Failure Mode, Ferroelastic Behavior and Toughening Effect of Bismuth Titanate Ferroelectric Ceramics Under Uniaxial Compression Load

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Materials and Design 152 (2018) 54–64

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

Failure mode, ferroelastic behavior and toughening effect of bismuth


titanate ferroelectric ceramics under uniaxial compression load
Yu Chen a,b, Jiageng Xu c, Shaoxiong Xie d, Rui Nie b, Jing Yuan b, Qingyuan Wang a,d,e,⁎, Jianguo Zhu b,⁎⁎
a
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China
b
College of Materials Science and Engineering, Sichuan University, Chengdu 610065, China
c
School of Architecture and Civil Engineering, Chengdu University, Chengdu 610106, China
d
College of Architecture and Environment, Sichuan University, Chengdu 610065, China
e
Failure Mechanics and Engineering Disaster Prevention and Mitigation Key Laboratory of Sichuan Province, Chengdu 610065, China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Two types of microcracks (opening and


closed) initiated in the matrix of
Bi4Ti3O12 ceramic during mechanical
loading.
• The underlying micromechanism of
ferroelastic domain switching was de-
scribed by the non-180o domain wall
motion.
• The constitutive relation of ceramics
under uniaxial compression load was
obtained by a domain distribution
model.
• The expression of ferroelasticity in-
duced toughening effect (ΔГ/Г) was de-
duced from the indentation fracture
mechanics.

a r t i c l e i n f o a b s t r a c t

Article history: A typic BLSF ceramics: bismuth titanate (Bi4Ti3O12), whose failure mode, ferroelastic behavior and toughening effect
Received 20 January 2018 under uniaxial compression load were comprehensively assessed. Firstly, a series of crystallographic parameters in-
Received in revised form 9 April 2018 cluding lattice constants (a, b, c and v), orthorhombicity (g) and single-crystal distortion (S0lattice) were calculated
Accepted 19 April 2018
from XRD patterns. The ferroelastic behavior referred to the 90° switching of domains with compression stress
Available online 24 April 2018
was identified by the nonlinear stress-strain curve, and the underlying micromechanism was more accurately de-
Keywords:
scribed by the non-180° domain wall motion. SEM observation on the fracture surface of the broken sample reveals
Bi4Ti3O12 ceramics its failure mode in terms of microcrack initiation and propagation behaviors. Further, a simplified domain distribu-
Ferroelastic domain switching tion model was constructed for unpoled ferroelectric ceramics based on a constitutive framework, which deduced
Microcracks their constitutive relation subjected to uniaxial compression load. The ferroelasticity induced toughening was ver-
Compression load ified by indentation test, and further analyzed by a micromechanics model of crack propagation. The expression of
Toughening effect toughening effect (ΔГ/Г) was also deduced from the indentation fracture mechanics. Finally, the evolution in me-
chanical properties (including coercive stress (σc), switching strain (εzz), apparent elastic modulus (E) and compres-
sion strength (σcf)) of Bi4Ti3O12 ceramics with the doping content of W was also quantified by the compression test.
© 2018 Elsevier Ltd. All rights reserved.

⁎ Correspondence to: Q. Wang, School of Mechanical Engineering, Chengdu University, Chengdu 610106, China.
⁎⁎ Corresponding author.
E-mail addresses: wangqy@scu.edu.cn, (Q. Wang), nic0400@scu.edu.cn (J. Zhu).

https://doi.org/10.1016/j.matdes.2018.04.055
0264-1275/© 2018 Elsevier Ltd. All rights reserved.
Y. Chen et al. / Materials and Design 152 (2018) 54–64 55

1. Introduction 2. Experimental procedure

Ferroelectric materials are one important category of functional mate- 2.1. Preparation of ceramics
rials with the piezoelectric effect which enables the conversion between
mechanical energy and electrical energy, which have been commonly W/Cr co-doped Bi4Ti3O12 ceramics with a chemical composition of
used as sensitive element for many electric devices including piezoelectric Bi4Ti3−xWxO12+x + 0.2 wt% Cr2O3 (x = 0, 0.025, 0.05, 0.075 and 0.1, ab-
sensors, electromechanical transducer, non-volatile memories, etc. [1]. It breviated them as BITWx-0.2Cr) were fabricated by a traditional solid
is the essential attribute of ferroelectrics that the domain state can be reaction process, according to the process technology described in our
modified by electrical or mechanical loadings of sufficient magnitude, previous works [24,25]. In addition, pure Bi4Ti3O12 (BIT) ceramics
which gives rise to the reorientation of polarization/strain along external were also fabricated by the same process in this work. All these BIT-
loadings. Specifically, domains can be switched 180° or 90° from the orig- based ceramics were made into cylinders with a size of φ 10 mm × h
inal direction by an electric field, which is called “ferroelectricity (or ferro- 10 mm, as well as fine polished before the mechanical testing.
electric behavior)”, and domains can be switched only 90° by a
mechanical stress, which is so-called “ferroelasticity (or ferroelastic be- 2.2. Characterization of ceramics
havior)” [2]. In most piezoelectric sensors and actuators, ferroelectric ma-
terials are prone to fatigue due to cyclic electrical or mechanical loadings. 2.2.1. Microstructural characterization
The fatigue manifests its effect as a reduction in domain switching or do- The phase structure of ceramics was characterized by a powder X-
main wall movement and subsequent premature failure of devices [3,4]. ray diffractometer (XRD; DX2700, Dandong, China) employing Cu-Kα
Recent research reported a discussion of switching kinetics for ferroelec- radiation (λ = 1.5418 Å). The lattice strain and lattice spacing were
trics, demonstrating experimentally that nucleation can be the more im- determined by a single peak fitting approach using a specific code
portant step, rather than domain wall forward-growth or sideways (Oringin 9.1 ware, OriginLab, USA) according to Gaussian functions.
velocities [5]. The microstructural morphology of ceramics was observed by a field-
Ferroelastic domain textures generated by an interchanging of the emission scanning electron microscopy (FE-SEM; JSM-7001F, JEOL,
crystallographic degenerate and preferred axes result in a macroscopic Japan).
strain component [6], while both lattice strain and ferroelastic strain con-
tribute to the large signal macroscopic strain response during an applied 2.2.2. Uniaxial compression test
bipolar electric field [7]. The spontaneous strain generated by domain We focus on the uniaxial compression stress states since ferroelec-
switching can be simulated directly by dislocations, Xie et al. [8] extended tric ceramics can be realized comparably easily for a brittle material
the dislocation simulation method of domain switching toughening to pi- with low tensile strength, thus the uniaxial compression test was per-
ezoelectric coupling field. 90° ferroelectric/ferroelastic domain reorienta- formed on a universal testing machine (8501, INSTRON, UK). A com-
tion was observed in a cantilever comprised of a 500 nm thick lead pression preload of 5 MPa was applied for the specimen fixation. The
zirconate titanate (Pb(Zr, Ti)O3, PZT) film on a 3 μm thick elastic layer loading mode was in the displacement control at a stroke rate of
composite of SiO2 and Si3N4 [9]. For both electrical and mechanical 1 mm/min. The macroscopic strain of specimen was measured using
switching, ferroelastic switching is found to occur most readily at the an extensometer with a gauge length of 5 mm mounted to its middle,
highly active needle points in ferroelastic domains [10]. The dynamic while the applied load of machine was automatically recorded by an in-
properties of domain reorientation under mechanical cyclic loading in ternal force sensor. The failure stress referring to the compression fail-
bulk PZT ceramics are studied by Pojprapai et al. [11]. A distinctive ure of materials is defined as their compression strength (σcf), which
ferroelastic creep was observed at 293 K during the mechanical deforma- is calculated according to the following equation:
tion of lanthanum strontium cobalt ferrite under uniaxial compression
[12]. Comparison of neutron diffraction and Raman spectroscopic studies P
σ cf ¼ ð1Þ
of the ferroelastic behavior of ceria-stabilized zirconia at elevated temper- S
atures was conducted by Bolon et al. [13]. On the other hand, the reorien-
where P is the crushing load (N) and S is the cross-sectional area of the
tation of ferroelastic domains around the crack tip is also proposed as a
specimen (mm2).
toughening mechanism for ferroelectric ceramics [14–17]. Moreover,
the domain switching of ferroelectric ceramics is also found to be related
2.2.3. Vickers indentation test
with the crystalline structure of materials. Li et al. [18] used a combined
Because only small samples could be obtained for ferroelectric ce-
theoretical and experimental approach to establish a relation between
ramics using this fabrication method, the fracture toughness was esti-
crystallographic symmetry and the ability of a ferroelectric polycrystalline
mated by microindentation techniques using a Vickers indenter
ceramics to switch. Fu et al. [19] investigated the domain switching and
(AKASHI AVK-A, Kanagawa, Japan). The fracture toughness with the
lattice strains in (Ka, Na)NbO3-based lead-free ceramics across
form of the work of fracture (Г), which is related to the energy release
orthorhombic-tetragonal phase boundary.
rate, is preferred over the fracture toughness KIC (=ГE)1/2 because it is
Bismuth titanate, Bi4Ti3O12 (BIT), was identified as a typical bismuth
the more relevant parameter and independent of the elastic modulus
layer-structured ferroelectrics (BLSFs or Aurivillius phase) with high
(E), and Vickers hardness (Hv) could be calculated by two equations as
Curie temperature (Tc = 675 °C) and large spontaneous polarization
follows,
(Ps (a-axis) = 49 μC/cm2) [20]. Many works have been devoted to the
research on the structural, optical, dielectric and electrical properties 2
of BIT [21–23]. However, many important mechanical problems 2 d
Γ ¼ 2ξ P ð2Þ
seem to be less discussed for BLSFs. In our previous works, W/Cr co- c3
doped Bi4Ti3O12 ceramics have been developed to achieve both good
P
electrical and mechanical properties, aiming at the high-temperature H v ¼ 0:464 2
ð3Þ
d
piezoelectric application [24–27]. In this work, some mechanical
problems with respected to ferroelastic domain switching were where P is the indentation load, d is the half-length diagonal of the
elaborately investigated for this kind of ceramics under uniaxial com- Vickers indentation, c is the half-length of the penny-shaped crack and
pression load, with laying stress on both underlying mechanism and ξ (0.016) is the non-dimensional coefficient. In this experiment, the in-
theoretical models of the microcrack propagation and the ferroelasticity dentation load of 19.6 N was applied and the holding time was 15 s.
toughening. Cracks resulting from the Vickers indentation were measured by SEM
56 Y. Chen et al. / Materials and Design 152 (2018) 54–64

immediately to avoid slow crack growth after printing, started by the crystallizes into an orthorhombic system below the Curie temperature,
stress field that acts upon loading. The crack extent was measured with two orthogonal single-crystal distortion directions relative to the
with an image analysis software (Leica Application Suite EZ; Leica prototype tetragonal phase, and [0 1 0] is defined as [h k l⁎] [6]. S0lattice
Microsystems Ltd, Germany). of BIT could be thus calculated by the following formula (a N b),

3. Results and discussions a−b


S0lattice ¼ ð5Þ
b
3.1. Crystallographic structures of BIT-based ceramics
According to the Rietveld 0refinement for the diffraction pattern, the
A room-temperature XRD pattern, associated with a cross-section cell parameters of Bi4Ti3O12 ceramic calculated are basically in agree-
SEM image was derived from the BIT ceramic, which are shown in ment with the values determined by Ivanov et al. [31]. All these crystal-
Fig. 1 (the data of BITWx-0.2Cr ceramics are similar to this and have lographic parameters of BIT-based ceramics are listed in Table 1. As can
been shown in [24]). In XRD pattern, all sets of reflections could be be seen, a remarkable decrease of g (6.82 → 4.99) reveals a less ortho-
well indexed in the parent compound of Bi4Ti3O12 (JCPDS card No. 72- rhombic distortion in BIT after the incorporation of Cr2O3 (x = 0),
1019) with an pseudo-orthorhombic structure and a space group of which may be due to the released rotation state of oxygen octahedra
B2cb (41), and there is no obvious second phase existing in the matrix. by the substitution of Cr3+ for Ti4+. Except for an ephemeral increase
As observed from the SEM image that, a dense microstructure con- (4.99 → 5.53) of g in the introduction of less W6+ (x = 0.025), this factor
structed by many randomly orientated plate-like grains has been devel- seems to decrease with increasing the doping content of W6+ (x N
oped in the BIT ceramic, even, a typic layer structure can be seen from 0.025) linearly, which may be due to the gradually released tilting
the fractured surface of several coarse grains indistinctly. Furthermore, state of oxygen octahedra by the substitution of W6+ for Ti4+. In addi-
the (1 1 7) plane shows the highest diffracted intensity, which agrees tion, it is also found that after introducing Cr2O3 into BIT, both (2 2
with the fact that the randomly orientated BLSFs usually reflect the 0) and (1 1 15) reflection shift to a lower 2θ angle, which could be relat-
strongest diffraction in their (1 1 2m + 1) crystallographic planes ed to the increase of lattice parameters according to the Bragg equation
[28–30], which are closely related to their layer structure along the c- and the formula about interplanar spacing. Especially, (1 1 15) reflection
axis. In addition, many reflections with crystallographic index of (h k is observed to disappear at x = 0.1, which may be caused by the struc-
l)/(k h l) such as (2 0 0)/(0 2 0) are observed to split into two peaks, ture anomaly (possibly the second phase like Bi6Ti3WO18 or Bi6Ti5WO22
which is due to the typical orthorhombic structure with a slight differ- [32]) of BITW0.1-0.2Cr with a high W content. Based on the analysis
ence between the lattice constants of a and b. Here, orthorhombicity above, for the ion-doped BIT system, the structure with a higher
(g) is used for evaluating the orthorhombic degree of the structure, orthorhombicity (g) tends to contain a larger single-crystal distortion
which could be calculated by the following formula, (S0lattice), which could be correlated with the larger displacements of
cationic polarization within a more distorted structure.
2ða−bÞ
g¼ ð4Þ
aþb 3.2. Stress-strain response of the unpoled BIT ceramic

On the other hand, for ferroelectrics, domain switching can yield a Fig. 2 shows the stress-strain response of the unpoled BIT ceramic
maximum degree of orientation that corresponds to the maximum subjected to uniaxial compression load. It can be seen from
number of possible directions of ferroelastic structural distortions rela- Fig. 2(a) that, the total stress-strain curve includes three segments
tive to a prototype unit cell. The single-crystal distortion (S0lattice) is de- with different slope, which indicates a typical nonlinear deformation
fined as the difference of the ferroelastic unique axis (herein referred to process of the ceramic. Specifically, when a sufficiently small loading
as [h k l⁎]) relative to the crystallographic degenerate axis. BIT was exerted on the ceramic from the starting point O, lattice ions will
be moved only slightly from their equilibrium positions and the ceramic
begin to elastically deform (Eini = 44 GPa) in this process (OA stage).
After the load stress exceeds the minimum coercive stress (σc, min =
45 MPa) of ferroelastic domains at the point A, their switching process
will be initiated, leading the ceramic into a ferroelastic deformation pro-
cess, where certain ions reach to new equilibrium positions such that
the longer c-axis of domains orientates to a direction perpendicular to
the compression load (90° switching). In this process (AB stage), both
the switching strain (εzz) of those reorientated ferroelastic domains
due to the interchanging of ferroelastic axis and the lattice strain (εlat)
of several reconstituted lattice planes due to the d-spacing change of
bismuth layer structure, as well as a slight elastic strain (εela) resulted
from those unmoved domains whose longer c-axis are originally per-
pendicular to the compression stress, contribute to the macroscopic
total strain of ceramic (Δεtol = εmax − εmin = εela + εzz + εlat =
0.60%). Both domain switching and lattice reconstitution are inelastic
and irreversible, their summations constitute the main part of remanent
strain (εrem = 0.65%) which was evaluated by the intercept of the load-
ing curve of BC. Here, the remanent strain represents the macroscopic
average of the spontaneous strains of unit cells when the domain struc-
ture is basically preserved (after backswitching) after unloading, be-
cause the loading curve of BC derived from the failure stress to the
switching-performed stress, thus the remanent strain evaluated by
Fig. 1. XRD patterns of the Bi4Ti3O12 ceramic registered at room temperature (the inserted
SEM image describes the microstructure of its polished and hot-etched cross section, the
this curve approximates to its actual value which should be determined
gray vertical lines at the bottom of the figure represent the peak positions and relative by the unloading curve [12]. For ferroelectric ceramics, the driving force
intensities derived from the JCPDS card.) for domain switching is the minimization of the elastic free energy
Y. Chen et al. / Materials and Design 152 (2018) 54–64 57

Table 1
Crystallographic parameters of BIT-based ceramics.

BIT BITW0-0.2Cr BITW0.025-0.2Cr BITW0.05-0.2Cr BITW0.075-02Cr BITW0.1-0.2Cr

Crystal system Orthorhombic


Space group B2cb (41)
a (Å) 5.4457 5.4699 5.4578 5.4485 5.4464 5.4567
b (Å) 5.4087 5.4427 5.4277 5.4199 5.4215 5.4400
c (Å) 32.8378 32.9354 32.9155 32.8742 32.8320 32.9274
V (Å) 967 982 975 971 969 979
g (×10−3) 6.82 4.99 5.53 5.26 4.58 3.07
S0lattice (×10−3) 6.84 5.00 5.55 5.28 4.59 3.07

Note: Cell parameters of BITWx-0.2Cr ceramics were derived from our previous works [24].

within the frontal zone [33]. Because the volume change of unit cell dur- temperature due to the release of grain microstresses. These thermal
ing domain switching is zero thus only shear stresses will initiate the microcracks seem to remain closed in the process of loading since per-
domain switching. As a result, if the matrix is not able to accommodate pendicular to the applied stress. And the large opening microcracks,
the twinning shear induced by the ferroelastic domain switching, many such as the two indicated by the blue arrow, may be initiated at some
holes may form at the blocked domain walls and then develop into pore points of high tensile stress or evolved from those blocked
microcracks. ferroelastic domains in switching during the mechanical loadings. As
On the other hand, during ferroelastic domain switching, the appar- parallel to the applied stress, these mechanical microcracks tended to
ent elastic modulus of materials tends to be gradually improved by the open, some of them further developed into macroscopic flaws, as exhib-
compressive stress due to the anisotropic elastic properties of noncubic ited in the inserted figure which shows the appearances of the broken
unit cells such as Bi4Ti3O12 [34]. After most of possible (or achievable) sample. Here: thermal microcracks closure, with consequent stiffening
domains were switched at point B (σc, max = 73 MPa), the ceramic of the body, and mechanical microcracks opening, with damage intro-
seems to react more elastically (Eld = 70 GPa) to the applied stress duced to the material. Those two processes are considered to act in com-
until to the compression failure at point C (σcf = 281 MPa; εtol = petition at lower loads, which may be responsible for the nonlinear step
1.06%), since the displacement linearly increases with the load again. (as marked by the green circle) at point B in Fig. 2(a).
In this process (BC stage), with increasing the load, microcracks start Further, single peak fitting is an additional method by which to ex-
closing and the apparent elastic modulus increases (E = 44 GPa tract the lattice strain, however, its actual value should be given by an
→ 70 GPa); if further increasing the load, damage takes over, which average of all lattice strains parallel to the compressive stress, weighted
leads the ceramic to cracking finally. The underlying micromechanism by their relative presence [36]. For Bi4Ti3O12 with the ferroelastic unique
can be further stated as follows [35]: since the cracks perpendicular to axis of [0 1 0], the interplanar spacing change of (0 2 0) plane could be
the applied stress tend to close, with consequent stiffening of the considered as an estimate of bulk-averaged lattice strain of materials
body, while those parallel to the stress tend to open, with damage intro- approximately parallel to the compression axis, thus the subtraction of
duced to the material. Additionally, two decoupled strain-stress curves, the lattice strain from the total irreversible strain can yield the
which could reflect the elastic deformation resulted from the reversible ferroelastic switching strain for the ferroelectric ceramics subjected to
ionic balance exercise (black line) and the inelastic deformation caused the compressive stress. For the BIT ceramic after compression at σc,
by the ferroelastic domain switching and lattice plane reconstitution max, the lattice strain can be evaluated as 0.44% (see Section 3.3), and
(blue one), respectively, are depicted with the total deformation curve the total irreversible strain (0.55%) was calculated by the macroscopic
(red one) in Fig. 2(b). total strain (0.60%) subtracting the elastic strain ((σc, max − σc, min) /
In order to identify the failure mode of ceramic, SEM observation Eave = 0.05%), thus the ferroelastic switching strain was determined
was conducted by focusing on the fracture surface of the broken sample. as 0.11% finally. According to the volume-weighted model proposed in
In Fig. 3, the ceramic behaves in the mode of transgranular fracture and [6], the calculated value of switching strain approximates 3/4 of its sat-
a direct evidence of microcrack creation can be identified by those clear- urated value (εzz = 0.212 · S0lattice = 0.15%) for the BIT ceramic. In fact,
ly visible microcracks. The very thin crack, such as the one indicated by the realistically or attainable achievable domain switching textures are
the red arrow, may have nucleated upon cooling from the Curie a fraction of this saturation texture because of the adverse

Fig. 2. Stress-strain response of the unpoled Bi4Ti3O12 ceramic subjected to uniaxial compression load, (a) total stress-strain curve and (b) decoupled strain-stress curves.
58 Y. Chen et al. / Materials and Design 152 (2018) 54–64

a result, in the cross section parallel to the compression stress, succes-


sional 90° switching of ferroelastic domains is practically in promoting
the gradual shift of 90° domain walls between c-axis domains and a/b-
axis ones situated at this plane. The corresponding bulk contributions
to the driving force are caused by variations of the ferroelectric/
ferroelastic anisotropy energy [40]. Ending at point B, the domain
state induced by strong uniaxial compression stresses is transversely
isotropic without macroscopic polarization, those domains switched
left an irreversible deformation to the material, since the inelastic strain
reaches to the maximum value.
The switching process of ferroelastic domain yields a maximum de-
gree of orientation that corresponds to the maximum number of possi-
ble directions of ferroelastic structural distortions relative to a prototype
unit cell. In order to estimate the fraction of ferroelastic domains
switched, using Kamlah-Jiang's constitutive framework for reference
[41], a simplified domain distribution model was built for unpoled fer-
roelectric ceramics in Fig. 5.
Here, the direction of uniaxial compression load is assumed to coin-
cide with the z-axis, a region within a grain of the ferroelectric polycrys-
tal where the c-axis of the unit cells (spontaneous strain axis) have the
Fig. 3. SEM image focusing on the fracture surface of the Bi4Ti3O12 ceramic broken by the
compression load.
same orientation to the loading direction is called a c-axis domain, all
the domains within the cones of 45° angle with the z-axis being the
cone axis are considered to be c-axis domain for simplicity. The unpoled
ferroelectric ceramics are in the thermally depoled reference state, by
microstructural interactions in real ceramics (such as interactions be- the cutoff of the spherical surface by these cones, the fraction of c-axis
tween neighboring grains and domain-wall pinning due to point de- domains is assumed to be 1/3, if we consider that the state of relative
fects, etc.), as well as the number of possible ferroelastic structural polarization of domains will have no influence on the remanent distor-
distortion directions relative to the prototype phase. tion of the lattice, and there is a simple linear relation between the accu-
mulation of irreversible strain with the fraction of switched domains,
3.3. Ferroelastic behavior and constitutive equation of unpoled ferroelectric the constitutive equation of an unpoled ferroelectric ceramic subjected
ceramics to a uniaxial compression load could be expressed as follows,

For most of perovskite-type ferroelectrics, when they are cooled


σ
through the paraelectric-ferroelectric phase transition temperature, S¼ þ Ssat ð1−3βÞ ð6Þ
E
the combination of electric and elastic boundary conditions loading on
the crystal usually leads to a complex domain structure with many
90° walls (to minimize the elastic energy) and 180° walls (to reduce here, β represents the fraction of c-axis domains (1/3 ≥ β ≥ 0), which is
the depolarizing electric fields). 90° walls are both ferroelectric and decreasing with increasing load, Ssat is the maximum value of the mac-
ferroelastic as they differ both in orientation of the spontaneous polari- roscopic remanent strain of the ceramic which is assumed for a domain
zation vector (Ps) and the spontaneous strain tensor (xs). In a single state of highest order with respect to a certain axis (i.e. β = 0), S is the
crystal of Bi4Ti3O12, there has been reported that the ferroelectric do- total strain, σ is the applied load and E is the elastic modulus. Now, we
main switching with electric fields along the c axis involves a large considered a point close to the failure point C where most of possible
change in the extinction angle with illumination along b [37], and (or achievable) domains have been switched by the stress, after
these ferroelastic domain walls are established after the rearrangement substituting the corresponding parameters derived from the stress-
of non-ferroelastic 180° domains accompanied by Ps(c) [38]. As a result, strain curve into the Eq. (6), the fraction of c-axis domains (β) which
for the polycrystalline ceramics of Bi4Ti3O12, their ferroelastic domain still stay in the cones of 45° angle is 4.66%, thus the corresponding
switching are more accurately described by non-180° (mainly 90°) do- switching fraction of c-axis domains is 86.01% (=1 − 3β), as for
main wall motion, whose micromechanism could be described by Fig. 4. whole domains, the switching fraction is 28.67% (=1/3 − β), which in-
Here, in order to simplify the problem of ferroelastic domain dicates that nearly all of c-axis domains switched to the direction per-
switching in polycrystalline Bi4Ti3O12 ceramics, as well as in view of pendicular to the compression stress.
the comparability of tetragonal phase with orthorhombic phase in the On the other hand, the ferroelastic (or 90°) domain switching caused
interchange between long axis (c) and short axis (a/b), we assumed by the mechanical loading has been determined previously by quantify-
the orthorhombic Bi4Ti3O12 as the tetragonal phase in the later discus- ing the intensity changes of specific peaks in a diffraction pattern, e.g.
sion and analysis with respect to its ferroelastic behaviors. The fresh- the “c”(0 0 l) and “a” (h 0 0) peaks in tetragonal perovskite ceramics
sintered ferroelectric ceramics could be considered as the thermally [42,43]. The orthorhombic Bi4Ti3O12 is also of perovskite units, Fig. 6
depoled piezoceramics with randomly orientated ferroelectric/ shows XRD patterns of the Bi4Ti3O12 ceramic before and after compres-
ferroelastic domains (i.e. initially isotropic), which could be described sion at σc, max (73 MPa). As can be seen from the enlarged drawing on
by the state at point O. Starting at point A, a mechanical stress cannot the left side, after the ceramic was compressed, the intensity of the
trigger a unique switching direction for the spontaneous strain of a peak corresponding to (0 0 l) planes decreases, indicating the grains
unit cell like that, all the c-axis will prefer a position close to a plane per- with c-axis oriented along with the pressing direction have been
pendicular to the compression stress, the distribution of these c-axis do- switched by the compress stress and added to the low degree of (0 0
mains within this plane will be random. It is worthy to noting that l) favored orientation observed [44]. Further, to obtain a more quantita-
domain orientations with the c-axis in the range 0° b α b 30° from the tive analysis of the switching behavior and domain texture of BIT, the in-
compression axis are easier to switch than those with the c-axis at an tensities measured in the diffraction experiment were used to calculate
angle of 30° b α b 60° [39], and the domains with a smaller coercive a density value which describes domain preferences in the unit multiple
stress are earlier to switch than those with a larger coercive stress. As of a random distribution (MRD) and which is expressed through the
Y. Chen et al. / Materials and Design 152 (2018) 54–64 59

Fig. 4. Micromechanism of the ferroelastic behavior of unpoled ferroelectric ceramics.

following equation [45], shown in the enlarged drawing on the right side. According to their rel-
  ative intensity variation, MRD of the BIT ceramic can be identified as
I 0012 =IR0012 0.4033, which is closed to that of unpoled PZT ceramics after cyclic com-
MRD ¼     ð7Þ pression at 140 MPa/1 Hz/100 times [46]. In addition, the interplanar
I0012 =IR0012 þ 2  I 200 =IR200 spacing of (0 2 0) planes has a decrease after compression, indicating
that the compression stress applied along the [0 1 0] direction of some
where I0012 and I200 are, respectively, the integrated intensities of the (0 unit cells has caused lattice strain for the structure. Here, the diffraction
0 12) and (2 0 0) peaks after poling or mechanical loading. IR0012 and data obtained from the X-ray diffractometer were used to ascertain lat-
IR200 are the intensities obtained from the sample with random domain tice strain, which is calculated from the difference between the lattice
texture (i.e. before compression). The evolution in the diffraction pat- spacing before and after loading, expressed as [11]:
tern of (0 0 12)/(2 0 0)/(0 2 0) peaks before and after compression is

0
dhkl −dhkl
εhkl ¼ 0
ð8Þ
dhkl

Fig. 5. Simplified domain distribution model of unpoled ferroelectric ceramics based on a Fig. 6. XRD patterns of the Bi4Ti3O12 ceramic before and after compression at σc, max
constitutive framework. (73 MPa).
60 Y. Chen et al. / Materials and Design 152 (2018) 54–64

where εhkl is the lattice strain (denoted by εlattice in this work), and dhkl domain switching is referred to as the process zone. In the contacting
and d0hkl are the lattice spacings before and after loading. Therefore, the process of sharp indenter with the material surface, the maximum ten-
value of lattice strain can be determined as 0.44% for the BIT ceramic sile stress occurs at the elastic/plastic interface, with fall-off within the
after compression. plastic zone to a negative value at the indenter/specimen contact and
within the surrounding elastic region to zero remote from the contact
3.4. Toughening effect resulted from ferroelastic behavior [48], which initiates the nucleation of crack at the vertex angle of inden-
tation. When cracks begin to propagate along the diagonal line of inden-
Under the application of a Vickers indenter, the indentations and tation and grow into the process zone, those domains previously located
cracks produced in the BIT ceramic before and after compression are in the process zone at the crack tip are now located in the crack wake. As
shown in Fig. 7. SEM observation reveals that a symmetric rhombic in- the polar rotation of ferroelastic domains switched by 90° has caused a
dentation has been introduced into the ceramic, and the resulting cracks strain mismatch of up to 1% (0.12% for BIT in this case) there, compres-
tend to propagate along the diagonal direction of indentation. However, sion stresses will be created perpendicular to the crack plane because of
most of cracks seem to be shorter in the ceramic after compression, and the preferred orientation of the longer axis of ferroelastic domains in
some of them have even disappeared. According to Eq. (2), the fracture this direction. When the crack extends, these compression stresses
toughness (Г) calculated from the length of cracks and indentations was tend to act in the crack wake and cause shielding of the crack tip from
determined as 32.21 J/m2 and 36.68 J/m2 for the BIT ceramic before and the applied load as in the case of the second phase toughening the ma-
after compression, respectively, which presents an obvious toughening trix [49]. In addition to compression stresses, tensile stresses are present
effect (ΔГ = 4.47 J/m2). On the other hand, the Vickers hardness (Hv) parallel to the crack plane. However, the tensile stresses parallel to the
was found to be 3.81 GPa and 3.49 GPa for the BIT ceramic before and crack plane are, in a first approximation, not of importance for the stress
after compression, respectively. It is well known that the hardness rep- intensity factor at the crack tip. Moreover, TEM observation from C.
resents the ability of materials to resist the plastic deformation. The pro- Mercer et al. [50]. has presented compelling evidence for the irrevers-
cess zone of ferroelastic domain switching is more suitable for the ible deformation of ferroelastic domains in a process zone on either
movement of lattice dislocations due to the residual strain existing side of a crack formed by indentation. The ferroelasticity induced tough-
there, which tends to decrease the hardness of the ceramic. Moreover, ening is supposed to have a quadratic relationship with the volume per-
after compression, the crack extension seems to be more anisotropic cent of T′ phase within YSZ ceramics, which is also identified by the
in view of their uneven length, the anisotropy of the indentation crack Vickers indentation test [51].
length and corresponding apparent fracture toughness could be related Fracture toughness, defined as the resistance to fast crack propaga-
with the interaction of domain switching and residual strain, which has tion at a critical stress level (KIC), is widely used for mechanical charac-
been observed in perovskite membrane materials [47]. terization of brittle ceramics. In view of the absorption of fracture
Here, this toughening effect is derived from the ferroelastic domain energy during ferroelastic domain switching, the crack tip process
switching, whose micromechanism can be illustrated by Fig. 8. As de- zone is conducive to resisting the crack extension, resulting in an elevat-
scribed by this micromechanics model for the propagation of indenta- ed fracture toughness for materials, thus the elevation of the steady-
tion cracks within ferroelectric ceramics. When ferroelectric ceramics state toughness Г (J/m2) would scale as [51]:
were compressed by a loading stress exceeding their coercive stress,
ferroelastic domain (90°) switching has occurred (domains only orien- ΔΓ ¼ 2  f  σ c  εc  h ð9Þ
tated to the direction perpendicular to the compression load, which
seems like those colorized bloc), the frontal region involved with in which σc is the coercive stress related to switching of the orientation

Fig. 7. SEM images of indentations and cracks produced in the Bi4Ti3O12 ceramic by a Vickers indenter. (a), (b) and (c) were derived from the upper surface of the sample before
compression; (d), (e) and (f) were derived from the lower surface of the sample after compression (73 MPa). Three testpoints were used for the average measurements.
Y. Chen et al. / Materials and Design 152 (2018) 54–64 61

Fig. 8. Micromechanics model for the propagation of indentation cracks within ferroelectric ceramics.

of ferroelastic domains, εc is the associated switching strain, σc.εc is the 3.5. Mechanical properties of unpoled BITWx-0.2Cr ceramics
dissipated energy per unit volume associated with ferroelastic toughen-
ing, h is the width of the processing zone (it is assumed to remain con- Fig. 9 shows the stress-strain response of unpoled BITWx-0.2Cr ce-
stant in front of the crack tip during crack propagation), and f is the ramics subjected to uniaxial compression load. Clearly, the compression
volume fraction of domains that switch within the zone. Here, Mehta behavior of BITWx-0.2Cr ceramics are similar to that of BIT ceramic, also
and Virkar [52] used a similar approach for estimating the size of do- including the inelastic deformation caused by the ferroelastic domain
main switching process zone with that of the crack tip plastic zone as switching. However, some specific difference can be observed from
follows, their stress-strain curves, as indicated by those saffron circles. Such as,
BITW0-0.2Cr has the highest compression strength while BITW0.025-
 2 0.2Cr received the lowest one, BITW0.05-0.2Cr has a smaller ferroelastic
K IC
h ≈ 0:08 ð10Þ deformation while BITW0.075-0.2Cr possesses a larger one. The inserted
σc
image gives out the change in the apparent elastic modulus (Eini and Eld)
of each sample during the initial process and loading process. As can be
And for the brittle materials such as ferroelectric ceramics under the
seen, all the others except for BITW0.025-0.2Cr present an increased elas-
application of a sharp indenter, another fracture toughness with the
tic modulus in the loading process, which is similar to the case of BIT.
form of KIC could be calculated by the following formula based on the
The unexpected increase in the elastic modulus may be attributed to
criterion of fracture mechanics [53],
the very small ratio of microcracks (perpendicular to the applied stress)
  which tend to close during loading. Therefore, the premature failure of
K IC ¼ 0:016ðE=HÞ1=2 P=C 3=2 ð11Þ BITW0.025-0.2Cr is considered to be dominated by the plenty of

Here, P is the load, E is the Elastic modulus, Hv is the hardness, and C


is the length of the crack measured from the centre of indent. According
to the expression of fracture toughness Г from Eq. (2), we can deduce
out the relationship between the two kinds of fracture toughness as fol-
lows,

ΓE
K IC 2 ¼ ð12Þ
0:928

Ultimately, substituting Eqs. (10) and (12) into Eq. (9), we can get
the expression of toughening effect (ΔГ/Г) with respect to the
ferroelastic domain switching as follows,

ΔΓ 1
¼ 0:172  f  εc  E  ð13Þ
Γ σc

for the BIT ceramic, substituting its corresponding parameters (f =


86.01%; σc, average = 59 MPa; εc = 0.11%; Eaverage = 59 GPa) derived
from its stress-strain curve (Fig. 2(a)) into Eq. (13), the toughening ef-
fect ΔГ/Г can be calculated to reach at 16.55%, which is slightly higher Fig. 9. Stress-strain response of unpoled BITWx-0.2Cr ceramics subjected to uniaxial
than its experimental data of 13.88%. compression load.
62 Y. Chen et al. / Materials and Design 152 (2018) 54–64

microcracks (parallel to the applied stress) which tend to open during observed from their stress-strain curves. Finally, the brittle fracture usu-
loading. ally occurs in ferroelectric ceramics during their elastic compression, the
According to the classic Griffith theory [54], the strength of a materi- failure stress is mainly contingent on their elastic modulus, with the ad-
al can be described as follows: dition of partial influence exerted by the nonlinear ferroelastic behavior
[55]. This mechanism could be reflected by the consistent varying trend
sffiffiffiffiffiffiffiffiffiffiffiffiffi between the two parameters when x is below 0.05, and the slight devi-
2Eγ f 0
σ f −σ i ¼ ð14Þ ation when x is above 0.05 (both BITW0.075-0.2Cr and BITW0.1-02Cr
πc have a larger switching strain), as shown in Fig. 10(d). In addition, as
marked by the red rectangular frame, BITW0.05-0.2Cr with a medium
where σf − σi stands for the actual fracture strength (σf is the fracture W content possesses both good microcosmic switching properties
strength without internal stress and σi is the part contributed on frac- (small coercive stress and large switching strain), and good macroscopic
ture strength by internal stress). c is the length of the critical crack, γf0 mechanical properties (high elastic modulus and compression
is the fracture surface energy without internal stress and E is the elastic strength).
modulus. Based on this theory, both the lower elastic modulus due to Fig. 11 gives out both testing fracture toughness (Г) and calculated
few “closed microcracks”, and more internal stress centralized by toughening effect (ΔГ/Г) for BITWx-0.2Cr ceramics with different W
many “opening microcracks”, might contribute to the much lower content (x). As can be seen, BITW0.05-0.2Cr gains the highest fracture
strength of BITW0.025-0.2Cr as compared with others. toughness of 58.21 J/m2, as well as expected to be of a larger toughening
In order to quantify the evolution of mechanical properties for effect (13.54%) than the others in the case of compression. A higher frac-
unpoled BITWx-0.2Cr ceramics, eight mechanical parameters related ture toughness may be contributed by a larger grain size for the quasi-
to the ferroelastic behavior were extracted from their stress-strain brittle fracture of ceramics, according to the prediction of ceramic frac-
curves, which have been depicted as a function of the doping content ture with normal distribution pertinent to grain size [56,57]. Here,
of W (x) in Fig. 10. As can be seen from this complex pattern that, firstly, BITW0.05-0.2Cr indeed has the largest grain size among BITWx-0.2Cr,
the variation of coercive stress (average) with x is just opposite to that of which has been reported in one of our previous works [27]. Moreover,
orthorhombicity (Fig. 10(a)), which shows that ferroelastic domains the reduced piezo-response is also expected relative to the bulk of the
with a higher structural orthorhombicity tend to be switched by a grain, due to the restricted domain-wall motion at grain boundaries.
smaller compression stress. And then, the varying trend of switching Therefore, such ferroelectric ceramics with coarse grains tend to present
strain with x seems to agree with that of single-crystal distortion well a higher piezoelectric ability after poling, which also agrees in the
(Fig. 10(b)), which has been decided by their tie-in equation of εzz = highest d33 ~ 28 pC/N of BITW0.05-0.2Cr ceramics as found by us in [25].
0.212 · S0lattice based on a volume-weighted average of the single-
crystal distortions of each domain over the entire orientation space. Fur- 4. Conclusion
ther, more switched domains certainly contribute to more irreversible
deformation for ferroelectric ceramics, which has been presented by For a typic number of bismuth layer-structured ferroelectric (BLSFs)
the similar trend of switching fraction to remanent strain in Fig. 10(c). ceramics: bismuth titanate (Bi4Ti3O12, BIT), its ferroelastic behavior sub-
Moreover, the remanent strain resulted from partial ferroelastic domain jected to uniaxial compression was detected by the nonlinear variation
switching and several lattice-plane reconstitution contribute about 60% of the stress-strain curve. The switching strain could approximate 3/4 of
to the total strain of these BIT-based ferroelectric ceramic, which can be its saturated value while the switching fraction of c-axis domains was

Fig. 10. Mechanical properties of unpoled BITWx-0.2Cr ceramics as a function of the doping content of W (x), (a) σc vs g; (b) εzz vs S0lattice; (c) f vs εrem; (d) E vs σcf.
Y. Chen et al. / Materials and Design 152 (2018) 54–64 63

[8] C. Xie, Q.H. Fang, Y.W. Liu, et al., Dislocation simulation of domain switching tough-
ening in ferroelectric ceramics, Int. J. Solids Struct. 50 (9) (2013) 1325–1331.
[9] G. Esteves, C.M. Fancher, M. Wallace, et al., In situ X-ray diffraction of lead zirconate
titanate piezoMEMS cantilever during actuation, Mater. Des. 111 (2016) 429–434.
[10] P. Gao, J. Britson, C.T. Nelson, et al., Ferroelastic domain switching dynamics under
electrical and mechanical excitations, Nat. Commun. 5 (2014) 3801.
[11] S. Pojprapai, Z. Luo, B. Clausen, et al., Dynamic processes of domain switching in lead
zirconate titanate under cyclic mechanical loading by in situ neutron diffraction,
Acta Mater. 58 (6) (2010) 1897–1908.
[12] W. Araki, J. Malzbender, Ferroelastic deformation of La0.58Sr0.4Co0.2Fe0.8O3−δ under
uniaxial compressive loading, J. Eur. Ceram. Soc. 33 (4) (2013) 805–812.
[13] A.M. Bolon, T.A. Sisneros, A.B. Schubert, et al., Comparison of neutron diffraction and
Raman spectroscopic studies of the ferroelastic behavior of ceria-stabilized zirconia
at elevated temperatures, J. Eur. Ceram. Soc. 35 (2) (2015) 623–629.
[14] A.V. Virkar, J.F. Jue, P. Smith, et al., The role of ferroelasticity in toughening of brittle
materials, Phase Transit. 35 (1) (2016) 27–46.
[15] S.M. Denkhaus, M. Vögler, N. Novak, et al., Short crack fracture toughness in (1
− x)(Na1/2Bi1/2)TiO3–xBaTiO3 relaxor ferroelectrics, J. Am. Ceram. Soc. 100 (2017).
[16] M. Vögler, J.E. Daniels, K.G. Webber, et al., Absence of toughening behavior in
0.94(Na1/2Bi1/2) TiO3-0.06 BaTiO3 relaxor ceramic, Scr. Mater. 136 (2017) 115–119.
[17] Y. Wang, H. Yang, R. Liu, et al., Ferroelastic domain switching toughening in spark
plasma sintered t′-yttria stabilized zirconia/La2Zr2O7 composite ceramics, Ceram.
Int. 43 (15) (2017) 13020–13024.
[18] J.Y. Li, R.C. Rogan, E. Üstündag, et al., Domain switching in polycrystalline ferroelec-
tric ceramics, Nat. Mater. 4 (10) (2005) 776–781.
Fig. 11. Testing fracture toughness (Г) and calculated toughening effect (ΔГ/Г) of unpoled [19] J. Fu, R. Zuo, Y. Xu, et al., Investigations of domain switching and lattice strains in
BITWx-0.2Cr ceramics with different W content (x). (Na, K) NbO3-based lead-free ceramics across orthorhombic-tetragonal phase
boundary, J. Eur. Ceram. Soc. 37 (3) (2017) 975–983.
[20] T. Jardiel, A.C. Caballero, M. Villegas, Aurivillius ceramics: Bi4Ti3O12-based piezoelec-
28.67%. SEM observation on the fracture surface reveals that it behaved trics, J. Ceram. Soc. Jpn. 116 (1352) (2008) 511–518.
[21] Y.J. Wong, J. Hassan, S.K. Chen, et al., Combined effects of thermal treatment and Er-
in the mode of transgranular fracture and two types of microcracks substitution on phase formation, microstructure, and dielectric responses of
(opening and closed) initiated in the matrix during mechanical loading. Bi4Ti3O12 Aurivillius ceramics, J. Alloys Compd. 723 (2017) 567–579.
The ferroelastic domain switching was also been demonstrated by the [22] A. Bettaibi, R. Jemai, M.A. Wederni, et al., Effect of erbium concentration on the
structural, optical and electrical properties of a Bi4Ti3O12 system, RSC Adv. 7 (36)
diffraction evolution of (0 0 12)/(2 0 0)/(0 2 0) peaks after compression.
(2017) 22578–22586.
The ferroelastic induced toughening effect (ΔГ/Г) reached to 16.55%. [23] F. Rehman, L. Wang, H.B. Jin, et al., Effect of Fe/Ta doping on structural, dielectric,
Among W/Cr co-doped Bi4Ti3O12 (BITWx-0.2Cr) ceramics, BITW0.025- and electrical properties of Bi4Ti2.5Fe0.25Ta0.25O12 ceramics, J. Am. Ceram. Soc. (2)
0.2Cr gained the smallest coercive stress (σc = 56 MPa) and lowest (2017) 602–611.
[24] Y. Chen, Z. Pen, Q. Wang, et al., Crystalline structure, ferroelectric properties, and
compression strength (σcf = 134 MPa) while BITW0.05-0.2Cr possessed electrical conduction characteristics of W/Cr co-doped Bi4Ti3O12 ceramics, J. Alloys
the highest fracture toughness (Г = 58.21 J/m2) and a significant poten- Compd. 612 (2014) 120–125.
tial toughening effect (ΔГ/Г = 13.54%). This research could not only [25] Y. Chen, D. Liang, Q. Wang, et al., Microstructures, dielectric, and piezoelectric prop-
erties of W/Cr co-doped Bi4Ti3O12 ceramics, J. Appl. Phys. 116 (7) (2014), 074108. .
help us to further understand the correlation between macroscopic [26] Y. Chen, C. Miao, S. Xie, et al., Fracture behaviors and ferroelastic deformation in W/
properties and ferroelastic domain switching of BLSFs, but also effec- Cr co-doped Bi4Ti3O12 ceramics, J. Am. Ceram. Soc. 99 (6) (2016) 2103–2109.
tively promote the possible application of BITWx-0.2Cr ceramics in [27] Y. Chen, C. Miao, S. Xie, et al., Microstructural evolutions, elastic properties and me-
chanical behaviors of W/Cr Co-doped Bi4Ti3O12 ceramics, Mater. Des. 90 (2016)
high-temperature piezoelectric devices. 628–634.
[28] I.A. Md, M.A. Gafur, S.M. Islam, Sintering characteristics of La/Nd doped Bi4Ti3O12
bismuth titanate ceramics, Sci. Sinter. 47 (2015) (2015) 175–186.
Data availability
[29] S. Kumar, K.B.R. Varma, Dielectric relaxation in bismuth layer-structured BaBi4Ti4O15
ferroelectric ceramics, Curr. Appl. Phys. 11 (2) (2011) 203–210.
The datasets generated during the current study are available from [30] P. Fang, Z. Xi, W. Long, X. Li, J. Li, Structure and electrical properties of SrBi2Nb2O9-
the corresponding author on reasonable request. based ferroelectric ceramics with lithium and cerium modification, J. Alloys
Compd. 575 (2013) 61–64.
[31] S.A. Ivanov, T. Sarkar, E.A. Fortalnova, et al., Composition dependence of the multi-
Acknowledgement functional properties of Nd-doped Bi4Ti3O12 ceramics, J. Mater. Sci. Mater. Electron.
28 (11) (2017) 7692–7707.
[32] T. Jardiel, M. Villegas, A. Caballero, et al., Solid-state compatibility in the system
This work was supported by the Applied Basic Research Program Bi2O3–TiO2–Bi2WO6, J. Am. Ceram. Soc. 91 (1) (2008) 278–282.
from Sichuan Province (2017JY0091), National Natural Science Founda- [33] G.G. Pisarenko, V.M. Chushko, S.P. Kovalev, Anisotropy of fracture toughness of pie-
zoelectric ceramics, J. Am. Ceram. Soc. 68 (5) (1985) 259–265.
tion of China (Grant No. 11702037, 11572057 and 51332003), China
[34] K.G. Webber, E. Aulbach, T. Key, et al., Temperature-dependent ferroelastic
Postdoctoral Science Foundation Funded Project (2017M623025), Spe- switching of soft lead zirconate titanate, Acta Mater. 57 (15) (2009)
cial Funding for Post Doctoral Research Projects from Sichuan Province 4614–4623.
((2017, presided over by Yu Chen). [35] I. Pozdnyakova, G. Bruno, A.M. Efremov, et al., Stress-dependent elastic properties of
porous microcracked ceramics, Adv. Eng. Mater. 11 (12) (2010) 1023–1029.
[36] M.R. Daymond, The determination of a continuum mechanics equivalent elastic
References strain from the analysis of multiple diffraction peaks, J. Appl. Phys. 96 (8) (2004)
4263–4272.
[1] M. Kobune, T. Kuriyama, R. Furotani, et al., Ferroelectric materials and their applica- [37] S.E. Cummins, L.E. Cross, Electrical and optical properties of ferroelectric Bi4Ti3O12
tions, Jpn. J. Appl. Phys. 54 (10S) (2015). single crystals, J. Appl. Phys. 39 (5) (1968) 2268–2274.
[2] B. Kaltenbacher, P. Krejčí, A thermodynamically consistent phenomenological model for [38] Y. Kitanaka, S. Katayama, Y. Noguchi, et al., Electric-field-stabilized ferroelastic do-
ferroelectric and ferroelastic hysteresis, J. Appl. Math. Mech. 96 (7) (2016) 874–891. main walls in monoclinic Bi4Ti3O12 crystals, Jpn. J. Appl. Phys. 46 (10) (2007)
[3] D. Lupascu, J. Rödel, Fatigue in bulk lead zirconate titanate actuator materials, Adv. 7028–7030.
Eng. Mater. 7 (10) (2005) 882–898. [39] J.L. Jones, M. Hoffman, S.C. Vogel, Ferroelastic domain switching in lead zirconate ti-
[4] N. Chaiyo, D.P. Cann, N. Vittayakorn, Lead-free (Ba, Ca)(Ti, Zr)O3 ceramics within the tanate measured by in situ neutron diffraction, Mech. Mater. 39 (4) (2007)
polymorphic phase region exhibiting large, fatigue-free piezoelectric strains, Mater. 283–290.
Des. 133 (2017) 109–121. [40] H. Kessler, H. Balke, A continuum analysis of the driving force of ferroelectric/
[5] J.F. Scott, A review of ferroelectric switching, Ferroelectrics 503 (1) (2016) 117–132. ferroelastic domain wall motions, J. Mech. Phys. Solids 54 (1) (2006)
[6] J.L. Jones, M. Hoffman, K.J. Bowman, Saturated domain switching textures and 113–127.
strains in ferroelastic ceramics, J. Appl. Phys. 98 (2) (2005), 024115. . [41] M. Kamlah, Q. Jiang, A constitutive model for ferroelectric PZT ceramics under uni-
[7] M.C. Ehmke, N.H. Khansur, J.E. Daniels, et al., Resolving structural contributions to axial loading, Smart Mater. Struct. 8 (4) (1999) 441.
the electric-field-induced strain in lead-free (1 − x)Ba(Zr0.2Ti0.8)O3- [42] J.L. Jones, M. Hoffman, S.C. Vogel, Domain switching under cyclic mechanical loading
x(Ba0.7Ca0.3)TiO3, piezoceramics, Acta Mater. 66 (3) (2014) 340–348. in lead zirconate titanate, J. Am. Ceram. Soc. 89 (11) (2006) 3567–3569.
64 Y. Chen et al. / Materials and Design 152 (2018) 54–64

[43] S. Pojprapai, J.L. Jones, T. Vodenitcharova, et al., Investigation of the domain [51] X. Ren, W. Pan, Mechanical properties of high-temperature-degraded yttria-
switching zone near a crack tip in pre-poled lead zirconate titanate ceramic via in stabilized zirconia, Acta Mater. 69 (2014) 397–406.
situ X-ray diffraction, Scr. Mater. 64 (1) (2011) 1–4. [52] Karun Mehta, Anil V. Virkar, Fracture mechanisms in ferroelectric-ferroelastic lead
[44] A.A. Azlan, W. Krengvirat, A.F.M. Noor, et al., Sintering and characterization of rare zirconate titanate (Zr:Ti = 0.54: 0.46) ceramics, J. Am. Ceram. Soc. 73 (3) (1990)
earth doped bismuth titanate ceramics prepared by soft combustion synthesis, 567–574.
Sintering of Ceramics - New Emerging Techniques, InTech, 2012. [53] G.R. Anstis, P. Chantikul, B.R. Lawn, et al., A critical evaluation of indentation tech-
[45] J.L. Jones, E.B. Slamovich, K.J. Bowman, Domain texture distributions in tetragonal niques for measuring fracture toughness. I.—direct crack measurements, J. Am.
lead zirconate titanate by x-ray and neutron diffraction, J. Appl. Phys. 97 (3) Ceram. Soc. 64 (9) (1981) 533–538.
(2005) 1194. [54] B. Lawn, Fracture of Brittle Solids, Cambridge University Press, 1993.
[46] S. Pojprapai, J.L. Jones, A.J. Studer, et al., Ferroelastic domain switching fatigue in lead [55] H. Cao, A.G. Evans, Nonlinear deformation of ferroelectric ceramics, J. Am. Ceram.
zirconate titanate ceramics, Acta Mater. 56 (7) (2008) 1577–1587. Soc. 76 (4) (1993) 890–896.
[47] B.X. Huang, J. Malzbender, The effect of an oxygen partial pressure gradient on the [56] C. Zhang, X. Hu, T. Sercombe, et al., Prediction of ceramic fracture with normal dis-
mechanical behavior of perovskite membrane materials, J. Eur. Ceram. Soc. 34 (7) tribution pertinent to grain size, Acta Mater. 145 (2018) 41–48.
(2014) 1777–1782. [57] Bibi Malmal Moshtaghioun, et al., Grain size dependence of hardness and fracture
[48] R. Hill, The Mathematical Theory of Plasticity, Clarendon Press, Oxford, 1950. toughness in pure near fully-dense boron carbide ceramics, J. Eur. Ceram. Soc. 36
[49] F. Meschke, A. Kolleck, G.A. Schneider, R-curve behaviour of BaTiO3 due to stress- (7) (2016) 1829–1834.
induced ferroelastic domain switching, J. Eur. Ceram. Soc. 17 (9) (1997) 1143–1149.
[50] C. Mercer, J.R. Williams, D.R. Clarke, et al., On a ferroelastic mechanism governing
the toughness of metastable tetragonal-prime (t′) yttria-stabilized zirconia, Proc.
R. Soc. Lond. A Math. Phys. Eng. Sci. 463 (2081) (2007) 1393–1408.

You might also like