Вы находитесь на странице: 1из 15

Journal of the Mechanics and Physics of Solids

49 (2001) 26892703
www.elsevier.com/locate/jmps
Resistance curves for mixed mode interface crack
growth between dissimilar elasticplastic solids
Viggo Tvergaard

Department of Solid Mechanics, Technical University of Denmark, Building 404, DK-2800 Lyngby,
Denmark
Abstract
For crack growth along an interface joining two elasticplastic solids resistance curves are
analysed numerically, under conditions of small scale yielding. As mismatch of elastic properties
is accounted for, the corresponding oscillating stress singularity elds are applied as boundary
conditions on the outer edge of the region analysed. The fracture process is represented in terms
of a cohesive zone model, for which the work of separation per unit area and the peak stress
required for separation are basic parameters. Eects of dierent kinds of mismatch between the
two materials are analysed, including cases where the substrate remains elastic as well as cases
where both materials yield plastically. For all the cases studied it is shown that plastic ow
near the crack tip results in much larger resistance to crack growth when mode II conditions
dominate at the crack tip than when mode I conditions dominate. ? 2001 Elsevier Science Ltd.
All rights reserved.
Keywords: A. Crack tip plasticity; Fracture process; B. elasticplastic material
1. Introduction
It is well understood that plastic work during crack growth in metals contributes
signicantly to the fracture toughness, so that the macroscopic work of fracture is much
larger than the work absorbed by the the local fracture process required to separate
the crack surfaces. This has been studied in more detail in a number of elasticplastic
crack growth computations (Tvergaard and Hutchinson, 1992, 1993, 1994a), in which
the fracture process has been modelled by a traction-separation law along the crack
plane with a specied work of separation per unit area, while the material has been
modelled as elasticplastic. One of these studies (Tvergaard and Hutchinson, 1993)

Tel.: +45-4525-4273; fax: +45-4593-1475.


E-mail addresses: viggo@mek.dtu.dk (V. Tvergaard).
0022-5096/01/$ - see front matter ? 2001 Elsevier Science Ltd. All rights reserved.
PII: S0022- 5096( 01) 00074- 6
2690
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
has focussed on the special case of crack growth along an interface joining an elastic
plastic solid to a solid which does not yield plastically.
Experimental studies of interface crack growth, where plastic yielding occurs in at
least one of the solids joined at the interface (Cao and Evans, 1989; Liechti and Chai,
1992; ODowd et al., 1992), have shown a strong dependence on the mode mixity,
such that the measured fracture toughness is much higher in cases where mode II
loading dominates at the crack tip than in cases where mode I loading dominates. This
was also predicted by Tvergaard and Hutchinson (1993), and it was concluded that
the observed strong dependence on the mode of loading can be explained by plastic
yielding. However, it is noted that also models of asperity interaction, representing
eects of friction due to rough crack surfaces, have shown a somewhat increased
toughness in cases where mode II loading dominates (Evans and Hutchinson, 1989;
Jensen, 1990).
A traction-separation law could be used to represent cleavage fracture by atomic
separation, but the stress peaks during atomic separation are very high, and it has
been found by Tvergaard and Hutchinson (1992, 1993) that crack growth predictions
based on a cohesive zone model of the fracture process are limited by the fact that
according to continuum plasticity crack solutions the maximum stresses near the crack
tip do not exceed about ve times the ow stress. However, in these cases the length
scale of the fracture process is typically much smaller than the dislocation spacing or
the elastic cell size within the dislocation structure, and therefore continuum plasticity
would not be realistic near the crack tip, assuming that the crack tip does not emit
dislocations. To model the elastic core region near the crack tip Suo et al. (1993) have
proposed the use of a long elastic strip of material, inside which the crack propagates,
and Beltz et al. (1996) have proposed a self-consistent way of identifying the width
of the strip. Wei and Hutchinson (1996, 1997) have used this elastic strip model in
steady-state analyses to study interface toughness of metalceramic joints under mixed
mode loading, and Tvergaard (1999) has used the model to determine crack growth
resistance curves for the same type of cases. The conclusion of these studies is that also
for cleavage fracture by atomic separation plastic yielding of the material outside the
elastic core region can signicantly increase the fracture toughness. And also in these
situations the predicted fracture toughness is higher in cases where mode II loading
dominates than when mode I loading dominates.
In the present paper the resistance curve behaviour is analysed for crack growth along
the interface between two elasticplastic solids. It is assumed that fracture occurs by a
ductile mechanism, involving voids or small poorly bonded regions on the interface, so
that fracture can be represented by a traction-separation law with a peak stress that does
not exceed either of the two ow stresses by a large factor. The analyses are restricted
to conditions of small scale yielding, with mixed mode loading conditions applied far
away from the crack tip. Thus, the elastic crack tip elds with an oscillating singularity
are applied as boundary conditions, as in the analyses of Tvergaard and Hutchinson
(1993) for a rigid substrate, but in the present studies the stress and strain elds are
determined on both sides of the interface. The analyses are used to study eects of
dierences between the elastic properties of the two materials as well as eects of
dierences between the plastic properties.
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2691
Fig. 1. Interface crack with dierent elasticplastic material properties on each side of the crack plane.
2. Problem formulation
In the plane strain interface crack problem analysed by Tvergaard and Hutchinson
(1993) one of the materials is taken to be elasticplastic while the other material
is rigid. With both materials represented as deformable the present study includes the
previous one as the limit where both Youngs modulus and the yield stress tend towards
innity for the substrate. But the present study also accounts for the symmetric problem
of crack growth through a homogeneous solid (as in Tvergaard and Hutchinson, 1992)
and for all mixed mode problems with dierent properties of the two materials.
2.1. Small scale yielding formulation
The equations governing the elastic crack tip singularity eld of a semi-innite in-
terface crack have been given by Rice (1988); see also Tvergaard and Hutchinson
(1993). This elastic crack problem was done long ago by many authors (e.g. England,
1965). The crack has tractions acting on the interface, which are given in terms of the
two stress intensity factor components, K
I
and K
II
, by
o
22
+ io
12
=(K
I
+ iK
II
)(2r)
1}2
r
ic
. (2.1)
Here, r is the distance from the tip, i =

1, c is the oscillation index


c =
1
2
ln
_
1 [
1 + [
_
(2.2)
and [ is the second Dundurs parameter
[ =
1
2
j
1
(1 2v
2
) j
2
(1 2v
1
)
j
1
(1 v
2
) + j
2
(1 v
1
)
. (2.3)
The shear moduli are j
1
=E
1
}((1 + v
1
)) and j
2
=E
2
}(2(1 + v
2
)), where E
1
, v
1
are the
elastic constants for material No. 1, i.e. the material located at x
2
0 (see Fig. 1),
2692
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
while E
2
, v
2
are the elastic constants for material No. 2. The relation between the energy
release rate and the magnitude |K| of stress intensity factors is
G =
1
2
(1 [
2
)
_
1 v
2
1
E
1
+
1 v
2
2
E
2
_
|K|
2
, |K| =
_
K
2
I
+ K
2
II
. (2.4)
With a reference length L chosen to characterize the remote eld an L-dependent
measure of mode mixity is dened by
tan =
Im[(K
I
+ iK
II
)L
ic
]
Re[(K
I
+ iK
II
)L
ic
]
(2.5)
which reduces to the more familiar measure, tan =K
II
}K
I
, when c =0. By using
Eq. (2.1) in Eq. (2.5) it is seen that tan =o
12
}o
22
at r =L on the interface. Thus,
measures the relative proportion of shear to normal stress on the interface a distance
L from the tip, as predicted by the elastic solution. The displacement components
associated with the singularity eld, with amplitude |K|, are specied in Tvergaard
and Hutchinson (1993). The mode mixity

at the radius r is given in terms of the
L-dependent measure of the mode mixity, , as

= + c ln(r}L). (2.6)
The materials considered here are taken to be elasticplastic with the true stress-logarithmic
strain curve in uniaxial tension specied by
c =
_
o}E for o 6o
Y
,
(o
Y
}E)(o}o
Y
)
1}N
for o o
Y
.
(2.7)
Here, o
Y
is the initial yield stress and N is the power hardening exponent, while E
and v are Youngs modulus and Poissons ratio, respectively. The tensile behaviour is
generalized to multiaxial stress states assuming isotropic hardening and using the Mises
yield surface. For material No. 1 the parameters are denoted E
1
, v
1
, o
Y1
and N
1
, while
for material No. 2 they are E
2
, v
2
, o
Y2
and N
2
(see Fig. 1).
In the numerical analyses nite strains are accounted for, using a convected coor-
dinate, Lagrangian formulation of the eld equations, in which q
i)
and G
i)
are metric
tensors in the reference conguration and the current conguration, respectively, with
determinants q and G, and p
i)
=1}2(G
i)
q
i)
) is the Lagrangian strain tensor. The
contravariant components t
i)
of the Kirchho stress tensor on the current base vectors
are related to the components of the Cauchy stress tensor o
i)
by t
i)
=
_
G}qo
i)
. Then,
in the nite-strain generalization of J
2
-ow theory discussed by Hutchinson (1973),
an incremental stressstrain relationship is obtained of the form t
i)
=L
i)k!
p
k!
(e.g. see
Tvergaard and Hutchinson, 1993).
2.2. Traction-separation law at interface
A traction-separation law proposed by Needleman (1987) for separation due to the
normal stress on an interface was generalized by Tvergaard (1990) to also account for
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2693
Fig. 2. Specication of traction-separation relation.
separation due to tangential separation. A special version of this generalized model, for
which the traction-separation law is governed by a potential, was used by Tvergaard
and Hutchinson (1993). Here, o
n
and o
t
denote the normal and tangential components
of the relative displacement of the crack faces across the interface in the zone where
the fracture processes are occurring, as indicated in Fig. 2. When o
c
n
and o
c
t
are criti-
cal values of these displacement components and a single non-dimensional separation
measure is dened as
z =
_
_
o
n
o
c
n
_
2
+
_
o
t
o
c
t
_
2
_
1}2
(2.8)
the tractions drop to zero when z =1. With o(z) displayed in Fig. 2, a potential from
which the tractions are derived is dened as
1(o
n
o
t
) =o
c
n
_
z
0
o(z

) dz

. (2.9)
The normal and tangential components of the traction acting on the interface in the
fracture process zone are given by
1
n
=
c1
co
n
=
o(z)
z
o
n
o
c
n
, 1
t
=
c1
co
t
=
oz
z
o
t
o
c
t
o
c
n
o
c
t
. (2.10)
The traction law under a purely normal separation (o
t
=0), is 1
n
=o(z) where z =o
n
}o
c
n
.
Under a purely tangential displacement (o
n
=0), 1
t
=(o
c
n
}o
c
t
)o(z) where z =o
t
}o
c
t
. The
peak normal traction under pure normal separation is o, and the peak shear traction
is (o
c
n
}o
c
t
) o in a pure tangential separation. The work of separation per unit area of
interface is given by Eq. (2.9) with z =1. For the separation function o(z) specied
in Fig. 2,
I
0
=
1
2
oo
c
n
(1 z
1
+ z
2
). (2.11)
The parameters governing the separation law are therefore the work I
0
of the fracture
process, the peak stress quantity o and the critical displacement ratio o
c
n
}o
c
t
, together
with the factors z
1
and z
2
governing the shape of the separation function. It has been
found by Tvergaard and Hutchinson (1992, 1993) that the details of the shape of the
2694
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
separation law are relatively unimportant, and that the two most important parameters
characterizing the fracture process in this model are I
0
and o.
2.3. Reference values
As mentioned above, the remote loading is specied in terms of the three parameters
|K|, , and L. Using Eqs. (2.4) and (2.11) a reference stress intensity factor is dened
as
K
0
=
_
1 v
2
1
E
1
+
1 v
2
2
E
2
_
1}2
_
2I
0
1 [
2
_
1}2
. (2.12)
Here, K
0
represents the value of |K| needed to advance the interface crack in the
absence of any plasticity. This value is independent of since a potential is used to
generate the relation of tractions to crack face displacements of the interface. A length
quantity R
0
, which scales with the size of the plastic zone in material No. 1 (when
|K|

= K
0
), is dened as
R
0
=
1
3
_
K
0
o
Y1
_
2
=
2
3
_
1 v
2
1
E
1
+
1 v
2
2
E
2
_
1
I
0
(1 [
2
)o
2
Y1
. (2.13)
While the mode mixity measure refers to the distance L from the tip, it is natural to
dene a reference measure of mixity,
0
, based on the reference length R
0
. By using
Eq. (2.6), and
0
are related by

0
= + c ln(R
0
}L). (2.14)
In the case to be analysed here the material parameters in materials 1 and 2, and thus
the value of the oscillation index c according to Eq. (2.2), vary from case to case, and
this means that the phase shift c ln(R
0
}L) in Eq. (2.14) will dier from case to case.
3. Numerical procedure
The numerical solutions are obtained by a crack growth procedure analogous to that
used by Tvergaard and Hutchinson (1993). Thus, a nite element approximation of the
displacement elds is used in a linear incremental method, with a Cartesian coordinate
system x
i
as reference, and with the displacement components on the reference base
vectors denoted by u
i
. The Lagrangian strain increments are given by
p
i)
=
1
2
( u
i, )
+ u
), i
+ u
k
, )
u
k, )
+ u
k
, i
u
k, )
), (3.1)
where ( )
, )
denotes the covariant derivative in the reference frame. Furthermore, the
incremental principle of virtual work takes the form
_
J
{ t
i)
op
i)
+ t
i)
u
k
, i
u
k, )
} dJ =
_
S

1u
i
dS
__
J
t
i)
op
i)
dJ
_
S
1
i
u
i
dS
_
, (3.2)
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2695
Fig. 3. Mesh used for some of the crack growth analyses.
where J and S are the reference volume and surface of the region analysed, 1
i
are the
nominal traction components, and the bracketed terms are included to prevent drifting
of the solution away from the true equilibrium path.
An example of the mesh used for the computations is shown in Fig. 3, where it
is seen that a uniform mesh region is used in the range where crack growth along
the interface is studied. The length of one square element inside the uniform mesh is
denoted A
0
, and the initial crack tip is located at x
1
=0. Most of the computations
are carried out with 60 6 quadrilaterals in the uniform mesh along the interface, and
a few computations have used a 120 6 mesh to allow for more crack growth. The
elements used are quadrilaterals each built-up of four triangular, linear-displacement
elements. The outer radius of the region analysed is chosen to be A
0
}A
0
=8000, in
order that the plastic zone size should not exceed A
0
}10.
On the free crack surfaces, for x
2
=0 and x
1
0, zero tractions, 1
1
=1
2
=0, are pre-
scribed, both for materials 1 and 2 (see Fig. 1). On the remaining part of the interface,
x
2
=0 and x
1
0, the displacements and tractions are related by the traction-separation
law (2.10). During the initial part of the crack growth resistance curve an increment of
|K| is prescribed, but this procedure becomes unstable when |K| approaches its asymp-
tote. Then, a RayleighRitz nite-element method (Tvergaard, 1976) is used to control
a monotonic increase in displacement dierences across the crack tip. Thus, nodal dis-
placements at the top and bottom of the interface are used as mode amplitudes in
the RayleighRitz nite-element procedure, and for each increment of the numerical
solution an appropriate dierence between such displacements in the fracture process
region is chosen as the parameter to be prescribed.
In all computations the properties of material No. 1 are specied as o
Y1
}E
1
=0.003,
v
1
=1}3 and N
1
=0.1, while dierent values are considered for the properties of
2696
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
material No. 2. In the traction-separation law the values o
c
n
}o
c
t
=1, o
c
n
=0.1A
0
, z
1
=0.15
and z
2
=0.50 are used, while o}o
Y1
is varied.
At some stages of the deformation the value of the J-integral is calculated on a
number of contours around the crack tip to check agreement with the prescribed am-
plitude |K| of the edge displacements, representing the |K| eld in the elastic material
far away from the tip. The J-integral (Rice, 1968) was formulated in the nite-strain
context by Eshelby (1970)
J =
_
I
{W dx
2
1
i
u
i, 1
ds}, W =
_
p
i)
0
t
i)
op
i)
, (3.3)
where I is some path in the reference conguration from the lower crack surface to
the upper crack surface, ds is an arc length element along the path, and 1
i
are the
nominal tractions on the boundary of the region enclosed by the contour I. The path
independence of J is conrmed by these calculations, and good agreement is found
with the value of the energy release rate G given by Eq. (2.4) in terms of |K|.
It is noted that the present problem formulation allows for switching the material
properties between regions 1 and 2, since full elasticplastic material behaviour is ac-
counted for in both regions. It has been checked for a few cases that the identical crack
growth resistance curve is predicted when dissimilar material properties are switched
between the two regions, while the sign of the applied mode mixity is changed at the
same time.
4. Results
The resistance curve analyses to be presented rst focus on the eect of elastic mod-
ulus mismatch in cases where material No. 2 does not yield plastically, while material
No. 1 is elasticplastic. In all these analyses the elasticplastic material parameters
are taken to be o
Y1
}E
1
=0.003, v
1
=1}3 and N
1
=0.1, while the elastic substrate has
o
Y2
and v
2
=1}3. The values of E
2
}E
1
to be considered are 1, 2, 6 and 1000.
For each elastic modulus mismatch a number of dierent values of the peak stress to
ow stress ratio, o}o
Y1
, are considered, and the eect of the mode mixity is studied
by considering dierent values of
0
in Eq. (2.14). This mixity measure should give
a good indication of the mode mixity in the process region at the crack tip, since it is
based on the reference length R
0
, which scales with the plastic zone size at the onset
of crack growth.
Fig. 4 shows four calculated crack growth resistance curves, which all correspond
to the same values of the material parameters in the elasticplastic material, while
material No. 2 does not yield (o
Y2
). The same traction separation law is used
with o}o
Y1
=3.0 in all four computations, and the value of the angular measure
0
is
near 0

in all cases, so that the crack tip conditions are close to mode I . Thus, the only
dierence between these four cases is the value of Youngs modulus, E
2
, in material
No. 2. It is seen that for E
2
}E
1
=1 plasticity has only a small eect on the interface
toughness, increasing the value of |K|}K
0
to 1.12 after a small amount of crack growth,
while plasticity has a larger eect for E
2
}E
1
=2 and E
2
}E
1
=6. In the case of a nearly
rigid material No. 2, E
2
}E
1
=1000, the resistance curve shows that |K|}K
0
increases to
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2697
Fig. 4. Interface crack growth resistance curves for o}o
Y1
=3.0, with o
Y1
}E
1
=0.003 and o
Y2
.
about 2.38, and here this maximum toughness is reached after more crack growth, at
a}R
0
=2.3. For the four dierent values of E
2
}E
1
considered in Fig. 4 the inuence
of the bond strength, o}o
Y1
, and of the mode mixity,
0
, is illustrated in Fig. 5, by
showing the steady state toughness |K|
SS
}K
0
.
In Fig. 5a there is no elastic mismatch, so that [ =0 results in c =0. It is seen that
plasticity gives a strong sensitivity to the mode mixity so that the minimum fracture
toughness for a given value of o}o
Y1
occurs somewhere near
0
=0

, while a notice-
able mode II contribution results in much increased fracture toughness. For the smallest
interface strength analysed, o}o
Y1
=1.5, there is an interval between
0
15

and

0
15

where the toughness is constant, |K|


SS
}K
0
=1, thus illustrating that in the
elastic range a strong mode II contribution does not increase the toughness. The in-
creased toughness found for values of
0
outside this interval is a result of plasticity.
The curves in Fig. 5a also show that for values of o}o
Y1
greater than 4.0 the toughness
increases rapidly with small increases of the peak stress in the traction-separation law.
Figs. 5bd show a picture very similar to that in Fig. 5a. In these three gures the
values of the oscillation index c are 0.0266, 0.0575 and 0.0811, respectively, so in all
these cases the oscillating singularity plays a role, and there is a dierence between
the mode mixity applied at the outer boundary of the region analysed, and that acting
in the process region near the crack tip. Fig. 4 has illustrated, for a given value of
o}o
Y1
, that the toughness increases when the value of E
2
}E
1
is increased, and it is
seen in Fig. 5 that this is a general trend for all values of the interface strength. The
case of E
2
}E
1
=1000 in Fig. 5d approaches that of an elasticplastic solid bonded to
a rigid substrate, as studied by Tvergaard and Hutchinson (1993), and it is seen that
the results in Fig. 5d are close to those found for the rigid substrate.
The computational model has also been used to study cases where both materials
undergo plastic yielding. For cases where E
2
}E
1
=2, as in Fig. 5b, examples of crack
growth resistance curves are shown in Fig. 6. In these cases the values of
0
are near 0

(as in Fig. 4), so that the resistance curves shown correspond to situations where mode
I conditions are dominant near the crack tip. The dashed curve for o}o
Y1
=3.0 and
o
Y2
is very similar to one of the resistance curves shown in Fig. 4, while the other
2698
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
Fig. 5. Steady-state interface toughness as a function of the local mixity measure
0
, for o
Y1
}E
1
=0.003
and o
Y2
, considering dierent values of o}o
Y1
. (a) E
2
}E
1
=1, (b) E
2
}E
1
=2, (c) E
2
}E
1
=6, (d)
E
2
}E
1
=1000.
dashed curve, for o
Y2
}o
Y1
=1, illustrates that additional plastic yielding in the substrate
adds noticeably to the fracture toughness. The solid curves in Fig. 6 are obtained for
o}o
Y1
=4.0, with values 2.0, 1.5 and 1.25 of the ow stress ratio, o
Y2
}o
Y1
, and with
the curve for o
Y2
included for reference. The resistance curve for o
Y2
}o
Y1
=1.25
shows a large increase in the maximum fracture toughness, compared to the result for
an elastic substrate (o
Y2
}o
Y1
).
The inuence of plasticity in the substrate is further illustrated in Fig. 7 by curves
of steady state toughness vs. mode mixity. As in Fig. 6 the ratio of Youngs moduli
is E
2
}E
1
=2, and a number of dierent values of the ow stress ratio o
Y2
}o
Y1
are
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2699
Fig. 6. Interface crack growth resistance curves for various values of o}o
Y1
and o
Y2
}o
Y1
, with E
2
}E
1
=2
and o
Y1
}E
1
=0.003.
Fig. 7. Steady-state interface toughness as a function of the local mixity measure
0
, for E
2
}E
1
=2 and
o
Y1
}E
1
=0.003. Dierent values of o
Y2
}o
Y1
and o}o
Y1
are considered.
considered. The three curves for o
Y2
in Fig. 7 are identical to three of the curves
shown in Fig. 5b, while the remaining curves in Fig. 7 illustrate the increase in fracture
toughness resulting from plastic yielding in the substrate. If we focus on o
Y2
}o
Y1
=2.0,
no dierence from o
Y2
is found for o}o
Y1
=3.0, whereas a small dierence is
found for o}o
Y1
=3.5, and a clear increase of the toughness is found for o}o
Y1
=4.0
in the whole range of
0
values analysed. For o
Y2
}o
Y1
=1.5 the same tendency is
observed, with a much stronger toughness increase at the higher interface strength.
For o
Y2
}o
Y1
=1.25 the curve corresponding to the highest interface strength is just
visible at the top of the diagram in Fig. 7. Thus, with the same elasticplastic material
2700
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
Fig. 8. Steady-state interface toughness as a function of the local mixity measure
0
, for E
2
}E
1
=1 and
o
Y1
}E
1
=0.003. Dierent values of o
Y2
}o
Y1
and o}o
Y1
are considered.
behaviour for material No. 1 the curves in Figs. 6 and 7 show a strong sensitivity to
whether or not plastic yielding takes place in the substrate. The toughness is increased
by plastic ow in the substrate, and this increase is larger the lower the initial yield
stress in the substrate. Also, the sensitivity to yielding in the substrate is much larger
in cases where the bond strength is higher, i.e. for a higher value of o}o
Y1
.
Curves of steady state toughness vs. mode mixity, analogous to those in Fig. 7, are
shown in Fig. 8 for a case where there is no elastic mismatch, E
2
}E
1
=1, so that the
only mismatch is that due to a dierence in ow stress. Again, the two curves for
o
Y2
}o
Y1
are identical to two of the curves shown in Fig. 5a, and plastic ow in
the substrate results in an increased fracture toughness. The curves for o
Y2
}o
Y1
=1.5
illustrate that also here the eect of plastic yielding in the substrate is much larger if
the interface strength is higher.
The dependence of the steady-state fracture toughness on the value of o}o
Y1
is
shown in Fig. 9 for a number of dierent cases of material mismatch. The values
used to plot these curves correspond to the minima of the toughness vs. mode mixity
curves as those shown in Figs. 5, 7 and 8. It is noted that analogous curves have been
calculated by Tvergaard and Hutchinson (1994b, 1996) in studies of the toughness of
an interface along a thin ductile layer joining elastic solids, but these results represent
remote mode I loading rather than the minimum of the toughness vs. mixity curve.
The solid curves in Fig. 9, corresponding to o
Y2
}o
Y1
, illustrate how increasing
elastic stiness of the substrate gives increasing fracture toughness, for a given value of
o}o
Y1
. Curves representing plastic yielding of the substrate are shown for o
Y2
}o
Y1
=2
and o
Y2
}o
Y1
=1. Both sets of curves illustrate the trends found in Figs. 7 and 8 that
plastic yielding of the substrate increases the toughness, and the curves quantify how
this behaviour varies with the value of o}o
Y1
.
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2701
Fig. 9. Steady-state interface toughness as a function of o}o
Y1
, for various values of E
2
}E
1
and o
Y2
}o
Y1
.
The values shown correspond to minima of curves like those in Figs. 7 and 8.
5. Discussion
The mixed mode toughness of an interface joining an elasticplastic solid to a
solid which does not yield plastically has been analysed by Tvergaard and Hutchinson
(1993), based on the simplifying assumption that the non-yielding substrate is rigid.
The present study extends this work to account for deformations of the substrate, still
with full consideration of mode mixity eects and of the oscillating singularity due
to elastic mismatch, and still using a traction separation law to represent the fracture
process along the interface. One of the results obtained is the quantied understanding
of the eect of the elastic stiness of a non-yielding substrate. In addition, results are
obtained for the eect of plastic yielding of the substrate.
The results for the inuence of elasticity of the substrate show that the resistance
curves are much like those found for a rigid substrate. Thus, for material parameters
where the interface toughness is only slightly higher than the elastic toughness the
peak toughness is reached after a very small amount of crack growth, but when the
normalised toughness |K|
R
}K
0
attains values higher than 3.0 or 4.0 the peak toughness is
reached after a great deal of crack growth, corresponding to several times the reference
size R
0
for the plastic region. However, as shown by Fig. 4, and also by the solid
curves in Fig. 9, the interface toughness is strongly dependent on the elastic modulus
ratio E
2
}E
1
, for xed values of all other material parameters. A similar dependence
has been noted in studies of interface crack growth along a thin ductile layer joining
elastic solids (Tvergaard and Hutchinson, 1994b, 1996), but those studies are limited
to remote mode I loading.
Crack growth along an interface joining an elasticplastic solid to an elastic solid
is relevant to metalceramic systems, and the elastic modulus ratio E
2
}E
1
=2 and
E
2
}E
1
=6 considered in Figs. 5b and c, respectively, are of an order of magnitude
representative of such material systems. The case for E
2
}E
1
=1 in Fig. 5a could repre-
sent two blocks of the same metal bonded together, where one of the blocks has been
hardened to have a much higher yield stress. The last case, E
2
}E
1
=1000 (Fig. 5d),
2702
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
has been included here to closely approach the results obtained previously for a rigid
substrate. A common feature of these studies of the eect of mode mixity is that a
signicant mode II contribution near the crack tip gives a signicant increase of the
toughness, as was also found in the case of a rigid substrate. In all four gures the
curve for o}o
Y1
=1.5 has a central interval around
0
=0

, in which |K|
ss
}K
0
=1.0.
These are intervals in which the behaviour is entirely elastic, but as soon as plasticity
appears at the ends of these intervals, it is again seen that the toughness increases with
increasing mode II contribution in the near tip region.
When also the substrate undergoes plastic yielding, more energy is dissipated during
crack growth as the near-tip plastic regions sweep over the material, and this results in
a higher interface toughness. This is illustrated by the resistance curves in Fig. 6 and
is seen in the curves of steady-state toughness vs. local mode mixity (Figs. 7 and 8),
which also show that the minima of the curves remain near
0
=0

, i.e. in the range


where the near-tip behaviour is completely dominated by mode I conditions. These
curves also show that the sensitivity to reductions of the yield stress ratio, o
Y2
}o
Y1
, is
much stronger for higher values of the interface strength, o}o
Y1
, which is associated
with the increasing slope of the curves in Fig. 9 for increasing value of o}o
Y1
.
The congurations with plastic yielding on both sides of the interface include as a
special case the symmetric solution for mode I crack growth along an interface be-
tween blocks of identical elasticplastic solids, as studied by Tvergaard and Hutchinson
(1992). Indeed, the curves in Fig. 9 for E
2
}E
1
=1 tend clearly towards this previous
result for the symmetric case, when considering the trend from o
Y2
}o
Y1
=2.0 through
o
Y2
}o
Y1
=1.5 and o
Y2
}o
Y1
=1.25. But the non-symmetric cases in Figs. 69 are also
relevant to a number of applications, where crack growth can occur along an interface
between blocks of dissimilar metals. Such dissimilar metal blocks can be bonded by
a number of dierent methods, such as cold welding, thermal diusive bonding, or a
thin adhesive layer. The non-symmetric analyses can also be relevant to some cases
of crack growth along a weld, where signicant overmatching of the weld material
gives highly dissimilar yield stress but no elastic dierences on either side of the crack
(Niordson, 1999).
It is nally noted that crack growth predictions by the present model are limited by
the fact that according to continuum plasticity crack solutions the maximum stresses
near the tip do not exceed about ve times the ow stress. Therefore, dierent models
(Suo et al., 1993) must be used when fracture occurs by atomic separation. Such
procedures using a thin elastic strip of material around the crack tip have been employed
by Tvergaard (1999) to determine the resistance curve behaviour under mixed mode
loading, at an interface between an elasticplastic solid and a rigid solid.
References
Beltz, G.E., Rice, J.R., Shih, C.F., Xia, L., 1996. A self-consistent model for cleavage in the presence of
plastic ow. Acta Metall. Mater. 44, 39433954.
Cao, H.C., Evans, A.G., 1989. An experimental study of the fracture resistance of bimaterial interfaces.
Mech. Mater. 7, 295304.
England, A.H., 1965. A crack between dissimilar media. J. Appl. Mech. 32, 400402.
V. Tvergaard / J. Mech. Phys. Solids 49 (2001) 26892703
2703
Eshelby, J.D., 1970. In: Kanninen, M.F. et al. (Eds.), Inelastic Behavior of Solids. McGraw-Hill, New York,
pp. 77115.
Evans, A.G., Hutchinson, J.W., 1989. Eect of non-planarity on the mixed mode fracture resistance of
bimaterial interfaces. Acta Metall. 37, 909916.
Hutchinson, J.W., 1973. Finite strain analysis of elasticplastic solids and structures. In: Hartung, R.F. (Ed.),
Numerical Solution of Nonlinear Structural Problems. ASME, New York, pp. 17.
Jensen, H.M., 1990. Mixed mode interface fracture criteria. Acta Metall. Mater. 38, 26372644.
Liechti, K.M., Chai, Y.S., 1992. Asymmetric shielding in interfacial fracture under in-plane shear. J. Appl.
Mech. 59, 295.
Needleman, A., 1987. A continuum model for void nucleation by inclusion debonding. J. Appl. Mech. 54,
525531.
Niordson, C.F., 1999. Analysis of steady-state ductile crack growth along a laser weld. DCAMM Rep. 626.
ODowd, N.P., Stout, M.G., Shih, C.F., 1992. Fracture toughness of alumina}niobium interfaces: experiments
and analyses. Philos. Mag. A66, 1037.
Rice, J.R., 1968. A path independent integral and the approximate analysis of strain concentration by notches
and cracks. J. Appl. Mech. 35, 379386.
Rice, J.R., 1988. Elastic fracture mechanics concepts for interfacial cracks. J. Appl. Mech. 55, 98103.
Suo, Z., Shih, C.F., Varias, A.G., 1993. A theory for cleavage cracking in the presence of plastic ow. Acta
Met. Mater. 41, 15511557.
Tvergaard, V., 1976. Eect of thickness inhomogeneities in internally pressurized elasticplastic spherical
shells. J. Mech. Phys. Solids 24, 291304.
Tvergaard, V., 1990. Eect of bre debonding in a whisker-reinforced metal. Mater. Sci. Eng. A125,
203213.
Tvergaard, V., 1999. Eect of plasticity on cleavage crack growth resistance at an interface. J. Mech. Phys.
Solids 47, 10951112.
Tvergaard, V., Hutchinson, J.W., 1992. The relation between crack growth resistance and fracture process
parameters in elasticplastic solids. J. Mech. Phys. Solids 40, 13771397.
Tvergaard, V., Hutchinson, J.W., 1993. The inuence of plasticity on mixed mode interface toughness. J.
Mech. Phys. Solids 41, 11191135.
Tvergaard, V., Hutchinson, J.W., 1994a. Eect of 1-stress on mode I crack growth resistance in a ductile
solid. Int. J. Solids Struct. 31, 823833.
Tvergaard, V., Hutchinson, J.W., 1994b. Toughness of an interface along a thin ductile layer joining elastic
solids. Philos. Mag. A 70, 641656.
Tvergaard, V., Hutchinson, J.W., 1996. On the toughness of ductile adhesive joints. J. Mech. Phys. Solids
44, 789800.
Wei, Y., Hutchinson, J.W., 1996. Mixed mode interface toughness of metal}ceramic joints. Material Research
Society Symposium, Vol. 409, Materials Research Society, Pittsburgh, pp. 163170
Wei, Y., Hutchinson, J.W., 1997. Nonlinear delamination mechanics for thin lms. J. Mech. Phys. Solids
45, 11371159.

Вам также может понравиться