Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

ARTICLE IN PRESS

International Journal of Plasticity xxx (2009) xxx–xxx

Contents lists available at ScienceDirect

International Journal of Plasticity


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

Damage of ductile materials deforming under multiple


plastic or viscoplastic mechanisms
J. Besson
Centre des Matériaux, Mines ParisTech, UMR CNRS 7633, BP 87, 91003 Evry Cedex, France

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

Article history: This work presents a model to represent ductile failure (i.e. failure
Received 11 September 2008 controlled by nucleation, growth and coalescence) of materials
Received in final revised form 17 March whose irreversible deformation is controlled by several plastic or
2009
viscoplastic deformation mechanisms. In addition work hardening
Available online xxxx
may result from both isotropic and kinematic hardening. Damage
is represented by a single variable representing void volume frac-
tion. The model uses an additive decomposition of the plastic strain
Keywords:
Ductile rupture
rate tensor. The model is developed based on the definition of dam-
Deformation mechanisms age dependant effective scalar stresses. The model is first devel-
Kinematic hardening oped within the generalized standard material framework and
expressions for Helmholtz free energy, yield potential and dissipa-
tion potential are proposed. In absence of void nucleation, the evo-
lution of the void volume fraction is governed by mass
conservation and damage does not need to be represented by state
variables. The model is extended to account for void nucleation. It
is implemented in a finite element software to perform structural
computations. The model is applied to three case studies: (i) failure
by void growth and coalescence by internal necking (pipeline steel)
where plastic flow is either governed by the Gurson–Tvergaard–
Needleman model or the Thomason model, (ii) creep failure (Grade
91 creep resistant steel) where viscoplastic flow is controlled by
dislocation creep or diffusional creep and (iii) ductile rupture after
pre-compression (aluminum alloy) where kinematic hardening
plays an important role.
Ó 2009 Elsevier Ltd. All rights reserved.

E-mail address: jacques.besson@ensmp.fr

0749-6419/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2009.03.001

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

2 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

1. Introduction

Constitutive models able to represent the strength and toughness of ductile materials have found
increasing interest and application in the context of the local description of fracture (Berdin et al.,
2004; Pineau, 2006). Models include a description of plastic behavior and damage growth. Two
mains approaches exist to derive constitutive equations representative of ductile damage. The first
one relies on a micromechanical description of the material which was initiated by the work of Gur-
son (1977) where a porous material containing spherical voids embedded in a plastic matrix was
considered. This original model was extended to account for plastic hardening and to give a realistic
description of material ductility (so called Gurson–Tvergaard–Needleman model (GTN); Tvergaard
and Needleman, 1984). This approach, and in particular the GTN model, has received considerable
attention and has found applications in the integrity assessment of structures (see, e.g. Nègre
et al., 2004; Dotta and Ruggieri, 2004). The second approach is based on a more phenomenological
description of damage and is often referred to as ‘‘continuum damage mechanics” (CDM) (Lemaitre,
1985, 1996; Lemaitre and Desmorat, 2005; Hammi and Horstemeyer, 2007; Celentano and Chab-
oche, 2007). This approach is derived based on a thermodynamic framework but it can be shown
that the micromechanically based models can also be described within this framework (Besson
and Guillemer-Neel, 2003).
As they rely on a realistic description of the material microstructure, micromechanically based
models have been enriched to account for microstructural features such as void shape, void spac-
ing (Gologanu et al., 1993, 1994; Pardoen and Hutchinson, 2000; Monchiet et al., 2008), void
rotation (Bordreuil et al., 2003) and void coalescence (Thomason, 1990; Benzerga et al., 1999;
Pardoen and Hutchinson, 2000). However, the description of the matrix behavior (i.e. the undam-
aged material) remains relatively simple as isotropic plastic hardening is generally assumed.
Viscoplasticity can easily be accounted for Tvergaard and Needleman (1986); Tvergaard (1990)
and specific expressions for the flow surface have been proposed in Leblond et al. (1994)
depending on the strain rate sensitivity. However, plastic deformation mechanisms can be much
more complex due to (i) the mixed isotropic/kinematic nature of hardening and (ii) the fact that
various physical mechanisms may contribute to the overall deformation (see Chaboche (2008) for
review). Coupling of ductile damage and mixed hardening using micromechanical models has
been treated by several authors using different formulations (Mear and Hutchinson, 1985; Arndt
et al., 1997; Leblond et al., 1995; Besson and Guillemer-Neel, 2003). Note that this coupling is
easily accounted for using the CDM approach (Lemaitre and Desmorat, 2005). Accounting for
multiple deformation mechanisms together with ductile damage does not appear to have been
treated.
The aim of this work is to propose a general micromechanically based framework allowing to de-
scribe ductile damage growth in a plastic/viscoplastic material whose deformation is controlled by
different mechanisms. Mixed isotropic/kinematic hardening is also accounted for. The text is orga-
nized as follows. The model is first presented in Section 2. The model is derived with the General-
ized Standard Material framework (Halphen and Nguyen, 1975) which requires to define state
variables and a related free energy function (Section 2.3). Rate equations for the state variables
are derived from a flow and a dissipation potential (Section 2.4).
Three different applications of the modeling framework are then proposed. The first one (Section
3) deals with the modeling of ductile failure by void growth and coalescence by internal necking
following the approach proposed in Benzerga et al. (1999); Pardoen and Hutchinson (2000, 2003)
in the limiting case where cavities are assumed to remain spherical (Zhang and Niemi, 1995). The
model is used to represent both deformation modes: (i) diffuse plasticity around the voids, (ii) local-
ized plasticity in the intervoid ligament.The second application (Section 4) deals with creep failure
in the case where creep deformation results from two different physical phenomena: (i) dislocation
creep (at high stresses) and (ii) diffusional creep (at low stresses).Finally the model is applied in Sec-
tion 5 to simulate experiments reported in Bao and Treitler (2004) where notched bars were frac-
tured after compressive prestraining. This third example shows the ability of the model to deal
with mixed isotropic/kinematic hardening.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 3

2. Model for porous materials with multiple deformation mechanisms

The model which is developed in this section is based both on constitutive equations for undam-
aged materials deforming under multiple mechanisms (Besson et al., 2001a; Lemaitre and Chaboche,
1990; Chaboche, 2008) and on an extension of models for porous ductile materials including kine-
matic hardening (Besson and Guillemer-Neel, 2003). The present model is designed in such a way that
these models are retrieved in cases (i) where damage is absent or (ii) where a single deformation
e
mechanism is present. The strain rate tensor e_ is defined as the sum of the elastic strain rate e_ and
of several contributions to the plastic strain rate e_ pi with i ¼ 1 . . . N where N is the number of
(visco)plastic deformation mechanisms contributing to the overall plastic strain:
X
N
e_ ¼ e_ e þ e_ pi ¼ e_ e þ e_ p ð1Þ
i¼1
p P
where e_ ¼ Ni¼1 e_ pi is the plastic strain rate tensor.
In the case of void growth dominated ductile failure, damage is identified to the porosity, f. Damage
is consequently isotropic in the following. The model is first developed for void growth damage and is
extended to include void nucleation (Section 2.5).

2.1. Effective stresses

The model is based on the definition of scalar effective stress ðaH Þ associated with stress tensors ðaÞ.
The effective scalar stress aH is a representation of the matrix loading when subjected to a macro-
scopic stress a (either Cauchy stress or back stress in the following) for a given damage level. The effec-
tive scalar stress is defined explicitly or implicitly by the following equation:
Sða; f ; aH Þ ¼ Sðaeq ; akk ; f ; aH Þ ¼ 0 ð2Þ
In the following the aH will be expressed using the first two invariants of the stress tensor a: its von
Mises invariant aeq and its trace akk . The resulting model is therefore isotropic. To reflect the damaging
effect of porosity, aH must be an increasing function of f:
aH ða; f1 Þ < aH ða; f2 Þ if f 1 < f2 ð3Þ
In the following, it will be assumed that aH is an homogeneous function of a of degree 1 and differen-
tiable. Consequently, the Euler’s theorem applies so that:
@aH
a: ¼ aH ð4Þ
@a
In the case where multiple deformation mechanisms are accounted for a function, Si , must be defined
for each mechanism. Various forms for S can be chosen depending on the selected model; examples
will be given in the sequel. Note that the Rousselier model (Rousselier, 1987) cannot be used within
the present framework as it does not allow for the definition of an effective stress which is an homo-
geneous function of degree 1. A modification of the Rousselier model has been proposed by Tanguy
and Besson (2002) which is consistent with the present framework.

2.2. Damage evolution

Due to mass conservation, porosity is related to the plastic strain tensor so that
p
f_ ¼ ð1  f Þtre_ ð5Þ
A porosity, fi , may be associated with each mechanism. It represents the contribution of each mech-
anism to the overall damage. Its evolution is given by:
X
N
f_ i ¼ ð1  f Þtre_ pi so that fi ¼ f ð6Þ
i¼1

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

4 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

Indeed damage caused by one single mechanism may affect the other mechanisms so that it is neces-
sary to define a damage variable f i resulting from the interaction:
X
N
f ¼ Hfij f j ð7Þ
i
j¼1

where Hfij is the corresponding interaction matrix. In most cases damage fi will affect all deformation
mechanisms so that Hfij ¼ 1 should be used. In that case f i ¼ f . Effective stresses for mechanism i are
then defined using the new damage variable as:

Si ða; f i ; aiH Þ ¼ 0 ð8Þ


where aiH is the effective stress corresponding to a for mechanism i.

2.3. State variables and free energy

The material is described within the Generalized Standard Material (GSM) framework (Halphen
and Nguyen, 1975). This framework was initially developed to deal with (visco)plasticity and was also
used in the case of CDM to incorporate damage considered as a reduction of sound area fraction
(Lemaitre and Desmorat, 2005). Recent works, including this paper, have proposed various extensions
of the framework including void volume fraction as a damaging factor (Besson and Guillemer-Neel,
2003; Chaboche et al., 2006; Brünig, 2003; Brünig et al., 2008).
The elastic strain tensor ee is taken as a state variable. Hardening for each deformation mechanism
is characterized by state variables representing isotropic and kinematic hardening. The material is
characterized by N isotropic hardening variables p1 ; . . . ; pN (each one is associated to a single deforma-
tion mechanism) and by N k kinematic variables a1 ; . . . ; aNk . Several kinematic variables may be at-
tached to a single mechanism. The free energy is then given as:

1 X
N
W ¼ ee : K : ee þ ð1  f Þ i Þ þ Wa ða1 ; . . . ; aI ; . . . ; aNk Þ
Gi ðp ð9Þ
2 i¼1

where K is the elastic stiffness tensor. It may depend on damage; the dependence can then be ob-
tained from micromechanical analyses (see, e.g. Kachanov, 1994). In practice this dependence is often
neglected as the stress level is controlled by the plastic flow potential of the voided material. On the
other hand, the dependence is essential in the case of CDM models as it allows to define a strain energy
release rate which is used to derive the damage rate (Lemaitre and Desmorat, 2005). Note that these
models do not usually consider plastic volume change which controls damage in the present case so
that an additional damage evolution law must be used.
The ‘‘thermodynamic forces” associated with the state variables are given by the partial derivative
of W with respect to the state variables as follows:

@W @W @Gi @ Wa
r¼ ¼ K : ee ; Ri ¼ ¼ ð1  f Þ ¼ ð1  f ÞRiH ; XI ¼ ð10Þ
@ ee @pi @pi @ aI

i Þ is intro-
To represent cross hardening between mechanisms a new isotropic hardening variable ðp
duced for each mechanism:
X
N
i ¼
p Hpij pj ð11Þ
j¼1

where Hpij is the interaction matrix. The function Gi is then defined as:
Z pi
Gi ðpi Þ ¼ i Þdpi
RiH ðp ð12Þ
0

where stress RiH represents matrix isotropic hardening for mechanism i. Each kinematic hardening
variable, aI is associated to a unique deformation mechanism, i. The indicator function viI is used in

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 5

the following; it is equal to 1 if the variable aI is associated to mechanism i and to 0 otherwise. For
each mechanism, a back stress X i is then defined as:
Nk
X
Xi ¼ viI X I ð13Þ
I¼1

It represents the translation of the yield surface.


In absence of cross interaction between kinematic variables the back stress X I is constructed as fol-
lows. An intermediate tensor, vI is constructed as:
2 CI
vI ¼ aI ð14Þ
3 1f
where C I > 0 is a material coefficient. X I is such as:
2 X IH
vI ¼ X IH ð15Þ
3 XI
where X IH is the effective stress associated with X I using Eq. (2). It can be shown (Besson and Guill-
emer-Neel, 2003) that the corresponding free energy term is expressed as:
Nk
X  
Wa ¼ WaI with WaI ¼ max aI : x  ð1  f Þx2H =ð2C I Þ ð16Þ
x
I¼1

The derivation of the model including kinematic hardening is summarized in Appendix A. In the case
where cross hardening between kinematic variables is needed, the intermediate stress, vI , is expressed
as:
J¼N
2 C I Xk k
vI ¼ H aJ ð17Þ
3 1  f J¼1 IJ

where HkIJ is the interaction matrix. No expression for the corresponding free energy was found and its
existence is postulated.

2.4. Flow and dissipation potentials: rate equations

For each mechanism, a flow potential is defined as:


 i Þ  ki
/iH ¼ BiH  RiH ðp with Bi ¼ r  X i ð18Þ
ki represents the radius of the initial elastic domain and RiH its increase caused by isotropic hardening.
k þ RiH is the flow stress. In the case of a plastic (resp. viscoplastic) mechanism, yielding occurs for
/iH ¼ 0; /_ iH ¼ 0 (resp. /iH P 0).
The evolution of state variables is given by defining a dissipation potential wi for each mechanism.
It is given by:

Nk
!
X D
wi ¼ ð1  f Þ BiH  RiH  ki þ viI I X 2IH ð19Þ
I¼1
2C I

where DI > 0 are material coefficients. The evolution of plastic strain and state variables is then given,
using normality rule, as the partial derivatives of the dissipation potential with respect to the ‘‘forces”:
@wi @BiH @BiH
e_ pi ¼ k_ i ¼ ð1  f Þk_ i ¼ ð1  f Þk_ i ¼ ð1  f Þk_ i ni ð20Þ
@r @r @Bi
@w
p_ i ¼ k_ i i ¼ k_ i ð21Þ
@Ri
 
@w 3 DI 2 @X
a_ I ¼ k_ i i ¼ ð1  f Þk_ i viI ni  vI with vI ¼ X IH IH ð22Þ
@X I 2 CI 3 @X I

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

6 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

ni denotes the flow direction. Applying Euler’s theorem (Eq. (4)) one notices that:
e_ pi : Bi ¼ ð1  f Þp_ i BiH ð23Þ
which expresses the equality between the plastic work at the macroscopic level (left-hand side) and
the microscopic level (right-hand side). The calculation of p_ i is done using the consistency condition
for plastic materials (i.e. /iH ¼ 0 and /_ iH ¼ 0). In the case of viscoplastic materials it is given by some
‘‘creep” function ðXi Þ of /iH . For instance, in the case of a Norton law: Xi ð/iH Þ ¼ h/iH =K i ini where K i and
ni are material parameters.
The dissipation g can then be calculated:
X
N Nk
X
g ¼ r : e_ p  p_ i Ri  a_ I : X I
i¼1 I¼1
X
N X
N
¼r: ð1  f Þp_ i ni  ð1  f Þp_ i RiH
i¼1 i¼1

XNk XN  
3 DI
 ð1  f Þp_ i viI ni  vI : X I
I¼1 i¼1
2 CI
Nk
!
X
N X 3 DI
¼ ð1  f Þ p_ i ðr  X i Þ : ni  RiH þ v vI : X I
i¼1 I¼1
2 iI C I
Nk
!
X X D X
¼ ð1  f Þ p_ i BiH  RiH þ viI I X 2IH ¼ gi ð24Þ
i I¼1
C I i

Noting that BiH  RiH > BiH  RiH  ki ¼ /i P 0 proves that dissipation is always positive. Note also that
dissipation can be expressed as the sum of independent contributions associated with each mecha-
nism gi > 0.

2.5. Void nucleation

Strain controlled nucleation of new voids can be added by modifying Eq. (5) for each mechanism
which then is replaced, following Chu and Needleman (1980), by:

f_ i ¼ ð1  f Þtre_ pi þ Ain p_ i ¼ f_ ig þ f_ in ð25Þ


Ain are material parameters representing the nucleation rate for a given mechanism. fig represents the
porosity due to void growth caused by mechanism i. fin represents damage nucleated at particles due
to mechanism i and corresponds to an apparent porosity (Besson et al., 2000). When nucleation is used
the model can no longer be formulated within the GSM framework at least when using the above sim-
ple nucleation law (Besson and Guillemer-Neel, 2003). Total growth and nucleation porosities (fg and
fn ) are expressed as:

X
N X
N
fg ¼ fig and f n ¼ fin and f ¼ fg þ fn ð26Þ
i¼1 i¼1

fg corresponds to the actual void volume fraction which could be estimated measuring the material
density (Pardoen et al., 1998).

2.6. Numerical implementation and finite element simulations

The model was implemented in the general purpose finite element software Zset (Besson and
Foerch, 1997; Foerch et al., 1997). A fully implicit scheme is used to integrate the state variables
over the load history following the method presented in Besson et al. (2001b). Finite strains are trea-
ted using a corotational reference frame (Sidoroff and Dogui, 2001) defined so that the stress rate
corresponds to the Jauman rate. All calculations are performed using quadratic interpolation

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 7

elements together with reduced integration (i.e. 8 nodes axisymmetric or plane strain elements with
4 Gauss points). Convergence is checked according to the procedure defined in Bron and Besson
(2006).
Models including damage lead to material softening and mesh size dependence when standard fi-
nite element techniques are used. In that case, it is necessary to fix the mesh size in order to obtain
results transferable from one sample to another. This technique is often used in the case of ductile rup-
ture (Rousselier, 1987; Ruggieri and Dodds, 1996; Gullerud et al., 2000; Bron and Besson, 2006)
although improved so called ‘‘non-local” models can be used to obtain mesh independent results
(Mediavilla et al., 2006; Enakoutsa et al., 2007). Consequently, the mesh size and in particular the
mesh height in the direction perpendicular to the crack plane (Siegmund and Brocks, 1999) should
be considered as an adjustable material parameter. In this work a constant mesh size was used in re-
gions where cracks propagate.

3. Application to ductile failure by void growth and coalescence by internal necking

3.1. Thomason model for coalescence

The coalescence model proposed by Thomason (1968, 1985a,b) assumes that coalescence starts
when plastic deformation localizes in the intervoid ligaments. This hypothesis is well supported by
unit cell calculations (Koplik and Needleman, 1988; Brocks et al., 1995). The model was initially devel-
oped for arbitrary ellipsoidal voids whose principal directions correspond to the principal directions of
the macroscopic stress tensor. Following Zhang and Niemi (1995), it will be assumed that voids re-
main spherical. A more detailed description including the effect of void shape can also be used as
in Pardoen and Hutchinson (2000). The material is assumed to be represented by a regular array of
cylindrical cells (height: 2H, diameter: 2L) containing a spherical void (diameter: 2R) as shown in
Fig. 1. The average stress tensor r acting on the cell is assumed to be axisymmetric with rzz > rrr .
The axial equilibrium of the cell is written as:

pL2 rzz ¼ C f pðL2  R2 Þrf ð27Þ


The force in the loading direction expressed in the ligament region (right-hand side) must be equal to
the force applied on the cell boundary (left-hand side). The cavity induces a stress concentration in the
ligament thus increasing the load carrying capacity of the ligament. This effect is described by the
plastic constraint factor C f in the previous equation. In the present case C f is given by (Thomason,
1985a; Benzerga et al., 1999):
pffiffiffiffiffiffiffiffiffi
C f ¼ 0:1ð1=v  1Þ2 þ 1:2 1=v with v ¼ R=L ð28Þ

Fig. 1. Thomason model for spherical voids. Plasticity is localized in the gray region.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

8 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

rf represents the flow stress in the ligament. It should be noted that the original model was developed
for a perfectly plastic matrix but it can be easily extended to work hardening materials like in this
work. Finally, Eq. (27) can be rearranged as:
3
C th rf ¼ rzz with C th ¼ ð1  v2 ÞC f ð29Þ
2
Finally, v must be computed. It can be shown that (Benzerga et al., 1999; Pardoen and Hutchinson,
2003):
  1=3
3 3
v¼ f k0 exp jezz ð30Þ
2 2
where ezz is the axial component of the deviator of the strain tensor. k0 is the initial value of the ratio
H=L which represents the void spacial arrangement. This parameter plays an important role on duc-
tility and toughness (Pardoen and Hutchinson, 2003). j is a fitting parameter used to represent that
the material is, in fact, not made of a perfect assembly of similar unit cells. The original model corre-
sponds to j ¼ 1.

3.2. Extension of Thomason model to arbitrary stress states

Although derived in the case of an axisymmetric cell, the Thomason model can be extended in or-
der to be applied to any stress state. Under axisymmetric conditions (Fig. 1), one gets
rzz ¼ 2=3req þ 1=3rkk . Using the formalism developed in Section 2, the effective scalar quantity is
now defined as:

2 1
SCþ ða; f ; aH Þ ¼ aeq þ akk  C th aH ð31Þ
3 3
For compressive stress states, a symmetric equation is obtained as (Pardoen and Hutchinson,
2003):
2 1
SC ða; f ; aH Þ ¼ aeq  akk  C th aH ð32Þ
3 3
An interpolating equation can be proposed to avoid using two definitions of the effective stress (and
therefore two distinct yield surfaces):

 a  a 1=a
2 1  2 1 
SC ða; f ; aH Þ ¼  aeq þ akk  þ  aeq  akk   C th aH ð33Þ
3 3 3 3

where a is a numerical factor chosen such that a  1 to obtain values for aH close to those obtained
with SCþ and SC .
The evaluation of f is still given by mass conservation (Eq. (5)). In order to compute v (Eq. (30)) it is
now necessary to evaluate the deformation along the cell axis of symmetry. Assuming radial loading
ezz can be computed as the maximum principal value of the strain tensor deviator. In addition when
porosity remains small, one gets: 3=2ezz ’ p. This approximation will be used in the following as it
simplifies the implementation and gives a realistic description of the decrease of the intervoid liga-
ment leading to failure. A material failure condition can also be obtained from the model. Failure oc-
curs when the intervoid ligament becomes very small, that is when v ¼ 1. In practice the material will
be assumed broken when v ¼ vc where vc is a critical value close to 1 which is used to ease the cal-
culations (in the following vc ¼ 0:95 will be used).
The ratio of volume to shear deformation is given by the ratio: se ¼ ð@ rH =@ rkk Þ=ð@ rH =@ req Þ. In the
case of the extension of the Thomason model, this ratio is constant and equal to 1/2. This corresponds
to a deformation state where deformation occurs along one direction with no deformation along the
perpendicular directions. This is fully consistent with the results of unit cell calculations in the coales-
cence regime (Brocks et al., 1995).

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 9

3.3. Full model for void growth and coalescence

The deformation of the porous material results from the competition between two mechanisms
(Pardoen and Hutchinson, 2000, 2003): (i) the deformation is homogeneous and the material behavior
is governed by the Gurson–Tvergaard–Needleman (GTN) model (Tvergaard and Needleman, 1984), (ii)
the deformation is localized between intervoid ligaments and the material is governed by the ex-
tended Thomason model. The transition between both modes is obtained using a two surface model.
This implementation of the model allows to have both modes active at the same time and to return to
the void growth mode after the localized mode has been activated. Previous implementations of the
model (Pardoen and Hutchinson, 2000, 2003; Benzerga et al., 1999) assumed that only one mode
could be activated at the same type and that the localized mode could not be deactivated once trig-
gered. Note that two deformation modes are present but that only one deformation mechanism is ac-
tive (dislocation glide). This implies that Hpij ¼ 1. Following the proposed framework, the GTN model
corresponds to the definition of an effective stress given by:
 
a2eq q2 akk
SGTN ða; f ; aH Þ ¼ þ 2q1 fH cosh  1  q21 fH2 ð34Þ
a2H 2 aH
q1 and q2 are parameters accounting for void interactions (Tvergaard and Needleman, 1984; Faleskog
et al., 1998). fH is a function of the porosity introduced on a purely phenomenological basis in Tverg-
aard and Needleman (1984) to represent void coalescence. It is often expressed as (fR : porosity at fail-
ure, fc : porosity at the onset of coalescence):
(
f if f < fc
fH ¼  ð35Þ
1 f fc
fc þ q1
 fc fR fc
otherwise

As coalescence is now fully described it is no longer necessary to use and adjust the fH function in the
GTN model. The model allows to use fH ¼ f and to obtain realistic ductilities. Note, in addition, that the
use of the fH in the GTN model does not modify the shape of the yield function so that the specific
deformation state characteristic of coalescence (i.e. pure uniaxial extension along a single direction)
is not reproduced.
Fig. 2 compares the different yield surfaces corresponding to Eqs. (31)–(34) in the akk =aH –aeq =aH
plane for material parameters corresponding to a highly damaged volume element. The GTN model

Fig. 2. Yield surface combining the GTN model and the Thomason model in the akk =aH —aeq =aH plane. The yield surfaces
corresponding to Eqs. (31)–(34) are indicated. The gray region indicate the elastic domain in the case where the GTN model is
used together with the approximate formulation for the coalescence model (Eq. (33)) (parameter values used in this example:
C th ¼ 0:8; a ¼ 10; q1 ¼ 1:5; q2 ¼ 1; f ¼ 0:1).

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

10 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

Table 1
Model parameters used for the simulation of tensile bars using the void growth and coalescence ductile failure model.

Young’s modulus 200 GPa


Poisson’s ratio 0.3
Plasticity k ¼ 530 MPa
RðpÞ ¼ 190ð1  expð36pÞÞ þ 81p MPa
Damage f0 ¼ 1:5  104
GTN Thomason
q1 ¼ 1:5 vc ¼ 0:95
q2 ¼ 1:1 k0 ¼ 1
fH ¼ f a ¼ 50
Interaction Hpij ¼ 1; Hfij ¼ 1

prevails at low stress triaxiality ratios and the Thomason model at high triaxialities. At the intersection
of both surfaces both mechanisms are active and the flow direction lies in a cone. The approximate
coalescence model (Eq. (33)) allows to avoid the vertex on the pressure axis which is present for
the original model (Eqs. (31) and (32)) thus simplifying the numerical treatment of the vertex (Lorentz
et al., 2008). Note that the vertex is present in the case of the Rousselier (1991) model which can rep-
resent the fact that the deformation always keeps a shear component during coalescence (Besson
et al., 2001b) although it was not originally designed to model this feature. Using a high value for
parameter a allows to stay very close to the original yield surface. a ¼ 50 is used in the following.

3.4. Application

The material used for the application is a low alloyed modern steel (Rivalin et al., 2000). Due to its
very low sulfur and phosphorus content, the inclusion volume fraction is low and equal to 1:5  104 .
Immediate debonding between the inclusions and the matrix will be assumed so that the inclusion
content corresponds to the initial porosity ðf0 Þ of the material. Its hardening behavior is phenomeno-
logically represented in tension by: RðpÞ ¼ 190ð1  expð36pÞÞ þ 81pðMPaÞ and k ¼ 530 MPa. Harden-
ing is assumed to be purely isotropic. The GTN model is used to model the void growth stage with
q1 ¼ 1:5: q2 was adjusted to obtained the experimental ductility on smooth and notched tensile bars
ðq2 ¼ 1:1Þ. The void spacing parameter k0 was taken equal to 1.0 which is a realistic value for a highly
homogeneous material. All material parameters used for the simulation are gathered in Table 1.

Fig. 3. Load (F: force, S0 : initial minimum cross section) vs. diameter reduction for j ¼ 1 and j ¼ 0 (i.e. GTN model). Contour
maps indicate damage ðf Þ for various values of j. A deformed mesh showing cup-cone fracture is also shown corresponding to
the GTN model.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 11

The model is applied to the simulation of necking and cup-cone fracture in a smooth tensile bar.
Simulation techniques are similar to those used in Besson et al. (2001b). Results of the simulations
are shown in Fig. 3 for several values of the fitting parameter j. For j ¼ 0, the deformation mode cor-
responding to internal necking is never activated so that the results coincide with those obtained
using the ‘‘standard” GTN model. In that case cup-cone fracture is obtained. For large values of j
and in particular for the original Thomason model, cup-cone fracture is inhibited. This indicates that
the change from one deformation mode to another leads to highly localized damage which inhibits the
formation of a damaged band over several elements (Besson et al., 2001b, 2003). This does not imply
that the complete model coupling the GTN and Thomason models prohibits cup-cone formation as the
softening character leads to a strong mesh size dependency. A ‘‘non-local” model (Lorentz and Andri-
eux, 1999; Mediavilla et al., 2006) which eliminates mesh size dependency should be used to precisely
evaluate the role of the constitutive equations on cup-cone formation and more generally on crack
path deviations. However, the present results indicate that cup-cone formation is more difficult using
the full (GTN+Thomason) model.

4. Application to creep failure

High temperature creep in metals and ceramics is often characterized by different regimes depend-
ing on the stress level: e.g. dislocation creep and diffusional creep (Frost and Ashby, 1982). In that case,
a viscoplastic model incorporating at least two independent deformation mechanisms is required to
describe the material behavior (Contesti and Cailletaud, 1989). The proposed framework allows to
straightforwardly incorporate damage in this model family. Note that most of the literature work de-
voted to the modeling of damage development during creep is based on CDM (see, e.g. Othman et al.,
1994; Perrin and Hayhurst, 1999).
In the case of viscous materials, void growth is poorly described by the Gurson model which was
initially developed for rate independent materials (Budianski et al., 1982). A semi-phenomenological
model was proposed by Leblond et al. (1994) which allows to propose expressions for the macroscopic
viscoplastic potentials of materials containing cavities for matrix materials with a strain rate sensitiv-
ity between 1 (linear viscous solids) and 0 (rate independent plasticity). Based on this description the
effective stress, rH , is defined by the following equation:
 
r2eq 1 rkk 1M 2
SLPS ¼ þ q 1 f H jM q  1  q21 f ð36Þ
r2H 2 2 rH 1þM H
with
1M 1
jM ðXÞ ¼ hM ðxÞ þ and hM ðXÞ ¼ ð1 þ MX 1þM Þ1=M ð37Þ
1 þ M hM ðxÞ
where M is the strain rate sensitivity.1 Two limiting cases are of particular interest. For M ! 0 the GTN
model is retrieved as: limM!0 jM ðxÞ ¼ 2 coshðxÞ. For M ! 1: jM ðxÞ ¼ 1 þ x2 so that Eq. (36) corresponds
to the ‘‘classical” elliptic model (Shima and Oyane, 1976). fH is a function of f used to reproduce coa-
lescence (Eq. (35)).
The model was used by Gaffard et al. (2005) to simulate the lifetime of a tempered martensitic
stainless steel (type 9Cr1Mo-NbV (P/T91)). Model parameters were identified using smooth and
notched creep specimens tested at 625 °C under various load levels. Elongations as a function of time,
times to failure but also locations of failure initiation were used to fit the model parameters. Details of
the procedure are given in Gaffard et al. (2005). A simplified version of the model is used in the present
study to illustrate the capacity of the model. Two deformation mechanisms are considered: disloca-
tion creep and diffusional creep. For each mechanism a Norton law is used:
p_ m ¼ e_ 0m ðr=r0 Þnm ð38Þ

1
For a Norton like creep law expressed as p_ ¼ e_ 0 ðreq =r0 Þn ; M ¼ 1=n:e_ 0 and r0 are material parameters.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

12 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

Fig. 4. Application of the model in the case of creep rupture of a P91 martensitic steel: (a) rupture time as a function of the
applied mean stress for each individual mechanism (dashed lines) and for the full model incorporating both diffusional and
dislocation creep (symbols). (b) Rupture strain as a function of the applied mean stress. (c) Rupture location as a function of the
applied mean stress.

where e_ 0m ; r0 and nm are material parameters. Due to the high concentration of reinforcing particles in
the material (see, e.g. Eggeler et al., 1989), nucleation of new voids is the predominant damage mech-
anism. Each deformation mechanism is assumed to contribute to nucleation. Nucleation rate is related
to the deformation rate but also depends on the stress state. It is expressed for each mechanism m as:
 
fnm ¼ Am m am
n þ Bn s p_ m ð39Þ
where Am m
n ; Bn and am are material parameters. s is the stress triaxiality ratio. GTN model parameters q1
and q2 were chosen equal to classically used values. No fH function was used in the case of the diffu-
sional creep mechanism as a creep exponent equal to 1 provides a high stability which prevents void
coalescence (Ashby et al., 1979). For the viscoplastic regime mechanism, the apparent porosity fH is
defined by Eq. (35).
Fig. 4 illustrates the use of the model in the case of an axisymmetric notched bar (minimum diam-
eter: R ¼ 10 mm, notch radius: 2 mm). Model parameters used for the simulations are gathered in Ta-
ble 2. Fig. 4a gives the evolution of the time to failure as a function of the applied mean stress (i.e. force
divided by the initial cross section). Failure is defined as the time at which a single material point has
reached the failure condition (i.e. fH ¼ 1=q1 ). Dashed lines indicate values of the time to failure
assuming that only one mechanism is active and symbols show the time to failure for the full model.
At low stresses (i.e. <70 MPa) failure is dominated by diffusional creep and for high (i.e. >200 MPa) dis-
location creep controls failure. For intermediate stresses, both deformation mechanisms play a role on
failure.
Fig. 4b gives the evolution of the rupture stain as a function of the applied mean stress. In actual
tests, failure occurs for slightly longer times and the strain to failure is larger as high stress deforma-
tion mechanisms may be triggered during final failure (i.e. dislocation creep and eventually plasticity).
It is shown that strain to failure decreases with decreasing applied stress (or increasing time to failure)
as it is commonly reported in the literature (see, e.g. Wilshire and Scharning, 2007). The decrease of
ductility with decreasing applied stress is directly related to the high values of Am and Bm for the dif-
fusional creep. Using one single deformation mechanism leads to a constant value for the rupture
strain.
Fig. 4c gives the evolution of the rupture initiation location ðr=RÞ as a function of the applied mean
stress. At low stresses, the apparent creep exponent n2 is close to 1 so that stress redistribution does not
occur: failure takes place at the free surface of the minimum cross section. As the applied stress in-
creases, n increases and stress redistribution occurs so that the failure initiation site moves towards

2
_
Defined as: n ¼ @ logðpÞ=@ logðrÞ.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 13

Table 2
Model parameters for the simulation of creep failure based on two deformation mechanisms (dij : Kronecker delta).

Dislocation creep Diffusional creep


Young’s modulus 145 GPa
Poisson’s ratio 0.3
1 1
e_ 0 4:8  106 h 3:0  107 h
n 8.1 1.
r0 100 MPa
q1 1.5
q2 1
Am 0.01 12
Bm 0.15 15
am 2
fH fc ¼ 0:1; fR ¼ 0:2 fH ¼ f
Interaction Hpij ¼ dij ; Hfij ¼ 1

Table 3
Parameters of the GTN model to represent failure of the 2024-T351 aluminum alloy under compression–tension loading.

Elastic properties: Young’s modulus*:70 GPa, Poisson’s ratio*:0.3


Plastic behavior
k ðMPaÞ Q 1 ðMPaÞ b1 Q 2 ðMPaÞ b2 C 1 ðMPaÞ D1 C 2 ðMPaÞ D2
304 202 6.4 11 56 436 14.4 2858 49
Damage behavior
f0 * fc fR * A0n pn * q1 * q2 *
0 0.035 0.2 0.18 0.15 1.5 1
*
Indicates parameters which were fixed a priori.

the center of the minimum cross section. This trend was experimentally evidenced in Gaffard et al.
(2005).

5. Application to ductile failure after prestraining

Metal sheets are often prestrained during the last stages of processing and during forming opera-
tions. This is the case of aluminum sheets used to manufacture aircraft fuselage or of steel plates used
to produce pipeline elements with the UOE process (Kostryzhev et al., 2007). Prestrain both hardens
and damages materials so that fracture properties (e.g. ductility and toughness) of the final product
may differ from those of the unstrained material (Enami, 2005a,b). As hardening is often both isotropic
and kinematic, prestrain can induce anisotropic hardening properties in addition to the texture related
plastic anisotropy. The proposed model is applied to experiments published (specimen A, Fig. 5) by
Bao and Treitler (2004). Tests are performed on aluminum (2024-T351) notched bars (minimum
diameter: 12 mm, notch radius: 12 mm) which are first prestrained in compression using various im-
posed displacements over a gauge length equal to 25.4 mm. The material contains a rather high con-
centration of brittle second phase particles at which micro-cracks are nucleated under both tension
and compression. The GTN model is used here; it is slightly modified to enforce that fg P 0 under com-
pression. The effective stress is given by:
 
a2eq 1 akk
SGTNc ða; f ; aH Þ ¼ þ 2q f
1 H cosh g q  1  q21 fH2 ¼ 0 ð40Þ
a2H 2 2 aH
where parameter g is defined as:

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

14 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

Fig. 5. Experimental (solid lines) and simulated (symbols) load–displacement curves for specimen type A used in Bao and
Treitler (2004).

(
1 if f g > 0 or akk P 0
g¼ ð41Þ
0 if f g ¼ 0 and akk < 0

Note that this modification is not needed in absence of damage nucleation as in that case f_ g ¼ 0 for
f ¼ fg ¼ 0. Experimental data were fitted using the following expression for isotropic hardening:
RðpÞ ¼ Q 1 ð1  expðb1 pÞÞ þ Q 2 ð1  expðb2 pÞÞ ð42Þ
Two kinematic hardening variables were used corresponding to parameters ðC 1 ; D1 Þ and ðC 2 ; D2 Þ. A
large number of fitting parameters (8) was used to represent hardening to obtain a good agreement
with experiments for both isotropic and kinematic hardening. In addition, the example also outlines
the possibility of using multiple kinematic hardening variables. Fitting of hardening parameters was
performed using an automatic procedure, which used the Nelder–Mead algorithm, to minimize the
quadratic difference between the experimental and simulated load–displacement curves. Damage
nucleation and growth was neglected at this stage. Standard values for q1 ; q2 and fR were used. It
was assumed that all second phase particles were cracked at the necking strain (about 15%). Conse-
quently nucleation parameter An is expressed as:
(
A0n if p < pn
An ¼ ð43Þ
0 otherwise

Finally fc and A0n were tuned to represent experimental data for the test carried out in pure tension
only. Model parameters are gathered in Table 3. Parameters were then validated by simulating pre-
strained specimens. Results of the simulations are shown in Fig. 5 showing a good agreement be-
tween experimental data and simulation. In particular, the displacement at which a macro-crack
is initiated in the specimen (corresponding to a sharp load drop) is well represented for each pre-
strain level.
In the present application, the model is used to represent prestraining only. It could however be
applied to cyclic loading as in Pirondi et al. (2006) where cyclic tests up to failure were also con-
ducted on notched bars for a number of cycles less than 100. The authors tried to simulate the
experiments using either the Leblond et al. (1995) model which is suitable for ductile rupture
and a CDM based model (Pirondi and Bonora, 2003) which can be adapted for fatigue failure. None
of both models could satisfactorily represent all test results which probably indicates that failure re-
sults from both ductile and fatigue damage. Cyclic testing with more than 1000 cycles can be de-
scribed phenomenologically using CDM as in Kang et al. (2009); the present model is clearly not
adapted to this case.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 15

6. Summary and conclusions

In this work, micromechanical models representing ductile failure have been extended to the case
of materials in which several mechanisms are responsible for deformation. Mixed isotropic/kinematic
hardening is also considered. The model representing the undamaged material is described within the
Generalized Standard Material framework (Halphen and Nguyen, 1975; Chaboche, 2008). It is shown
that the model including ductile damage is also described within the GSM framework at least in ab-
sence of damage nucleation. The model relies on the definition of effective scalar stresses representing
the effect of the macroscopic stresses and of damage on the matrix load. Note that using this concept
several existing models, such as the well known GTN model, can be described within the same general
framework.
Three applications are given to illustrate the use of the model. The first one deals with ductile
failure by void growth and void coalescence by internal necking. Void growth corresponds to homo-
geneous deformation of the matrix material and internal necking to a localized mode. Each mode
corresponds to a yield surface. Indeed matrix deformation mechanisms are similar in both cases.
The second example deals with creep failure where deformation is controlled by a high stress
mechanism (dislocation climb) and a low stress mechanism (diffusional creep). In the third exam-
ple, only one deformation mechanism is used but mixed isotropic/kinematic hardening is intro-
duced to represent failure after pre-compression. Note that the proposed general model allows to
deal with several mechanisms and mixed hardening. Although the model can be applied to cyclic
loading, material degradation still corresponds to ductile mechanisms so that it should not be ap-
plied to actual fatigue failure for which specific models exist (Kang et al., 2009; Lemaitre and Desm-
orat, 2005). The model could also be applied to glassy and semi-crystalline polymers in which back
stresses develop together with voids during straining (Steenbrink et al., 1997; Steenbrink and Van
der Giessen, 1997). Recent studies on polymeric materials (Challier et al., 2006; Zaïri et al., 2008), as
they deal only with monotonic loading, do not account for kinematic hardening although it can lead
to earlier flow localization (Mear and Hutchinson, 1985). Note that in that case, Eq. (22) should be
modified to account for the strong hardening observed in glassy polymers following Steenbrink
et al. (1997).
Interesting developments of the model include the introduction of anisotropy which may be caused
by both anisotropic plasticity of the matrix and anisotropic void shape and distribution. Anisotropy
may result from anisotropic void shape (Gologanu et al., 1997; Kailasam et al., 2000) and spacing (Par-
doen and Hutchinson, 2000). Including void shape effects in the present framework is possible
although it would represent an important work. The simplest method consist in using a single void
shape factor for all deformation mechanisms. Its evolution should then be given by the macroscopic
inelastic strain rate tensor. Due to specific crystallographic texture, the matrix material may also have
anisotropic plastic properties which can be represented using phenomenological yield surfaces such
as those proposed in Hill (1950); Barlat et al. (1991); Karafillis and Boyce (1993); Bron and Besson
(2004). These anisotropic plasticity models can easily be introduced in the present formalism by
replacing the von Mises invariant in Eq. (2) by any of anisotropic stress measure. This was already
done in the case of a single deformation mechanism in Grange et al. (2000); Rivalin et al. (2000); Bru-
net et al. (2005); Monchiet et al. (2008) with the Hill criterion and in Bron and Besson (2006) with a
more complex yield criterion (Bron and Besson, 2004).

Appendix A. Derivation of the damage model with kinematic hardening

In this section, a summary of the derivation of the damage model with kinematic hardening pro-
posed in Besson and Guillemer-Neel (2003) is provided. A single kinematic hardening variable ðaÞ is
used but the extension to several variables is straightforward. The model is first derived for the Green
(1972) model which allows to obtain explicit expressions. The effective stress is expressed as:
 1=2
3
rH ¼ r:M:r ðA:1Þ
2

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

16 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

where M is a fourth order tensor. It can be shown using Eq. (23) (which in the initial model is postu-
lated and not the result of the model itself as for the proposed framework) that the effective plastic
strain rate p_ is related to the plastic strain rate tensor by:
 1=2
1 2
p_ ¼ e_ p : M1 : e_ p ðA:2Þ
1f 3
The term ðWa Þ corresponding the kinematic hardening in the expression of the free energy for the
undamaged material is expressed as: Wa ¼ 13 C a : a. It may be rewritten as: Wa ¼ 12 C a2e where
ae ¼ ð23 a : aÞ1=2 : ae is the von Mises invariant associated to a (which corresponds to a strain tensor).
In the case of the porous material, it is proposed to use Eq. (A.2) to relate a to the effective measure
ae . The resulting expression for Wa is then 12 ð1  f ÞC a2e ; the ð1  f Þ factor is due to the fact that ae is
interpreted (as p) as a measure of the matrix strain. Wa is finally expressed as:
1 C
Wa ¼ a : M1 : a ðA:3Þ
3 1f
In the more general case (e.g. in the case of the GTN model) there is no explicit link between the plastic
strain rate tensor and p (Eq. (A.2)) so that the previous definitions must be adapted. An intermediate
tensor v is introduced as:
2 C
v¼ a ðA:4Þ
3 1f
In the case of the Green model, is can be shown that:

1 @X 2H
v¼ ðA:5Þ
3 @X
This equation is assumed to hold in the general case so that X is numerically found solving the follow-
ing equation:

2 C 1 @X 2H
a¼ ðA:6Þ
3 1f 3 @X
It is however important to prove the existence of the free energy term Wa associated to a as for the
Green model. This is done as follows. One can note that there exists a scalar function GX of X such that:
a ¼ @GX =@X with GX ¼ ð1  f ÞX 2H =ð2CÞ The free energy Wa corresponding to a is then given as the
Legendre transform of GX :
Wa ¼ max ða : x  GX ðxÞÞ ðA:7Þ
x

To derive the dissipation potential ðwÞ, the von Mises invariant used for the undamaged model is sim-
ply replaced by the effective back stress X H (Eq. (19)).

References

Arndt, S., Svendsen, B., Klingbeil, D., 1997. Modellierung der Eigenspannungen and der Rißspitze mit einem Schägigungsmodell.
Technische Mechanik 17 (4), 323–332.
Ashby, M.F., Gandhi, C., Taplin, D.M.R., 1979. Fracture mechanism maps and their construction for fcc metals and alloys. Acta
Metall. 27, 699–729.
Bao, Y., Treitler, R., 2004. Ductile crack formation on notched al2024-t351 bars under compression tension loading. Mater. Sci.
Eng. A 384, 385–394.
Barlat, F., Lege, D.J., Brem, J.C., 1991. A six-component yield function for anisotropic materials. Int. J. Plasticity 7, 693–712.
Benzerga, A., Besson, J., Pineau, A., 1999. Coalescence-controlled anisotropic ductile fracture. J. Eng. Mater. Technol. 121, 121–
229.
Berdin, C., Besson, J., Bugat, S., Desmorat, R., Feyel, F., Forest, S., Lorentz, E., Maire, E., Pardoen, T., Pineau, A., Tanguy, B., 2004.
Local Approach to Fracture. Presses de l’Ecole des Mines, Paris.
Besson, J., Cailletaud, G., Chaboche, J.-L., Forest, S., 2001a. Mécanique non-linéaire des matériaux. Hermes.
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. Eng. Fract. Mech. 67 (2), 169–190.
Besson, J., Foerch, R., 1997. Large scale object-oriented finite element code design. Comp. Meth. Appl. Mech. Eng. 142, 165–187.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx 17

Besson, J., Guillemer-Neel, C., 2003. An extension of the Green and Gurson models to kinematic hardening. Mech. Mater. 35, 1–
18.
Besson, J., Steglich, D., Brocks, W., 2001b. Modeling of crack growth in round bars and plane strain specimens. Int. J. Solids Struct.
38 (46–47), 8259–8284.
Besson, J., Steglich, D., Brocks, W., 2003. Modeling of plane strain ductile rupture. Int. J. Plasticity 19 (10), 1517–1541.
Bordreuil, C., Boyer, J.C., Sallé, E., 2003. On modelling the growth and the orientation changes of ellipsoidal voids in a rigid plastic
matrix. Model. Simul. Mater. Sci. Eng. 11 (3), 365–380.
Brocks, W., Sun, D.Z., Hönig, A., 1995. Verification of the transferability of micromechanical parameters by cell model
calculations with visco-plastic materials. Int. J. Plasticity 11, 971–989.
Bron, F., Besson, J., 2004. A yield function for anisotropic materials. Application to aluminium alloys. Int. J. Plasticity 20, 937–
963.
Bron, F., Besson, J., 2006. Simulation of the ductile tearing for two grades of 2024 aluminum alloy thin sheets. Eng. Fract. Mech.
73, 1531–1552.
Brunet, M., Morestin, F., Walter-Leberre, H., 2005. Failure analysis of anisotropic sheet-metals using a non-local plastic damage
model. J. Mater. Process. Technol. 170, 457–470.
Brünig, M., 2003. An anisotropic ductile damage model based on irreversible thermodynamics. Int. J. Plasticity 19 (10), 1679–
1713.
Brünig, M., Chyra, O., Albrecht, D., Driemeier, L., Alves, M., 2008. A ductile damage criterion at various stress triaxialities. Int. J.
Plasticity 24, 1731–1755.
Budianski, 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, pp. 13–45.
Celentano, D.J., Chaboche, J.L., 2007. Experimental and numerical characterization of damage evolution in steels. Int. J. Plasticity
23, 1739–1762.
Chaboche, J.L., 2008. A review of some plasticity and viscoplasticity constitutive theories. Int. J. Plasticity 24, 1642–1693.
Chaboche, J.L., Boudifa, M., Saanouni, K., 2006. A CDM approach of ductile damage with plastic compressibility. Int. J. Fracture
137 (1–4), 51–75.
Challier, M., Besson, J., Laiarinandrasana, L., Piques, R., 2006. Damage and fracture of PVDF at 20 °C. Eng. Fract. Mech. 73, 79–90.
Chu, C.C., Needleman, A., 1980. Void nucleation effects in biaxially stretched sheets. J. Eng. Mater. Technol. 102, 249–256.
Contesti, E., Cailletaud, G., 1989. Description of creep–plasticity interaction with non-unified constitutive equations. Nucl. Eng.
Des. 116, 265.
Dotta, F., Ruggieri, C., 2004. Structural integrity assessments of high pressure pipelines with axial flaws using a micromechanics
model. Int. J. Pressure Vessels Piping 81, 761–770.
Eggeler, G., Earthman, J.C., Nilsvang, N., Ilschner, B., 1989. Microstructural study of creep rupture in a 12% chromium ferritic
steel. Acta Metall. 37 (1), 49–60.
Enakoutsa, K., Leblond, J.B., Perrin, G., 2007. Numerical implementation and assessment of a phenomenological nonlocal model
of ductile rupture. Comp. Meth. Appl. Mech. Eng. 196 (13–16), 1946–1957.
Enami, K., 2005a. Ductile crack initiation behavior in steels with compressive prestrain. J. Mar. Sci. Technol. 10, 10–41.
Enami, K., 2005b. The effects of compressive and tensile prestrain on ductile fracture initiation in steels. Eng. Fract. Mech. 72 (7),
1089–1105.
Faleskog, J., Gao, X., Shih, C.F., 1998. Cell model for nonlinear fracture analysis – I. Micromechanics calibration. Int. J. Fracture 89,
355–373.
Foerch, R., Besson, J., Cailletaud, G., Pilvin, P., 1997. Polymorphic constitutive equations in finite element codes. Comp. Meth.
Appl. Mech. Eng. 141, 355–372.
Frost, H.J., Ashby, M.F., 1982. Deformation-mechanism Maps: the Plasticity and Creep of Metals and Ceramics. Pergamon Press,
Oxford.
Gaffard, V., Besson, J., Gourgues-Lorenzon, A.F., 2005. Creep failure model of a tempered martensitic stainless steel integrating
multiple deformation and damage mechanisms. Int. J. Fracture 133 (2), 139–166.
Gologanu, M., Leblond, J.B., Devaux, J., 1993. Approximate models for ductile metals containing non-spherical voids – case of
axisymmetric prolate ellipsoidal cavities. J. Mech. Phys. Solids 41 (11), 1723–1754.
Gologanu, M., Leblond, J.B., Devaux, J., 1994. Approximate models for ductile metals containing non-spherical voids – case of
axisymmetric oblate ellipsoidal cavities. J. Eng. Mater. Technol. 116, 290–297.
Gologanu, M., Leblond, J.B., Perrin, G., Devaux, J., 1997. Recent extensions of Gurson’s model for porous ductile metals. In:
Suquet, P. (Ed.), Continuum Micromechanics, CISM Lectures Series. Springer, New York.
Grange, M., Besson, J., Andrieu, E., 2000. An anisotropic Gurson model to represent the ductile rupture of hybrided Zircaloy-4
sheets. Int. J. Fracture 105 (3), 273–293.
Green, R.J., 1972. A plasticity theory for porous solids. Int. J. Mech. Sci. 14, 215–224.
Gullerud, A.S., Gao, X., Dodds Jr., R.H., Haj-Ali, R., 2000. Simulation of ductile crack growth using computational cells: numerical
aspects. Eng. Fract. Mech. 66, 65–92.
Gurson, A.L., 1977. Continuum theory of ductile rupture by void nucleation and growth. Part I – Yield criteria and flow rules for
porous ductile media. J. Eng. Mater. Technol. 99, 2–15.
Halphen, B., Nguyen, Q.S., 1975. Sur les matriaux standards généralisés. J. Mécanique 14 (1), 39–63.
Hammi, Y., Horstemeyer, M.F., 2007. A physically motivated anisotropic tensorial representation of damage with separate
functions for void nucleation growth and coalescence. Int. J. Plasticity 23, 1641–1678.
Hill, R., 1950. The Mathematical Theory of Plasticity. Clarendon Press, Oxford.
Kachanov, M., 1994. Elastic solids with many cracks and related problems. Adv. Appl. Mech. 30, 259–445.
Kailasam, M., Aravas, N., Ponte Castañeda, P., 2000. Porous metals with developing anisotropy: constitutive models,
computational issues and applications to deformation processing. Comput. Model. Eng. Sci. 1 (2), 105–118.
Kang, G., Liu, Y., Ding, J., Gao, Q., 2009. Uniaxial ratcheting and fatigue failure of tempered 42CrMo steel: damage evolution and
damage-coupled visco-plastic constitutive model. Int. J. Plasticity 25 (5), 838–860.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001
ARTICLE IN PRESS

18 J. Besson / International Journal of Plasticity xxx (2009) xxx–xxx

Karafillis, A.P., Boyce, M.C., 1993. A general anisotropic yield criterion using bounds and a transformation weighting tensor. J.
Mech. Phys. Solids 41, 1859–1886.
Koplik, J., Needleman, A., 1988. Void growth and coalescence in porous plastic solids. Int. J. Solids Struct. 24 (8), 835–853.
Kostryzhev, A.G., Strangwood, M., Davis, C.L., 2007. Influence of microalloying precipitates on Bauschinger effect during UOE
forming of line pipe steels. Mater. Technol. 22 (3), 166–172.
Leblond, J.B., Perrin, G., Devaux, J., 1995. An improved Gurson-type model for hardenable ductile metals. Eur. J. Mech. 14A (4),
499–527.
Leblond, J.B., Perrin, G., Suquet, P., 1994. Exact results and approximate models for porous viscoplastic solids. Int. J. Plasticity 10
(3), 213–235.
Lemaitre, J., 1985. A continuous damage mechanics model for ductile fracture. J. Eng. Mater. Technol. 107, 83–89.
Lemaitre, J., 1996. A Course on Damage Mechanics. Springer Verlag, Berlin.
Lemaitre, J., Chaboche, J.L., 1990. Mechanics of Solid Materials. Cambridge University Press, Cambridge, UK.
Lemaitre, J., Desmorat, R., 2005. Engineering Damage Mechanics. Springer.
Lorentz, E., Andrieux, S., 1999. A variational formulation for nonlocal damage models. Int. J. Plasticity 15 (2), 119–138.
Lorentz, E., Besson, J., Cano, V., 2008. Numerical simulation of ductile fracture with the Rousselier constitutive law. Comp. Meth.
Appl. Mech. Eng. 197 (21–24), 1965–1982.
Mear, M.E., Hutchinson, J.W., 1985. Influence of yield surface curvature on flow localization in dilatant plasticity. Mech. Mater. 4,
395–407.
Mediavilla, J., Peerlings, R.H.J., Geers, M.G.D., 2006. Discrete crack modelling of ductile fracture driven by non-local softening
plasticity. Int. J. Numer. Meth. Eng. 66 (4), 661–688.
Monchiet, V., Cazacu, O., Charkaluk, E., Kondo, D., 2008. Macroscopic yield criteria for plastic anisotropic materials containing
spheroidal voids. Int. J. Plasticity 24, 1158–1189.
Nègre, P., Steglich, D., Brocks, W., 2004. Crack extension in aluminium welds: a numerical approach using the Gurson–
Tvergaard–Needleman model. Eng. Fract. Mech. 71, 2365–2383.
Othman, A.M., Dyson, B.F., Hayhurst, D.R., Lin, J., 1994. Continuum damage mechanics modelling of circumferentially notched
tension bars undergoing tertiary creep with physically based constitutive equations. Acta Metall. Mater. 42 (3), 597–611.
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 Mater. 46 (2), 541–552.
Pardoen, T., Hutchinson, J.W., 2000. An extended model for void growth and coalescence. J. Mech. Phys. Solids 48 (12), 2467–
2512.
Pardoen, T., Hutchinson, J.W., 2003. Micromechanics-based model for trends in toughness of ductile metals. Acta Mater. 51,
133–148.
Perrin, I.J., Hayhurst, D.R., 1999. Continuum damage mechanics analyses of type IV creep failure in ferritic steel crossweld
specimens. Int. J. Pressure Vessels Piping 76 (9), 599–617.
Pineau, A., 2006. Development of the local approach to fracture over the past 25 years: theory and applications. Int. J. Fracture
138 (1–4), 139–166.
Pirondi, A., Bonora, N., 2003. Modeling ductile damage under fully reversed cycling. Comput. Mat. Sci. 26, 129–141.
Pirondi, A., Bonora, N., Steglich, D., Brocks, W., Hellmann, D., 2006. Simulation of failure under cyclic plastic loading by damage
models. Int. J. Plasticity 22, 2146–2170.
Rivalin, F., Besson, J., Di Fant, M., Pineau, A., 2000. Ductile tearing of pipeline-steel wide plates – II. Modeling of in-plane crack
propagation. Eng. Fract. Mech. 68 (3), 347–364.
Rousselier, G., 1987. Ductile fracture models and their potential in local approach of fracture. Nucl. Eng. Des. 105, 97–111.
Rousselier, G., 1991. Application de l’analyse de stabilité d’une perturbation à la localisation de la déformation dans un matériau
dilatable adoucissant. C. R. Acad. Sci. Paris 313 (Série II), 1367–1373.
Ruggieri, C., Dodds, R.H., 1996. A transferability model for brittle fracture including constraint and ductile tearing effects: a
probabilistic approach. Int. J. Fracture 79, 309–340.
Shima, S., Oyane, M., 1976. Plasticity theory for porous metals. Int. J. Mech. Sci. 18, 285–291.
Sidoroff, F., Dogui, A., 2001. Some issues about anisotropic elastic–plastic models at finite strain. Int. J. Solids Struct. 38, 9569–
9578.
Siegmund, T., Brocks, W., 1999. Prediction of the work of separation and implications to modelling. Int. J. Fracture 99, 97–116.
Steenbrink, A.C., Van der Giessen, E., 1997. Void growth in glassy polymers: effect of yield properties on hydrostatic expansion.
Int. J. Damage Mech. 7 (7), 317–330.
Steenbrink, A.C., Van der Giessen, E., Wu, P.D., 1997. Void growth in glassy polymers. J. Mech. Phys. Solids 45 (3), 405–437.
Tanguy, B., Besson, J., 2002. An extension of the Rousselier model to viscoplastic temperature dependent materials. Int. J.
Fracture 116 (1), 81–101.
Thomason, P.F., 1990. Ductile Fracture of Metals. Pergamon Press, Oxford.
Thomason, P.F., 1968. A theory for ductile fracture by internal necking of cavities. J. Inst. Metals 96, 360–365.
Thomason, P.F., 1985a. A three-dimensional model for ductile fracture by the growth and coalescence of microvoids. Acta
Metall. 33 (6), 1087–1095.
Thomason, P.F., 1985b. Three-dimensional models for the plastic limit – loads at incipient failure of the intervoid matrix in
ductile porous solids. Acta Metall. 33 (6), 1079–1085.
Tvergaard, V., 1990. Material failure by void growth to coalescence. Adv. Appl. Mech. 27, 83–151.
Tvergaard, V., Needleman, A., 1984. Analysis of the cup-cone fracture in a round tensile bar. Acta Metall. 32, 157–169.
Tvergaard, V., Needleman, A., 1986. Effect of material rate sensitivity on failure modes in the Charpy V-notch test. J. Mech. Phys.
Solids 34 (3), 213–241.
Wilshire, B., Scharning, P.J., 2007. Creep ductilities of 9–12% chromium steels. Scripta Mater. 56, 1023–1026.
Zaïri, F., Naït-Abdelaziz, M., Gloaguen, J.M., Lefebvre, J.M., 2008. Modelling of the elasto-viscopastic damage behaviour of glassy
polymers. Int. J. Plasticity 24, 945–965.
Zhang, Z.L., Niemi, E., 1995. A new failure criterion for the Gurson–Tvergaard dilational constitutive model. Int. J. Fracture 70,
321–334.

Please cite this article in press as: Besson, J. Damage of ductile materials deforming under multiple plastic or
viscoplastic mechanisms. Int. J. Plasticity (2009), doi:10.1016/j.ijplas.2009.03.001

You might also like