You are on page 1of 25

The growth and coalescence of ellipsoidal voids in plane strain under

combined shear and tension


F. Scheyvaerts
a,b
, P.R. Onck
b
, C. Tekoglu
a
, T. Pardoen
a,n
a
Institute of Mechanics, Materials and Civil engineering, Universite catholique de Louvain, Place Sainte Barbe 2, B-1348 Louvain-la-Neuve, Belgium
b
Zernike Institute for Advanced Materials, University of Groningen, Micromechanics of Materials, Nijenborgh 4, 9747 AG Groningen, The Netherlands
a r t i c l e i n f o
Article history:
Received 15 July 2009
Received in revised form
30 September 2010
Accepted 9 October 2010
Available online 19 November 2010
Keywords:
Ductility
Fracture mechanisms
Voids and inclusions
Constitutive behaviour
Shear
a b s t r a c t
New extensions of a model for the growth and coalescence of ellipsoidal voids based on
the Gurson formalism are proposed in order to treat problems involving shear and/or
voids axis not necessarily aligned with the main loading direction, under plane strain
loading conditions. These extensions are motivated and validated using 3Dnite element
void cell calculations with overall plane strain enforced in one direction. The starting
point is the Gologanumodel dealing withspheroidal voidshape. Avoidrotationlawbased
on homogenization theory is coupled to this damage model. The predictions of the model
closely agree with the 3D cell calculations, capturing the effect of the initial void shape
andorientationonthe voidrotationrate. Anempirical correctionis also introducedfor the
change of the void aspect ratio in the plane transverse to the main axis of the void
departing from its initially circular shape. This correction is needed for an accurate
prediction of the onset of coalescence. Next, a new approach is proposed to take strain
hardening into account within the Thomason criterion for internal necking, avoiding the
use of strain hardening-dependent tting parameters. The coalescence criterion is
generalized to any possible direction of the coalescence plane and void orientation.
Finally, the model is supplemented by a mathematical description of the nal drop of the
stress carrying capacity during coalescence. The entire model is developedfor plane strain
conditions, setting the path to a 3D extension. After validation of the model, a parametric
study addresses the effect of shear on the ductility of metallic alloys for a range of
microstructural and ow parameters, under different stress states. In general, the
presence of shear, for identical stress triaxiality, decreases the ductility, partly explaining
recent experimental results obtained in the low stress triaxiality regime.
& 2010 Elsevier Ltd. All rights reserved.
1. Introduction
The rst generation of micromechanics-based models for ductile fracture was developed in the late 1960s, early 1970s,
based on closed-form solutions for the growth of isolated voids within a plastically deforming material (e.g. McClintock,
1968; Rice and Tracey, 1969). Simple critical strain or stress based void nucleation criteria, and a critical porosity condition at
cracking initiation were added to these models in order to predict the fracture strain. These solutions are still very useful
today for enlightening the key trends in the relationships between the ductility of metallic alloys, ow properties,
microstructure, and loading conditions (e.g. see Pineau and Pardoen, 2007).
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/jmps
Journal of the Mechanics and Physics of Solids
0022-5096/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jmps.2010.10.003
n
Corresponding author. Tel.: +32 10 472417; fax: +32 10 474028.
E-mail address: thomas.pardoen@uclouvain.be (T. Pardoen).
Journal of the Mechanics and Physics of Solids 59 (2011) 373397
The second generation of micromechanical models was initiatedby the work of Gurson(1977) andconsistedof a yieldlocus
for plastic dilatational materials as well as evolution laws for the internal variables, allowing the formulation of a complete
constitutive model for porous materials, withporosity levels typically lower thana fewpercent. Several extensions, drivenby
the need(i) to deliver quantitative predictions and(ii) to simulate the complete fracture process, were proposedby Tvergaard
and Needleman (1984) (see Tvergaard, 1990 for a detailed review), followed by many others (e.g. Becker et al., 1987, 1989b;
Koplik and Needleman, 1988; Homand McMeeking, 1989; Steglich and Brocks, 1997; Faleskog et al., 1998, Gao et al., 1998).
Simple empirical void nucleation and coalescence laws were used and the voids were considered to remain spherical. Up to
the early 1990s, the interest for these models was mainly driven by the fracture mechanics community to simulate crack
growth in ductile materials and solve the transferability problems when small scale yielding cannot be insured either in
laboratory specimens or in structural components (e.g. see applications by DEscatha and Devaux, 1979; Needleman and
Tvergaard, 1987; Devaux et al., 1989; Xia and Shih, 1995; Xia et al., 1995; Brocks et al., 1995; Ruggieri et al., 1996;
Koppenhoefer and Dodds, 1998).
Nevertheless, this second generation of models suffered from several limitations: (i) poor predictions under low stress
triaxiality conditions typical of metal forming operations andthinplate cracking (see e.g. Pardoenet al., 1999); (ii) problemto
encompass bothlowandlarge stress triaxiality regimes without tuning the material parameters; and(iii) difcultyto address
fracture in alloys with complex microstructures. These limitations have motivated new developments over the last 1015
years leading to a third generation of models (see recent review by Benzerga and Leblond, 2010):
+ The versions of the Gurson model based on the spherical void assumption do not behave well under low or intermediate
stress triaxiality. Low triaxiality implies void shape changes, see seminal analysis by Budiansky et al. (1982). Void shape
has been introduced explicitly by Gologanu et al. (1993, 1994, 1997) for a perfectly plastic response. This model has then
been extended to strain hardening and strain rate sensitive materials, and validated (e.g. by Pardoen and Hutchinson,
2000, 2003; Kl ocker and Tvergaard, 2003; Benzerga et al., 2004; Flandi and Leblond, 2005; Pardoen, 2006; Lassance et al.,
2007).
+ Materials with heterogeneous microstructures involving different phases at different scales lead to the competition or co-
operation between different damage mechanisms that have been included in or treated with the Gurson formalism (or
similar), suchas the competitionbetweenintergranular versus transgranular ductile fracture (see e.g. Becker et al., 1989a;
Pardoen et al., 2003; Morgeneyer et al., 2008), the presence of soft and hard phases (Pardoen and Brechet, 2004; Steglich
et al., 2008; Chehab et al., 2010), void distribution effects (e.g. Besson et al., 2000), the presence of a second population of
voids (Perrin and Leblond, 2000; Enakoutsa et al., 2005), multiple primary inclusion populations (Lassance et al., 2007),
and void locking by inclusions (Siruguet and Leblond, 2004).
+ The condition for void coalescence depends on the local void conguration and loading conditions. A simple critical
porosity criterion is not adequate in many instances. Micromechanics based coalescence models, following the ideas of
Thomason (1985a, b, 1990) and Brown and Embury (1973) have been extended and integrated in the Gurson formalism
(see Zhang and Niemi, 1995; Benzerga, 2002; Pardoen and Hutchinson, 2000; Zhang et al., 2000; Benzerga et al., 2002,
2004; Pardoen and Brechet, 2004; Pardoen et al., 1998, 2004; Huber et al., 2005; Fabr egue and Pardoen, 2008).
+ The Gurson model relies on isotropic hardening while many metals exhibit plastic anisotropy and kinematic hardening.
Several authors have improved the description of the matrix material to account for these effects, e.g. (for plastic
anisotropy) Liao et al. (1997), Grange et al. (2000), Benzerga and Besson (2001), Benzerga et al. (2002, 2004), Wang et al.
(2004), Keralavarma and Benzerga (2008, 2010), Monchiet et al. (2008), (for kinematic hardening) Mear and Hutchinson
(1985), Besson and Guillemer-Neel (2003), M uhlich and Brocks (2003), see also the analysis by Leblond et al. (1995) for a
more adequate incorporation of isotropic strain hardening.
One key issue which has not received much attention is the effect of the void orientation and of the change of this
orientation during deformation, especially under shear dominated loading. Void rotation takes place in several problems
such as in cutting and shear dominated forming operations, mixed mode ductile cracking, and, at the microstructural scale,
when voids growon inclined soft grain boundaries or withinshear bands. Analysis of the growth of spherical voids in linearly
viscous materials under shear has been investigated by Fleck and Hutchinson (1986), showing that the voids were elongating
and rotating towards a penny shape. Later, Fleck et al. (1989) have addressed the problemof the contact of the void with the
inner inclusion which, under shear, prevents the void from closing. Gologanu et al. (1997) as well as Ponte Castaneda and
Zaidman(1994) introducedvoidrotations laws intheir constitutive model. Gologanuet al. (1997) proposedtoaccount for the
void rotation by imposing the void to follow the rotation of the surrounding material. FE cell calculations for voids growing
under shear have beenperformedby Bordreuil et al. (2003), LeblondandMottet (2008), andTvergaard (2008,2009). Recently,
Nahshon and Hutchinson (2008) have heuristically extended the Gurson model in order to account for shear effects by
making use of the third invariant of stress (see implementation by Nielsen and Tvergaard, 2009, as well as the model
formulatedbyXue, 2007). One of the motivations for these last developments was the experimental evidence for a decrease of
the fracture strain at low stress triaxiality in the presence of shear, in contrast with the expected continuous increase of
ductility with decreasing stress triaxiality in the absence of shear (Barsoum and Faleskog, 2007). Recently, Leblond and
Mottet (2008) have proposed a coalescence model combining shear and tensile localization.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 374
The goal of this work is todevelopandvalidate a set of extensions of a constitutive model for the growthandcoalescence of
ellipsoidal voids within a J2 plastically hardening matrix, accounting for void orientation and rotation, coupled to the void
growth and shape evolution. The foundations of the model are based on the work of Gologanu et al. (1993, 1994, 1997). More
precisely, the new extensions are the following: (i) introduction of a new state variable to account for void rotation,
(ii) development of an empirical correction for the loss of spheroidicity, (iii) incorporation of the strain hardening in the
internal necking coalescence criterion, (iv) generalization of the internal necking criterion to general loading conditions,
(v) coupling with a geometrical model for the internal necking process and the associated drop of load carrying capacity,
presented in detail in Scheyvaerts et al. (2010). The model is for 3D spheroidal voids, but is formulated only for overall plane
strain conditions, one of the purposes being the implementation in a 2D plane strain FE code, see Scheyvaerts (2009) and
Pardoen et al. (2010).
The paper is organized as follows. Section 2 describes a new set of 3D nite element cell calculations under overall plane
strain conditions with the purpose (i) to showthe maineffects of shear on the damage evolution, and (ii) to provide a basis for
validationof the constitutive model. The constitutive model is elaboratedinSection3. Section4provides anassessment of the
model against the 3D void cell calculations. Finally, Section 5 discusses various aspects of the model and proposes general
views about the ductility of metallic alloys under complex loading conditions.
2. FE void cell calculations under combined shear and tension
2.1. Description of the void cell simulations
The unit cell is a cube made of an elastic-plastic material containing a single spheroidal void. The origin of the reference
system {e
i
}
i =1,2,3
is located at the center of the void, see O in Fig. 1a. The initial dimensions of the cell along these axes are
2L
10
, 2L
20
, 2L
30
withL
10
=L
30
. Plane strainis enforcedon average inthe e
3
direction, hence L
3
=L
30
. We will refer in this paper to
overall plane strain conditions. The initial cell aspect ratio l
0
=L
20
/L
10
characterizes the void distribution in the {e
1,
e
2
}
plane. An additional orthonormal axis system, {n
i
}
i =1,2,3
, is associated with the void with n
2
corresponding to the symmetry
axis of the spheroid and n
3
=e
3
. We will explain in Section 2.2 the procedure followed to determine the reference system
attached to the void. Initially, the void axes and reference axes coincide (see Fig. 1b). The spheroidal void is characterized by
the initial dimensions of the symmetry axis 2R
20
and of the transverse axes 2R
10
=2R
30
. The initial aspect ratio of the void is
dened as W
0
=R
20
/R
10
. The initial volume fraction is given by f
0
=k
g
(R
2
10
R
20
)=(L
2
10
L
20
), where k
g
depends on the arrangement
of voids: for a simple cubic array, k
g
=p=6.
The response of the matrix obeys isotropic Hookean elasticity with Youngs modulus E and Poisson ratio n. The J
2
ow
theory describes the plastic behaviour using the hardening law
s
M
s
0
= 1
Ee
p
M
s
0
_ _n
, (1)
where s
0
is the initial yield stress, s
M
is the current yield stress, e
M
p
is the accumulated plastic strain, and n is the strain-
hardening exponent. A constant ratio E/s
0
=300, typical of a wide range of metallic alloys, is used in this study.
The cell is deformed by a combination of normal displacements U
1
and U
2
in the e
1
and e
2
directions, respectively, as well
as by U
t
applied uniformly on the upper surface parallel to e
1
. The components of the macroscopic velocity gradient L
ij
in the
Fig. 1. (a) Finite element mesh used for the 3D void cell calculations, here for an initially prolate void with a void aspect ratio W
0
=6, initial volume fraction
f
0
=510
3
, and void distribution factor l
0
=1; (b) description of the loading and boundary conditions, as well as of the void reference system and rotation
angle; plane strain is applied in the out of plane direction.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 375
reference system {e
i
}
i =1,2,3
read
L
ij
=
_
U
1
=2L
1
_
U
t
=L
2
0
0
_
U
2
=2L
2
0
0 0 0
_
_
_
_
_
_, (2)
so that the macroscopic strain rate D
ij
and spin O
ij
components follow to be
D
ij
=
_
U
1
=2L
1
_
U
t
=2L
2
0
_
U
t
=2L
2
_
U
2
=2L
2
0
0 0 0
_
_
_
_
_
_, O
ij
=
0
_
U
t
=2L
2
0

_
U
t
=2L
2
0 0
0 0 0
_
_
_
_
_
_ (3)
Due to the symmetries of the void distribution and loading conguration, the problem can be limited to simulating a
quarter of the representative cubic volume element. Periodic boundary conditions are applied by enforcing appropriate
constraint among opposing faces. The detailed description of the boundary conditions is given in Appendix A.
The loading state is characterized by a constant ratio
_
U
t
=
_
U
2
. Except for simple shear calculations, a constant stress
triaxiality T =s
h
=s
e
(where s
h
is the hydrostatic stress and s
e
is the von Mises stress) is imposed. An iterative Newton-
Raphson procedure is used to determine the DU
1
to be applied in addition to the prescribed DU
2
and DU
t
values in order to
enforce the target triaxiality T (see Pardoen and Hutchinson, 2000).
Unless stated otherwise, the components of the overall strain tensor e
ij
are dened with respect to reference axis system
{e
i
}
i =1,2,3
and are calculated by integrating the strain rates D
ij
over time. The overall stress components are calculated by
volume averaging throughall the elements of the cell. The current void volume fraction f is calculatedby subtracting fromthe
total volume of the deformed unit cell (known fromthe applied displacements and initial dimensions) the sumof the volume
of all the elements. The extraction of the other quantities of interest (void orientation, shape, etc.), is described when
presenting the results of the calculations. The simulations have beenperformedusing the nite element software ABAQUS 6.5
(2004), with elements C3D8. The convergence of the results with respect to mesh renement until the coalescence stage has
been carefully veried. A typical converged mesh for an initial porosity f
0
equal to 10
3
contains 4112 elements.
2.2. Results of the void cell calculations
Attention is rst devoted to the evolution of the orientation of the cavity under simple shear. Next, the effect of an applied
shear component onthe voidgrowthrate, onthe material ductility (denedas the strainat the onset of voidcoalescence), and
on the unloading rate during coalescence is addressed for various loading conditions and material parameters.
2.2.1. Simple shear (under plane strain)
The rotation of a cavity is analyzed rst for simple shear conditions. The evolution of the other parameters will be
addressed later for more general loading cases. The void is assumed to rotate solely in the {e
1,
e
2
} plane. The angle y (see
Fig. 1(b)), denes the relative orientation with respect to the reference axis system, i.e.
n
1
=cosye
1
sinye
2
n
2
=sinye
1
cosye
2
n
3
=n
1
Qn
2
=e
3
_

_
(4)
The initial orientationy
0
is equal to either 0 or p/2. The deformationgradient is of the formF=I+ge
1
Qe
2
where g quanties the
amount of shear (U
t
=g L
20
, U
1
=U
2
=0). The elasto-plastic matrix is characterized by E/s
0
=300, n=0.3, n=0.1, and initial porosity
f
0
=10
2
. Three initial void shapes are studied (W
0
=1, 1/3, 3), with two different initial orientations for the oblate (W
0
=1/3)
and prolate (W
0
=3) cases: (i) perpendicular to the shear direction (n
2
=e
2
, or equivalently y
0
=0) and (ii) parallel to it (n
2
= e
1
, or
y
0
=p/2). Fig. 2 shows the undeformed (in blue) and the deformed (in red) congurations corresponding to these ve simulations.
In case (a) of an initially prolate void, the symmetry axis n
2
rotates in order to align with the principal strain direction irrespective
of theinitial orientation. Inthecase(b) of aninitiallyoblatevoid, theoppositeoccurs, i.e. theaxis n
2
rotates awayfromtheprincipal
strain direction. Hence, the symmetry axes of prolate and oblate cavities rotate in opposite directions under the same loading
conditions. Intheparticular caseof aninitiallyspherical void, thelocal system{n
i
}
i =1,2,3
andy areinitiallyundened. As soonas the
material deforms, the cavity elongates in the direction of maximum stretching, evolving towards a prolate shape (see Fig. 2c).
The following procedure is proposed to dene and determine the angle y from the FE cell calculations when the void
deforms and rotates, departing from the perfect ellipsoidal shape. The direction n
2
is chosen as to provide the largest (resp.
smallest) radius R
2
for an initially prolate (resp. oblate) void shape (see Fig. 3). The direction n
1
is perpendicular to n
2
in the
{e
1,
e
2
} plane, giving the radius R
1
. The radius R
3
is determined by the position of the node initially located at (0,0, R
30
). Fig. 4
shows the evolution of y as a function of the shear strain, demonstrating the change in the rotation behaviour for different
initial void aspect ratio W
0
and angle y
0
.
The stress triaxiality corresponding to these simple shear calculations is very small, close to 0. More precisely, for an initial
porosity f
0
equal to10
3
, the stress triaxialityT is equal to 0andslowlyincreases due tothe presence of the growing voidupto
T=0.1 for an effective strain equal to 1. Coalescence is never attained for the range of strains that can be applied before
unacceptable mesh distortions appear; the overall effective strain must typically not be larger than 1.5 (see Tvergaard 2008,
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 376
2009, for recent simulations of void growth under simple shear, involving remeshing and frictional effects which makes
possible to reach coalescence).
2.2.2. Combined sheartension loading
The following void cell calculations were performed at a constant T, including a shear contribution characterized by a
constant ratio
_
U
t
=
_
U
2
.
n
2
n
2
n
2
n
2

0
= 0

0
= 0

0
= /2

0
= /2
Fig. 2. Sectioninthe plane x
3
=0 of a unit-cell deformedbya shear straing=0.5, under simple shear loading, withf
0
=10
2
, l
0
=1, s
0
/E=0.003, andn=0.1. The
inuence of the initial orientation of the void, i.e. y
0
=0or p/2, on void rotation is observed for an initially (a) prolate void with W
0
=3, (b) oblate void with
W
0
=1/3, and (c) spherical void with W
0
=1.
R
1
R
2

R
1
R
2

R
1
R
2

Fig. 3. Evolution of the void shape in the plane normal to the plane strain direction, with the direction n
1
and n
2
dened within the last deformed
conguration for T=1, U
t
/U
2
=1, f
0
=10
3
, k
0
=1, and for three different initial void aspect ratios (a) W
0
=1/6, (b) W
0
=1, and (c) W
0
=6.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 377
2.2.2.1. Loss of spheroidicity. The void is initially spheroidal. With deformation, the void departs fromits circular shape in the
{n
1
,n
3
} plane and becomes elliptical, as schematically illustrated in Fig. 5 for a low stress triaxiality condition. The out-of-
plane radius (R
3
) is larger than the in-plane radius (R
1
). Note that in Fig. 5, R
1
is taken with respect to the reference axis e
1
insteadof n
1
for illustrationpurposes. Fig. 6shows the evolutionof the ratioR
1
/R
3
as a functionof the axial straine
22
upto void
coalescence, for various material parameters and loading conditions, without shear. Fig. 6a shows that the evolution of R
1
/R
3
is almost independent of the initial void shape (W
0
=1/6, 1 or 6) except that void coalescence starts later when the voids are
prolate. The minimumof R
1
/R
3
corresponds to the peak inthe stress straincurve. The growthrate
_
R
1
thenbecomes equal to
_
R
3
so that the R
1
/R
3
ratio starts increasing again. Changing the stress triaxiality T, as shown in Fig. 6b, leads to very different
evolutions of the shape factor R
1
/R
3
, from a relatively constant value at high T, to signicant departure from the circular
section at low T. Adding a shear component or changing the initial porosity (results not shown) does not affect the con-
clusions, only stress triaxiality matters.
2.2.2.2. Effect of shear on ductility. The following results aim at comparing calculations performed without shear and with a
shear component U
t
/U
2
=1. The results presented in Fig. 7 are for a material with n=0.1, f
0
=10
3
, l
0
=1, and W
0
=1/6, 1, or 6,
with T=1. Fig. 7a displays variations of the stress s
22
as a function of strain e
22
. The onset of void coalescence, detected by a
sudden change in the evolution of _ e
22
as a functionof _ e
11
(see also Koplik and Needleman, 1988), is indicated by a small circle.
Adding a shear component leads to a decrease of the peak stress and advances the initiation of the coalescence process. The
inuence of the shear component is amplied for large void aspect ratios W
0
.
After the onset of void coalescence, the loss of stress carrying capacity is almost linear (see Fig. 7a), until the voids become
so large that the mesh becomes highly distorted and the quality of the solution signicantly deteriorates. In an analysis using
-1.5
-1
-0.5
0
0.5
1
1.5
2
0


(
r
a
d
)

W
0
= 1
W
0
= 1/3
W
0
= 1/3
W
0
= 3

0
= 0

0
= /2
W
0
= 3
0.1 0.2 0.3 0.4 0.5 0.6
Fig. 4. Variation of the void orientation angle as a function of the applied shear strain for a simple shear loading, with f
0
=10
2
, l
0
=1, W
0
=3, 1/3 or 1, and
y
0
=0 or p=2.
Fig. 5. Schematic description of the evolution of the void section in the plane transverse to the main void axis under low stress triaxiality conditions and
plane strain enforced in the direction e
3
. Note that the void is initially spherical, i.e. the initial shape in this plane is circular.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 378
remeshing techniques for extreme void expansion at an interface between a metal and a ceramic, Tvergaard (1997) has
shown that the overall stress drop remains linear until complete loss of stress carrying capacity. No signicant effect of shear
on the unloading slope is detected in Fig. 7a.
The originof the effect of the shear distortiononthe overall behaviour canbe elucidatedbyanalyzing the evolutioninFig. 7b
and c of the porosity f and void shape factor S=ln (W), where Wis dened in an average sense as W=2R
2
/(R
1
+R
3
). The inuence
of the void shape on void growth has already been extensively studied by, e.g. Becker et al. (1989b), Pardoen and Hutchinson
(2000), Lassance et al. (2006): oblate voids grow faster than prolate voids. Fig. 7b and c show that the presence of a shear
distortion induces a faster void growth explaining the smaller coalescence strain e
C
22
in the presence of shear. Fig. 7d plots the
variation of the void rotation angle y as a function of the matrix rotation angle g, for U
t
/U
2
=1. Fig. 7d shows also that an oblate
void rotates much faster than the surrounding material. Prolate voids follow the material rotation, while spherical voids
behaviour differs by the fact that their orientation is only dened once the elongation starts along the principal strain direction.
0.5
0 0 0.5 1 1.5 2.5 2 0.2 0.4 0.6 0.8 1 1.2
0.6
0.7
0.8
0.9
1
Void cells
Model
R
1
/
R
3

22

22
T=1, f
0
=10
-3
,
0
=1
W
0
=1/6
W
0
=1
W
0
=6
0
0.2
0.4
0.6
0.8
1
Void cells
Model
f
0
=10
-3
, W
0
=1,
0
=1
T=0.577
T=1
T=1.5
T=2
R
1
/
R
3
Fig. 6. Variation of the ratio between the in-plane and out-of-plane radii of the void, R
1
=R
3
=1=W
t
, as a function of the axial strain, up to the onset of void
coalescence. Comparison between FE void cell results and the predictions of the model. Inuence of the (a) initial void aspect ratio W
0
and (b) applied stress
triaxiality T.
0
0.5
1
1.5
2
2.5
3

2
2
/

0
0
0.05
0.1
0.15
f
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
S
=
l
n
(
W
)

22
-1
-0.5
0
0.5
1
1.5


(
r
a
d
)

0 0 0.2 0.4 0.6 1 1.2 0.8 0.1 0.2 0.3 0.4 0.5 0.6
Fig. 7. Voidcell results obtainedwithU
t
=U
2
or without shear (U
t
=0), for three voidaspect ratios W
0
=1/6, 1, or 6, T=1, f
0
=10
3
, l
0
=1, s
0
/E=0.003, and n=0.1:
(a) stressstrain response; (b) porosity vs. axial strain; (c) void aspect ratio vs. axial strain; and (d) void rotation vs. matrix material rotation.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 379
Fig. 8a summarizes the effect of shear on the material ductility dened as the equivalent strain at the onset of coalescence
e
c
eq
for different initial void aspect ratios. The ductility increases with increasing W
0
(see Pardoen and Hutchinson, 2000).
Adding a shear contribution decreases the ductility. Similarly, Fig. 8b shows the effect of shear on the ductility of
materials with three different volume fractions: f
0
=10
2
, 10
3
, and 10
4
, for initially spherical voids. Again, the effect
of shear is obvious with a slightly more pronounced weakening effect when f
0
is small. Note also that the strains at
coalescence reported here for 3Dcalculations under overall plane strain (e.g. e
C
22
=0.88 for T=1, no shear and W
0
=1) are larger
than the one obtained by Pardoen and Hutchinson (2000) for axisymmetric conditions (e.g. e
C
22
=0.77 for T=1, no shear
and W
0
=1).
Fig. 9 compares the variation of the ductility as a function of the stress triaxiality for different levels of shear distortion
givenby U
t
/U
2
=0, 0.5, 1, 2, 5for stress triaxiality varying betweenthe smallest triaxiality at whichcoalescence is detectedand
T=3, in a material with f
0
=10
2
, W
0
=1 and l
0
=1. The loss of ductility caused by an additional shear contribution increases
withdecreasing stress triaxiality. For instance, doubling the shear contributionleads to a decrease of the ductility by less than
3%at T=2, whereas it causes upto a 30%drop at T=0.577. This last result agrees withrecent experiments reportedby Barsoum
and Faleskog (2007) showing that the increase of ductility with decreasing stress triaxiality stops at low stress triaxiality
when superposing a shear component. A qualitative trend line of the experimental variation of the ductility is provided in
Fig. 9. In the experiments, the amount of shear increases when decreasing the stress triaxiality as indicated by the arrow in
Fig. 9. Even though we cannot directly make an one to one comparison and relate the amount of shear prescribed by the cell
calculations to the shear in their experiments, the present analysis shows that a signicant loss of ductility at low stress
triaxiality can be attributed to the rotation of the voids which decreases their spacing and brings them into a conguration
favorable for coalescence. This observationhas alsobeenmade recently by Tvergaard(2008). Now, mechanisms relatedto the
initiationof shear bands can play an additional role in this problem, but this will depend also on the geometry of the structure
or specimen and on the loading conguration.
Fig. 8. Variation of the equivalent and tensile strain components at the onset of coalescence e
c
eq
and e
c
22
, respectively, as a function of the shear component
with T=1, n=0.1, and l
0
=1, for (a) different initial void aspect ratio values and f
0
=10
3
and (b) for different initial porosity values and W
0
=1.
Fig. 9. Variation of the equivalent strain at the onset of coalescence as a function of the stress triaxiality for different shear contributions, with n=0.1,
f
0
=10
2
, W
0
=1, and l
0
=1. The dashed line showing the trend is added in order to build a qualitative link with the experimental results of Barsoum and
Faleskog (2007).
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 380
3. Constitutive model and implementation
3.1. General assumptions
The model presented in this section describes the response of an elasticplastic solid in which small internal voids
nucleate, grow and coalesce leading to the complete loss of integrity of the material, under overall plane strain conditions.
Gologanu et al. (1993, 1994, 1997) have extended the Gurson model (1977) to include the evolution of the void aspect ratio,
assuming a spheroidal void shape. This model, with the extension to strain hardening materials and supplemented with a
criterion for the onset of coalescence, has been analyzed and validated for a cylindrical RVE and axisymmetric loading
conditions by PardoenandHutchinson(2000), see also Benzerga et al. (2004), Huber et al. (2005), Gallais et al. (2007), Nielsen
et al. (2010), Simar et al. (2010) for validationagainst experimental data. These studies didnot analyze (i) voidorientationand
rotation effects which can be important in several applications (see introduction), (ii) the loss of spheroidicity demonstrated
by the cell calculations in Section 2.2 (which does obviously not appear under axisymetric conditions), and (iii) the
dependence of the orientation of the critical inter-void ligament with the loading conditions, and void geometry. Note that
Gologanuet al. (1997) proposed to account for the void rotationby imposing the void to rotate together with the surrounding
material (
_
n
2
=Xn
2
, with X the macroscopic continuum spin) which is, as shown in Section 2, not correct under certain
loading and void orientation conditions. New developments have been made to address these three limitations. Also the
introduction of strain hardening in the condition for the onset of coalescence is improved in order to avoid the use of tting
parameters. Furthermore, a model (described in Scheyvaerts, 2009) for the internal necking evolution after the onset of
coalescence is added and adapted to the rest of the constitutive model.
The model is extended for plane strain conditions only with a motivation towards an implementation in a FE code for
treating 2D plane strain cracking problems (Scheyvaerts, 2009). A global orthonormal system of reference {e
1
, e
2
, e
3
} is
dened. Alocal orthonormal axis system{n
1
, n
2
,n
3
} is associated with the cavity, where n
2
corresponds to the symmetry axis
of the representative spheroid and n
3
is the axis in the out-of-plane direction (see Fig. 10). The void radii are R
1
=R
3
=R
r
and R
2
,
with aspect ratio W=R
2
/R
r
. A correction for different evolutions of R
1
and R
3
will be developed in order to accurately predict
the onset of coalescence. This correction does not enter the void growth model. The void is assumed to rotate in the {e
1
, e
2
}
plane and only one variable (y) describes its orientation.
The behaviour of the material surrounding the void is isotropic linear elastic as long as s
e
os
M
, with s
M
the current yield
stress, and J
2
elastoplastic during plastic deformation. The hardening law is given by Eq. (1). The model involves 10 state
variables: the six components of the stress tensor r, the average yield stress of the matrix material s
M
, the porosity f, the void
aspect ratio S dened by S=ln (W), the angle y, dened in Eq. (4). The relative void spacing and the in plane void aspect ratio
R
1
/R
3
will enter as two additional variables, when dealing with the coalescence of the voids only.
The different parts of the model are described nowwith special focus on the newdevelopments, i.e. void growthin Section
3.2, the onset of coalescence inSection3.3, andunloading duringcoalescence inSection3.4. Note that noeffort was made here
to address the important issue of void nucleation (see Teko glu and Pardoen, 2010). The voids are considered to be present in
the material before loading. There is no difculty upgrading the present framework with an existing void nucleation law.
3.2. Void growth
3.2.1. General description of the model
The void growth model consists of a owpotential supplemented by evolution laws for the internal variables and the ow
rule for the plastic strain increment, details are given elsewhere (see Pardoen and Hutchinson, 2000, or Pardoen, 2006). The
ow potential F is dened as
F=
C
s
2
M
:sZs
hg
X:2q(g1)(gf )cosh k
s
hg
s
M
_ _
(g1)
2
q
2
(gf )
2
0 (5)
where s is the deviatoric stress tensor, s
hg
a generalizedhydrostatic stress denedby a
2
(s
11
s
33
)(1a
2
)s
22
, andXa tensor
dened by
2
3
n
2
Qn
2

1
3
n
1
Qn
1

1
3
n
3
Qn
3
. The evolution laws for the porosity, void aspect ratio, and principal void axis (or
similarly its orientation y) are, respectively
_
f =(1f )tr(D
p
), (6)
_
W
W
=
3
2
(1h
1
) D
p

1
3
tr(D
p
)I
_ _
: (n
2
Qn
2
)h
2
tr(D
p
), (7)
_
n
2
=xn
2
(8)
withthe rotationtensor xinEq. (8) tobe denedinthe next subsection. The details of the expressions relatingthe parameters
appearing in Eqs. (5) to (8) (e.g. C, Z, q, g, k, h
1
, h
2
, a
2
) to the state variables are provided in Gologanu et al. (1997) or Pardoen
andHutchinson(2000). The total strainrate consists of the sumof anelastic part andplastic part, D=D
el
+D
P
. The plastic rate of
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 381
deformation tensor D
P
is given by the ow rule
D
P
=L
@F
@r
(9)
In order to evaluate the proportionality factor L in Eq. (9), the relationship between the plastic dissipation within the
matrix material and the plastic work rate of the voided material proposed by Gurson (1977) is used together with the
hardening law (1):
s
M
_ e
p
M
(1f ) =r:D
p
(10)
The consistency condition
_
F =0 determines _ s
M
. For numerical convenience, the model is integrated using a viscoplastic
formulation with a very small rate sensitivity parameter in order to recover time independent results (see e.g. Peirce et al.,
1984; Onck and van der Giessen, 1999).
3.2.2. Void rotation law
Kaisalam and Ponte Castaneda (1998) proposed a general constitutive theory for the effective behaviour and
microstructure evolution in heterogeneous materials consisting of randomly oriented and distributed ellipsoidal inclusions
(or pores), undergoing general three-dimensional deformation. The theory is based on a variational homogenization
technique. The special case of porous metals was considered by Kaisalam and Ponte Castaneda (1998) and Kailasam et al.
(1997, 2000). The homogenized continuumis assumed to be locally orthotropic, where the local axes of orthotropy coincides
Fig. 10. (a) Initial conguration of the void distribution, geometry and orientation; (b) current conguration; (c) current void geometry and orientation; and
(d) denition of the angle and dimensions entering the generalized void coalescence criterion.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 382
with the principal axes of the representative ellipsoid. In our model, however, the void distribution characterized by the
position of the void centers evolves with the material, i.e. it is governed by the deformation rate D and spin X; whereas only
the void orientation follows the local axes of orthotropy during plastic ow, i.e. it is governed by the rotation tensor x, as in
Kaisalamand Ponte Castaneda (1998). For details on the derivation of xin terms of the macroscopic plastic deformation rate
D
P
and the macroscopic continuum spin X, the reader is referred to Kaisalam and Ponte Castaneda (1998) and references
therein. Here, we strictly follow the derivation by Aravas and Ponte Castaneda (2004) and only summarize the nal
expressions for completeness. The rotation tensor x can be written as
x=XX
P
(11)
with X
P
the plastic spin, i.e. the spin of the material axes relative to the void axes. In plane strain, in which deformation
takes place in the {e
1
, e
2
} plane, X
p
is of the form
X
P
=o
P
(n
1
Qn
2
n
2
Qn
1
) (12)
with
o
P
=e
1
(C:D
P
)e
2

1
2
1W
2
1W
2
(n
1
Qn
2
n
2
Qn
1
) : A:D
P
when Wa1 (13)
and
o
P
=0 when W=1 (14)
Here, A and C are concentration tensors, relating the deformation rate and spin in the void to the macroscopic
deformation rate and spin (see Aravas and Ponte Castaneda, 2004). The combination of the variational homogenization
method with the micromechanics based Gurson model leads to a hybrid formalismwhich aims at integrating the best of both
frameworks.
3.2.3. Correction for the loss of spheroidicity
The correction for the loss of spheroidicity is introduced in the present section because it concerns the void growth stage
even though it will only play a role in the void coalescence model. The 3D unit cell calculations with overall plane strain
conditions have shown that an initially spheroidal cavity loses its symmetry of revolution around the axis n
2
at low stress
triaxiality, growing faster in the out of plane direction n
3
compared to the in plane direction n
1
(see Section 2.2). With the
material contraction being constrained in the direction n
3
=e
3
, it is counterbalanced by a larger void growth rate in that
direction to contribute to volume conservation. This observation is transposed mathematically into the following heuristic
expression for the kinematics, written with respect to the initial conguration as
_ u
n
1
L
10

_ u
n
3
L
30
=p
_
R
1
R
10

_
R
3
R
30
_ _
(15)
where _ u
n
1
and _ u
n
3
are the applied displacement rates in the void reference systemand p is a constant that will be identied by
comparison with the cell calculations. All quantities R
i
and L
i
are dened with respect to the void reference system. Imposing
R
10
=R
30
and the plane strain condition, gives
_
u
n
1
=p(
_
R
1

_
R
3
)=
R
10
L
10
_ _
(16)
This expression can be reformulated in terms of non-dimensional variables and of the macroscopic plastic deformation
rate D
P
n
, whichis denedwithrespect to the voidreference system. The transverse aspect ratios are denedas W
t
=R
3
=R
1
and
W
in
=R
2
=R
1
, the transverse voiddistributionindexis denedas l
t
=L
3
=L
1
andthe relative voidspacingina directioni (i =13)
is dened as w
i
=R
i
=L
i
. Evolutions laws for all these non-dimensional quantities can be written in terms of D
P
n
, and of
_
f =f and
_
W=W which are direct outcomes of the constitutive model described in Section 3.2.1. These relationships are provided in
Appendix B. The evolution law for W
t
, which is the quantity of interest here writes:
_
W
t
W
t
=C(1W
t
)K (17)
with
K = D
P
n11
D
P
n33
l
t
l
t0
_ _
w
10
pw
1
and C =
(
_
f =f D
p
nkk

_
W=WK((2W
t
)=(1W
t
))
14W
t
W
2
t
(1W
t
) (18)
Fig. 6b compares the evolution of W
t
directly extracted from the cell calculations to predictions obtained by integrating
(17) using p =
1
2
with all the quantities (f, etc.), extracted from the cell calculations. The agreement is conspicuous for all the
calculations performed in this work, p =
1
2
providing the best t. Note that the correction proposed here is far from being
perfect froma theoretical point of view. Ideally, an extension of the Gologanu model (which is limited to spheroidal voids) to
general ellipsoid void shapes is the most attractive, as recently initiated by Leblond and Gologanu (2008). A more in depth
theoretical analysis of the foundations of this correctionandof the reasons for being so accurate is left for future investigation.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 383
Once again, this correction does not enter the constitutive model but will only be used later in connection with the void
coalescence criterion.
3.3. Generalized criterion for the onset of void coalescence
3.3.1. New extension to strain hardening
Thomasons criterion (1990) is a condition for the onset of plastic localization in the ligament between neighboring voids,
leading to internal necking. The original criterion, developed for perfectly plastic materials, writes
s
n
s
0
=(1Zw
2
) a
1w
wW
_ _
2
1:2

1
w
_ _
, (19)
where a is a constant equal to 0.1, and Z is a geometric factor which depends on the void arrangement and which is dened
suchthat the term1=(1Zw
2
) represents the ratio of the total ligament area including the voiddividedby the non-porous area
(e.g., in the 3Dcubic primitive packing, Z=p/4). According to this criterion, coalescence occurs when the stress normal to the
ligament, s
n
, normalized by the yield stress of the matrix material, s
0
, reaches a value for which a localized mode of plastic
yielding conned in the ligament between the voids becomes more favorable than the more diffuse mode of plastic yielding
all around the growing cavities. This critical value is a function of the ligament geometry, described by the void aspect ratio,
W, andthe relative inter-distance betweenvoids along the ligament, w. (Note also that inthe original model R
1
=R
3
andL
1
=L
3
.)
Adependence of the parameter a on the strain hardening exponent n was proposed by Pardoen and Hutchinson (2000) based
on tting with a large number of axisymmetric FE cell calculations: a =0:10:217n4:83n
2
(with n dened by a hardening
law of the type (1) and 0rnr0:3). In addition, s
0
was replaced by the current mean yield stress of the matrix material s
M
,
estimated using Eq. (10). A new approach to introduce strain hardening in the Thomasons criterion is proposed hereafter.
Here, it is proposedthat the magnitude of the stress associatedwiththe localizedmode of plastic yielding in the inter-void
ligament is, in strain hardening materials, controlled by the local value of the current yield stress at the most deformed
location in the ligament s
loc
M
instead of the current average yield stress s
M
. Eq. (19) thus becomes
s
n
s
loc
M
=(1Zw
2
) 0:1
1w
wW
_ _
2
1:2

1
w
_ _
(20)
with
s
loc
M
=s
0
1
E(e
p
M
)
loc
s
0
_ _
n
-s
0
1
Ee
loc
M
s
0
_ _
n
and e
loc
M
=
_

2
3
_ e
loc
ij
_ e
loc
ij
_
dt (21)
while the parameter a of Eq. (19) is then expected to remain constant. The most deformed location in the ligament is in the
minimum section at the surface of the void, as already shown by Fabr egue and Pardoen (2008). The local strain rate
components in this region, _ e
loc
ij
, can be related to the applied strain rate and to the rate of change of void dimensions based on
simple geometrical arguments as givenin Fabr

egue and Pardoen(2008, 2009) but transposed here to plane strain conditions:
_ e
loc
33
=
_
R
3
R
3
=
1
3
_
f
f

_
W
in
W
in
2
_
W
t
W
t
_ e
22
_ e
11
_ _
, (22)
_ e
loc
22
=
_
R
2
R
2
P
e
4R
2
-
1
3
_ e
22
_ e
11

_
f
f
2
_
W
in
W
in

_
W
t
W
t
_ _
p
2

1
2
1
W
2
in
1
_ _

_
_
_
_
_
(23)
_ e
loc
11
=_ e
loc
22
_ e
loc
33
(24)
The term
_
W
in
=W
in
in Eqs. (22) and (23) is calculated using the equations provided in Appendix B, while
_
W
t
=W
t
is given in
Eq. (17). Eq. (24) states plastic volume conservation.
3.3.2. Generalization to unspecied coalescence direction
Criterion (20) is nowgeneralized to conditions where the void axis is not aligned with the principal loading direction and
the ligament where coalescence takes place is not necessarily transverse to the maximum principal stress, as
s
n
(d)
s
loc
M
(d)
=(1Zw
2
eff
(d)) 0:1
1w
eff
(d)
w
eff
(d)W
eff
(d)
_ _
2
1:2

1
w
eff
(d)

_
_
_
_
, for d c 0; p=2
_
(25)
where the void shape W
eff
d ( ) and relative void spacing w
eff
d ( ) are functions of the orientation of the ligament d dened with
respect to e
1
in the [e
1,
e
2
] plane and the stress s
n
(d) is the stress normal to the ligament (see Fig. 10d). The generalization is
thus made for coalescence occurring only within the [e
1
,e
2
] plane. The coalescence criterion has to be checked in every
directiond. Several choices canbe made todene W
eff
(d) andw
eff
(d). After different attempts anderrors basedonassessing the
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 384
predictions of the model towards the cell calculations, the following assumptions have beenselected regarding the denition
of W
eff
(d) and w
eff
(d):
+ The two quantities are dened based on an effective void radius and void spacing within the plane [e
1
,e
2
].
+ The void has an elliptical shape in the deformationplane described R
1
and R
2
and an orientation given by the state variable
y, as shown in Fig. 10b. The generalized aspect ratio is dened as W
eff
(d) =R
n
(d)=R
tr
(d), with R
n
(d) and R
tr
(d) located on the
ellipse with axes R
2
=R
n
(d =y) and R
1
=R
tr
(d =y). Hence, if C =cos
2
(dy) and D=sin
2
(dy)
R
tr
(d)
R
1
=(C(W
in
)
2
D)
1=2
= C W
1W
t
2
_ _
2
D
_ _
1=2
, (26)
R
n
(d)
R
1
= D W
1W
t
2
_ _
2
C
_ _
1=2
(27)
and thus
W
eff
(d) =
C(W((1W
t
)=2))
2
D
D(W((1W
t
)=2))
2
C
_ _
1=2
(28)
+ The centers of the voids are assumed to be initially statistically homogenously distributed in the deformation plane [e
1
,e
2
]
following an ellipsoidal symmetry (see Kaisalamand Ponte Castaneda, 1998), with half-minor axis L
10
and half-major axis
L
20
, so that l
0
=L
20
=L
10
(see Fig. 10a). As explained in Section 3.2.2, material compatibility requires that the void center
distribution follows the material deformation and remains elliptical. Fig. 10b describes the current state. The current
orientation of the axis L
2
with respect to e
2
is given by the angle c (c
0
=0). The void distribution index l
gen
is dened by
the ratio L
2
=L
J
1
, with L
J
1
the axis perpendicular to L
2
and included in the ellipse built on L
1
and L
2
. The axis L
J
1
is dened
perpendicular to the main axis. If j is the complementary of the angle between the axis L
1
and L
2
, then
cos(j)
L
J
1
_ _
2
=
1
L
1
_ _
2

sin(j)
L
2
_ _
2
(29)
+ The effective relative void spacing is then dened as w
eff
(d) =R
tr
(d)=L
tr
(d) with L
tr
(d) located on the ellipse of axis L
2
and
L
J
1
=L
tr
(d =c) (see Fig. 10d), hence
L
J
1
L
tr
(d)
=(cos
2
(dc) l
2
gen
sin
2
(dc))
1=2
(30)
Introducing (26), (29) and (30) into
w
eff
(d) =
R
tr
(d)
L
tr
(d)
=
R
tr
(d)
R
1
R
1
L
1
L
1
L
J
1
L
J
1
L
tr
(d)
(31)
leads to
w
eff
(d) =
f l
kW
t
Wexp(e
11
)
_ _
1=3
cos
2
(dc)l
2
gen
sin
2
(dc)
cos
2
(j)l
2
gen
sin
2
(j)
_ _
1=2
C W
1W
t
2
_ _
2
D
_ _
1=2
(32)
The evolution laws for l are given in Appendix B. Finally, the stress component normal to a direction given by the angle d
writes
s
d
r : ( ~ e
d2
Q ~ e
d2
) (33)
with
~
e
di
=
~
R
k
i
(d)e
k
and with
~
R(d) being the rotation matrix around e
3
~
R(d) =
cos(d) sin(d) 0
sin(d) cos(d) 0
0 0 1
_
_
_
_
_
_ (34)
3.4. Coalescence model
After condition (25) has been met for one orientation d=d
c
, plastic localization takes place in the ligament separating the
two cavities with the strain rate parallel to the ligament becoming very small, leading to a state of uniaxial straining (see e.g.
Koplik and Needleman, 1988). The only non-zero plastic strain component develops along the normal n
dc
, so that the plastic
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 385
strain rate tensor in Eq. (9) becomes
D
P

_
e
P
n
(n
dc
Qn
dc
) (35)
where _ e
P
n
is the plastic strain rate normal to the ligament. The elastic contribution is neglected so that the total normal strain
rate _ e
n
is assumed to be equal to _ e
P
n
.
Asimple geometric model has beenworkedout for deriving the unloading slope, requiredtobring a material element from
the onset of void coalescence up to nal fracture. This slope controls the amount of energy dissipated during coalescence. The
details of the derivation are provided in Scheyvaerts et al. (2010), only the basic assumptions and equations are summarized
here. The normal stress to the ligament s
n
is assumed to decrease linearly with the normal straining e
n
, from (e
c
n
,s
c
n
) at
coalescence to (e
f
n
,0) at fracture. This linear evolution has been veried by Tvergaard (1997) up to very large strains using a
remeshing procedure. Hence, the slope writes s
nc
=De
n
, where De
n
=e
f
n
e
c
n
is the additional strain increment needed to
bring a material element from the onset of coalescence to full failure. The strain increment De
n
is estimated by enforcing
uniaxial straining and by making several geometric assumptions regarding the shape of the void during coalescence. The
solution depends on the value of a shape factor g
1
, dened as g
1
=1=(2w
c
W
c
), with the quantities dened at the onset of
coalescence as W
effc
(d) W
c
and w
effc
(d) w
c
(all derivations are given in Scheyvaerts et al., 2010):
if g
1
r1
De
n
=ln 1g
1
w
nc
w
c
exp(e
c
tr
)
w
c
(3g
2
1
)(1w
c

1g
2
1
_
)((2exp(e
c
tr
))=w
c
w
2
c
(1g
2
1
)
3=2
)
w
c
exp(e
c
tr
)((R
trf
=R
trc
)w
2
c
(1g
2
1
)
3=2
)(12=p)(1w
c

1g
2
1
_
)
_

_
_

_
_
_
_
_
_
_ (36)
if g
1
41
De
n
=ln 1
p
6
w
nc
w
2
c
exp(e
c
tr
) (42exp(e
c
tr
))=w
c
1
_
_ _
(37)
with
R
trf
=R
trc
=(2exp(e
c
tr
))=w
c
(38)
The parameter w
nc
=R
nc
=L
nc
, is the relative spacing in the direction normal to the critical ligament at the onset of
coalescence.
4. Verication of the model
The predictions related to each extension of the constitutive model are assessed rst separately by comparison to the FE
cell results, while ending with the analysis of the accuracy of the full constitutive model when assembling all the different
parts. Note that the coalescence model for nal fracture has been veried elsewhere (Scheyvaerts et al., 2010).
4.1. Assessment of the void rotation law
In order to validate the void rotation model independently of the void growth model, Fig. 11a compares, for plane strain
simple shear conditions, the rotation extracted from the 3D unit-cell calculation (see Section 2.2) to the rotation estimated
with Eqs. (8) and (11) to (14) for an initial cavity with the main axis oriented along e
2
(y
0
=0) with three different initial
shapes W
0
=1/3, 1 and 3. The rotation tensor is evaluated using the current values of the porosity f, void shape W, plastic
Fig. 11. Variationof the voidorientationangle y as a functionof the shear straing under simple shear loading withs
0
/E=0.003, n=0.1, f
0
=10
2
, andl
0
=1, for
three different initial void shapes W
0
=3, 1/3, or 1, and for an initial void orientation (a) y
0
=0 and (b) y
0
=p/2. Comparison between the evolution of the void
rotation (i) extracted fromsingle void cell computations; (ii) predicted by the rotation model with other parameters coming fromthe cell calculations; and
(iii) predicted by the full void growth model.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 386
deformation rate D
P
, and deformation gradient F (of the form F=I+ge
1
Qe
2
), directly extracted from the FE cell simulations.
Fig. 11b provides the same comparison for an initial cavity axis coinciding with e
1
(y
0
=p=2). The agreement between the
rotation directly computed by the FE calculations and the calculated curve is conspicuous, regardless of the initial void shape
and rotation direction.
4.2. Assessment of the correction for the loss of spheroidicity
The predictions of the model for the evolution of 1=W
t
R
1
=R
3
based on Eqs. (16)(18) as a function of the strain level are
compared to the void cell results in Fig. 6 in the absence of shear. The results are very satisfactory, especially in Fig. 6b, which
shows that the model captures the effect of the appliedtriaxiality, the mainfactor affecting the loss of spheroidicity. The same
level of agreement is obtained in the presence of shear (results not shown).
4.3. Assessment of the generalized internal necking condition
First, the new extension of the Thomason model regarding strain hardening is validated for n=0.1 and 0.3 using void cell
calculations performed under constant T ranging from 0.577 to 2 and for a wide range of initial congurations. No shear is
involved here. Fig. 12 compares the equivalent strain at the onset of coalescence directly extracted from the FE cell
calculations, to the strain at which criterion (25) is satised with the parameters (w, W, f, s
n
) taken from the void cell
computations, and s
loc
M
calculated using Eqs. (21)(24)
1
. (Note that particularization of Eqs. (28) and (32) when d=01 and
y =0 leads to W
eff
=W
in
and w
eff
=w
1
). These predictions obtained with the coalescence model are very close to the FE cell
calculation results in all cases, justifying the newapproach used with the Thomason model based on a local estimation of the
current yield stress. This is a progress in the formulation of a versatile void coalescence condition as it does not depend
anymore on the tting parameter a(n) such as in the former version where the denition of n is also tied to one specic
hardening law(Pardoenand Hutchinson, 2000). This newformulationof the Thomasonmodel has beenrecently investigated
by Yerra et al. (2010) in the context of single crystal plasticity with similar good agreement. Note that the quality of the
predictions deteriorates quite signicantly (results not shown) when an average w is used instead of w
1
and an average W is
used instead of W
t
. This justies the necessity of the correction for the loss of spheroidicity elaborated in Section 3.2.3.
Next, Fig. 13(a) and (b) show the evolution of the right and left hand side terms of the void coalescence criterion (25) with
deformation for different amount of shear and stress triaxiality. In the void cell calculations, coalescence always takes place at
01. Hence, the criterion for the onset of coalescence was checked only at d=01. Again, all the quantities entering the coalescence
condition are directly extracted from the void cell simulations (except s
loc
M
) in order to assess this part of the model alone.
Coalescence is predicted to occur when the two curves intersect. This point must be compared to the onset of coalescence
directly predicted by the cell calculation, indicated by the dotted line. The quality of the predictions remains very good in the
presence of a shear component. The loss of ductility is properly captured. Note that as the void rotates and y changes, the use of
the complete denition of the effective void aspect ratio and effective relative void spacing as given by Eqs. (28) and (32) is
0
0.5
1
1.5
0 0.5 1 1.5
T=0.577, n=0.1
T=1, n=0.1
T=1.5, n=0.1
T=2, n=0.1
T=1.5, n=0.3

c
e
q

f
r
o
m

a
p
p
l
y
i
n
g

c
r
i
t
e
r
i
o
n

(
2
5
)

c
eq
directly from void cell calculation
E/
0
=0.003
W
0
=1/6,1, 6
f
0
=10
-2
, 10
-3
,10
-4

0
=1/2, 1, 2
Fig. 12. Results from FE void cell computations performed at T=0.577, 1, 1.5, and 2 (without shear) for E/s
0
=0.003, n=0.1 or 0.3, and for various initial
microstructures: f
0
=10
2
, 10
3
and 10
4
, W
i
=1/6, 1 and 6, l
0
=1/2, 1 and 2. Comparisonbetween the equivalent strain at the transitionto a uniaxial mode of
straining, and the strain at which criterion (25) is satised with all the parameters taken from the FE void cell computations, except s
loc
M
calculated using
Eqs. (21)(24).
1
When the model tends to predict a larger coalescence strain than given by the void cells, the parameters entering in Eq. (25) and extracted fromthe FE
void cell calculations were extrapolated fromtheir values before the onset of coalescence as if the localizationdid not happen inthe void cells, i.e. as if diffuse
plasticity carries on in the cell.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 387
necessary without any possible simplication. Hence, checking coalescence at d=01 provides a true vericationof the complete
model andis not a special limit case. Nevertheless, further works will aimat checking the validityof the generalizedcriterionfor
other morecomplex voiddistributions, as addressedfor instance byThomsonet al. (2003) (see Scheyvaerts, 2009, for additional
discussions about the orientation of the localization band as predicted by the generalized model).
4.4. General assessment
The complete extended model for void growth and coalescence is assessed now. First, Fig. 11 shows the rotation predicted
when using the full enhanced Gurson model to integrate Eq. (8). No signicant extra error on the predicted rotation rate is
Fig. 13. Evolutionof the right-and left-handsides of the voidcoalescence criterion(25) as a functionof the axial strain, for f
0
=10
3
, W
0
=1, andl
0
=1, (a) with
T=1andU
t
/U
2
=0, 0.5and1; (b) withU
t
/U
2
=1andT=1, 1.5and2. Dottedlines indicate the strainat the onset of coalescence predictedby the cell calculation; a
circle indicates the predictions of the coalescence model when applied directly on the data coming from the FE cell calculations; the arrow indicates the
prediction made by the full enhanced Gurson model.
Fig. 14. Comparison between void cell calculations and the model for n=0.1, f
0
=10
3
, l
0
=1 W
0
=1, T=1 and 1.5, and U
t
/U
2
=0 and 1: (a) axial stress vs axial
strain curves; (b) porosity variation; (c) in-plane/out-of-plane radii ratio evolution up to the onset of coalescence; and (d) void rotation.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 388
introduced when estimating f and W with the void growth model. In Fig. 13, the onset of coalescence predicted by the full
enhanced Gursonmodel is indicated by anarrow. The agreement is againvery good, witha slight overestimationof the actual
localization strain at high shear and low triaxiality.
Fig. 14 validates the model predictions for two different stress triaxialities T=1 and 1.5, with n=0.1, f
0
=10
3
, W
0
=1, and l
0
=1.
The evolutionof the (a) stress s
22
/s
0
, (b) porosityf, (c) in-plane/out-of plane radii ratioW
t
1
are givenas a functionof e
22
. The void
orientation y is given as a function of the shear strain e
12
in Fig. 14d. An arrow indicates the predicted onset of coalescence in
Fig. 14b and d. In Fig. 14c, W
t
1
is represented up to the onset of coalescence. Fig. 14a shows excellent agreement between the
predictedowcurve andthe nite element voidcell results without shear, as alreadyreportedbyPardoenandHutchinson(2000).
Whenshear is added, the ductilitypredictedbythe full model slightlyoverestimates the value givenbythe voidcell results. This is
partly due to an underestimation of the void growth rate before coalescence in the presence of shear (see Fig. 14b), and partly to
inaccuracies inthepredictionof thetransversevoidaspect ratio, W
t
1
, especiallyat T=1.5, U
t
/U
2
=1(seeFig. 14c). Thepeakstresses
and the unloading slopes are accurately reproduced irrespective of the stress triaxiality and magnitude of shear. According to
Fig. 14c, the major difference between the in-plane radius and the out-of-plane one is well reproduced by the model. Fig. 14d
shows that the void orientation is well predicted by the model up to a strain value at which the void rotation accelerates. This
acceleration is not correctly captured. The assessment is also made for initially very elongated (W
0
=6) and very at (W
0
=1/6)
voids, withdifferent initial porosity values (results not shown). As a general observation, whenthe ductility is underestimated(vs.
overestimated) in the absence of shear, the underestimation (vs. overestimation) is smaller (vs. larger) when shear is added.
5. Additional results and discussion
Materials oftenshowpreferential orientationof secondphase particles giving rise topreferential initial voidshapes and/or
orientations. Fig. 15 presents the predictions of the extended Gurson model for prolate (W
0
=6) or oblate (W
0
=1=6) cavities
with the main axis oriented in a direction y
0
. The initial porosity is modied in order to keep the relative void spacing w
0
constant and equal to 0.068 for W
0
=6, y
0
=0 (f
0
=10
3
) and W
0
=1=6, y
0
=p=2 (f
0
=610
3
) for one case; and w
0
equal to
0.226 for W
0
=1=6, y
0
=0 (f
0
=10
3
) and W
0
=6, y
0
=p=2 (f
0
=1/610
3
) for the other case. The stress triaxiality is equal
Fig. 15. Effect of the initial void orientation under T=1 for l
0
=1, W
0
=1/6 and 6; (a) axial stress vs. axial strain for w
0
=0.0684 or 0.226 and y
0
=0 or y
0
=p=2;
(b) relative void spacing vs. strain for w
0
=0.0684 or 0.226 and y
0
=0 or y
0
=p=2; (c) relative void spacing vs. strain for w
0
=0.15 and y
0
=p=4 or y
0
=p=4.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 389
to 1. No shear is involved. Fig. 15(a) and (b) showthe evolution of the axial stress and relative void spacing, respectively, as a
function of e
22
.
Fig. 15(a) and (b) indicate that the morphology of the void has an important effect, when considering identical initial void
spacing and initial in plane void aspect ratio when dened with respect to the global coordinate system. On the one hand,
whenthe voidhas the long axis parallel to the mainloading direction, i.e. for W
0
=6, y
0
=0 (f
0
=10
3
) andW
0
=1=6, y
0
=p=2
(f
0
=610
3
), the oblate case gives a muchsmaller ductility, mainlydue toa large transverse voidgrowthrate (voidopening).
On the other hand, when the void has the long axis oriented perpendicular to the main loading direction, i.e. for W
0
=1=6,
y
0
=0 (f
0
=10
3
) and W
0
=6, y
0
=p=2 (f
0
=1/610
3
), the difference in ductility is not so signicant. An intermediate result
is shown in Fig. 15c where the long axis of the voids is initially oriented at y
0
=p=4 from the main loading direction. In this
case, in addition to the initial orientation of the void, void rotation also plays a role.
Now, when comparing the two initial orientations for the prolate voids, the y
0
=p=2 case is associated with a small inter-
void ligament length (see Fig. 15b) and leads to lower ductility. Intuitively, an increase of the ductility with the y
0
=p=2
orientation with respect to the y
0
=0 orientation would be expected for the oblate void, again because of smaller w
0
. The
oppositeis observed. Thereasonis that anoblatevoidopens mainlyinthe directionof its small principal radius. The coalescence
is essentially controlled by the relative void spacing which increases much faster for y
0
=p=2 compared to y
0
=0 (see Fig. 15b).
Comparing now the prolate and oblate cases with the same initial orientation with each other, Fig. 15a shows that the
material containing prolate cavities is the most ductile when y
0
=0 due to much smaller w
0
. However, when y
0
=p=2
the difference inductilitybetweenthe twovoidaspect ratios is muchsmaller. Indeed, eventhoughthe relative voidspacing in
the oblate case is initially much smaller, it grows so fast in the transverse direction, as shown in Fig. 15b, that it eventually
becomes the largest.
This analysis shows that one should be careful when considering only one aspect ratio to describe the void geometry
without accounting for the orientation: elongated cigar-type voids and penny-shape voids evolve very differently.
Depending onthe material andforming process, one canndinreal materials these two types of limit geometries. Finally, it is
important to realize that void rotation takes place even in the absence of imposed shear loading as illustrated with the results
for y
0
=p=4 in Fig. 15c.
Fig. 16 describes generic predictions involving a coupling between the effect of an imposed shear component and the
inuence of the initial void shape and porosity. The calculations are performed at T=1, for U
t
/U
2
=0, 0.5, and 1. The results are
presented, in (aceg), for three different initial void aspect ratios (W
0
=1/6, 1, 6) and, in (bdfh), for three different initial
porosities (f
0
=10
4
, 10
3
, 10
2
). Fig. 16b shows that the initial porosity has almost no inuence on the void rotation rate.
Fig. 16(c) and (d) show again that shear leads to softer material response and advances the onset of coalescence (see also
Fig. 16(e) and (f)). Fig. 16(g) demonstrates that there is almost no inuence of shear on the evolution of the aspect ratio of
initially spherical and prolate voids, while the rotation of oblate voids has a considerable inuence on the shape evolution.
Indeed, after some amount of deformation, the void orients such that the principal loading direction does not contribute to
the opening of the voidanymore, leading eventoa tendency for the voidtoclose. Fig. 16hshows that f
0
has noinuence onthe
void shape evolution.
Fig. 17 shows the variation of the critical porosity at the onset of coalescence f
c
, normalized by f
c
under no shear, as a
functionof the shear component, for different f
0
andW
0
. For all voidaspect ratios, f
c
increases withincreasing shear. The effect
of shear on f
c
depends on the initial void shape W
0
. This gure shows that the use of a critical porosity is not a viable
coalescence criterion when, in addition to changing the stress triaxiality, various levels of shear enter the problemthat must
be modelled.
In addition to the inuence of a shear component on the evolution of the state variables during void growth, the presence
of shear also affects the onset of coalescence. Remember rst that only the stress normal to the ligament enters the
coalescence criterion (25). The assessment proposed in the paper shows that this is a viable approximation as long as the
shear contribution is not too large. Now, in principle, the shear stress component parallel to the ligament should also play a
role in triggering the localization especially when the coalescence band is close to maximum shear direction and the shear
stress is large, hence, calling for additional theoretical developments (see recent paper by Leblond and Mottet, 2008, with a
generalized treatment of the coalescence by internal necking and coalescence in shear; see also McVeigh et al., 2007).
6. Conclusions
Compared to earlier versions of the Gurson model, several extensions have been proposed within a plane strain
framework:
(1) twoadditional variables have beenintroduced, inorder todeal withloadingconditions involvingshear andvoids withthe
mainaxis not orientedparallel to the mainloading direction. The rst one is the voidorientationangle, y, the secondis the
transverse aspect ratio of the initially spheroidal void, W
t
. The evolution laws for y, borrowed from homogenization
theory, and for W
t
reproduce very well the results of void cell simulations. Knowing the evolution of W
t
, other quantities,
such as the relative void spacing in different directions, can be calculated. The evolution of the angle y directly affects the
void growth and coalescence process. The next step in the generalization is to allow any rotation in 3D and general
ellipsoidal void shapes (rst steps in this direction are under progress by Leblond and Gologanu, 2008).
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 390
Fig. 16. Effect of shear (U
t
/U
2
=0, 0.5, or 1) on the variation (as a function of the axial strain) of the (a, b) void orientation angle; (c, d) axial stress; (e, f)
porosity; (g, h) void aspect ratio, for l
0
=1, n=0.1 at T=1. For (a, c, e, g), the initial void shape is equal to W
0
=1/6, 1, or 6 and for (b, d, f, h) the initial porosity is
equal to f
0
=10
4
, 10
3
, or 10
2
.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 391
(2) The newformulationof the Thomasonvoidcoalescence criterion whichis basedona local yieldstress value takenat the
most deformed point of the ligament gives very accurate predictions of the coalescence strain while avoiding the use of
a strain hardening dependent tting parameter.
(3) The generalization of the Thomason criterion to orientation dependent void coalescence permits a check of the condition
in every direction of space. Even though the present coalescence model does not explicitly account for the shear stress
applied parallel to the ligament, it captures quite well the effect of a shear strain component on the onset of void
coalescence, and thus on the ductility. The use of a properly dened effective void aspect ratio and effective void spacing
corresponding to the direction and plane of coalescence instead of average quantities is needed for accurate void
coalescence predictions. Two extra extensions are to introduce a shear stress effect inthe coalescence conditionas well as
validating the models towards FE cell calculations with neighboring voids located at different angles (see McVeigh et al.,
2007).
The comparison with the void cell simulations has established that, for a wide range of loading conditions including (or
not) shear, the full model is able (i) to account for variations in all the characteristic parameters of the representative volume
element (ii) to capture quite precisely the onset and direction of coalescence. As a general conclusion regarding the
contribution of shear to the fracture process, there is a clear decrease of the ductility when increasing the amount of shear at
the same level of stress triaxiality, except when the rotation of the voids tends to limit the decrease of the relative void
spacing. This agrees with recent experimental results on combined shear/tensile loadings (e.g. Barsoumand Faleskog, 2007)
which could also involve additional geometry dependent plastic localization.
Acknowledgements
Fruitful discussions about the present studywithJ.-B. LeblondandG. Perrinas well as the veryconstructive comments made
bythe Referees aregratefullyacknowledged. F.S. acknowledges the support of the WalloonRegion(DGTRE) andthe Fonds Social
Europe enthrough a FIRST EUROPE Objectif 3 project under the contract EPH3310300R0302/215284. The support of the CISM
(Institut de calcul intensif et de stockage de masse) of UCL is gratefully acknowledged. This research was carried out under the
University Attraction Poles (IAP) Programme, nanced by the Belgian Science Policy under contract P6/24.
Appendix A. Periodic boundary conditions for the FE unit cell calculations
Fig. A1 provides the nomenclature used to dene the different surfaces and edges of the cell to which the boundary
conditions are applied.
Edge-top-middle
u
1
(0,L
2
,x
3
) =
1
2
U
t
(A:1)
u
2
(0,L
2
,x
3
) =
1
2
U
2
(A:2)
1
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
0 0.5 1 1.5 2 2.5
f
0
=10
-2
f
0
=10
-4
U
t
/U
2
f
c
/
f
c
(
U
1
2
=
0
)
W
0
=1/6
T=1,
0
=1,
0
=0
W
0
=1
W
0
=6
Fig. 17. Variation of the critical porosity at the onset of coalescence as a function of the relative amount of shear U
t
=U
2
, for different initial porosity f
0
=10
2
or 10
4
and void shapes W
0
=1/6, 1, or 6, with T=1, n=0.1, y
0
=0 and l
0
=1.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 392
Surface-top-left/surface-top-right
u
1
(x
1
,L
2
,x
3
)u
1
(x
1
,L
2
,x
3
) =U
t
(A:3)
u
2
(x
1
,L
2
,x
3
)u
2
(x
1
,L
2
,x
3
) =U
2
(A:4)
u
3
(x
1
,L
2
,x
3
)u
3
(x
1
,L
2
,x
3
) =0 (A:5)
Edge-top-left/edge-top-right
u
1
(7L
1
,L
2
,x
3
) =
1
2
(U
t
7U
1
) (A:6)
u
2
(7L
1
,L
2
,x
3
) =
1
2
U
2
(A:7)
u
3
(L
1
,L
2
,x
3
)u
3
(L
1
,L
2
,x
3
) =0 (A:8)
Surface-left/surface-right
u
1
(L
1
,x
2
,x
3
)u
1
(L
1
,x
2
,x
3
) =U
1
(A:9)
u
2
(L
1
,x
2
,x
3
)u
2
(L
1
,x
2
,x
3
) =0 (A:10)
u
3
(L
1
,x
2
,x
3
)u
3
(L
1
,x
2
,x
3
) =0 (A:11)
Edge-bottom-left/edge-bottom-right
u
1
(7L
1
,0,x
3
) = 7
1
2
U
1
(A:12)
u
2
(7L
1
,0,x
3
) =0 (A:13)
u
3
(L
1
,0,x
3
)u
3
(L
1
,0,x
3
) =0 (A:14)
Fig. A1. Finite element meshfor the1/4of theunit cell for acubic primitivelatticeshowingthedifferent regions ontheexternal surfaces onwhichtheperiodic
boundary conditions are applied. The origin of the reference coordinate system corresponds to the center of the void in the undeformed conguration.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 393
Edge-bottom-middle
u
1
(0,0,x
3
) =0 (A:15)
u
2
(0,0,x
3
) =0 (A:16)
Surface-bottom-left/surface-bottom-right
u
1
(x
1
,0,x
3
) u
1
(x
1
,0,x
3
) =0 (A:17)
u
2
(x
1
,0,x
3
) u
2
(x
1
,0,x
3
) =0 (A:18)
u
3
(x
1
,0,x
3
)u
3
(x
1
,0,x
3
) =0 (A:19)
Surface-back
u
3
(x
1
,x
2
,0) =0 (A:20)
Surface-front
u
3
(x
1
,x
2
,L
3
) =
1
2
U
3
(A:21)
Note that for the plane strain analyses addressed in this paper U
3
=0.
Appendix B. Evolution laws of the non dimensional geometrical parameters
The void is characterizedby three radii R
1
, R
2
, andR
3
andthe representative cell by three lengths L
1
, L
2
, andL
3
. Several non-
dimensional parameters can be used to characterize the geometry:
the porosity f =k
g
R
1
R
2
R
3
L
1
L
2
L
3
, (B:1)
the average void aspect ratio W=
2R
2
(R
1
R
3
)
, (B:2)
the transverse void aspect ratio W
t
=
R
3
R
1
, (B:3)
the in plane void aspect ratio W
in
=
R
2
R
1
, (B:4)
the average void distribution parameter l =
2L
2
(L
1
L
3
)
, (B:5)
the transverse void distribution parameter l
t
=
L
3
L
1
, (B:6)
the relativevoid spacing along each axis w
1
=
R
1
L
1
, w
2
=
R
2
L
2
, w
3
=
R
3
L
3
, (B:7)
The geometry is fully characterized by ve of these quantities. The evolution of the void distribution parameters follows
the overall deformation of the material, hence
_
l
t
l
t
=
_
L
3
L
3

_
L
1
L
1
=D
p
n33
D
p
n11
, (B:8)
_
l
l
=D
P
n22

D
P
n11
D
P
n33
l
t
1l
t
, (B:9)
The evolutions of f and W, which can be expressed as
_
f
f
=
_ w
1
w
1

_ w
2
w
2

_ w
3
w
3
=
_
R
1
R
1

_
R
2
R
2

_
R
3
R
3
D
P
nkk
(B:10)
and
_
W
W
=
_
R
2
R
2

(
_
R
1
=R
1
)(
_
R
3
=R
3
)(W
t
)
1W
t
(B:11)
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 394
are predicted by the constitutive model Eqs. (6) and (7). The postulated Eq. (15) provides a fth equation from which the
evolution laws for W
t
, W
in
and for the w
i
s can be derived. The starting point is Eq. (B.10) which leads, after some elementary
algebra, to
_ w
1
w
1
=KCD
P
n11
(B:12)
_
W
t
W
t
=C(1W
t
)K (B:13)
where
K = D
P
n11
D
P
n33
l
t
l
t0
_ _
w
10
pw
1
(B:14)
C =
((
_
f =f )D
P
nkk
(
_
W=W)K((2W
t
)=(1W
t
))
14W
t
W
2
t
(1W
t
) (B:15)
References
ABAQUS/Standard Version 6.5, 2004. Users Manual. Hibbit, Karlsson & Sorensen, Inc., Providence, RI, US.
Aravas, N., Ponte Castaneda, P., 2004. Numerical methods for porous metals withdeformation-induced anisotropy. Computer Methods inAppliedMechanics
and Engineering 193, 37673805.
Barsoum, I., Faleskog, J., 2007. Rupture mechanisms in combined tension and shearexperiments. International Journal of Solids and Structures 44,
17681786.
Becker, R., 1987. The effect of porosity distribution on ductile failure. Journal of the Mechanics and Physics of Solids 35, 577599.
Becker, R., Needleman, A., Suresh, S., Tvergaard, V., Vasudevan, A.K., 1989a. Ananalysis of ductile failure by grainboundary voidgrowth. Acta Metallurgica 37,
99120.
Becker, R., Smelser, R., Richmond, O., Appleby, E., 1989b. The effect of void shape on void growth and ductility in axisymmetric tension tests. Metallurgical
Transactions A 20A, 853861.
Benzerga, A.A., Besson, J., 2001. Plastic potentials for anisotropic porous solids. European Journal of Mechanics A/Solids 20A (3), 397434.
Benzerga, A.A., 2002. Micromechanics of coalescence in ductile fracture. Journal of the Mechanics and Physics of Solids 50, 13311362.
Benzerga, A.A., Besson, J., Batisse, R., Pineau, A., 2002. Synergistic effects of plastic anisotropy and void coalescence on fracture mode in plane strain.
Modelling and Simulation in Material Science and Engineering 10, 73102.
Benzerga, A.A., Besson, J., Pineau, A., 2004. Anisotropic ductile fracturePart II: theory. Acta Materialia 52, 46394650.
Besson, J., Devillers-Guerville, L., Pineau, A., 2000. Modeling of scatter and size effect in ductile fracture: application to thermal embrittlement of duplex
stainless steels. Engineering Fracture Mechanics 67, 169190.
Benzerga, A.A., Leblond, J.-B., 2010. Ductile fracture by void growth to coalescence. Advances in Applied Mechanics 44, 169305.
Besson, J., Guillemer-Neel, C., 2003. An extension of the Green and Gurson models to kinematic hardening. Mechanics of Materials 35, 118.
Bordreuil, C., Boyer, J.-C., Salle , E., 2003. On modeling the growth and the orientation changes of ellipsoidal voids in a rigid plastic matrix. Modelling and
Simulation in Materials Science and Engineering 11, 365380.
Brocks, W., Sun, D.Z., H onig, A., 1995. Vericationof the transferabilityof micromechanical parameters bycell model calculations withviscoplastic materials.
International Journal of Plasticity 11, 971989.
Brown, L.M., Embury, J.D., 1973. The initiation and growth of voids at second phase particles. In: Proceedings of the Third International Conference on the
Strength of Metals and Alloys, ICSMA 3, Cambridge, England, pp. 164169.
Budiansky, B., Hutchinson, J.W., Slutsky, S., 1982. Void growth and collapse in viscous solids. In: Hopkins, H.G., Sewell, M.J. (Eds.), Mechanics of Solids. The
Rodney Hill 60th Anniversary Volume. Pergamon Press, Oxford, pp. 1345.
Chehab, B., Bre chet, Y., Ve ron, M., Jacques, P.J., Parry, G., Glez, J.-C., Mithieux, J.-D., Pardoen, T., 2010. Micromechanics of the high temperature damage in a
dual phase stainless steel. Acta Materialia 58, 626637.
DEscatha, Y., Devaux, J.C., 1979. Numerical studyof initiation, stable crackgrowthandmaximumloadwitha ductile fracture criterionbasedonthe growthof
holes. In: Landes, J.D., Begley, J.A., Clarke, G.A. (Eds.), Elastic Plastic Fracture, 668. ASTM STP, pp. 229248.
Devaux, J.C., Mudry, F., Pineau, A., Rousselier, G., 1989. Experimental and numerical validation of a ductile fracture local criterion based on a simulation of
cavity growth. In: Landes, J.D., Saxena, A., Merkle, J.G. (Eds.), Non Linear Fracture Mechanics, vol. IIElasticPlastic Fracture, ASTM STP 995. American
Society for Testing and Materials, Philadelphia, pp. 723.
Enakoutsa, K., Leblond, J.-B., Audoly, B., 2005. Inuence of continuous nucleation of secondary voids upon growth and coalescence of cavities in porous
ductile metals. In: Carpinteri, A. (Ed.), Proceedings of the 11th International Conference on Fracture, March 2025, Turin, Italy, CD-Rom.
Fabr egue, D., Pardoen, T., 2008. Aconstitutive model for elastoplastic solids containing primary and secondary voids. Journal of the Mechanics and Physics of
Solids 56, 719741.
Fabr egue, D., Pardoen, T., 2009. Corrigendum to A constitutive model for elastoplastic solids containing primary and secondary voids. Journal of the
Mechanics and Physics of Solids 57, 869870.
Faleskog, J., Gao, X., Shih, C., 1998. Cell model for nonlinear fracture analysisI. Micromechanics calibration. International Journal of Fracture 89, 355373.
Flandi, L., Leblond, J.-B., 2005. Anewmodel for porous nonlinear viscous solids incorporating voidshape effectsI: Theory. EuropeanJournal of Mechanics A/
Solids 24, 537551.
Fleck, N.A., Hutchinson, J.W., 1986. Void growth in shear. Proceedings of the Royal Society of London A407, 435458.
Fleck, N.A., Hutchinson, J.W., Tvergaard, V., 1989. Softening by void nucleation and growth in tension and shear. Journal of the Mechanics and Physics of
Solids 37, 515540.
Gallais, C., Simar, A., Fabr egue, D., Denquin, A., Lapasset, G., de Meester, B., Brechet, Y., Pardoen, T., 2007. Multiscale analysis of the strength and ductility of
friction stir welded 6056 Al joints. Metallurgical and Materials Transactions A 38A, 964981.
Gao, X., Faleskog, J., Shih, C.F., 1998. Cell model for nonlinear fracture analysisII. Fracture-process calibration and verication. International Journal of
Fracture 89, 374386.
Gologanu, M., Leblond, J.B., Devaux, J., 1993. Approximate models for ductile metals containing nonspherical voidscase of axisymmetric prolate ellipsoidal
cavities. Journal of the Mechanics and Physics of Solids 41, 17231754.
Gologanu, M., Leblond, J.-B., Devaux, J., 1994. Approximate models for ductile metals containing non-spherical voidscase of axisymmetric oblate
ellipsoidal cavities. ASME Journal of Engineering Materials and Technology 116, 290297.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 395
Gologanu, M., Leblond, J.-B., Perrin, G., Devaux, J., 1997. Recent extensions of Gursons model for porous ductile metals. In: Suquet, P. (Ed.), Continuum
Micromechanics. CISM Lectures Series. Springer, NewYork, pp. 61130.
Grange, M., Besson, J., Andrieu, E., 2000. An anisotropic Gurson model to represent the ductile rupture of hydrided Zircaloy4 sheets. International Journal of
Fracture 105 (3), 273293.
Gurson, A., 1977. Continuumtheory of ductile rupture byvoidnucleationandgrowth: Part IYieldcriteria andowrules for porous ductile media. Journal of
Engineering Materials and Technology 99, 215.
Hom, C.L., Mc Meeking, R.M., 1989. Void growth in elasticplastic materials. Journal of Applied Mechanics 56, 309317.
Huber, G., Brechet, Y., Pardoen, T., 2005. Void growth and void nucleation controlled ductility in quasi eutectic cast aluminium alloys. Acta Materialia 53,
27392749.
Kailasam, M., Ponte Castaneda, P., Willis, J.R., 1997. The effect of particle size, shape, distribution and their evolution on the constitutive response of
nonlinearly viscous compositesII. Examples. Philosophical Transactions of the Royal Society of London A 355, 18531872.
Kaisalam, M., Ponte Castaneda, P., 1998. Ageneral constitutive theory for linear andnonlinear particulate media withmicrostructure evolution. Journal of the
Mechanics and Physics of Solids 46, 427465.
Kailasam, M., Aravas, N., Ponte Castaneda, P., 2000. Porous metals withdeveloping anisotropy: constitutive models, computational issues and applications to
deformation processing. Computational Modelling in Engineering Science 1, 105118.
Keralavarma, S.M., Benzerga, A.A., 2008. An approximate yield criterion for anisotropic porous media. Comptes Rendus Me canique 336, 685692.
Keralavarma, S.M., Benzerga, A.A., 2010. Aconstitutive model for plastically anisotropic solids withnon-spherical voids. Journal of the Mechanics andPhysics
of Solids 58, 874901.
Kl ocker, H., Tvergaard, V., 2003. Growth and coalescence of non-spherical voids in metals deformed at elevated temperature. International Journal of
Mechanical Sciences 25, 12831308.
Koplik, J., Needleman, A., 1988. Void growth and coalescence in porous plastic solids. International Journal of Solids and Structures 24, 835853.
Koppenhoefer, K.C., Dodds, R.H., 1998. Ductile crack growth in pre-cracked CVN specimens: numerical studies. Nuclear Engineering and Design 180,
221241.
Lassance, D., Fabr egue, D., Delannay, F., Pardoen, T., 2007. Micromechanics of room and high temperature fracture in 6xxx Al alloys. Progress in Materials
Science 52, 62129.
Lassance, D., Scheyvaerts, F., Pardoen, T., 2006. Growth and coalescence of penny-shaped voids in metallic alloys. Engineering Fracture Mechanics 73,
10091034.
Leblond, J.B., Gologanu, M., 2008. External estimate of the yield surface of an arbitrary ellipsoid containing a confocal void. Comptes Rendus Me canique 336,
813819.
Leblond, J.B., Mottet, G., 2008. Atheoretical approach of strain localization within thin planar bands in porous ductile materials. Comptes Rendus Me canique
336, 176189.
Leblond, J.B., Perrin, G., Devaux, J., 1995. An improved Gurson-type model for hardenable ductile metals. European Journal of Mechanics A/Solids 14,
499527.
Liao, K.C., Pan, J., Tang, S.C., 1997. Approximate yield criteria for anisotropic porous ductile sheet metals. Mechanics of Materials 26, 213226.
McClintock, F.A., 1968. A criterion for ductile fracture by the growth of holes. Journal of Applied Mechanics 35, 363371.
McVeigh, C., Vernerey, F., Liu, W.K., Moran, B., Olson, G., 2007. An interactive micro-void shear localization mechanismin high strength steels. Journal of the
Mechanics and Physics of Solids 55, 225244.
Mear, M., Hutchinson, J., 1985. Inuence of yield surface curvature on ow localization in dilatant plasticity. Mechanics of Materials 4, 395407.
Monchiet, V., Cazacu, O., Charkaluk, E., Kondo, D., 2008. Macroscopic yield criteria for plastic anisotropic materials containing spheroidal voids. International
Journal of Plasticity 241, 1581189.
Morgeneyer, T.F., Starink, M.J., Sinclair, I., 2008. Evolution of voids during ductile crack propagation in an aluminium alloy sheet toughness test studied by
synchrotron radiation computed tomography. Acta Materialia 56, 16711679.
M uhlich, U., Brocks, W., 2003. On the numerical integration of a class of pressure-dependent plasticity models including kinematic hardening.
Computational Mechanics 31, 479488.
Nahshon, K., Hutchinson, J.W., 2008. Modication of the Gurson model for shear failure. European Journal of Mechanics A/Solids 27, 117.
Needleman, A., Tvergaard, V., 1987. An analysis of ductile rupture modes at a crack tip. Journal of the Mechanics and Physics of Solids 35, 151183.
Nielsen, K.L., Tvergaard, V., 2009. Effect of a shear modied Gurson model on damage development in a FSWtensile specimen. International Journal of Solids
and Structures 46, 587601.
Nielsen, K.L., Simar, A., de Meester, B., Pardoen, T., 2010. The effect of stage IV hardening on localization and damage development in friction stir welded
6005A aluminum alloy. International Journal of Solids and Structures 47, 23592370.
Onck, P.R., van der Giessen, E., 1999. Growth of an initially sharp crack by grain boundary cavitation. Journal of the Mechanics and Physics of Solids 47,
99139.
Pardoen, T., 2006. Numerical simulation of low stress triaxiality ductile fracture. Computers and Structures 84, 16411650.
Pardoen, T., Doghri, I., Delannay, F., 1998. Experimental and numerical comparison of void growth models and void coalescence criteria for the prediction of
ductile fracture in copper bars. Acta Materialia 46, 541552.
Pardoen, T., Dumont, D., Deschamps, A., Brechet, Y., 2003. Grain boundary versus transgranular ductile failure. Journal of the Mechanics and Physics of Solids
51, 637665.
Pardoen, T., Brechet, Y., 2004. Inuence of microstructure-driven strain localization on the ductile fracture of metallic alloys. Plilosophical Magazine 84,
269297.
Pardoen, T., Hachez, F., Marchioni, B., Blyth, H., Atkins, A.G., 2004. Mode I fracture of sheet metal. Journal of the Mechanics and Physics of Solids 52,
423452.
Pardoen, T., Hutchinson, J., 2000. An extended model for void growth and coalescence. Journal of the Mechanics and Physics of Solids 48, 24672512.
Pardoen, T., Hutchinson, J., 2003. Micromechanics-based model for trends in toughness of ductile metals. Acta Materialia 51, 133148.
Pardoen, T., Marchal, Y., Delannay, F., 1999. Thickness dependence of cracking initiation criteria in thin aluminum plates. Journal of the Mechanics and
Physics of Solids 47, 20932123.
Pardoen, T., Scheyvaerts, F., Simar, A., Teko glu, C., Onck, P.R., 2010. Multiscale modeling of ductile failure in metallic alloys. Comptes Rendus Physique 11,
326345.
Peirce, D., Shih, C.F., Needleman, A., 1984. A tangent modulus method for rate dependent solids. Computers and Structures 18, 875887.
Perrin, G., Leblond, J.-B., 2000. Accelerated void growth in porous ductile solids containing two populations of cavities. International Journal of Plasticity 16,
91120.
Pineau, A., Pardoen, T., 2007. Failure mechanisms of metals. In: Milne, I., Ritchie, R.O., Karihaloo, B.L. (Eds.), Comprehensive Structural Integrity Encyclopedia,
vol. 2. Elsevier (Chapter 6, on-line).
Ponte Castaneda, P., Zaidman, M., 1994. Constitutive models for porous materials with evolving microstructure. Journal of the Mechanics and Physics of
Solids 42, 14591497.
Rice, J.R., Tracey, D.M., 1969. On the ductile enlargement of voids in triaxial stress. Journal of the Mechanics and Physics of Solids 17, 201217.
Ruggieri, C., Panontin, T.L., Dodds Jr., R.H., 1996. Numerical modeling of ductile crack growth in 3-Dusing computational cell elements. International Journal
of Fracture 82, 6795.
Scheyvaerts, F., 2009. Multiscale modelling of ductile fracture in heterogeneous metallic alloys, Ph.D. Thesis, Universite catholique de Louvain.
Scheyvaerts, F., Onck, P., Pardoen, T., 2010. A new model for void coalescence by internal necking. International Journal of Damage Mechanics 19, 95126.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 396
Simar, A., Nielsen, K.L., de Meester, B., Tvergaard, V., Pardoen, T., 2010. Micro-mechanical modellingof damage in6005Aaluminiumbasedona physical strain
hardening law including stage IV. Engineering Fracture Mechanics 77, 24912503.
Siruguet, K., Leblond, J.B., 2004. Effect of void locking by inclusions upon the plastic behavior of porous ductile solidsI: Theoretical modeling and numerical
study of void growth. Intern. Journal of Plasticity 20, 225254.
Steglich, D., Brocks, W., 1997. Micromechanical modelling of the behaviour of ductile materials including particles. Computational Materials Science 9, 717.
Steglich, D., Brocks, W., Heerens, J., Pardoen, T., 2008. Anisotropic ductile damage modelling of Al2024 alloys. Engineering Fracture Mechanics 75,
36923706.
Teko glu, C., Pardoen, T., 2010. A micromechanics based damage model for composite materials. International Journal of Plasticity 26, 549569.
Thomason, P.F., 1985a. Three-dimensional models for the plastic limit-load at incipient failure of the intervoid matrix in ductile porous solids. Acta
Metallurgica 33, 10791085.
Thomason, P.F., 1985b. A three-dimensional model for ductile fracture by the growth and coalescence of microvoids. Acta Metallurgica 33, 10871095.
Thomason, P.F., 1990. Ductile Fracture of Metals. Pergamon Press, Oxford.
Thomson, C.I.A., Worswick, M.J., Pilkey, A.K., Lloyd, D.J., 2003. Void coalescence within periodic clusters of particles. Journal of the Mechanics and Physics of
Solids 51, 127146.
Tvergaard, V., 1990. Material failure by void growth to coalescence. Advances in Applied Mechanics 27, 83151.
Tvergaard, V., 1997. Studies of void growth in a thin ductile layer between ceramics. Computational Mechanics 20, 186191.
Tvergaard, V., 2008. Shear deformation of voids with contact modeled by internal pressure. International Journal of Mechanical Sciences 50, 14591465.
Tvergaard, V., 2009. Behaviour of voids in a shear eld. International Journal of Fracture 158, 4149.
Tvergaard, V., Needleman, A., 1984. Analysis of the cup and cone fracture in a round tensile bar. Acta Metallurgica 32, 157169.
Wang, D.A., Pan, J., Liu, S.D., 2004. An anisotropic Gurson yield criterion for porous ductile sheet metals with planar anisotropy. International Journal of
Damage Mechanics 13, 733.
Xia, L., Shih, C., Hutchinson, J., 1995. A computational approach to ductile crack growth under large scale yielding conditions. Journal of the Mechanics and
Physics of Solids 43, 389413.
Xia, L., Shih, C.F., 1995. Ductile crack growthI. A numerical study using computational cells with microstructurally based length scales. Journal of the
Mechanics and Physics of Solids 43, 233259.
Xue, L., 2007. Damage accumulation and fracture initiation in uncracked ductile solids subject to triaxial loading. International Journal of Solids and
Structures 44, 51635181.
Yerra, S.K., Teko glu, C., Scheyvaerts, F., Delannay, L., Van Houtte, P., Pardoen, T., 2010. Void growth and coalescence in single crystals. International Journal of
Solids and Structures 47, 10161029.
Zhang, Z.L., Niemi, E., 1995. A new failure criterion for the GursonTvergaard dilational constitutive model. International Journal of Fracture 70, 321334.
Zhang, Z.L., Thaulow, C., Odegard, J., 2000. A complete Gurson model approach for ductile fracture. Engineering Fracture Mechanics 67, 155168.
F. Scheyvaerts et al. / J. Mech. Phys. Solids 59 (2011) 373397 397

You might also like