy of the BO6 octahedrons from the bismuth layer, lead-
(fatigue endurance)196 make them major candidates for
ing to an easy movement of the octahedral cations in the
the development of nonvolatile ferroelectric memories.197
direction of applied external electric field.214 This facilitates
Relaxor compositions of the BLSF material family can be poling these compounds.
formally divided into three groups. To the first group belong The solid solutions between BLSFs with ferroelectric
compounds with the A-site completely or partially occupied and relaxor properties, for example, (1x)SrBi2Nb2O9–x
by Ba2+ ions BaBi2Nb2O9198–201, BaBi2Ta2O9,198,202 BaBi4 BaBi2Nb2O9215,216 and (1x)CaBi2Nb2O9–xBaBi2Nb2O9217
Ti4O15,203,204 and Ba2Bi4Ti5O18.205 Materials where cations with or compounds from the third group with small concentration
16 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

Table IV. Dielectric and Relaxor Characteristics of Some Relaxors with the Aurivillius-Type Structure. Parameters are the Same
as for Table I
Composition em (105 Hz) Tm, K (105 Hz) DTm, K (103–106 Hz) c Tf , K Ea, eV f0, Hz Ref.
13
BaBi2Nb2O9 550 450 115 1.94 97 0.63 10 199
390 460 90 100 0.56 4.7 9 1013 217
BaBi2Ta2O9 525 345 60 113 0.25 8 9 1012 202
Sr0.5Ba0.5Bi2Nb2O9 669 506 40 1.6 371 0.194 1.5 9 1012 216
Ca0.2Ba0.8Bi2Nb2O9 380 583 32 1.73 525 0.11 1.54 9 1011 217
K0.5La0.5Bi2Nb2O9 480 480 27 2 428 0.04 1010 206
K0.5La0.5Bi2Ta2O9 387 437 11 1.93 396 0.06 1.3 9 1013 207
BaBi4Ti4O15 2350 685 20 1.9 635 0.02 5.8 9 1011 203
1150 675 35 1.7 204
BaBi4Ti3Fe0.5Nb0.5O15 750 580 16 558 0.002 6.5 9 108 282
Ba2Bi4Ti5O18 2100 625 30–40 205
SrBi1.65La0.35Nb2O9 258 339 9 326 0.0088 2.2 9 109 209
SrBi1.8Pr0.2Nb2O9 471 290 7 1.45 224
SrBi1.6Nd0.4Nb2O9 195 342 15 331 0.012 109 208
Bi2LaTiNbO9 225 443 65 1.98 296 0.15 1.7 9 1011 210
Bi2.75La1.25Ti3O12 310 250 15–20 211
SrBi3LaTi4O15 413 510 3 211
Sr2Bi3.5La0.5Ti5O18 470 380 6 211
SrBi6.5La1.5Ti7O27 660 640 5 211

of substituting elements208,210,211 show a continuous cross- Analysis of the frequency dispersion of the dielectric per-
over from ferroelectric to relaxor behavior. Chen et al. have mittivity indicates some substantial differences between
analyzed the threshold La content, x0, which is necessary to BaBi2Nb2O9 and BaBi2Ta2O9 on the one hand and canonical
induce relaxor behavior in Sr-containing BLSFs with differ- relaxors, like PMN, on the other. Namely, the freezing tem-
ent numbers of perovskite-like blocks, m.211 The x0 decreases perature is much below the maximum in e(T) dependences,
both by increasing m and by increasing relative Sr content in Tm(105 Hz)  Tf ~ 350 K in BaBi2Nb2O9 [Ref. 199] against
the perovskite units. ~80 K in PMN. The activation energy, Ea, is about one
There are only few reports considering the influence of the order of magnitude larger in comparison with the typical
crystallographic anisotropy on relaxor properties in BLSF value for PMN-based relaxors. On cooling, the relaxation
compounds. Irie et al. reported on electrical properties of time distribution shifts to larger times, but without significant
Ba2Bi4Ti5O18 single crystals.214 Yan et al. have studied the broadening typical of canonical relaxors.199,202 These factors
orientation dependence of the relaxor behavior for grain-ori- point out a relatively small size of polar clusters and weak
ented BaBi2Nb2O9 ceramics.218 In both cases, the relaxor interaction between them in Bi-layered relaxors with m = 2,
parameters, such as DTm(f), the freezing temperature, and when only two perovskite layers are located between the
the activation energy of the Vogel–Fulcher fit are isotropic, (Bi2O2)2+ layers. The growth of PNRs is probably restricted
in spite of a strong anisotropy of dielectric permittivity and to two dimensions, in the (a,b) plane. Therefore, they remain
polarization. small and the transition to a NR state is delayed. The situa-
Similar to materials with the perovskite structure, the tion can be different in BLSF compounds with larger m,
relaxor behavior in the Bi-layered compounds is attributed to where the larger number of perovskite layers may result in a
a structural disorder, which is either inherent or introduced crossover from quasi-two-dimensional to three-dimensional
by doping. The first situation is realized in BLSF relaxors growth of PNRs. Indeed, BaBi4Ti4O15 and Ba2Bi4Ti5O18
with Ba2+ cations on the A-sites. Withers and co-workers show DTm, Tm  Tf, and Ea comparable to values observed
suggested that the major contribution to ferroelectric polari- for PMN and related relaxors. Moreover, the polarization
zation in BLSFs with m = 2 is due to the displacement of A- hysteresis loops below Tf indicate that a transition to the fer-
cations in the perovskite units.219,220 Hence, disorder on the roelectric state occurs at low temperatures. The interactions
A-sites may significantly affect the polar properties of these between PNRs are substantially enhanced as compared to
materials. It was originally proposed by Smolenskii et al. BaBi2Nb2O9 and BaBi2Ta2O9.
that the relaxor behavior of BaBi2Nb2O9 and BaBi2Ta2O9 Structural analysis shows that Bi3+ and Ba2+ ions have
originates from anti-site defects, where the Bi3+ ions enter specific anionic environments in BaBi4Ti4O15 and cannot
A-sites, while Ba2+ are incorporated into (Bi2O2)2+ layers share identical atomic positions.223 The strong overbonding
(Fig. 19).198 X-ray and neutron diffraction investigations observed for the Ba2+ cations on A-sites reflects structural
have confirmed this assumption221,222. It was shown in calcu- stresses in the structure.221,223 A certain stress relief can be
lations that when Ba2+ occupies the A-site exclusively, the assured considering local configurations with compositional
corresponding valence is larger than two and therefore fluctuations, i.e. Ba- and Bi-rich zones.223 Existence of such
“overbonding” of the A-site cation occurs, whereas cation regions with different degrees of structural distortion, where
disordering reduces this overbonding.221 Ba-rich zones are potentially nonpolar, can also be responsi-
When a bismuth ion enters an A-site, it can easily go off- ble for the relaxor behavior.
center (contrary to Ba2+) leading to local polarization and In K0.5La0.5Bi2Nb2O9 and K0.5La0.5Bi2Ta2O9, selected area
stimulating the appearance of PNRs. Concurrently, the cat- electron diffraction reveals a local 1:1 cationic ordering on
ion redistribution also leaves positive point charges on the A-sites, which is driven by the substantial size and charge
perovskite blocks where Bi3+ replaces Ba2+ and negative differences between K+ and La3+ (ionic radius RLa3+ =
point charges in the fluorite-like layers where Ba2+ replaces 0.136 nm and RK+ = 0.164 nm).206 It was argued that simi-
Bi3+. The RFs resulting from these point charges as well as lar to PMN and related compounds, where B-site ordering
mechanical stresses due to inhomogeneous cation distribution has been observed, these chemically ordered regions might be
effectively break up the long-range interaction and prevent the origin of local RFs promoting the relaxor behavior in
macroscopic ferroelectric ordering. these BLSF compounds.207
January 2012 Lead-Free Relaxors 17

The mechanism of the transformation from the relaxor However, contrary to the lead-free perovskite relaxors
into ferroelectric behavior in the case of substitution of Bi3+
1. the low-temperature state is rather long-range than
with lanthanides in fluorite-like (Bi2O2)2+ layers yet remains
short-range-ordered
unclear. Raman spectroscopy revealed a pronounced cation
2. A-site disorder plays the crucial role in uniaxial TTB
disorder and related compositional inhomogeneity in the
relaxors, while
fluorite-like layers with increasing La3+ content.211,213,224,225
3. polar behavior is still attributed to the B-site cations
It seems that such type of disorder deteriorates ferroelectric
like in the structurally other systems.
long-range ordering in BLSF compounds.225 This is likely
related to the fact that in spite of the same charges and very Like in materials with the perovskite and Aurivillius-type
similar ionic radii of Bi3+ and La3+, the latter ion does not structures, a prerequisite of relaxor behavior in TTB com-
possess a stereochemically active lone electron pair. Conse- pounds is structural disorder related to the occupation of the
quently, La3+ would prefer to be more central than the Bi3+ same crystallographic site by different ions (Table V). This
position. Such centralization would result in an additional may be A-site disorder, like in Sr5xBa5(1-x)(Nb10O30 (x 
need for contraction of the ideal a-parameter of the fluorite- 0.6)41,228 and Ba2Bi22Nb10O30 (Ref.229) (here  denotes a
like unit.226 This destabilizes the perfect [BiLaO2]2+ disorder vacant site) when Sr and Ba (Bi and Ba) occupy the A2-site,
resulting in a marked preference of La3+ for the A-site of or B-site disorder, like in Ba6M3+Nb9O30 (M = Ga, Sc, In,
the perovskite block instead.226 In such a way, La-doping or Co),230–232 Ba2ReTi3Nb2O15 (Re = Bi3+ or La3+),233 and
enhances the disorder on the A-site, which to a certain extent Sr2K(Nb1xTax)5O15 (0.16  x  1).16 The relaxor behav-
even exists in ferroelectric BLSF compounds being related to ior was also reported for oxyfluorides with the TTB struc-
anti-site defects, when part of the A-cations enter (Bi2O2)2+ ture, where the substitution of F for O2 is coupled with a
layers, while the Bi3+ ions occupy A-sites.221,222 cationic substitution to ensure electroneutrality. In this case,
relaxor behavior is provoked either by cationic disorder like
in BaNa2Nb5O14F15 with the A2-site being filled both by Ba2+
(3) Uniaxial Relaxors with Tetragonal Tungsten Bronze and Na+ or by solely anionic disorder like in K3LiNb5O14F
Structure (here K+ and Li+ occupy the A- and C-sites, respectively).16
Relaxor behavior has been observed in a number of materials Among the most studied relaxor TTB systems is Sr5x
crystallizing in the tetragonal tungsten bronze (TTB) struc- Ba5(1x) Nb10O30 or otherwise SrxBa1xNb2O6 (SBN). For
ture.15,16 The TTB structure A12A24C4B12B28X30 consists of SBN, the A1 positions are occupied by Sr2+ only, the A2-
layers of BX6 octahedra sharing corners to form three types sites are filled with both Sr2+ and Ba2+, whereas the C chan-
of interstitial channels: square A1, pentagonal A2, and trian- nels remain empty (Fig. 20). As there are only five Sr and Ba
gular C channels (Fig. 20).227 The pentagonal and square atoms for six A1 and A2 positions, one of the A-sites
channels are usually occupied by large alkali, alkaline earth, remains unoccupied.234
and rare earth cations (e.g., Na+, K+, Sr2+, Ba2+, La3+, Upon cooling, SBN undergoes an order–disorder phase
Bi3+, etc.). The smaller C channels are either empty or filled transition from a paraelectric 4/mmm to a polar 4mm state
by small cations like Li+. The highly charged cations of with the spontaneous polarization oriented parallel to the c-
transient metals, like Ti4+, Nb5+, and Ta5+, reside on octa- axis.235 The polarization is caused by the off-center displace-
hedral B-sites. One of the B-site cations in the octahedral site ments of the B-site (Nb5+) cations. For increasing Sr/Ba
has to be ferroelectrically active. The presence of crystallo- molar ratio, a crossover from ferroelectric to relaxor behavior
graphically nonequivalent A- and B-sites as well as extra C- occurs in compositions with x > 0.6.41,228 Single crystal
sites provides extra degrees of freedom for the manipulation SBN40 (x = 0.40) exhibits a relatively sharp peak in the tem-
of the structure offering huge flexibility of composition and perature dependence of e(T), whose position, Tm  410 K,
properties.16 does not shift with probing frequency as is typical of ferroelec-
The TTB-relaxors display the following properties similar trics. The Tm coincides with the Curie temperature, TC.228,236
to other relaxors: For SBN61 and SBN75 single crystals, the maximum of the
dielectric susceptibility is much broader and the peak position
1. Typical Tm(f) shift following the Vogel–Fulcher law.
shifts to lower temperatures when decreasing the frequency of
2. Dynamic PNRs much above Tm.
the probing field as is typical of relaxors.228,237
3. Coexistence of static and dynamic PNRs at Tm.
The crossover to the relaxor behavior in SBN with larger
x is considered to be related to an enhancement of structural
and related charge disorder.41,238 This disorder is primarily
due to randomly distributed vacancies on A-sites. The miss-
ing charges on these vacancies are the most intense sources
of RFs. Occupation of A2-sites by different cations intro-
duces disorder of the oxygen ion positions due to different
Ba–O and Sr–O bonding lengths, whereas the presence of
vacancies on both A1 and A2-sites enhances this disorder.234
Accommodation misfits of the different oxygen octahedra
give rise to local buckling and tilting deformation creating
localized electric multipole moments. These are also sources
of random electric fields. More structural disorder results in
stronger RFs. From statistical considerations, the most
ordered structure is expected in SBN20, where all A2-sites
are occupied solely by Ba2+ cations, while the Sr2+ ions and
vacancies are randomly distributed on the A1-sites. In other
compositions the entropy of the cation distribution, which is
a measure of the structural disorder, increases with increasing
Sr/Ba ratio and reaches a maximum level at x ~ 0.65.238 It
has also been argued that vacancies on A2-sites have a stron-
Fig. 20. Tetragonal tungsten bronze crystal structure, ger impact on the polar properties of SBN, than those at
Sr0.61Ba0.39Nb2O6, A1-sites are occupied by Sr, A2-sites by Sr and A1-sites.238 As the vacancy tends to increasingly occupy the
Ba, among six A-sites one is vacant (here the A2-site); triangular C- A2-sublattice as x increases,238,239 the corresponding “active”
channels are empty. charge disorder will be enhanced at large Sr2+ content. The
18 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

Table V. Dielectric and Relaxor Characteristics of Some Relaxors with the Tetragonal Tungsten Bronze Structure. Parameters
Have the Same Meaning as for Table I
Composition Tm, K (103 Hz) DTm, K (102–105 Hz) Tf, K Ea, eV f0, Hz Ref.

A-site K2LaNb5O15 165 22 16


disorder K2BiNb5O15 220 80 15
BaLaNb5O15 203 55 229
BaBiNb5O15 295 75 107 0.33 4.7 9 1011 229
BaLa2/31/3NaNb5O15 210 34 16
Sr3Ba2Nb10O30 320 12 299 0.028 5.8 9 109 237
Sr3.75Ba1.25Nb10O30 361 3 347 0.014 8 9 108 253
B-site Ba2Bi(Nb3Ti2)O15 200 30 229
disorder Ba2La(Nb3Ti2)O15 190 20 233
Sr2La(Nb3Ti2)O15 221 14 190 0.046 2.1 9 1010 283
Sr2Nd(Nb3Ti2)O15 243 35 148 0.17 8.5 9 1011 260
Ba6GaNb9O30 80 20 56 0.077 6.7 9 1011 230, 231
Ba6InNb9O30 280 50 158 0.198 4.1 9 1011 230, 231
Ba6ScNb9O30 230 25 153 0.105 9.1 9 1010 230, 231
Ba6CoNb9O30 135 70 20 0.31 3.6 9 1013 232
Sr2Na(Nb1xTax)5O15 222 (x = 0.4) 20 (x = 0.40) 15
(0.40  x  1) 120 (x = 1) 25 (x = 1)
Sr2K(Nb1xTax)5O15 263 (x = 0.16) 21 (x = 0.16) 16
(0.16  x  1) 80 (x = 1)
Ba2K(Nb1xTax)5O15 268 (x = 0.35) 16 (x = 0.35) 15
(0.35  x  1) 80 (x = 1)
A- and Ba5SmTi3Ta7O30 150 25 117 0.045 4.1 9 109 284
B-site Sr5LaNb7Ti3O30 240 20 184 0.11 5.3 9 1012 261
disorder
X-site K3LiNb5O14F 210 40 16
disorder
A- and Ba4.51.5Nb5O29F 175 30 16
X-site Ba2xNa1+xNb5O15xFx 250 (x = 0.31) 40 (x = 0.31) 285
disorder (0.31  x  1) 140 (x = 1) 17 (x = 1)
Sr2xK1+xNb5O15xFx 290 (x = 0.20) 15 (x = 0.2) 15
(0.20  x  1) 112 (x = 1) 16 (x = 1)
B- and Ba2NaNb5xTixO15xFx 265 (x = 0.31) 15 (x = 0.31) 285
X-site (0.31  x  0.50) 168 (x = 0.50) 30 (x = 0.50)
disorder
 denotes vacant site.

enhancement of structural and charge disorder is reflected in critical slowing-down, the dynamic PNRs are expected to
changes of the polar structure of SBN.238 merely transform into a random field controlled metastable
Even for ferroelectric SBN compositions, the antiparallel domain state.247 Indeed, statistics and topology of experi-
domains are substantially distorted having fractal-shape mentally observed domain structures in SBN below the tran-
boundaries.240 They are strongly anisotropic with a size sition temperature corroborate the theoretical predictions for
along the polar axis of several tens of micrometers, which is RFIM systems.240
about two orders of magnitude larger than that perpendicu- The evolution of PNRs in the vicinity of the transition
lar to the polar axis.240,241 The shape of the domains espe- temperature has been probed by piezoresponse force micros-
cially in the plane normal to the polarization vector is copy and dielectric spectroscopy.240,247,248 The transition
dictated by the random electric fields, which pin and distort from the high-temperature ER state to the low-temperature
domain walls, thus forming a complex pattern. The polar dis- long-range ferroelectric state does not take place at a fixed
order is enhanced in compositions with larger x correspond- temperature. There is a certain temperature range above TC
ing to a strengthening of RFs with increasing Sr content where the system contains both small dynamic and large
(Fig. 21). Namely, the domains become smaller; the mean quasi-static PNRs. The latter ones are precursors of domains
domain size in SBN75 is only about 75 nm.240 Simulta- in the low-temperature ferroelectric state. The intermediate
neously, the fractal dimension of domain boundaries state has a nonergodic character, which was verified by the
increases, which indicates more pronounced “irregularity” of observation of aging (via an isothermal decay of the dielec-
the polar structures in relaxor SBN.238 tric permittivity) in the same temperature range above TC,
Like in other relaxor systems, the relaxor behavior of SBN where frozen PNRs were visualized.248,249 This aging is
is attributed to the existence of PNRs, which is evidenced related to isothermal growth of quasi-static PNRs at the
above Tm by the measurements of the optical index of refrac- expense of the dynamic ones.
tion,242 birefringence,243 and dynamic244 and inelastic Brill- The appearance of frozen PNRs in the close vicinity above
ouin245 light scattering. The uniaxial character of the order TC may explain a long time disputed controversy between
parameter together with the existence of the quenched RFs the experimentally observed critical behavior in 3D-RFIM
due to the randomly distributed A-site vacancies allows con- systems and theoretical predictions.247,250 Namely, the values
sidering SBN to be a member of the three-dimensional ran- of critical exponents found in experiments are close to those
dom-field Ising model (3D-RFIM) universality class.29,246 In theoretically predicted for the two-dimensional Ising model,
this case, contrary to the perovskite relaxors, a transition to but deviate strongly from those of the 3D-RFIM sys-
a long-range-ordered ground state is predicted below the tems.250,251 Kleemann et al. suggested that the appearance of
Curie temperature, TC. In reality, however, due to extreme quasi-stable PNRs excludes true equilibrium criticality when
January 2012 Lead-Free Relaxors 19

(a) (b)

(c) (d)

Fig. 21. Piezoresponse force microscopy images taken on c-cut SrxBa1xNb2O6 single crystals below the transition temperatures. Compositions
with x = 0.40 (a) and 0.50 (b) show the ferroelectric behavior, while those with x = 0.61 (c) and 0.75 (d) the relaxor one (Ref. 240).

approaching TC.247 While these large PNRs are frozen on a modulated structure, respectively. While the incommensurate
finite time scale, the unfrozen interfaces can be considered as modulation has been often reported for TTB structures, the
regions with very short-ranged correlation of RFs, which are commensurate phase has not been observed before. It was
not able to select sizable polar clusters via field energy gain. argued that both the type of the oxygen octahedra tilting and
In these essentially 2D regions forming a percolating network the periodicity of the tilting modulation (commensurate or
through the sample, a phase transition may take place under incommensurate) are determined by the competing bonding
the constraint of a weakly disordered quasi-staggered field. requirements of the cations in the pentagonal (Ba) and square
At the same time, the 2D Ising model criticality of the inter- (M) tunnels.256 The commensurate TTB phase is considered
face system is preserved.247 to be more favorable when A1-sites are occupied by ions with
The relaxor properties become more pronounced in SBN small ionic radii (  1.3 Å).233 The lack of commensurate peri-
doped by trivalent cations, like rare-earth elements on the A- odicity in Ba2BiTi2Nb3O15 and Ba2LaTi2Nb3O15 may contrib-
site252,253 or by Cr3+ on the B-site254 probably reflecting ute to the disruption of long-range dipolar coupling in these
stronger RF activity related to heterovalent doping. On the compositions resulting in relaxor rather than classic ferroelec-
other hand, the substitution of either Sr2+ or Ba2+ by Na+ tric behavior.233 According to Zhu et al., the polar behavior
in Sr0.6xBa0.4Na2xNb2O6 (x  0.15) and Sr0.6Ba0.4yNa2y in this family of tungsten bronzes is dominated not only by
Nb2O6 (y  0.1), respectively, results in the suppression of the ionic radius of the rare-earth elements, but rather by the
relaxor properties and stabilization of the ferroelectric radius difference between A1 and A2 cations.258 These argu-
state.255 In this case, each Sr/Ba substitution by 2Na cations ments were supported by investigation of (BaxSr1x)2Nd-
reduces the disorder in the crystal lattice due to the reduction Ti2Nb3O15 ceramics showing crossover from the ferroelectric
of the vacancy concentration. to relaxor behavior as the radius difference between A1 (Nd)
Recently, increasing interest has reached TTB relaxors with and A2 (Ba/Sr) sites decreases.259
disorder in the B-sites. It was observed that among the family Analysis of the relaxor behavior of the tungsten bronzes
of materials with the general formula Ba2MTi2Nb3O15 family Ba6MNb9O30 (M = Ga3+, Sc3+, or In3+) reveal that
(M = Bi3+, La3+, Nd3+, Sm3+, Gd3+),233,256,257 the com- all characteristic temperatures, including the freezing temper-
pounds with large M cations, Bi3+ and La3+, exhibit relaxor ature and temperature of the maximum of the dielectric per-
properties, whereas those with smaller M = Nd3+, Sm3+, mittivity, increase with increasing radius of the M cation (see
Gd3+ behave like classical ferroelectrics. X-ray diffraction Table V).230,231 Such a trend correlates with the increased
confirmed that Ba2+ and M occupy A2 and A1-sites, respec- tetragonality of the unit cell, and hence, with an enhance-
tively, while the B-sites are randomly filled with Nb5+ and Ti4 ment of the local polarization. The latter is mainly related to
+
cations. Similar to other tungsten bronzes, the polarization the displacement of the B1-site cations (M or Nb) along the
is caused by off-centrality of the B-cations. The ferroelectric c-axis.231 Taking into account that the ratio of B1 to B2 sites
behavior of Ba2MTi2Nb3O15 with smaller M-ions indicates is 1:4 in the TTB unit cell, Rotaru et al. have argued that
that the B-site disorder per se is probably not enough to these materials have a dilute dipolar structure arising from
ensure relaxor behavior in these compounds. The electron dif- “polar” chains of B1 ions spatially separated within a weakly
fraction experiments reveal that the ferroelectric and relaxor polar matrix formed by the crystallographic distribution of
Ba2MTi2Nb3O15 have a commensurate and incommensurate the B2-sites.231 Correspondingly, the formation of the low-
20 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1

temperature domain state as in SrxBa1xNb2O6 is considered


to be restricted. Furthermore, it was suggested that the
behavior of the Ba6MNb9O30 compounds is more close to
the dipolar glasses than to the relaxor ferroelectrics.231 The
linear shape of P(E) dependences as well as a presumable
absence of PNRs seem to confirm this supposition.
Some TTB ceramics show rather complex temperature
dependences of the dielectric permittivity with two (some-
times even more) anomalies. Often the low-temperature
anomaly shows relaxor-like frequency dispersion, while the Fig. 22. Temperature dependence of the dielectric permittivity of
high-temperature one does not. The origin of these anomalies the 0.96K0.5Na0.5NbO3–0.04BiScO3 ceramic measured at a frequency
is still debated and might be different for different TTB fami- of 10 kHz. Reprinted with permission from Ref. 139. Copyright
lies. The anomalies were observed in some ceramics with 2008, American Institute of Physics.
compositions intermediate between compounds with ferro-
electric and relaxor behavior, for example, Sr4Nd2Ti4Nb6O30
(Ref. 260) or Ba2Pr0.6Nd0.4FeNb4O16 (Ref. 104) [Fig. 11(b)]. conditioning, electromagnetic interference suppression, com-
It was supposed that the high-temperature maximum corre- mutation in power electronic circuits, etc.51 Usually high-
sponds to the ferroelectric Curie temperature, while the low- energy density applications demand high permittivity, low hys-
temperature anomaly is associated with the freezing of teresis loss, and high dielectric breakdown strength. Ogihara
PNRs. While Castel et al. have proposed the coexistence of et al. have shown that relaxor BaTiO3–BiScO3 ceramics are
regions with long and short range orders,104 probably due to feasible for such applications. In particular, they reported that
nanoscale compositional segregation, Zhu et al. have sug- an energy density of about 6 J/cm3 at a field of 73 kV/cm can
gested that the long-range-ordered state is disrupted into be achieved in single layer 0.7BaTiO3–0.3BiScO3 capacitors,
polar clusters due to a transition to an incommensurate state which is superior to the characteristic of commercial capaci-
on cooling.258 The low-frequency relaxation might also be tors.51 Moreover, the high-energy density values were con-
related to a freeze-out of the polarizability in the (a,b)- firmed up to 300°C making this material an attractive
plane,261 for example, related to concerted rotations of candidate for applications at elevated temperatures.51
oxygen octahedra.262 While the ferroelectric properties are The dc electric-field tuned dielectric materials are candi-
determined by ionic displacements along the polar c-axis, dates for microwave applications, including voltage-con-
concerted rotations of oxygen octahedra, which may carry trolled oscillators, tunable filters, phase shifters, etc.265–267
weaker dipole moments, can give rise to the isotropic dielec- Usually the materials have to show a large dielectric tuna-
tric response within the (a,b)-plane.262 bilty, =100%·[e(0)e(E)]/e(0), and low dielectric loss. Among
The reports on a possible anisotropy of the relaxor prop- the most studied compounds are presently ferroelectric (Ba,
erties in the uniaxial systems are controversial. Simon et al. Sr)TiO3268 and (Pb,Sr)TiO3266. However, recently, the lead-
have reported that the Sr0.75Ba0.25Nb2O6 single crystals free relaxor systems Ba(Zr,Ti)O350,269and Ba(Sn,Ti)O3270
behave like relaxors along the [001] direction, while exhibit- have attracted, increased attention due to their excellent
ing a ferroelectric transition in the [100] direction in the same characteristics. The BTZr (x = 0.30) ceramics exhibit tunabil-
temperature range.263 On the contrary, qualitatively, the ity as high as 45% (at E = 40 kV/cm) and correspondingly
same relaxor behavior was found for the [001] and [100] low loss, tgd ~ 0.002 at room temperature.269 The tunability
directions of Sr0.61Ba0.39Nb2O6 single crystals, (V. V. is even higher at lower temperatures (up to 90% at 230 K).
Shvartsman, unpublished data) as well as for the textured For increasing Zr content, the tunability decreases, but still
Sr0.63Ba0.37Nb2O6 ceramics with the grains-oriented preferen- remains above 20% for BTZr (x = 0.45).50
tially along and normal to the polar axis, respectively.264
At the end of this section, we would like to point out that
contrary to the lead-free perovskite relaxors, where B-site (2) Electromechanical Properties
disorder is most important for inducing the relaxor behavior, While ferroelectric Ba(Ti,Zr)O3 ceramics with a moderate Zr
the A-site disorder plays the crucial role in uniaxial TTB content (~5%) exhibit excellent piezoelectric properties,
relaxors, in spite of the fact that the polar behavior in both d33 ~ 236 pC/N and field-induced strain ~0.18% (at 40 kV/
cases is attributed to B-site cations. cm),271 the relaxor BT-based compounds with larger Zr con-
tent usually show poor electromechanical response. Neverthe-
less, a progressive change of the piezoelectric coefficient with
IV. Technological Perspectives composition may be used for fabrication of bending actua-
(1) Dielectric Properties tors based on monolithic ceramics with a gradient of compo-
Large and relatively temperature-insensitive dielectric permit- sition, for example, BaTi1xSnxO3 (0.075  x  0.15).272
tivity makes relaxors obvious candidates for capacitors appli- The situation is different for bismuth sodium titanate-
cations. Of particular interest are compounds showing a based relaxors. The (1x)BNT–xBT solid solutions close to
maximum of the dielectric constant e at room and elevated the morphotropic phase boundary between the rhombohedral
temperatures. There is demand for the latter materials, for and tetragonal ferroelectric phases (x ~ 0.06–0.08) show good
example, in automotive and aerospace industry. While most piezoelectric properties, d33 ~ 150 pC/N.126 The electrome-
of BaTiO3-derived relaxors exhibit a peak of the dielectric chanical response is drastically enhanced by a partial substi-
permittivity below room temperature, systems based on tution of BNT by KNN. The 0.92Bi0.5Na0.5TiO3–0.06BaTiO3
KNN or BNT as well as a number of relaxors with the Auri- –0.02K0.5Na0.5NbO3 ceramics can deliver giant strain
villius structure are promising candidates for high-temperature ~0.45% (at E = 80 kV/cm) under both unipolar and bipolar
applications. For example, 0.96K0.5Na0.5NbO3–0.04BiScO3139 electrical loading,53 surpassing the maximum strain values
and 0.82(0.94Bi1/2Na1/2TiO3–0.06BaTiO3)–0.18K0.5Na0.5NbO3175 obtained in commercial ferroelectric Pb(Zr,Ti)O3 ceramics
relaxor ceramics were shown to have permittivity values of (Fig. 23). It was suggested that these superior properties are
over 2000 with variation of De/e less than 10% in the tempera- due to the relaxor character of the BNT–BT–KNN com-
ture range from 100°C to 300°C (Fig. 22). A large, tempera- pounds.176 The origin of the giant strain is mostly a conse-
ture stable dielectric constant combined with high electrical quence of the disappearance of the remnant strain due to the
resistivity is necessary for the development of high-energy den- presence of nonpolar regions in the relaxor phase at zero
sity capacitors, which are used in power electronic applica- electric field. These regions destabilize and randomize an
tions, such as filtering, voltage smoothing, dc blocking, power electrically induced ferroelectric order and bring the system
January 2012 Lead-Free Relaxors 21

Fig. 23. Unipolar field-induced strain in (1x)Bi0.5Na0.5TiO3–


0.06BaTiO3–xK0.5Na0.5NbO3 ceramics with x = 0, 0.02, and 0.04 in
comparison with a unipolar strain–electric-field curve of commercial
soft PZT. Reprinted with permission from Ref. 53. Copyright 2007,
American Institute of Physics.

back to its unpoled state once the applied electric field is


Fig. 24. The three families of relaxor ferroelectrics with different
removed.176
dimensionality of the order parameter (polarization) and different
anisotropy of the respective polar nanoregions. For relaxors with the
perovskite structure PNRs with rhombohedral or tetragonal
(3) Optical Properties symmetry of polarization may exist depending on the individual
Some relaxors with TTB structure, like Sr1xBaxNb2O6 show compound.
excellent electrooptic properties. Large linear electro-optic
coefficients, r33 = 1340 9 1012 m/V for composition with
x = 0.75, make them potential materials for light modulation
charge disorder, size disorder, or even at nominally the same
and laser frequency multiplication required for optical com-
effective ionic charge a difference in the existence of a lone
munication systems.227 The SBN single crystals furthermore
electron pair at an ion or not may be sufficient sources of
exhibit a substantial photorefractive effect, optically induced
disordering. Polar nanoregions generated from such disorder
change of refractive index, which can be used in holographic
are found in all relaxors. Strong anisotropy of macroscopic
memory, phase conjugation, image amplification, optical
material properties does not translate into comparable
resonators, and other nonlinear optical applications.54
anisotropy of the relaxor properties. Thus, all forms of
longer range structural ordering do not directly correlate
V. Conclusions with the polar nanostate. It is thus inherently restricted to
atomic scale processes irrespective of the overall crystal sym-
Lead-free relaxor ferroelectrics offer all property features one
metry. The large variety of lead-free relaxor systems has
can possibly think of in a ferroelectric relaxor system. Their
allowed us to arrive at such a clear statement.
basic features are summarized in Fig. 24. Here, we assume
Large and potentially “giant” material parameters have
that the inherent symmetry of the local order parameter, i.e.
been found and device technology is waiting for the final sur-
polarization, results in a corresponding shape anisotropy of
passing of the output thresholds set by the lead-containing
the polar nanoregions. Indeed, strongly anisotropic needle-
compositions. Fundamentally, there is no reason why the lat-
like nano-domains have been observed in the one-dimen-
ter should continue to dominate device technology on the
sional relaxors like (Sr,Ba)Nb2O6 with tetragonal tungsten
long run.
bronze structure.273 It was suggested that in 2D relaxors with
Aurivillius structure growth of PNRs is restricted to the
plane of the perovskite layer, but no experimental proves References
have been given so far. It is often accepted that PNRs in 3D 1
Y. Murakami, H. Kasai, J. J. Kim, S. Mamishin, D. Shindo, S. Mori, and
relaxor, for example, in PbMg1/3Nb2/3O3, are of ellipsoidal A. Tonomura, “Ferromagnetic Domain Nucleation and Growth in Colossal
shape.9,274 However, there are evidences of planar-type corre- Magnetoresistive Manganite,” Nature Nanotechnol., 5, 37–41 (2010).
2
lation of local polarization in some relaxors with the perov- R. P. Rairigh, G. Singh-Bhalla, S. Tongay, T. Dhakal, A. Biswas, and A.
skite structure.275–277 For example, results of three- F. Hebard, “Colossal Magnetocapacitance and scale-Invariant Dielectric
Response in Phase-Separated Manganites,” Nat. Phys., 3, 551–5 (2007).
dimensional mapping of diffuse X-ray scattering intensities 3
J. Lee, K. Fujita, K. McElroy, J. A. Slezak, M. Wang, Y. Aiura, H.
performed for PbZn1/3Nb2/3O3–PbTiO3 single crystals indi- Bando, M. Ishikado, T. Masui, J.-X. Zhu, A. V. Balatsky, H. Eisaki, S. Uch-
cate that PNRs have a pancake-like shape resulting from ida, and J. C. Davis, “Interplay of Electron–Lattice Interactions and Super-
{110} planar correlations of <1–10> polarizations.275 It was conductivity in Bi2Sr2CaCu2O8+8” Nature, 442, 546 (2006).
4
Z. Kutnjak, J. Petzelt, and R. Blinc, “The Giant Electromechanical
suggested that <111>-polarized PNRs could be a result of Response in Ferroelectric Relaxors as a Critical Phenomenon,” Nature, 441,
the overlap of three different {110}-type pancakes.275 Local 956–9 (2006).
5
symmetry of PNRs may be different for different relaxor sys- E. Dagotto, “Complexity in Strongly Correlated Electronic Systems,”
tems, for example, rhombohedral or tetragonal for BaTiO3 Science, 309, 257–62 (2005).
6
G. A. Smolenskii and V. A. Isupov, “Segnetoelektricheskie svoistva tver-
and Bi1/2Na1/2TiO3-based relaxors, respectively. It is interest- dykh rastvorov stannata bariya v titanate bariya,” Zh. Tekh. Fiz., 24, 1375–86
ing, however, that one, two, and three dimensional systems (1954).
7
offer surprisingly similar material behavior at least at the G. A. Smolenskii, V. A. Isupov, A. I. Agranovskaya, and S. N. Popov,
macroscopic scale. “Ferroelectrics with Diffuse Phase Transitions,” Sov. Phys. Solid State, 2,
2584–94 (1961).
The relaxor state in all systems relies on some atomic and 8
L. E. Cross, “Relaxor Ferroelectrics,” Ferroelectrics, 76, 241–67 (1987).
field disorder. For each of the systems, the structural or 9
A. A. Bokov and Z. –G. Ye, “Recent Progress in Relaxor Ferroelectrics
atomic origin may be different: site disorder, substitutional with Perovskite Structure,” J. Mater. Sci., 41, 31–52 (2006).
22 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1
10 43
A. J. Bell, “Calculations of Dielectric Properties from the Superparaelec- F. M. Jiang and S. Kojima, “Relaxation Mode in 0.65Pb(Mg1/3Nb2/3)O3-
tric Model of Relaxors,” J. Phys. Condens. Matter, 5, 8773–92 (1993). 0.35PbTiO3 Relaxor Single Crystals Studied by Micro-Brillouin Scattering,”
11
S. -E. Park and T. R. Shrout, “Ultrahigh Strain and Piezoelectric Behavior Phys. Rev. B, 62, 8572–5 (2000).
44
in Relaxor Based Ferroelectric Single Crystals,” J. Appl. Phys., 82, 1804–11 V. V. Shvartsman and A. L. Kholkin, “Domain Structure of 0.8PbMg1/3
(1997). Nb2/3O3-0.2PbTiO3 Studied by Piezoresponse Force Microscopy,” Phys. Rev.
12
K. Uchino, Piezoelectric Actuators and Ultrasonic Motors. Kluwer B, 69, 014102 (2004).
45
Academic, Boston, 1996. G. Xu, H. Hiraka, G. Shirane, and K. Ohwada, “Dual Structures in (1-x)
13
G. A. Samara, “The Relaxational Properties of Compositionally Disor- Pb(Zn1/3Nb2/3)O3-xPbTiO3 Ferroelectric Relaxors,” Appl. Phys. Lett., 84, 3975
dered ABO3 Perovskites,” J. Phys.: Condens. Matter, 15, R367–411 (2003). –7 (2004).
14 46
R. A. Cowley, S. N. Gvasaliya, S. G. Lushnikov, B. Roessli, and G. M. G. Xu, D. Viehland, J. F. Li, P. M. Gehring, and G. Shirane, “Evidence
Rotaru, “Relaxing with Relaxors: A Review of Relaxor Ferroelectrics,” Adv. of Decoupled Lattice Distortion and Ferroelectric Polarization in the Relaxor
Mater., 60, 29–327 (2011). System PMN-xPT,” Phys. Rev. B, 68, 212410 (2003).
15 47
J. Ravez and A. Simon, “Some Solid State Chemistry Aspects of Lead- V. V. Kirillov and V. A. Isupov, “Relaxation Polarization of PbMg1/3Nb2/3
Free Relaxor Ferroelectrics,” J. Sol. State Chem., 162, 260–5 (2001). O3 (PMN) – a Ferroelectric with a Diffused Phase Transition,” Ferroelectrics,
16
J. Ravez and A. Simon, “Lead-Free Relaxor Ferroelectrics with “TTB” 5, 3–9 (1973).
48
Structure,” C. R. Chimie, 5, 143–8 (2002). R. Blinc and R. Pirc, “Spherical Random-Bond–Random-Field Model of
17
G. Burns and F. H. Dacol, “Glassy Polarization Behavior in Ferroelectric Relaxor Ferroelectrics,” Phys. Rev. B, 60, 13470–8 (1999).
49
Compounds Pb(Mg1/3Nb2/3)O3 and Pb(Zn1/3Nb2/3)O3,” Solid State Commun., EU-Directive 2002/95/EC, “Restriction of the Use of Certain Hazardous
48, 853–6 (1983). Substances in Electrical and Electronic Equipment (RoHS),” Off. J. Eur.
18
V. Westphal, W. Kleemann, and M. Glinchuk, “Diffuse Phase Transitions Union, 46 [L37] 19–23 (2003).
50
and Random-Field-Induced Domain States of the “relaxor” Ferroelectric T. Maiti, R. Guo, and A. S. Bhalla, “Enhanced Electric Field Tunable
PbMg1/3Nb2/3O3,” Phys. Rev. Lett., 68, 847–50 (1992). Dielectric Properties of BaZrxTi1-xO3 Relaxor Ferroelectrics,” Appl. Phys.
19
M. E. Lines and A. M. Glass, Principles and Applications of Ferroelectrics Lett., 90, 182901 (2007).
51
and Related Materials. Clarendon Press, Oxford, 1979. H. Ogihara, C. A. Randall, and S. Trolier-McKinstry, “High-Energy
20
S. B. Vakhrushev, B. E. Kvyatkovsky, A. A. Naberezhnov, N. M. Okuneva, Density Capacitors Utilizing 0.7BaTiO3–0.3BiScO3 Ceramics,” J. Am. Ceram.
and B. P. Toperverg, “Glassy Phenomena in Disordered Perovskite-Like Single Soc., 92, 1719–24 (2009).
Crystals,” Ferroelectrics, 90, 173–6 (1989). 52
J. R. Gomah-Pettry, S. Said, P. Marchet, and J. –P. Mercurio, “Sodium-
21
A. Naberezhnov, S. Vakhrushev, B. Dorner, D. Strauch, and H. Moud- Bismuth Titanate Based Lead-Free Ferroelectric Materials,” J. Eur. Ceram.
den, “Inelastic Neutron Scattering Study of the Relaxor Ferroelectric PbMg1/3 Soc., 24, 1165–9 (2004).
Nb2/3O3 at High Temperatures,” Eur. Phys. J. B, 11, 13–20 (1999). 53
S. –T. Zhang, A. B. Kounga, E. Aulbach, H. Ehrenberg, and J. Rödel,
22
K. Hirota, S. Wakimoto, and D. E. Cox, “Neutron and X-Ray Scattering “Giant Strain in Lead-Free Piezoceramics Bi0.5Na0.5TiO3-BaTiO3-
Studies of Relaxors,” J. Phys. Soc. Jpn., 75, 111006 (2006). K0.5Na0.5NbO3,” Appl. Phys. Lett., 91, 112906 (2007).
23 54
G. Xu, G. Shirane, J. R. D. Copley, and P. M. Gehring, “Neutron Elastic M. D. Ewbank, R. R. Neurgaonkar, W. K. Gory, and J. Feinberg,
Diffuse Scattering Study of PbMg1/3Nb2/3O3,” Phys. Rev. B, 69, 064112 (2004). “Photorefractive Properties of Strontium-Barium Niobate,” J. Appl. Phys., 62,
24
D. La-Orauttapong, J. Toulouse, J. L. Robertson, and Z.-G. Ye, “Diffuse 374–80 (1987).
Neutron Scattering Study of a Disordered Complex Perovskite PbZn1/3 55
B. Wul, “Barium Titanate – A New Ferroelectric,” Nature, 157, 808 (1946).
56
Nb2/3O3 Crystal,” Phys. Rev. B, 64, 212101 (2001). A. von Hippel, “Ferroelectricity, Domain Structure, and Phase Transitions
25
C. Stock, R. J. Birgeneau, S. Wakimoto, J. S. Gardner, W. Chen, Z.-G. of Barium Titanate,” Rev. Mod. Phys., 22, 221–37 (1950).
57
Ye, and G. Shirane, “Universal Static and Dynamic Properties of the Struc- N. Abdelmoula, H. Chaabane, H. Khemakem, R. von der Muhll, and A.
tural Transition in PbZn1/3Nb2/3O3,” Phys. Rev. B, 69, 094104 (2004). Simon, “Relaxor or Classical Ferroelectric Behavior in A-Site Substituted
26
M. Yoshida, S. Mori, N. Yamamoto, Y. Uesu, and J. M. Kiat, “TEM Perovskite Type Ba1-x(Sm0.5Na0.5) xTiO3,” Solid State Sci., 8, 880–7 (2006).
58
Observation of Polar Domains in Relaxor Ferroelectric Pb(Mg1/3Nb2/3)O3,” R. Farhi, M. El Marssi, A. Simon, and J. Ravez, “Relaxor-Like and Spec-
Ferroelectrics, 217, 327–33 (1998). troscopic Properties of Niobium Modified Barium Titanate,” Eur. Phys. J. B,
27
P. K. Davies and M. A. Akbas, “Chemical Order in PMN-Related Relax- 18, 605–10 (1999).
59
ors: Structure, Stability, Modification, and Impact on Properties,” J. Phys. H. Ogihara, C. A. Randall, and S. Trolier-McKinstry, “Weakly Coupled
Chem. Solids, 61, 159 (2000). Relaxor Behavior of BaTiO3–BiScO3 Ceramics,” J. Am. Ceram. Soc., 92, 110–
28
W. Kleemann, “Random-Field Induced Antiferromagnetic, Ferroelectric 8 (2009).
60
and Structural Domain States,” Int. J. Mod. Phys. B, 7, 2469–507 (1993). S. Tinberg and S. Troiler-McKinstry, “Structural and Electrical Character-
29
W. Kleemann, “The Relaxor Enigma - Charge Disorder and Random ization of xBiScO3 -(1-x)BaTiO3 Thin Films,” J. Appl. Phys., 101, 024112
Fields in Ferroelectrics,” J. Mater. Sci., 41, 129–36 (2006). (2007).
30 61
I. Imry and S. K. Ma, “Random Field Instability of the Ordered State of S. S. N. Bharadwaja, J. R. Kim, H. Ogihara, L. E. Cross, S. Trolier-
Continious Symmetry,” Phys. Rev. Lett., 35, 1399–401 (1975). McKinstry, and C. A. Randall, “Critical Slowing Down Mechanism and
31
D. Viehland, S. J. Jang, L. E. Cross, and M. Wuttig, “Freezing of the Reentrant Dipole Glass Phenomena in (1-x)BaTiO3-xBiScO3 (0.1 
Polarization Fluctuations in Lead Magnesium Niobate Relaxors,” J. Appl. x  0.4): The High Energy Density Dielectrics,” Phys. Rev. B, 83, 024106
Phys., 68, 2916–21 (1990). (2011).
32 62
A. E. Glazounov and A. K. Tagantsev, “Direct Evidence for Vögel–Ful- A. N. Salak, M. P. Seabra, and V. M. Ferreira, “Relaxor Behavior of the
cher Freezing in Relaxor Ferroelectrics,” Appl. Phys. Lett., 73, 856–8 (1998). 0.9BaTiO3–0.1La(Mg1/2Ti1/2)O3 Solid Solution,” J. Am. Ceram. Soc., 87, 216–
33
A. Levstik, Z. Kutnjak, C. Filipic, and R. Pirc, “Glassy Freezing in Relax- 20 (2004).
63
or Ferroelectric Lead Magnesium Niobate,” Phys. Rev. B, 57, 11204–11 (1998). A. N. Salak, M. P. Seabra, and V. M. Ferreira, “Evolution from
34
E. V. Colla, Yu. E. Koroleva, N. M. Okuneva, and S. B. Vakhrushev, Ferroelectric to Relaxor Behavior in the (1  x)BaTiO3  xLa(Mg1/2Ti1/2)O3
“Long-Time Relaxation of the Dielectric Response in Lead Magnoniobate,” System,” Ferroelectrics, 318, 185–92 (2005).
64
Phys. Rev. Lett., 74, 1681–4 (1995). T. Maiti, R. Guo, and A. S. Bhalla, “Structure-Property Phase Diagram
35
E. V. Colla, L. K. Chao, M. B. Weissman, and D. Viehland, “Aging of BaZrxTi1-xO3 System,” J. Am. Ceram. Soc., 91, 1769–80 (2008).
65
in a Relaxor Ferroelctric: Scaling and Memory Effects,” Phys. Rev. Lett., 85, I. A. Santos and J. A. Eiras, “Phenomenological Description of the
3033–6 (2000). Diffuse Phase Transition in Ferroelectrics,” J. Phys.: Condens. Matter, 13,
36
Z. Kutnjak, C. Filipic, R. Pirc, A. Levstik, R. Farhi, and M. El Marssi, 11733–40 (2001).
66
“Slow Dynamics and Ergodicity Breaking in a Lanthanum-Modified Lead K. Uchino and S. Nomura, “Critical Exponents of the Dielectric Con-
Zirconate Titanate Relaxor System,” Phys. Rev. B, 59, 294–301 (1999). stants in Diffused-Phase-Transition Crystals,” Ferroelectric Lett., 44, 55–61
37
Z. -G. Ye and H. Schmid, “Optical, Dielectric and Polarization Studies of (1982).
67
the Electric Field-Induced Phase Transition in Pb(Mg1/3Nb2/3)O3 [PMN],” J. Ravez and A. Simon, “Lead-Free Ferroelectric Relaxor Ceramics
Ferroelectrics, 145, 83–108 (1993). Derived from BaTiO3,” Eur. Phys. J. AP, 11, 9–13 (2000).
38 68
T. Kim, J. N. Hanson, A. Gruverman, A. I. Kingon, and S. K. Streiffer, H. Khemakhem, A. Simon, R. von der Mühll, and J. Ravez, “Relaxor
“Ferroelectric Behavior in Nominally Relaxor Lead Lanthanum Zirconate or Classical Ferroelectric Behavior in Ceramics with Composition Ba1-x
Titanate Thin Films Prepared by Chemical Solution Deposition on Copper NaxTi1-xNbxO3,” J. Phys.: Condens. Matter, 12, 5951–9 (2000).
69
Foil,” Appl. Phys. Lett., 88, 262907 (2006). J. Ravez and A. Simon, “Le premier relaxeur ferroélectrique oxyfluoré,”
39
F. Chu, I. M. Reaney, and N. Setter, “Spontaneous (Zero-Field) Relaxor- Phys. Status Solidi A, 159, 517–22 (1997).
70
to-Ferroelectric-Phase Transition in Disordered Pb(Sc1/2Nb1/2)O3,” J. Appl. F. Bahri, A. Simon, H. Khemakem, and J. Ravez, “Classical or Relaxor
Phys., 77, 1671–6 (1995). Ferroelectric Behavior of Ceramics with Composition Ba1-xBi2x/3TiO3,” Phys.
40
O. Noblanc, P. Gaucher, and G. Calvarin, “Structural and Dielectric Status Solidi A, 184, 459–64 (2001).
71
Studies of Pb(Mg1/3Nb2/3)O3-PbTiO3 Ferroelectric Solid Solutions Around the J. Ravez and A. Simon, “New Lead-Free Relaxor Ceramics Derived from
Morphotropic Boundary,” J. Appl. Phys., 79, 4291–7 (1996). BaTiO3 by Cationic Heterovalent Substitutions in the 12 C.N. Crystallo-
41
A. M. Glass, “Investigation of the Electrical Properties of Sr1-xBaxNb2O6 graphic Site,” Solid State Sci., 2, 525–9 (2000).
72
with Special Reference to Pyroelectric Detection,” J. Appl. Phys., 40, 4699–713 A. Simon, J. Ravez, and M. Maglione, “The Crossover from a Ferroelec-
(1969). tric to a Relaxor State in Lead-Free Solid Solutions,” J. Phys.: Condens.
42
B. Dkhil, J. M. Kiat, G. Calvarin, G. Baldinozzi, S. B. Vakhrushev, and Matter, 16, 963–70 (2004).
73
E. Suard, “Local and Long Range Polar Order in the Relaxor-Ferroelectric A. A. Bokov, M. Maglione, and Z. G. Ye, “Quasi-Ferroelectric State in
Compounds PbMg1/3Nb2/3O3 and PbMg0.3Nb0.6Ti0.1O3,” Phys. Rev. B, 65, Ba(Ti1-xZrx)O3 Relaxor: Dielectric Spectroscopy Evidence,” J. Phys.: Condens.
024104 (2002). Matter, 19, 092001 (2007).
January 2012 Lead-Free Relaxors 23
74 104
V. V. Shvartsman, J. Zhai, and W. Kleemann, “The Dielectric Relaxation E. Castel, M. Josse, D. Michau, and M. Maglione, “Flexible Relaxor
in Solid Solutions BaTi1-xZrxO3,” Ferroelectrics, 379, 77–85 (2009). Materials: Ba2PrxNd1xFeNb4O15 Tetragonal Tungsten Bronze Solid Solu-
75
V. Mueller, H. Beige, and H. -P. Abicht, “Non-Debye Dielectric Disper- tion,” J. Phys.: Condens. Matter, 21, 452201 (2009).
105
sion of Barium Titanate Stannate in the Relaxor and Diffuse Phase-Transition R. Farhi, M. El Marssi, A. Simon, and J. Ravez, “A Raman and Dielec-
State,” Appl. Phys. Lett., 84, 1341–3 (2004). tric Study of Ferroelectric Ba(Ti1-xZrx)O3 Ceramics,” Eur. Phys. J. B, 9, 599–
76
V. V. Shvartsman, J. Dec, Z. K. Xu, J. Banys, P. Keburis, and W. Klee- 604 (1999).
106
mann, “Crossover from Ferroelectric to Relaxor Behavior in BaTi1-xSnxO3 N. K. Karan, R. S. Katiyar, T. Maiti, R. Guo, and A. S. Bhalla, “Raman
Solid Solutions,” Phase Transitions, 81, 1013–21 (2008). Spectra of Zr4+-rich BaZrxTi1-xO3 (0.5  x  1.00) Phase Diagram,”
77
C. Lei, A. A. Bokov, and Z. -G. Ye, “Ferroelectric to Relaxor Crossover J. Raman Spectrosc., 40, 370–5 (2009).
107
and the Dielectric Phase Diagram in BaTiO3-BaSnO3 System,” J. Appl. Phys., M. P. Fontana and M. Lambert, “Linear Disorder and Temperature
101, 084105 (2007). Dependence of Raman Scattering in BaTiO3,” Solid State Commun., 10, 1–4
78
S. G. Lu, Z. K. Xu, and H. Chen, “Tunability and Relaxor Properties of (1972).
108
Ferroelectric Barium Stannate Titanate Ceramics,” Appl. Phys. Lett., 85, 5319 G. A. Barbosa, A. Chaves, and S. P. S. Porto, “Temperature Dependence
–21 (2004). of the Raman Cross Sections in BaTiO3 and SrTiO3,” Solid State Commun.,
79
D. Hennings, A. Schnell, and G. Simon, “Diffuse Ferroelectric Phase 11, 1053–5 (1972).
109
Transitions in Ba(Ti1-yZry)O3 Ceramics,” J. Am. Ceram. Soc., 65, 539–44 S. Miao, J. Pokorny, U. M. Pasha, O. P. Thakur, D. C. Sinclair, and
(1982). I. M. Reaney, “Polar Order and Diffuse Scatter in Ba(Ti1-xZrx)O3 Ceramics,”
80
L. Geske, H. Beige, H. –P. Abicht, and V. Mueller, “Electromechanical J. Appl. Phys., 106, 114111 (2009).
110
Resonance Study of the Diffuse Ferroelectric Phase Transition in BaTi1-x C. Laulhé, F. Hippert, R. Bellissent, A. Simon, and G. J. Cuello, “Local
SnxO3 Ceramics,” Ferroelectrics, 314, 97–104 (2005). Structure in BaTi1-xZrxO3 Relaxors from Neutron Pair Distribution Function
81
X. Y. Wei, Y. Feng, and X. Yao, “Dielectric Relaxation Behavior in Bar- Analysis,” Phys. Rev. B, 79, 064104 (2009).
111
ium Stannate Titanate Ferroelectric Ceramics with Diffused Phase Transition,” C. Laulhé, A. Pasturel, F. Hippert, and J. Kreisel, “Random Local Strain
Appl. Phys. Lett., 83, 2031–3 (2003). Effects in Homovalent-Substituted Relaxor Ferroelectrics: A First-Principles
82
K. -Y. Oh, K. Uchino, and L. E. Cross, “Optical Study of Domains in Ba Study of BaTi0.74Zr0.26O3,” Phys. Rev. B, 82, 132102 (2010).
112
(Ti,Sn)O3 Ceramics,” J. Am. Ceram. Soc., 77, 2809–16 (1994). I.-K. Jeong, C. Y. Park, J. S. Ahn, S. Park, and D. J. Kim, “Ferroelec-
83
V. V. Shvartsman, W. Kleemann, J. Dec, Z. K. Xu, and S. G. Lu, “Dif- tric-Relaxor Crossover in Ba(Ti1-xZrx)O3 Studied Using Neutron Total Scat-
fuse Phase Transition in BaTi1-xSnxO3 Ceramics: An Intermediate State tering Measurements and Reverse Monte Carlo Modelling,” Phys. Rev. B, 81,
Between Ferroelectric and Relaxor Behavior,” J. Appl. Phys., 99, 124111 214119 (2010).
113
(2006). Y. Liu, R. L. Whiters, X. Y. Wei, and J. D. F. Gerald, “Structure
84
A. N. Salak, V. V. Shvartsman, M. P. Seabra, A. L. Kholkin, and V. Diffused Scattering and Polar Nano-Regions in the Ba(Ti1-xSnx)O3 Relaxor
M. Ferreira, “Ferroelectric-To-Relaxor Transition Behavior of BaTiO3 Ferroelectric System,” J. Sol. St. Chem., 180, 858–65 (2007).
114
Ceramics Doped with La(Mg1/2Ti1/2)O3,” J. Phys. Condens. Matter, 16, Y. Liu, R. L. Whiters, B. Nguyen, and K. Elliot, “Structurally Frustrated
2785–94 (2004). Polar Nanoregions in BaTiO3-Based Relaxor Ferroelectric Systems,” Appl.
85
S. Kamba, M. Kempa, V. Bovtun, J. Petzelt, K. Brinkman, and N. Setter, Phys. Lett., 91, 152907 (2007).
115
“Soft and Central Mode Behavior in PbMg1/3Nb2/3O3 Relaxor Ferroelectric,” R. Gomes, M. Lambert, and A. Guinier, “The Chain Structure of
J. Phys.: Condens. Matter, 17, 3965–74 (2005). BaTiO3 and KNbO3,” Solid State Commun., 6, 715–9 (1968).
86 116
A. K. Tagantsev, “Vogel-Fulcher Relationship for the Dielectric Permittiv- J. Harada, M. Watanabe, S. Kodera, and G. Honjo, “Diffuse Streak Dif-
ity of Relaxor Ferroelectrics,” Phys. Rev. Lett., 72, 1100–3 (1994). fraction Pattern of Electron and X-Rays Due to Low Frequency Optical Mode
87
T. Maiti, R. Guo, and A. S. Bhalla, “The Evolution of Relaxor Behavior in Tetragonal BaTiO3,” J. Phys. Soc. Jpn., 20, 630–1 (1965).
in Ti4+ Doped BaZrO3 Ceramics,” J. Appl. Phys., 100, 114109 (2006). 117
A. S. Chaves, F. C. S. Barreto, R. A. Nogueira, and B. Zeks, “Thermo-
88
A. Kumar, I. Rivera, and R. S. Katiyar, “Investigation of Local Structure dynamic of an Eight-Sight Order-Disorder Model for Ferroelectrics,” Phys.
of Lead-Free Relaxor Ba(Ti0.70Sn0.30)O3 by Raman Spectroscopy,” J. Raman Rev. B, 13, 207–12 (1976).
118
Spectrosc., 40, 459–62 (2009). K. A. Müller and W. Berlinger, “Microscopic Probing of Order-Disorder
89
M. Nagasawa, H. Kawaji, T. Tojo, and T. Atake, “Absence of The Heat Versus Displacive Behavior in BaTiO3 by Fe3+ EPR,” Phys. Rev. B, 34, 6130–
Capacity Anomaly in the Pb-Free Relaxor BaTi0.65Zr0.35O3,” Phys. Rev. B, 74, 6 (1986).
119
132101 (2006). B. Zalar, A. Lebar, J. Seliger, R. Blins, V. V. Laguta, and M. Itoh, “NMR
90
J. Kreisel, P. Bouvier, M. Maglione, B. Dkhil, and A. Simon, “High-Pres- Studies of Disorder in BaTiO3 and SrTiO3,” Phys. Rev. B, 71, 064107 (2005).
120
sure Raman Investigation of the Pb-Free Relaxor BaTi0.65Zr0.35O3,” Phys. J. Hlinka, T. Ostapchuk, D. Nuzhnyy, J. Petzelt, P. Kuzel, C. Kadlec,
Rev. B, 69, 092104 (2004). P. Vanek, I. Ponomareva, and L. Bellaiche, “Coexistence of the Phonon
91
Ph. Sciau, G. Calvarin, and J. Ravez, “X-ray Diffraction Study of and Relaxation Soft Modes in the Terahertz Dielectric Response of Tetragonal
BaTi0.65Zr0.35O3 and Ba0.92Ca0.08Ti0.75Zr0.25O3 Compositions: Influence of BaTiO3,” Phys. Rev. Lett., 101, 167402 (2008).
121
Electric Field,” Solid State Commun., 113, 77–82 (2000). H. Ko, S. Kojima, T.-Y. Koo, J. H. Jung, C. J. Won, and N. J. Hur,
92
Y. Moriya, H. Kawaji, T. Tojo, and T. Atake, “Specific-Heat Anomaly “Elastic Softening and Central Peaks in BaTiO3 Single Crystals Above the
Caused by Ferroelectric Nanoregions in Pb(Mg1/3Nb2/3)O3 and Pb(Mg1/3Ta2/3) Cubic-Tetragonal Phase-Transition Temperature,” Appl. Phys. Lett., 93,
O3 Relaxors,” Phys. Rev. Lett., 90, 205901 (2003). 102905 (2008).
93 122
M. V. Gorev, V. S. Bondarev, I. N. Flerov, Ph. Sciau, and J. -M. Savariault, M. Maglione, R. Böhmer, A. Loidl, and U. T. Höchli, “Polar Relaxation
“Heat Capacity Study of Double Perovskite-Like Compounds BaTi1-xZrxO3,” Mode in Pure and Iron-Doped Barium Titanate,” Phys. Rev. B, 40, 11441–4 (1989).
123
Phys. Solid St., 47, 2304–8 (2005). R. Yan, Z. Guo, R. Tai, H. Xu, X. Zhao, D. Lin, X. Li, and H. Luo,
94
N. de Mathan, E. Husson, G. Calvarin, and A. Morell, “Structural Study “Observation of Long Range Correlation Dynamics in BaTiO3 Near TC by
of a Poled PbMg1/3Nb2/3O3 Ceramic at Low Temperature,” Mater. Res. Bull., Photon Correlation Spectroscopy,” Appl. Phys. Lett., 93, 192908 (2008).
124
26, 1197–1172 (1991). G. Burns and F. H. Dacol, “Polarization in the Cubic Phase of BaTiO3,”
95
C. Laulhé, F. Hippert, J. Kreisel, M. Maglione, A. Simon, J. L. Haze- Solid State Commun., 42, 9–12 (1982).
125
mann, and V. Nassif, “EXAFS Study of Lead-Free Relaxor Ferroelectric E. Dul’kin, J. Petzelt, S. Kamba, E. Mojaev, and M. Roth, “Relaxor-like
BaTi1-xZrxO3 at the Zr K Edge,” Phys. Rev. B, 74, 014106 (2006). Behavior of BaTiO3 Crystals from Acoustic Emission Study,” Appl. Phys.
96
A. A. Belik, S. Iikubo, K. Kodama, N. Igawa, S. Shamoto, M. Maie, T. Lett., 97, 032903 (2010).
126
Nagai, Y. Matsui, S. Y. Stefanovich, B. I. Lazoryak, and E. Takayama-Muro- J. Rödel, W. Jo, K. T. P. Seifert, E.-M. Anton, T. Granzow, and D.
machi, “BiScO3: Centrosymmetric BiMnO3-Type Oxide,” J. Am. Chem. Soc., Damjanovic, “Perspective on the Development of Lead-Free Piezoceramics,”
128, 706–7 (2006). J. Am. Ceram. Soc., 92, 1153–77 (2009).
97 127
G. Shirane, and K. Suzuki, “On The Phase Transition in Barium-Lead M. D. Maeder, D. Damjanovic, and N. Setter, “Lead Free Piezoelectric
Titanate,” J. Phys. Soc. Jpn., 6, 274–8 (1951). Materials,” J. Electroceram., 13, 385–92 (2004).
98 128
K. Datta, P. A. Thomas, and K. Roleder, “Anomalous Phase Transi- T. Takenaka, H. Nagata, and Y. Hiruma, “Current Developments
tions of Lead-Free Piezoelectric xNa0.5Bi0.5TiO3 - (1-x)BaTiO3 Solid Solutions and Prospective of Lead-Free Piezoelectric Ceramics,” Jpn. J. Appl. Phys., 47,
with Enhanced Phase Transition Temperatures,” Phys. Rev. B, 82, 224105 3787–801 (2008).
129
(2010). M. Ahtee and A. M. Glazer, “Lattice Parameters and Tilted Octahedra
99
R. Seshadri and N. A. Hill, “Visualizing the Role of Bi 6s “lone pairs” in Sodium-Potassium Niobate Solid Solutions,” Acta Crystallogr., Sect. A, 32,
in the Off-Center Distortion in Ferromagnetic BiMnO3,” Chem. Mater., 13, 434–46 (1976).
130
2892–9 (2001). L. Egerton and D. M. Dillon, “Piezoelectric and Dielectric Properties of
100
R. J. Bratton and T. Y. Tien, “Phase Transitions in the System BaTiO3- Ceramics in the System Potassium Sodium Niobate,” J. Am. Ceram. Soc., 42,
KNbO3,” J. Am. Cream. Soc., 50, 90–3 (1967). 438–42 (1959).
101 131
I. P. Raevskii, L. M. Proskuryakova, L. A. Reznichenko, E. K. Y. Saito, H. Takao, I. Tani, T. Nonoyama, K. Takatori, T. Homma, T.
Zvorykina, and L. A. Shilkina, “Obtaining Solid Solutions in the NaNbO3- Nagaya, and M. Nakamura, “Lead-Free Piezoceramics,” Nature, 432, 84–7 (2004).
132
BaTiO3 System and Investigation of its Properties,” Rus. Phys. J., 21, 259– K. Wang and J. F. Li, “Domain Engineering of Lead-Free Li-Modified
61 (1978). (K,Na)NbO3 Polycrystals with Highly Enhanced Piezoelectricity,” Adv. Funct.
102
H. Y. Guo, C. Lei, and Z. –G. Ye, “Re-Entrant Type Relaxor Behav- Mater., 20, 1924–9 (2010).
133
ior in (1-x)BaTiO3-xBiScO3 Solid Solution,” Appl. Phys. Lett., 92, 172901 H. Birol, D. Damjanovic, and N. Setter, “Preparation and Characteriza-
(2008). tion of (K0.5Na0.5)NbO3 Ceramics,” J. Eur. Ceram. Soc., 26, 861–6 (2006).
103
C. Lei, and Z. –G. Ye, “Re-Entrant-Like Relaxor Behavior in the New 134
Y. Guo, K. Kakimoto, and H. Ohsato, “Structure and Electrical Proper-
0.99BaTiO3–0.01AgNbO3 Solid Solution,” J. Phys.: Condenc. Matter, 20, ties of Lead-Free (Na0.5K0.5)NbO3-BaTiO3 Ceramics,” Jpn. J. Appl. Phys., 43,
232201 (2008). 6662–6 (2004).
24 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1
135 164
Y. Guo, K. Kakimoto, and H. Ohsato, “Dielectric and Piezoelectric Y. Hiruma, Y. Imai, Y. Watanabe, H. Nagata, and T. Takenaka, “Large
Properties of Lead-Free (Na0.5K0.5)NbO3–SrTiO3 Ceramics,” Solid State Electrostrain Near The Phase Transition Temperature of (Bi0.5Na0.5)
Commun., 129, 279–84 (2004). TiO3–SrTiO3 Ferroelectric Ceramics,” Appl. Phys. Lett., 92, 262904 (2008).
136 165
V. Bobnar, J. Bernard, and M. Kosec, “Relaxorlike Dielectric Properties S. I. Raevskaya, L. A. Reznitchenko, I. P. Raevski, V. V. Titov, S. V.
and History-Dependent Effects in the Lead-Free K0.5Na0.5NbO3–SrTiO3 Cera- Titova, and J.- L. Dellis, “Lead Free Niobate Ceramics with Relaxor-Like
mic System,” Appl. Phys. Lett., 85, 994–6 (2004). Properties,” Ferroelectrics, 340, 107–12 (2006).
137 166
M. Kosec, V. Bobnar, M. Hrovat, J. Bernard, B. Malic, and J. Holc, V. Dorcet, P. Marchet, O. Peña, and G. Trolliard, “Properties of the
“New Lead-Free Relaxors Based on the K0.5Na0.5NbO3–SrTiO3 Solid Solu- Solid Solutions (1-x)Na0.5Bi0.5TiO3-xBiFeO3,” J. Magn. Magn. Mater., 321,
tion,” J. Mat Res., 19, 1849–54 (2004). 1762–6 (2009).
138 167
V. Bobnar, J. Holc, M. Hrovat, and M. Kosec, “Relaxorlike Dielectric D. Lin, C. Xu, Q. Zheng, Y. Wei, and D. Gao, “Piezoelectric and Dielec-
Dynamics in the Lead-Free K0.5Na0.5NbO3-SrZrO3 Ceramic System,” J. Appl. tric Properties of Bi0.5Na0.5TiO3–Bi0.5Li0.5TiO3 Lead-Free Ceramics,”
Phys., 101, 074103 (2007). J. Mater. Sci. - Mater. Electron., 20, 393–7 (2009).
139 168
H. Du, W. Zhou, F. Luo, D. Zhu, S. Qu, Y. Li, and Z. Pei, “High Y. Hiruma, H. Nagata, and T. Takenaka, “Formation of Morphotropic
Tm Lead-Free Relaxor Ferroelectrics with Broad Temperature Usage Phase Boundary and Electrical Properties of (Bi1/2Na1/2)TiO3–Ba(Al1/2Nb1/2)
Range: 0.04BiScO30.96(K0.5Na0.5)NbO3,” J. Appl. Phys., 104, 044104 O3 Solid Solution Ceramics,” J. Jpn. Appl. Phys., 48, 09KC08 (2009).
(2008). 169
S. Said and J. –P. Mercurio, “Relaxor Behavior of Low Lead and Lead
140
H. Du, W. Zhou, F. Luo, D. Zhu, S. Qu, and Z. Pei, “Phase Structure, Free Ferroelectric Ceramics of the Na0.5Bi0.5TiO3–PbTiO3 and Na0.5Bi0.5TiO3–
Dielectric Properties, and Relaxor Behavior of (K0.5Na0.5)NbO3–(Ba0.5Sr0.5) K0.5Bi0.5TiO3 Systems,” J. Eur. Ceram. Soc., 21, 1333–6 (2001).
170
TiO3 Lead-Free Solid Solution for High Temperature Applications,” J. Appl. Y. Hiruma, H. Nagata, and T. Takenaka, “Phase Diagrams and Electri-
Phys., 105, 124104 (2009). cal Properties of (Bi1/2Na1/2)TiO3 - Based Solid Solutions,” J. Appl. Phys.,
141
Z. Kutnjak, C. Filipic, A. Levstik, and R. Pirc, “Glassy Dynamics of 104, 124106 (2008).
171
Rb0.40(ND4)0.60D2PO4,” Phys. Rev. Lett., 70, 4015–8 (1993). Y. Li, W. Chen, Q. Xu, J. Zhou, X. Gu, and S. Fang, “Electromechanical
142
T. Takenaka, K. Maruyama, and K. Sakata, “(Bi1/2Na1/2)TiO3-BaTiO3 and Dielectric Properties of Na0.5Bi0.5TiO3–K0.5Bi0.5TiO3–BaTiO3 Lead-Free
System for Lead-Free Piezoelectric Ceramics,” Jpn. J. Appl. Phys., 30, 2236–9 Ceramics,” Mat Chem. Phys., 94, 328–32 (2005).
172
(1991). Y. Wu, H. Zhang, Y. Zhnag, J. Ma, and D. Xie, “Lead-Free Piezoelectric
143
C. Xu, D. Lin, and K. W. Kwok, “Structure, Electrical Properties and Ceramics with Composition of (0.97x)Na1/2Bi1/2TiO3-0.03NaNbO3-x
Depolarization Temperature of (Bi0.5Na0.5)TiO3-BaTiO3 Lead-Free Piezoelec- BaTiO3,” J. Mater. Sci., 38, 987–94 (2003).
tric Ceramics,” Solid State Sci., 10, 934–40 (2008). 173
Y. -M. Li, W. Chen, Q. Xu, J. Zhou, H. –J. Sun, and M. –S. Liao,
144
C. Ma, X. Tan, E. Dul’kin, and M. Roth, “Domain Structure-Dielectric “Dielectric and Piezoelectric Properties of Na0.5Bi0.5TiO3-K0.5Bi0.5TiO3-NaN-
Property Relationship in Lead-Free (1-x)(Bi1/2Na1/2)TiO3-xBaTiO3 Ceramics,” bO3 Lead-Free Ceramics,” J. Electroceram., 14, 53–8 (2005).
174
J. Appl. Phys., 108, 104105 (2010). C. Zhou, X. Liu, W. Li, C. Yuan, and G. Chen, “Structure and Electrical
145
F. Cordero, F. Craciun, F. Trequattrini, E. Mercadelli, and C. Galassi, Properties of Bi0.5Na0.5TiO3-Bi0.5K0.5TiO3-BiCoO3 Lead-Free Piezoelectric
“Phase Transitions and Phase Diagram of the Ferroelectric Perovskite Ceramics,” J. Mater. Sci., 44, 3833–40 (2009).
175
(Na0.5Bi0.5)1-xBaxTiO3 by Anelastic and Dielectric Measurements,” Phys. Rev. R. Dittmer, W. Jo, D. Damjanovic, and J. Rödel, “Lead-Free High-Tem-
B, 81, 144124 (2010). perature Dielectrics with Wide Operational Range,” J. Appl. Phys., 109,
146
V. Dorcet, G. Trolliard, and P. Boullay, “Reinvestigation of Phase Tran- 034107 (2011).
176
sitions in Na0.5Bi0.5TiO3 by TEM. Part I: First Order Rhombohedral to W. Jo, T. Granzow, E. Aulbach, J. Rödel, and D. Damjanovic, “Origin
Orthorhombic Phase Transition,” Chem. Mater., 20, 5061–73 (2008). of the Large Strain Response in (K0.5Na0.5)NbO3-Modified (Bi0.5Na0.5)
147
V. A. Isupov, “Ferroelectric Na0.5Bi0.5TiO3 and K0.5Bi0.5TiO3 and Their TiO3-BaTiO3 Lead-Free Piezoceramics,” J. Appl. Phys., 105, 094102 (2009).
177
Solid Solutions,” Ferroelectrics, 315, 123–47 (2005). A. Aydi, H. Khemakhem, C. Boudaya, R. von der Muhll, and A. Simon,
148
G. A. Smolenskii, V. A. Isupov, A. I. Agranovskaya, and N. N. Krainik, “New Ferroelectric and Relaxor Ceramics in the Mixed Oxide System NaN-
“New Ferroelectrics of Complex Composition IV,” Sov. Phys. Solid State, 2, bO3-BaSnO3,” Solid State Sci., 6, 333–7 (2004).
178
2651–4 (1961). R Jimenez, M. L. Sanjuan, and B. Jimenez, “Stabilization of the Ferro-
149
G. O. Jones and P. A. Thomas, “Investigation of the Structure and Phase electric Phase and Relaxor-Like Behavior in Low Li Content Sodium
Transitions in the Novel A-Site Substituted Distorted Perovskite Compound Niobates,” J. Phys. Condens. Matter, 16, 7493–510 (2004).
179
Na0.5Bi0.5TiO3,” Acta Crystallogr., Sect. B, 58, 168–78 (2002). S. I. Raevskaya, V. V. Titov, M. A. Malitskaya, L. A. Reznitchenko, M.
150
S. B. Vakhrushev, V. A. Isupov, B. E. Kvyatkovsky, N. M. Okuneva, I. A. Seredkina, I. P. Raevski, and J. -L- Dellis, “Some Properties of the Relax-
P. Pronin, G. A. Smolensky, and P. P. Syrnikov, “Phase Transitions and Soft or-Like Behavior in Sodium Niobate-Based Binary Solid Solutions,” Ferroelec-
Modes in Sodium Bismuth Titanate,” Ferroelectrics, 63, 153–60 (1985). trics, 374, 122–7 (2008).
151
C. –S. Tu, I. G. Siny, and V. H. Schmidt, “Sequence of Dielectric Anom- 180
I. P. Raevski and S. A. Prosandeev, “A New, Lead Free, Family of
alies and High-Temperature Relaxation Behavior in Na1/2Bi1/2TiO3,” Phys. Perovskites with a Diffuse Phase Transition: NaNbO3 Based Solid Solutions,”
Rev. B, 49, 11550–9 (1994). J. Phys. Chem. Solids, 63, 1939–50 (2002).
152 181
K. Sakata and Y. Masuda, “Ferroelectric and Antiferroelectric Properties H. D. Megaw, “The Seven Phases of Sodium Niobate,” Ferroelectrics, 7,
of (Na0.5Bi0.5)TiO3-SrTiO3 Solid Solution Ceramics,” Ferroelectrics, 7, 347–9 87–9 (1974).
182
(1974). L. E. Cross and B. J. Nicholson, “The Optical and Electrical Properties
153
J. Suchanicz, “The Low-Frequency Dielectric Relaxation Na0.5Bi0.5TiO3 of Single Crystals of Sodium Niobate,” Phil. Mag., 46, 453–66 (1955).
183
Ceramics,” Mater. Sci. Eng. B, 55, 114–8 (1998). C. N. W. Darlington and K. S. Knight, “High Temperature Phases of
154
A. M. Balagurov, E. Yu. Koroleva, A. A. Naberezhnov, V. P. Sak- NaNbO3 and NaTaO3,” Acta Crystallogr., Sect. B, 55, 24–30 (1999).
184
hnenko, B. N. Savenko, N. V. Ter-Oganessian, and S. B. Vakhrushev, “The A. Molak, “The Influence of Reduction in Valency of Nb Ions on the
Rhombohderal Phase with Incommensurate Modulation in Na1/2Bi1/2TiO3,” Antiferroelectric Phase Transition in NaNbO3,” Solid State Commun., 62, 413
Phase Trans., 79, 163–73 (2006). –7 (1987).
155 185
J. Suchanicz, K. Roleder, A. Kania, and J. Hanaderek, “Electrostrictive I. P. Raevskii, L. A. Reznichenko, V. G. Smotrakov, V. V. Eremkin, M.
Strain and Pyroeffect in the Region of Phase Coexistence in Na0.5Bi0.5TiO3,” A. Malitskaya, E. M. Kuznetsova, and L. A. Shilkina, “A New Phase Transi-
Ferroelectrics, 77, 107–10 (1988). tion in Sodium Niobate,” Tech. Phys. Lett., 26, 744–6 (2000).
156 186
J. Kreisel, A. M. Glazer, P. Bouvier, and G. Lucazeau, “High-Pressure S. K. Mishra, N. Choudhury, S. L. Chaplot, P. S. R. Krishna, and
Raman Study of a Relaxor Ferroelectric: The Na0.5Bi0.5TiO3 Perovskite,” R. Mittal, “Competing Antiferroelectric and Ferroelectric Interactions in
Phys. Rev. B, 63, 174106 (2001). NaNbO3: Neutron Diffraction and Theoretical Studies,” Phys. Rev. B, 76,
157
J. Hlinka, J. Petzelt, S. Kamba, D. Noujni, and T. Ostapchuk, “Infrared 024110 (2007).
187
Dielectric Response Of Relaxor Ferroelectrics,” Phase Transitions, 79, 41–78 Z. X. Shen, X. B. Wang, M. H. Kuok, and S. H. Tang, “Raman Scatter-
(2006). ing Investigations of the Antiferroelectric-Ferroelectric Phase Transition of
158
I. P. Aleksandrova, A. A. Sokhovsky, Yu. N. Ivanov, Yu. E. Yablons- NaNbO3,” J. Raman Spectrosc., 29, 379–84 (1998).
188
kaya, and S. B. Vakhrushev, “Local and Average Structure of Relaxor S. Lanfredi, M. H. Lente, and A. J. Eiras, “Phase Transition at Low
Na1/2Bi1/2TiO3 from the Point of View of NMR,” Ferroelectrics, 378, 6–22 Temperature in NaNbO3 Ceramic,” Appl. Phys. Lett., 80, 2731–3 (2002).
189
(2009). J. Cheng, and D. Feng, “TEM Studies of Phase and Domains in NaN-
159
S. A. Sheets, A. N. Soukhojak, N. Ohashi, and Y.-M. Chiang, “Relaxor bO3 at Room Temperature,” Phys. Status Solidi A, 109, 171–85 (1988).
190
Single Crystals in the (Bi1/2Na1/2)1-xBaxZryTi1-yO3 System Exhibiting High S. O. Lisitsyna, I. P. Raevskii, and G. A. Geguzina, “Classification of
Electrostrictive Strain,” J. Appl. Phys., 90, 5287–95 (2001). Binary Sodium Niobate Solid Solutions According to the Properties of Their
160
I. W. Chen, “Structural Origin of Relaxor Ferroelectrics – Revisited,” Second Component,” Sov. Phys. – Tech. Phys., 31, 673–5 (1986).
191
J. Phys. Chem. Solids, 61, 197–208 (2000). M. P. Ivliev, I. P. Raevskii, L. A. Reznichenko, S. I. Raevskaya, and
161
J. Suchanicz, J. Kusz, H. Böhm, H. Duda, J. P. Mercurio, and K. Kon- V. P. Sakhnenko, “Phase States and Dielectric Properties of Sodium–Potassium
ieczny, “Structural and Dielectric Properties of (Na0.5Bi0.5)0.70Ba0.30TiO3 Niobate Solid Solutions,” Phys. Sol. St., 45, 1984–9 (2003).
192
Ceramics,” J. Eur. Ceram. Soc., 23, 1559–64 (2003). B. Aurivillius, “Mixed Bismuth Oxides with Layer Lattices. 1. The Struc-
162
J. Suchanicz, J. Kusz, and H. Böhm, “Structural and Electric Characteris- ture Type of CaNb2Bi2O9,” Ark. Kemi, 1, 463 (1950).
193
tics of (Na0.5Bi0.5)0.50Ba0.50TiO3 and (Na0.5Bi0.5)0.20Ba0.80TiO3 Ceramics,” S. E. Cummins and L. E. Cross, “Electrical an Optical Properties of
Mater. Sci. Eng. B, 97, 154–9 (2003). Ferroelectric Bi4Ti8O12 Single Crystals,” J. Appl. Phys., 39, 2268–74 (1968).
163 194
D. Rout, K.-S. Moon, S.-J. L. Kang, and I. W. Kim, “Dielectric H. Amorı́n, V. V. Shvartsman, A. L. Kholkin, and M. E. V. Costa,
and Raman Scattering Studies of Phase Transitions in the (100-x)Na0.5Bi0.5 “Ferroelectric and Dielectric Anisotropy in High-Quality SrBi2Ta2O9 Single
TiO3-xSrTiO3 System,” J. Appl. Phys., 108, 084102 (2010). Crystals,” Appl. Phys. Lett., 85, 5667–9 (2004).
January 2012 Lead-Free Relaxors 25
195 225
E. C. Subbarao, “A Family of Ferroelectric Bismuth Compounds,” M. Osada, M. Tada, M. Kakihana, T. Watanabe, and H. Fubakubo,
J. Phys. Chem. Solids, 23, 665–76 (1962). “Cation Distribution and Structural Instability in Bi4-xLaxTi3O12,” Jpn.
196
C. A. Paz de Araujo, J. D. Cuchiaro, L. D. McMillan, M. C. Scott, and J. Appl. Phys., 40, 5572–5 (2001).
226
J. F. Scott, “Fatigue-Free Ferroelectric Capacitors with Platinum Electrodes,” C. H. Hervoches and P. Lightfoot, “Cation Disorder in Three-Layer Au-
Nature, 374, 627–9 (1995). rivillius Phases: Structural Studies of Bi2-xSr2+xTi1-xNb2+xO12 (0 < x < 0.8)
197
J. F. Scott, Ferroelectric memories. Springer, Berlin-Heidelberg, 2000. and Bi4-xLaxTi3O12 (x = 1 and 2),” J. Sol. St. Chem., 153, 66–73 (2000).
198 227
G. A. Smolenskii, V. A. Isupov, and A. I. Agranovskaya, “Ferroelectrics Y. Xu, Ferroelectric materials and their applications. Elsevier, North-
of the Oxygen-Octahedral Type with Layered Structure,” Sov. Phys. Solid Holland, Amsterdam, 1999.
State, 3, 651–5 (1961). 228
T. Łukasiewicz, M. A. Swirkowicz, J. Dec, W. Hofman, and W. Szyrski,
199
A. L. Kholkin, M. Avdeev, M. E. V. Costa, J. L. Baptista, and S. N. “Strontium-Barium Niobate Single Crystals, Growth and Ferroelectric Proper-
Dorogovtsev, “Dielectric Relaxation in Ba-Based Layered Perovskites,” Appl. ties,” J. Cryst. Growth, 310, 1464–9 (2008).
229
Phys. Lett., 79, 662–4 (2001). A. Simon and J. Ravez, “Lead Free Relaxors with “TTB” Structure Con-
200
M. Adamczyk, Z. Ujma, and M. Paweczyk, “Dielectric Properties of taining Either Lanthanum or Bismuth,” Phys. Status Solidi A, 199, 541–5 (2001).
230
BaBi2Nb2O9 Ceramics,” J. Mater. Sci., 41, 5317–22 (2006). D. C. Arnold and F. D. Morrison, “B-Cation Effects in Relaxor and Fer-
201
D. Nuzhnyy, S. Kamba, P. Kužel, S. Veljko, V. Bovtun, M. Savinov, roelectric Tetragonal Tungsten Bronzes,” J. Mat. Chem., 19, 6485–8 (2009).
231
J. Petzelt, H. Amorı́n, M. E. V. Costa, A. L. Kholkin, Ph. Boullay, and M. A. Rotaru, D. C. Arnold, A. Daoud-Aladine, and F. D. Morrison, “Ori-
Adamczyk, “Dynamics of the Phase Transitions in Bi-Layered Ferroelectrics gin and Stability of the Dipolar Response in a Family of Tetragonal Tungsten
with Aurivillius Structure: Dielectric Response in the Terahertz Spectral Bronze Relaxors,” Phys. Rev. B, 83, 184302 (2011).
232
Range,” Phys. Rev. B, 74, 134105 (2006). P. P. Liu, X. L. Zhu, and X. M. Chen, “Relaxor Ferroelectric and Mag-
202
V. V. Shvartsman, M. E. V. Costa, M. Avdeev, and A. L. Kholkin, netic Properties of Ba6CoNb9O30 Ceramics with Tungsten Bronze Structure,”
“Relaxor Behavior of BaBi2Ta2O9 and BaBi2Nb2O9 Ceramics,” Feroelectrics, J. Appl. Phys., 106, 074111 (2009).
233
296, 187–97 (2003). M. C. Stennett, I. M. Reaney, G. C. Miles, D. I. Woodward, A. R. West,
203
D. Ben Jannet, M. El Maaoui, and J. P. Mercurio, “Ferroelectric Versus C. A. Kirk, and I. Levin, “Dielectric and Structural Studies of Ba2M-
Relaxor Behavior in Na0.5Bi4.5Ti4O15-BaBi4Ti4O15 Solid Solutions,” J. Electro- Ti2Nb3O15 (BMTNO15, M = Bi3+, La3+, Nd3+, Sm3+, Gd3+) Tetragonal
ceram., 11, 101–6 (2003). Tungsten Bronze-Structured Ceramics,” J. Appl. Phys., 101, 104114 (2007).
204 234
S. K. Rout, E. Sinha, A. Hussian, J. S. Lee, C. W. Ahn, I. W. Kim, and P. B. Jamieson, S. C. Abrahams, and J. L. Bernstein, “Ferroelectric
S. I. Woo, “Phase Transition in ABi4Ti4O15 (A=Ca, Sr, Ba) Aurivillius Oxides Tungsten Bronze-Type Crystal Structures. I. Barium Strontium Niobate
Prepared Through a Soft Chemical Route,” J. Appl. Phys., 105, 024105 (2009). Ba0.27Sr0.75Nb2O5.78,” J. Chem. Phys., 48, 5048–57 (1968).
205 235
H. Irie, M. Miyayama, and T. Kudo, “Electrical Properties of a Bismuth J. R. Oliver, R. R. Neurgaonkar, and L. E. Cross, “A Thermodynamic
Layer-Structured Ba2Bi4Ti5O18 Single Crystal,” J. Am. Ceram. Soc., 83, 2699– Phenomenology for Ferroelectric Tungsten Bronze Sr0.6Ba0.4Nb2O6 (SBN:60),”
704 (2000). J. Appl. Phys., 64, 37–47 (1988).
206
C. Karthik, N. Ravishankar, K. B. R. Varma, M. Maglione, R. 236
S. Miga, W. Kleemann, J. Dec, and T. Łukaciewicz, “Three-Dimensional
Vondermuhll, and J. Etourneau, “Relaxor Behavior of K0.5La0.5Bi2Nb2O9 Random-Field Ising Model Phase Transition in Virgin Sr0.4Ba0.6Nb2O6: Over-
Ceramics,” Appl. Phys. Lett., 89, 042905 (2006). coming Aging,” Phys. Rev. B, 80, 220103 (2009).
207
C. Karthik, N. Ravishankar, M. Maglione, R. Vondermuhll, J. Etour- 237
J. Dec, W. Kleemann, and T. Łukasiewicz, “Critical Slowing Down in the
neau, and K.B.R. Varma, “Relaxor Behavior of K0.5La0.5Bi2Ta2O9 Ceramics,” Single Crystalline Relaxor SBN75,” Phase Transit., 79, 505–11 (2006).
Solid State Commun., 139, 268–72 (2006). 238
V. V. Shvartsman, J. Dec, S. Miga, T. Łukaciewicz, and W. Kleemann,
208
L. Sun, C. Feng, L. Chen, and S. Huang, “Effect of Substitution of “Ferroelectric Domains in SrxBa1-xNb2O6 Single Crystals (0.4  x  0.75),”
Nd3+ for Bi3+ on the Dielectric Properties and Structures of Ferroelectrics, 376, 197–204 (2008).
239
SrBi2xNdxNb2O9 Bismuth Layer-Structured Ceramics,” J. Appl. Phys., 101, T. S. Chernaya, B. A. Maksimov, I. V. Verin, L. I. Ivleva, and V. I. Si-
084102 (2007). monov, “Crystal Structure of Ba0.39Sr0.61Nb2O6,” Crystallogr. Rep., 42, 375–80
209
S. Huang, L. Sun, C. Feng, and L. Chen, “Relaxor Behavior of Layer (1997).
Structured SrBi1.65La0.35Nb2O9,” J. Appl. Phys., 99, 076104 (2006). 240
V. V. Shvartsman, W. Kleemann, T. Łukasiewicz, and J. Dec, “Polar
210
Z. Zhou, X. Dong, S. Huang, and H. Yan, “Dielectric Relaxation of La3+ Structure in SrxBa1-xNb2O6 Single Crystals Studied by Piezoelectric Force
-Modified Bi3TiNbO9 Aurivillius Phase Ceramics,” J. Am. Ceram. Soc., 89, Microscopy,” Phys. Rev. B, 77, 054105 (2008).
241
2939–42 (2006). U. Voelker and K. Betzler, “Domain Morphology from k-space Spectros-
211
X. –B. Chen, R. Hui, J. Zhu, W. –P. Lu, and X. –Y. Mao, “Relaxor copy of Ferroelectric Crystals,” Phys. Rev. B, 74, 132104 (2006).
242
Properties of Lanthanum-Doped Bismuth Layer-Structured Ferroelectrics,” A. S. Bhalla, R. Guo, L. E. Cross, G. Burns, F. H. Dacol, and R. R.
J. Appl. Phys., 96, 5697–700 (2004). Neurgaonkar, “Measurements of Strain and the Optical Indices in the Ferro-
212
Y. Shimakawa, Y. Kubo, Y. Nakagawa, S. Goto, T. Kamiyama, H. electric Ba0.4Sr0.6Nb2O6: Polarization Effects,” Phys. Rev. B, 36, 2030–5
Asano, and F. Izumi, “Crystal Structure and Ferroelectric Properties of (1987).
243
ABi2Ta2O9 (A= Ca, Sr, and Ba),” Phys. Rev. B, 61, 6559–64 (2000). P. Lehnen, W. Kleemann, Th. Woike, and P. Pankrath, “Phase Transi-
213
P. Fang, H. Fan, J. Li, and F. Liang, “Lanthanum Induced Larger tions in Sr0.61Ba0.39Nb2O6:Ce3+: II. Linear Birefringence Studies of Spontane-
Polarization and Dielectric Relaxation in Aurivillius Phase SrBi2-xLaxNb2O9 ous and Precursor Polarization,” Eur. Phys. J. B, 14, 633–7 (2000).
244
Ferroelectric Ceramics,” J. Appl. Phys., 107, 064104 (2010). W. Kleemann, P. Licinio, Th. Woike, and P. Pankrath, “Dynamic Light
214
H. Irie, M. Miyayama, and T. Kudo, “Structure Dependence of Ferro- Scattering at Domains and Nanoclusters in a Relaxor Ferroelectric,” Phys.
electric Properties of Bismuth Layer-Structured Ferroelectric Single Crystals,” Rev. Lett., 86, 6014–7 (2001).
J. Appl. Phys., 90, 4089–94 (2001). 245
J. –H. Ko and S. Kojima, “Intrinsic and Extrinsic Central Peaks in the
215
S. Huang, C. Feng, L. Chen, and Q. Wang, “Relaxor Behavior of Brillouin Light Scattering Spectrum of the Uniaxial Ferroelectric Relaxor
Sr1-xBaxBi2Nb2O9 Ceramics,” J. Am. Ceram. Soc., 89, 328–31 (2006). Sr0.61Ba0.39Nb2O6,” Appl. Phys. Lett., 91, 082903 (2007).
216 246
B. Behera, E. B. Araújo, and A. F. Junior, “Structural and Dielectric W. Kleemann, J. Dec, P. Lehnen, R. Blinc, B. Zalar, and R. Pankrath,
Properties of Relaxor Sr0.5Ba0.5Bi2Nb2O9 Ceramic,” Adv. Appl. Ceram., 109, “Uniaxial Relaxor Ferroelectrics: The Ferroic Random-Field Ising Model
1–5 (2010). Materialized at Last,” Europhys. Lett., 57, 14–9 (2002).
217 247
H. Zhang, H. Yan, and M. J. Reece, “Microstructure and Electrical W. Kleemann, J. Dec, V. V. Shvartsman, S. Kutnjak, and T.h. Braun,
Properties of Aurivillius Phase (CaBi2Nb2O9)1-x(BaBi2Nb2O9)x Solid Solu- “Two-Dimensional Ising Model Criticality in a Three-Dimensional Uniaxial
tion,” J. Appl. Phys., 108, 014109 (2010). Relaxor Ferroelectric with Frozen Polar Nanoregions,” Phys. Rev. Lett., 97,
218
H. Yan, H. Zhang, R. Ubic, M. Reece, J. Liu, and Z. Shen, “Orientation 065702 (2006).
248
Dependence of Dielectric and Relaxor Behavior in Aurivillius Phase J. Dec, V. V. Shvartsman, and W. Kleemann, “Domainlike Precursor
BaBi2Nb2O9 Ceramics Prepared by Spark Plasma Sintering,” J. Mater. Sci: Clusters in the Paraelectric Phase of the Uniaxial Relaxor Sr0.61Ba0.39Nb2O6,”
Mater. Electron., 17, 657–61 (2006). Appl. Phys. Lett., 89, 212901 (2006).
219
R. L. Withers, J. G. Thompson, and A. D. Rae, “The Crystal Chemistry 249
J. Dec, W. Kleemann, S. Miga, V. V. Shvartsman, T. Łukasiewicz, and
Underlying Ferroelectricity in Bi4Ti3O12, Bi3TiNbO9, and Bi2WO6,” J. Solid M. Swirkowicz, “Aging, Rejuvenation, and Memory Effects in the Domain
State Chem., 94, 404–17 (1991). State of Sr0.75Ba0.25Nb2O6, Phase,” Transit., 80, 131–40 (2007).
220 250
A. D. Rae, J. G. Thompson, and R. L. Withers, “Structure Refinement W. Kleemann, “Absence of True Critical Exponents in Relaxor Ferroelec-
of Commensurately Modulated Bismuth Strontium Tantalite,” Bi2SrTa2O9, trics: The Case for Nanodomain Freezing,” J. Phys.: Condens. Matter, 18,
Acta Crystallogr., Sect. B, 48, 818 (1992). L523–6 (2006).
221 251
A. M. Blake, M. J. Falconer, M. McCreedy, and P. Lightfoot, “Cation J. F. Scott, “Absence of True Critical Exponents in Relaxor Ferroelectrics:
Disorder in Ferroelectric Aurivillius Phases of the Type Bi2ANb2O9 (A=Ba, The Case for Defect Dynamics,” J. Phys.: Condens. Matter, 18, 7123–34 (2006).
252
Sr, Ca),” J. Mater. Chem., 7, 1609–13 (1997). T. R. Volk, V. Y. Salobutin, L. I. Ivleva, N. M. Polozkov, R. Pankrath,
222
Ismunandar and B. J. Kennedy, “Effect of Temperature on Cation Disor- and M. Woehlecke, “Ferroelectric Properties of Strontium Barium Niobate
der in ABi2Nb2O9 (A=Sr, Ba),” J. Mater. Chem., 9, 541–4 (1999). Crystals Doped with Rare-Earth Metals,” Phys. Sol. State, 42, 2129–36 (2000).
223 253
J. Tellier, Ph. Boullay, M. Manier, and D. Mercurio, “A Comparative J. Dec, W. Kleemann, Th. Woike, and R. Pankrath, “Phase Transitions
Study of the Aurivillius Phase Ferroelectrics CaBi4Ti4O15 and BaBi4Ti4O15,” in Sr0.61Ba0.39Nb2O6:Ce3+: I. Susceptibility of Clusters and Domains,” Eur.
J. Sol. St. Chem., 177, 1829–37 (2004). Phys. J. B, 14, 627–32 (2000).
224 254
S. Huang, C. Feng, L. Chen, and X. Wen, “Dielectric Properties of SrBi2-x P. Lehnen, E. Beckers, W. Kleemann, T. Woike, and R. Pankrath, “Fer-
PrxNb2O9 Ceramics (x = 0, 0.04 and 0.2),” Solid State Commun., 133, 375–9 roelectric Domains in the Uniaxial Relaxor System SBN: Ce, Cr, and Co,”
(2005). Ferroelectrics, 253, 567–75 (2001).
26 Journal of the American Ceramic Society—Shvartsman and Lupascu Vol. 95, No. 1
255 271
A. Belous, O. V’yunov, D. Mishchuk, S. Kamba, and D. Nuzhnyy, Z. Yu, C. Ang, R. Guo, and A. S. Bhalla, “Piezoelectric and Strain Prop-
“Effect of Vacancies on the Structural and Relaxor Properties of (Sr,Ba,Na) erties of Ba(Ti1-xZrx)O3 Ceramics,” J. Appl. Phys., 92, 1489–93 (2002).
272
Nb2O6,” J. Appl. Phys., 102, 014111 (2007). R. Steinhausen, H. Th. Langhammer, A. Z. Kouvatov, C. Pientschke, H.
256
I. Levin, M. C. Stennet, G. C. Miles, D. I. Woodward, A. R. West, and Beige, and H.-P. Abicht, “Bending Actuators Based on Monolithic Barium
I. M. Reaney, “Coupling Between Octahedral Tilting and Ferroelectric Order Titanate-Stannate Ceramics as Functional Gradient Materials,” Mat. Sci.
in Tetragonal Tungsten Bronze-Structured Dielectrics,” Appl. Phys. Lett., 89, Forum, 494, 167–74 (2005).
273
122908 (2006). D. Viehland, Z. Xu, and W.-H. Huang, “Structure-Property Relation-
257
X. L. Zhu, X. M. Chen, X. Q. Liu, and X. G. Li, “Ferroelectric Phase ships in Strontium Barium Niobate: I. Needle-Like Nanopolar Domains and
Transition and Low-Temperature Structure Fluctuations in Ba4Nd2Ti4Nb6O30 Metastably-Locked Incommensurate Structure,” Philos. Mag. A, 71, 205–17
Tungsten Bronze Ceramics,” J. Appl. Phys., 105, 124110 (2009). (1995).
258 274
X. L. Zhu, X. M. Chen, X. Q. Liu, and X. G. Li, “Dielectric Relaxations, N. Takesue, Y. Fujii, and H. You, “X-Ray Diffuse Scattering Study on
Ultrasonic Attenuation, and their Structure Dependence in Sr4(LaxNd1-x)2 Ionic-Pair Displacement Correlations in Relaxor Lead Magnesium Niobate,”
Ti4Nb6O30 Tungsten Bronze Ceramics,” J. Mater. Res., 23, 3112–21 (2008). Phys. Rev. B, 64, 184112 (2001).
259 275
X. L. Zhu, S. Y. Wu, and X. M. Chen, “Dielectric Anomalies in (BaxSr1-x)4 G. Xu, Z. Zhong, H. Hiraka, and G. Shirane, “Three-Dimensional Map-
Nd2Ti4Nb6O30 Ceramics with Various Radius Differences Between A1- and A2- ping of Diffuse Scattering in Pb(Zn1/3Nb2/3)O3-xPbTiO3,” Phys. Rev. B, 70,
Site Ions,” Appl. Phys. Lett., 91, 162906 (2007). 174109 (2004).
260 276
X. L. Zhu, X. M. Chen, and X. Q. Liu, “Dielectric Abnormity of J. Kreisel, P. Bouvier, B. Dkhil, P. A. Tomas, A. M. Glazer, T. R. Wel-
Sr4Nd2Ti4Nb6O30 Tungsten Bronze Ceramics Over a Broad Temperature berry, B. Chaabane, and M. Mezouar, “High-Pressure X-Ray Scattering of
Range,” J. Mater. Res., 22, 2217–22 (2007). Oxides wih a Nanosclae Local Structure: Application to Na1/2Bi1/2TiO3,”
261
X. L. Zhu, X. Q. Liu, and X. M. Chen, “Crystal Structure and Dielectric Phys. Rev. B, 68, 014113 (2003).
Properties of Sr5RTi3Nb7O30 (R = La, Nd, Sm, and Eu) Tungsten Bronze 277
G. Young, J. Toulouse, R. Erwin, S. M. Shapiro, and B. Hennion,
Ceramics,” J. Am. Ceram. Soc., 94, 1829–36 (2011). “Pretransitional Diffuse Neutron Scattering in the Mixed Perovskite Relaxor
262
J. -H. Ko, S. Kojima, S. G. Lushnikov, R. S. Katiyar, T. –H. Kim, and K1-xLixTaO3,” Phys. Rev. B, 62, 14736 (2000).
J. –H. Ro, “Low-Temperature Transverse Dielectric and Pyroelectric Anoma- 278
Z. Yu, C. Ang, Z. Jing, P. M. Vilarinho, and J. L. Baptista, “Dielectric
lies of Uniaxial Tungsten Bronze Crystals,” J. Appl. Phys., 92, 1536–43 (2002). Properties of Ba(Ti,Ce)O3 from 102 to 105 Hz in the Temperature Range
263
A. Simon, G. T. Joo, M. Maglione, and J. Ravez, “Anisotropic Polar 85–700 K,” J. Phys.: Condens. Matter, 9, 3081–8 (1997).
279
State of Sr0.75Ba0.25Nb2O6 Single Crystal,” Solid State Sci., 9, 52–6 (2007). L. Cui, Y.-D. Hou, S. Wang, C. Wang, and M. -K. Zhu, “Relaxor
264
M. Venet, J. de Los, S. Guerra, I. A. Santos, J. A. Eiras, and D. Garcia, Behavior of (Ba,Bi)(Ti,Al)O3 Ferroelectric Ceramic,” J. Appl. Phys., 107,
“Diffuse Phase Transition and Relaxor Behavior of Textured 054105 (2010).
280
Sr0.63Ba0.37Nb2O6 Ceramics,” J. Phys.: Condens. Matter, 19, 026207 (2007). H. Chaabane, N. Abdelmoula, H. Khemakhem, A. Simon, D. Michau,
265
X. X. Xi, H. –C. Li, W. Si, A. A. Sirenko, I. A. Akimov, J. R. Fox, A. and M. Maglione, “Dielectric And Ferroelectric Properties of Lead-Free Ba1-x
M. Clark, and J. Hao, “Oxide Thin Films for Tunable Microwave Devices,” LaxTi1-xFexO3 Ceramics,” Phys. Status Solidi A, 208, 180–5 (2011).
281
J. Electroceram., 4, 393–405 (2000). S. Kumar and K. B. R. Varma, “Relaxor Behavior of BaBi4Ti3-
266
Y. Somiya, A. S. Bhalla, and L. E. Cross, “Study of (Sr, Pb)TiO3 Ceram- Fe0.5Nb0.5O15 Ceramics,” Solid State Commun., 147, 457–60 (2008).
282
ics on Dielectric and Physical Properties,” Int. J. Inorg. Mat., 3, 709–14 X. L. Zhu, X. M. Chen, X. Q. Liu, and Y. Yuan, “Dielectric Characteris-
(2001). tics and Diffuse Ferroelectric Phase Transition in Sr4La2Ti4Nb6O30 Tungsten
267
C. M. Jackson, J. H. Kobayashi, A. Lee, C. Pettiette-Hall, J. F. Burch, Bronze Ceramics,” J. Mater. Res., 21, 1787–92 (2006).
283
R. Hu, R. Hilton, and J. McDade, “Novel Monolithic Phys Shifter Combining M. X. Cao, X. L. Zhu, X. Q. Liu, and X. M. Cheng, “Crystal Structure
Ferroelectrics and High Temperature Superconductors,” Microwave Opt. Tech- and Ferroelectric Behaviors of Ba5SmTi3Ta7O30 and Ba4Sm2Ti4Ta6O30 Tung-
nol. Lett., 5, 722–6 (1992). sten Bronze Ceramics,” J. Am. Ceram. Soc., 93, 782–6 (2010).
268 284
J. -W. Liou and B. -S. Chiou, “Effect of Direct-Current Biasing on the H. El Alaoui-Belghiti, A. Simon, M. Elaatmani, J. M. Reau, and
Dielectric Properties of Barium Strontium Titanate,” J. Am. Ceram. Soc., 80, J. Ravez, “TKWB-Type Lead-Free Oxyfluoride Relaxors Derived from
3093–9 (1997). Ba2NaNb5O15,” Phys. Status Solidi A, 187, 549–56 (2001).
269 285
Z. Yu, C. Ang, R. Guo, and A. S. Bhalla, “Dielectric Properties and V. V. Shvartsman, A. L. Kholkin, A. Orlova, D. Kiselev, A. A. Bogomo-
High Tunability of Ba(Ti0.7Zr0.3)O3 Ceramics under dc Electric Field,” Appl. lov, and A. Sternberg, “Polar Nanodomains and Local Ferroelectric Phenom-
Phys. Lett., 81, 1285–7 (2002). ena in Relaxor Lead Lanthanum Zirconate Titanate Ceramics,” Appl. Phys.
270
T. Wang, X. M. Chen, and X. H. Zheng, “Dielectric Characteristics and Lett., 86, 202907 (2005). h
Tunability of Barium Stannate Titanate Ceramics,” J. Electroceram., 11, 173–8
(2003).

Vladimir V. Shvartsman studied solid state physics at Moscow State Engineering Physics Institute and
obtained a PhD in physical chemistry from Karpov Institute of Physical Chemistry in Moscow in
2000. His PhD thesis was devoted to sintering and investigations of ferroelectric and relaxor ceramics
for electrocaloric applications. Thereafter he joined the group of Dr. Andrei Kholkin at the Depart-
ment of Ceramic and Glass Engineering at Aveiro University, Portugal, where he has specialized in
piezoresponse force microscopy (PFM) studies of polar dielectrics, especially relaxors. These studies he
continued at the group of Prof. Wolfgang Kleemann at the Physical Faculty of University Duisburg-
Essen, where he moved in the year 2005. Since 2007 he has been involved in investigations of multi-
ferroic and magnetoelectric materials. In 2009 he joined the Institute for Materials Science in Civil
Engineering at University of Duisburg-Essen, where he leads several research projects including devel-
opment of new lead-free piezoelectrics and composite multiferroic materials with a core-shell structure.

Doru C. Lupascu studied physics at Braunschweig and Paris and obtained a PhD in nuclear solid state
physics from University of Göttingen. Already at this time, oxides were his major research field: NMR
studies on high-Tc superconductors and gg angular correlation spectroscopy on rare earth oxides. He
joined the Department of Materials Science at Darmstadt University of Technology in 1996 and found
his research field of passion there, namely ferroelectrics, their fatigue characteristics and their mechan-
ics. His habilitation for Materials Science in 2003 lead to a first tenured professorship at the Institute
of Materials Science at Dresden University of Technology in 2006. Already in 2008, his journey con-
tinued to the Institute for Materials Science in Civil Engineering at University of Duisburg-Essen. The
institute intends to foster the research on functional materials and their incorporation into classical
building materials and structures. Present research fields of the institute span ferroelectrics, magneto-
electrics, organic solar cells, thermal insulation, and durability of concrete.
Copyright of Journal of the American Ceramic Society is the property of Wiley-Blackwell and its content may
not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written
permission. However, users may print, download, or email articles for individual use.

Вам также может понравиться