You are on page 1of 17

Journal of the Mechanics and Physics of Solids

51 (2003) 11711187
www.elsevier.com/locate/jmps
The size eect on void growth
in ductile materials
B. Liu
a
, X. Qiu
b
, Y. Huang
a,
, K.C. Hwang
b
, M. Li
c
, C. Liu
d
a
Department of Mechanical and Industrial Engineering, University of Illinois, 1206 W. Green Street,
Urbana, IL 61801, USA
b
Department of Engineering Mechanics, Tsinghua University, Beijing 100084, China
c
Alcoa Technical Center, Aluminum Company of America, Alcoa Center, PA 15069, USA
d
Materials Science and Technology Division, Los Alamos National Laboratory,
Los Alamos, NM 87545, USA
Received 1 March 2002; accepted 19 February 2003
Abstract
We have extended the RiceTracey model (J. Mech. Phys. Solids 17 (1969) 201) of void
growth to account for the void size eect based on the Taylor dislocation model, and have found
that small voids tend to grow slower than large voids. For a perfectly plastic solid, the void size
eect comes into play through the ratio c!}R
0
, where ! is the intrinsic material length on the
order of microns, c the remote eective strain, and R
0
the void size. For micron-sized voids and
small remote eective strain such that c!}R
0
60.02, the void size inuences the void growth rate
only at high stress triaxialities. However, for sub-micron-sized voids and relatively large eective
strain such that c!}R
0
0.2, the void size has a signicant eect on the void growth rate at all
levels of stress triaxiality. We have also obtained the asymptotic solutions of void growth rate at
high stress triaxialities accounting for the void size eect. For c!}R
0
0.2, the void growth rate
scales with the square of mean stress, rather than the exponential function in the RiceTracey
model (1969). The void size eect in a power-law hardening solid has also been studied.
? 2003 Elsevier Science Ltd. All rights reserved.
Keywords: Voids; Size eect; Void growth rate; Strain gradient plasticity; Taylor dislocation model
1. Introduction
The micro-indentation hardness experiments have repeatedly shown that the indenta-
tion hardness of metallic materials displays strong dependence on the indentation depth

Corresponding author. Tel.: +1-217-265-5072; fax: +1-217-244-6534.


E-mail address: huang9@uiuc.edu (Y. Huang).
0022-5096/03/$ - see front matter ? 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0022-5096(03)00037-1
1172 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
(e.g., Nix, 1989, 1997; Stelmashenko et al., 1993; Ma and Clarke, 1995; Poole et al.,
1996; McElhaney et al., 1998; Swadener et al., 2002). The indentation hardness may
increase by a factor of two or even three as the indentation depth decreases to mi-
crons and sub-microns. Similar size dependence of plastic behavior of materials at the
micron scale has also been observed in micro-twist of thin copper wires (Fleck et al.,
1994), micro-bend of thin nickel foils (Stolken and Evans, 1998; Shrotriya et al., 2003)
and thin aluminum beams (Haque and Saif, 2003), in an aluminum matrix reinforced
by SiC particles (Lloyd, 1994; Xue et al., 2002a), and in a micro-electro-mechanical
system (MEMS) named Digital Micromirror Device (DMD) (Douglass, 1998; Xue
et al., 2002b). Direct dislocation simulations have also repeatedly shown strong size
dependence of metallic materials at the micron scale under various loading conditions
(Cleveringa et al., 1997, 1998, 1999a, b, 2000; Needleman, 2000).
Classical theories of plasticity fail to explain the observed size dependence of ma-
terial behavior at the micron and sub-micron scales since their constitutive models
possess no internal material length. This has led to the development of strain gradient
plasticity theories (e.g., Fleck and Hutchinson, 1993, 1997, 2001; Fleck et al., 1994;
Nix and Gao, 1998; Gao et al., 1999a, b; Huang et al., 1999, 2000a, b; Acharya and
Bassani, 2000; Acharya and Beaudoin, 2000; Qiu et al., 2001; Gurtin, 2002; Hwang
et al., 2002, 2003) based on the concept of geometrically necessary dislocations in
dislocation mechanics (Nye, 1953; Cottrell, 1964; Ashby, 1970; Arsenlis and Parks,
1999; Gurtin, 2000). The Taylor dislocation model (Taylor, 1934, 1938), which links
the ow stress and the dislocation density, is the basis for some strain gradient plastic-
ity theories (e.g., Gao et al., 1999b; Huang et al., 1999, 2000a, b; Qiu et al., 2001), and
these theories have shown excellent agreement with the micro-indentation experiments,
such as with McElhaney et al.s (1998) of copper hardness data (Huang et al., 2000b),
Stelmashenko et al.s tungsten data (Qiu et al., 2001), and Saha et al.s (2001) data
for an aluminum thin lm on a glass substrate. These theories also agree well with
other micron scale experiments, such as Fleck et al.s (1994) micro-twist and Stolken
and Evans (1998) and Haque and Saifs (2003) micro-bend experiments (Gao et al.,
1999a; Haque and Saif, 2003), with Lloyds (1994) particle-reinforced metal-matrix
composite materials (Xue et al., 2002a), as well as with Douglasss (1998) MEMS
named DMD (Xue et al., 2002b). Jiang et al. (2001) and Shi et al. (2001) have shown
that these strain gradient plasticity theories based on the Taylor dislocation model can
explain cleavage fracture in ductile materials observed in experiments (Bagchi et al.,
1994; Elssner et al., 1994; Bagchi and Evans, 1996).
The ductile failure of metallic materials results from the nucleation, growth and
coalescence of microvoids. For a rigidperfectly plastic solid under high stress triax-
iality, the pioneer work of Rice and Tracey (1969) established that the void growth
rate increases exponentially with the mean stress surrounding the void. Attention was
limited to a spherical void in an innite rigidperfectly plastic solid characterized by
the Mises yield condition, o
e
= o
Y
, where the eective stress o
e
was related to the
deviatoric stresses s
i)
=o
i)

1
3
o
kk
o
i)
by o
e
=(3s
i)
s
i)
}2)
1}2
, and o
Y
was the yield stress.
The solid was subjected to remote axisymmetric tension, o

11
, o

22
=o

11
, and o

33
, sat-
isfying the Mises yield condition o

33
o

11
=o
Y
. The non-vanishing strain rates in the
remote eld were c

33
=2 c

11
=2 c

22
, and the corresponding eective strain rate was
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1173
c = (2 c

i)
c

i)
}3)
1}2
= c

33
. Rice and Tracey (1969) gave the void growth rate as
D =

J
3 cJ
= 0.2833 exp
_
3o
m
2o
Y
_
, (1)
where J was the void volume and o
m
= o

kk
}3 was the mean stress in the remote
eld. The RiceTracey model (1969) has been widely used to study void growth in
ductile materials, and has been extended to viscoplastic materials (e.g., Budiansky
et al., 1982), to account for the eect of low stress triaxiality (e.g., Huang, 1991),
and to establish yield criteria for isotropic (e.g., Gurson, 1977; Tvergaard, 1990),
anisotropic (e.g., Chen et al., 2000) and viscoplastic solids containing microvoids
(Chen et al., 2003).
Recent experimental investigations (e.g., Schluter et al., 1996; Khraishi et al., 2001)
and numerical studies (e.g., Fond et al., 1996; Fleck and Hutchinson, 1997; Shu, 1998;
Huo et al., 1999; Zhang et al., 1999; Zhang and Hsia, 2001) have shown that void
growth in ductile materials also depends strongly on the void size. The micron- and
sub-micron-sized voids tend to grow slower than large voids under the same stress level.
This size eect is consistent with the aforementioned experiments of micro-indentation,
micro-twist, micro-bend, particle-reinforced composite, and MEMS, i.e., smaller is
stronger. The RiceTracey model (1969) as well as other void growth models (e.g.,
Budiansky et al., 1982; Huang, 1991) cannot account for the void size eect since
the classical plasticity theories possess no intrinsic material lengths. Since the Taylor
dislocation model (Taylor, 1934, 1938) can characterize the plastic deformation rather
accurately at the micron and sub-micron scales, it is used in the present paper to extend
the RiceTracey model (1969) of the void growth rate accounting for the void size
eect. Section 2 provides a summary of the RiceTracey model of void growth, while
Section 3 gives the Taylor dislocation model, which relates the ow stress to the plastic
strain and plastic strain gradient. Section 4 studies the void size eect in a perfectly
plastic solid, while Section 5 gives the analytical expression of the void growth rate
for the high stress triaxiality limit. The void size eect in a power-law hardening solid
is studied in Section 6.
2. Variation principle for a rigidperfectly plastic solid
Rice and Tracey (1969) established a variation principle for an innite, rigid
perfectly plastic solid containing a single void. The variation principle, also adopted by
Budiansky et al. (1982) and Huang (1991) to study void growth in ductile materials,
involves the minimization of the following functional of velocity u,
1( u) =
_
O
[s
i)
( c) s

i)
] c
i)
dO o

i)
_
S
O
n
i
u
)
dS
O
, (2)
where O is the innite volume exterior to the void, S
O
the void surface, n
i
the unit
normal on S
O
pointing into the solid; u is the velocity eld satisfying the incompress-
ibility condition u
k, k
=0, strain rates c
i)
=( u
i, )
+ u
), i
)}2; s
i)
and s

i)
are the deviatoric
stresses and their counterparts in the remote eld, respectively, and are given in terms
1174 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
of the strain rates via plastic ow normality,
s
i)
=
o
Y
c
i)
(
3
2
c
k!
c
k!
)
1}2
, s

i)
=
o
Y
c

i)
(
3
2
c

k!
c

k!
)
1}2
. (3)
Here the Mises yield condition o
e
= o
Y
has been used, and o
Y
is the tensile yield
stress.
For prescribed remote strain rates c

i)
, an additional velocity eld u

i
is dened such
that
c
i)
= c

i)
+ c

i)
, c

i)
=
1
2
( u

i, )
+ u

), i
). (4)
Among all admissible additional elds satisfying u

= o(R
3}2
) as the distance to the
void center R, the exact eld minimizes the functional 1 (Rice and Tracey, 1969).
Similar to Rice and Tracey (1969), Budiansky et al. (1982) and Huang (1991), we
decompose the admissible velocity eld u
i
into the following three parts,
(i) a uniform velocity eld c

i)
x
)
associated with the uniform remote strain rates c

i)
,
where x
)
are the Cartesian coordinates; this eld does not change the void volume
due to incompressibility of the remote eld, c

kk
= 0;
(ii) a volume-changing, spherical symmetric additional velocity eld u
D
i
correspond-
ing to a change in the void volume but no change of shape, i.e., u
D
R
=(R
0
}R)
2

R
0
,
u
D
0
= u
D
[
=0 in spherical coordinates (R, 0, [), where R
0
is the void radius and R
the distance to the center of spherical void; and
(iii) a shape-changing additional velocity eld, u
E
i
, decaying at remote distances, which
changes the void shape but not its volume, i.e., u
E
i
(R ) = o(R
3}2
),
_
S
O
u
E
R
dS
O
= 0, where integration is over the void surface S
O
.
Rice and Tracey (1969) showed that the above shape-changing additional velocity
eld (iii) has essentially no eect on the void growth rate at high stress triaxiality.
The velocity eld is then written as
u
i
= c

i)
x
)
+
R
2
0
R
3

R
0
x
i
= c

i)
x
)
+ cD
R
3
0
R
3
x
i
, (5)
where c =
_
2 c

i)
c

i)
}3 is the eective strain rate in the remote eld,
D =

R
0
cR
0
=

J
3 cJ
(6)
is the void growth rate consistent with Rice and Tracey (1969) given in Eq. (1), and
J is the void volume. The strain rates can be obtained from (5) as
c
i)
= c

i)
+ cD c
D
i)
(7)
and the non-vanishing components of c
D
i)
are
c
D
RR
=2 c
D
00
=2 c
D
[[
=2
R
3
0
R
3
(8)
in the spherical coordinates.
The parameter to be determined in the velocity eld (5) is the void growth rate D.
For a rigidperfectly plastic solid, the minimization of velocity functional 1 in Eq.
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1175
(2) with respect to D gives the following non-linear integral equation for D (Rice and
Tracey, 1969)
_
O
[s
i)
(D) s

i)
] c
D
i)
dO = o
m
R
0
_
S
O
dS
O
, (9)
where s
i)
(D) are obtained from (3), (7), and the Mises yield condition o
e
=o
Y
; s

i)
are
obtained similarly in terms of the remote strain rates c

i)
; the non-vanishing components
of c
D
i)
are given in Eq. (8), o
m
= o

kk
}3 is the remote mean stress, and R
0
the void
radius. Note that in computing derivatives for the minimization, we have used
cs
i)
(D)
cD
c
i)
= 0 (10)
since the stress derivative is tangent to the yield surface and is therefore normal to c
i)
.
Rice and Tracey (1969) obtained the solution of (9), which gives the void growth rate
in Eq. (1) under high stress triaxialities.
In the present study of the void size eect on the void growth rate D, we modify the
RiceTracey model to replace the tensile yield stress o
Y
by the ow stress obtained
from the Taylor dislocation model, as discussed in the following sections.
3. The Taylor dislocation model
The Taylor dislocation model (Taylor, 1934, 1938) gives the shear ow stress t in
terms of the dislocation density j by
t = :jb

j = :jb

j
S
+ j
G
, (11)
where j is the shear modulus, b the Burgers vector, : an empirical material constant
around 0.3 (e.g., Taylor, 1934, 1938; Wiedersich, 1964), j
S
and j
G
are densities of
statistically stored dislocations and geometrically necessary dislocations, respectively.
The above relation has been rewritten to give the tensile ow stress in terms of the
plastic strain and plastic strain gradient (Nix and Gao, 1998; Huang et al., 2000b).
The tensile ow stress o is related to the shear ow stress t in Eq. (11) by o = Mt,
where M is the Taylor coecient; M =

3 for an isotropic solid, and M =3.08 for a


face-centered-cubic (FCC) crystal (Bishop and Hill, 1951a, b; Kocks, 1970).
The density of geometrically necessary dislocations is related to the gradient of
plastic deformation by (Ashby, 1970; Nix and Gao, 1998; Huang et al., 2000b)
j
G
= r
p
p
b
, (12)
where b is the Burgers vector, p
p
the eective plastic strain gradient to be given later,
r is the Nye factor (Arsenlis and Parks, 1999) to account for the eect of discrete slip
systems on the distribution of geometrically necessary dislocations, and r is around
1.9 for FCC crystals (Arsenlis and Parks, 1999). The density of statistically stored
dislocations is determined from the relation between stress o and plastic strain c
p
in
uniaxial tension, o =o
ref
[
p
(c
p
), where o
ref
is a reference stress (e.g., yield stress o
Y
).
This is because the density of geometrically necessary dislocations vanishes in uniaxial
1176 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
tension since there is no strain gradient [see (12)]. The Taylor dislocation model (11)
then gives
o
ref
[
p
(c
p
)
M
= :jb

j
S
, (13)
where the tensile ow stress has been substituted by the uniaxial stressplastic strain
relation. The substitution of (12) and (13) into (11) gives the ow stress in terms of
eective plastic strain and eective plastic strain gradient,
o =
_
o
2
ref
[
2
p
(c
p
) + 18:
2
j
2
bp
p
= o
ref
_
[
2
p
(c
p
) + !p
p
, (14)
where c
p
=
_
c
p
dt is the eective plastic strain, c
p
=
_
2
3
c
p
i)
c
p
i)
the eective plastic strain
rate, and c
p
i)
the plastic strain rate tensor; the eective plastic strain gradient p
p
is given
later;
! = 18:
2
_
j
o
ref
_
2
b (15)
is the intrinsic material length associated with the strain gradient eect, and is on the
order of microns (Huang et al., 2000b). It should be pointed out that, even though
the intrinsic material length ! depends on o
ref
, the ow stress in Eq. (14) depends
only on the overall uniaxial stressstrain curve o
ref
[
p
(c
p
) and strain gradient p
p
, and
is independent of the choice of o
ref
.
Gao et al. (1999b) developed dislocation models to calculate the density of geomet-
rically necessary dislocations in bending, torsion and void growth. Based on dislocation
models, they determined the eective plastic strain gradient p
p
as
p
p
=
_
1
4
p

i)k
p

i)k
, p

i)k
= c
p
)k, i
+ c
p
ik, )
c
p
i), k
, (16)
where c
p
i)
=
_
c
p
i)
dt is the plastic strain tensor, and c
p
i)
is the corresponding plastic strain
rate tensor. It is noted that Fleck and Hutchinson (1997) also determined the eective
strain gradient in terms of the second order invariants of the strain gradient tensor, but
the coecients are dierent from
1
4
in Eq. (16).
In the following sections, we use the ow stress in Eq. (14) based on the Taylor
dislocation model to replace the tensile yield stress o
Y
in Eq. (3). Therefore, such
an approach does not involve higher order stresses, contrary to some strain gradient
plasticity theories. Moreover, since the ow stress in Eq. (14) depends on the plastic
strain and strain gradient but not their increments, the variational principle (2) still
holds in the present study.
4. The size eect on void growth in a perfectly plastic solid
We investigate the void size eect on void growth rate in a perfectly plastic solid
based on the Taylor dislocation model. Follow Rice and Tracey (1969), we neglect the
shape-changing additional velocity eld such that the strain rates c
i)
and velocities u
i
are still given by (7) and (5), and are linearly dependent on the void growth rate D.
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1177
Similar to (3), normality of plastic ow gives the deviatoric stresses in terms of strain
rates as
s
i)
(D) =
o
e
c
i)
(D)
[
3
2
c
k!
(D) c
k!
(D)]
1}2
, s

i)
=
o
Y
c

i)
(
3
2
c

k!
c

k!
)
1}2
, (17)
where the dependence on the void growth rate D is made explicit; o
e
is the eective
stress, and the eective stress in the remote eld is the yield stress o
Y
, as shown later.
The yield condition in the Taylor dislocation model gives the eective stress as
o
e
= o, (18)
where o is the ow stress in Eq. (14) accounting for the strain gradient eect. For a
perfectly plastic solid, o
ref
[
p
(c
p
) = o
Y
, the ow stress becomes
o =
_
o
2
Y
+ 18:
2
j
2
bp
p
= o
Y
_
1 + !p
p
, (19)
which diers from the yield stress o
Y
by the eective plastic strain gradient term, !p
p
,
where ! is the intrinsic material length in Eq. (15), and p
p
is given in Eq. (16). Unlike
Rice and Tracey (1969), the ow stress in Eq. (19) depends on the displacement eld
through plastic strain gradient p
p
.
If we assume proportional deformation, the displacement eld can be obtained from
the velocity eld (5) as
u
i
= c

i)
x
)
+
R
2
0
R
3
u
0
x
i
= c

i)
x
)
+ cD
R
3
0
R
3
x
i
, (20)
where c

i)
and c =
_
2c

i)
c

i)
}3 are the strains and eective strain in the remote eld,
respectively; u
0
is the increase of average void radius, and D=u
0
}(cR
0
) is identical to
the void growth rate in Eq. (6) under proportional deformation. Since elastic deforma-
tion is neglected (Rice and Tracey, 1969; Budiansky et al., 1982; Huang, 1991), the
eective plastic strain gradient p
p
can be obtained from the above displacement eld
via (16),
p
p
= 3
_
5
2
R
2
0
R
4
u
0
= 3
_
5
2
R
3
0
R
4
cD. (21)
Its substitution into (19) gives the ow stress as
o = o
Y

1 + 3
_
5
2
R
2
0
R
4
u
0
! = o
Y

1 + 3
_
5
2
_
R
0
R
_
4
c!
R
0
D. (22)
It is straightforward to verify that the above degenerates to yield stress o
Y
in the remote
eld (R ).
As discussed in Section 3, the variation principle (2) still holds after the strain
gradient eect is accounted for in the present study for a rigidperfectly plastic solid.
For the velocity eld in Eq. (5), minimization of velocity functional (2) also leads
to (9), except that the deviatoric stresses s
i)
are obtained from (17)(19) to account
for the strain gradient eect. By changing the integration variables to z = R
3
0
}R
3
and
1178 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
=cos 0, we rewrite (9) to the following integral equation for the void growth rate D,
_
1
0
dz
_
1
0
_

_
(1 + 4Dz 3
2
)
_
1 + 3
_
5
2
c!
R
0
Dz
4}3
z
_
1 + 2Dz + 4D
2
z
2
6Dz
2

1 3
2
z
_

_
d =
3o
m
o
Y
, (23)
where o
m
= o

kk
}3 and c are the remote mean stress and eective strain, respectively,
o
Y
the yield stress, ! the intrinsic material length in Eq. (15), and R
0
the void radius.
The void size eect comes into play through the ratio c!}R
0
. For void size much larger
than the intrinsic material length which is on the order of microns, R
0
!, the void
size eect disappears. Eq. (23) then degenerates to Rice and Traceys (1969) analysis
and gives the void growth rate in Eq. (1) under high stress triaxialities. It is observed
that, for vanishingly small remote eective strain (c 0), the void size eect also
disappears since the ratio c!}R
0
0.
Eq. (23) can be simplied to the following integral equation after the integration
with respect to is carried out analytically,
_
1
0

1 + 3
_
5
2
c!
R
0
Dz
4}3
_
|1 2Dz|
1 2Dz 12D
2
z
2

6Dz
sin
1
_
6Dz
1 + 2Dz + 4D
2
z
2
_
dz
4Dz
2
=
3o
m
o
Y
. (24)
It can be shown that the above integrand is non-singular around z = 0. The solution
of (24) takes the form
D = D
_
o
m
o
Y
,
c!
R
0
_
, (25)
which once again indicates that the void size eect comes into play through the
non-dimensional combination of c!}R
0
.
Fig. 1 shows the void growth rate D versus normalized mean stress o
m
}o
Y
for
c!}R
0
=0, 0.004, 0.01 and 0.02. The limit of c!}R
0
=0 corresponds to Rice and Traceys
(1969) void growth rate in classical plasticity. The void growth rate decreases with
increasing c!}R
0
, which suggests that, at a given remote eective strain c and intrinsic
material length !, small voids tend to grow slower than large voids. This is consis-
tent with the aforementioned experimental observations (Schluter et al., 1996; Khraishi
et al., 2001) and numerical studies of void growth (Fond et al., 1996; Fleck and
Hutchinson, 1997; Shu, 1998; Huo et al., 1999; Zhang et al., 1999; Zhang and Hsia,
2001). It should be pointed out that the inuence of void size depends strongly on
the mean stress level. At relatively small mean stresses (e.g., o
m
}o
Y
1) as in most
laboratory experiments, void size has little or essentially no eect on the void growth
rate. At the large mean stress o
m
}o
Y
=4.5 of cavitation instability (Huang et al., 1991;
Tvergaard et al., 1992) observed in highly constrained ductile materials (Ashby et al.,
1989), small voids grow signicantly slower than large voids. Even at an intermediate
mean stress o
m
}o
Y
= 2.6 which is a representative value for a mode-I crack tip in a
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1179
Fig. 1. The void growth rate D versus normalized mean stress o
m
}o
Y
for a perfectly plastic solid and
c!}R
0
= 0, 0.004, 0.01 and 0.02, where D is dened in Eq. (6), o
m
and c are the mean stress and eective
strain in the remote eld, respectively, o
Y
the tensile yield stress, ! the intrinsic material length in Eq. (15),
and R
0
the void radius. The curve for c!}R
0
= 0 corresponds to the RiceTracey model (1969) of void
growth in classical plasticity.
perfectly plastic solid, the void growth rate for c!}R
0
=0.02 is still more than 30% less
than the prediction of classical plasticity (c!}R
0
= 0).
The eect of void size can be magnied by the remote eective strain c since void
size comes into play through the ratio c!}R
0
. We may take aluminum subjected to mean
stress o
m
}o
Y
= 2.6 as an example. The intrinsic material length ! for aluminum is a
few microns (Xue et al., 2002a; Haque and Saif, 2003). For a relatively small eective
strain c =0.002, the void size eect is signicant (c!}R
0
=0.02) for a sub-micron-sized
void (!}R
0
= 10), but is much less signicant (c!}R
0
= 0.004) for a micron-sized void
(!}R
0
=2). However, at a larger eective strain c=0.01, the size eect for a micron-sized
void (!}R
0
= 2) becomes signicant (c!}R
0
= 0.02).
5. High stress-triaxiality approximations
Follow Rice and Tracey (1969), we establish the asymptotic expression of void
growth rate under high stress triaxiality accounting for the void size eect in this
section. As shown in Fig. 1 [and also (1)], high stress triaxiality o
m
}o
Y
1 leads to
large void growth rate, D1. Under high stress-triaxiality limit the governing equation
(24) for void growth rate can be simplied to give analytic or semi-analytic solutions.
We rst study the void growth rate for small ratio of c!}R
0
as in Fig. 1, c!}R
0
1.
This corresponds to relatively small remote eective strain (c up to a few percent) and
micron-sized voids (!}R
0
less than or on the order of 1). It can be shown that, for
1180 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
Fig. 2. Comparison of the void growth rate in a perfectly plastic solid with its asymptotic solution (26) for
high stress triaxialities o
m
}o
Y
1 (D1) and small ratio c!}R
0
1. All parameters and notations are the
same as in Fig. 1. The asymptotic solution (26) involves no tting parameters.
c!}R
0
1 and D1, the asymptotic analysis of (24) gives
Dexp
_
9

10
16
c!
R
0
D
_
= 0.2833 exp
_
3o
m
2o
Y
_
for
c!
R
0
1. (26)
Its right-hand side is Rice and Traceys (1969) void growth rate under high stress
triaxiality given in Eq. (1). For c!}R
0
=0, (26) degenerates to Rice and Traceys (1969)
solution in classical plasticity. The algebraic equation (26) for D is much simpler than
the integral equation (24) and involves no tting parameters. Its solution, marked as
the asymptotic solution in Fig. 2, has good agreement with the numerical results of
(24) for the same set of c!}R
0
.
Eq. (26) has the following approximate but explicit solution for small ratio c!}R
0
D =
2D
RT
1 +
_
1 + 8.506(c!}R
0
)D
RT
for
c!
R
0
1, (27)
where D
RT
=0.2833 exp(3o
m
}2o
Y
) is Rice and Traceys (1969) void growth rate given
in Eq. (1). The right-hand side of (27) comes from the asymptotic analysis of (26) for
c!}R
0
1, but the coecient 8.506 inside the square root is determined by tting the
numerical solution of (24). As shown in Fig. 3, (27) also gives a good approximate
solution of the void growth rate to account for the void size eect for c!}R
0
1. This
approximate solution shows that, once the void size eect is accounted for, the void
growth rate is no longer an exponential function of mean stress as the RiceTracey
model (1969).
It should be emphasized that the asymptotic solution (26) and approximate solution
(27) hold strictly for c!}R
0
1. For a sub-micron-sized void !}R
0
= 10 and a large
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1181
Fig. 3. Comparison of the void growth rate in a perfectly plastic solid with the approximate solution (27)
for high stress triaxialities o
m
}o
Y
1 (D1) and small ratio c!}R
0
1. All parameters and notations are
the same as in Fig. 1. The approximate solution (27) involves one tting parameter. The void growth rate
is no longer an exponential function of mean stress as the RiceTracey model (1969) after the void size
eect is accounted for.
remote eective strain c = 0.1, the ratio c!}R
0
becomes 1 such that (26) and (27) are
not applicable anymore. We have also analyzed the asymptotic limit of (24) for D1
and c!}R
0
on the order of 1, and obtained the following asymptotic solution for the
void growth rate without any parameter tting
D =
4
45
_
c!
R
0
_
2
_
o
m
o
Y
_
2
for
c!
R
0
0.2. (28)
This asymptotic solution suggests that, for suciently small voids and large eective
strain such that c!}R
0
0.2, the void growth rate D scales with the square of mean
stress instead of the exponential function in Eq. (1) at high stress triaxialities. Fig. 4
shows the void growth rate D versus the normalized mean stress o
m
}o
Y
for c!}R
0
=1.
Both numerical solution of (24) and asymptotic solution (28) are shown, along with
the void growth rate in classical plasticity (c!}R
0
= 0). For a sub-micron-sized void
!}R
0
=10 and large eective strain c =0.1, the void size eect becomes signicant even
at relatively small mean stresses (e.g., o
m
}o
Y
1), as seen from the large dierence
between the solid curve for c!}R
0
= 1 and that for classical plasticity c!}R
0
= 0. The
asymptotic solution (28) is a good representation of the void growth rate only at large
mean stresses. This is because, for c!}R
0
= 1, the void size eect has signicantly
reduced D to 5 even at a large mean stress o
m
}o
Y
= 4.5 for cavitation instability, and
such a value of 5 does not satisfy D1 required in the asymptotic analysis.
1182 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
Fig. 4. The void growth rate D versus normalized mean stress o
m
}o
Y
for a perfectly plastic solid and a
large ratio c!}R
0
= 1. All notations are the same as in Fig. 1. The curve for c!}R
0
= 0 corresponds to the
RiceTracey model (1969) in classical plasticity. The asymptotic solution (28) for high stress triaxialities
o
m
}o
Y
1 (D1) and large ratio c!}R
0
(on the order of 1) is also shown. The asymptotic solution (28),
involving no tting parameters, scales with the square of mean stress, (o
m
}o
Y
)
2
, instead of the exponential
function in the RiceTracey model (1969) for c!}R
0
on the order of 1.
6. Void growth in a power-law hardening solid
We investigate the void size eect on void growth rate in a power-law hardening
solid in this section. The relation between stress and plastic strain in uniaxial tension
can be written as o
ref
[
p
(c
p
) = o
ref
c
N
p
, where o
ref
is a reference stress and N (1) is
the plastic work hardening exponent. The strain rates c
i)
, velocities u
i
, displacements
u
i
and eective plastic strain gradient p
p
are still given by (7), (5), (20) and (21),
respectively. For monotonically increasing and proportional deformation, the eective
plastic strain c
p
is obtained from the displacements in Eq. (20) as
c
p
= c
_
1 + 2(1 3 cos
2
0)D
R
3
0
R
3
+ 4D
2
R
6
0
R
6
, (29)
where c is the eective strain in the remote eld, and 0 is the spherical angle.
The deviatoric stresses are given by (17) except that the eective stress in the remote
eld is no longer the yield stress o
Y
. The yield condition in the Taylor dislocation
model (18) still holds. The ow stress o in Eq. (14) accounting for the strain gradient
eect becomes
o =
_
[o
ref
c
N
p
]
2
+ 18:
2
j
2
bp
p
= o
ref
_
c
2N
p
+ !p
p
, (30)
where ! =18:
2
(j}o
ref
)
2
b is the intrinsic material length for a power-law hardening
solid, and c
p
and p
p
are given in Eqs. (29) and (16), respectively.
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1183
For a strain hardening solid Rice and Tracey (1969) showed that, among all admis-
sible velocity elds, the exact eld minimizes the following functional of velocity
1( u) =
_
O
[s
i)
( c) s

i)
] c
i)
dO
_
O
[s
i)
( c

) s

i)
] c

i)
dO o

i)
_
S
O
n
i
u
)
dS
O
.
(31)
The above functional degenerates to (2) for a perfectly plastic solid. For the velocity
eld in Eq. (5), minimization of velocity functional (31) give an integral equation
for the void growth rate D. By changing the integration variables to z = R
3
0
}R
3
and
= cos 0, this integral equation becomes
_
1
0
dz
_
1
0
(1 + 4Dz 3
2
)
_
[1 + 2Dz(1 3
2
) + 4D
2
z
2
]
N
+ 3
_
5
2
c
12N
!
R
0
Dz
4}3
z
_
1 + 2Dz + 4D
2
z
2
6Dz
2
d =
3o
m
o
f
, (32)
where o
m
= o

kk
}3 and c are the mean stress and eective strain in the remote eld,
respectively, o
f
= o
ref
c
N
is the ow stress corresponding to the remote eective strain
c, N the plastic work hardening exponent, ! the intrinsic material length in strain
gradient plasticity, and R
0
the void radius. The void size eect comes into play through
the ratio c
12N
!}R
0
. This ratio c
12N
!}R
0
degenerates to c!}R
0
for a perfectly plastic
solid (N = 0). Therefore, the void size eect disappears for voids much larger than
the intrinsic material length (R
0
!) or for vanishingly small remote eective strain
(c 0).
The void growth rate governed by (32) takes the form
D = D
_
o
m
o
f
,
c
12N
!
R
0
, N
_
= D
_
o
m
o
ref
c
N
,
c
12N
!
R
0
, N
_
. (33)
Fig. 4 shows the void growth rate D versus normalized mean stress o
m
}o
f
for c
12N
!}
R
0
=0, 0.024, 0.048 and 0.12 and N =0.2. The limit of c
12N
!}R
0
=0 corresponds to
classical plasticity. Similar to Fig. 1, small voids tend to grow slower than large voids,
which is consistent with the experimental and numerical studies of void growth (Fond
et al., 1996; Schluter et al., 1996; Fleck and Hutchinson, 1997; Huo et al., 1999;
Zhang et al., 1999; Khraishi et al., 2001; Zhang and Hsia, 2001). The void size has
essentially no eect on the void growth rate at small mean stresses (e.g., o
m
}o
Y
1),
but at relatively large mean stresses (e.g., o
m
}o
Y
2) the void size eect may become
signicant. The void size eect is also magnied at a large remote eective strain c
because of the ratio c
12N
!}R
0
(Fig. 5).
7. Concluding remarks
We have extended the RiceTracey (1969) model of void growth to account for the
void size eect. The following objectives have been achieved in this paper.
1184 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
Fig. 5. The void growth rate D versus normalized mean stress o
m
}o
f
for a power-law hardening solid and
c
12N
!}R
0
= 0, 0.024, 0.048 and 0.12, where D is dened in Eq. (6), o
m
and c are the mean stress and
eective strain in the remote eld, respectively, o
f
the ow stress in the remote eld corresponding to the
eective strain c, N = 0.2 the plastic work hardening exponent, ! the intrinsic material length in Eq. (15),
and R
0
the void radius. The curve for c
12N
!}R
0
=0 corresponds to the RiceTracey model (1969) of void
growth in classical plasticity.
(i) We have studied the size eect on void growth in ductile materials based on the
Taylor dislocation model. It is established that small voids tend to grow slower
than large voids. This is consistent with the prior experimental and numerical
studies of void growth, which cannot be explained by classical plasticity theories.
(ii) For a perfectly plastic solid, we have found that the void size eect comes into
play through the ratio c!}R
0
, where ! is the intrinsic material length given in Eq.
(15) and is on the order of microns, c the remote eective strain, and R
0
the
void size. This suggests that the void size eect can be magnied by the remote
eective strain. For micron-sized voids and small remote eective strain such that
c!}R
0
60.02, the void size inuences the void growth rate only at high stress
triaxialities. For sub-micron-sized voids and relatively large eective strain such
that c!}R
0
0.2, the void size has a signicant eect on the void growth rate at
all levels of stress triaxiality.
(iii) We have obtained analytically the asymptotic solutions of void growth rate at
high stress triaxialities accounting for the void size eect. Even for c!}R
0
60.02,
the void growth rate already deviates from the RiceTracey model (1969) of void
growth based on classical plasticity. For c!}R
0
0.2, the void growth rate scales
with the square of mean stress (o
m
}o
Y
)
2
, rather than the exponential function in
the RiceTracey model (1969), exp(3o
m
}2o
Y
).
(iv) For a power-law hardening solid, we have observed similar size eect on the void
growth rate, i.e., small voids grow slower than large voids. The void size comes
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1185
into play through the ratio c
12N
!}R
0
, where N is the plastic work hardening
exponent.
Acknowledgements
YH acknowledges the support from NSF (grant CMS-0084980 and a supplemental
to grant CMS-9896285 from the NSF International Program). KCH acknowledges the
support from NSFC.
References
Acharya, A., Bassani, J.L., 2000. Lattice incompatibility and a gradient theory of crystal plasticity. J. Mech.
Phys. Solids 48, 15651595.
Acharya, A., Beaudoin, A.J., 2000. Grain-size eect in viscoplastic polycrystals at moderate strains. J. Mech.
Phys. Solids 48, 22132230.
Arsenlis, A., Parks, D.M., 1999. Crystallographic aspects of geometrically necessary and statistically stored
dislocation density. Acta Mater. 47, 15971611.
Ashby, M.F., 1970. The deformation of plastically non-homogeneous alloys. Philos. Mag. 21, 399424.
Ashby, M.F., Blunt, F.J., Bannister, M., 1989. Flow characteristics of highly constrained metal wires. Acta
Metall. 37, 1857.
Bagchi, A., Evans, A.G., 1996. The mechanics and physics of thin lm decohesion and its measurement.
Interface Sci. 3, 169193.
Bagchi, A., Lucas, G.E., Suo, Z., Evans, A.G., 1994. A new procedure for measuring the decohesion energy
of thin ductile lms on substrates. J. Mater. Res. 9, 17341741.
Bishop, J.F.W., Hill, R., 1951a. A theory of plastic distortion of a polycrystalline aggregate under combined
stresses. Philos. Mag. 42, 414427.
Bishop, J.F.W., Hill, R., 1951b. A theoretical derivation of the plastic properties of a polycrystalline
face-centered metal. Philos. Mag. 42, 12981307.
Budiansky, B., Hutchinson, J.W., Slutsky, S., 1982. Void growth and collapse in viscous solids. In: Hopkins,
H.G., Sewll, M.J. (Eds.), Mechanics of Solids, The Rodney Hill 60th Anniversary Volume. Pergamon
Press, Oxford, pp. 1345.
Chen, B., Wu, P.D., Xia, Z.C., MacEwen, S.R., Tang, S.C., Huang, Y., 2000. A dilatational plasticity theory
for aluminum sheets. In: Chuang, T.-J., Rudnicki, J.W. (Eds.), Multiscale Deformation and Fracture in
Materials and Structures, the James R. Rice 60th Anniversary Volume. Kluwer Academic Publishers,
Dordrecht, The Netherlands, pp. 1730.
Chen, B., Huang, Y., Liu, C., Wu, P.D., MacEwen, S.R., 2003. A dilatational plasticity theory for viscoplastic
materials. Mech. Mater., in press.
Cleveringa, H.H.M., van der Giessen, E., Needleman, A., 1997. Comparison of discrete dislocation and
continuum plasticity predictions for a composite material. Acta Mater. 45, 31633179.
Cleveringa, H.H.M., van der Giessen, E., Needleman, A., 1998. Discrete dislocation simulations and size
dependent hardening in single slip. J. Phys. IV 8, 8392.
Cleveringa, H.H.M., van der Giessen, E., Needleman, A., 1999a. A discrete dislocation analysis of bending.
Int. J. Plasticity 15, 837868.
Cleveringa, H.H.M., van der Giessen, E., Needleman, A., 1999b. A discrete dislocation analysis of residual
stresses in a composite material. Philos. Mag. A 79, 893920.
Cleveringa, H.H.M., van der Giessen, E., Needleman, A., 2000. A discrete dislocation analysis of mode I
crack growth. J. Mech. Phys. Solids 48, 11331157.
Cottrell, A.H., 1964. The Mechanical Properties of Materials. Wiley, New York, p. 277.
Douglass, M.R., 1998. Lifetime estimates and unique failure mechanisms of the digital micromirror device
(DMD). Annual ProceedingsReliability Physics Symposium (Sponsored by IEEE), March 31April 2,
1998, pp. 916; also http://www.dlp.com/dlp/resources/whitepapers/pdf/ieeeir.pdf.
1186 B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187
Elssner, G., Korn, D., Ruehle, M., 1994. The inuence of interface impurities on fracture energy of UHV
diusion bonded metal-ceramic bicrystals. Scripta Metall. Mater. 31, 10371042.
Fleck, N.A., Hutchinson, J.W., 1993. A phenomenological theory for strain gradient eects in plasticity. J.
Mech. Phys. Solids 42, 18251857.
Fleck, N.A., Hutchinson, J.W., 1997. Strain gradient plasticity. In: Hutchinson, J.W., Wu, T.Y. (Eds.),
Advances in Applied Mechanics, Vol. 33. Academic Press, New York, pp. 295361.
Fleck, N.A., Hutchinson, J.W., 2001. A reformulation of strain gradient plasticity. J. Mech. Phys. Solids 49,
22452271.
Fleck, N.A., Muller, G.M., Ashby, M.F., Hutchinson, J.W., 1994. Strain gradient plasticity: theory and
experiment. Acta Metall. Mater. 42, 475487.
Fond, C., Lobbrecht, A., Schirrer, R., 1996. Polymers toughened with rubber microspheres: an analytical
solution for stresses and strains in the rubber particles at equilibrium and rupture. Int. J. Fract. 77, 141
159.
Gao, H., Huang, Y., Nix, W.D., 1999a. Modeling plasticity at the micrometer scale. Naturwissenschaften 86,
507515.
Gao, H., Huang, Y., Nix, W.D., Hutchinson, J.W., 1999b. Mechanism-based strain gradient plasticityI.
Theory. J. Mech. Phys. Solids 47, 12391263.
Gurson, A.L., 1977. Continuum theory of ductile rupture by void nucleation and growth: part 1yield
criteria and ow rules for porous ductile media. ASME Trans. J. Eng. Mater. Tech. 99, 215.
Gurtin, M.E., 2000. On the plasticity of single crystals: free energy, microforces, plastic-strain gradients. J.
Mech. Phys. Solids 48, 9891036.
Gurtin, M.E., 2002. A gradient theory of single-crystal viscoplasticity that accounts for geometrically
necessary dislocations. J. Mech. Phys. Solids 50, 532.
Haque, A., Saif, M.T.A., 2003. Strain gradient eect in nanoscale thin lm. Acta Mater., in press.
Huang, Y., 1991. Accurate dilatation rate for spherical voids in triaxial stress elds. ASME Trans. J. Appl.
Mech. 58, 10841086.
Huang, Y., Hutchinson, J.W., Tvergaard, V., 1991. Cavitation instabilities in elasticplastic solids. J. Mech.
Phys. Solids 39, 223241.
Huang, Y., Gao, H., Hwang, K.C., 1999. Strain-gradient plasticity at the micron scale. In: Ellyin, F., Provan,
J.W. (Eds.), Progress in Mechanical Behavior of Materials, Vol. III. pp. 10511056.
Huang, Y., Gao, H., Nix, W.D., Hutchinson, J.W., 2000a. Mechanism-based strain gradient plasticityII.
analysis. J. Mech. Phys. Solids 48, 99128.
Huang, Y., Xue, Z., Gao, H., Nix, W.D., Xia, Z.C., 2000b. A study of micro-indentation hardness tests by
mechanism-based strain gradient plasticity. J. Mater. Res. 15, 17861796.
Huo, B., Zheng, Q.-S., Huang, Y., 1999. A note on the eect of surface energy and void size to void growth.
Euro. J. Mech.A/Solids 18, 987994.
Hwang, K.C., Jiang, H., Huang, Y., Gao, H., Hu, N., 2002. A nite deformation theory of strain gradient
plasticity. J. Mech. Phys. Solids 50, 8189.
Hwang, K.C., Jiang, H., Huang, Y., Gao, H., 2003. Finite deformation analysis of mechanism-based strain
gradient plasticity: torsion and crack tip eld. Int. J. Plasticity 19, 235251.
Jiang, H., Huang, Y., Zhuang, Z., Hwang, K.C., 2001. Fracture in mechanism-based strain gradient plasticity.
J. Mech. Phys. Solids 49, 979993.
Khraishi, T.A., Khaleel, M.A., Zbib, H.M., 2001. Parametric-experimental study of void growth in
superplastic deformation. Int. J. Plasticity 17, 297315.
Kocks, U.F., 1970. The relation between polycrystal deformation and single crystal deformation. Metall.
Trans. 1, 11211144.
Lloyd, D.J., 1994. Particle reinforced aluminum and magnesium matrix composites. Int. Mater. Rev. 39, 1
23.
Ma, Q., Clarke, D.R., 1995. Size dependent hardness in silver single crystals. J. Mater. Res. 10, 853863.
McElhaney, K.W., Vlassak, J.J., Nix, W.D., 1998. Determination of indenter tip geometry and indentation
contact area for depth-sensing indentation experiments. J. Mater. Res. 13, 13001306.
Needleman, A., 2000. Computational mechanics at the mesoscale. Acta Mater. 48, 105124.
Nix, W.D., 1989. Mechanical properties of thin lms. Metall. Trans. 20A, 22172245.
B. Liu et al. / J. Mech. Phys. Solids 51 (2003) 11711187 1187
Nix, W.D., 1997. Elastic and plastic properties of thin lms on substrates nanoindentation techniques. Mater.
Sci. Eng. A234236, 3744.
Nix, W.D., Gao, H., 1998. Indentation size eects in crystalline materials: a law for strain gradient plasticity.
J. Mech. Phys. Solids 46, 411425.
Nye, J.F., 1953. Some geometrical relations in dislocated crystals. Acta Metall. Mater. 1, 153162.
Poole, W.J., Ashby, M.F., Fleck, N.A., 1996. Micro-hardness of annealed and work-hardened copper
polycrystals. Scripta Metall. Mater. 34, 559564.
Qiu, X., Huang, H., Nix, W.D., Hwang, K.C., Gao, H., 2001. Eect of intrinsic lattice resistance in strain
gradient plasticity. Acta Mater. 49, 39493958.
Rice, J.R., Tracey, D.M., 1969. On the ductile enlargement of voids in triaxial stress elds. J. Mech. Phys.
Solids 17, 201217.
Saha, R., Xue, Z., Huang, Y., Nix, W.D., 2001. Indentation of a soft metal lm on a hard substrate: strain
gradient hardening eects. J. Mech. Phys. Solids 49, 19972014.
Schluter, N., Grimpe, F., Bleck, W., Dahl, W., 1996. Modelling of the damage in ductile steels. Comp.
Mater. Sci. 7, 2733.
Shi, M., Huang, Y., Jiang, H., Hwang, K.C., Li, M., 2001. The boundary layer eect on the crack tip eld
in mechanism-based strain gradient plasticity. Int. J. Fract. 112, 2341.
Shrotriya, P., Allameh, S.M., Lou, J., Buchheit, T., Soboyejo, W.O., 2003. On the measurement of the
plasticity length scale parameter in LIGA nickel foils. Mech. Mater. 35, 233243.
Shu, J.Y., 1998. Scale-dependent deformation of porous single crystals. Int. J. Plasticity 14, 10851107.
Stelmashenko, N.A., Walls, M.G., Brown, L.M., Milman, Y.V., 1993. Microindentation on W and Mo
oriented single crystals: an STM study. Acta Metall. Mater. 41, 28552865.
Stolken, J.S., Evans, A.G., 1998. A microbend test method for measuring the plasticity length scale. Acta
Mater. 46, 51095115.
Swadener, J.G., George, E.P., Pharr, G.M., 2002. The correlation of the indentation size eect measured
with indenters of various shapes. J. Mech. Phys. Solids 50, 681694.
Taylor, G.I., 1934. The mechanism of plastic deformation of crystals. Part Itheoretical. Proc. Roy. Soc.
(London) A145, 362387.
Taylor, G.I., 1938. Plastic strain in metals. J. Inst. Metals 62, 307324.
Tvergaard, V., 1990. Material failure by void growth to coalescence. Adv. Appl. Mech. 27, 83147.
Tvergaard, V., Huang, Y., Hutchinson, J.W., 1992. Cavitation instabilities in a power hardening elasticplastic
solid. Euro. J. Mech. A/Solids 11, 215231.
Wiedersich, H., 1964. Hardening mechanisms and the theory of deformation. J. Metals 16, 425430.
Xue, Z., Huang, Y., Li, M., 2002a. Particle size eect in metallic materials: a study by the theory of
mechanism-based strain gradient plasticity. Acta Mater. 50, 149160.
Xue, Z., Saif, T.M.A., Huang, Y., 2002b. The strain gradient eect in micro-electro-mechanical systems
(MEMS). J. Microelectromechanical Systems 11, 2735.
Zhang, S., Hsia, K.J., 2001. Modeling the fracture of a sandwich structure due to cavitation in a ductile
adhesive layer. ASME Trans. J. Appl. Mech. 68, 93100.
Zhang, K.S., Bai, J.B., Francois, D., 1999. Ductile fracture of materials with high void volume fraction. Int.
J. Solids Struct. 36, 34073425.

You might also like