You are on page 1of 41

Fracture behaviors of piezoelectric materials

T.Y. Zhang
*
, C.F. Gao
Department of Mechanical Engineering, Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong
Abstract
Theoretical analyses and experimental observations of the failure and fracture behaviors of piezoelectric materials
are presented. The theoretical analyses are based on the Stroh formalism. A strip dielectric breakdown model is pro-
posed to estimate the eect of electrical non-linearity on the piezoelectric fracture of electrically insulated cracks. The
reviewed experiments include the indentation fracture test, the bending test on smooth samples, the fracture test on pre-
notched or pre-cracked samples, the environment-assisted fracture test, etc. For electrically insulated cracks, the
experimental results show a complicated fracture behavior under combined electrical and mechanical loading. Fracture
data are greatly scattered when a static electric eld is applied. For electrically conducting cracks, the experimental
results demonstrate that static electric elds can fracture poled and depoled lead zirconate titanate (PZT) ceramics. A
charge-free zone model is introduced to understand the failure behavior of conducting cracks in the depoled lead
zirconate titanate ceramics under electrical and/or mechanical loading. These theoretical and experimental results
indicate that fracture mechanics concepts are useful in the study of the failure behaviors of piezoelectric materials.
2003 Elsevier Ltd. All rights reserved.
Keywords: Piezoelectric materials; Crack; Theoretical analysis; Experimental observation
1. Introduction
Piezoelectric ceramics have become preferred
materials for a wide variety of electronic and
mechatronic devices due to their pronounced pie-
zoelectric, dielectric, and pyroelectric properties.
However, piezoelectric ceramics are brittle and
susceptible to cracking at all scales ranging
from electric domains to devices. Various defects,
such as domain walls, grain boundaries, aws and
pores, impurities and inclusions, etc., exist in pie-
zoelectric ceramics. The defects cause geometric,
electric, thermal, and mechanical discontinuities
and thus induce high stress and/or electric eld
concentrations, which may induce crack initiation,
crack growth, partial discharge, and cause dielec-
tric breakdown, fracture and failure. Due to the
importance of the reliability of these devices, there
has been tremendous interest in studying the
fracture and failure behaviors of such materials
[14]. In the study of piezoelectric fractures and
failures, available solutions in purely elastic media
have been extended to the corresponding prob-
lems in piezoelectric materials, for example, crack
problems [58], inclusion problems [9], Greens
function problems [10], problems concerning inter-
actions between dislocation and cracks and/or
*
Corresponding author. Tel.: +852-2358-7192; fax: +852-
2358-1543.
E-mail address: mezhangt@ust.hk (T.Y. Zhang).
0167-8442/$ - see front matter 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.tafmec.2003.11.019
Theoretical and Applied Fracture Mechanics 41 (2004) 339379
www.elsevier.com/locate/tafmec
inclusions in piezoelectric media [11], thermo-
electro-elastic problems [12], dynamic fracture
problems [13], and problems related to failure
criteria [14,15]. However, the fracture behaviors of
piezoelectric ceramics are more complex than that
in conventional materials, because of the non-
linear nature of the mechanical and electrical
behaviors and the complicated coupling relation-
ship between mechanical and electric elds. There
are new challenges in studying the fracture be-
havior of piezoelectric ceramics, e.g., the deter-
mination of the electric boundary conditions on
crack faces [16], the eect of electric elds on the
piezoelectric fracture behavior [17,18], the calcu-
lation of the global and local energy release rates
[1922], etc. More challenging tasks include non-
linear simulations of the fracture behavior of
piezoelectric ceramics because domain switching
causes the non-linearity between the polarization
and the electric eld strength, the non-linearity of
the strain versus the electric eld strength, and
the non-linearity between stress and strain. Theo-
retically, mechanism-based micromechanical ap-
proaches and phenomenological methods are all
adopted to solve the non-linear fracture problems
of piezoelectric ceramics and much eort has been
dedicated to the modeling of the non-linear failure
behavior of ferroelectric and ferroelastic ceramics
[23,24]. At the moment, although there are volu-
minous theoretical studies on the fracture of pie-
zoelectric materials, the number of experimental
studies is still limited. Due to the complexity of the
failure behavior of piezoelectric ceramics, many
failure and/or fracture criteria and mechanisms
have been proposed, each of which can success-
fully explain and predict a failure mode under a
certain circumstance.
The present work provides a review of theo-
retical results and experimental observations. A
strip dielectric breakdown model, similar to the
strip polarization saturation model [25] is pro-
posed and a charge-free zone model is introduced.
Domain switching toughening is also summarized.
Matrix and vector notations in this article are
more consistent with those in the book [26]. The
basic equations are given for a linear piezoelectric
solid in the form of the Stroh formalism. The
solution of an elliptical cylinder cavity is given for
an innite piezoelectric solid under anti-plane
deformations. The eld intensity factors and the
energy release rate are then presented in explicit
forms when the cavity degenerates into a crack.
These results are further extended to the case of a
generalized two-dimensional (2D) deformation in
an innite piezoelectric material with an elliptical
cylinder cavity. Given also is a general method for
solving crack problems, with impermeable crack, a
conducting crack, a permeable crack and a semi-
permeable crack boundary conditions within the
framework of linear electro-elasticity. The strip
dielectric breakdown model is introduced to esti-
mate the eect of the electrical non-linearity on the
piezoelectric fracture. The non-linearity induced
by domain switching and its eect on the fracture
behavior is discussed. Experimental observations
are then described. The charge-free zone model is
introduced to understand the failure behavior of
conducting cracks in dielectric materials under
electrical and/or mechanical loading.
2. Stroh formalism
In a rectangular coordinate system, x
i
i
1; 2; 3, the complete set of basic equations for a
linear piezoelectric solid is [27]
r
ij;j
0; D
i;i
0; 1
c
ij

1
2
u
i;j
u
j;i
; E
i
u
;i
; 2
r
ij
C
ijkl
c
kl
e
kij
E
k
; D
k
e
kij
c
ij
e
kl
E
l
; 3
where u
i
, u, r
ij
, c
ij
, D
j
and E
i
are the displacement,
the electric potential, the stress, the strain, the
electric displacement and the electric eld, res-
pectively, and C
ijkl
, e
ijk
and e
ij
stand for the elas-
tic constants, the piezoelectric constants and the
dielectric constants, respectively. The material
constants have the following symmetries
C
ijkl
C
jikl
C
klij
; e
kij
e
kji
; e
kl
e
lk
:
Moreover, C
ijkl
and e
kl
are positive denite in the
sense that
C
ijkl
u
i;j
u
k;l
> 0; e
kl
E
k
E
l
> 0
for any arbitrary real non-zero u
i;j
and E
k
.
340 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
2.1. Stroh formalismfor standard boundary con-
ditions
For generalized two-dimensional deformations
in which u
i
i 1; 2; 3 and u depend on x
1
and x
2
only, a general solution to Eq. (1) is
u u
1
; u
2
; u
3
; u
T
af x
1
px
2
; 4
where the superscript T represents the transpose,
f x
1
px
2
is an analytic function, p is a complex
number, and a is a four-element column constant
vector, the value of which depends on the material
constants and the sample orientation. Eqs. (1)(3)
can be satised by Eq. (4) for an arbitrary
f x
1
px
2
if
QpR R
T
p
2
Ta 0; 5
where the 4 4 matrices, Q, R and T, are given by
Q
C
i1k1
e
11i
e
T
11i
e
11
_ _
; R
C
i1k2
e
21i
e
T
12i
e
12
_ _
;
T
C
i2k2
e
22i
e
T
22i
e
22
_ _
; i; k 1; 2; 3:
The condition for non-trivial solutions of Eq. (5)
requires
jQpR R
T
p
2
Tj 0: 6
Eq. (6) has eight roots p
a
and p
a
a 1; 2; 3; 4,
where p
a
cannot be real because of the positive
deniteness of the strain energy and electric energy
densities. It is convenient to calculate p
a
by solving
the following standard eigen-equation:
Nf pf; 7
where
N
N
1
N
2
N
3
N
T
1
_ _
; f
a
b
_ _
;
N
1
T
1
R
T
; N
2
T
1
N
T
2
;
N
3
RT
1
R
T
Q N
T
3
; and b R
T
pTa:
Finally, the general solution to Eqs. (1)(3) can be
expressed as
u Afz

Afz; 8
/ Bfz

Bfz; 9
where
A a
1
; a
2
; a
3
; a
4
; B b
1
; b
2
; b
3
; b
4
;
fz
a
f
1
z
1
; f
2
z
2
; f
3
z
3
; f
4
z
4

T
;
z
a
x
1
p
a
x
2
;
and / is the generalized stress function such that
R
2
r
2j
; D
2

T
/
;1
; R
1
r
1j
; D
1

T
/
;2
:
10
The A and B matrices have the following rela-
tionship [26]
B
T
A
T

B
T

A
T
_ _
A

A
B

B
_ _

I 0
0 I
_ _
; 11
where I is a 4 4 unit matrix.
In addition, two matrices, Y and H, used in the
following analysis are dened by
Y iAB
1
and H 2ReY: 12
Matrix Y is a Hermitian matrix and can be parti-
tioned into [27]
Y
Y
e
Y
31
Y
13
Y
44
_ _
; 13
where the upper left block, Y
e
, is a 3 3 matrix,
and Y
44
is a real element. For a stable material, Y
e
is positive denite and Y
44
< 0 [27], which leads to
H
44
< 0.
Dierentiating Eqs. (8) and (9) with respect to
x
1
gives
u
;1
AFz

AFz; 14
/
;1
BFz

BFz; 15
where Fz dfz=dz. Eqs. (14) and (15) can be
re-expressed as
/
;1
Kz Kz; 16
u
;1

1
i
YKz

YKz; 17
where Kz BFz.
If thermal eects on deformations of piezo-
electric materials are considered, the Stroh for-
malism, i.e., Eqs. (8) and (9), can be extended to
[26]
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 341
u 2ReAfz cgz; 18
/ 2ReBfz dgz; 19
where c and d are two complex constant vectors,
and gz is the thermal complex potential deter-
mined by solving the thermal boundary value
problem. Re-writing Eqs. (18) and (19) gives
/ 2Rek
t
z; 20
u 2ImYk
t
z Mgz; 21
where
k
t
z Bfz dgz; M Yd ic: 22
2.2. Stroh formalismfor mixed boundary condi-
tions
The standard Stroh formalism introduced is
appropriate for impermeable or permeable cracks,
where the displacement vector, u, is expressed in
terms of the displacements and electric potential.
In this section, a mixed-form Stroh formalism is
introduced for solving conducting crack problems
in which a hybrid displacement vector, ^u, and a
hybrid stress function vector,
^
/, are constructed as
^u u
1
; u
2
; u
3
; /
4

T
; 23
^
/ /
1
; /
2
; /
3
; u
T
: 24
The hybrid displacement and stress function vec-
tors can be expressed in terms of the original dis-
placement and stress function vectors:
^u I
u
u I
/
/;
^
/ I
/
u I
u
/; 25
where I
u
and I
/
are two diagonal matrices:
I
u
hh1; 1; 1; 0ii; I
/
hh0; 0; 0; 1ii:
Substituting Eqs. (8) and (9) into Eq. (25) leads
to
^u

Afz
a


Afz
a
; 26
^
/

Bfz
a


Bfz
a
; 27
where

A I
u
A I
/
B;

B I
/
A I
u
B:
Eqs. (26) and (27) may be called the Stroh for-
malism for the mixed boundary value problem.
The nature of

A and

B was studied for the case of
purely anisotropic media [28] as well as for the case
of piezoelectric materials [2,29]. Matrixes

A and

B
have a similar feature as that of matrixes A and B,
e.g.

B
T

A
T

B
T

A
T
_ _

A

A

B

B
_ _

I 0
0 I
_ _
:
Similarly, one can dene

Y and

H as

Y i

B
1
;

H 2Re

Y: 28
Matrix

Y is a positive Hermitian matrix, indicat-
ing

Y
44
> 0 and

H
44
> 0, which is dierent from
the features of Y
44
and H
44
.
Similar to (14) and (15), the dierentiation
form,
^
/
;1


Kz

Kz; 29
^u
;1

1
i

Kz

Kz; 30
of Eqs. (26) and (27) with respect to x
1
is conve-
nient in applications where

Kz

BFz. After
^
/
is determined from given mixed-form boundary
conditions, the eld variables can be obtained
from

R
2
r
2j
; E
1

^
/
;1
;

R
1
r
1j
; E
2

T

^
/
;2
:
31
2.3. Stroh formalismfor anti-plane deformations
Solutions for anti-plane deformations are rela-
tively simple and explicit, thereby easily giving
insights into the physical picture of the frac-
ture behavior of piezoelectric ceramics. In this
sub-section, anti-plane deformations are studied
through an approach similar to the Stroh formal-
ism, which may be called the Stroh formalism for
anti-plane deformations. A transversely isotropic
material such as a poled ceramic with the poling
direction along the positive x
3
direction and the
isotropic plane in the x
1
x
2
plane is considered in
this sub-section. The anti-plane deformation is
determined by the out-of-plane displacement and
342 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
the in-plane electric eld and both are functions of
x
1
and x
2
, i.e.,
u
1
u
2
0; u
3
u
3
x
1
; x
2
; u ux
1
; x
2
:
32
Inserting Eq. (32) into Eq. (3) leads to
r
3k
c
44
ou
3
ox
k
e
15
ou
ox
k
; 33
D
k
e
15
ou
3
ox
k
e
11
ou
ox
k
; k 1; 2: 34
Substituting Eqs. (33) and (34) into Eq. (1) results
in
B
0
r
2
u 0; 35
where r
2
o
2
=ox
2
1
o
2
=ox
2
2
and
u u
3
; u
T
; B
0

c
44
e
15
e
15
e
11
_ _
: 36
Because e
2
15
c
44
e
11
6 0, B
1
0
exists and thus Eq.
(35) is equivalent to
r
2
u 0: 37
The general solution to Eq. (37) is
u fz fz with z x
1
ix
2
; 38
where fz is a complex function. Similarly, stress
vectors are dened as
R
1
r
31
; D
1

T
/
;2
; R
2
r
32
; D
2

T
/
;1
:
39
Substituting Eq. (38) into Eqs. (33) and (34) yields
/
;2
B
0
of
ox
1
B
0
of
ox
1
; 40
/
;1
B
0
of
ox
2
B
0
of
ox
2
: 41
Eq. (40) or Eq. (41) gives
/ iB
0
fz iB
0
fz: 42
Eqs. (38) and (42) can be re-written as
u Afz

Afz; 43
/ Bfz

Bfz; 44
where
A I; B iB
0
: 45
Eqs. (43) and (44) are the general solution of anti-
plane deformations. Furthermore, substituting
Eq. (45) into Eq. (12) yields
Y B
1
0

1
e
2
15
c
44
e
11
e
11
e
15
e
15
c
44
_ _

1
1 q
c
1
44
qe
1
15
qe
1
15
e
1
11
_ _
; 46
H
2
e
2
15
c
44
e
11
e
11
e
15
e
15
c
44
_ _

2
1 q
c
1
44
qe
1
15
qe
1
15
e
1
11
_ _
; 47
where
q
e
2
15
c
44
e
11
:
Eqs. (46) and (47) show that both Y
22
and H
22
are
negative.
Comparing the Stroh formalism for anti-plane
deformations to that for general deformations, one
will nd that if the solutions for anti-plane defor-
mations can be obtained, the corresponding solu-
tions for general deformations can be written out
without diculty.
3. Solutions derived from the elliptical-cavity-
approach
An elliptical cylinder cavity in a piezoelectric
material under anti-plane loading or general load-
ing is rst studied to take the electric eld inside
the cavity into account. When the cavity is reduced
to a sharp crack, the solutions for the cavity will be
reduced to the solutions for a mode III crack or a
general mode crack. Since the exact boundary
conditions are adopted along the cavity surface,
the results for the cavity can serve as a benchmark
to test the validity of various analysis approaches
to a mathematically sharp crack, which will be
discussed in the next section.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 343
3.1. Anti-plane deformations
3.1.1. An elliptical cylinder cavity under anti-plane
deformations
Consider an elliptical cylinder cavity in an
innite transversely isotropic piezoelectric material
under uniform remote anti-plane shear and in-
plane electric eld loading, i.e., R
21
r
1
32
; D
1
2

T
and R
11
r
1
31
; D
1
1

T
, as shown in Fig. 1. Assume
that the poling direction is along the positive x
3
-
axis, while the isotropic plane is in the x
1
x
2
plane.
The cavity is free of traction on its surface and is
lled with air or is in vacuum, whose dielectric
constant is e
0
. In this case, the electric eld inside
the cavity is uniform, that is
/
0
D
0
2
x
1
D
0
1
x
2
i
2
; 48
u
0
E
0
1
x
1
E
0
2
x
2
; 49
where i
2
0; 1
T
, and D
0
k
and E
0
k
are the compo-
nents of electric displacement and electric eld
inside the cavity, respectively, which are constant
and to be determined by the loading condition.
Inside the material, the complex function, fz,
involved in Eqs. (43) and (44) takes the form:
fz c
1
z f
0
z; 50
where c
1
is a constant related to the loading
condition at innity and f
0
z is an unknown
complex function that nulls at innity, i.e.,
f
0
1 0. Substituting Eq. (50) into Eqs. (43) and
(44) and noting that Y is real result in
/ R
21
x
1
R
11
x
2
2Rek
0
z; 51
u e
11
x
1
e
21
x
2
2YImk
0
z; 52
where
e
11
2c
1
13
; E
1
1

T
; e
21
2c
1
23
; E
1
2

T
;
k
0
z Bf
0
z:
On the cavity surface, the continuous conditions
require
/
0
/; u u
0
: 53
A mapping function, z zf,
zf Rf mf
1
; R
a b
2
; m
a b
a b
;
maps the ellipse in the z-plane into a unit circle in
the f-plane. In the f-plane, Eq. (53) can be re-
written, by using Eqs. (48), (49), (51) and (52), as
D
0
2
x
1
r D
0
1
x
2
r D
1
2
x
1
r D
1
1
x
2
r
2Rek
0
r
2
; 54
E
0
1
x
1
r E
0
2
x
2
r E
1
1
x
1
r E
1
2
x
2
r
2YImk
0
r
2
; 55
where r e
ih
, and [ ]
2
denotes the second row of
the [ ] vector. Solving Eqs. (54) and (55) results in
k
0
z
1
2
f
1
zp; 56
where
f
1
z
z

z
2
a
2
b
2
p
a b
; 57
p ar
1
32
ibr
1
31
; aD
1
2
D
0
2
ibD
1
1
D
0
1

T
;
58
D
1
2
D
0
2

bD
1
2
e
0
E
1
2
e
0
ReY
21
ar
1
32
ibr
1
31

b ae
0
Y
22
;
59
D
1
1
D
0
1

aD
1
1
e
0
E
1
1

a be
0
Y
22
: 60
For a crack-like aw with b d
c
and d
c
=a ( 1,
omitting be
0
in Eqs. (59) and (60) leads to
D
1
2
D
0
2

kD
1
2

H
21
H
22
r
1
32
1 k
; 61
D
0
1
e
0
E
1
1
; 62
1
x
2
x
a
b

2 32
, D

1 31
, D
Fig. 1. An elliptical cavity in a piezoelectric solid subjected to
remote loading.
344 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
where H
ij
is given by Eq. (47), and
k a=b; a d
c
=a; b e
0
=e
eff
;
e
eff
Y
1
22
e
11
e
2
15
=c
44
:
63
The parameters, a, b, and e
eff
were introduced and
discussed in [30,31]. For simplicity, one may use k
to present the ratio of a=b. The value of k ranges
from zero to innity, corresponding to electrically
permeable and impermeable cracks, respectively.
3.1.2. Field intensity factors of a mode-III crack
For this case, Eq. (56) becomes
k
0
z
1
2
f
1
zp
c
; 64
where
f
1
z z
_

z
2
a
2
p
__
a; p
c
aP
21
; 65
P
21
r
1
32
; D
1
2
D
0
2

T
: 66
According to the denitions of
k
III
lim
x
1
!a

2px
1
a
_
r
32
x
1
and
k
D
lim
x
1
!a

2px
1
a
_
D
2
x
1
; 67
the eld intensity factor vector is
k k
III
; k
D

pa
p
r
1
32
; D
1
2
D
0
2

T
: 68
Explicitly, the stress intensity factor is
k
III
r
1
32

pa
p
; 69
which is independent of k. However, the electrical
displacement intensity factor is a function of k. By
inserting Eq. (61) into Eq. (68) and noting from
Eq. (47) that H
21
=H
22
e
15
=c
44
, the intensity
factor of the electric displacement is given by
k
D

kD
1
2

e
15
c
44
r
1
32
1 k

pa
p
: 70
For an impermeable crack, k ! 1 and Eq. (70) is
reduced to
k
D
D
1
2

pa
p
: 71
For a permeable crack, k ! 0, and combining Eqs.
(69) and (70) gives
k
D

e
15
c
44
r
1
32

pa
p

e
15
c
44
k
III
: 72
Eq. (72) indicates that the intensity factor of the
electric displacement for a permeable crack is in-
duced by the piezoelectric eect.
3.1.3. Energy release rate
There are four isothermal thermodynamic
functions: free energy, f , electric enthalpy, h, me-
chanical enthalpy, w, and full Gibbs free energy, g
[2,32]. Each of the four isothermal thermodynamic
functions corresponds to an isothermal potential,
P, that is dened as the dierence between the
generalized work, W , done by mechanical and
electrical loads and the generalized mechanical and
electrical energy, U, stored inside the body of
interest. In a dierential form, one has
oP oW oU: 73
If electric enthalpy is adopted, the generalized work
and the generalized stored energy are given by
W
_
C
r
ij
n
j
u
i
D
i
n
i
u dC; 74
U
_ _
X
hdX 75
where C denotes a contour, as shown in Fig. 2, X
is the integration domain closed by C, u
i
is the
mechanical displacement, D
i
n
i
is the prescribed
boundary value of the electric displacement, and h
is the electric enthalpy per unit volume such that
h
1
2
r
ij
e
ij

1
2
D
i
E
i
: 76
Dene the energy release rate, G, as
G
oP
oA
; 77
where A is the area of the crack faces.
1
x
2
x

Fig. 2. The integration contour to calculate the energy release


rate.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 345
From Eqs. (73)(75), it is found that
G
1
2
_
C
d/
T

ou
oa
: 78
According to the condition for self-similar
enlargement of the cavity, i.e., b=a or D
0
2
remains
xed during crack propagation, it can be shown
from Eq. (78) that
G
pa
4
R
T
21
HP
21
; 79
where P
21
is given in Eq. (66), and
R
T
21
r
1
32
; D
1
2

T
: 80
Substituting Eqs. (80), (47) and (66) into Eq. (79)
leads to
G
pa
4
H
11
_

H
2
21
H
22
1
1 k
_
r
1
32

2
H
22
k
1 k
D
1
2

2
H
21
2k
1 k
D
1
2
r
1
32
: 81
For an impermeable crack, Eq. (81) becomes
G
pa
4
H
11
r
1
32

2
H
22
D
1
2

2
2H
21
D
1
2
r
1
32
: 82
Inserting Eq. (47) into Eq. (82) results in
G
pa
2
e
11
r
1
32

2
2e
15
D
1
2
r
1
32
c
44
D
1
2

2
e
11
c
44
e
2
15
_ _
;
83
which is consistent with a previous result [33]. For
a permeable crack, Eq. (81) becomes
G
pa
4
H
11
_

H
2
21
H
22
_
r
1
32

2
: 84
Using Eq. (47), it is found that
H
11
_

H
2
21
H
22
_

2
c
44
: 85
Inserting Eq. (85) into Eq. (84) produces
G
pa
2c
44
r
1
32

2
; 86
which is also consistent with a previous result [30].
In summary, the energy release rate is dependent
on the value of k, as shown in Eq. (81).
3.2. General two-dimensional deformations
3.2.1. An elliptical cylinder cavity under generalized
2D deformation
If the sample orientation is arbitrary and all
eld variables in the solid are independent of x
3
,
the problem becomes a general two-dimensional
problem. In this case, the far-eld mechanical
electric loading condition becomes R
21
r
1
21
; r
1
22
;
r
1
23
; D
1
2

T
and R
11
r
1
11
; r
1
12
; r
1
13
; D
1
1

T
. Similarly,
there is no traction on the cavity surface and the
cavity is lled with air or is in vacuum. For the
generalized 2D problem, the complex function,
fz

, has the form


fz

hhz
a
iic
1
f
0
z

; 87
where hh ii indicates a diagonal matrix, f
0
z

is a
complex vector, and again f
0
1 0.
Following a similar procedure as that used in
solving anti-plane deformations, the electric elds
inside the cavity and the complex potentials in the
material [34] can be obtained. The electric elds
inside the cavity are
D
1
2
D
0
2

bD
1
2
e
0
E
1
2
e
0
Re

3
j1
Y
4j
M
1
j
b ae
0
Y
44
; 88
D
1
1
D
0
1

aD
1
1
e
0
E
1
1
e
0
Im

3
j1
Y
4j
M
1
j
a be
0
Y
44
; 89
where Y is determined from Eq. (12), and
M
1
ar
1
21
ibr
1
11
; ar
1
22
ibr
1
12
; ar
1
23
ibr
1
13

T
:
90
The complex potentials in the material are
/ R
21
x
1
R
11
x
2
2Rek
0
z

; 91
u e
11
x
1
e
21
x
2
2ImYk
0
z

; 92
where
e
11
c
1
11
; c
1
12
x
1
3
; 2c
1
13
; E
1
1

T
;
e
21
c
1
21
x
1
3
; c
1
22
; 2c
1
23
; E
1
2

T
;
k
0
z

Bf
0
z


1
2
hhf
1
a
z
a
iip;
93
346 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
f
1
a
z
a

z
a

z
2
a
c
2
a
_
a ip
a
b
; c
2
a
a
2
p
2
a
b
2
; 94
p ar
1
21
ibr
1
11
; ar
1
22
ibr
1
12
; ar
1
23
ibr
1
13
; aD
1
2
D
0
2
ibD
1
1
D
0
1

T
: 95
The generalized strain vectors, e
11
and e
21
, are
determined from r
1
ij
and D
1
j
through the consti-
tutive equations and an additional condition of
the value of the remote rotation, which is usually
taken as zero, i.e., x
1
3
0.
3.2.2. Field intensity factors
For a crack-like aw with an initial width,
b d
c
, the electric eld inside the crack becomes
D
1
2
D
0
2

kD
1
2

1
H
44

3
j1
H
4j
r
1
2j
1 k
; 96
D
0
1
e
0
E
1
1
e
0

3
j1
ImY
4j
r
1
2j
; 97
where k is still dened by Eq. (63), but in this case
e
eff
in k should be replaced with e
eff
Y
1
44
. Then,
the complex function, k
0
z

, is given by
k
0
z


1
2
hhf
1
a
z
a
iip
c
; 98
where
p
c
aP
21
; 99a
P
21
r
1
21
; r
1
22
; r
1
23
; D
1
2
D
0
2

T
; 99b
f
1
a
z
a

1
a
z
a
_

z
2
a
a
2
_ _
: 100
The eld intensity factor vector takes the form of
k

pa
p
P
21
: 101
Substituting Eq. (99b) together with Eq. (96) into
Eq. (101), there gives
k
j

pa
p
r
1
2j
; j 1; 2; 3; 102
k
D

pa
p
k
1 k
D
1
2
_

1
H
44
1
1 k

3
j1
H
4j
r
1
2j
_
:
103
Clearly,
k
1
k
II
; k
2
k
I
; and k
3
k
III
:
For an impermeable crack, k ! 1, and Eqs. (102)
and (103) give
k
j

pa
p
r
1
2j
; j 1; 2; 3 and k
D

pa
p
D
1
2
;
104
respectively. For a permeable crack, k 0, and
Eqs. (102) and (103) give
k
j

pa
p
r
1
2j
; j 1; 2; 3; 105
k
D

pa
p 1
H
44

3
j1
H
4j
r
1
2j
: 106
Comparing Eqs. (102)(106) to Eqs. (68)(72), it is
found that the results for a general crack are
similar to these for a mode III crack.
3.2.3. Energy release rate
Similar to the case for a mode III crack, the
energy release rate for a general crack can be ex-
pressed as
G
pa
4
R
T
21
HP
21
: 107
where H and P
21
are determined from Eq. (12)
and Eq. (99b), respectively, and
R
21
r
1
21
; r
1
22
; r
1
23
; D
1
2

T
: 108
For an impermeable crack, D
0
2
0, and Eq. (107)
gives
G
pa
4
R
T
21
HR
21
: 109
When electric loading is only applied, Eq. (109)
becomes
G
pa
4
H
44
D
1
2

2
: 110
Since H
44
< 0, Eq. (110) shows that G is always
negative for an electrically impermeable crack
loaded by a remote uniform electric displacement.
For a permeable crack, k 0 and Eq. (96) be-
comes
D
1
2
D
0
2

1
H
44

3
j1
H
4j
r
1
2j
; 111
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 347
and then Eq. (107) is reduced to
G
pa
4
P
T
21
HP
21
: 112
Namely,
G
pa
4
r
T
21
H

r
21
; 113
where
r
21
r
1
21
; r
1
22
; r
1
23

T
; H

fH

ij
g
33
H
ij
_

H
i4
H
4j
H
44
_
33
:
Obviously, H

is positive denite. This means that


G is always positive for a permeable crack.
4. Solutions derived from the crack-approach
The elliptical-cavity-approach is used in Section
3 to solve crack problems in which the exact
electric boundary conditions can be used and the
crack solutions are obtained by reducing the cavity
to a crack. For general crack problems, however, it
may be dicult to obtain solutions if the elliptical-
cavity-approach is adopted. This section presents
a general method, which directly solves general
crack problems in an innite piezoelectric solid.
Some basic, important and foundational solutions
are obtained by using the general method within
the linear framework. Although these results are
given only for a single crack, it is easy to extend
these results to more complicated crack geometries
or loading conditions.
Consider a generalized 2D problem of a single
crack, as shown in Fig. 3. The crack is assumed to
be free of traction and external charge, and the
material is subjected to remote uniform loading. In
this case, inserting Eq. (87) into Eqs. (8) and (9)
leads to
/
;1
R
21
K
0
z K
0
z; 114
u
;1
e
11

1
i
YK
0
z YK
0
z; 115
where K
0
z BF
0
z, and R
21
is given by Eq.
(108). On the x
1
-axis, the continuous conditions of
generalized traction require
/

;1
/

;1
for 1 < x
1
< 1: 116
Substituting Eq. (114) into Eq. (116) leads to
[35]
K
0
z K
0
z 0; 117
where

K
0
z K
0
z. Using Eq. (117), one has
from Eqs. (114) and (115) that
/

;1
R
21
K

0
x
1
K

0
x
1
; 118
iDu
;1
HK

0
x
1
K

0
x
1
: 119
On the crack faces, the generalized stress boundary
condition is
K

0
x
1
K

0
x
1
/

;1
x
1
R
21
for a < x
1
< a:
120
In addition, the singled-value condition of dis-
placements and electric potential requires
_
a
a
K

0
x
1
K

0
x
1
dx
1
0: 121
4.1. Solutions for an impermeable crack
For an impermeable crack, the boundary con-
ditions on the crack faces are
r

2j
r

2j
0 and D

2
D

2
0
for a < x
1
< a;
122
which are equivalent to
/

;1
x
1
0 for a < x
1
< a: 123
x
1
2 x
2a
1 1
,
j D

2
,
j D
2

Fig. 3. A crack in a piezoelectric solid subjected to remote


loading.
348 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
Inserting Eq. (123) into Eq. (120) gives
K
0
z
1
2
R
21
1
_

z

z
2
a
2
p
_
: 124
With Eq. (124), it can be shown that
ka

pa
p
R
21
; G
pa
4
R
T
21
HR
21
;
Du
1
2
HR
21

a
2
x
2
1
_
:
125
4.2. Solutions for a conducting crack
The boundary conditions along the faces of a
conducting crack are
r

2j
r

2j
0 and E

1
E

1
0
for a < x
1
< a;
126
which are equivalent to
^
/

;1
x
1
0 for a < x
1
< a: 127
Similarly, one has

K
0
z
1
2

R
21
1
_

z

z
2
a
2
p
_
: 128
Then, the eld intensity factor vector, the energy
release rate and the generalized crack opening are
given by
^
ka k
j
; k
E
1

pa
p

R
21
;
^
G
pa
4

R
T
21

R
21
;
D^u
1
2

R
21

a
2
x
2
1
_
;
129
where

R
21
r
1
21
; r
1
22
; r
1
23
; E
1
1

T
:
Note that
^
G is always positive for a conducting
crack because

H is positive denite.
4.3. Solutions for a permeable crack
The boundary conditions along the faces of a
permeable crack are [16]
r

2j
r

2j
0 for a < x
1
< a
and D

2
D

2
; E

1
E

1
for 1 < x
1
< 1: 130
Eqs. (130) and (119) indicate
H
4
K

0
x
1
K

0
x
1
0 for 1 < x
1
< 1;
131
where H
4
stands for the fourth row of H. The
solution to Eq. (131) is [35]
H
4
K
0
z 0: 132
One has from Eq. (132) that
K
40
z
1
H
44

3
j1
H
4j
K
j0
z: 133
On the other hand, Eq. (118) can be re-written as
r

2j
x
1
r
1
2j
K

j0
x
1
K

j0
x
1
; j 1; 2; 3;
134
D

2
x
1
D
1
2
K

40
x
1
K

40
x
1
: 135
With the use of Eq. (133), Eq. (135) is reduced to
D

2
x
1
D
1
2

1
H
44

3
j1
H
4j
K

j0
x
1
K

j0
x
1
:
136
From Eq. (134), the boundary conditions along
the crack faces are expressed by
K

j0
x
1
K

j0
x
1
r
0
2j
x
1
r
1
2j
for a < x
1
< a:
137
Substituting Eq. (137) into Eq. (136) leads to
D
0
2
x
1
D
1
2

1
H
44

3
j1
H
4j
r
0
2j
x
1
r
1
2j
: 138
Eq. (138) shows that when the stress, r
0
2j
x
1
, is
given on crack faces, the boundary value of electric
displacement can be easily written out. For the
present case, r
0
2j
x
1
0 and thus Eq. (138) gives
D
0
2
D
1
2

1
H
44

3
j1
H
4j
r
1
2j
: 139
Similarly, one can obtain another component of
the electric eld inside the crack as
E
0
1
E
1
1

3
j1
ImY
4j
r
1
2j
: 140
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 349
Using Eq. (139), one is able to re-express the
boundary conditions as
/

;1
i
4
D
0
2
; i
4
0; 0; 0; 1
T
for a < x
1
< a:
141
Since D
0
2
is a constant in this case, the solution to
Eq. (120) can be easily obtained by inserting Eq.
(141) into Eq. (120). It is
K
0
z
1
2
R
21
i
4
D
0
2
1
_

z

z
2
a
2
p
_
: 142
Then, the eld intensity factor vector is given by
ka

pa
p
R
21
i
4
D
0
2
: 143
Its components are
k
j
a r
1
2j

pa
p
; 144
k
D
a
1
H
44

3
j1
H
4j
k
j
a: 145
In addition, the crack opening and the energy re-
lease rate take the forms of
Du
1
2
HR
21
i
4
D
0
2

a
2
x
2
1
_
;
G
pa
4
P
T
21
HP
21
;
146
respectively, where P
21
and R
21
are given by Eqs.
(99b) and (108).
According to a similar procedure described
above, one can obtain solutions for a permeable
crack under more complicated loading conditions
and/or with a more complex crack geometry. For
example, when a far-eld uniform heat ow,
q
1
2
, is added in Fig. 3, the electric elds inside a
thermally impermeable and electrically permeable
crack can be obtained from using Eqs. (20) and
(21) as [36]
D
0
2
D
1
2

1
H
44

3
j1
H
4j
r
1
2j

ImM
4

j
t
Y
44
q
1
2
x
1
; 147
E
0
1
E
1
1

3
j1
ImY
4j
r
1
2j

ReM
4

j
t
q
1
2
x
1
; 148
and the eld intensity factor vector is
ka k
T
a k
r
a k
r
D
a k
T
D
a; 149
where
k
T
a 2

p
a
_
Bd dc
1
;
k
r
a

pa
p
R
21
;
k
r
D
a i
4

pa
p 1
H
44

3
j1
H
4j
r
1
2j
;
k
T
D
a i
4

pa
p ImM
4
j
t
Y
44
aq
1
2
;
in which j
t
, d and c
1
are parameters related to the
thermal and elastic properties of the material. (See
[36] for details.)
Another example is a permeable circular-arc
crack in an innite, transversely isotropic piezo-
electric material under uniform remote anti-plane
shear and in-plane electric eld, as shown in Fig. 4.
Assume that the poling direction is along the x
3
-
axis, while the isotropic plane is in the x
1
x
2
plane.
It can be shown that, in this case, the electric dis-
placement inside the crack is [37]
D
0
2r
h D
1
2
c
D
r
1
32
sin h D
1
1
c
D
r
1
31
cos h
for h
0
< h < h
0
; 150
and the eld intensity factor vector is
k
r
a 2

pRsin h
0
_
r
1
32
sin
h
0
2
_
r
1
31
cos
h
0
2
_
;
151
k
D
a c
D
k
r
a; 152
2
x
1
x
o

1
D
2
D
23

13

o i
e R a =
Fig. 4. The case of a circular-arc crack.
350 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
where
c
D

e
15
c
44
:
4.4. Solutions for a semi-permeable crack
A semi-permeable crack is a crack, along which
faces the electrically semi-permeable boundary
conditions [38],
r

2j
r

2j
0; D

2
D

2
e
0
Du
Du
2
; 153
are applied, where Du is the electric potential drop
across the crack and Du
2
is the crack opening.
Based on Eqs. (153), if one assumes in advance
that D
0
2
is a constant, Eqs. (141)(143) and (146)
are still valid, but D
0
2
in the related equations
should be re-determined by Eq. (153).
For a generalized 2D problem with a crack, as
shown in Fig. 3, one has from Eq. (146) that
Du
1
2
H
4
R
21
i
4
D
0
2

a
2
x
2
1
_
; 154
Du
2

1
2
H
2
R
21
i
4
D
0
2

a
2
x
2
1
_
: 155
Inserting Eqs. (154) and (155) into Eq. (153) leads
to
D
0
2
e
0
H
4
R
21
i
4
D
0
2

H
2
R
21
i
4
D
0
2

; 156
which is equivalent to
H
24
D
0
2

2
e
0
H
44
H
2
R
21
D
0
2
e
0
H
4
R
21
0:
157
The solution of Eq. (157) is
D
0
2

1
2H
24
H
2
R
21
_
e
0
H
44

H
2
R
21
e
0
H
44

2
4e
0
H
24
H
4
R
21
_
_
:
158
Substituting Eq. (158) into Eqs. (143) and (146)
gives the solutions of the eld intensity factors and
the energy release rate for the semi-permeable
boundary condition.
Caution should be used in employing the semi-
permeable boundary conditions when only electric
loading is applied. In this case, the value of D
0
2
should be determined from Eq. (96), which is
D
1
2
D
0
2

kD
1
2
1 k
; 159
when no mechanical loads are applied. Eq. (159)
yields the solution of
D
0
2

D
1
2
1 k
: 160
Eq. (160) is equivalent to Eq. (3.66b) in [2].
However, if the value of D
0
2
is still determined from
Eq. (156), then
D
0
2
e
0
H
44
H
24
: 161
The value of D
0
2
given by Eq. (161) depends only
on the material properties and the sample orien-
tation, but it is independent of the applied electric
eld. Clearly, the result contradicts the solution
derived from the elliptical-cavity-approach, in
which the exact boundary conditions are used.
Under purely electrical loading, one has from Eqs.
(103) and Eq. (107) that
k
D

k
1 k

pa
p
D
1
2
and
G
pa
4
k
1 k
H
44
D
1
2

2
: 162
In summary, the theoretical results within the
framework of linear electro-elasticity show that
when an electric loading is applied solely, (i) G is
always negative for an impermeable crack or a
semi-permeable crack, (ii) G 0 for a permeable
crack, and (iii) G is always positive for a con-
ducting crack. A negative value of G indicates that
an impermeable crack or a semi-permeable crack
will not propagate under purely electrical loading,
if the material still follows linear electro-elasticity.
From the denition of G or from the energy point
of view, a negative value of G means that when
there is a virtual extension of the crack length, the
change in the work done by external loads is less
than the change in the internal energy of the
material.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 351
A negative energy release rate can be under-
stood if one considers induced charges along the
crack surfaces. That is, under applied loading,
positive (or negative) and negative (or positive)
bound charges emerge on the upper and lower
surfaces of the crack, respectively. The charges at
the upper and lower surfaces are opposite in sign
and equal in magnitude of the intensity, D
0
2
, where
D
0
2
is the normal component of electric displace-
ment along the crack faces. The charges with
opposite signs generate an attractive Columbic
force that has a tendency to close the crack. The
attractive Columbic force is similar to the ber
bridge force in the bridge-crack model for ber-
reinforced composites. Recently, the experimental
work shows that an applied electric eld always
has a tendency to retard crack growth [39].
5. Strip dielectric breakdown model
The results described above are based on linear
electro-elasticity, showing that the electric eld
might be singular at the crack tip. For an electri-
cally conductive or impermeable crack, the singu-
larity of electric eld is dependent on the applied
electric eld parallel or perpendicular to the crack
line, while for an electrically permeable crack, the
crack-tip singularity of electric eld is related to the
stress singularities via the piezoelectric eect. Even
at a permeable crack tip, the electric eld strength is
about 1000 times higher in magnitude than the
applied electric eld for most piezoelectric materi-
als, which have about a 1000 times higher dielectric
constant in magnitude than the crack interior (air
or vacuum). Under such a high local electric eld,
electrical non-linearity may occur near the crack
tip. The strip polarization saturation (PS) model
[25] was developed to explore the eects of electri-
cal non-linearity on piezoelectric fracture. The PS
model takes the advantage that the constitutive
relationship between the electric displacement and
the electric eld strength is similar to that between
the stress and the strain. For that reason, the PS
model corresponds to a mechanical Dugdale model
in which the strain remains a constant value as the
stress increases. Comprehensive comments on the
PS model were given in [40]. From the energy point
of view, the electric displacement behaves like the
strain, while the electric eld strength functions like
the mechanical stress. Therefore, the PS model
does not correspond to the classical Dugdale model
in which the stress is equal to the yield strength
along the strip in front of a crack tip.
A strip dielectric breakdown (DB) model is
proposed here. The DB model is exactly analogous
to the classical Dugdale model from the energy
point of view because, in the DB model, the electric
eld strength on a strip adjacent to a crack tip is
taken as a constant. The physical arguments for the
DB model are described as follows. As discussed
above, a very high electric eld exists near a crack
tip when the piezoelectric material is subjected to
mechanical and/or electric loads. The local elec-
tric eld may be much higher than the dielectric
breakdown strength. The dielectric breakdown
strength is dened as a critical electric eld at which
dielectric discharge occurs and leads to dielectric
breakdown. The characteristics of the breakdown
strength are similar to those of the mechanical
fracture strength [41], which is very sensitive to
solid defects such as aws, voids and cracks,
thereby indicating that partial discharge may occur
at the crack tip due to the high electric eld. As a
result, a local partial discharge zone or electric
breakdown zone, like a plastic deformation zone, is
formed adjacent to the crack tip, and in the partial
discharge zone the electric eld cannot exceed the
dielectric breakdown strength, E
b
, which is analo-
gous to the yield strength in mechanically plastic
deformation. In analogy with the mechanical
Dugdale model, the dielectric breakdown region is
assumed to be a strip along the cracks front line,
where the electric eld strength is equal to the
dielectric breakdown strength, E
b
.
The simplied electro-elasticity constitutive
equations and an electrically impermeable crack,
which were used in the PS model, were adopted
here in order to compare the present results with
those from the PS model. The simplied consti-
tutive equations make the formulation simple and
give more insights into the physical picture. Under
the small-scale yielding condition, which is used
in the present study, the partial discharge zone is
much shorter than the crack length and thus the
crack can be considered to be semi-innitely long.
352 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
Consider a semi-innite impermeable crack
perpendicular to the poling direction in a trans-
versely isotropic piezoelectric material. The coor-
dinate system is set up such that the crack tip is at
its origin and the crack is on the negative x-axis,
as shown in Fig. 5. To reduce the number of the
independent material constants to a minimum, the
constitutive equation (3) is simplied to [25]
r
11
r
22
r
33
r
23
r
13
r
12
_

_
_

_
M
1 0 0 0
1 0 0 0
1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0
_

_
_

_
e
11
e
22
e
33
2e
23
2e
13
2e
12
_

_
_

_
e
0 0 1
0 0 1
0 0 1
0 1 0
1 0 0
0 0 0
_

_
_

_
E
1
E
2
E
3
_
_
_
_
_
_
; 163
D
1
D
2
D
3
_
_
_
_
_
_
e
0 0 0 0 1 0
0 0 0 1 0 0
1 1 1 0 0 0
_
_
_
_
e
11
e
22
e
33
2e
23
2e
13
2e
12
_

_
_

_
j
1 0 0
0 1 0
0 0 1
_
_
_
_
E
1
E
2
E
3
_
_
_
_
_
_
; 164
where e
i;j
(i; j 1, 2, 3) denote mechanical strains,
* means that the corresponding constant will not
appear in the model. Only three positive indepen-
dent material constants, M, e and j, are used to
represent, on a qualitative basis, the elastic, pie-
zoelectric and dielectric properties of the material.
Note that the constitutive equations expressed in
Eqs. (163) and (164) are for the poling direction
along the x
3
direction.
Furthermore, the material is constrained to
move only in the y-direction [25], i.e.,
u
x
0; u
y
u
y
x; y; u ux; y: 165
Re-arranging the constitutive equations by substi-
tuting Eqs. (2) and (165) into Eqs. (163) and (164)
gives
r
yx
Mu
;x
eu
;x
;
r
yy
Mu
;y
eu
;y
;
D
x
eu
;x
ju
;x
;
D
y
eu
;y
ju
;y
:
166
If u and u are expressed as imaginary parts of two
analytic functions, i.e.,
u ImUz; u ImUz; 167
the equilibrium conditions of Eq. (1) are satised
automatically. In this case, the constitutive equa-
tions and kinematic equations are given by
e
yy
i2e
yx
U
0
z;
E
y
iE
x
U
0
z;
168
r
yy
ir
yx
MU
0
z eU
0
z;
D
y
iD
x
eU
0
z jU
0
z:
169
The stress, strain, electric eld strength, and elec-
tric displacement intensity factors are dened by
K
r
lim
z!0

2pz
p
r
yy
; K
e
lim
z!0

2pz
p
e
yy
;
K
E
lim
z!0

2pz
p
E
y
; K
D
lim
z!0

2pz
p
D
y
:
170
Eq. (170) also satises the constitutive relation-
ships given by Eq. (169). From the denitions of
the intensity factors, we can express the mechani-
cal and electrical elds near the crack in terms of
the four intensity factors, i.e.,
r
yy

K
r

2pr
p cos
h
2
; r
yx

K
r

2pr
p sin
h
2
;
e
yy

K
e

2pr
p cos
h
2
; 2e
yx

K
e

2pr
p sin
h
2
;
E
y

K
E

2pr
p cos
h
2
; E
x

K
E

2pr
p sin
h
2
;
D
y

K
D

2pr
p cos
h
2
; D
x

K
D

2pr
p sin
h
2
;
171
x
a o
y(pole)
Fig. 5. A semi-innite crack with a dielectric breakdown strip.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 353
where r is the distance from the crack tip and h is
the polar angle. Using Eq. (171) and the J-integral,
J
_
C
hn
1
r
ij
n
j
u
i;1
D
i
E
1
n
i
dC; 172
where h is the electric enthalpy per unit volume
dened by Eq. (76), one obtains the relationship
between the energy release rate and the eld
intensity factors,
J
1
2
K
r
K
e
K
D
K
E
: 173
It should be emphasized that Eq. (173) is valid as
long as Eq. (171) is valid along the path of the J-
integral. In linear analysis, i.e., without consider-
ing the breakdown zone, Eq. (173) gives
J
a

1
2
K
a
r
K
a
e
K
a
D
K
a
E
; 174
where the superscript a denotes the applied. The
applied strain intensity factor and the electric dis-
placement intensity factor can be expressed in
terms of the applied stress intensity factor and
electric intensity factor by using the constitutive
Eq. (169), i.e.,
K
a
e

K
a
r
eK
a
E
M
; K
a
D
j 1
_

e
2
Mj
_
K
a
E

e
M
K
a
r
:
175
Inserting Eq. (175) into Eq. (173) leads to the
expression of the applied energy release rate with
two independent loading parameters as
2J
a

1
M
K
a
r

2
j 1
_

e
2
Mj
_
K
a
E

2
: 176
Eq. (176) is consistent with the results of Eq. (83)
or Eq. (109) for an impermeable crack under mode
III or general mode loading, correspondingly.
An electric dislocation is studied here prior to
discussion of the DB model. In general, a piezo-
electric dislocation may have eight characteristics,
as discussed in [2], which are three jumps in three
mechanical displacements, which yield the Burgers
vector, a jump in electric potential, three line for-
ces per unit length along three coordinate direc-
tions, and a line charge per unit length. The
electric dislocation studied here has only one
characteristic, the potential jump, Du. The electric
dislocation produces an electric eld, but it does
not produce any mechanical displacement. The
complex potentials for such an electric dislocation
located at z
d
in an innite domain without any
cracks are given by
U 0;
U b
u
lnz z
d
;
177
where
b
u

Du
2p
: 178
The electric dislocation does not produce any
strain eld, but it induces a stress eld through the
piezoelectric eect.
When an electric dislocation is located at z
d
near
a semi-innite impermeable crack, the boundary
conditions,
r
yy
0; D
y
0; for x < 0; 179
should be satised. The complex potentials satis-
fying the boundary conditions take the following
form:
U 0;
U b
u
ln

z
p


z
d
p
b
u
ln

z
p

z
d
p
;
180
where the overbar denotes the conjugate of a
complex variable. The electric dislocation pro-
duces a stress eld and an electric eld, which are
given by
E
y
iE
x

b
u
2

z
d
p


z
d
p

z
p
z

z
p


z
d
p


z
d
p


z
d
z
d
p

;
r
yy
ir
yx
eE
y
iE
x
;
D
y
iD
x
jE
y
iE
x
:
181
When the electric dislocation is located on the
x-axis, Eq. (181) is reduced to
E
y
iE
x
b
u

x
d
p

z
p
z x
d

;
r
yy
ir
yx
eb
u

x
d
p

z
p
z x
d

;
D
y
iD
x
jb
u

x
d
p

z
p
z x
d

:
182
Substituting Eq. (182) into the intensity factor
denition of Eq. (170) leads to
354 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
K
d
E

Du

2px
d
p ;
K
d
r
eK
d
E
;
K
d
D
jK
d
E
;
183
where the superscript (d) denotes the electric
dislocation. Eq. (183) gives the stress, electric and
electric displacement intensity factors produced by
the electric dislocation in front of the semi-innite
impermeable crack.
Fig. 5 shows the proposed DB model, where 0a
denotes the dielectric breakdown size. If the dis-
tribution function of electrical dislocations is de-
noted by f x
0
, the number of dislocations located
at x
0
in the interval dx
0
is f x
0
dx
0
. The electrical
eld, E
y
, in the dielectric breakdown zone is de-
scribed by
K
a
E

2px
p b
u
_
a
0
f x
0

x
0
p

x
p
x x
0

dx
0
E
b
for 0 6x 6a:
184
The rst term on the left in Eq. (184) represents
the applied electric eld, while the second term
on the left stands for the induced electric eld. The
uniqueness condition for a distribution of f x
0
,
which has zero value at 0 and a, is given by [42]
K
a
E
2

2a
p
E
b
=

p
p
; 185
which is the equation determining the strip length,
i.e., Eq. (185) gives
a
p
2
K
a
E
2E
b
_ _
2
: 186
The solution to Eq. (184) is
f x
0

E
b
p
2
b
u
ln

a
p

a x
0
p

a
p

a x
0
p
_ _
: 187
The electrical dislocations produce an electric
intensity factor, which is calculated by
K
i
E

2p
p
b
u
_
a
0
f x
0

x
0
p dx
0
K
a
E
; 188
where the terms with the superscript i denote the
eld variables induced by the dielectric breakdown
strip. Using Eqs. (183) and (188) gives the induced
stress intensity factor and electric displacement
intensity factor:
K
i
r
eK
a
E
;
K
i
D
jK
a
E
:
189
The local electric intensity factor is the sum of the
applied electric intensity factor and the induced
electric intensity factor, i.e., K
l
K
a
K
i
, and
thus, one has
K
l
r
K
a
r
eK
a
E
;
K
l
e
K
a
e
;
K
l
E
0;
K
l
D
K
a
D
jK
a
E
:
190
Substituting Eq. (175) into Eq. (190) and then into
Eq. (173) gives the local energy release rate as
2J
l
DB

1
M
K
a
r
eK
a
E

2
; 191
where J
l
DB
indicates the local J-integral based on
the strip dielectric breakdown model. In this case,
the global J-integral is still given by Eq. (176).
Following a similar procedure described above,
the local J-integral can be derived from the PS
model. In the PS model, the electrical displace-
ment, D
y
, in the polarization saturation is de-
scribed by modifying (184). It is
K
a
D

2px
p jb
u
_
a
0
f x
0

x
0
p

x
p
x x
0

dx
0
D
s
for 0 6x 6a;
192
where D
s
is the saturated value of the electric dis-
placement. Then, the local intensity factors are
calculated to be
K
l
r
K
a
r
eK
a
E
K
a
D
=j;
K
l
e
K
a
e
;
K
l
E
K
a
E
K
a
D
=j;
K
l
D
0:
193
Finally, the local energy release rate, J
l
PS
, based on
the PS model [25] takes the form of
2J
l
PS

1
M
1
_

e
2
Mj
_
K
a
r
_
eK
a
E
_
2
: 194
In fact, Eq. (194) can be directly written out from
Eq. (39) in [25] when K
a
r


pa
p
r
1
and K
a
E

pa
p
E
1
are considered, where a is the semi-length
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 355
of a nite crack, E
1
and r
1
are the applied electric
eld and stress, respectively, as shown in Fig. 6.
Comparing Eq. (194) with Eq. (191), it is found
that
J
l
PS
J
l
DB
1
e
2
Mj
> 1: 195
This means that the PS model gives a higher value
of the local energy release rate than that derived
from the DB model. However, Eqs. (191) and
(194) both indicate that positive electric eld will
assist an applied mechanical stress to propagate
the impermeable crack if the local J-integral is
adopted as a failure criterion, while a negative
electric eld will retard crack propagation if
ejK
a
E
j < K
0
r;C
, where K
0
r;C
is the fracture toughness
of the material under purely mechanical loading.
The relationship between the applied stress inten-
sity factor and the applied electric intensity factor
is identical in the two models. In this sense, the DB
model gives the same result as the PS model.
Without taking the small-scale yielding condi-
tion into account in the DB model, one has to
consider a nite crack with the same geometry and
loading as shown in Fig. 6. In this case, inserting
K
a
r


pa
p
r
1
and K
a
E


pa
p
E
1
into Eq. (191)
leads to
J
l
DB

pa
2M
eE
1
r
1

2
: 196
However, the strip size in Fig. 6 is determined by
c a and c is given by
c
a
sec
p
2
E
1
E
b
_ _
: 197
When c a ( a, which corresponds to pE
1
=
2E
b
! 0 and the small-scale yielding condition,
the strip size is reduced to
c a
p
2
a
2
E
1
2E
b
_ _
2
: 198
Eq. (198) is consistent to Eq. (186).
Finally, it should be noted that though the
present analysis is only restricted to the case of an
impermeable crack for the sake of explicitness, the
present results may easily be extended to the cases
of permeable cracks and semi-permeable cracks.
These are left to future work.
6. Domain switching
In piezoelectric ceramics, many material prop-
erties vary non-linearly with applied electrical
and/or mechanical elds. Typical examples of the
non-linear relationships are the non-linear PE
(polarizationelectric eld) hysteresis loop and the
cE (strainelectric eld) buttery loop. Experi-
mental observations and theoretical analysis have
proved that domain switching causes these non-
linear behaviors. Domain switching changes the
internal stress and electric elds, and thus, may
play a substantial role in the failure behavior.
Therefore, many domain-switching-based models
of fracture have been developed analogous with
phase-transformation-based models of fracture. A
brief review of domain-switching-based models of
fracture is presented below.
A piezoelectric ceramic possesses a domain
structure below its Curie temperature. Due to
crystalline anisotropy, the direction of the spon-
taneous polarization of a domain is along one of
the polar axes. Loading on piezoelectric materials
produces a high local electric eld and a stress eld
near the tips of cracks or electrodes. A suciently
high local eld may cause 180- and/or 90-domain
switching, in which a domain switches its orien-
tation by 180 or 90 from one polar axis to an-
other polar axis to reduce the energy by aligning its
orientation as close as possible to the eld. When
the applied electric eld reverses its direction, 180-
switching will occur, which causes the polarization
hysteresis loop and the strain buttery loop.
Switching a domain by 180 causes little or no
mechanical distortion because the spontaneous
Fig. 6. A nite crack with a polarization saturation strip.
356 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
strain has about 180 symmetry. The spontaneous
strain, which will be dened in terms of lattice
constants later, is a measure of the lattice distor-
tion induced by spontaneous polarization. Thus,
180-domain switching is activated primarily by
the electric eld. However, 90-domain switching
involves a change in the orientation of the spon-
taneous strain by about 90 and inducing a high
local strain eld. As a direct consequence, both
applied mechanical and electrical elds can cause
domains to switch by 90, which in turn changes
the internal stress eld. If the state prior to the
switching is assumed to be stress-free, 90-domain
switching generates an internal stress eld. The
induced internal stress eld may help or retard the
applied mechanical eld to fracture a piezoelec-
tric sample, depending on its distribution. If the
internal stress is high enough, the switching-
induced stress itself can damage the sample. Sim-
ilarly, if the 90-switching occurs in the vicinity of
a mechanically and/or electrically loaded crack tip,
the switching-induced internal stress eld may,
depending on the nature of the internal stresses,
shield or anti-shield the crack tip from the
applied loads, resulting in switching-toughening or
switching-weakening. Switching-toughening may
be regarded as a kind of phase-transformation-
toughening [43] because polarization switching in
ferroelectric ceramics may be treated as a twinning
process.
In the development of domain-switching-based
models of fracture, the key problems include the
establishment of a switching criterion, the rene-
ment of simulations of the switching process,
the determination of the switching zone, and the
clarication of the failure criterion, etc. Thermo-
dynamic approaches can be used to establish
switching criteria and calculate the driving force of
the polarization switching [44]. An energy-based
criterion for an individual domain states that a
domain switches when the electric work plus the
mechanical work exceed a critical value [23].
According to the criterion, a 180-switching, where
stresses are irrelevant, is activated if
E
i
DP
i
P2P
s
E
c
; 199
where E
i
and DP
i
are, respectively, the electric eld
strength and the polarization switch vectors, P
s
is
the magnitude of the spontaneous polarization,
and E
c
is the absolute value of the coercive eld.
For a 90-switching, the work done by mechanical
stresses has to be taken into account, and the cri-
terion is
r
ij
De
ij
E
i
DP
i
P2P
s
E
c
: 200
The spontaneous strain may be estimated from
e
s

c
0
a
0
a
0
; 201
where c
0
and a
0
are the lattice constants of a
ferroelectric phase. Normalizing the changes in
strains as De
ij
e
s
e
ij
, one may dene a stress
type criterion for 90-switching as
r
ij
e
ij
Pr
c

P
s
E
c
e
s
2
_

E
i
E
c
DP
i
P
s
_
; 202
where r
c
is called the stress threshold, which is a
function of the electric eld [45]. Note that 90-
switching may occur without the presence of ex-
ternal electric elds.
The switching criterion given in Eq. (200) rep-
resents a simplied version of an energetic switch-
ing criterion but it neglects any interactions
between domains. As a step towards taking the
interactions into account, a polarization-switching
criterion for polycrystalline ferroelectric ceramics
[46] was developed from the potential energy of an
innite homogeneous piezoelectric medium con-
taining a single ellipsoidal transforming inclusion.
The switching criterion accounts for the eects of
the mean electric and stress elds arising from the
constraints presented by surrounding domains as
well as the externally applied mechanical and
electrical loads, and it can be expressed as
DU
I
V
I
DW
c
0; 203
where DU
I
is the potential energy change of the
inclusion, V
I
is the volume of the inclusion, and
DW
c
stands for the energy barrier per unit volume,
which must be overcome during switching. Insert-
ing explicit expressions of DU
I
and V
I
for 180- and
90-switching into Eq. (203) can produce explicit
forms of the switching criteria [46].
A rigorous derivation of the energy release rate
during the polarization switching [19] extends the
switching criterion Eq. (200) to a general form:
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 357
r
1
r
2
2
Dc
r

E
1
E
2
2
DP
r

r
1
r
2
2
DS

E
1
E
2
2
De
E
1
r
2
E
2
r
1
2
Dd PJ; 204
where Dc
r
; DP
r
; De; DS and Dd are the local chan-
ges in the remnant strain, remnant polarization,
compliance, permittivity and piezoelectric cou-
pling tensor due to switching, respectively, and the
superscripts 1 and 2 represent the eld states be-
fore and after switching. J is the energy barrier for
switching of a crystallite. Since the improved cri-
terion takes into account the dierent values of the
elds before and after switching as well as the
changes in the material tensors, it can be used to
study the stability of switching.
A simple domain-switching model for electri-
cally conductive and electrically impermeable
cracks was proposed on the basis of step-like
switching and ignoring the eect of stress on the
electric eld [47]. Since the electric eld in front of
the crack is high enough to cause domain switch-
ing, the switching zone is determined completely
by the electric eld and the switching occurs when
E > E
c
. In the absence of stress, the switching
strain is zero when E < E
c
, or e
s
when E > E
c
.
Thus, an internal stress eld is generated due to the
switching. Then, the internal stress eld contrib-
utes to the driving force or resistance in the crack
propagation. This model captures the essence of
the switching behavior and gives insight into the
source of the stress eld that drives the crack
growth.
The stress intensity factor criterion and the
stress criterion are widely used failure criteria in
domain-switching models. The local stress inten-
sity factors can be approximated by the sums of
the applied factors and the switching-induced
factors, which can be evaluated in the spirit of
transformation toughening. In the phase-trans-
formation-toughening theory, a transformation
wake shields the crack-tip stress intensity factor.
By analogy with the phase-transformation-tough-
ening theory, a switched wake shields the crack-tip
stress intensity factor [45]. The stress intensity
factor induced by 90-switching may be generally
expressed as
DK gX; 205
g
Y e
2
s
P
s
E
c
; 206
where Y is a generalized Youngs modulus and X is
related to the electrical and mechanical loads as
well as the domain orientation prior to switching.
Switching-toughening in mono-domain and mul-
tiple-domain ferroelectrics was studied in [45]
based on two basic assumptions: the stress-assisted
switching zone is conned within the specimen and
the electric eld is uniform. When the stress
intensity factor induced by 90-switching is avail-
able, the local stress intensity factor is the sum of
the applied intensity factor plus the induced one,
i.e.,
K
l
K
a
DK: 207
When the local stress intensity factor exceeds the
critical value, K
C
, i.e.,
K
l
PK
C
; 208
the crack will grow. Eq. (208) is the commonly
used failure criterion in domain-switching models.
A mechanistic explanation for the electric-eld-
induced fatigue crack growth was developed based
on a domain-switching model of an electrically
impermeable crack [48,49]. The high electric eld at
the impermeable crack tip drives domains to switch
and, consequently, the domain switching generates
an internal stress eld. A cyclic electric load causes
a domain switching sequence that generates a cyclic
internal stress eld, which mechanically fatigues
ferroelectric ceramics. In this case, the term X in
Eq. (205) is induced completely by the applied
electric eld and the crack is driven solely by
DK. Similarly, microcrack nucleation may also
occur during repeated domain switching, which de-
grades the electric properties of ferroelectric cera-
mics [5052].
In general, a domain-switching model requires
the development of new constitutive equations and
extensive numerical calculations. For example, a
non-linear constitutive model was proposed for
ferroelectric polycrystals under a combination of
mechanical stress and electric eld [53] by employ-
ing a self-consistent analysis, as an extension of
358 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
the self-consistent crystal plasticity scheme [5456],
to address ferroelectric switching and to estimate
the macroscopic response of tetragonal crystals
under a variety of loading paths. In a qualitative
way, this model captures several observed features
of ferroelectrics such as the shapes of the dielectric
hysteresis and the buttery loops, the Bauschinger
eect under mechanical and electrical loads and
the depolarization of a polycrystal by compressive
stresses [53].
Along with the development of dierent
switching models, many researchers carried out
numerical calculations on the eects of domain
switching on the fracture toughness. The Preisach
hysteresis model for each grain was used in
numerical simulations of non-linear behavior of
PLZT [23]. The response of the bulk ceramic to
applied loads was predicated by averaging the re-
sponse of individual grains that were considered to
be statistically random in orientation [23]. The
non-linear electric displacement versus electric
eld hysteresis loops [57] and stressstrain curves
[58] were simulated by using a nite element
model, in which each nite element represented
one crystallite and had its own principal crystal-
lographic directions, but was assigned randomly.
The material responses to the applied electric and/
or mechanical elds were averaged over the
switching responses of many randomly oriented
elements [57,58]. A continuous domain orientation
distribution function (ODF) could be used to de-
velop a domain-switching model for ferroelectrics
[59]. Using the ODF-based domain-switching
model and neglecting electromechanical interac-
tions between dierent domains, various mechan-
ical and electrical properties were simulated such
as virgin curves, and hysteresis and buttery
curves for uniaxial electromechanical loading
and unloading histories [59]. Recently, a novel
domain-switching model was developed with the
consideration of electromechanical interactions in
a self-consistent manner and the use of an ODF
[60,61]. More recently, the domain-switching-
induced enhancement of the steady-state fracture
toughness in a polycrystalline ferroelastic ceramic
was analyzed with a new phenomenological con-
stitutive law [62]. The constitutive law accounts for
the strain saturation, asymmetry in tension versus
compression, Bauschinger eects, reverse switch-
ing and strain re-orientation. In the analysis, crack
growth is assumed to proceed at a critical level
of the crack-tip mechanical energy release rate,
which is equivalent to the failure criterion of the
stress intensity factor. The predicated steady-state
toughness enhancement is in the range of DK
I
%
60140% [62], which is much higher than the re-
sults reported in [24], in which the maximum
contribution from the crack-tip-stress-induced do-
main rotation to the fracture toughness of a fer-
roelectric material is less than 10% even under the
most favorable conditions.
Numerical simulations of domain switching are
rather time-consuming. Several analytical methods
have been used in the determination of a switching
zone in order to simplify the simulation process.
The simple switching criterion proposed in [23]
was employed to analyze the complete eld prob-
lem around a crack tip with a non-linear switching
zone where the piezoelectric coupling eects are
neglected [45,63]. The domain-switching-based
fracture mechanics of ferroelectric/piezoelectric
ceramics was also theoretically investigated for a
conducting crack [64], an impermeable crack [65]
and a permeable crack in a piezoelectric compact
tension (CT) specimen [17]. Taking into account
the electrical non-linearity near the crack tip,
the simple switching criterion [23] was employed
to estimate the shape and size of the domain-
switching zone and then to develop a non-linear
ferroelectric fracture model, which might provide
a theoretical explanation for the experimental ob-
servation of various fracture behaviors of ferro-
electric materials [18].
It should be noted that most works available on
analytical modeling of domain switching are rst-
order approximations, since the related analyses
do not take into account the interactions between
domains during the switching process. That is,
under a given loading condition, the crack-tip eld
is rstly determined based on linear theory and
then the prole of the switching zone is obtained
by a simple switching criterion [23]. This zone can
be modeled as an Eshelby inclusion and the change
of the crack-tip stress intensity factors induced by
the inclusion can be evaluated by the Eshelby
technique or the Duhamel analogy for thermal
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 359
stress analysis. For simplicity, many analytical
models assume that the switching is the entire re-
sponse of an inclusion and there is no partial
transformation allowed. More assumptions are
that once the switching occurs the reverse switch-
ing and strain re-orientation cannot also be al-
lowed. In many cases, the electric characteristics
are ignored, such as the electric boundary condi-
tions between domains and the electrically direct
contribution to the local crack-tip energy release
rate, i.e., the electrical energy release rate. In fact,
domain switching is a gradual process; all the unit
cells in a domain do not switch simultaneously.
The process of domain switching may start at one
or a few crystalline cells near a crack tip. Local
switching will lead to a re-distribution of stress
and electric elds. Thus, the process of domain
switching is strictly a kinetic process involving
changes in the material microstructure. The size
and time scale eects should be reected in the
process. On this subject, numerical simulations of
the non-linear behavior of ferroelectrics are con-
ducted by using a phase-eld model [66], which
will be described in detail at the end of this section.
Polarization domain switching is a result of new
domain nucleation, if any one of the existing do-
mains does not meet the minimum energy condi-
tion, and domain wall motion. Domain wall
motion is equivalent to partial switching. Studying
domain wall kinetics will help to clarify the
mechanisms of the failure processes in piezoelec-
tric ceramics. In addition, the use of a simple do-
main wall kinetics model can greatly simplify the
mathematics involved in domain-switching mod-
els. Under applied electric and mechanical elds,
domains with low-energy orientations grow and
domains with high-energy orientations shrink,
resulting from domain wall motion. The displace-
ment of 90 domain walls causes a shear defor-
mation. Thus, both domain wall displacement and
volume deformation aect the material properties
of ferroelectric ceramics [67]. It has been studied
how to estimate the force constant and the eective
mass of a 90 domain wall in ferroelectric ceramics
with consideration of the domain size and the
grain size such that the domain wall motion at the
microstructure scale is linked to the macroprop-
erties of the ferroelectric ceramics [44]. The do-
main wall motion model explains well the internal
friction as well as the dielectric dispersion of fer-
roelectric ceramics [68]. A domain wall kinetics
model was developed to explain the eects of
temperature and the electric eld on the bending
strength of PZT-841 ceramics [69]. Either an ap-
plied stress or an electric eld can drive domain
walls to move [70]. In the domain wall kinetics
model, the total elastic compliance is divided into
two parts, one part is attributed to the volume
deformation inside the domains and the other is
attributed to the domain wall motion, which has
been veried by the phase eld simulations [66].
Experimental observations show that there is a
threshold stress below which the domain wall will
not move under an applied stress [69]. Experi-
mental observations also indicate that the thresh-
old stress is exponentially related to temperature
[69]. Domain wall motion, in return, changes the
internal stress eld, which is directly connected to
the spontaneous strain. The spontaneous strain, as
a function of temperature for PZT-841 ceramics,
was measured by X-ray diraction and given by
[69]
e
s
T 0:062 1
_

T
T
c
_
0:142 1
_

T
T
c
_
2
0:170 1
_

T
T
c
_
3
0:077 1
_

T
T
c
_
4
;
209
where T
c
(%272 C) is the Curie temperature. The
measured total elastic compliance ts the theoret-
ical predictions perfectly [2,69]. For polycrystalline
ceramics, cracks nucleate usually at grain bound-
aries. Using the stress failure criterion, failure oc-
curs when the total stress, including the applied
stress and the internal stress, exceeds the fracture
strength. In this way, the domain wall kinetics
model gives the apparent fracture strength, r
f
, as
a function of temperature. It is
r
f

r
B
1 ~ je
s
T expT=T
0

; 210
where r
B
denotes the intrinsic strength, which is a
constant, and ~ j and T
0
are constants. r
B
, ~ j and T
0
can be determined experimentally (see [2,69] for
details). Thus, the bending strength is a function of
360 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
temperature and Eq. (210) has been veried
experimentally [69].
The velocity, _ u, of domain wall motion depends
on the electric eld and is given by the empirical
equation [69]
_ u v
0
exp1 E
0
=jEj; 211
where E
0
is the activation eld strength and v
0
is
the domain wall velocity under E
0
. The domain
wall kinetics model links the velocity of domain
wall motion to the internal stress and gives [69]
r
f

r
B
r
E
0
exp1 E
0
=jEj
1 ks
D
; 212
where r
E
0
is the equivalent internal stress induced
by the activation eld of E
0
, k is a constant, and s
D
denotes the domain wall compliance, which varies
only with temperature. Eq. (212) gives the fracture
strength as a function of the applied electric eld.
For simplicity, ks
D
was assumed to be indepen-
dent of the applied electric eld. Taking r
B
99:3
MPa, Eq. (212) was used to t the experimental
data (see Fig. 11). The outcome was that r
E
0
35:7
MPa and E
0
14:2 kV/cm for the positive elds,
and r
E
0
22:8 MPa and E
0
14:0 kV/cm for the
negative elds. It is interesting to note that the
activation eld under positive electric loading is
almost the same as that under negative electric
loading.
The domain wall kinetics model is essentially the
same as a domain switching model. Both models
consider the internal stress eld induced by applied
mechanical and electrical loads. Depending on its
nature, the internal stress eld may assist or resist
the applied loads to fracture the samples. For
bending of smooth samples, the failure initiates
at locations of high internal tensile stress. Grain
boundaries, domain boundaries and other defects
are the potential locations for cracks to initiate.
Phase-eld model based on the time-dependent
GinzburgLandau equation can also be used to
model the non-linear behavior of ferroelectrics due
to polarization switching [71,72]. In the phase-eld
model, the nature of the phase transformation as
well as the microstructures is described by a set of
continuous order-parameter elds. The temporal
microstructure evolution is obtained by solving
kinetics equations that govern the time-dependence
of the spatially inhomogeneous order parameters.
Compared with the switching models above, the
important feature of this model does not make any
prior assumptions about the transient morpho-
logies and microstructures that may appear during
a phase transformation path. However, little
attention has been paid to the polarization switch-
ing of ferroelectrics subjected to external electrical
and/or mechanical loading using the phase-eld
model. In particular, how to combine phase-eld
simulations with nite element calculations still
remains a great challenge.
7. Experimental observations
7.1. Electrically insulated cracks
7.1.1. Indentation fracture tests
The fracture behavior of piezoelectric ceramics
under combined static mechanical and electrical
loads was studied using indentation-induced frac-
ture, three- or four-point bending, and fracture
testing on pre-cracked or pre-notched specimens.
The early indentation tests on PZT-8 ceramics in a
static electric eld were under an indentation load
of 4.9 N and the tests were repeated 1015 times
on dierent locations under a given applied electric
eld [73]. The crack length was measured from the
corner of the indent to the crack tip with either
optical micrographs or via a calibrated eyepiece.
The results indicate that the electric eld does not
change the average crack length of the cracks
parallel to the poling direction [73]. For the cracks
perpendicular to the poling direction, however,
analysis of the standard deviations in the data
suggests that the variations in the crack growth
due to poling and eld application are statistically
signicant. For example, the average crack length
is 21.50 lm under a negative electric eld of )4.7
kV/cm and the associated standard deviation is
4.82 lm, leading to a relative error of 22.4%. The
results show that the cracks perpendicular to the
poling direction become longer or shorter under a
positive or negative electric eld, indicating that
the positive eld assists the applied mechanical
load in propagating the crack, whereas the nega-
tive eld retards the crack propagation. The
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 361
positive electric eld of 4.7 kV/cm increases the
crack length perpendicular to the poling direction
by almost 70%, while a negative electric eld of
)4.7 kV/cm reduces it by 30% as compared to the
case of no applied electric eld [73]. However, an
opposite trend [74] was observed in the indentation
fracture tests on PZT EC-65 ceramics under the
load ranging from 4.9 to 11.76 N, showing that for
all indentation loads and in both the perpendicular
and parallel directions, a positive electric eld of 5
kV/cm resulted in shorter cracks than a negative
electric eld of )5 kV/cm did. In both directions,
the dierence in crack lengths measured under
positive and negative electric elds was larger at
smaller indentation loads and this dierence
diminished with increasing indentation load [74].
Using indentation loads of 9.8 and 49 N, the
indentation fracture tests under various applied
electric elds [75] reveal that the eect of an ap-
plied electric eld on the fracture behavior depends
on the magnitude of the indentation load. Under
the indentation load of 9.8 N, the indentation-
induced crack lengths perpendicular to the poling
direction were 103, 107, 109, 117, and 126 lm,
corresponding to the applied electric elds of )7.5,
)5.0, 0.0, 5.0, and 7.5 kV/cm, indicating a similar
trend as observed in [73]. But, under the indenta-
tion load of 49 N, the indentation-induced crack
lengths perpendicular to the poling direction were
316, 315, 302, 320, and 349 lm, corresponding to
the applied electric elds of )5.0, )2.5, 0.0, 2.5,
and 5.0 kV/cm, showing that both negative and
positive elds facilitated crack propagation. The
experimental results indicate that the eect of
electric eld on the indentation fracture behavior
depends on the level of the indentation load.
Moreover, it was reported in [75] that under the
indentation load of 49 N, the indentation-induced
crack lengths parallel to the poling direction were
176, 172, 168, 167, and 182 lm, corresponding to
the applied electric elds of )5.0, )2.5, 0.0, 2.5,
and 5.0 kV/cm. One may use the data under 5.0
kV/cm as an example to demonstrate the inden-
tation load-dependent phenomenon. Under the
indentation load of 9.8 N, the electric eld of )5.0
kV/cm (or 5.0 kV/cm) decreases (or increases) the
crack length perpendicular to the poling direction
by 2 lm (or about 8 lm), whereas the crack length
parallel to the poling direction remains almost
unchanged. Under the indentation load of 49 N,
however, the electric elds of )5.0 and 5.0 kV/cm
increase the crack length perpendicular to the
poling direction by 14 and about 38 lm and the
crack length parallel to the poling direction by
8 and 14 lm, respectively. The experimental
results indicate that the applied electric eld also
aects the cracking behavior parallel to the poling
direction, if the indentation load is high. A wedge
model [76] was proposed to explain the load-
dependent behavior. Three assumptions made in
the wedge model were: (1) Negative electric elds
caused 180-domain switching in the wedge region
under the indenter; (2) The magnitude of the pie-
zoelectric constant, d
33
, was reduced after domain
switching due to uncompleted switching; and (3)
The greater the indentation load, the more com-
plete the 180-domain switching and the smaller
the reduction in the piezoelectric constant. Thus, a
reduction factor links the wedge force with the
applied electric eld and the reduction factor
changes with the applied mechanical load. The
wedge model in conjunction with the mechanical
strain energy release rate as a failure criterion
successfully explains the load-dependent pheno-
menon [76].
The results of indentation fracture tests with a
39.2 N load on PLZT ceramics [77] show that a
positive electric eld enhances the growth of cracks
perpendicular to the poling direction in PLZT
ceramics. This eect seemed to saturate at 500 kV/
cm, about 1.2 times the coercive eld. When in-
dented under a high electric eld, the ferroelectric
ceramics developed a series of microcracks in the
stress eld of the indentation. The microcracks
were not extensions of the radial or lateral crack
systems. The eect of negative eld was not re-
ported in [77], however.
The indentation fracture tests on PZT-841
ceramics with a load of 49.0 N were conducted in
such a way that under each level of the electric
elds of 4 kV/cm, about 10 tests were performed
and the average values were reported [78]. Fig. 7
shows the variation of K
IC
with the applied electric
eld. Under pure mechanical loading, the average
K
IC
was 1.01 0.06 MPa m
1=2
. The mean value of
K
IC
was reduced by either a positive or a negative
362 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
applied electric eld. A reduction of 0.21 or 0.10
MPa m
1=2
resulted from the respective application
of a negative or positive eld of 4 kV/cm. These
facts indicate that a negative eld had a stronger
inuence on the average K
IC
than a positive
eld did. The eect of electric eld on the frac-
ture behavior of ferroelectric barium titanate was
investigated with the indentation fracture tech-
nique by using a 40 N load [79]. The results indi-
cate that the curve of the measured crack lengths
as a function of the applied electric eld might be
similar to strain hysteresis, i.e., the buttery curve.
Curves of cracks parallel and perpendicular to the
electric eld direction are symmetric to each other.
The crack lengths for cracks perpendicular to the
poling direction were longer under either positive
or negative electric elds than those under no ap-
plied electric eld [79]. At this point, the results
[79] are consistent with the observations in [78].
Atomic force microscopy (AFM) and Kelvin
probe microscopy (KFW) were used to investigate
an indentation crack in a poled PZT ceramic
subjected to an electric eld [39]. The indentation
crack was introduced rst. Then, the crack open-
ing displacement was measured with AFM and the
electric potential dierence across the crack was
determined with KFM. Consequently, the stress
intensity factor, the electric displacement intensity
factor, and the crack-tip energy release rate could
be estimated from the measured crack opening
displacement and the measured electric potential
dierence across the crack. From the applied
electric eld and the measured eld interior to the
crack, the dielectric constant of the crack interior
was determined to be 40. The theoretical results
indicate, as described in Section 3, that the eect of
an electric eld on the fracture behavior of an
electrically insulated crack depends on the value of
k, which is directly linked to the dielectric constant
of the crack interior. In practical situations in
terms of crack length, applied load and electric
eld level, the retardation of crack growth by the
electric eld is negligible when the dielectric con-
stant of the crack interior is higher than 20 [39].
7.1.2. Fracture tests on pre-cracked or pre-notched
samples
Using pre-cracked or pre-notched samples is
more standard than the indentation fracture tests
to measure fracture toughness. Fracture tests were
conducted on pre-notched compact tension (CT)
samples of PZT-4 [80] and PZT-841 ceramics [78].
In the CT tests on the PZT-841 ceramics [78],
about 10 samples were tested at each level of
electric elds, except that 33 samples were tested
at the electric eld of 15 kV/cm to study the dis-
tribution of the fracture toughness. The energy
release rate was calculated with nite element
analysis and then converted into the mode I stress
intensity factor. Fig. 7 shows the variation of K
IC
with the applied electric eld for PZT-4 ceramics
(a) and for PZT-841 ceramics (b). The results from
the PZT-4 ceramics [80] revealed nearly a linear
Fig. 7. The eects of static electric elds on the fracture
toughness measured from pre-notched compact tension samples
of (a) PZT-4 ceramics (with experimental data from Park and
Sun [80]), and (b) PZT-841 ceramics. The fracture toughness is
normalized by its values under purely mechanical loading.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 363
eect of the electric eld on the fracture load.
The results from the PZT-841 ceramics show that
the applied electric eld increases the scattering
of the measured apparent fracture toughness. The
applied electric eld, either positive or negative,
reduces the mean of the apparent fracture tough-
ness [78]. Under purely mechanical loading, the
average K
IC
for PZT-841 was 1.12 0.05 MPa m
1=2
,
being almost the same as the result obtained from
the indentation fracture tests. A negative eld of
7.5 kV/cm reduces the average K
IC
by 0.25
MPa m
1=2
, while the same strength positive eld
reduces the average K
IC
by 0.10 MPa m
1=2
. Apply-
ing a positive electric eld of 15 kV/cm reduces
further the averaged K
IC
to 0.92 0.14 MPa m
1=2
,
resulting in a relative reduction of 18%. Applying
an electric eld generally causes considerable scat-
ter in the measured fracture toughness data. The
largest scatter in the fracture toughness is induced
by an electric eld of +15 kV/cm. Fig. 8 shows the
distribution of K
IC
under the eld of +15 kV/cm for
33 samples. The ratio of the standard deviation
of 0.14 MPa m
1=2
to the associate mean 0.92
MPa m
1=2
leads to a relative error of about 15%. As
described below, the relative error is about 33% for
the bending strength at an electric eld of 10 kV/cm
[81], which is over double the relative error of
about 15% observed in the CT tests.
The electric fracture properties of poled PZT
ceramics P-7 were investigated by single-edge pre-
cracked beam (SEPB) tests [82]. A sharp pre-crack
was introduced to each sample by Vickers inden-
tation, in which the indent diagonals aligned one
by one closely along a line perpendicular to the
poling direction. Then, the indented sample was
compressed to form a sharp pre-crack from the
indents. Finite element calculations were carried
out to determine the stress intensity factor under
mechanical and/or electrical loading, in which the
relationship between the mechanical energy release
rate and the stress intensity factor was employed.
The experimental results indicate that the pre-
crack length monotonically increases with the
indentation load for a given number of indents.
When the number of indents was 11, the pre-crack
lengths
1
were 307, 405, 515, 723, and 994 lm,
corresponding to the indentation loads of 1.96,
2.94, 4.90, 9.80, and 19.6 N. The fracture tough-
ness without any electric eld was found to in-
crease from 0.97 to 1.58 MPa m
1=2
as the pre-crack
lengths increased from 307 to 994 lm. However,
for a given indentation load of 19.6 N, an increase
in the number of intents increased pre-crack
length, but did not change the fracture toughness
of 1.58 MPa m
1=2
without electric eld. A positive
electric eld of 200 kV/m increased the fracture
toughness from 1.58 to 1.76 MPa m
1=2
, while a
negative electric eld of )100 kV/m decreased it to
1.50 MPa m
1=2
. The critical mechanical energy re-
lease rate was not constant either with or without
an electric eld. The experimental results [82] show
that the critical values of the mechanical energy
release rate are 16.3, 18.0 and 22.5 N/m for the
electric elds of )100, 0.0, and 200 kV/m, respec-
tively. The experimental observations, numerical
calculations and theoretical analysis indicate that
the energy release rate criteria derived from the
permeable boundary conditions are superior to the
fracture criteria derived from the impermeable
boundary conditions [82].
Fig. 8. The distribution density of fracture toughness under a
static electric eld of 15 kV/cm for CT tests on pre-notched
PZT-841 samples.
1
The data were copied from Table II in [82] and by changing
the units of meter to millimeter.
364 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
7.1.3. Bending tests
It is convenient to measure the bending strength
of a ceramic. The bending strength of PZT
ceramics also varies with an applied electric eld.
Axisymmetric bending tests were conducted on
ferroelectric PZT-19 disks with a diameter of 20
mm and a thickness of 1 mm, where the poling
direction was along the thickness [83]. The experi-
mental results show that the distribution of the
electrical strength is similar to that of the me-
chanical strength, implying that local sites of fail-
ure in the electric and mechanical elds may be
identical and governed by the microstructure of
the ceramics. Weak electric elds of %10 kV/cm,
either positive or negative, slightly strengthen
the ceramics, whereas strong elds weaken them.
Fig. 9 gives the experimental results on the PZT-19
ceramics with each point representing the average
of 1520 measurements [83]. As can be seen in Fig.
9, there exists a mutual inuence of the electric and
mechanical agents. Three-point bending tests on
PZT samples were carried out under positive and
negative electric elds with the poling direction
towards the jig surface [84]. The results [84] show
that either a positive or a negative electric eld,
even at values less than 10 kV/cm, reduces the
bending strength of PZT ceramics. The eect of
a static electric eld on the bending strength of
PZT-841 ceramics was comprehensively studied
[69,81,85]. In the three-point bending tests, the
poling direction was perpendicular to the jig sur-
face and a static electric eld up to 20 kV/cm was
applied across the sample, either parallel (positive)
or anti-parallel (negative) to the poling direction.
Over 50 samples were tested under each electric
eld [69,81,85]. Fig. 10 illustrates the probability
of the bending strength for both the positive
electric elds (Fig. 10a) and negative electric elds
(Fig. 10b). Fig. 11 plots the mean with its error bar
of the bending strength versus electric eld.
Applying an electric eld scattered the data tre-
mendously. The maximum scattering occurred
under the electric eld of 10 kV/cm. Without any
applied electric eld, the bending strength along
the poling direction, i.e., 88.0 MPa, is compara-
tively smaller than that perpendicular to the poling
direction, i.e., 97.8 MPa, due to the anisotropy of
the material. Under the electric elds of )3.33 and
+3.33 kV/cm, the bending strengths are, respec-
tively, 89.9 and 89.8 MPa, slightly higher than 88.0
MPa. However, the statistical u-test analysis at a
95% signicance level shows that the electric elds
of 3.33 kV/cm do not change the bending
strength. The bending strengths under the applied
elds of )6.7 and +6.7 kV/cm are respectively 81.4
and 77.7 MPa, signicantly lower than that with-
out any electric eld. The bending strength is fur-
ther reduced under a higher electric eld, either
positively or negatively applied, as shown in Fig.
11. According to the experimental results, the do-
main wall kinetics model, as described above, was
developed [69] to understand the failure behavior.
The experimental data agree well with the predic-
tions from the domain wall kinetics model, as
shown by the solid curve in Fig. 11.
7.2. Electrically conducting cracks
Fig. 12 schematically shows the similarity be-
tween a conductive crack under electrical loading
and a normal crack under mechanical loading. To
ensure that the electric eld inside the conductive
crack remains zero, electric charges in the con-
ductive crack surfaces must re-arrange themselves
to produce an induced eld that has the same
magnitude as the applied one but with the opposite
Fig. 9. The eects of static electric elds on the bending
strength of PZT-19 (with experimental data from Zhoga and
Shpeizman [83]).
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 365
sign. As a result, the charges in the upper and
lower crack surfaces near the crack tip have the
same sign, as shown in Fig. 12. The charges with
the same sign repel each other and then have a
tendency to propagate the crack. The contour-
independent J-integral used in fracture mechanics
can also apply to conductive cracks [2] and the J-
integral results for the purely mechanical and
electrical loadings are illustrated in Fig. 12.
However, there are few reports about experi-
mental observations on conductive cracks in pie-
zoelectric ceramics. The electro-mechanical fracture
toughness of conductive cracks in PZT-PIC ceram-
ics was investigated by four-point bending tests on
pro-notched bars, in which the poling direction
was toward the jig surface and an NaCl solution
was lled into the notch to make the crack con-
ducting [86]. Wide scattering was found under a
large applied electric eld of jK
E
j > 50 kV/m
1=2
. It
seems that the critical stress intensity factor in-
creases, as the intensity factor of electric eld
strength changes from 30 to )90 kV/m
1=2
. When the
electrical intensity factor is in the range of )15 to 15
kV/m
1=2
, a model based on domain switching near
the crack tip can explain the experimental results
[86].
Fig. 11. Dependence of the average bending strength of PZT-
841 on static electric elds. The solid curves are based on the
domain wall kinetics (DWK) model.
Fig. 10. Probability distribution of the bending strength of PZT-841 ceramics under (a) positive electric elds and (b) negative electric
elds. Over 50 samples were tested under each static electric eld level.
366 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
Fracture tests of electrically conductive cracks
were conducted on depoled PZT-4 ceramics [87].
The depoling was achieved by thermally annealing
the PZT-4 ceramics at 400 C for 40 min and then
cooling them down in a furnace to room tempera-
ture. After depoling, the ceramics lost their piezo-
electricity and became macroscopic dielectrics that
were isotropic in elasticity and dielectricity. No
detectable piezoelectric constants in the depoled
ceramics were measured with the piezo d
33
tester
with 1% accuracy (Pennebaker, Model 8000). The
isotropic dielectric constants were measured using
an impedance analyzer (Hewlett Packard, 4192A
Lf) at the standard frequency of 1 MHz. The mean
value of the dielectric constants with the standard
deviation was j 9:794 0:205 10
9
F/m. The
Youngs modulus and the Poisson ratio were
measured by compressive tests using strain gauges
on a universal testing machine (MTS Sintech 10/D).
Fig. 13 shows typical stressstrain curves obtained
from the compressive tests. Clearly, the loading and
unloading curves are all non-linear, showing hys-
teresis loops. However, under small loads, the
stress was linearly proportional to the longitudinal
and lateral strains, from which we determined the
Youngs modulus and the Poisson ratio. The
average Youngs modulus and the average Poisson
ratio from four loading curves were Y 51 GPa
2
and m 0:43, respectively, for the depoled ceram-
ics. Since it is dicult to make pre-cracks in ceramic
samples, pre-notched compact tension (CT) sam-
ples were used in the fracture tests under purely
mechanical, purely electrical and mixed mechanical
and electrical loads. All samples had widths of
2w
1
10 mm, heights of w
0
10 mm and thick-
nesses of d 3 mm. A pre-notch or crack was cut
in each sample with a 0.15 mm thick diamond saw
2a
conductive crack
2a

a C J =
M
M 2
E
a E C J =
E
E 2
Fig. 12. A comparison of a normal crack loaded by uniform mechanical stress, r, and a conductive crack loaded by uniform electrical
eld, E, where J
M
and J
E
denote, respectively, the mechanical and electrical J-integrals, C
M
p=Y

, Y

Y for plane stress and


Y

Y =1 m
2
for plane strain, and C
E
pj=2 with Y , m, and j being Youngs modulus, Poisson ratio, and dielectric constant,
respectively.
0.000 0.001 0.002 0.003 0.004 0.005 0.006
0
50
100
150
200
250
300
4.0x10
-4
8.0x10
-4
1.2x10
-3
1.6x10
-3
2.50x10
-4
5.00x10
-4
7.50x10
-4

Strain -
11
S
t
r
a
i
n

-

2
2
S
t
r
e
s
s

1
1

(
M
P
a
)
Strain
11
(m/m)
2
4
3
1
Fig. 13. Typical stressstrain curves, r
11
versus e
11
, ob-
tained from the compressive test, in which two cycles of loading
and unloading were conducted, where the subscripts 11 and
22 denote the direction parallel with and perpendicular to the
applied load direction, respectively, and the arrows and the
number nearby show the loading history, 1234. The Youngs
modulus was determined from the loading curves under small
loads. The inset gure shows the variation of e
22
with e
11
during loading under small loads, which slope yields the Pois-
son ratio.
2
The values of the elastic constant and the Poisson ratio,
Y 25:3 0:34 GPa and m 0:38 0:006, reported in [87]
were incorrect.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 367
and further sharpened with a 0.12 mm diameter
wire saw, leading to a notch radius of 60 lm at the
notch tip. The total notch length varied from
sample to sample and ranged from 4.8 to 6.8 mm,
which corresponded to the ligament, l, ranging
from 5.2 to 3.2 mm. Comparing the notch length
with the notch radius reveals that the notches are all
deep and sharp. After the cutting, the samples were
cleaned ultrasonically in distilled water for 2 min.
To create conductive cracks, silver paint was lled
into the notch (crack) to make it function as an
electrode. Under purely electric loading, a static
voltage parallel to the crack direction was applied
to the sample and manually increased until the
sample failed. To avoid electric sparking, the
samples were put in silicone oil during the electrical
fracture test. To compare the electrical toughness
with the mechanical toughness, nominal fracture
tests were carried out under purely mechanical
loading. In the tests, the positive pole of the power
supplier was always connected to the notch elec-
trode of the sample. All tests were conducted at
room temperature and 58 and 44 samples were
respectively tested under purely electrical and
mechanical loads. The critical mechanical loads
and electrical voltages at failure were recorded to
calculate and the critical value of the stress intensity
factor, the critical value of the intensity factor of
electric eld strength, which was called here the
electric intensity factor, and the critical values
of the energy release rate. The stress intensity fac-
tor and the electric intensity factor were numeri-
cally calculated with nite element analysis
based on the linear theory as described in Section 2.
From fracture mechanics, the critical value of
the mode I stress intensity factor at fracture
under the plane-strain condition is called the frac-
ture toughness. For the used CT samples of the
PZT samples, mechanical loading is in mode I
and the plane-strain condition is satised. Thus,
the critical value of the stress intensity factor
under purely mechanical loading is termed the
mechanical fracture toughness. Similarly, the crit-
ical value of the electric intensity factor under
purely electrical loading is termed the electrical
fracture toughness. Clearly, more work is needed to
standardize successfully the electrical fracture
toughness.
Fig. 14 shows the critical stress intensity factor
as a function of the crack length, where a solid
circle represents an experimental datum and here-
after the same symbol is used in the following
gures without notation. The mean and stan-
dard deviation of the critical stress intensity factor
is K
0
r;C
0:946 0:062 MPa m
1=2
under purely
mechanical loading. The value of K
0
r;C
0:946
MPa m
1=2
is slightly higher than the fracture
toughness, K
0
r;C
0:9 MPa m
1=2
, of PZT-PIC ce-
ramics obtained by the four-point bending tests
[86] and the fracture toughness, K
0
r;C
0:934
MPa m
1=2
, of the poled PZT-4 ceramics [88], but
slightly lower than the fracture toughness, K
0
r;C

1.11.7 MPa m
1=2
, of lead zirconate titanate ce-
ramics reported in [89].
An applied electric eld could fracture the
depoled ceramic samples with conductive deep and
sharp notches. The fracture was accompanied by
dielectric discharging. Fig. 15 shows a typical
example of the fracture of a conductive deep
notch. When the deep notch was free of silver
paint, the sample could survive an electric voltage
of 20 kV, which was the maximum voltage that the
used power suppler could provide. After lling the
notch with silver paint, the notch became electri-
cally conductive. Then, an applied electric eld of
about 12 kV fractured the sample. The crack
propagated along the notch or pre-crack direction,
4.6 4.8 5.0 5.2 5.4 5.6 5.8 6.0
0.6
0.7
0.8
0.9
1.0
1.1
1.2
K
0



(
M
P
a

m
0
.
5
)
Notch length (mm)

,
C
S
t
r
e
s
s

i
n
t
e
n
s
i
t
y

f
a
c
t
o
r
Fig. 14. The mechanical fracture toughness versus the notch
length.
368 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
as shown in Fig. 15. As described above, the
charges in the upper and lower surfaces of a con-
ductive crack have the same sign. The charges with
the same sign repel each other and generate a
Columbic force that opens and propagates the
conductive crack. For the CT samples of the brittle
ceramics, the fracture process is unstable because
the intensity factors increase monotonically with
the crack length, which means that failure will
occur once the crack propagation is triggered
mechanically and/or electrically. The fracture pro-
cess and the dielectric breakdown may occur
simultaneously in the CT samples, even if the
charge emission from the conductive notch (crack)
tip may happen prior to the process of fracture and
dielectric breakdown. It is the sample geometry and
the loading condition that makes the failure be-
come two-dimensional. The experimental results
all reveal the two-dimensional failure behavior, as
shown in Fig. 15.
The experimental results under purely electrical
loading are plotted in Fig. 16, showing the rela-
tionship of the critical electric intensity factor as a
function of the crack length. The mean and stan-
dard deviation of the electric fracture toughness
is K
0
E;C
218:34 19:48 kV/m
1=2
for the depoled
ceramics. Like the mechanical fracture toughness,
the electric fracture toughness is a material con-
stant independent of the crack length and thus can
serve as a failure criterion for conductive cracks
(or deep notches) in dielectrics under purely elec-
trical loading. The relative error in the experi-
mental data of the electric fracture toughness is
about 9%, which is higher than that of the
mechanical fracture toughness of about 7%. The
experimental errors may be attributed to the fol-
lowing reasons: (1) the notch is not sharp enough,
(2) there are many defects such as grain boundaries
and pores in the PZT ceramics fabricated by sin-
tering PZT powders, and (3) there may exist elec-
tric defects in the PZT ceramics, which cause
partially electric discharging, etc.
To compare the electrical fracture toughness
with the mechanical fracture toughness, the frac-
ture toughnesses are converted to the critical values
of the energy release rates by using the relation-
ships of G
M
K
r

2
=Y

and G
E
jK
E

2
=2, where
Y

Y 1 m
2
for plane strain and Y

Y for
plane stress, Y and m denote Youngs modulus and
the Poisson ratio, respectively, k is the dielectric
constant of the material. The mean value of the
critical energy release rate under purely electric
loading, G
E;C
233:5 N/m, is about 16.3 times
higher than that, G
M;C
14:3 N/m, under purely
mechanical loading, which is attributed to electrical
plastic deformation, such as electrical discharge,
etc., that occurs at the tip of the conductive notch
forming an electrical plastic zone. Actually,
Fig. 15. A typical optical microscopic picture of the fractured
CT sample under purely electrical loading, showing the two-
dimensional failure behavior.
4.0 4.5 5.0 5.5 6.0 6.5 7.0
120
140
160
180
200
220
240
260
280
E
l
e
c
t
r
i
c

i
n
t
e
n
s
i
t
y

f
a
c
t
o
r

K
0 E
,
C


(
k
V
m
-
0
.
5
)
Notch length (mm)
Fig. 16. The electric mechanical fracture toughness versus the
notch length.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 369
dielectric breakdown accompanies the fracture
under purely electrical loading. The fracture sur-
faces are at on the samples fractured under mech-
anical loading, while electrical loading yields rough
fracture surfaces [87]. In the electrically fractured
samples, the discharge may locally burn the sam-
ple. It is the electrical plastic deformation that con-
sumes more energy and thus leads to the high
resistance to fracture under electrical loading.
Furthermore, fracture tests on conductive cracks
were conducted on poled PZT-4 piezoelectric ce-
ramics and the poling direction was parallel to the
pre-notch [88] with compact tension samples pre-
pared in the same way as described above. The tests
were conducted under (1) purely mechanical load-
ing and (2) purely positive electric loading. Similar
results to those for the depoled PZT-4 ceramics
were obtained for the poled PZT-ceramics.
As a natural extension, the fracture tests on CT
samples of depoled PZT-4 ceramics were carried
out under combined electric and mechanical
loading [90]. When the critical stress intensity
factor, K
a
r;C
, is normalized by the critical stress
intensity factor, K
0
r;C
, under purely mechanical
loading, and the critical electric intensity factor,
K
a
E;C
, is normalized by the critical electric inten-
sity factor, K
0
E;C
, under purely electric loading, the
relationship of the normalized electric intensity
factor versus the normalized stress intensity factor
is plotted in Fig. 17. The experimental data can be
approximately described by the equation of
K
a
r;C
K
0
r;C
_ _
2

K
a
E;C
K
0
E;C
_ _
2
1; 213
which is shown as the curve in Fig. 17. Despite of
the data scattering, the experimental results indi-
cate that Eq. (213) serves as a failure criterion for
the depoled PZT ceramics with conductive cracks
under electrical and/or mechanical loading. A
charge-free zone (CFZ) model is developed to ex-
plain the experimental observations [90]. The CFZ
model will be introduced in Section 8.
7.3. Other related tests
To investigate the delayed fracture behavior of
PZT-5 ceramics with a Zr/Ti ratio of 52/48 at a
constant load under dierent environments, a
series of environment-induced fracture experi-
ments were conducted on single-edge notched
tensile specimens with each having the dimensions
of 0.9 8 40 mm
3
[91,92]. The samples were poled
along the thickness or length direction under an
electric eld of 3.0 kV/mm. A pre-crack or pre-
notch was cut in each sample perpendicular to the
length direction, along which a mechanical load
was applied. The experimental results show that
stress corrosion cracking (SCC) occurs when
the PZT-5 ceramics are in moist air, silicon oil,
methanol and formamide. When the poling was
along the thickness direction, i.e., the crack sur-
faces were parallel to the poling direction, the
normalized threshold stress intensity factor,
K
k
ISCC
=K
k
IC
, was 0.66 in water, 0.73 in methanol, or
0.75 in formamide, where K
k
IC
1:34 0:25
MPa m
1=2
was the fracture toughness of the PZT-5
ceramics. When the poling was along the length
direction, i.e., the crack surfaces were perpendic-
ular to the poling direction, the normalized
threshold stress intensity factor, K
?
ISCC
=K
?
IC
, was
0.46 in water, 0.59 in formamide, 0.49 in silicon
oil, or 0.64 in moist air, where K
?
IC
0:92 0:10
MPa m
1=2
, demonstrating that the failure behavior
0.0 0.5 1.0 1.5
/
K
a E
,
C
0 E
,
C
0.0
0.5
1.0
1.5
N
o
r
m
a
l
i
z
e
d

e
l
e
c
t
r
i
c

i
n
t
e
n
s
i
t
y
f
a
c
t
o
r

K
/K
a
,C
0
,C
Normalized stress intensity
factor K
Fig. 17. Experimental results for the failure of electrically
conductive cracks in the depoled PZT-4 ceramics under com-
bined mechanical and electrical loading.
370 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
is anisotropic. The anisotropy factor of the frac-
ture toughness was K
k
IC
=K
?
IC
1:4 for the PZT-5
ceramics, whereas the anisotropy factors of
K
k
ISCC
=K
?
ISCC
were 2.1 in water and 1.8 in form-
amide. The domain-preferential orientation along
the poling direction in the poled ceramics might
cause the anisotropy of the K
IC
and K
ISCC
and the
anisotropy in the resistance of the PZT-5 ceramics
against SCC. In addition to SCC, hydrogen-
induced delayed fracture (HIC) occurred in elec-
trically hydrogen-charged PZT-5 ceramics [93]. To
charge hydrogen electrically into the PZT-5
ceramics, a nickel layer was plated on the samples.
The HIC test was carried out during dynamic
hydrogen charging under a sustained load. The
hydrogen concentration as a function of the
charging current density was determined by col-
lecting hydrogen gas escaped from a sample at
70 C after 100 h of charging at room temperature.
The results show that the threshold stress intensity
factor decreases linearly with the logarithm of
the hydrogen concentration. The threshold stress
intensity factor of hydrogen-induced delayed
fracture reveals anisotropy, i.e., K
k
IH
> K
?
IH
. How-
ever, the normalized threshold stress intensity
factor of HIC revealed an isotropic behavior, i.e.,
K
k
IH
=K
k
IC
K
?
IH
=K
?
IC
.
The anisotropic fracture behavior was also ob-
served in measurements of the fracture resistance
curves (R-curves) of PZT-PIC 151 ceramics with
CT specimens for dierent poling directions and
grain sizes [94]. The CT samples had dimensions of
50 48 3 mm
3
and the poling direction was along
the thickness, the crack growth or the loading
direction. For the ceramics with an average grain
size of 6.4 lm, the resistances to fracture at the
plateaus of the R-curves were 1.54, 1.22, and 1.13
MPa m
1=2
for the poling along the thickness, the
crack growth and the loading directions, respec-
tively, whereas the value at the R-curve plateau
was 1.30 MPa m
1=2
in the unpoled ceramics.
Compact-tension samples were also used to
investigate the eect of an electric eld on the R-
curves of BaTiO
3
and PZT-PIC 151, where the
electric eld was applied along the thickness
direction [95]. The results indicate that the electric
eld increases the starting and plateau values
of the R-curve and makes the curve shorter. A
domain-switching model can successfully explain
the observed increase to fracture resistance phe-
nomenon [95].
Failure processing is in principle dependent on
time, although little work on the time-dependent
failure process has been reported. Time-dependent
cracking in PZT-5 ceramics was reported under a
sustained electric eld or mechanical load [96]. Fig.
18 illustrates the sequence of the time-dependent
cracking under a sustained electric eld of 6 kV/
cm, where (a) and (b) show an indentation crack,
Fig. 18. Initiation and propagation of cracks under a constant
electric eld [96]. (Permission granted by the American Institute
of Physics.)
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 371
(c) is the image taken after 2 h showing no detected
changes, (d) shows the crack propagation from
point A to point B after 10 h of electric loading, (e)
indicates the crack propagation from point B to
point C in the following 1 h, and (f) illustrates a
microcrack appearing in the direction normal to
the electric eld. There was a threshold eld for the
delayed fracture. The magnitude of the threshold
eld was about half of the coercive eld of the
PZT-5 ceramics [96].
The electric-eld-induced fatigue crack growth
in ferroelectric ceramics is also observed [97] under
cyclic electric elds. The experimental results show
that the electric-eld-induced fatigue displays dis-
tinct characteristics under dierent amplitudes of
the cyclic electric loading. The crack growth rate is
non-linearly related to the cyclic electric load and
there exists a threshold for the crack growth. The
crack did not grow when the amplitude of the
applied cyclic eld was lower than the threshold
eld. The experimental results give the threshold
eld, E 0:797E
c
, where E
c
denotes the coer-
cive eld [97].
The eld-driven in situ transmission electron
microscopy (TEM) technique was developed to
examine micromechanisms of eld-induced crack-
ing in ferroelectric ceramics under cyclic electric
elds [98]. Microcracks were found to initiate from
pores at the triple junctions and propagate along
the grain boundary in a lead zirconate titanate
ceramic. The crack growth rate followed a power-
law dependence on incompatibility-stress intensity
[98]. The TEM observations might suggest that
dielectric breakdown of the grain boundary phase
was conducive to grain boundary cavitation and
the crack growth followed the progressive accu-
mulation of cavity density with electric cycling.
The piezoelectric eect causes an electroded
region to expand or contract under an applied
electric eld, while the external or unelectroded
region will constrain the deformation of the elect-
roded region, thereby generating strain incom-
patibilities and a stress eld. The electrically
induced stress eld is similar to a thermal stress
eld, but the electrically induced stress eld is
much easier to control than a thermal stress eld.
The electrically induced stress eld in a pre-
cracked CT sample could cause crack deection,
which was observed in situ with an optical
microscope [99,100]. Under electric loading, small
electrode widths leads to straight cracks with
two transitions between stable and unstable crack
growth, while large electric widths result in curved
cracks with four transitions. Poling the sample
prior to the experiment altered the crack path and
introduced an anisotropy in the R-curve behavior
as well as in the achievable strain mismatch
[99,100].
In summary, the experimental results demon-
strate that the failure behavior of piezoelectric
ceramics is complex, especially under combined
mechanical and electrical loads. The experimental
results on electrically insulated cracks are not
consistent. The major reason is the great scatter of
experimental data, which is caused by the defects
in the tested ceramics and by multiple mechanisms.
In addition to the non-linearities in the relation-
ships of polarization versus electric eld and strain
versus electric eld, the stressstrain curves are
non-linear during both loading and unloading, as
shown in Fig. 13. The non-linear loading and
unloading curves indicate domain-switching and
reverse switching. The domain-reverse switching
makes it more challenging to develop a domain-
switching-based model of fracture. Furthermore,
dielectric breakdown may occur locally in front of
an electrically insulated crack, as described in the
dielectric breakdown model. Electrical discharge
may occur within an insulated crack, which makes
the crack become electrically conductive. All the
failure mechanisms may function jointly in a
fracture test. Therefore, well-designed and well-
controlled experiments are required to clarify the
role of electric eld in the failure behavior. On one
hand, the degree of the data scatter in experiments
should be reduced. On the other hand, sucient
tests must be repeated to have a large sampling
number such that statistical analysis can be carried
out to provide reliable experimental results.
The preliminary results on electrically conduc-
tive cracks (or notches) are encouraging. They
evidence the existence of electric fracture toughness
under purely electrical loading. The signicance of
the existence of electrical fracture toughness is that
it enables people to use the concepts from frac-
ture mechanics to understand electrically induced
372 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
failures, and it provides designers of electronic and
electromechanical devices with a useful material
property. The advantage of applying the concepts
of fracture mechanics to dielectric failure lies in the
ability to predict the critical electric eld at which
a dielectric ceramic material containing a con-
ductive crack or an internal electrode fails. The
critical electric eld is a function of the crack
dimension or the length of the electrode, while the
electrical fracture toughness is a material con-
stant. Thus, one can predict the critical electric
eld when information on the sample geometry
and the electric fracture toughness is available.
8. The charge-free zone model
Recently, a charge-free zone (CFZ) model [90]
was proposed to understand the failure behavior
of conductive cracks in dielectric ceramics under
electrical and/or mechanical loading. The CFZ
model treats dielectric ceramics mechanically brit-
tle and electrically ductile. Charge emission and
charge trapping were assumed to occur in the CFZ
model, which consume more work and thus lead to
a high value of the electric toughness. In the CFZ
model, the local electric intensity factor has a non-
zero value and consequently there is a non-zero
local electric energy release rate, which contributes
to the driving force to propagate the conductive
crack. The merit of the CFZ model lies in the
ability to apply the Grith criterion directly to
link the local energy release rate to the fracture
toughness in a completely brittle manner. As a
result, an explicit failure criterion results from the
CFZ model to predict the failure behavior of
conductive cracks in dielectric ceramics under
electrical and/or mechanical loading and the theo-
retical predictions agree perfectly with the experi-
mental observations [90].
The CFZ model is based on the eld limiting
space charge (FLSC) model [101] and analogy with
the dislocation-free zone (DFZ) model [42,102
106] and the tip emission adjusted zone (TEAZ)
model [107109] in the plastic fracture mechanics.
In the FLSC model [101], the charge mobility had
only two extremes. If the electric eld, E, is lower
than a critical value of E
c
, the value of the charge
mobility is assumed to be zero in the dielectric
material, whereas the charge mobility has a nite
value when E > E
c
. Based on the assumption used
in the FLSC model, the level of the electric eld
remains the critical value in the space charge re-
gion, thereby allowing one to calculate the space
charge distribution. The electric eld at the tip of
an electrically conductive crack is extremely high
and theoretically approaches innity. That is why
an intensity factor of electric eld strength, i.e., the
electric intensity factor, is adopted to gauge the tip
eld. When the electric intensity factor reaches a
critical value, charges could be emitted from the
tip. Various emission mechanisms, such as the
Schottky emission [110] and the FowlerNordheim
emission [111], may function jointly at the tip. The
emitted charges may form a charge cloud around
the tip and thus shield the tip from the applied
electric eld. At the onset of the failure of an
electrically conductive crack in a dielectric body,
the charge cloud should reach a critical level. In
the proposed charge-free zone model, this critical
level of the charge cloud was investigated based
on the concepts of fracture mechanics. To simplify
the analysis, charges are treated as line charges and
the charge cloud is modeled as a charge strip in the
CFZ model.
8.1. A single line charge near an electrical conduc-
tive crack
Consider a single line charge near an electrical
conductive crack rst. As described above, the
electric potential u can be expressed by the imag-
inary part of a complex potential U that is an
analytic function of z x iy, i.e.,
/ ImUz: 214
Thus, the electric eld is given by
iE
x
iE
y
U
0
z: 215
For a line charge located at z
d
in an innite
dielectric medium, the complex potential is given
by
U
0


B lnz z
d
;

B iB and B
q
2pj
;
216
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 373
where q denotes the line charge per unit length and
j is the dielectric constant. Consider a conducting
crack with a semi-innite length occupying x < 0
in a dielectric. The electric boundary condition
requires
E
x
0 for x < 0 217
along the crack faces. When a line charge is near
the conducting crack, the complex potential satis-
fying the boundary conditions takes the following
form
U

B ln

z
p


z
d
p


B ln

z
p

z
d
p
; 218
where the overbar denotes the conjugate of a
complex variable. This line charge produces an
electric eld, which is given by
E
x
iE
y

B
2

z
p
1

z
p


z
d
p
_

1

z
p


z
d
p
_
: 219
When the line charge is located on the x-axis, Eq.
(219) is reduced to
E
x
iE
y

q
2pj

x
d
p

z
p
z x
d

: 220
The electric intensity factor is dened as
K
E
lim
z!0

2pz
p
E
x
: 221
From the denition, the tip eld can be expressed
as
E
x
iE
y

K
E

2pz
p : 222
Substituting Eq. (220) into Eq. (221) yields the
electric intensity factor induced by the line charge,
K
E

q
j

2px
d
p : 223
For the line charge near an electrical conductive
crack, the applied tip eld and the image eld exert
forces on the line charge. The image eld is cal-
culated by taking away the corresponding eld of
the same line charge in an innite body from the
eld of the line charge with the crack, which is
given by
E
x;i
iE
y;i

B
2

z
p
1

z
p


z
d
p
_

1

z
p


z
d
p
_
:
224
When the line charge is located on the x-axis at x
d
,
the image force per unit length is calculated from
f
i
qE
i
. Letting z x
d
in Eq. (224) gives
f
i

q
2
2pj
1
2x
d
: 225
This image force always has the tendency to push
the charge back towards the crack. On the other
hand, the applied tip eld exerts a driving force, f
a
,
per unit length on the line charge, which is given by
f
a

K
E
q

2px
d
p : 226
The sign of f
a
must be positive to emit a charge
from the crack, thereby indicating that a positive
(or negative) value of K
E
will have the tendency to
emit a positive (or negative) charge. Furthermore,
the driving force must be larger than the image
force in order to emit a charge from the tip. For a
given applied K
E
, however, there exists a critical
distance from the crack tip, as shown by x
0
in Fig.
19. When the distance from the tip is smaller than
x
0
, the image force dominates, while the driving
force dominates if the distance is larger than x
0
.
With the FLSC model, a charge moves forward in
the region of x
1
x
2
, as shown in Fig. 19, because
the total electric eld is higher than E
c
in the re-
gion. Thus, when a charge is emitted from the tip,
it must be emitted to a distance larger than x
1
.
Then, the charge moves forward until it reaches x
2
,
beyond which the total eld is lower than the
critical level of E
c
and the charge mobility becomes
zero. The analysis indicates that, microscopically,
a charge-free zone is formed adjacent to the tip in
f
x
f
i
~-1/(2x)
f
a
~1/x
0.5
f
a
+ f
i
f
c
x
0
x
1
x
2
f
c
=qE
c
Conductive
crack
o
Fig. 19. A single line charge in front of an electrically con-
ductive crack subjected to the image force and the tip driving
force.
374 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
the charge emission process. The situation is sim-
ilar to the dislocation emission from a crack tip
[42,102109].
8.2. The charge-free zone model
In addition to the image force and the driving
force, the interaction force between charges must
be taken into account for many line charges. When
more and more charges are emitted from the crack
tip, these charges will entrap in the region of ba, as
shown in Fig. 20, where ob denotes the CFZ size. If
we dene f x
0
to be the charge distribution func-
tion, the number of line charges located at x
0
in the
interval dx
0
is f x
0
dx
0
. The equilibrium condition
that the electrical eld E
x
equals the critical value,
E
c
, in the charge trap zone is described by
K
a
E

2px
p B
_
a
b
f x
0

x
0
p

x
p
x x
0

dx
0
E
c
; b 6x 6a:
227
The rst term on the left in Eq. (227) represents the
applied electric eld, while the second term on
the left stands for the electric eld induced by the
electric charges. The uniqueness relationship for a
distribution f x
0
, which has zero value at b and a,
is given by [42]
K
a
E
2

2pa
p
E
c
E
p
2
; k
_ __
p; 228
where E
p
2
; k is the complete elliptic integral of the
second kind and k

1 b=a
_
. Thus, the solution
to Eq. (227) is given as
f x
0

2E
c
b
p
2
B

a x
0
x
0
x
0
b

P
p
2
;
x
0
a b
ax
0
b
; k
_ _
;
229
where P
p
2
; n
2
; k is the complete elliptic integral of
the third kind. The electrical charges produce an
electric intensity factor, which is calculated by
K
i
E

2p
p
B
_
a
b
f x
0

x
0
p dx
0
2

2
p
_
E
c

a
p
E
p
2
; k
_ _ _

b
p
F
p
2
; k
_ __
;
230
where the superscript i denotes the charges, and
F
p
2
; k is the complete elliptic integral of the rst
kind. The local electric intensity factor is the sum of
the applied electric intensity factor plus the electric
intensity factor induced by the charges, i.e., K
l
E

K
a
E
K
i
E
. Using Eqs. (230) and (228) results in
K
l
E
XK
a
E
; 231
X

b
a
_
F
p
2
;

1
b
a
_ _ _
E
p
2
;

1
b
a
_ _ _ : 232
If the ratio of b=a is assumed to be a constant, then
the parameter X is a constant. Eq. (231) indicates
that the local electric intensity factor is linearly
proportional to the applied electric intensity fac-
tor. The local energy release rate for a dielectric
ceramic is expressed in terms of the intensity fac-
tors as
G
l
K
a
r

2
=Y


j
2
K
l
E

2
: 233
Under purely mechanical loading or purely elec-
trical loading, an application of the Grith crite-
rion to Eq. (233) yields
G
l
C

K
0
r;C

2
Y

C; 234
G
l
C

jK
l
E

2
2

jK
0
E;C

2
X
2
2
C: 235
As described above, both K
0
r;C
and K
0
E;C
are deter-
mined experimentally. Eq. (234) indicates that the
E
E
c
o
b
a
x
Fig. 20. The eld distribution in front of a conductive crack,
wherein ob is the size of the charge-free zone and ba denotes the
charge zone.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 375
value of C is identical to the mechanical toughness,
G
M;C
, which was G
M;C
14:3 N/m for the PZT-4
ceramics. Then, the value of X is calculated from
Eq. (235) with the value of K
0
E;C
, i.e.,
X
2

2C
jK
0
E;C

2

2K
0
r;C

2
jY

K
0
E;C

2
: 236
Using the experimental data, the value of
X 0:245 is determined for the depoled PZT-4
ceramics mentioned earlier. Consequently, the
value of the ratio of b=a is calculated from Eq.
(232) to be b=a 0:0028
3
.
If the critical value E
c
is available, the value of a
can be evaluated from Eq. (228) and then the value
of b. The size of the charge zone is dened as ba in
the present model. It is the size of the charge
zone that represents the degree of electric plastic
deformation. The larger the size of the charge
zone, the more severe the electric plastic defor-
mation is. As discussed above, the value of C is
determined from the mechanical fracture tough-
ness and the CFZ size is determined from the
electric fracture toughness. Next, applying the
Grith criterion to Eq. (233) yields
G
l
C
K
a
r;C

2
=Y


j
2
K
l
E;C

2
C: 237
The subscript C is labeled to indicate the fracture
equilibrium condition. Then, using Eqs. (231),
(234), (235), and (237) leads to the failure crite-
rion, Eq. (213), which has been observed experi-
mentally.
9. Concluding remarks
The present review article is a continuation of
the overview article published in 2002 [2]. Since
many publications before the submission of the
previous overview in the Fall 2000 are summarized
in the previous overview, most of them are not
included in the present review article. Readers may
refer to the previous overview for detail, especially
for the topics on thermodynamic approaches,
interface cracks and three-dimensional problems.
As mentioned in Section 1, the matrix and vector
notations used here are similar to those used in the
book [26], but dierent from the notations used in
the previous overview. Hopefully, the changes in
the notation help readers to easily get into the
anisotropic approach of electro-elasticity from
the anisotropic approach of elasticity. Meanwhile,
the changes in the notation will not be too much
inconvenient for readers of the previous overview.
In addition to summarizing experimental results
published recently, some important experimental
results are reported again, even the experimental
results are not consistent with one another. The
purpose of doing this is to remind readers that the
fracture behavior is complex. More careful and
well-designed and well-controlled experiments are
required to clarify the mechanisms of the failure
behavior of piezoelectric ceramics under combined
mechanical and electrical loading. Mechanical
incompatibilities induced by a defect in a purely
elastic solid produce an internal stress eld [112].
Similarly, electrical incompatibilities induced by a
defect in a piezoelectric ceramic will produce an
internal electric eld and/or an internal stress eld.
The strip dielectric breakdown model and the
charge-free zone model are developed based on the
mechanical and electrical incompatibilities. Study-
ing internal electric and mechanical elds produced
by various defects will create insights into the
failure mechanisms of piezoelectric ceramics.
Acknowledgements
This work was supported by a grant
(N_HKUST602/01) from the Research Grants
Council of the Hong Kong Special Administrative
Region, China. TYZ thanks the Croucher Foun-
dation for the Croucher Senior Research Fellow-
ship Award, which gave him more research time
by releasing him from teaching duties.
References
[1] N. Setter, R. Waser, Electroceramics materials, Acta
Mater. 48 (2000) 151178.
[2] T.Y. Zhang, M.H. Zhao, P. Tong, Fracture of piezoelec-
tric ceramics, Adv. Appl. Mech. 38 (2002) 148289.
3
The b=a value of 0.087 reported in [90] was incorrect.
376 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
[3] M. Kamlah, Ferroelectric and ferroelastic piezoceramics:
modeling of electromechanical hysteresis phenomena,
Continuum Mech. Thermodyn. 13 (2001) 219268.
[4] D.A. Hall, Nonlinearity in piezoelectric ceramics, J.
Mater. Sci. 36 (2001) 45754601.
[5] G.C. Sih, J.Z. Zuo, Multiscale behavior of crack initia-
tion and growth in piezoelectric ceramics, Theor. Appl.
Fract. Mech. 34 (2000) 123141.
[6] C.Q. Ru, S.X. Mao, Eect of microcracking on electric-
eld-induced stress intensity factors in dielectric ceramics,
Philos. Mag. 83 (2003) 277294.
[7] X.F. Li, Electroelastic analysis of an internal interface
crack in a half-plane consisting of two bonded dissimilar
piezoelectric quarter-planes, Mechanica 38 (2003) 309
323.
[8] M. Liu, K.J. Hsia, Interfacial cracks between piezoelectric
and elastic materials under in-plane electric loading, J.
Mech. Phys. Solids 51 (2003) 921944.
[9] T.M. Michelitsch, H.J. Gao, V.M. Levin, Dynamic
Eshelby tensor and potentials for ellipsoidal inclusions,
Proc. Roy. Soc. Lond. A 459 (2002) 863890.
[10] L.A. Filshtinskii, M.L. Filshtinskii, Green function for a
composite piezoceramic plane with an interfacial crack,
Prikl. Mat. Mekh. 58 (1994) 159166 (in Russian).
[11] X.F. Wu, S. Cohn, Y.A. Dzenis, Screw dislocation
interacting with interfacial and interface cracks in piezo-
electric bimaterials, Int. J. Eng. Sci. 41 (2003) 667682.
[12] B.L. Wang, Y.W. Mai, Thermal shock fracture of
piezoelectric materials, Philos. Mag. A 83 (2003) 631657.
[13] C.Y. Li, G.J. Weng, Yoe-type moving crack in a
functionally graded piezoelectric material, Proc. Roy.
Soc. Lond. A 458 (2002) 381399.
[14] J.Z. Zuo, G.C. Sih, Energy density theory formulation
and interpretation of cracking behavior for piezoelectric
ceramics, Theor. Appl. Fract. Mech. 34 (2000) 1733.
[15] S. Lin, F. Narita, Y. Shindo, Comparison of energy
release rate and energy density criteria for a piezoelectric
layered composite with a crack normal to interface,
Theor. Appl. Fract. Mech. 39 (2003) 229243.
[16] V.Z. Parton, B.A. Kudryavtsev, Electromagnetoelasticity,
Gordon and Breach Science Publishers, 1988.
[17] A. Ricoeur, M. Kuna, Inuence of electric elds on the
fracture of ferroelectric ceramics, J. Eur. Ceram. Soc. 23
(2003) 13131328.
[18] H.G. Beom, S.N. Atluri, Eect of electric elds on
fracture behavior of ferroelectric ceramics, J. Mech. Phys.
Solids 51 (2003) 11071125.
[19] H. Kessler, H. Balke, On the local and average energy
release in polarization switching phenomena, J. Mech.
Phys. Solids 49 (2001) 953978.
[20] S. Li, On global energy release rate of a permeable crack
in a piezoelectric ceramic, J. Appl. Mech. 70 (2003) 246
252.
[21] F. Guiu, M. Alguero, M.J. Reece, Crack extension force
and rate of mechanical work of fracture in linear
dielectrics and piezoelectrics, Philos. Mag. A 83 (2003)
873888.
[22] R.M. McMeeking, The energy release rate for a Grith
crack in a piezoelectric material, Eng. Fract. Mech. 71
(2004) 11691183.
[23] S.C. Hwang, C.S. Lynch, R.M. McMeeking, Ferroelec-
tric/ferroelastic interactions and a polarization switching
model, Acta Metall. Mater. 43 (1995) 20732084.
[24] M.J. Reece, F. Guiu, Toughening Produced by crack-tip-
stress-induced domain reorientation in ferroelectric and/
or ferroelastic materials, Philos. Mag. A 82 (2002) 2938.
[25] H. Gao, T.Y. Zhang, P. Tong, Local and global energy
release rates for an electrically yield crack in piezoelectric
ceramics, J. Mech. Phys. Solids 45 (1997) 491510.
[26] T.C.T. Ting, Anisotropic Elasticity: Theory and Appli-
cations, Oxford Science Publications, New York, 1996.
[27] D.M. Barnett, J. Lothe, Dislocations and line charges in
anisotropic piezoelectric insulators, Phys. Status Solidi
(b) 67 (1975) 105111.
[28] M.Z. Wang, T.C.T. Ting, G.P. Yan, The anisotropic elastic
semi-innite strip, Q. Appl. Math. 51 (1993) 283297.
[29] C. Hausler, H. Balke, An interface crack between a
piezoelectric medium and an electric conductor, in: P.
Haupt, T. Kersten, V. Ulbricht (Eds.), Contributions to
modeling and identication, Gesamthochschule, Kassel,
2001 (in German).
[30] T.Y. Zhang, P. Tong, Fracture mechanics for a mode III
crack in a piezoelectric material, Int. J. Solids Struct. 33
(1996) 343359.
[31] T.Y. Zhang, C.F. Qian, P. Tong, Linear electro-elastic
analysis of a cavity or a crack in a piezoelectric material,
Int. J. Solids Struct. 35 (1998) 21212149.
[32] G.P. Cherepanov, Mechanics of Brittle Fracture, Nauka,
Moscow, 1974 (in Russian) English translation: McGraw-
Hill, 1979.
[33] Y.E. Pak, Crack extension force in a piezoelectric mate-
rial, Trans. ASME, J. Appl. Mech. 57 (1990) 647653.
[34] C.F. Gao, Further study of the generalized 2D problem
of an elliptical hole or a crack in piezoelectric media,
Mech. Res. Commun. 27 (2000) 429434.
[35] N.I. Muskhelishvili, Some Basic Problems of the Mathe-
matical Theory of Elasticity, Noordho, Groningen,
1953, Translated by J.R.M. Radok.
[36] C.F. Gao, Y.T. Zhao, M.Z. Wang, An exact and explicit
treatment of an elliptic hole problem in thermopiezoelec-
tric media, Int. J. Solids Struct. 39 (2002) 26652685.
[37] C.F. Gao, H. Balke, Fracture analysis of circular-arc
interface cracks in piezoelectric materials, Int. J. Solids
Struct. 40 (2003) 35073522.
[38] T.H. Hao, Z.Y. Shen, A new electric boundary condition
of electric fracture mechanics and its application, Eng.
Fract. Mech. 47 (1994) 793802.
[39] G.A. Schneider, F. Felten, R.M. McMeeking, The
electrical potential dierence across cracks in PZT mea-
sured by Kelvin probe microscopy and the implications
for fracture, Acta Mater. 51 (2003) 22352241.
[40] R.M. McMeeking, Towards a fracture mechanics for
brittle piezoelectric and dielectric materials, Int. J. Fract.
108 (2001) 2541.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 377
[41] L.A. Dissado, J.C. Fothergill, Electrical Degradation and
Breakdown in Polymers, Peter Peregrinus Ltd., London,
1992.
[42] B.S. Majumdar, S.J. Burns, Grith crack shielded by a
dislocation pile-up, Int. J. Fract. 21 (1983) 229240.
[43] R.M. McMeeking, A.G. Evans, Mechanics in transfor-
mation-toughening in brittle materials, J. Am. Ceram.
Soc. 65 (1982) 242246.
[44] G. Arlt, N.A. Pertsev, Force constant and eective mass
of 90 domain walls in ferroelectric ceramics, J. Appl.
Phys. 70 (1991) 22832289.
[45] W. Yang, T. Zhu, Switch-toughening of ferroelectrics
subjected to electric elds, J. Mech. Phys. Solids 46 (1998)
291311.
[46] R.M. McMeeking, S.C. Hwang, On the potential energy
of a piezoelectric inclusion and the criterion for ferro-
electric switching, Ferroelectrics 200 (1997) 151173.
[47] C.S. Lynch, W. Yang, L. Collier, Z. Suo, R.M.
McMeeking, Electric eld induced cracking in ferroelectric
ceramics, Ferroelectrics 166 (1995) 1130.
[48] W. Yang, T. Zhu, Fracture and fatigue of ferroelectrics
under electric and mechanical loading, Fatig. Fract. Eng.
Mater. Struct. 21 (1998) 13611369.
[49] T. Zhu, W. Yang, Fatigue crack growth in ferroelectrics
driven by cyclic electric loading, J. Mech. Phys. Solids 47
(1999) 8197.
[50] Q. Jiang, E.C. Subbarao, L.E. Cross, Eect of composi-
tion and temperature on electric fatigue of La-doped lead
zirconate titanate ceramics, J. Appl. Phys. 75 (1994)
74337443.
[51] Q. Jiang, E.C. Subbarao, L.E. Cross, Grain size depen-
dence of electric fatigue behavior of hot pressed PLZT
ferroelectric ceramics, Acta Metall. Mater. 42 (1994)
36873694.
[52] Y. Zhang, Q. Jiang, Twinning-induced stress and electric
eld concentrations in ferroelectric ceramics, J. Am.
Ceram. Soc. 78 (1995) 32903296.
[53] J.E. Huber, N.A. Fleck, C.M. Landis, R.M. McMeeking,
A constitutive model for ferroelectric polycrystals,
J. Mech. Phys. Solids 47 (1999) 16631697.
[54] R. Hill, A self-consistent mechanics of composite mate-
rials, J. Mech. Phys. Solids 13 (1965) 213222.
[55] R. Hill, Continue micro-mechanics of elastoplastic
polycrystals, J. Mech. Phys. Solids 13 (1965) 89
101.
[56] J.W. Hutchinson, Elasticplastic behavior of polycrystal-
line metals and composites, Proc. Roy. Soc. Lond. A 319
(1970) 247272.
[57] S.C. Hwang, R.M. McMeeking, A nite element model of
ferroelectric polycrystals, Ferroelectrics 211 (1998) 177
194.
[58] S.C. Hwang, R.M. McMeeking, A nite element model of
ferroelastic polycrystals, Int. J. Solids Struct. 36 (1999)
15411556.
[59] T. Michelitsch, W.S. Kreher, A simple model for the
nonlinear material behavior of ferroelectrics, Acta Mater.
46 (1998) 50855094.
[60] W. Lu, D.N. Fang, K.C. Hwang, Micromechanics of
ferroelectric domain switching behavior Part I: Coupled
electromechanical eld of domain inclusions, Theor.
Appl. Fract. Mech. 37 (2001) 2938.
[61] W. Lu, D.N. Fang, K.C. Hwang, Micromechanics of
ferroelectric domain switching behavior Part II: Consti-
tutive relations and hysteresis, Theor. Appl. Fract. Mech.
37 (2001) 3947.
[62] C.M. Landis, On the fracture toughness of ferroelastic
materials, J. Mech. Phys. Solids 51 (2003) 13471369.
[63] T. Zhu, W. Yang, Toughness variation of ferroelectrics by
polarization switch under non-uniform electric eld, Acta
Mater. 45 (1997) 46954702.
[64] R.K.N.D. Rajapakse, X. Zeng, Toughening of conduct-
ing cracks due to domain switching, Acta Mater. 49
(2001) 877885.
[65] X.L. Han, X.J. Li, S.X. Mao, Toughening and weakening
in ferroelectric ceramics by domain-switching process
under mixed electric and mechanical loading, Metall.
Mater. Trans. A 33 (2002) 28352845.
[66] J. Wang, S.Q. Shi, L.Q. Chen, Y. Li, T.Y. Zhang, Phase
eld simulations of ferroelectric and ferroelastic polari-
zation switching, Acta Mater., in press.
[67] R. Herbiet, U. Robels, H. Dederichs, A. Arlt, Domain
wall and volume contributions to material properties of
PZT ceramics, Ferroelectrics 98 (1989) 107121.
[68] N.A. Pertsev, G. Arlt, Forced translational vibrations 90
domain walls and the dielectric dispersion in ferroelectric
ceramics, J. Appl. Phys. 74 (1993) 41054112.
[69] R. Fu, T.Y. Zhang, Inuence of temperature and electric
eld on the bending strength of lead zirconate titanate
ceramics, Acta Mater. 48 (2000) 17291740.
[70] G. Arlt, H. Dederichs, R. Herbiet, 90 domain wall
relaxation in tetragonally distorted ferroelectric ceramics,
Ferroelectrics 74 (1987) 3753.
[71] D. Ricinschi, C. Harnagea, C. Papusoi, L. Mitoseriu, V.
Tura, M. Okuyama, Analysis of ferroelectric switching
in nite media as a Landau-type phase transition, J.
Phys.-Condens. Matter 10 (1998) 477492.
[72] L.Q. Chen, C. Wolverton, V. Vaithyanathan, Z.K. Liu,
Modeling solid-state phase transformation and micro-
structure evolution, MRS Bull. (2001) 197202.
[73] A.G. Tobin, Y.E. Pak, Eect of electric elds on fracture
behavior of PZT ceramics, Proc. SPIEInt. Soc. Opt.
Eng. 1916 (1993) 7886.
[74] H.Y. Wang, R.N. Singh, Crack propagation in piezo-
electric ceramics: eects of applied electric elds, J. Appl.
Phys. 81 (1997) 74717479.
[75] C.T. Sun, S.B. Park, Measuring fracture toughness of
piezoelectrics by Vickers indentation under the inuence
of electric elds, Ferroelectrics 248 (2000) 7995.
[76] L.Z. Jiang, C.T. Sun, Analysis of indentation cracking in
piezoelectrics, Int. J. Solids Struct. 38 (2001) 1903
1918.
[77] C.S. Lynch, Fracture of ferroelectric and relaxor electro-
ceramics: inuence of electric eld, Acta Mater. 46 (1998)
599608.
378 T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379
[78] R. Fu, T.Y. Zhang, Eect of an applied electric eld on
the fracture toughness of lead zirconate titanate ceramics,
J. Am. Ceram. Soc. 83 (2000) 12151218.
[79] G.A. Schneider, V. Heyer, Inuence of the electric eld
on Vickers indentation crack growth in BaTiO
3
, J. Eur.
Ceram. Soc. 19 (1999) 12991306.
[80] S.B. Park, C.T. Sun, Fracture criteria for piezoelectric
ceramics, J. Am. Ceram. Soc. 78 (1995) 14751480.
[81] R. Fu, T.Y. Zhang, Eects of an applied electric eld
on the modulus of rupture of poled Lead Zirconate
Titanate ceramics, J. Am. Ceram. Soc. 81 (1998) 1058
1060.
[82] Y. Shindo, H. Murakami, K. Horiguchi, F. Narita,
Evaluation of electric fracture properties of piezoelectric
ceramics using the nite element and single-edge pre-
cracked-beam methods, J. Am. Ceram. Soc. 85 (2002)
12431248.
[83] L.V. Zhoga, V.V. Shpeizman, Failure of ferroelectric
ceramics in electric and mechanical elds, Sov. Phys.
Solid State 34 (1992) 25782583.
[84] H. Makino, N. Kamiya, Eect of dc electric eld on
mechanical properties of piezoelectric ceramics, Jpn. J.
Appl. Phys. 33 (1994) 53235327.
[85] R. Fu, T.Y. Zhang, Statistical studies of the modulus of
rupture of poled PZT-841 ceramics under combined
mechanicalelectric loadings, In: Fifth International con-
ference on the Fundamentals of Fracture, NIST, USA,
1997, pp. 130131.
[86] V. Heyer, G.A. Schneider, H. Balker, J. Drescher, H.A.
Bahr, A fracture criterion for conducting cracks in
homogeneously poled piezoelectric PZT-PIC 151 cera-
mics, Acta Mater. 46 (1998) 66156622.
[87] T.H. Wang, T.Y. Zhang, Electrical fracture toughness for
electrically conductive deep notches driven by electric
elds in depoled lead zirconate titanate ceramics, Appl.
Phys. Lett. 79 (2001) 41984200.
[88] R. Fu, C.F. Qian, T.Y. Zhang, Electrical fracture
toughness for conductive cracks driven by electric elds
in piezoelectric materials, Appl. Phys. Lett. 76 (2000) 126
128.
[89] S.W. Freiman, G.S. White, Intelligent ceramic materials-
issues of brittle-fracture, J. Intell. Mater. Syst. Struct. 6
(1995) 4954.
[90] T.Y. Zhang, T.H. Wang, M.H. Zhao, Failure behavior
and failure criterion of conductive cracks (deep notches)
in thermally depoled PZT-4 ceramics, Acta Mater. 51
(2003) 48814895.
[91] Y. Wang, W.Y. Chu, Y.J. Su, L.J. Qiao, Stress corrosion
cracking of PZT piezoelectric ceramics, Mater. Lett. 57
(2003) 11561159.
[92] Y. Wang, W.Y. Chu, Y.J. Su, L.J. Qiao, Anisotropy of
stress corrosion cracking of piezoelectric ceramics, Mat-
ter. Sci. Eng. B 95 (2002) 263267.
[93] Y. Wang, W.Y. Chu, L.J. Qiao, Y.J. Su, Hydrogen-
induced delayed fracture of PZT ceramics during dynamic
charging under constant load, Matter. Sci. Eng. B 98
(2003) 15.
[94] S.L. dos Santos e Lucato, D.C. Lupascu, J. R odel, Eect
of poling direction on R-curve behavior in lead zirconate
titanate, J. Am. Ceram. Soc. 83 (2000) 424426.
[95] A. Kolleck, G.A. Schneider, F.A. Meschke, R-curve
behavior of BaTiO
3
and PZT ceramics under the inu-
ence of an electric eld applied parallel to the crack front,
Acta Mater. 48 (2000) 40994113.
[96] Y. Wang, W.Y. Chu, K.W. Gao, Y.J. Su, L.J. Qiao,
Delayed fracture of lead zirconate titanate ferroelectric
ceramics under sustained electric eld, Appl. Phys. Lett.
82 (2003) 15831586.
[97] B. Liu, D.N. Fang, K.C. Hwang, Electric-eld-induced
fatigue crack growth in ferroelectric ceramics, Mater.
Lett. 54 (2002) 442446.
[98] X.L. Tan, J.K. Shang, In-situ transmission electron
microscopy study of electric eld-induced grain-boundary
cracking in lead zirconate titanate, Philos. Mag. A 82
(2002) 14631478.
[99] S.L. dos Santos e Lucato, H.A. Bahr, V.B. Pham, D.C.
Lupascu, H. Balke, J. R odel, U. Bahr, Crack deection in
piezoelectric ceramics, J. Eur. Ceram. Soc. 23 (2003)
11471156.
[100] S.L. dos Santos e Lucato, H.A. Bahr, V.B. Pham, D.C.
Lupascu, H. Balke, J. R odel, U. Bahr, Electrically driven
cracks in piezoelectric ceramics: experiments and fracture
mechanics analysis, J. Mech. Phys. Solids 50 (2002) 2333
2353.
[101] H.R. Zeller, W.R. Schneider, Electrofracture mechanics
of dielectric aging, J. Appl. Phys. 56 (1984) 455459.
[102] S.M. Ohr, An electron microscope study of crack tip
deformation and its impact on the dislocation theory of
fracture, Mater. Sci. Eng. 72 (1985) 135.
[103] S.J. Chang, S.M. Ohr, Dislocation-free zone model of
fracture, J. Appl. Phys. 52 (1981) 71747181.
[104] S.H. Dai, J.C.M. Li, Dislocation-free zone at the crack
tip, Scr. Metall. 16 (1982) 183188.
[105] J. Weertman, I.H. Lin, R. Thomson, Double slip plane
crack model, Acta Metall. 31 (1983) 473482.
[106] I.H. Lin, R. Thomson, Cleavage, dislocation emission
and shielding for cracks under general loading, Acta
Metall. 34 (1985) 187206.
[107] J.R. Rice, D.E. Hawk, R.J. Asaro, Crack tip elds in
ductile crystals, Int. J. Fract. 42 (1990) 301321.
[108] W. Zielinski, M.J. Lii, H. Huang, P.G. Marsh, W.W.
Gerberich, Crack-tip dislocation emission arrangements
for equilibrium, 1: In-situ TEM observations of Fe-2Wt-
percent-Si, Acta Metall. Mater. 40 (1992) 28612871.
[109] H. Huang, W.W. Gerberich, Quasi-equilibrium modeling
of the toughness transition during semibrittle cleavage,
Acta Metall. Mater. 42 (1994) 639647.
[110] T. Mihara, H. Watenable, Electronic conduction charac-
teristics of solgel ferroelectric Pb(Zr0.4Ti0.6)O
3
thin-lm
capacitors. 1, Jpn. J. Appl. Phys. 34 (1995) 56645673.
[111] R.H. Fowler, L. Nordheim, Electron emission in intense
electricelds, Proc. Roy. Soc. Lond. A1928(1928) 173181.
[112] J.D. Eshelby, The continuum theory of lattice defects,
Solid State Phys. 3 (1956) 79144.
T.Y. Zhang, C.F. Gao / Theoretical and Applied Fracture Mechanics 41 (2004) 339379 379

You might also like