You are on page 1of 35

Journal Pre-proof

Effects of crystal chemistry and local random fields on relaxor and


piezoelectric behavior of lead-oxide perovskites

Mikhail V. Talanov , Alexei A. Bokov , Mikhail A. Marakhovsky

PII: S1359-6454(20)30300-1
DOI: https://doi.org/10.1016/j.actamat.2020.04.035
Reference: AM 15982

To appear in: Acta Materialia

Received date: 27 August 2019


Revised date: 17 April 2020
Accepted date: 20 April 2020

Please cite this article as: Mikhail V. Talanov , Alexei A. Bokov , Mikhail A. Marakhovsky , Effects of
crystal chemistry and local random fields on relaxor and piezoelectric behavior of lead-oxide per-
ovskites, Acta Materialia (2020), doi: https://doi.org/10.1016/j.actamat.2020.04.035

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd on behalf of Acta Materialia Inc.


Effects of crystal chemistry and local random fields on relaxor and

piezoelectric behavior of lead-oxide perovskites

Mikhail V. Talanov,*a Alexei A. Bokov b and Mikhail A. Marakhovskyc

a
Research Institute of Physics, Southern Federal University, Rostov-on-Don 344090, Russia.

b
Department of Chemistry and 4D LABS, Simon Fraser University, Burnaby, British Columbia,

V5A 1S6, Canada.

c
Institute of High Technologies and Piezotechnics, Southern Federal University, Rostov-on-Don

344090, Russia.

* Corresponding author. E-mail: mvtalanov@gmail.com (M.V. Talanov)

Random electric fields caused by the disordered distribution of heterovalent cations over

the equivalent crystallographic positions are often considered to be the main reason of the

relaxor behavior and the associated giant electromechanical response. Some models relate the

development of the relaxor state to the crystal chemical properties of the constituent ions. In this

work the functions of random fields and ferroactive cations are compared in perovskite solid

solutions in the desirable for technological applications composition range of morphotropic

phase boundary. The crystal structure, dielectric and ferroelectric properties, piezoelectric effect

and electrostriction are investigated in technologically important ceramic solid solutions

PbzBa1−z(Mg1/3Nb2/3)m(Zn1/3Nb2/3)y(Ni1/3Nb2/3)nTixO3 with substitutions in A (Pb2+ by Ba2+)- and

B (Ti4+ by [B2+1/3Nb5+2/3]4+) perovskite sublattices. Two composition sections of the system are

designed; in both of them the concentration of ferroactive ions, Pb2+ and Ti4+, respectively,

varies within the same range of 15%, but the strength of quenched random electric fields changes

significantly only in B-substituted ceramics. Similar behavior is found in both sections, including

the same sequence of structural phase transitions and transformations from normal ferroelectric

to relaxor state and similar evolution of properties with concentration. The results clearly

1
demonstrate that the role of crystal chemical factors in the development of relaxor behavior and

electromechanical performance of lead-oxide materials is not less important than that of

quenched random fields. The mechanisms of the observed behavior are discussed in the frame of

existing theoretical models.

1. Introduction

For more than 70 years since the discovery of ferroelectricity in a BaTiO3 (BT) crystal in

1945 [1], oxides with the ABO3 perovskite structure have been attracting the attention of

researchers due to their numerous applications in various fields of science and technology. It was

shown that hybridization of the 3d state of the titanium ion and the 2p state of the oxygen ion is

essential for the ferroelectric instability in perovskites [2]. The microscopic nature of Ti4+ local

displacements from the center of the oxygen octahedra is associated with a second order Yahn-

Teller effect (pseudo-Yahn-Teller effect) due to participation of unoccupied wave function

dπ(t2g) in the chemical bonding [3]. Another situation occurs in prototypic perovskite

ferroelectric PbTiO3 (PT) where a local off-centering tendency of Pb2+ cations with lone-pair

electrons leads to Pb−O covalent interactions [4]. At the same time, Pb−O hybridization in

PbTiO3 has an indirect effect on Ti−O interaction and is a key factor of much larger

ferroelectricity in PbTiO3 as compared to BaTiO3 [2,5]: the spontaneous polarization at room

temperature and the temperature of ferroelectric phase transition from the cubic to tetragonal

phase are 75 μC/cm2 and 766 K, respectively, in the first compound, and 26 μC/cm2 and 393 K

in the second one [6,7]. Therefore, it is not surprising that solid solutions of (1-

x)PbTiO3−xPbZrO3 (PZT) system containing both lead and titanium ions have become the most

common piezoelectric materials. For technological applications, the optimal composition with

the maximum values of the piezoelectric parameters corresponds to the morphotropic phase

boundary (MPB) (x ≈ 0.52) which separates the rhombohedral (Rh) and tetragonal (T) regions on

the x–T phase diagram [8].

2
The next stage in the development of piezoelectric materials was the use of lead-

containing relaxor ferroelectrics Pb(Mg1/3Nb2/3)O3 (PMN) and Pb(Zn1/3Nb2/3)O3 (PZN) as

components of solid solutions, in particular, with PT. These solid solutions demonstrate ultra-

high values of piezoelectric responses (d33> 2500−4000 pC/N and d33 ≈ 1500 pC/N in the case of

single crystals and ceramics, respectively) near the MPB which coincides in concentration with

ferroelectric-to-relaxor crossover [9−11]. This fact makes it possible to use relaxor-based

materials in a wide range of devices: actuators, sensors, transducers, energy storage devices, self-

powered non-volatile ferroelectric transistor memory as well as in magnetoelectric composite

interfaces [12−17].

From the point of view of phenomenological theory, the phase transition between two

adjacent ferroelectric phases observed near the MPB is essentially associated with a decrease in

the anisotropy of the energy surface potential [18], or, in other words, with the flattening of the

Gibbs free energy profile [19]. Based on the first-principles calculations, Fu and Cohen [20]

showed that the giant piezoelectric effect in these materials can be related to the mechanism of

polarization vector rotation during the electric field (E) induced phase transitions from Rh to T

phase through a series of intermediate low-symmetry phases with close values of free energy. In

1999 the monoclinic phase with the Cm space group (MA) was experimentally observed by

Noheda et al. [21] in PZT solid solutions near the MPB. Subsequently, monoclinic and

orthorhombic intermediate phases were detected in the x−T, x−P and E−T phase diagrams of

relaxor−PT solid solutions [22−33].

Extrinsic contribution to the electric field induced strains related to domain switching

(domain wall motion) is added to the intrinsic contribution in materials with a ferroelectric

domain structure [34]. The coexistence of different crystalline phases in the vicinity of the MPB

can lead to a significant increase in possible domain configurations. As a result, domain

switching processes are simplified and the piezoelectric response increases. In relaxor

ferroelectrics, the situation is complicated due to the development of a nanodomain structure. It

3
is believed that upon cooling of relaxor ferroelectrics below the Burns temperature (TB) the

dynamical nanometer scale polar regions appear in the paraelectric matrix and the crystal

transforms into an ergodic relaxor state [35]. Further cooling leads to an increase in the number

and volume of the polar nanoregions (PNR), and freezing of their dynamics below the Vogel-

Fulcher temperature (TVF), which accompanies the transition to the non-ergodic relaxor state in

the case of canonical relaxor ferroelectrics, e.g. PMN [36−40]. Chemical modification, external

pressure or electric field can induce reversible and irreversible transitions between the ergodic

relaxor, non-ergodic relaxor and ferroelectric states. PNRs make a significant contribution to the

piezoelectric and dielectric properties of relaxor−PT solid solutions [41–45]. In particular, it was

shown that the contribution of PNR to the dielectric and piezoelectric properties can reach up to

50–80% of room temperature values [41].

Apparently, the relaxor state originates from the disordered distribution of different ions

on equivalent crystallographic positions [37,38,46]. The disorder leads to strong random

quenched (static) electric fields in the case of heterovalent substitution, while such fields are

absent in isovalent solid solutions [44,47,48]. Another type of microscopic models that describes

disordered systems is quenched random bonds (interactions), which are typical for spin glasses

[49]. Both of these types of models were actively developed for magnetic systems and are

described by various Hamiltonians (random fields couple to the order parameter linearly and

random bonds couple to the order parameter quadratically) and various phase diagrams [50-52].

However, despite the numerous analogies between magnets and ferroelectrics [53], they still

belong to different classes of universality (due to the different microscopic nature of the

interactions in them), which does not allow the use of well-known models of disordered magnets

to describe the properties of relaxor ferroelectrics. In addition, it is extremely difficult to identify

experimental criteria for separating relaxors with quenched random fields and quenched random

bonds, because both of them can appear in one object. Therefore, the idea of quenched random

4
fields of different strengths in solid solutions with different substitution patterns (hetero- or

isovalent) can be very useful for interpreting experimental results [44, 47].

There are a lot of relaxor ferroelectric systems with a hetero- or isovalent arrangement of

cations both in A (Pb1-xLax(Zr0.65Ti0.35)1-x/4O3 (PLZT), (1-x)Na1/2Bi1/2TiO3 (NBT)−xBaTiO3) and

in B (PbMg1/3Nb2/3O3−PT, BaTi1-xZrxO3, PbSc1/2Ta1/2O3−PT) crystallographic positions

(sublattices) of perovskite structure. However, in most of binary systems involving relaxor

ferroelectrics, the MPB manifests itself in a quite narrow range of concentrations, which

complicates experimental investigation and also hinders the choice of compositions with the

required set of functional parameters. One way to solve this problem is to design

multicomponent solid solutions systems, which significantly enhances the possibility of choosing

functional materials by increasing the dimensionality of the MPB and expanding the composition

range of its existence. This direction in materials science had been developing since the 70s of

the 20th century, and by now a significant amount of multicomponent systems based on PZT,

PMN−PT, PZN−PT, etc. have been investigated [54−61].

A new approach was suggested by the model of x−T phase diagram of piezoelectric solid

solutions proposed by Damjanovic [62]. In this work easy paths for polarization change are

highlighted: polarization rotation and polarization extension. The first mechanism is mainly

responsible for enhancing the piezoelectric response at the morphotropic phase transitions

between two polar phases, as is observed in numerous binary solid solutions such as PZT,

PMN−PT, PZT−PT, etc. The second mechanism is implemented close to phase transitions

between polar and non-polar phases. The combination of both mechanisms within the same

chemical composition (system) looks very appealing from the viewpoint of practical perspective.

Establishing the role of Ti4+, Pb2+ ions and usual in relaxors [B2+1/3Nb5+2/3]4+ complexes

in the formation of electromechanical properties of multicomponent solid solutions near the

MPB, as well as determining the effect of random electric fields on the relaxor-ferroelectric

crossover are important tasks. We investigate this problem using the example of PbzBa1-

5
z(Mg1/3Nb2/3)m(Zn1/3Nb2/3)y(Ni1/3Nb2/3)nTixO3 multicomponent solid solutions, where Ba2+ ions

are replaced by Pb2+ ions in the A-sublattice and in the B-sublattice Ti4+ ions substitute the

[B2+1/3Nb5+2/3]4+ complexes. Our previous studies [63] revealed that the solid solutions of this

system demonstrate large values of permittivity, dielectric tunability and field-induced strain

making them promising candidates for the use in field-drive actuators, tunable devices and

hydrophones. Section 3.1 presents general characterization of the samples and defines the

boundaries of the existence of crystalline phases and polar states (based on X-ray diffraction and

dielectric spectroscopy). Section 3.2 presents the results of the study of electric field-induced

longitudinal strain and polarization. Section 3.3 is devoted to the determination of piezoelectric

and electrostriction coefficients and general relationships of their concentration dependences in

A- and B-substituted solid solutions. The observed features of electromechanical responses and

their possible origins are summarized in section 3.4. It is shown that, despite different crystal

chemical motifs involved in the formation of the relaxor state in A- and B-substituted solid

solutions, the revealed features of dielectric and electromechanical responses appear to be

identical.

2. Experimental

Solid solutions were obtained by solid-phase synthesis using the columbite method [64].

The synthesis of columbite-type compounds MgNb2O6, NiNb2O6, ZnNb2O6 from high-purity

MgO, NiO, ZnO and Nb2O5 oxides involved two stages: for ZnNb2O6 and MgNb2O6 calcinations

were performed at T1 = 1273 K and T2 = 1373 K for 6 h and 4 h, respectively, and for NiNb2O6

at T1 = 1273 K and T2 = 1513 K for 6 h and 2 h, respectively. Solid solutions of the final

compositions were prepared by synthesis from a preformed columbite-type compounds and PbO,

TiO2 and BaCO3 at T = 1223 K for 4 h. Sintering was carried out at temperatures of

1453−1493 K. The relative density of samples was about 95% of the theoretical value.

Experimental samples were disks with the diameter of 10 mm and thickness of 1 mm. Silver

electrodes were deposited by double firing onto disk surfaces.


6
X-ray diffraction investigations were performed by the powder diffraction method using a

DRON-3 diffractometer (CoKα radiation, Bragg–Brentano focusing). Dielectric measurements

were performed using a LCR-meter Agilent E4980A in the temperature range of 298−873 K at

frequencies of the measuring electric field f = 0.1−1000 kHz. The polarization (P−E) and

longitudinal strain (S−E) dependences (f = 50 Hz, bipolar mode) were studied using an

automated test bench operating according to the Sawyer-Tower scheme and equipped with an

optical system for recording strain. Small-signal d33 measurements were performed on polarized

samples using a Berlincourt d33-meter APC YE2730A.

3. Results and discussion

3.1 Phase diagram and relaxor characteristics of the system

Solid solutions of a multicomponent system

PbzBa1−z(Mg1/3Nb2/3)m(Zn1/3Nb2/3)y(Ni1/3Nb2/3)nTixO3 are studied, where x = 0.25‒0.40, y =

0.0842‒0.1130, m = 0.1298‒0.4844, n = 0.1266‒0.4726, z = 0.85‒1.00, m+y+n =1 − x and y/m/n

is constant for all compositions. Based on the results of preliminary X-ray diffraction

investigations [63,65,66], which are partially presented in Supplementary Fig. S1, we chose two

composition sections which demonstrate a similar sequence of structural phase transitions. In the

first section the chemical composition changes only in the B-sublattice: Ti4+ ions replace

[Mg2+1/3Nb5+2/3]4+, [Zn2+1/3Nb5+2/3]4+ and [Ni2+1/3Nb5+2/3]4+ with increasing x. In the second

section only the A-sublattice is modified by replacing Pb2+ with Ba2+ ions. The sections intersect

at the composition with x = 0.30 and z = 0.95, where the solid solution has high functional

parameters: d33 ~ 600 pC/N and dielectric constant (ε) ~ 7000.

Fig. 1(a) and (b) present room-temperature phase-transition sequences determined in our

previous X-ray diffraction studies of these solid solutions [63,65,66]. With increasing x (Ti4+

concentration), as well as z (Pb2+ concentration), the systems transform from the cubic (C) to

pseudocubic (Psc) phase, then to heterophase state (T + Psc) and then to tetragonal (T) phase.

7
However, a single-phase region with T-symmetry is not reached in A-substituted solutions in the

considered concentration range (Fig. 1(b)). The symmetry of the Psc phase could not be

accurately determined using the profile fitting methods because of strong diffuse scattering and

very small distortion of the cubic perovskite unit cell. Based on the literature data on the

structures of binary solid solutions PMN−PT, PZN−PT and PNN−PT [22−29,31−33], we assume

that near the C phase, i.e. at comparatively small x and z, the Psc phase has Rh symmetry

(PscRh), while it is monoclinic (PscM) at comparatively large x and z.

Fig. 1. (a and b) Phase diagrams of the solid solutions obtained using X-ray diffraction at room
temperature. (c and d) Dependences of parameters δ and ∆ on concentrations x (Ti4+) and z
(Pb2+). (e and f) Phase diagrams of the solid solutions obtained using the results of dielectric
spectroscopy. PE is a paraelectric phase, ER is an ergodic relaxor phase, FE is a normal
ferroelectric phase, RFE is a relaxor ferroelectric phase.
We studied earlier the temperature dependences of the relative dielectric permittivity, and

dielectric loss of the same solid solutions [67,68] and found a significant increase in the

temperatures of the permittivity maxima (Tm) with increasing x or z. This behavior is illustrated

8
in Supplementary Figs. S2 and S3. The ε−T dependences changed their shape from diffused,

with a significant shift of Tm toward high temperatures with increasing f, which is typical of

relaxor ferroelectrics, to a relatively sharp shape with frequency-independent Tm as characteristic

of normal ferroelectrics. In order to distinguish different polar states (normal ferroelectric phase,

ergodic relaxor phase, etc.) and to build x/z−T phase diagrams, we determined the main

characteristic temperatures: TB, Curie-Weiss temperatures (TCW), Tm (at f = 1 kHz) and TVF (see

details in Fig. S4−S6). The TCW was determined from the Curie-Weiss law which describes the

temperature behavior of ε in the paraelectric phase:

ε = εadd + C/(T-TCW), (1)

where C is the Curie constant and εadd is an additional parameter that corresponds to the possible

contribution from temperature-independent non-ferroelectric polarization mechanisms. We

found that εadd ˂˂ ε in all studied compositions. Temperature TB was determined from the

deviation of the temperature dependence of ε from the Curie-Weiss law during cooling.

To determine TVF we used the equation characteristic of relaxor ferroelectrics:

f = f0 exp [-Ea/k(Tm-TVF)]. (2)

~
In classical relaxor ferroelectrics, for example in PMN, f0 1013 is the frequency of

attempts to overcome the potential barrier, Ea is the activation energy, k is the Boltzmann

constant and TVF corresponds to the freezing temperature of the dipole dynamics and the

transition to non-ergodic relaxor state. However, such a correspondence is not always valid, and

the transition to non-ergodic relaxor state may not be observed, despite the fulfilment of relation
~
(2) [69]. Many relaxor ferroelectrics, for example, PZN and (1−x) PMN−xPT with x 0.25,

undergo on cooling a diffuse transition from an ergodic relaxor phase to the ferroelectric phase

(RFE as in Fig. 1(e) and (f)), and not to non-ergodic relaxor phase [70]. The transition to the

ferroelectric phase in (1−x)PMN−xPT can be not only diffuse, but also sharp, depending on the

concentration x [70].

9
In connection with the foregoing, we used additional parameters to establish the

boundaries between different polar states. To quantify the degree of diffuseness of the dielectric

maxima, we used the shape parameter, δ, from a quadratic law [71]:

εA/ε = 1+[(T-TA)2/2δ2], (3)

where TA (<Tm), εA (>ε at T=Tm) and δ are fitting parameters. The Tm shift with changing f is

described by the parameter ∆ = Tm (f=100 kHz) − Tm (f=1 kHz), which is convenient as a

measure for the characteristic relaxor dispersion [59,72]. We found that both parameters

decrease significantly with increasing concentrations x (Ti4+) and z (Pb2+) from the values typical

of relaxor ferroelectrics (δ > 40 and ∆ ≥ 5 at x ≤ 0.35 and z ≤ 0.975) to those characteristic of

ferroelectrics with a diffuse phase transition and normal ferroelectrics (Fig. 1(c) and (d)).

Based on the dielectric spectroscopy data, phase diagrams of both composition sections

were constructed (Fig. 1(e) and (f)), which are almost identical:

− the value of TB, near which the transition from the paraelectric to ergodic relaxor state

occurs, is practically independent of the concentration;

− with increasing x and z, the relaxor characteristics weaken, i.e.  and  values decrease,

while TCW and Tm values converge. However, the situation when TCW <Tm, characteristic of

normal ferroelectrics is not achieved in the studied concentration range;

− at a certain critical concentration (x = 0.35 or z = 0.975), the kink in the concentration

dependences of Tm is observed, as well as the fusion/separation of the Tm and TVF dependences.

As x and z decrease, the difference between Tm and TVF increases Similar features were

previously observed in phase diagrams of other relaxor-based systems, i.e. Ba(Ti1−xZrx)O3,

Ba1−x/2□x/2(Ti1−xNbx)O3 [73] and Ba(Ti1−xSnx)O3 [74]. In those works it was noted that the critical

concentration is related to a crossover from relaxor ferroelectric to normal ferroelectric

compositions.

Therefore, at room temperature a transformation from relaxor ferroelectric to normal

ferroelectric state is observed with increasing Ti4+ and Pb2+ concentrations at x  0.35 and z 
10
0.975, respectively. Let us discuss some important details. Solid solutions with x  0.25 and z 

0.90, for which TVF is close to room temperature, are very similar in their structure and dielectric

parameters. For example, for solid solution with z = 0.90, the parameters δ and ∆ are 52.2 and 9,

respectively, and for ceramics with x = 0.25, these parameters are 52.0 and 9, respectively. The

parameters of the unit cells of these solid solutions differ by less than 0.05%. Another important

feature is that they return to their original state after the application of sufficiently high electric

fields, which is characteristic of relaxors in the ergodic phase [37]. In particular, the differences

between ε values before and after the application of the electric field (E > 20 kV/cm) in the case

of all solid solutions of relaxors that are in the ergodic state at room temperature does not exceed

1-5%, and their ε−E dependences are characterized by a hysteresis-free bell-shaped curves (Fig.

S4 and S5) [65,75]. For solid solutions that are in the relaxor ferroelectric phase at room

temperature, the difference in ε values reaches 10–25%, which may be associated with the

transformation from the heterophase (T + PscM) to single-phase state [75,76] and/or with

irreversible changes in ferroelectric domain structure. Note that with a long exposure of
~ ~
sufficiently strong electric fields (t 30 min, E 30 kV/cm), part of the volume of the cubic

sample can irreversibly transform to the normal ferroelectric state, which leads to non-zero d33

values (section 3.3).

Phase diagram of multicomponent system under study is similar to well-known binary

system (1−x)PMN−xPT. In the latter case the MPB and the crossover between the relaxor and

normal ferroelectric states were detected at concentration x ~ 0.30, near which the values of Tm,

TVF, TCW increase with increasing x and TB does not depend on x. A characteristic feature of the

studied multicomponent system is a more significant decrease in Tm (by more than 50 K at x =

0.30) as compared to PMN−PT [77], and, as a result, an increase in the dielectric permittivity at

room temperature from ε ~ 3000 in PMN−PT with x = 0.3072 to ε ~ 7000 in the studied solid

solutions at x = 0.300 and z = 0.950. Another feature of our multicomponent system is a wider

concentration range of the cubic phase in comparison with PMN−PT. It is known that pure PMN

11
retains macroscopically cubic symmetry down to 5 K [78,79]. Even a small (x  0.05) addition of

PT leads to the formation of the rhombohedral phase [80]. In the case of multicomponent system,

significant compositional disorder in B-and A-sublattices leads to stabilization of the cubic phase

and strong smearing of the X-ray reflexes. The latter impedes the structure identification,

especially at boundary compositions (in particular, with x = 0.25 and z = 0.90), which we

designated as relaxors in the ergodic state (at room temperature), but which exhibit non-ergodic

behavior (non-zero d33 after polarization).

3.2 P−E and S−E dependences

Fig. 2 shows the dependences of P and S on AC electric field strength. The evolution of

dielectric and electromechanical hysteresis loops observed with increasing concentrations of

both x (Ti4+) and z (Pb2+), is similar to that observed in relaxor−PT system with decreasing

temperature [81,82]. Ceramics with x < 0.28 and z < 0.90 are characterized by slim P−E loops

and ―sprout‖-like S−E dependences, which are typical of relaxor ferroelectrics in ergodic state

[83,84]. There is practically no hysteresis in these ceramics, which is an important advantage

when using them in actuators. As the composition-driven transitions to the heterophase state and

then to the T-phase (in B-substituted ceramics) occur, the hysteresis and the values of coercive

electric field (Ec) increase. In addition, there is an increase of remanent polarization (Prem) (Fig.

3(a) and (b)) and negative strain (Sneg) (inset in Fig. 2(c), Fig. 3(c) and (d)). As a result, at x ≥

0.30 and z ≥ 0.95, almost rectangular P−E loops and butterfly-like S–E dependences are

observed, which are typical of normal ferroelectrics and relaxor ferroelectrics in nonergodic

state. Therefore, changes in the character of P–E and S–E dependences are consistent with the

results of X-ray diffraction and dielectric spectroscopy considered in the previous section.

12
Fig. 2. (a and b) P−E and (c and d) S−E dependences in (a and c) B-substituted and (b and d) A-
substituted solid solutions of different compositions at room temperature. Double arrow in the
inset indicates a negative strain (Sneg).
The characteristics of P−E and S−E curves are presented in detail in Fig. 3. One can

observe a sharp change of Prem and the formation of a |Sneg| maximum near the boundary

between the heterophase (T + PscM) and single phase (PscM) regions at 0.275 < x < 0.300 and

0.925 < z <0.950. Concentration dependences of Smax (strain measured at E = 15 kV/cm) reveal

two maxima at x = 0.275 and x = 0.35 as well as at z = 0.90, and z = 0.975. Similar behavior of

Smax was observed earlier in unipolar S−E dependences of B-substituted solid solutions measured

in the quasi-static mode [63]. In both sections the first maximum of Smax is formed near the PscRh

phase region, and the position of the second maximum is close to the crossover between the

relaxor ferroelectric and normal ferroelectric.

3.3 Piezoelectric and electrostriction coefficients

From the phenomenological point of view, electric field-induced strain can be

represented as an expansion in terms of E or P [85,86]:


13
Sij = dkijEk + MijklEkEl +…, (4)

Sij = gkijPk + QijklPkPl +…, (5)

where dkij and gkij are the piezoelectric coefficients, Mijkl and Qijkl are the electrostriction

coefficients. However, the use of expression (4) in describing various contributions to the E-

induced strain is complicated by nonlinearities caused by domain switching. It is more rational,

in our case, to apply expression (5) which relates S and P. In ferroelectrics with centrosymmetric

paraelectric phase, only the electrostriction component contribute to the E-induced strain [87].

Therefore, expression (5) takes the form:

Sij = QijklPkPl. (6)

Fig. 4(a) and (b) show S−P dependences in studied solid solutions. At x < 0.275 and z ≤ 0.90,

quadratic (inset in Fig. 4(a)) hysteresis-free dependences are observed similar to those found in

PMN [88]. Using expression (6), the electrostriction coefficient Q33 was estimated to be 0.01–

0.02 m4/C2 (see Fig. 4(c) and (d)). It is comparable to the values obtained for [001] oriented

PMN crystal (0.018–0.027 m4/C2), [111] poled PMN−0.32PT (0.014 m4/C2) and PMN−0.37PT

(0.015 m4/C2) crystals [88,89], as well as lead-free ceramics of the NBT-BT-K1/2Na1/2NbO3

(0.021−0.027 m4/C2) system [90], but inferior to ceramics Ba(Zr0.2Ti0.8)O3−(Ba0.7Ca0.3)TiO3

(0.039−0.050 m4/C2) [91,92], Ba(Ti1-xSnx)O3 (0.0398−0.0515 m4/C2) [93] and textured ceramics

0.97Bi1/2(Na0.78 K0.22)1/2TiO3-0.03BiAlO3 (0.049 m4/C2) [94]. With a further increase of Ti4+ and

Pb2+ concentration, the hysteresis in S−P dependences appears. It does not allow to apply the

expression (6) over the entire range of P. In order to reduce the effect of domain switching

processes on the values of the determined coefficients, we used expression (6) in the high-field

(E >> Ec) region, just as it had been done by Li et al. [91]. To estimate the hysteresis of the S–P

dependences, we used the following formula: H = ∆S (at P = Pmax/2)/Smax. Maximum values of

H (30%) are achieved in heterophase solid solutions with x = 0.30 and z = 0.95 (inset in Fig.

4(b)).

14
Fig. 3. Dependences of (a and b) Pmax (polarization measured at E = 15 kV/cm) and Prem and
(c and d) Smax and Sneg of (a and c) B-substituted and (b and d) A-substituted solid solutions at
room temperature. Error bars were determined from the results of measurements of a series of
samples. In the case of Prem and Sneg, error bars are close to the size of the marker.

Fig. 4(c) and (d) show the dependences of Q33 and small-signal d33 on concentrations.

The behavior of the Q33(x/z) curves of both B-substituted and A-substituted solid solutions

correlates with the behavior of the Smax concentration dependences (Fig. 3(c) and (d)) having two

maxima: a diffused one near the PscRh phase at x = 0.275 and z = 0.90 and a sharper one near the

crossover between relaxor ferroelectric and normal ferroelectric at x = 0.350 and z = 0.975.

Concentration dependences of d33 are characterized by pronounced maxima near the phase

transition between PscRh and T + PscM, reaching d33 = 620 pC/N in the compositions with x =

0.30 and z = 0.95 (this composition is at the intersection of two sections of the system).

15
3.4 Discussion

Electromechanical response features of B-substituted and A-substituted solid solutions

that we found can be divided into three groups, depending on their appearance in the phase

diagrams.

Group I. In the vicinity of the MPB between PscRh and T + PscM regions, a pronounced

maximum of small-signal d33, sharp jumps in Prem and ǀSnegǀ, significant hysteresis of the P−E,

S−E and S−P dependences are observed. This behavior can be explained by the formation of

ferroelectric domains and, as a consequence, by enhancement of the contribution to macroscopic

responses from domain switching due to a large number of possible domain configurations (6T

and 24M domain orientations) and the miniaturization of the domain structure [95−105]. As a

result, the coercive field values are relatively small (Ec ~ 5 kV/cm), which simplifies domain

switching processes. Further growth of x and z leads to an increase in the fraction of the

tetragonal phase and the tetragonality of the unit cell [63,65] and, consequently, to the increase

in Ec. As a result, the mobility of domain walls is reduced and d33 is decreased.

E-field induced phase transition (or E-field driven change in the phase ratio) which we found

earlier [76] in the studied heterophase samples can also contribute to the formation of S−P

hysteresis. The most significant hysteresis is observed in the heterophase samples where the

proportion of the PscM phase is maximum, and, therefore, the T + PscM → T phase transition will

take place in a comparatively large volume.

Group II. The maxima of Smax and Q33 are observed near the transition to the PscRh phase

from the cubic phase of the ergodic relaxor. Enhanced large-field responses can be connected to

the polarization rotation during E-field induced phase transitions from the PscRh to T phase via

intermediate phases [20,31,106,107] (in our case it is PscM phase). Similar phase transitions were

observed in many relaxor−PT systems, for example, in single crystals and ceramics of

(1−x)PMN−xPT with x ≈ 0.30 (Rh → MA → MC → T and MA → MC, respectively) [25,108], in

single crystals of (1−x)PZN−xPT with x = 0.08 (Rh → MA → T or MA → MC → T) [27,109], and


16
in single crystals of Pb(In1/2Nb1/2)O3−PMN−PT [110]. This assumption is supported by the fact

that phase transition from the heterophase (T + PscM) to the single phase (T) state was detected

in ceramics with similar composition (x = 0.30) at values of E comparable with the present

experiment (E ~ 5 kV/cm) [76]. Similar E-field induced phase transition in the (1−x)PMN−xPT

ceramics with x ≈ 0.35 (MC + T → mainly T) occurs together with the build-up of the tetragonal

domain walls, which most likely allows easier switching between the domain variants [111].

Another possible reason for the electromechanical response enhancement is E-field

induced transition from the ergodic relaxor to the ferroelectric state, accompanied by the

transformation of PNRs (or nanodomains) into macrodomains. Associated with such transitions

are the maximum values of d33eff for compositions located near the boundary between the phases

of ergodic and nonergodic relaxors (or relaxor ferroelectric phase) [112−114]. The conclusion

about similar E-field induced transition in B-substituted solid solutions with x = 0.25 was made

by us earlier based on the fact that dielectric dispersion significantly decreases at E > 5 kV/cm

[115].

Group III. The maxima of Smax and Q33 are close to the boundary between the relaxor

ferroelectric and normal ferroelectric (or a ferroelectric with diffuse phase transition). Extrinsic

contributions can influence these features due to the crossover between various mechanisms of

the phase transition to the normal ferroelectric state. According to Ref. [116] in PMN−xPT

crystals with large x (close to normal ferroelectrics), the alignment of PNR dipole moments

dominates at a phase transition. In the case when x is small (relaxor ferroelectrics), an abrupt

growth of PNRs without any change in the directions of their dipole moments takes place. Based

on the found correlation between the boundaries of crystal phases and polar states, it can be

assumed that relaxor properties in the systems are associated with the Psc phase. The

heterophase T + PscM state can be considered as coexistence of comparatively large ferroelectric

T domains and PNRs (or nanodomains) with a different symmetry (PscM). Similar system was

considered as PNR-ferroelectric nanocomposites in the phase-field simulations by Li et al. [41].

17
It was shown that PNRs in a ferroelectric matrix can facilitate polarization rotation and enhance

the shear piezoelectric response. Possible role of preferentially oriented PNR growth in

simplifying the switching of T domains is also reported by Zuo et al. [117].

Fig. 4. (a and b) S–P dependences in (a) B-substituted and (b) A-substituted solid solutions. Inset
in panel (a) shows the strain of the sample with x = 0.25 as a function of P2; the best-fit linear
dependence is shown by the line. Inset in panel (b) presents the hysteresis in S–P dependences as
a function of z. (c and d) dependences of Q33 and d33 on concentrations of (c) Ti4+ and (d) Pb2+.
Despite obvious similarity of three groups of electromechanical response features in both

A-and B-substituted solid solutions, crystal chemical conditions for establishing relaxor state in

them are different. The substitution of Ti4+ ions for the complex (B1/3Nb2/3)4+ (B = Mg2+, Ni2+,

Zn2+) leads to an increase in compositional disorder in B-sublattice. Disordered distribution of

heterovalent B2+, Ti4+ and Nb5+ cations in crystallographically equivalent positions is commonly

expected to be a source of strong random quenched (static) electric field due to the local charge

imbalance. In particular, if in perovskite A2+B4+O23 structure two cations with different

oxidation states (e.g. Mg2+ and Nb5+) replace B cations (e.g. Ti4+), charged perovskite unit cells
18
[A2+ Mg2+O23]2 and [A2+ Nb5+O23]+ are created, giving rise to additional quenched (static)

electric field between these unit cells. In case of isovalent substitution, local electric fields may

also change randomly due to the difference in electronegativities and unit cell distortions

associated with the difference in ionic radii. However, random fields at isovalent substitution are

expected to be comparatively small. Strong random quenched electric field was suggested to

establish the relaxor state in lead-oxide perovskites [44,48]. When such a field is absent, as for

example, in PZT solid solutions, the materials do not typically exhibit relaxor behavior [44].

However, in the case of isovalent substitution of Pb2+ for Ba2+ ions studied in our work, random

electric fields should not change significantly; nevertheless the crossover to relaxor ferroelectric

behavior is observed. This result demonstrates that besides quenched random electric field,

which is the only reason for relaxor behavior according to some theories [48,118] some other

mechanisms can contribute to the development of the relaxor state.

Based on a study of the model Hamiltonian with an additional term describing quenched

random fields of different strengths, it was shown that the appearance and stability of ―random

field state‖ depends on the degree of compositional disorder [48]. In the studied in our work В-

substituted solid solutions not only the strength of random fields is expected to be changed, but

also the concentration of ferroactive (Ti4+) ions changes. However, as can be seen from Fig. 1(e)

and (f), isovalent substitution leads to the same type of phase diagram and similar changes of the

macroscopic responses as in the case of heterovalent substitutions. In both sections of solid

solutions, the concentration of ferroactive (Pb2+ and Ti4+) ions varies within the same range of

15% and the variation is accompanied by almost identical changes in structure and all studied

properties (Figs. 1−4, Supplementary Fig.S8). If we consider the composition near the MPB with

x = 0.30 and z = 0.95 (the point of intersection of two sections) as the "starting point", then the

change in both x and z by 5 mol.% leads to identical dielectric and structural parameters (see

section 3.1). Furthermore, the change in the concentration of Pb2+ in the entire studied range

gives rise to more significant differences of  and  parameters then in B-substituted ceramics

19
(Fig. 1). At the same time, the average size of ionic radius in A- and B-sublattices changes only

by 0.47% and 0.38%, respectively. This means that geometric factor (difference in ionic radii)

does not explain the observed changes in A-substituted ceramics, and peculiarities of covalent

chemical bonds may play a significant role in the development of relaxor, dielectric and

electromechanical properties of the studied solid solutions. We believe that the microscopic

mechanism responsible for establishing the relaxor state is related to random distribution of

chemical Ba-O and Pb−O bonds. Due to significant differences in the electronic structure of

stereoactive Pb2+ ions with a lone pair and Ba2+ with spherical (isotropic) electron density

distribution, corresponding A–O chemical bonds are significantly different in electronegativity

and, as a result, in covalence. In addition, as noted earlier the character of A–O chemical bonds

(hybridization between the states of lead and oxygen or completely ionic interaction between

barium and oxygen) may have an indirect effect on Ti–O interaction [2,5]. In this case the

overlap (hybridization) between the 3d state of titanium and the 2p state of oxygen contributes to

the suppression of short-range inter-ion repulsion and facilitates the formation of spontaneous

polarization [2].

Isovalent substitution of Ba for Pb is known not to lead to relaxor state in (Pb,Ba)TiO3

solid solutions [119]. Therefore, the development of relaxor state with decreasing z in our A-

substituted ceramics can be considered a combined effect of isovalent substitution in A-

sublattice and already existing disorder in B-sublattice which is, however, is not strong enough

to destroy FE state at z=1. In lead-containing relaxor ferroelectrics with heterovalent substitution

in B-sublattice, the local environment of oxygen ions can contribute to (in the case of two high-

valence neighbours) or prevent (in the case of two low-valence neighbours) displacements of

Pb2+ ions and the formation of short chemical bonds [120,121]. Disordered distribution in B-

sublattice of isovalent cations with different tendencies to off-centre displacements [120,122]

due to various electronic properties can lead to the relaxor state even without changing random

electric fields strength. Apparently, the relaxor state in A-substituted ceramics is the result of

20
superposition effect of constant random electric fields (due to fixed ratio of heterovalent ions in

B-sublattice) and the changes in the A−O bond character (due to substitution in A-sublattice)

which does not lead to strong random electric fields. Such behavior can be understood in the

frame of spherical random bond–random field model [123,124] which underlines the importance

of both random fields and random bonds for developing the relaxor phase. This model considers

the Hamiltonian

∑ ̅ ̅ ∑ ̅ ̅,

where ̅ is an order parameter field, related to the dipole moment of ith polar cluster in relaxor,

Jij are randomly frustrated interactions or bonds and ̅ are local quenched electric fields. The

bonds are characterized by Gaussian probability distribution with the mean value J0/N and the

variance J2/N and the random fields by the variance . If J0 < √ the theory predicts the

development of relaxor state at low temperatures, while at J0 > √ the ferroelectric state

becomes stable. In solid solutions with relatively large x and z the second condition is evidently

satisfied, and the material is ferroelectric. With decreasing x (concentration of ferroactive Ti4+

ions), the value of J0 should decrease while  and, possibly, J2, should increase, which results in

the fulfilment of the first condition and appearance of the relaxor state. With a decrease in z

(concentration of ferroactive Pb2+ ions), the value of  remains approximately the same and the

relaxor state appears due to decreasing J0 and increasing J2. Based on our experimental data, we

cannot separate the contribution of the random fields () and the disordered bonds (J0 and J2)

into the development of relaxor behavior and we cannot verify whether strong random electric

field is a necessary condition for the formation of the relaxor state in lead-containing relaxors.

But it can be concluded that the relaxor state with all characteristic features can be induced by

isovalent substitution of ferroactive Pb2+ ions and resulting compositional disorder. Moreover,

such a substitution may affect relaxor properties even more than heterovalent substitution in B-

21
sublattice within the same concentration range (compare Fig. 1(e) and (f)) which changes the

strength of random electric fields.

4. Conclusions

Electromechanical responses of lead-oxide relaxor systems were studied based on the

example of multicomponent ceramics PMN−PNN−PZN−PT with iso- and heterovalent

substitution in А(Pb2+ ions replace Ba2+ ions)- and B(Ti4+ ions replace the [B2+1/3Nb5+2/3]4+

complexes) sublattices of perovskite structure, respectively. Designing the multicomponent

system which combines the possibility of easy paths for both polarization rotation and

polarization extension, was aimed at creating functional materials with high piezo- and dielectric

parameters as well as for evaluation effect of random electric fields strength on the relaxor-

ferroelectric crossover. As a result, it was possible to achieve amplification of macroscopic

responses at room temperature in comparison with the PMN−PT binary system. In particular, ε

values of developed ceramics near the MPB with x = 0.30 and z = 0.95 are twice higher as

compared to the ceramics of the binary system (1−x)PMN−xPT at x = 0.30 (~ 7000 and ~ 3800

[77], respectively) with close values of d33 (~ 600 pC/N).

In both studied composition sections the concentration of ferroactive ions varies within

the same range of 15%, but in B-substituted ceramics the strength of random electric fields is

changed considerably too (the concentration of heterovalent ions almost doubles). However,

isovalent substitution in A-sublattice which does not contribute significantly to the change of

random electric fields strength leads to identical evolution of the structure and all properties and

induces the relaxor state. In the studied system, with increasing concentrations of both Pb2+ in A-

sublattice and Ti4+ in B-sublattice the same sequence of structural phase transitions at room

temperature is observed: C → PscRh → T + PscM → T as well as the same sequence of

transformations between different polar states: relaxor ferroelectric in the ergodic phase →

relaxor ferroelectric in the ferroelectric phase → normal ferroelectric. Concentration

dependences of dielectric and electromechanical properties are similar in both cases. Three

22
groups of electromechanical response features are highlighted, which are observed both in B-

substituted and A-substituted solid solutions: (group I) maxima of small-signal d33, sharp Prem

and ǀSnegǀ jumps, as well as significant hysteresis in P−E, S−E and S−P dependences near the

MPB between PscRh and T + PscM; (group II) maxima of Smax and Q33, observed near the PscRh

phase; (group III) the maxima of Smax and Q33 near the ferroelectric-to-relaxor crossover.

Therefore, in lead-containing perovskites the ferroelectric-to-relaxor crossover caused by

substitution with or without the change of random electric fields is accompanied by identical

transformations of the structure, dielectric and electromechanical properties. Our results

demonstrate that not only quenched random fields, but also crystal chemical features of the

constituent ions can greatly affect the formation of a relaxor state and ferroelectric,

electrostrictive and piezoelectric properties of lead-oxide relaxor materials. The mechanism of

relaxor properties engineering without the significant change of random electric fields strength

identified in this work can be used to create new functional materials with reduced lead content.

Acknowledgements

The study was funded by Russian Science Foundation according to the research project

No. 18-72-00030 (experimental measurements) and by the Ministry of Science and Higher

Education of the Russian Federation (samples preparation).

References

[1] B. Wul, Dielectric constants of some titanates, Nature 156 (1945) 480.
[2] R.E. Cohen, Origin of ferroelectricity in perovskite oxides, Nature 358 (1992)
136−138.
[3] I.B. Bersuker, Pseudo-Jahn–Teller effect—a two-state paradigm in formation,
deformation, and transformation of molecular systems and solids, Chem. Rev. 113
(2013) 1351 – 1390.
[4] A. Walsh, D. J. Payne, R. G. Egdell, G. W. Watson, Stereochemistry of post-
transition metal oxides: revision of the classical lone pair model, Chem. Soc. Rev. 40
(2011) 4455 – 4463.
23
[5] Y. Kuroiwa, S. Aoyagi, A. Sawada, J. Harada, E. Nishibori, M. Takata, M. Sakata,
Evidence for Pb-O covalency in tetragonal PbTiO3, Phys. Rev. Lett. 87 (2001)
217601.
[6] H. Wieder, Electrical behavior of barium titanate single crystals at low temperatures,
Phys. Rev. 99 (1955) 1161– 1165.
[7] V. G. Gavrilyachenko, V. G. Spinko, R. I. Martynenko, E. G. Fesenko, Spontaneous
polarization and coercive field of lead titanate, Sov. Phys.–Solid State 12 (1970)
1203.
[8] B. Jaffe, W. R. Cook, H. Jaffe, Piezoelectric Ceramics, Academic Press, London and
New York, 1971.
[9] S. E. Park, T. R. Shrout, Ultrahigh strain and piezoelectric behavior in relaxor based
ferroelectric single crystals, J. Appl. Phys. 82 (1997) 1804 – 1811.
[10] F. Li, D. Lin, Z. Chen, Z. Cheng, J. Wang, C. C. Li, Z. Xu, Q. Huang, X.
Liao, L.-Q. Chen, T. R. Shrout, S. Zhang, Ultrahigh piezoelectricity in ferroelectric
ceramics by design, Nat. Mater. 17 (2018) 349 – 354.
[11] F. Li, M.J. Cabral, B. Xu, Z. Cheng, E.C. Dickey, J. M. LeBeau, J. Wang, J.
Luo, S. Taylor, W. Hackenberger, L. Bellaiche, Z. Xu, L.-Q. Chen, T.R. Shrout, S.
Zhang, Giant piezoelectricity of Sm-doped Pb(Mg1/3Nb2/3)O3-PbTiO3 single crystals,
Science 364 (2019) 264 – 268.
[12] K. Uchino, Piezoelectric actuators and ultrasonic motors, Kluwer, Boston,
1997.
[13] S. Zhang, F. Li, High performance ferroelectric relaxor-PbTiO3 single
crystals: Status and perspective, J. Appl. Phys. 111 (2012) 031301.
[14] E. Sun, W. Cao, Relaxor-based ferroelectric single crystals: Growth, domain
engineering, characterization and applications, Prog. Mater. Sci. 65 (2014) 124 –
210.
[15] J. Hao, W. Li, J. Zhai, H. Chen., Progress in high-strain perovskite
piezoelectric ceramics, Mater. Sci. Eng. R 135 (2019) 1 – 57.
[16] X. Gao, J. Wu, Y. Yu, Z. Chu, H. Shi, Sh. Dong, Giant Piezoelectric
Coefficients in Relaxor Piezoelectric Ceramic PNN‐PZT for Vibration Energy
Harvesting, Adv. Funct. Mater. 28 (2018) 1706895.
[17] J. M. Hu, L. Q. Chen, C. W. Nan, Multiferroic heterostructures integrating
ferroelectric and magnetic materials, Adv. Mater. 28 (2016) 15 – 39.

24
[18] A. A. Heitmann, G. A. Rossetti, Thermodynamics of ferroelectric solid
solutions with morphotropic phase boundaries, J. Am. Ceram. Soc. 97 (2014) 1661 –
1685.
[19] D. Damjanovic, Contributions to the piezoelectric effect in ferroelectric single
crystals and ceramics, J. Am. Ceram. Soc. 88 (2005) 2663 – 2676.
[20] H. X. Fu, R. E. Cohen, Polarization rotation mechanism for ultrahigh
electromechanical response in single-crystal piezoelectrics, Nature 403 (2000) 281 –
283.
[21] B. Noheda, D. E. Cox, G. Shirane, J. A. Gonzalo, L. E. Cross, S. E. Park, A
monoclinic ferroelectric phase in the Pb(Zr1−xTix)O3 solid solution, Appl. Phys. Lett.
74 (1999) 2059 – 2061.
[22] B. Noheda, D. E. Cox, G. Shirane, J. Gao, Z.-G. Ye, Phase diagram of the
ferroelectric relaxor (1−x)PbMg1/3Nb2/3O3−xPbTiO3, Phys. Rev. B 66 (2002)
054104.
[23] Z. G. Ye, B. Noheda, M. Dong, D. Cox, G. Shirane, Monoclinic phase in the
relaxor-based piezoelectric/ferroelectric Pb(Mg1/3Nb2/3)O3−PbTiO3 system, Phys.
Rev. B 64 (2001) 184114.
[24] M. Ahart, S. Sinogeikin, O. Shebanova, D. Ikuta, Z.-G. Ye, H. K. Mao, R. E.
Cohen, R. J. Hemley, Pressure dependence of the monoclinic phase in
(1−x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 solid solutions, Phys. Rev. B 86 (2012) 224111.
[25] F. Bai, N. Wang, J. Li, D. Viehland, P. M. Gehring, G. Xu, G. Shirane, X-ray
and neutron diffraction investigations of the structural phase transformation sequence
under electric field in 0.7Pb(Mg1∕3Nb2∕3)-0.3PbTiO3 crystal, J. Appl. Phys. 96 (2004)
1620 – 1627.
[26] D. La-Orauttapong, B. Noheda, Z.-G. Ye, P. M. Gehring, J. Toulouse, D. E.
Cox, G. Shirane, Phase diagram of the relaxor ferroelectric
(1−x)Pb(Zn1/3Nb2/3)O3−xPbTiO3, Phys. Rev. B 65 (2002) 144101.
[27] K. Ohwada, K. Hirota, P. W. Rehrig, Y. Fujii, G. Shirane, Neutron diffraction
study of field-cooling effects on the relaxor ferroelectric Pb[(Zn1/3Nb2/3)0.92Ti0.08]O3,
Phys. Rev. B 67 (2003) 094111.
[28] S. P. Singh, A. K. Singh, D. Pandey, S. M. Yusuf, Dielectric relaxation and
phase transitions at cryogenic temperatures in 0.65[Pb(Ni1∕3Nb2∕3)O3]−0.35PbTiO3
ceramics, Phys. Rev. B 76 (2007) 054102.
[29] A. A. Bokov, Z.-G. Ye, Ferroelectric properties of monoclinic
Pb(Mg1/3Nb2/3)O3−PbTiO3 crystals, Phys. Rev. B 66 (2002) 094112.
25
[30] R. Haumont, A. Al-Barakaty, B. Dkhil, J. M. Kiat, L. Bellaiche, Morphotropic
phase boundary of heterovalent perovskite solid solutions: Experimental and
theoretical investigation of PbSc1∕2Nb1∕2 O3−PbTiO3, Phys. Rev. B 71 (2005) 104106.
[31] B. Noheda, D. E. Cox, G. Shirane, S. E. Park, L. E. Cross, Z. Zhong,
Polarization rotation via a monoclinic phase in the piezoelectric 92% PbZn1/3Nb2/3O3
– 8%PbTiO3, Phys. Rev. Lett. 86 (2001) 3891 – 3894.
[32] J. M. Kiat, Y. Uesu, B. Dkhil, M. Matsuda, C. Malibert, G. Calvarin,
Monoclinic structure of unpoled morphotropic high piezoelectric PMN-PT and PZN-
PT compounds, Phys. Rev. B 65 (2002) 064106.
[33] A. K. Singh, D. Pandey, Structure and the location of the morphotropic phase
boundary region in (1-x)[Pb(Mg1/3Nb2/3)O3]-xPbTiO3, J. Phys.: Condens. Matter 13
(2001) L931 – L936.
[34] D. Damjanovic, M. Demartin, Contribution of the irreversible displacement of
domain walls to the piezoelectric effect in barium titanate and lead zirconate titanate
ceramics, J. Phys.: Condens. Matter 9 (1997) 4943 – 4953.
[35] G. Burns, F. H. Dacol, Crystalline ferroelectrics with glassy polarization
behavior, Phys. Rev. B 28 (1983) 2527 – 2530.
[36] D. Viehland, S. J. Jang, L. E. Cross, M. Wuttig, Freezing of the polarization
fluctuations in lead magnesium niobate relaxors, J. Appl. Phys. 68 (1990) 2916 –
2921.
[37] A. A. Bokov, Z.-G Ye, Recent progress in relaxor ferroelectrics with
perovskite structure, J. Mater. Sci. 41 (2006) 31 – 52.
[38] G. A. Samara, The relaxational properties of compositionally disordered
ABO3 perovskites, J. Phys.: Condens. Matter 15 (2003) R367 – R411.
[39] B. Dkhil, P. Gemeiner, A. Al-Barakaty, L. Bellaiche, E. Dul‘kin, E. Mojaev,
M. Roth, Intermediate temperature scale T* in lead-based relaxor systems, Phys.
Rev. B 80 (2009) 064103.
[40] A. A. Bokov, Z.-G Ye, Double freezing of dielectric response in relaxor
Pb(Mg1∕3Nb2∕3)O3 crystals, Phys. Rev. B 74 (2006) 132102.
[41] F. Li, S. Zhang, T. Yang, Z. Xu, N. Zhang, G. Liu, J. Wang, J. Wang, Z.
Cheng, Z.-G. Ye, J. Luo, T. R. Shrout, L.-Q. Chen, The origin of ultrahigh
piezoelectricity in relaxor-ferroelectric solid solution crystals, Nature Comm. 7
(2016) 13807.

26
[42] F. Li, S. Zhang, Z. Xu, L.-Q. Chen, The contributions of polar nanoregions to
the dielectric and piezoelectric responses in domain‐engineered relaxor‐PbTiO3
Crystals, Adv. Funct. Mater. 27 (2017) 1700310.
[43] M.E. Manley, D. L. Abernathy, R. Sahul, D. E. Parshall, J. W. Lynn, A. D.
Christianson, P. J. Stonaha, E. D. Specht, J. D. Budai, Giant electromechanical
coupling of relaxor ferroelectrics controlled by polar nanoregion vibrations, Sci.
Adv. 2 (2016) e1501814.
[44] D. Phelan, C. Stock, J. A. Rodriguez-Rivera, S. Chi, J. Leão, X. Long, Y. Xie,
A. A. Bokov, Z.-G. Ye, P. Ganesh, P. M. Gehring, Role of random electric fields in
relaxors, Proc. Natl Acad. Sci. USA 111 (2014) 1754 – 1759.
[45] F. Li, S. Zhang, D. Damjanovic, L.-Q. Chen, T. R. Shrout, Local structural
heterogeneity and electromechanical responses of ferroelectrics: Learning from
relaxor ferroelectrics. Adv. Funct. Mater. 28 (2018) 1801504.
[46] S. Tinte, B. P. Burton, E. Cockayne, U. V. Waghmare, Origin of the relaxor
state in Pb(BxB′1−x)O3 perovskites, Phys. Rev. Lett. 97 (2006) 137601.
[47] W. Kleemann, Relaxor ferroelectrics: Cluster glass ground state via random
fields and random bonds, Phys. Status Solidi B 251 (2014) 1993 – 2002.
[48] J. R. Arce-Gamboa, G. G. Guzmán-Verri, Random electric field instabilities
of relaxor ferroelectrics, npj Quantum Mater. 2 (2017) 28.
[49] K. H. Fischer, J.A. Hertz, Spin glasses, Cambridge University Press, Cambridge,
1993.
[50] U. T. Höchli, K. Knorr, A. Loidl, Orientational glasses, Adv. Phys. 51 (2002) 589
– 798.
[51] K. Binder, A. P.Young, Spin glasses: Experimental facts, theoretical concepts, and
open questions, Rev. Mod. Phys. 58 (1986) 801 – 976.
[52] V. S. Dotsenko, Critical phenomena and quenched disorder, Phys. Usp. 38 (1995)
457 – 496.
[53] R. A. Cowley, S. N. Gvasaliya, S. G. Lushnikov, B. Roessli, G. M. Rotaru,
Relaxing with relaxors: a review of relaxor ferroelectrics, Adv. Phys. 60 (2011) 229 –
327.
[54] Е. G. Fesenko, A. Ya. Dantsiger, O. N. Razumovskaya, Novel piezoelectric
ceramic materials, Rostov University Press, Rostov-on-Don, 1983.
[55] J.-S. Park, J.-K. Lee, H. Park, K. S. Hong, Ferroelectric properties of
Pb(Zn1/3Nb2/3)O3–PbTiO3–RNbO3(R=Na, K) ceramics, J. Am. Ceram. Soc. 90
(2007) 3512 – 3516.
27
[56] N. Luo, S. Zhang, Q. Li, C. Xu, Z. Yang, Q. Yan, Y. Zhang, T. R. Shrout,
New Pb(Mg1/3Nb2/3)O3–Pb(In1/2Nb1/2)O3–PbZrO3–PbTiO3 quaternary ceramics:
Morphotropic phase boundary design and electrical properties, ACS Appl. Mater.
Interfaces 8 (2016) 15506 – 15517.
[57] Y. Chen, J. Fu, R. Zuo, Electric field induced irreversible change and
asymmetric butterfly strain loops in Pb(Zr,Ti)O3-Pb(Ni1/3Nb2/3)O3-Bi(Zn1/2Ti1/2)O3
quaternary ceramics, Ceram. Int. 44 (2018) 8514 – 8520.
[58] Y. Zhang, X. Zhua, J. Zhu, X. Zeng, X. Feng, J. Liao, Composition design,
phase transitions and electrical properties of Sr2+-substituted xPZN–0.1PNN–
(0.9−x)PZT piezoelectric ceramics, Ceram. Int. 42 (2016) 4080 – 4089.
[59] M.V. Talanov, A.A. Bush, K.E. Кamentsev, V.P. Sirotinkin, A.G. Segalla,
Structure‐property relationships in BiScO3–PbTiO3–PbMg1/3Nb2/3O3 ceramics near
the morphotropic phase boundary, J. Am. Ceram. Soc. 101 (2018) 683 – 693.
[60] D. Lin, C. Li, S. Ge, E. Gorzkowski, S. Zhou, W. Liu, F. Li, High
rhombohedral to tetragonal phase transition temperature and electromechanical
response in Pb(Yb1/2Nb1/2)O3-Pb(Sc1/2Nb1/2)O3-PbTiO3 ferroelectric system near the
morphotropic phase boundary, J. Eur. Ceram. Soc. 39 (2019) 2082 – 2090.
[61] Z. Liu, A. R. Paterson, H. Wu, P. Gao, W. Ren, Z.-G. Ye, Synthesis, structure
and piezo-/ferroelectric properties of a novel bismuth-containing ternary complex
perovskite solid solution, J. Mater. Chem. C 5 (2017) 3916 – 3923.
[62] D. Damjanovic, A morphotropic phase boundary system based on polarization
rotation and polarization extension, Appl. Phys. Lett. 97 (2010) 062906.
[63] M. V. Talanov, L. A. Shilkina, L. A. Reznichenko, Anomalies of the dielectric
and electromechanical responses of multicomponent ceramics on the basis of PMN–
PT near the morphotropic phase boundary, Sens. Actuator A Phys. 217 (2014) 62 –
67.
[64] L. Swarz, T. R. Shrout, Fabrication of perovskite lead magnesium niobate,
Mater. Res. Bull. 17 (1982) 1245 – 1250.
[65] M. V. Talanov, L. A. Shilkina, I. A. Verbenko, L. A. Reznichenko, Impact of
Ba2+ on Structure and Piezoelectric Properties of PMN–PZN–PNN–PT Ceramics
Near the Morphotropic Phase Boundary, J. Am. Ceram. Soc. 98 (2015) 838 – 847.
[66] M. V. Talanov, O. N. Razumovskaya, L. A. Shilkina, L. A. Reznichenko,
Effect of barium on the structure and dielectric properties of multicomponent
ceramics based on ferroelectric relaxors, Inorg. Mat. 49 (2013) 957 – 961.

28
[67] M. V. Talanov, S. P. Kubrin, A. A. Pavelko, L. A. Reznichenko, Dielectric
spectroscopy of Pb1–x Bax (Mg1/3Nb2/3)m(Zn1/3Nb2/3)y(Ni1/3Nb2/3)nTizO3 solid solutions
in a wide temperature interval, Phys. Sol. State 58 (2016) 1160 – 1165.
[68] M. V. Talanov, L. A. Reznichenko, Phase Diagrams of Solid Solutions of
Relaxor Ferroelectrics from the Dielectric Spectroscopy Data, Phys. Sol. State 60
(2018) 437 – 441.
[69] A. K. Tagantsev, Vogel-Fulcher relationship for the dielectric permittivity of
relaxor ferroelectrics, Phys. Rev. Lett. 72 (1994) 1100.
[70] Y.-H. Bing, A. A. Bokov, Z.-G. Ye, Diffuse and sharp ferroelectric phase
transitions in relaxors, Curr. Appl. Phys. 11 (2011) S14 – S21.
[71] A. A. Bokov, Z.-G. Ye, Phenomenological description of dielectric
permittivity peak in relaxor ferroelectrics, Solid State Commun. 116 (2000) 105 –
108.
[72] J. Zhuang, A. A. Bokov, N. Zhang, D. Walker, S. Huo, J. Zhang, W. Ren, Z.-
G. Ye, Impact of quenched random fields on the ferroelectric-to-relaxor crossover in
the solid solution(1−x)BaTiO3−xDyFeO3, Phys. Rev. B 98 (2018) 174104.
[73] A. Simon, J. Ravez, M. Maglione, The crossover from a ferroelectric to a
relaxor state in lead-free solid solutions, J. Phys.: Condens. Matter 16 (2004) 963–
970.
[74] C. Lei, A. A. Bokov, Z.-G. Ye, Ferroelectric to relaxor crossover and
dielectric phase diagram in the BaTiO3–BaSnO3 system, J. Appl. Phys. 101 (2007)
084105.
[75] M. V. Talanov, L. A. Shilkina, L. A. Reznichenko, Evolution of domain
processes during the transition from classical ferroelectric to relaxor ferroelectric,
Phys. Sol. State 54 (2012) 990 – 991.
[76] M. V. Talanov, O. A. Bunina, M. A. Bunin, I. N. Zakharchenko, L. A.
Reznichenko, Electric-field-induced phase transition in the relaxor ceramics based on
PMN-PT, Phys. Sol. State 55 (2013) 326 – 333.
[77] J. Kelly, M. Leonard, C. Tantigate, A. Safari, Effect of composition on the
electromechanical properties of (1‐x)Pb(Mg1/3Nb2/3)O3−XPbTiO3 ceramics, J. Am.
Ceram. Soc. 80 (1997) 957 – 964.
[78] P. Bonneau, P. Garnier, G. Calvarin, E. Husson, J. R. Gavarri, A. W. Heiwat,
A. Morell, X-ray and neutron diffraction studies of the diffuse phase transition in
ceramics, J. Solid State Chem. 91 (1991) 350 – 361.

29
[79] N. de Mathan, E. Husson, J.R. Gavarri, A.W. Heiwat, A. Morell, A structural
model for the relaxor PbMg1/3Nb2/3O3 at 5K, J. Phys.: Condens. Matter 3 (1991)
8159 – 8171.
[80] Z.-G. Ye, Y. Bing, J. Gao, A.A. Bokov, P. Stephens, B. Noheda, G. Shirane,
Development of ferroelectric order in relaxor (1−x)Pb(Mg1/3Nb2/3)O3–
xPbTiO3(0<~x<~0.15), Phys. Rev. B 67 (2003) 104104.
[81] J. C. Ho, K. S. Liu, I. N. Lin, Study of ferroelectricity in the PMN-PT system
near the morphotropic phase boundary, J. Mater. Sci. 28 (1993) 4497 – 4502.
[82] H.-Y. Chen, C.-S. Tu, C.-M. Hung, R. R. Chien, V. H. Schmidt, C.-S. Ku, H.-
Y. Lee, Poling effect and piezoelectric response in high-strain ferroelectric
0.70Pb(Mg1/3Nb2/3)O3–0.30PbTiO3 crystal, J. Appl. Phys. 108 (2010) 044101.
[83] W. Jo, R. Dittmer, M. Acosta, J. Zang, C. Groh, E. Sapper, K. Wang, J. Rödel,
Giant electric-field-induced strains in lead-free ceramics for actuator applications –
status and perspective, J. Electroceramics 29 (2012) 71 – 93.
[84] L. Jin, F Li, S Zhang, Decoding the fingerprint of ferroelectric loops:
Comprehension of the material properties and structures, J. Am. Ceram. Soc. 97
(2014) 1 – 27.
[85] J. F. Nye, Physical properties of crystals, Oxford Science Publications,
Oxford, 1993.
[86] R. E. Newnham, Properties of materials: Anisotropy, Symmetry, Structure,
Oxford University Press, Oxford, 2005.
[87] M. E. Lines, A. M. Glass, Principles and applications of ferroelectrics and
related materials, Clarendon, Oxford, 1979.
[88] F. Li, L. Jin, Z. Xu, D. Wang, S. Zhang, Electrostrictive effect in
Pb(Mg1/3Nb2/3)O3-xPbTiO3 crystals, Appl. Phys. Lett. 102 (2013) 152910.
[89] F. Li, L. Jin, Z. Xu, S. Zhang, Electrostrictive effect in ferroelectrics: An
alternative approach to improve piezoelectricity, Appl. Phys. Rev. 1 (2014) 011103.
[90] S.-T. Zhang, A. B. Kounga, W. Jo, C. Jamin, K. Seifert, T. Granzow, J. Rödel,
D. Damjanovic, High‐Strain Lead‐free Antiferroelectric Electrostrictors, Adv. Mater.
21 (2009) 4716 – 4720.
[91] F. Li, L. Jin, R. Guo, High electrostrictive coefficient Q33 in lead-free
Ba(Zr0.2Ti0.8)O3-x(Ba0.7Ca0.3)TiO3 piezoelectric ceramics, Appl. Phys. Lett. 105
(2014) 232903.
[92] L. Jin, R. Huo, R. Guo, F. Li, D. Wang, Ye Tian, Q. Hu, X. Wei, Z. He, Y.
Yan, G. Liu, Diffuse phase transitions and giant electrostrictive coefficients in lead-
30
free Fe3+-doped 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Ca0.3)TiO3 ferroelectric ceramics, ACS
Appl. Mater. Interfaces 8 (2016) 31109 – 31119.
[93] L. Jin, J. Qiao, L. Hou, Y. Tian, Q. Hu, L. Wang, X. Lu, L. Zhang, H. Du, X.
Wei, G. Liu, Y. Yan, A strategy for obtaining high electrostrictive properties and its
application in barium stannate titanate lead-free ferroelectrics, Ceram. Int. 44 (2018)
21816 – 21824.
[94] C. W. Ahn, G. Choi, I. W. Kim, J.-S. Lee, K. Wang, Y. Hwang, W. Jo, Forced
electrostriction by constraining polarization switching enhances the
electromechanical strain properties of incipient piezoceramics, npj Asia Mater. 9
(2017) e346.
[95] K. A. Schönau, L. A. Schmitt, M. Knapp, H. Fuess, R.-A. Eichel, H. Kungl,
M.J. Hoffmann, Nanodomain structure of Pb[Zr1−x Tix]O3 at its morphotropic phase
boundary: Investigations from local to average structure, Phys. Rev. B 75 (2007)
184117.
[96] K. A. Schönau, M. Knapp, H. Kungl, M. J. Hoffmann, H. Fuess, In situ
synchrotron diffraction investigation of morphotropic Pb[Zr1−xTix]O3 under an
applied electric field, Phys. Rev. B 76 (2007) 144112.
[97] L. A. Schmitt, K. A. Schönau, R. Theissmann, H. Fuess, H. Kungl, M. J.
Hoffmann, Composition dependence of the domain configuration and size in
Pb(Zr1−xTix)O3 ceramics, J. Appl. Phys. 101 (2007) 074107.
[98] J. Fu, R. Zuo, Z. Xu, High piezoelectric activity in (Na,K)NbO3 based lead-
free piezoelectric ceramics: Contribution of nanodomains, Appl. Phys. Lett. 99
(2011) 062901.
[99] J. Y. Li, R. C. Rogan, E. Üstündag, K. Bhattacharya, Domain switching in
polycrystalline ferroelectric ceramics, Nat. Mater. 4 (2005) 776 – 781.
[100] J. Ouyang, A. L .Roytburd, Theoretical modeling of coexisting tetragonal and
rhombohedral heterophase polydomain structures in lead zirconate titanate
ferroelectric films near the morphotropic phase boundary, Acta Mater. 54 (2006)
5565 – 5572.
[101] Y. U. Wang, Three intrinsic relationships of lattice parameters between
intermediate monoclinic MC and tetragonal phases in ferroelectric
Pb[(Mg1∕3Nb2∕3)1−xTix]O3 and Pb[(Zn1∕3Nb2∕3)1−xTix]O3 near morphotropic phase
boundaries, Phys. Rev. B 73 (2006) 014113.
[102] W.-F. Rao, Yu U. Wang, Bridging domain mechanism for phase coexistence
in morphotropic phase boundary ferroelectrics, Appl. Phys. Lett. 90 (2007) 182906.
31
[103] H. Wang, J. Zhu, N. Lu, A. A. Bokov, Z.-G. Ye, X. W. Zhang, Hierarchical
micro-/nanoscale domain structure in MC phase of (1−x)Pb(Mg1∕3Nb2∕3)O3–xPbTiO3
single crystal, Appl. Phys. Lett. 89 (2006) 042908.
[104] W.-Y. Chang, C.-C. Chung, Z. Yuan, C.-H. Chang, J. Tian, D. Viehland, J.-F. Li,
J. L. Jones, X. Jiang, Patterned nano-domains in PMN-PT single crystals, Acta Mater.
143 (2018) 166 – 173.
[105] C. Luo, W.-Y. Chang, M. Gao, C. Chih-Hao, J.-F. Li, D. Viehland, J. Tian,
X. Jiang, Multi-layered domain morphology in relaxor single crystals with nano-
patterned composite electrode, Acta Mater. 182 (2020) 10 – 17.
[106] H. Liu, J. Chen, L. Fan, Y. Ren, Z. Pan, K. V. Lalitha, J. Rödel, X. Xing,
Critical role of monoclinic polarization rotation in high-performance perovskite
piezoelectric materials, Phys. Rev. Lett. 119 (2017) 017601.
[107] H. Liu, J. Chen, L. Fan, Y. Ren, L. Hu, F. Guo, J. Deng, X. Xing, Structural
evidence for strong coupling between polarization rotation and lattice strain in
monoclinic relaxor ferroelectrics, Chem. Mater. 29 (2017) 5767 – 5771.
[108] D. Hou, T.-M. Usher, L. Fulanovic, M. Vrabelj, M. Otonicar, H. Ursic, B.
Malic, I. Levin, J.L. Jones, Field-induced polarization rotation and phase transitions
in 0.70Pb (Mg1/3Nb2/3)O3−0.30 PbTiO3 piezoceramics observed by in situ high-
energy x-ray scattering, Phys. Rev. B 97 (2018) 214102.
[109] B. Noheda, Z. Zhong, D. E. Cox, G. Shirane, S-E. Park, P. Rehrig, Electric-
field-induced phase transitions in rhombohedral Pb(Zn1/3Nb /3)1−xTixO3, Phys. Rev. B
65 (2002) 224101.
[110] Y. Wang, Z. Wang, W. Ge, C. Luo, J. Li, D. Viehland, J. Chen, H. Luo,
Temperature-induced and electric-field-induced phase transitions in rhombohedral
Pb(In1/2Nb1/2)O3−Pb(Mg1/3Nb2/3)O3−PbTiO3 ternary single crystals, Phys. Rev. B 90
(2014) 134107.
[111] M. Otonicar, H. Ursic, M. Dragomir, A. Bradesko, G. Esteves, J.L. Jones, A.
Bencan, B. Malic, T. Rojac, Multiscale field-induced structure of (1-
x)Pb(Mg1/3Nb2/3)O3–xPbTiO3 ceramics from combined techniques, Acta Mater. 154
(2018) 14 – 24.
[112] W. Zhao, R. Zuo, J. Fu, M. Shi, Large strains accompanying field-induced
ergodic phase-polar ordered phase transformations in Bi(Mg0.5Ti0.5)O3–PbTiO3–
(Bi0.5Na0.5)TiO3 ternary system, J. Eur. Ceram. Soc. 34 (2014) 2299 – 2309.
[113] D. Zheng, R. Zuo, Relaxor-normal ferroelectric phase transition and
significantly enhanced electromechanical strain behavior in Bi(Ni1/2Ti1/2)O3–
32
PbTiO3–Pb(Mg1/3Nb2/3)O3 ternary system close to the morphotropic phase boundary,
J. Eur. Ceram. Soc. 35 (2015) 3485 – 3493.
[114] F. Li, R. Zuo, D. Zheng, L. Li, Phase‐composition‐dependent piezoelectric
and electromechanical strain properties in (Bi1/2Na1/2)TiO3–Ba(Ni1/2Nb1/2)O3 lead‐
free ceramics, J. Am. Ceram. Soc. 98 (2015) 811 – 818.
[115] M. V. Talanov, A. V. Turik, L. A. Reznichenko, Reversible permittivity of
multicomponent PMN-PT-based ceramics, Tech. Phys. 58 (2013) 1608 – 1613.
[116] J.-H. Ko, D. H. Kim, S. Tsukada, S. Kojima, A. A. Bokov, Z.-G. Ye,
Crossover in the mechanism of ferroelectric phase transition of
Pb[(Mg1/3Nb2/3)1−xTix]O3 single crystals studied by Brillouin light scattering, Phys.
Rev. B 82 (2010) 104110.
[117] R. Zuo, F. Li, J. Fu, D. Zheng, W. Zhao, H. Qi, Electric field forced c-axis
oriented growth of polar nanoregions and rapid switching of tetragonal domains in
BNT-PT-PMN ternary system, J. Eur. Ceram. Soc. 36 (2016) 515 – 525.
[118] V. Westphal, W. Kleemann, M. D. Glinchuk, Diffuse phase transitions and
random-field-induced domain states of the ‗‗relaxor‘‘ ferroelectric PbMg1/3Nb2/3O3, Phys
Rev. Lett. 68 (1992) 847 – 850.
[119] S. Subrahmanyam, E. Goo, Diffuse phase transitions in the (PbxBa1-x)TiO3
system, J. Mater. Sci. 33 (1998) 4085 – 4088.
[120] I. Grinberg, A.M. Rappe, Local structure and macroscopic properties in
PbMg1∕3Nb2∕3O3−PbTiO3 and PbZn1∕3Nb2∕3O3−PbTiO3 solid solutions, Phys. Rev. B
70 (2004) 220101(R).
[121] I. Grinberg, P. Juha s, P.K. Davies, A.M. Rappe, Relationship between Local
Structure and Relaxor Behavior in Perovskite Oxides, Phys. Rev. Lett. 99 (2007)
267603.
[122] H. Tan, H. Takenaka, C. Xu, W. Duan, I. Grinberg, A.M. Rappe, First-
principles studies of the local structure and relaxor behavior of
Pb(Mg1/3Nb2/3)O3−PbTiO3-derived ferroelectric perovskite solid solutions, Phys.
Rev. B 97 (2018) 174101(R).
[123] R. Blinc, J. Dolinsek, A. Gregorovic, B. Zalar, C. Filipic, Z. Kutnjak, A. Levstik,
R. Pirc, Local Polarization Distribution and Edwards-Anderson Order Parameter of
Relaxor Ferroelectrics, Phys. Rev. Lett. 83 (1999) 424 – 427.
[124] R. Pirc, R. Blinc, Spherical random-bond–random-field model of relaxor
ferroelectrics, Phys. Rev. B 60 (1999) 13 470.

33
Graphical abstract

34

You might also like