Sie sind auf Seite 1von 18

International Journal of Solids and Structures xxx (2013) xxx–xxx

Contents lists available at ScienceDirect

International Journal of Solids and Structures


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

An implicit-gradient-enhanced incremental-secant mean-field


homogenization scheme for elasto-plastic composites with damage
L. Wu a,⇑, L. Noels a, L. Adam b, I. Doghri b,c
a
University of Liege, Department of Aeronautics and Mechanical Engineering, Computational & Multiscale Mechanics of Materials (CM3), Chemin des Chevreuils 1,
B-4000 Liège, Belgium
b
e-Xstream Engineering, Axis Park-Building H, Rue Emile Francqui 9, B-1435 Mont-Saint-Guibert, Belgium
c
Université Catholique de Louvain, Bâtiment Euler, 1348 Louvain-la-Neuve, Belgium

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

Article history: This paper presents an incremental-secant mean-field homogenization (MFH) procedure for composites
Received 27 March 2013 made of elasto-plastic constituents exhibiting damage. During the damaging process of one phase, the
Received in revised form 19 June 2013 proposed method can account for the resulting unloading of the other phase, ensuring an accurate pre-
Available online xxxx
diction of the scheme. When strain softening of materials is involved, classical finite element formula-
tions lose solution uniqueness and face the strain localization problem. To avoid this issue the model
Keywords: is formulated in a so-called implicit gradient-enhanced approach, with a view toward macro-scale sim-
Mean-field homogenization
ulations. The method is then used to predict the behavior of composites whose matrix phases exhibit
Composites
Damage
strain softening, and is shown to be accurate compared to unit cell simulations and experimental results.
Non-local Then the convergence of the method upon strain softening, with respect to the mesh size, is demon-
Incremental-secant strated on a notched composite ply. Finally, applications consisting in a stacking plate, successively with-
out and with a hole, are given as illustrations of the possibility of the method to be used in a multiscale
framework.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction strain states. Besides MFH methods, non-linear effects can be con-
sidered by other homogenization methods, such as the method of
Homogenization methods can predict the macro or meso-scopic cells proposed by Lissenden and Arnold (1997) and Aboudi et al.
response of heterogeneous materials from their micro-structure (2003), the unit cell finite element (FE)-based computations as per-
constituents properties, with an acceptable accuracy, while being formed by Wieckowski (2000), Segurado et al. (2002), Ji and Wang
much more computationally efficient than direct numerical simu- (2003) and Carrere et al. (2004), or again such as the multiscale FE2
lations. Among those methods, the mean-field homogenization method pioneered by Kouznetsova et al. (2002, 2004) as a non-
(MFH) approach is a (semi-) analytical framework for the modeling exhaustive list. Kanouté et al. (2009) and Geers et al. (2010) have
of multi-phase composites. MFH methods were first developed for presented overviews of the different homogenization methods.
linear elastic composite materials by extending Eshelby (1957) sin- Although multiscale homogenization methods have achieved
gle inclusion solution to multiple inclusions interacting in an aver- an acceptable level of accuracy to capture the non-linear behavior
age way in the composite material. Most common extensions of of composites (Pierard et al., 2007), accounting for material degra-
Eshelby (1957) solution are the Mori–Tanaka scheme developed dation, through damage or fracture models, remains highly chal-
by Mori and Tanaka (1973) and Benveniste (1987) and the self- lenging as discussed by Geers et al. (2010) and LLorca et al.
consistent scheme pioneered by Kröner (1958) and Hill (1965b). (2011). Besides the complexity of formulating such a multiscale
MFH schemes can also be developed in the non-linear range to ac- method, the governing partial differential equations lose ellipticity
count for non-linear behaviors of the composite material’s constit- at strain-softening onset, losing the uniqueness of the FE solution.
uents. Most of these extensions revolve around the definition of a Many enhanced physical or phenomenological models, which
so-called linear comparison composite (LCC) (Talbot and Willis, introduce higher-order terms in the continuum, were proposed
1985, 1987, 1992; Castañeda, 1991, 1992; Molinari et al., 2004), to avoid the strain localization issue, such as in the Cosserat model
which is a virtual linear composite whose constituents behaviors reformulated by De Borst (1991), the non-local model pioneered by
match the linearized behaviors of the real constituents for given Bažant et al. (1984) or the gradient model, as exploited by Zbib and
Aifantis (1989). Because of the presence of higher order terms, the
⇑ Corresponding author. Tel.: +32 4 366 94 53; fax: +32 4 366 95 05. interactions between neighboring material points are reflected in
E-mail address: L.Wu@ulg.ac.be (L. Wu). these models through an internal length, which is related to the

0020-7683/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
2 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

micro-structure and to the failure mechanisms of the material. An Another existing MFH method is the affine method, which ap-
overview of these methods was presented by De Borst et al. (1993). plies the mean-field homogenization on the total strain field as
One possible difficulty, with the existence of higher-order terms in proposed by Molinari et al. (1987, 2004) for visco-plastic materials,
the continuum, lies in the requirement of developing finite ele- and by Zaoui and Masson (2002) and Masson et al. (2000) for elas-
ments with higher-order derivatives continuity to evaluate explic- to-plastic materials. Chaboche et al. (2005) showed that this meth-
itly internal variable derivatives, such as the strain gradient. To od can lead to too stiff results when an anisotropic tangent
alleviate this complexity, the non-local kernel has been reformu- operator is considered in the homogenization process. The accu-
lated by Peerlings et al. (1996), Geers (1997) and Peerlings et al. racy of this method for visco-plastic composites has been im-
(1998) in an implicit way such that a new non-local variable, rep- proved by Pierard and Doghri (2006a), Mercier and Molinari
resentative of an internal variable and its derivatives, results from (2009) and Doghri et al. (2010). When considering damage, the af-
the resolution of a new boundary value problem. Besides the fine method potentially suffers from the same over-estimation lim-
advantage of using C 0 elements, although the elements have now itation as the incremental tangent method. The LCC can also be
one additional degree-of-freedom per node, this approach also defined from a secant operator, which is joining the origin to the
possesses the feature of being fully non-local as it is constructed current strain/stress state, as initially proposed by Berveiller and
on the basis of a weighted averaging integral under the form of a Zaoui (1978) for elasto-plastic materials. However, this secant
new partial differential equation, contrarily to non-local models method is limited to monotonic and proportional loading paths.
constructed on the incorporation of higher-order terms as for the The MFH methods previously described only consider first-sta-
models of Aifantis (1992) and Svedberg and Runesson (1997). tistical-moment values of the strain and stress fields during the
Although such high-order and non-local formulations have homogenization process. This can lead to poor predictions in the
been widely used in direct numerical simulations, their applica- elasto-plastic case, as shown by Moulinec and Suquet (2003). This
tions in multiscale computations are not commonly developed. motivated to consider the second-statistical-moment (Castañeda,
When considering semi-analytical homogenization processes, Liu 1996) during the homogenization process. Such methods have
and Hu (2005) have applied the Cosserat model in the Mori–Tana- been proposed for the secant formulations by Suquet (1995) and
ka procedure to study the particle size dependency of composite Castañeda (2002a,b) and for the incremental-tangent formulation
materials, Dascalu (2009) has connected the locally periodic mi- by Doghri et al. (2011). Suquet (1995) actually showed that the
cro-crack with the macroscopic damage through an asymptotic variational forms pioneered by Castañeda (1992) correspond to a
homogenization, and Knockaert and Doghri (1999) have intro- second-order secant formulation, which was called modified se-
duced, in a 1D framework, the gradients of the internal variables, cant. Finally, incremental variational formulations, which also cor-
which are obtained from a micro/macro homogenization proce- respond to a second-moment estimation, were recently proposed
dure, at the macro-scale computation. When considering numeri- for visco-elastic composites by Lahellec and Suquet (2007a,b), for
cal computational homogenizations, Massart et al. (2005, 2007) thermo-elastic composites by Lahellec et al. (2011), for elasto-(vis-
have considered the non-local approach at the micro-scale in the co-) plastic composites by Brassart et al. (2011, 2012) and for elas-
framework of the computational homogenization for the problem to-visco-plastic composites with isotropic and kinematic
of masonry, and Coenen et al. (2011, 2012) have extended this hardening laws by Lahellec and Suquet (2013). However introduc-
method in a more general setting. ing damage models in combination with second-statistical-
Recently, Wu et al. (2012) – the authors – have proposed a MFH moment MFH methods is still an open area.
analysis allowing softening at the micro-scale and at the macro- In order to remain accurate for materials exhibiting strain
scale to be captured without causing localization. In order to avoid softening, it is mandatory to allow the fibers to be elastically un-
the strain/damage localization caused by the matrix material soft- loaded when the stress field in the matrix decreases due to the
ening, the implicit non-local formulation (Peerlings et al., 2001; degradation process. Recently, Wu et al. (2013) proposed a new
Engelen et al., 2003) was adopted during the homogenization pro- first-moment incremental-secant MFH approach for elasto-plastic
cess. In this formulation, the non-local accumulated plastic strain materials. In the formulation, at a given strain/stress state of the
of the matrix is defined and depends on the local accumulated composite material, a virtual unloading step is applied on the com-
plastic strain and on its derivatives through the resolution of a posite level and the residual stresses are evaluated in both phases.
new boundary value problem following the developments of Peer- Thus a secant approach is applied from this unloaded stage to de-
lings et al. (1996), Geers (1997) and Peerlings et al. (1998). fine the LCC. This method was shown to be simple to implement
The formulation presented by Wu et al. (2012) was based on the and to predict results with an accuracy comparable to other MFH
incremental-tangent MFH pioneered by Hill (1965a). This formula- schemes for elasto-plastic materials, for short and long fiber com-
tion defines the LCC from linearized relations between the stress posite materials, and also to be fit for monotonic and non-mono-
and strain increments of the different constituents around their tonic loadings.
current strain states. Thus, the classical homogenization tech- It is intended in this paper to extend this incremental-secant
niques for linear responses can still be used on the strain incre- MFH to composites whose matrix exhibits a damaging process. Be-
ments to predict the behaviors of elasto-plastic composites, as cause of the virtual elastic unloading step involved in the formula-
developed by Pettermann et al. (1999), Doghri and Ouaar (2003), tion the method remains highly accurate during the softening
Doghri and Tinel (2005) and Pierard and Doghri (2006b). The stage of the composite, even for high fiber volume fractions. This
MFH scheme developed by Wu et al. (2012), for composites whose non-local multiscale method is also implemented in a finite ele-
matrix phase obeys an elasto-plastic law with non-local damage ment code to simulate problems at the laminate level.
enhancement, was shown to have a reduced accuracy for high vol- The paper is organized as follows. Section 2 presents generali-
ume fractions of fibers when compared to the direct numerical ties on the MFH, as well as the main ideas of the incremental-se-
simulations of a representative volume element (RVE), due to the cant MFH approach for elasto-plastic materials without damage.
recourse to an incremental-tangent formulation. Indeed, during Section 3 presents the extension of the incremental-secant MFH
the strain softening of the matrix, the fibers should see an elastic for materials exhibiting damage, after having recalled the mechan-
unloading due to the damaging process in the matrix, which can- ics of elasto-plastic materials exhibiting damage in a non-local set-
not be modeled using this incremental-tangent approach. There- ting. The prediction of the behavior of fiber-reinforced elasto-
fore, it appears clear that the method should be reformulated plastic matrix is demonstrated, in Section 4, to be more accurate
within a different MFH framework. than with the previously developed non-local incremental-tangent

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 3

n o1
method, as the softening response is better captured. The model is 1
B ¼ I þ S : ½ðC LCC
0 Þ : C LCC
I  I ; ð4Þ
also validated against experimental results for a metal matrix com-
posite material. In Section 5, the convergence of the method with  LCC Þ depends on the geometry of
where Eshelby (1957) tensor SðI; C 0
respect to the mesh size is illustrated on a notched sample. Finally the inclusions phase and on the virtual elastic operator C  LCC . The
0
the model is applied to study the response of laminates. First the  LCC and C
expressions of the tensors C  LCC depend on the chosen
0 I
material model is validated againt experiments by considering ten- MFH process, and will be particularized in the next subsection for
sile tests including unloading on unidirectional carbon-fibers rein- the incremental-secant method.
forced epoxy coupons. Then specimens made of the same laminate,
but having a hole to force the localization, are considered. In this 2.2. Incremental-secant MFH for elasto-plastic composites without
last example it can be seen that the damage-enhanced MFH cap- damage
tures the damage path oriented along the ply fiber directions in
agreement with experimental results. The development of an incremental-secant MFH formulation
was motivated by the overestimation of the macro-response ob-
2. Incremental-secant MFH for elasto-plastic composites tained by the previously developed incremental-tangent MFH
without damage when the matrix exhibits damage. Thus Wu et al. (2013) proposed
to consider an unloading step before applying the MFH process,
In this section concepts of the MFH for linear and non-linear which leads to an incremental-secant formulation with per-phase
composite materials are first presented before introducing the residual strains. Such a formulation has been developed for elas-
key ideas of the incremental-secant MFH. to-plastic materials without considering damage, and will be ex-
tended to the damage case in this paper. In this section, the main
2.1. Generalities on MFH ideas developed for elasto-plastic materials that do not exhibit
damage are recalled.
Fig. 1 represents a classical multiscale approach: at each macro- Let us consider a time interval ½tn ; tnþ1 . The total strain tensor
point X the macro-strain e is known and the macro-stress r  is at time t n reads en and the strain increment Denþ1 on the interval
sought through the resolution of a micro-scale boundary value results from the finite element resolution yielding
problem (BVP). Toward this end, the macro-point is viewed at
the micro-level as the center of a RVE of domain x 2 x and bound- enþ1 ¼ en þ Denþ1 ; ð5Þ
ary @ x. Then, the Hill–Mandell condition, expressing the equality at time t nþ1 , see Fig. 2(a). The key point of the incremental-secant
between energies at both scales, transforms the relation between method is to assume at time t n a residual strain tensor eres
n that cor-
macro-strains e and stresses r  into a relation between average responds to an elastic unloading from the stress state rn to a stress
R
strains hei and stresses hri over the RVE, with hf ðxÞi ¼ V1x x f ðxÞdV. state rres
n . Note that this residual stress does not necessarily corre-
For two-phase composite materials of matrix subdomain x0 spond to a zero-stress state. During the homogenization process,
and of inclusions subdomain xI -subscripts 0 and I refer respec- the residual stress for the homogenized composite material will
tively to the matrix and to the inclusions, the macro-strains e be null while the different phases can have non-zero values.
and stresses r can be expressed as
e ¼ v 0 heix þ v I heix ; and ð1Þ 2.2.1. Incremental-secant moduli for elasto-plastic phases
0 I
Considering an elasto-plastic phase of the composite material,
r ¼ v 0 hrix0 þ v I hrixI ; ð2Þ the secant linearization of the elasto-plastic material is thus car-
ried out in the time interval [t n ; t nþ1 ] with the strain increment
with the phase volume ratios satisfying v 0 þ v I ¼ 1 . For simplicity, Dernþ1 , such that
in the following developments, the notations hixi will be replaced
by i .
enþ1 ¼ eres r
n þ Denþ1 : ð6Þ
The relation between the average incremental strains in the two In this subsection, the indices related to the composite phases are
phases in the case of non-linear behaviors relies on the definition omitted to simplify the notations. The stress tensor is now defined
of a so-called linear comparison composite (LCC), which is charac- from the stress increment Drrnþ1 following the incremental-secant
terized by the expressions of the virtual elastic operators C  LCC of
0 approach depicted in Fig. 2(a):
the matrix phase and C  LCC of the inclusions phase I, leading to
I
rnþ1 ¼ rres r
n þ Drnþ1 : ð7Þ
DeI ¼ B ðI;C LCC  LCC
0 ; C I Þ : De0 : ð3Þ Sr
The stress increment is computed using C , the residual-incremen-
This equation describes the relation between the averages of the tal-secant operator of the elasto-plastic material, following
strain increments per phase through the strain concentration tensor
B . Considering the Mori and Tanaka, 1973 (M–T) assumption, the Drrnþ1 ¼ C Sr : Dernþ1 : ð8Þ
strain concentration tensor reads
During the elastic response, this residual-incremental-secant
operator corresponds to the elastic material tensor C el . During
the plastic flow, this operator is deduced by solving the following
system of equations.

 The von Mises stress criterion reads


 
f rnþ1 ; pnþ1 ¼ req
nþ1  Rðpnþ1 Þ  rY ¼ 0; ð9Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where f is the yield surface, req ¼ 32 devðrÞ : devðrÞ is the
equivalent von Mises stress, rY is the initial yield stress, and
RðpÞ P 0 is the isotropic hardening stress in terms of the accu-
mulated plastic strain p, which is an internal variable character-
Fig. 1. Multiscale method. izing the irreversible behavior.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
4 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

Fig. 2. Definition of the incremental-secant formulation, without considering damage. (a) Definition of the residual strain and stress and of the residual-secant operator. (b)
Definition of the zero-secant operator.

 The Cauchy stress at time t nþ1 is written as 2.2.2. Incremental-secant MFH scheme
res el r el p p The stress tensor in the different composite material phases can
rnþ1 ¼ r þ C : De
n nþ1  C : De ; with De ¼ DpN nþ1 ;
be computed from a secant approach, following
ð10Þ ( Sr
rres
n þC : Dernþ1 for the residual-incremental-secant method;
where N is the plastic flow direction. Within the incremental-se- rnþ1 ¼ S0 r
C : Denþ1 for the zero-incremental-secant method:
cant approach, this direction is set to
ð14Þ
 dev
3 rnþ1  rres
n
N nþ1 ¼  eq ; ð11Þ The secant operators can then be used in the MFH approach to de-
2 rnþ1  rres
n fine the LCC operators. Using these definitions of the LCCs, the MFH
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
stated by Eqs. (1)–(3) can thus be applied, and the system of equa-
which satisfies N : N ¼ 32 and Dp ¼ 23 Dep : Dep . On the one hand,
 dev   tions is rewritten as
if rres
n ¼ 0 this last equation corresponds to N nþ1 ¼ @f ð@rr; pÞ
nþ1
Dernþ1 ¼ v 0 Der0nþ1 þ v I DerInþ1 ; ð15Þ
and we have the classical relation. On the other hand, if
 res dev r nþ1 ¼ v 0 r0nþ1 þ v I rInþ1 ; ð16Þ
rn – 0; N nþ1 is a first-order approximation in terms of De
of the normal to the yield surface in the stress space. It has been with the stress tensors in the different phases deriving from
shown in Wu et al. (2013) that N nþ1 ¼ N trnþ1 , where the trial state Eq. (14), and the relation between the strain increments reading
‘‘tr’’ corresponds to an elastic increment (Dep;tr ¼ 0). The elasto-
plastic scheme consists of solving Eqs. (10) and
DerInþ1 ¼ B ðI;C S0 ; C SI Þ : Der0nþ1 ; ð17Þ
 
f rnþ1 ; pnþ1 ¼ 0, with pnþ1 ¼ pn þ Dp, in terms of Dp and req nþ1 . where C  S substitutes to either C
 Sr or to C
 S0 , the per-phase constant
 It has been demonstrated by Wu et al. (2013) that the residual- secant-operators resulting from the MFH process. The choice of the
incremental-secant operator of the linear comparison material, residual- or zero-incremental-secant operators will be discussed in
which can be computed from rnþ1 , is isotropic for J2-plasticity, the development of the MFH with damage. To complete these equa-
and that it reads
n ¼ v 0 r0n þ v I rIn ¼ 0.
tions, the unloaded state is defined using r  res res res

The detailed MFH process is described by Wu et al. (2013).


C Sr ¼ 3jr I vol þ 2lrs I dev : ð12Þ
r
The bulk modulus j of the virtual elastic material is directly ob- 3. Incremental-secant MFH for elasto-plastic composites with
tained as jel and the shear modulus lrs of the virtual elastic damage
material can be obtained by decomposing the increments of
the stress and strain tensors into their hydrostatic and deviatoric In this section the incremental-secant MFH method summa-
parts, as detailed in Appendix B.1. rized in Section 2.2 is extended to the non-local damage formula-
 The derivative of the operator (12) required during the upcom- tion. First the equations of elasto-plastic materials combined with
ing MFH process is obtained following Appendix B.1. a Lemaitre and Chaboche (1991) damage model are given in a non-
 The incremental-secant-operator (12) can be used to define the local setting. Then the secant approach for one-phase materials is
LCC during the MFH process. One alternative proposed by Wu presented before eventually developing the incremental-secant
et al. (2013) follows the suggestion in Fig. 2(b). In that case MFH. The improvement of the prediction accuracy compared to
the residual stress rres
n is omitted in the described approach the incremental-tangent MFH scheme will be illustrated for prob-
and the plastic flow direction (11) is rigorously normal to the lems involving matrix softening in the next section.
yield surface. For this latter approach, the zero-incremental-
secant operator of the linear comparison material reads 3.1. Elasto-plastic materials with non-local enhanced damage models
C S0 ¼ 3j0 I vol þ 2l0s I dev ; ð13Þ
The non-local implicit approach can be applied to damage mod-
where j0 and l0s are the elastic bulk and shear moduli of the vir- els in order to avoid the loss of ellipticity at the macro-scale. Note
tual elastic material, respectively, obtained as detailed in Appen- that formulating the MFH in a local form can be done by consider-
dix B.1. ing the elasto-plastic model with damage also in a local form. Only

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 5

Fig. 3. Definition of the incremental-secant formulation when damage is considered. (a) Definition of the residual strain and stress and of the residual-secant operator. (b)
Definition of the zero-secant operator.

the non-local case is considered herein, as the local form can be de- 1 el
Y¼ e : C el : eel ; ð22Þ
rived easily from the presented equations. 2
The damage is introduced with the usual assumption that the
strain tensors observed in the actual body and in its undamaged and a is an interpolation parameter ranging from 0 to 1.
representation are equivalent (Lemaitre, 1985), and the usual def- In this gradient enhanced damage model, the non-local accu-
~ is computed from the implicit formulation
mulated plastic strain p
inition of the effective stress reads
r ~  O  ðc g  Op
p ~Þ ¼ p; ð23Þ
r^ ¼ ; ð18Þ
ð1  DÞ
where c g is the characteristic lengths tensor as defined by Wu et al.
where r is the apparent Cauchy stress tensor and where 0 6 D < 1 (2013). For isotropic materials this tensor reads c g ¼ diagðcÞ, with
is the damage variable. 2
c ¼ l the square of the characteristic size. For composite materials,
Assuming an elasto-plastic material, which obeys J2-elasto- the anisotropy can be accounted for in this definition. As an exam-
plasticity, the von Mises stress criterion (9) is now written ple, for the matrix phase of a unidirectional (UD) continuous fiber-
f ðr
^ ; rÞ 6 0, where r is an internal variable related to the accumu- i 2
reinforced composite material, ci ¼ ðl Þ is different in the directions
lated plastic strain p and to the plastic multiplier k_ following parallel and transverse to the inclusions. The tensor is thus com-
r_ ¼ k_ ¼ ð1  DÞp,
_ see (Doghri, 1995) for details. However in this pa-
puted from the rotation tensor R describing the inclusions
per we use the classical approximation that consists in writing the orientation
J2-plasticity in the effective stress space: f ðr
^ ; pÞ 6 0.
 
During the plastic flow f ¼ 0; Dp > 0 and the plastic strain ten- c g ¼ RT  diag ci  R: ð24Þ
sor increment follows the plastic flow direction
Dep ¼ DpN; ð19Þ Relation (23) is completed by the natural boundary condition

where N is usually the normal to the yield surface in the effective ~Þ  n ¼ 0:


ðc g  Op ð25Þ

stress space N ¼ @@fr^ ¼ 32 devð^
r^ eq . Finally using the coupled damage con- The linearizations of the model with respect to e and p
~ are pro-
cept leads to the stress expression at time t nþ1 :
vided in Appendix C.
 
rnþ1 ¼ ð1  Dnþ1 ÞC el : enþ1  epnþ1 ¼ ð1  Dnþ1 ÞC el : eelnþ1 : ð20Þ
3.2. Incremental-secant moduli with damage
What remains to be defined is the evolution of the damage. Follow-
ing the technique proposed by Geers et al. (1998) to develop non-lo-
The secant formulation described in Section 2.2 is extended
cal damage laws, in this paper the non-local accumulated plastic
~ is applied to calculate the damage evolution in Lemaitre herein to the non-local damage enhanced elasto-plastic case. We
strain p
are still considering a time interval ½t n ; t nþ1 , with the total strain
and Chaboche (1991) incremental model1:
8 tensor en at time t n and the strain increment Denþ1 resulting from
< 0; ~ 6 pC ;
if p the FE resolution and satisfying (5).
DD ¼  s ð21Þ At time t n , one can define an elastic unloading from the stress
Y nþa
: ~; if p
Dp ~ > pC :
S0 state rn or from the effective stress state r ^ n – that corresponds
to a residual strain tensor eres
n , see Fig. 3. The main idea of the incre-
In this expression, pC is a plastic threshold for the damage evolution,
mental-secant method, is to define a LCC, subjected to a strain
S0 and s are the material parameters, Y is the strain energy release
increment Dernþ1 , satisfying (6), and from which the stress tensor
rate computed as
is computed. As for the case without damage, the two methods
illustrated in Fig. 3 will be considered: the residual-incremental-
1
In order to simplify the developments of the method, a scalar damage model is secant method, which evaluates the effective stress tensor from
assumed. Although to remain more general, a damage variable should be represented
by a tensor due to the existence of several mechanisms of damage, this is not
the residual effective stress arising upon virtual unloading, and
necessary for damage induced by meso- or micro-plasticity as pointed out by the zero-incremental-secant method, which evaluates the effective
Lemaitre and Desmorat (2005), which justifies the use of a simple model. stress tensor from a zero-stress state.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
6 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

 dev
3.2.1. Residual-incremental-secant approach el r
In the case of damage enhanced elasto-plastic materials, the se- 3 C : Denþ1
N nþ1 ¼  eq ; ð32Þ
cant formulation has to account for the effective and current stress 2 C el : Der
nþ1
tensors, with r^ n ¼ rn =ð1  Dn Þ at time t n . After the elastic unload-
ing process the residual stress in each phase can be computed as well as the system of equations to be solved as
from, see Fig. 3  
eq  eq
r^ nþ1  r^ res
n þ 3lel Dp ¼ r
^ tr ^ res
nþ1  rn ; and ð33Þ
res el  
r n ¼ ð1  Dn Þ ^ res
r n
^ n  ð1  Dn ÞC : De
¼ ð1  Dn Þr unload
n : ð26Þ f r^ nþ1 ; pnþ1 ¼ 0; ð34Þ
During the time interval [t n ; tnþ1 ] the stress reaches rnþ1 , and the with pnþ1 ¼ pn þ Dp.
damage reaches Dnþ1 . Following the method pictured on Fig. 3(a),  Knowing r ^ nþ1 , and using Eqs. (28)–(30) compute the residual-
the effective stress tensor at time t nþ1 can be rewritten as incremental-secant operator of the undamaged linear compari-
r^ nþ1 ¼ r^ res ^r son material from
n þ Drnþ1 and rnþ1 ¼ ð1  Dnþ1 Þr
^ nþ1 ; ð27Þ
where ^ rnþ1 ¼ C Sr : Dernþ1 ¼ C el : Dernþ1  2lel DpN nþ1 ;
Dr ð35Þ

^ rnþ1 ¼ C Sr : Dernþ1 :
Dr ð28Þ which becomes after using Eqs. (28)–(30)
2 3
In this last equation C Sr is the residual-incremental-secant operator I :C dev el

of the undamaged linear comparison material. ^ rnþ1 ¼ 6


Dr el
4C  3lel Dp 
7
eq 5 : Dernþ1 ¼ C
Sr
el r
During the elastic regime, the elastic tensor C el can be used as C : Denþ1
undamaged residual-incremental-secant operator. During elasto-
: Dernþ1 : ð36Þ
plastic flow, the effective stress tensor r ^ nþ1 is computed from the
el
unloaded state in the following way: For J2-elasto-plastic materials, since C is isotropic, the residual-
incremental-secant operator of the undamaged linear compari-
 Evaluate the trial effective stress tensor from an elastic son material C Sr is also isotropic. Moreover, as C el ¼ 3jel I vol þ
response: 2lel I dev , one can directly deduce

r^ trnþ1 ¼ r^ n þ C el : Denþ1 ¼ r^ res el r


n þ C : Denþ1 : ð29Þ C Sr ¼ 3jr I vol þ 2lrs I dev ; ð37Þ
 If the trial effective stress tensor does not respect the von Mises with
 tr 
criterion (9), i.e. f r ^ nþ1 ; pn > 0, then apply the plastic
correction
jr ¼ jel ; and ð38Þ
el 2 el 2
3l Dp 3 l Dp
r^ nþ1 ¼ r^ trnþ1  C el : Dep ; with Dep ¼ DpN nþ1 : ð30Þ lrs ¼ lel   eq ¼ lel   eq : ð39Þ
el
C : De r
nþ1
r^ nþ1  r^ res
n
In this last equation, N is the plastic flow direction, which,
extending the assumption discussed by Wu et al. (2013) from  Practically the undamaged shear modulus of the virtual undam-
undamaged elasto-plasticity to damage coupling, reads aged elastic comparison material can be obtained by decompos-
 dev ing the increments of the effective stress and strain tensors into
Sr r  dev
3 C : Denþ1 3 r^ nþ1  r^ res
n
their hydrostatic and deviatoric parts:
N nþ1 ¼  eq ¼  eq ; ð31Þ
2 C Sr : Der 2 r^ nþ1  r^ res
n ^ r ¼ Dr
Dr ^ m 1 þ D^s and Der ¼ Dem 1 þ De; ð40Þ
nþ1
 res dev where Dr ^ m ¼ 13 trðDr ^ r  Dr
^ r Þ; D^s ¼ Dr ^ m 1; Dem ¼ 13 trðDer Þ, and
and which satisfies N : N ¼ 32. When r ^n – 0; N nþ1 is a first
where De ¼ Der  Dem 1, see Appendix A for the notations. In-
order approximation of the normal to the yield surface in the
deed, the increments of the von Mises equivalent stress and
effective stress space, see Fig. 4. Using Eqs. (29)–(31), one can
strain are respectively given by
deduce

III
III

tr tr
n 1 n 1
el el
2 pN 2 pN
n n

II
II
res res
n n
f ( n , pn ) 0 f ( n , pn ) 0
f( n 1 , pn 1 ) 0 f( n 1 , pn 1 ) 0
I
I

Fig. 4. Plastic corrections in the effective stress space (a) Radial return mapping (b) Approximation (31).

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 7

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ( SDr


3 2 ^ res
ð1  Dnþ1 Þr n þC : Dernþ1 for the residual-incremental-secant;
^ eq
Dr ¼ Ds : Ds and De ¼
^ ^ eq
De : De ð41Þ rnþ1 ¼
2 3 C SD0 : Dernþ1 for the zero-incremental-secant:

and one has directly 2 ð51Þ

Dr^ eq Note that in case of elastic response, the secant undamaged opera-
lrs ¼ : ð42Þ tor is the elastic tensor C el .
3Deeq
As lengthy described by Wu et al. (2013) when no damage is
 Evaluate the damage Dnþ1 following the non-local Lemaitre– considered, for composites with inclusions exhibiting an elasto-
Chaboche model described in Section 3.1. plastic behavior with a hardening coefficient lower than or of the
 Compute the final stress rnþ1 from r
^ nþ1 . Using Eq. (27), one also
same order as the one of the elasto-plastic matrix material, using
deduces the residual-incremental-secant operator for both phases leads to
rnþ1 ¼ ð1  Dnþ1 Þr^ res Sr
: Dernþ1 ; accurate predictions. However, for composites whose inclusions
n þ ð1  Dnþ1 ÞC ð43Þ
|fflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflffl} phase is much stiffer than the matrix material, such as elastic
SDr
C
inclusions embedded in an elasto-plastic matrix or elasto-plastic
where the residual-incremental-secant operator of the damaged inclusions with a high hardening compared to the one of the
isotropic-linear comparison material reads embedding elasto-plastic matrix, the zero-incremental-secant
operator should be used in the matrix phase to avoid predicting
C SDr ¼ 3jDr I vol þ 2lDr
s I
dev
; ð44Þ over-stiff responses. With this restriction on the choice of the ma-
with the equivalent damaged bulk and shear elastic moduli jDr trix operator, the method has been shown to predict the macro-
and lDr stress with an accuracy level at least similar to those of the first-or-
s directly obtained from
der incremental-tangent MFH method, or even advanced MFH
jDr ¼ ð1  Dnþ1 Þjel ; and ð45Þ schemes using per-phase statistical-second-moments of strain
Dr^ eq and stress fields. In the present case, we keep the same combina-
lDr r
s ¼ ð1  Dnþ1 Þls ¼ ð1  Dnþ1 Þ : ð46Þ
tions of secant operators.
3Deeq
We need to clarify that the equivalent damaged bulk elastic
modulus jDr is not constant although the plastic flow is 3.3. MFH scheme with non-local damage
incompressible.
 Evaluate the derivation of the operator (44) following Appendix In this section the incremental-secant MFH scheme presented
B.2. in Section 2.2 is extended to the case of non-local elasto-plastic
materials with damage, using the incremental-secant operators
3.2.2. Zero-incremental-secant approach of the isotropic-linear comparison materials as defined in Sec-
As lengthy discussed by Wu et al. (2013), when defining the tion 3.2 to construct a LCC. Unless the expressions need to be par-
LCC, it can be advantageous to modify the residual-incremental-se- ticularized to the residual-incremental-secant or to the zero-
cant approach by neglecting the residual stress – but not the resid- incremental-secant forms, the isotropic linear comparison operator
ual strain – in the matrix phase. The modification follows the
C SD ¼ 3jD I vol þ 2lDs I dev ; ð52Þ
suggestion illustrated in Fig. 3(b) and consists in neglecting r^ res
n
SDr SD0
in the formalism described here above, in which case will substitute to either C or C . Similarly, l holds for either
D
s
SD0 S0 lDr
s or for ls , and j holds for either j
D0 D Dr
or for jD0 . In this paper,
C ¼ ð1  Dnþ1 ÞC
we consider a damage model only in the matrix phase.
¼ 3 ð1  Dnþ1 Þj0 I vol þ 2 ð1  Dnþ1 Þl0s I dev ð47Þ Considering a time interval ½tn ; tnþ1 , the system of governing
|fflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflffl}
jD0
equations of the homogenized material using the implicit gradi-
lD0
s
ent-enhanced elasto-plasticity is stated at time tnþ1 by
is the zero-incremental-secant operator of the damaged isotropic-
linear comparison material expressed in terms of the equivalent r  rTnþ1 þ f nþ1 ¼ 0 for the homogenized material; ð53Þ
 
damaged bulk and shear elastic moduli jD0 and lD0 ~nþ1  r  c g  rp
p ~nþ1 ¼ pnþ1 for the matrix material only; ð54Þ
s . These values
are readily obtained from
where f represents the body force vector, and where p ~ is a homog-
D0rm ð1  Dnþ1 Þr
^m enized representation of the non-local accumulated plastic strain of
j ¼ ¼ ¼ ð1  Dnþ1 Þjel ; and ð48Þ
3 Dem 3 Dem the matrix material. This set of equations can be solved using a
req ð1  Dnþ1 Þr
^ eq weak finite element form combined to a Newton–Raphson lineari-
lD0
s ¼ ¼ : ð49Þ
3Deeq 3Deeq zation procedure, see Wu et al. (2012) for details.
During that time interval, the known data are the macro-total
The linearization of C SD0 is given in Appendix B.2.
strain tensor en , the macro-strain increment Denþ1 and the internal
variables at t n , which include the usual internal variables of the
3.2.3. Incremental-secant approach summary
constituents material models and the residual variables computed
Two incremental-secant models have been considered, the first
from the elastic unloading step: the residual strains in the compos-
one accounting for the residual stress, see Fig. 3(a), and the other
ite, in the inclusions phase, in the matrix phase and the residual
one is formulated from a stress-free state, see Fig. 3(b), with
( stresses in the inclusions and matrix phases. The strain increment
Sr
r^ res
n þC : Dernþ1 for the residual-incremental-secant method; Denþ1 resulting from the iterations at the weak form level is differ-
r^ nþ1 ¼ S0 r
C : Denþ1 for the zero-incremental-secant method: ent from the strain increment Dernþ1 applied to the LCC used in the
ð50Þ MFH procedure. Combining (5) and (6) for the homogenized mate-
rial, one has
In the current stress state, these two models read
Dernþ1 ¼ en þ Denþ1  eres
n : ð55Þ

The explicit evaluation of eres


n is described in the details of the MFH
2
If Deeq ¼ 0, the indefiniteness is solved by considering lrs ¼ lel . process reported here below. Thus, based on the definition of C SD

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
8 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

(52) for damage-enhanced elasto-plastic materials, the operators 6. Check if the residual j F j6 Tol. If so exit the loop.
C  LCC of the LCCs can be defined to apply the MFH scheme
 LCC and C 7. Else, compute the Jacobian matrix at constant Dernþ1 , such that
0 I
stated by Eqs. (1)–(3). Eventually, the MFH process is summarized dF ¼ J : derI following Appendix D.
by the following equations 8. Correct the strain increment in inclusions
Dernþ1 ¼ v 0 Der0nþ1 þ v I DerInþ1 ; ð56Þ DerInþ1 DerInþ1 þ ceI with c eI ¼ J 1 : F; ð62Þ
r nþ1 ¼ v 0 r0nþ1 þ v I rInþ1 ; ð57Þ
then start a new iteration (go to step 1).
DerInþ1 ¼ B ðI;C SD S r
0 ; C I Þ : De0nþ1 ; ð58Þ  After convergence, compute
1. The homogenized stress
where the stress tensors are evaluated in each phase following Eq.
(51), without damage for the inclusions phase. To complete this r nþ1 ¼ v 0 r0nþ1 þ v I rInþ1 ; ð63Þ
set, at the unloaded state, we have
and the internal variables.
n ¼ v 0 r0n þ v I rIn ¼ 0:
r res res res
ð59Þ 2. The ‘‘consistent’’ linearization of the homogenized stress

In this paper, a ‘‘first-order moment’’ method is considered, and the  ¼ C alg : de þ C p~ dp
dr ~; ð64Þ
MFH process is described as follows. Note that during the MFH pro-  alg and C p~ given in E.
with the ‘‘consistent’’ tangent operators C
cess, the strain increment of the composite material Denþ1 , and the
~nþ1 – required to  An unloading step is thus applied here to fit the incremental-
non-local plastic strain increment of the matrix Dp
secant process, and the obtained results will be kept as history
evaluate the damage – are constant, as they result from the FE iter-
variables at time step tnþ1 . The required residual variables from
ations formulated from the strong form (53) and (54).
this unloading are the residual strains in the composite material
Practically, the resolution of Eqs. (55)–(59) is achieved as
eres res res
nþ1 , in the inclusions phase eInþ1 and in the matrix phase e0nþ1 , as
follows
well as the residual stress in the inclusions phase rres Inþ1 and the
effective residual stress in the matrix phase r ^ res
0nþ1 , respectively.
 Initialize the strain increment in the inclusions phase:
1. The residual strain eres  res
nþ1 (the strain at rnþ1 ¼ 0) of the composite
Dernþ1 ! DerInþ1 .
material is calculated from an unloading step, which is
 Follow the iterations process (upper indices (i) for values at iter-
assumed to be a purely elastic process. The unloading operator
ation i of time tnþ1 are omitted for simplicity):
of the damaged composite is
1. Call the constitutive material function of the real inclusions
h i
1
material with the strain tensor increment in the inclusions C elD v I C elI : B þ v 0 C elD 
nþ1 ¼ nþ1 0nþ1 : v I B þ v 0 I ; ð65Þ
phase DerInþ1 and the internal variables at time tn as input.
After having applied the constitutive model described in Sec- with
tion 3.2, but without considering damage, the output is the  1 1
updated stress rInþ1 , the internal variables at time tnþ1 , the B ¼ I þ S : ½ ð1  Dnþ1 ÞC el
0nþ1 : C el
Inþ1  I : ð66Þ
 alg ¼ @ rInþ1 ,4 see Appen-
‘‘consistent’’3 (anisotropic) operator C In þ1 @ DeI
nþ1
dix C for details, and the incremental-secant operator C  S for
Inþ1
Note that C elD el
0nþ1 ¼ ð1  Dnþ1 ÞC 0nþ1 is used in the Eshelby tensor of
the inclusions phase. In case there is no plastic flow, we use the elastic unloading step. The residual strain of the composite
 S ¼ C el .
C satisfying (59) can be calculated by
Inþ1 I
2. Compute the average strain in the matrix phase: 1
eres  unload ¼ enþ1  ðC elD
nþ1 ¼ enþ1  Denþ1 nþ1 Þ :r
 nþ1 : ð67Þ
Der0nþ1 ¼ ðDernþ1  v I DerInþ1 Þ=v 0 : ð60Þ
2. The residual strains in the inclusions and in the matrix phases
3. Call the constitutive material function of the real matrix mate- are computed following the M–T method, yielding
rial with the strain tensor increment in the matrix phase 1
Der0nþ1 and the internal variables at time t n as input. After hav- eres unload
Inþ1 ¼ eInþ1  DeInþ1 ¼ eInþ1  B : ½v I B þ v 0 I
ing applied the constitutive model described in Section 3.2, : Deunload ð68Þ
nþ1 ;
the output is the updated stress r0nþ1 , the damage Dnþ1 , the
internal variables at time tnþ1 , the ‘‘consistent’’ (anisotropic) 1
operators C alg and C  algD ¼ ð1  Dnþ1 ÞC  alg , see Appendix C for eres unload
0nþ1 ¼ e0nþ1  De0nþ1 ¼ e0nþ1  ½v I B þ v 0 I
0nþ1 0nþ1 0nþ1
details, and the incremental-secant operator C  SD for the : Deunload ð69Þ
0nþ1 nþ1 :
matrix phase. In case there is no plastic flow, we use
 SD ¼ ð1  Dnþ1 ÞC .
C el 3. Finally, the residual (effective) stresses in the inclusions and
0nþ1 0
4. Predict the Eshelby tensor IðI; C  SD Þ using the isotropic dam- matrix phases can be obtained, respectively, from
0nþ1
aged secant-operator of the matrix phase. rres el unload
Inþ1 ¼ rInþ1  C Inþ1 : DeInþ1 ; ð70Þ
5. For a time interval ½t n ; tnþ1  both Dernþ1 and Dp ~nþ1 are constant
and verifying Eq. (58) corresponds to satisfying F ¼ 0, where F r^ res
0nþ1 ¼ r0nþ1 
^ C el
0nþ1 : Deunload
0nþ1 : ð71Þ
is the stress residual vector developed in Appendix D, which
reads
 Note that by expressing Eqs. (67)–(69) at time t n , one has
1 1
F ¼ C SD : D er
 S : ðD er
 Der
Þ  C SInþ1 : DerInþ1 : e ¼ v 0 eres
res 0n þ v I eIn :
res
ð72Þ
0nþ1 Inþ1
v0 Inþ1 nþ1 n

ð61Þ Moreover, Eqs. (55) and (56) yield


v 0 Der0 nþ1
þ v I DerInþ1 ¼ Dernþ1 ¼ v 0 e0n þ v I eIn þ Denþ1  eres
n ; ð73Þ

which becomes, after using Eq. (72),


3
In this paper we will use the term ‘‘consistent’’ operator for the derivative of the    
stress state with respect to the deformation increment. Denþ1 ¼ v 0 Der0nþ1  e0n þ eres
0n þ v I DerInþ1  eIn þ eres
In : ð74Þ
4
The derivative with respect to DerI nþ1 is the same as the derivative with respect to
DeI nþ1 . This relation demonstrates that the process still satisfies

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 9

Denþ1 ¼ v 0 De0nþ1 þ v I DeInþ1 : ð75Þ ered herein. This consists of continuous elastic isotropic fibers
embedded in a matrix material, which follows the elasto-plastic
These relations can now be used in a FE implementation. More de-
behavior model enhanced by a non-local damage setting, as de-
tails about the FE implementation of a non-local MFH were dis-
scribed in Section 3.1. The material parameters are
cussed by Wu et al. (2012) for an incremental-tangent approach,
and can be directly adapted for this incremental-secant approach.
 Inclusions: EI ¼ 238 GPa and mI ¼ 0:26.
 Matrix: E0 ¼ 2:89 GPa; m0 ¼ 0:3; rY0 ¼ 35 MPa; h0 ¼ 73 MPa;
4. Model validation m0 ¼ 60; S0 ¼ 2 MPa; s ¼ 0:5, and pC ¼ 0.

In this section, the accuracy and the reliability of the proposed The matrix material follows the hardening law
incremental-secant method with non-local damage are first as-
R0 ðpÞ ¼ h0 ð1  em0 p Þ ð76Þ
sessed through the comparison with direct FE simulations on rep-
resentative unit cells of the micro-structure. In particular it is and the damage law (21). The volume fraction of the continuous fi-
shown that the new incremental-secant MFH method reaches a bers is v I ¼ 50%. The test consists in a transverse loading of the
higher accuracy than the incremental-tangent MFH scheme previ- composite under constrained strain (with plane-stress state in the
ously developed by Wu et al. (2012). The model is then validated other transverse direction and plane-strain state in the longitudinal
against experimental results for a metallic composite material. direction), followed by a complete unloading until reaching a zero-
strain state. As the inclusions remain elastic, the matrix obeys to the
4.1. Comparison with direct FE simulations zero-incremental-secant formulation while the inclusions phase
follows the residual-incremental-secant formulation. For the FE cell
The example studied by Wu et al. (2012), using an incremental- simulations, the characteristic length of the matrix phase is taken
2
tangent MFH and with FE simulations of a periodical cell, is consid- such that l ¼ 2 mm2. This value was calibrated from experimental

120
100
80
[MPa]

60
eq

40 FE (Wu et al., 2012)


20 MF, incr. tangent
MF, incr. secant
0
0 0.05 0.1

Fig. 5. Results for continuous-elastic fibers embedded in a matrix following an elasto-plastic behavior with damage. Comparison of the values obtained with the direct FE
simulations and with the MFH schemes. ((a) and (b)) Average stresses in the composite material along the loading direction and along the fiber direction. ((c) and (d)) Average
stresses in both phases along the loading direction. (e) Effective von Mises stress in the matrix phase. (f) Damage in the matrix phase – the minimum, average, and maximum
damage values are reported.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
10 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

0.02
MF, incr. secant
0.015

D
0.01

0.005

0
0 0.005 0.01 0.015 0.02

Fig. 6. SiC particles-reinforced AL–Si–Mg alloy. (a) Comparison between the damage-enhanced MFH prediction and the experimental results. (b) Damage evolution predicted
in the matrix.

results by Geers et al. (1999) for short fiber composites and is used tion is not uniform -parts are in tension and parts are in compres-
herein for the matrix phase. For the MFH results, as the fields re- sion, the average of the matrix effective von Mises stress is never
main uniform, the results are independent on this characteristic equal to zero. This is a limit of the first-order method as
 eq
length.
hreq ix0 – hrix0 . Note that results obtained by the incremen-
Fig. 5(a) and (b) compare the macro-stress evolution, respec-
tively along the loading direction and along the fibers direction, ob- tal-tangent MFH are actually not more accurate. Indeed in the
tained with the new incremental-secant MFH and the previously vicinity of the point at which the composite material stress along
developed incremental-tangent MFH to the FE results. Clearly, the loading direction vanishes, the average stress along the fibers
the tangent method overestimates the results. This is in agreement direction is overvalued with the incremental-tangent method, see
with what was stated by Wu et al. (2012): the accuracy of the Fig. 5(b), which explains why the von Mises stress in the matrix
incremental-tangent method decreases after strain softening of does not vanish. This is actually due to the unability of the incre-
the matrix as the fibers cannot see the elastic unloading which mental method to consider non-monotonic loadings as shown by
should arise due to the damaging process in the matrix. This is Wu et al. (2013).
clearly illustrated in Fig. 5(c) where the average stress along the Finally, Fig. 5(f) illustrates the damage evolution in the matrix.
loading direction in the fibers keeps increasing during the loading For the FE prediction, the minimum, average and maximum values
process. As expected, the new method does not suffer from this of the damage reached in the matrix are reported. It can be seen
limitation: the fibers are unloaded during the softening process, that the damage predicted by the MFH methods lies between the
see Fig. 5(c), and the method predicts the macro-stress with a high- average and maximum damage predicted by the FE simulations.
er accuracy, see Fig. 5(a). The average stress state in the matrix is This comes from our definition of the damage of the MFH, which
also predicted with a better accuracy as shown in Fig. 5(d). does not correspond to the average damage. During the decrease
Fig. 5(e) compares the average value of the effective equivalent of the strain to zero, the composite material is first unloaded and
von Mises stress in the matrix. For the FE simulations, the value is then enters into compression. This compression induces an in-
obtained by averaging the effective von Mises stress computed at crease of the accumulated plastic strain and of the damage (see
each integration point. For the first-order MFH schemes, by defini- Fig. 5(f)).
tion, this value is obtained as the equivalent von Mises value of the
average effective stress tensor in the matrix. Results obtained by 4.2. Comparison with experimental results
the three methods are comparable during the plastic flow, except
during the unloading in the vicinity of the point at which the com- The presented model is now applied to predict the macro-
posite material stress along the loading direction vanishes. At that mechanical response of a metal matrix composite material in ten-
stage, the stresses in the matrix and in the fibers phases are close sion. The considered material consists of SiC particles-reinforced
to zero for the FE simulations and for the incremental-secant MFH. AL–Si–Mg alloy (A356) with T61 ageing. The experimental results
As a consequence the effective von Mises stress in the matrix phase presented by Corbin and Wilkinson (1994) are used as the refer-
predicted by the incremental-secant MFH is also close to zero. As ence. The SiC spherical particules are modeled with a linear elastic
the stress distribution in the matrix phase of the direct FE simula- material without damage and their properties were reported by

Fig. 7. Sample with a notch. (a) Schematics of the sample test. (b) Force per unit ply length vs. displacement of the sample for different meshes.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 11

Fig. 8. The damage distribution in the matrix phase for the mesh – mesh size = 0.15 mm – obtained at ((a) and (b)) strain softening onset, ((c) and (d)) 0.825 mm-
displacement, and with the ((a) and (c)) incremental-secant MFH scheme, ((b) and (d)) incremental-tangent MFH scheme.

where p is the accumulated plastic strain and where the damage


4.38±0.05
evolution law is obtained through a curve fitting of the experimen-
tal data reported by Corbin and Wilkinson (1994) (Fig. 7 of this
reference).
25.3±0.2
The resulting experimental data obtained by Corbin and Wilkin-
50 160 son (1994) for a uni-axial tension on a material with v I ¼ 10% vol-
260 ume fraction of spherical inclusions are compared to the results
obtained by the presented MFH model in Fig. 6(a), where a good
Fig. 9. Geometry schematics of the tensile specimens (units in mm). agreement can be seen. The damage evolution in the matrix pre-
dicted by the homogenization model is illustrated in Fig. 6(b). Note
that in our model we assume that the particles are uniformly dis-
Christman et al. (1989). The constitutive equation of the matrix tributed, which is not always the case in real materials but remains
material follows a Ramburg–Osgood equation, reasonable for v I ¼ 10% volume fraction of inclusions.
 1=N
r r r
e¼ þa 0 ; ð77Þ
E E r0 5. Applications

where a rE0 ¼ 0:002 according to the definition of the yielding stress, In this section, the developed non-local MFH model is applied to
E is the elastic modulus, r0 is the yield strength defined at an offset study meso- and macro-scale problems. At first the convergence of
of 0:2%, and where N is the strain hardening exponent. The matrix the method with respect to the mesh size upon strain softening is
material experiences a linear plastic damage evolution according to demonstrated on a notched composite ply. Then, a composite lam-
the experimental result obtained by Corbin and Wilkinson (1994) inate is studied. In a first set of studies the predictions of the dam-
on the un-reinforced A356–T61 alloy. The material parameters are age-enhanced MFH model are compared with experimental results
on coupon tensile tests. These tests include an unloading stage to
 Inclusions: EI ¼ 450 GPa and mI ¼ 0:17. validate the damage model. Then the same laminate is used on a
 Matrix: E0 ¼ 70 GPa; m0 ¼ 0:33; r0 ¼ 220 MPa; N ¼ 0:1, and specimen with a hole. In this last example it can be seen that the
damage law D ¼ 0:875p, damage-enhanced MFH captures the damage path oriented along

Fig. 10. Mesh of the ½45 2 =  45 2 S laminate. Arrows indicate the grips of the extensometer.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
12 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

Fig. 11. Tensile stress–strain results predicted with the considered elasto-plastic model with damage for the epoxy matrix. (a) Stress–strain curve, the linear response, the
tensile strength, and the maximum tensile strain given by the manufacturer are reported in dotted lines. (b) Evolution of the accumulated plastic strain. (c) Evolution of the
damage.

Fig. 12. Results for 60% volume UD transverse-isotropic fibers reinforced epoxy tensile test with unloading and reloading on the ½45 2 =  45 2 S stacking sequence. The
experimental strain corresponds to the average strain value measured by a 50-mm long extensometer. The average values obtained for the experimental tests are reported
with their discrepancy range and are compared to the numerical curves.

the fibers directions in each ply, accordingly to experimental (2012) for comparison purpose. The fibers are assumed to be per-
results. pendicular to this section. Thus plane-strain conditions are applied
in the fiber direction (z-direction), and traction-free conditions are
5.1. Notched sample applied along the ply-thickness (y-direction).
Because of the symmetry, only one quarter of the structure is
In this section, the two-dimensional example proposed by Wu considered in the finite element simulations. The mesh of the quar-
et al. (2012), the authors, to study the convergence of the non-local ter specimen is refined in the weak zone of 8:0  8:0 mm (see the
damage-enhanced MFH method with respect to the mesh size, is shadow area in Fig. 7(a)), and four meshes are successively
considered. In particular the effect of the mesh size upon strain considered:
softening is analyzed using the developed incremental-secant
MFH formulation. Results are also compared to the ones previously  Coarsest mesh of 182 elements and with a mesh size of about
obtained by Wu et al. (2012) with the incremental-tangent MFH 0.43 mm at the notch;
method.  Intermediate mesh of 360 elements and with a mesh size of
The specimen, whose geometry is illustrated in Fig. 7(a), repre- about 0.3 mm at the notch;
sents the transverse section of a unidirectional reinforced ply. In  Fine mesh of 1120 elements and with a the mesh size of about
order to study the effect of the non-local damage model, notches 0.15 mm at the notch.
are added to force strain localization. The same material system  Finest mesh of 2540 elements and with a the mesh size of about
as in Section 4.1 is considered herein but with a fiber volume ratio 0.1 mm at the notch.
v I ¼ 0:3 and with the characteristic length l2 = 2.0 mm2. It should
be noted that these values are the ones studied by Wu et al. Bi-linear quadrangular elements with locking control are used.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 13

Fig. 13. Model of the ½45 2 =45 2 S -laminate with a hole. Units are expressed in mm, and the total thickness is 4.68 mm.

totally unloaded but not the inclusion phase explaining the kink
in Fig. 7(b). The incremental-secant method does not suffer from
this limitation and allows a complete unloading of the homoge-
nized material to be achieved.
This stress overestimation predicted by the incremental-tan-
gent method is also observed when analyzing the matrix damage
distributions obtained for the fine mesh with the two methods at
strain softening onset, Fig. 8(a) and (b), and at 0.825 mm-displace-
ment, Fig. 8(c) and (d). The damage reached with the incremental-
secant method is higher than with the incremental-tangent meth-
Fig. 14. Geometry schematics of the specimens used during experiments. od, especially during the strain softening part of the curve.

5.2. Tensile tests on laminates

In this section, the anisotropic gradient enhanced MFH model is


used to investigate the response of laminates made of a unidirec-
tional (UD) ½45 2 =45 2 S stacking sequence, and subjected to a
uni-axial tension. The geometry of the sample is illustrated in
Fig. 9.
In order to validate the damage model, experiments are con-
ducted in which the samples are loaded up to 0:56%-strain5 before
being unloaded and then reloaded a second time to reach the same
stress level. The specimens of carbon fibers reinforced epoxy com-
posite are manufactured from prepreg Hexply M10.1/38%/UD300/
HS (R), with a ½45 2 =45 2 S stacking sequence. The total thickness
of the specimens is 4.38 ± 0.05 mm. After curing, the resulting fiber
volume fraction is mI ¼ 60%.6 The specimens are cut from an auto-
clave consolidated unidirectional laminate panel of 300  300 mm2 ,
and their geometry schematic is shown in Fig. 9. To prevent gripping
damage, aluminum tabs are glued at both extremities of each spec-
Fig. 15. Results for 60% volume UD transverse-isotropic fibers reinforced epoxy imen. The static tensile tests are carried out on a 1185 no
tensile tests on the ½45 2 =  45 2 S stacking sequence with a hole. (a) Stress–strain H4573(ME002) Instran machine in displacement control mode with
curve. The experimental strain corresponds to the average strain value measured by
a constant cross-head speed of 2 mm/min, according to the specifica-
a 50-mm long extensometer. We have also reported the solution predicted for
independent plies, meaning as if the plate was fully delaminated. (b) Damage tion of ISO-527-4 standard. An extensometer measures the average
evolution in the internal and external plies. strain in the 50-mm band.
A numerical model is made by meshing the different plies of the
laminate sample, and is illustrated in Fig. 10. Each ply orientation is
The load–displacement curves are extracted for the four mesh meshed by 915 bi-linear quadrangular elements with locking con-
sizes and are presented in Fig. 7(b). When considering both the trol. In each ply, the MFH model is used as a material law with the
incremental-tangent and the incremental-secant MFH, a good con- proper fiber orientations. Constrained displacements are applied
vergence can be seen with the decrease of the mesh size. Indeed, along the x-axis at both laminate extremities and the deformations
the difference between the curves for mesh sizes of 0.15 mm and are measured from the displacements at the nodes corresponding
0.3 mm is much smaller than the difference between the curves to the extensometer grips positions (arrows on Fig. 10).
for mesh sizes of 0.3 mm and 0.43 mm, despite a larger difference In this application, the following length scales are considered:
in the mesh reduction. The results for a mesh-size of 0.1 mm are c3 ¼ 2:0 mm2 as suggested by Geers et al. (1999) for short fiber
ever closer to the 0.15 mm mesh-size curve. composites along the fiber directions, and c1 ¼ c2 ¼ 8:45
As expected, with the incremental-tangent MFH scheme the 105 mm2 in the transverse direction. In this direction c1 ¼ c2 is
predicted strain softening onset is over-estimated compared to 2
equal to l2 , where l is the maximum distance of possible interaction
the results obtained with the incremental-secant MFH method,
which allows the fibers to be unloaded during the damaging pro- 5
Before failure of the specimens happens.
cess. Also with the incremental-tangent method, during the soften- 6
The percentage is a mean value obtained from a microscopic imaging process of
ing part for a displacement close to 1 mm, the matrix phase is the cured laminates.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
14 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

Fig. 16. Snapshots of the damage distribution (logarithmic scale) for an average 0.65% strain-(a) Outer ply, (b) Inner ply.

between material points (Peerlings et al., 2001). In our case, this 5.3. Laminate with a hole
distance corresponds to the distance between fibers, which are
obstacles to the interaction, and can be statistically computed from In the previous section, the localization of damage during the
the volume fraction of fibers (60%) and from the fibers diameter experiments was not controlled. In this section, the response of
(10 lm). Note that an artificial spreading of the damaged zone an open hole specimen made of the same UD ½45 2 =45 2 S stack-
orthogonal to the direction of a possible crack propagation has ing sequence, and subjected to a uni-axial tension is studied. The
been observed by Geers et al. (1998, 1999) in the numerical simu- presence of the hole allows comparing the localization effects hap-
lations involving damage to fracture transition with constant l, but pening during the experiments and predicted by the numerical
this is beyond the scope of this paper. model. The geometry of the sample is illustrated in Fig. 13(a). For
In order to conduct the numerical simulations, the material the numerical model, although the plate is not symmetrical locally
properties, of the different phases are identified as follows. The car- because of the material anisotropy, the global behavior is symmet-
bon fibers are assumed to be linear elastic and transversely isotro- rical on the width of the sample as the stacking sequence is bal-
pic. Typical material constants, e.g. (Byström, 2009) for T300 anced. Thus the numerical model considers only half of the plate
carbon fibers are considered. The cured epoxy matrix properties re- with 1280  8 elements, see Fig. 13(b). This approximation is valid
ported by the manufacturer are a tensile modulus of 3.2 GPa, and a as the local behavior of interest issues from the hole, which is not
tensile strength of 85 MPa at 0.035-strain. By lack of elasto-plastic on the ‘‘pseudo’’ symmetrical axis. Bi-linear quadrangular elements
data, an exponential hardening law (76) and a power damage law with locking control are used.
(21) are considered. The material properties used are thus In this application, the numerical and material parameters are
identical as the ones reported in Section 5.2. The tests are con-
 Inclusions: Longitudinal Young modulus EL ¼ 230 GPa, trans- ducted on specimens, whose geometry schematic is shown in
verse Young modulus ET ¼ 40 ðGPaÞ, transverse Poisson ratio Fig. 14, manufactured as described in Section 5.2. The tensile tests
mTT ¼ 0:20, longitudinal-transverse Poisson ratio mLT ¼ 0:256, also follow the same protocol.
transverse shear modulus GTT ¼ 16:7 ðGPaÞ, and longitudinal- Numerical predictions and experimental results are compared
transverse shear modulus GLT ¼ 24 ðGPaÞ. in Fig. 15(a). The stress evolution, evaluated by dividing the ten-
 Matrix: E0 ¼ 3:2 GPa; m0 ¼ 0:3; rY0 ¼ 15 MPa; h0 ¼ 300 MPa; sile load by the cross-section area, is in excellent agreement up to
m0 ¼ 100; S0 ¼ 0:1 MPa; s ¼ 2, and pC ¼ 0. 0.2% deformation. Upon this state the stress is over-predicted,
with a maximum discrepancy near the fracture onset, see
The corresponding stress–strain curve of the epoxy matrix is illus- Fig. 15(a). The difference between the numerical prediction and
trated on Fig. 11(a), where the manufacturer data are also reported. the experimental results comes from the other damaging mecha-
With the considered law, it was not possible to reach the maxi- nisms such as delamination, fiber pull-out etc., which are not ta-
mum stress at an exact 0.035-strain, and the maximum strain is ken into account in the simulation. To illustrate this point we
reached during the softening part. Moreover, to account for the have also reported the solution predicted for independent plies
non-local framework for which the damage is computed from a on Fig. 15(a), meaning as if the plate was fully delaminated,
non-local accumulated plastic strain, the matrix law was fitted and it can be seen that the experimental results get closer from
using a damage threshold at pC =0.0025 instead of the used value this last curve during the tensile tests. Finally the damage evolu-
of 0. From the damage evolution on Fig. 11(c), it can be seen that tion is reported in Fig. 15(b), while Fig. 16 exhibits the damage
softening onset is reached for a damage value around 0.032. distributions in the external and internal plies. It can be seen that
Five samples are tested and their stress–strain curves are com- the maximum damage location of the numerical predictions is in
pared to the numerical predictions. The first loading strain–stress good agreement with the crack initiation location in the different
curves and the curves extracted from the unloading/second load- plies observed for the experimental results, see Fig. 17. Moreover,
ing7 are reported in Fig. 12(a) and (b). It can be seen that the model the damage bands propagate along directions parallel to the fi-
is in excellent agreement with the experiments. In particular, the bers orientations and the damage reaches a value close to one
MFH model with damage allows predicting the unloading with a when fracture is experimentally observed. Note that experimen-
high accuracy, and the residual strain is in perfect agreement with tally the cracks propagate with a 45-degree angle in the outer
the experiments, as it can be observed on Fig. 12(b). plies (Fig. 17) and with a 45-degree angle in the inner plies. These
Remark that for these tests, the values of the damage reached two orientations can be seen simultaneously in the numerical
are low because the geometry is perfectly regular and because predictions under the form of damage bands in both the inner
the non-local approach avoids spurious localization. For a real and in the outer plies (as the displacements are continuous at
geometry the imperfection will cause a shear band. In order to plies interfaces), see Fig. 16, but the high damage concentrations
demonstrate that the approach can initiate such a band we con- (close to one) follow the respective 45-degree direction in the
sider in the next section an application in which a strain gradient outer plies, see Fig. 16(a), and the 45-degree direction in the inner
naturally exists. plies, see Fig. 16(b).

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 15

Fig. 17. Broken open hole sample. The cracks follow the
45 -directions in each ply and originate from the predicted locations.

6. Conclusions  Dyadic products are designated by :

ðu  v Þij ¼ ui v j ; ða  bÞijkl ¼ aij bkl : ðA:2Þ


In this work, a new incremental-secant MFH process for
composites made of elasto-plastic constituents exhibiting dam-  Symbols 1 and I designate the second- and fourth-order sym-
age was proposed. In this approach, an unloading of the com- metric identity tensors respectively:
posite material is virtually performed to estimate the residual
strains in each phase before applying a secant approach on 1
1ij ¼ dij ; I ijkl ¼ ðdik djl þ dil djk Þ; ðA:3Þ
the strain increments, which differ in each phase. In order to 2
define the LCC two secant operators were defined. The first
where dij ¼ 1 if i ¼ j; dij ¼ 0 if i – j.
one, the residual-incremental-secant operator, is defined from
 The spherical and deviatoric operators are I vol and I dev
the phase residual stress. The second operator, the zero-incre-
respectively:
mental-secant operator, is defined from a stress-free state in
the phase. 1
I vol 1  1; I dev ¼ I  I vol ; ðA:4Þ
The main advantage of this secant approach is made obvious 3
when considering a composite whose matrix phase exhibits a dam-
aging process. In this case, the inclusions phase can be unloaded so that for symmetric tensors aij ¼ aji we have:
during the softening stage of the matrix, ensuring an accurate pre- 1 1
diction of the scheme. In particular, it was shown that for a com- I vol : a ¼ amm 1; I dev : a ¼ a  amm 1 ¼ devðaÞ: ðA:5Þ
3 3
posite with 50% volume ratio of fibers the new incremental-
secant-method is much more accurate than the incremental-tan-
gent approach previously developed. Appendix B. Derivation of the closed-form expressions for the
As classical FE formulations lose solution uniqueness and face incremental-secant method
the strain localization problem when strain softening of materials
is involved, the model was formulated in a so-called implicit B.1. Incremental-secant operator C S
non-local approach. It was shown in a numerical example that
the numerical simulations converge with the mesh size during During MFH process, in order to identify j and ls – standing for
the strain softening response. either jr and lrs or j0 and l0s – the elastic bulk and shear moduli of
Finally the model has been applied to study the response of the virtual elastic material, the increments of the stress and strain
laminates. First the material model has been validated by consider- tensors are decomposed into their hydrostatic and deviatoric parts:
ing tensile tests including unloading on coupons. Then specimens
Dr ¼ Drm 1 þ Ds and De ¼ Dem 1 þ De; ðB:1Þ
made of the same laminate, but with a hole forcing the localization,
have been considered. It was shown that the predicted damage where Drm ¼ 13 trðDrÞ; Ds ¼ Dr  Drm 1; Dem ¼ 13 trðDeÞ, and
evolution follows the fibers direction accordingly to the experi- De ¼ De  Dem 1. Note that Dr ¼ r when the zero-incremental-se-
mental results. cant operator C S0 is used in the MFH process.
In the future it is intended to account for more degradation pro- Following the developments of Wu et al. (2013), the elastic bulk
cesses, such as delamination, but also to formulate the incremen- modulus j ¼ jel remains constant because of the incompressible
tal-secant method in a second-statistical-moment framework to nature of the plastic flow, and
improve its accuracy even further.
Dreq
3ls ¼ ; ðB:2Þ
Acknowledgment Deeq
with
The research has been funded by the Walloon Region under the
agreement SIMUCOMP no 1017232 (CT-EUC 2010-10-12) in the  1=2  1=2
3 2
context of the ERA-NET+, Matera + framework. Dreq ¼ Ds : Ds and Deeq ¼ De : De : ðB:3Þ
2 3

Appendix A. Tensorial operations and notations The derivative of the tensor – with C S standing for either C Sr or
S0
C – is expressed as
 Dots and colons are used to indicate tensor products contracted " #
@C S 1 alg 2 De
over one and two indices respectively: ¼ 2I dev  D s : C  l ; ðB:4Þ
@e 6ls ðDeeq Þ2 3 s ðDeeq Þ2
u  v ¼ ui v i ; ða  uÞi ¼ aij uj ;
ða  bÞij ¼ aik bkj ; a : b ¼ aij bji ; where the ‘‘consistent’’ operator C alg is the derivative of the stress
ðC : aÞij ¼ C ijkl alk ; ðC : DÞijkl ¼ C ijmn Dnmkl : ðA:1Þ increment with respect to the strain increment, which is obtained
from the constitutive law of the material, see Appendix C.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
16 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

 el 2
B.2. Damaged incremental-secant operator C SD 2l
C alg ¼ C el  NN
@C SD
h
In order to apply the MFH scheme we need to know @e
-with   2  
C standing for either C SDr or C SD0 . One has directly
SD 2lel Dp 3 dev
  tr eq I NN : ðC:7Þ
r^ nþ1  r^ res
n
2
SD S
@C @C @D
¼ ð1  DÞ  CS  ; ðB:5Þ For the zero-incremental-secant approach presented in Sec-
@e @e @e
S tion 3.2.2 the same way of proceeding results in
where @C@e
has the same expression as in Appendix B.1, and where @D
@e
2 2  
is given in Appendix C. ð2lel Þ ð2lel Þ Dp 3 dev
~ is used during the damage evalua-
As the non-local variable p C alg ¼ C el  N  N   tr eq I NN ; ðC:8Þ
h0 r^ nþ1 2
tion, we also have to determine
with h0 ¼ 3lel þ dR > 0 and N ¼ N 0 . In this case the direction of the
@C SD @D S dp
normal corresponds strictly to the radial return mapping assump-
¼ C ; ðB:6Þ
~
@p ~
@p tion, and the classical expression of C alg is recovered, e.g. (Doghri,
where @D
is given in Appendix C. 2000, chapter 12).
~
@p
Moreover, it can be easily deduced from (21) that
Appendix C. Linearization of the Lemaitre–Chaboche ductile @Y @ ee
: : de ¼ ee : C alg : de; ðC:9Þ
damage model in the non-local form @ ee @ e
leading to
During a finite incremental process the constitutive equations
are discretized in time intervals [t n ; t nþ1 ] and are differentiated at @ DD @Y @ ee @ DD ðY Þs1 @Y
dDðe; p
~Þ : : d e þ d p ~ nþas
~ ¼ s Dp
tnþ1 . Remembering that for the non-local formulation the damage @Y @ ee @ e @p~ S0 @ ee
depends on both e and on p ~, this linearization reads e
 s s1
  @e Y nþa ðY Þ
@D @D : : de þ ~ ¼ asDp
dp ~ nþas ee : C alg
drnþ1 ¼ C algD : denþ1  r^ nþ1  : denþ1  r
^ nþ1 ~ ;
dp ðC:1Þ @e S0 S0
@e ~ nþ1
@p  s
Y nþa
: de þ ~:
dp ðC:10Þ
where S0
C algD ¼ ð1  Dnþ1 ÞC alg : ðC:2Þ
In these last expressions, C ¼ alg
¼ @r
^ nþ1 @r^ nþ1
is the derivative of Appendix D. Stress residual vector
@e @ Der
the effective stress increment with respect to the strain incre-
ment,8 which ensures quadratic convergence, as shown by Simo The equation to be satisfied at the
 end of the MFH procedure is
and Taylor (1985). Eq. (17). Multiplying Eq. (15) by B I;C
 SD ; C
0
 S and using (17) lead to
I

This expression is now derived for the residual-incremental-se- v 0 DerI


nþ1
þ v I B ðI;C SD S r   SD  S r
0 ; C I Þ : DeInþ1 ¼ B ðI;C 0 ; C I Þ : Denþ1 : ðD:1Þ
cant approach presented in Section 3.2.1. In particular the effect of
the assumption in the plastic flow direction is accounted for. From With the M–T assumption the strain concentration tensor follows
Eqs. (29) and (30) one has directly from (4), and Eq. (D.1) reads
h 1 i
@ Dp @N DerInþ1 þ v 0 S : C SD : C SI  I : DerInþ1 ¼ Dernþ1 ; ðD:2Þ
C alg ¼ C el  2lel N   2lel Dp : ðC:3Þ 0
@e @e
eq dev or again F ¼ 0 with
On the one hand, using @a@a ¼ 32 aaeq , the two Eqs. (33) and (34) of the
 
system related to the yield surface evolution lead to 1 1  r
F ¼ C SD r
0 : DeInþ1  S : DeInþ1  Dernþ1  C SI : DerInþ1 : ðD:3Þ
 dev 2 !1 3 v0
3 r^ nþ1  r^ res ^ nþ1 Þdev
1 3 ðr @R 5
 n
eq : 4 dp þ 3lel dp In order to satisfy F ¼ 0, Eq. (D.3) is linearized as9
2 r ^ nþ1  r
^ res
n
^ nþ1 Þeq
3 2 ðr @p
 tr dev @F @F @F @F
3 r
^ r ^ res dF ¼ : dDerI þ : dDer0 þ : dDer þ ~:
dp ðD:4Þ
¼  nþ1 n el @ eI @ e0 De ~
@p
 : C : de; ðC:4Þ
2 r
^ tr ^ res eq
nþ1  rn When solving F ¼ 0 at constant Der and constant p ~, as v 0 Der0 þ v I DerI
which can dev be rewritten, using Eqs. (31) and (32) and calling is also constant, the iteration process relies on dF ¼ J : deI with
ðr^ Þ
N 0 ¼ 32 ^nþ1 eq the normal to the yield surface, as
ðrnþ1 Þ @F @F @ e0  SD h i @ C SInþ1
J¼ þ : ¼ C 0nþ1 : I  S 1  C SInþ1 
@p 2lel @ eI @ e0 @ eI @ eI
¼ N; ðC:5Þ !
@e h v @  SD
C @  SD @D
C
 1 : DerInþ1 
I 0 nþ1
þ
0 nþ1
@adev
with h ¼ 13 N : N 0 @R
@p
þ 3lel . On the other hand, using @a
¼ I dev , v 0 @ e0 @D @ e0
Eq. (32) yields 2  3
DerInþ1  Dernþ1
@N 3 I dev : C el 2lel N  N : 4Der  S : 1 5  v I C SD  ðDer  Dernþ1 Þ
¼  eq   eq : ðC:6Þ Inþ1
v0 v 20 0nþ1 Inþ1
@ e 2 C : Der
el
C el : Dernþ1  
nþ1
@S @S @D vI
:: ðS 1  S 1 Þ :: þ  C SD : S 1 ; ðD:5Þ
Finally, combining Eqs. (C.3), (C.5) and (32) leads to the final expres- @ e0 @D @ e0 v 0 0nþ1
sion of the ‘‘consistent’’ operator

8
In this non-local formalism, we explicitly write the dependence of the stress in D
and e and linearize with respect to both terms. We use the symbol C alg for the 9
Note that the derivative with respect to Derr has the same expression as the
derivative of the effective stress increment with respect to the strain increment. derivative with respect to er .

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx 17

S  SD  SD
@C  SD
@C
@C dC 0 0
where results from Eq. (B.4), and where
I
@ eI
¼ @ enþ1
0
þ @Dnþ1 @@D
de0
0
e0 re- When damage is considered, the secant moduli with damage need
sults from Eq. (B.5). The derivative of the Eshelby tensor to be used (see Eqs. (44)–(49)) to compute the Eshelby tensor,
 
@S
¼ @@Se0 þ @D
@S @D
is reported in Appendix F. and its derivative reads
@ e0 @ e0
 
Once F ¼ 0 is satisfied, the effect on the strain increment in each @S @S @ m @ jD @ m @ lDs
phase of a variation dDer at constant Dp
~ can directly be obtained by ¼  þ
@ Der @ m @ jD @ Der @ lDs @ Der
constraining dF ¼ 0, and Eq. (D.4) becomes   D   
@S @m @j @ jD @D @ m @ lDs @ lDs @D
¼  þ þ þ : ðF:4Þ
@F @F @F @m @ jD @ Der @D @ Der @ lDs @ Der @D @ Der
0¼ : dDerI þ : dDer0 þ : dDer ðD:6Þ
@ eI @ e0 @ e
or again
References
@ eI @F
¼ J 1 : : ðD:7Þ Aboudi, J., Pindera, M.J., Arnold, S., 2003. Higher-order theory for periodic
@ e @ e
multiphase materials with inelastic phases. Int. J. Plast. 19, 805–847.
As under these circumstances der ¼ v 0 der0 þ v I derI , this last equation Aifantis, E., 1992. On the role of gradients in the localization of deformation and
fracture. Int. J. Eng. Sci. 30, 1279–1299.
is completed by Bažant, Z.P., Belytchko, T.B., Chang, T.P., 1984. Continuum theory for strain-
  softening. J. Eng. Mech. ASCE 110, 1666–1692.
@ e0 1 @ eI
¼ I  v1 : ðD:8Þ Benveniste, Y., 1987. A new approach to the application of Mori–Tanaka’s theory in
@ e v0 @ e composite materials. Mech. Mater. 6, 147–157.
Berveiller, M., Zaoui, A., 1978. An extension of the self-consistent scheme to
The same relations can be obtained for a linearization with re- plastically-flowing polycrystals. J. Mech. Phys. Solids 26, 325–344.
~ at constant Der .
spect to p Brassart, L., Stainier, L., Doghri, I., Delannay, L., 2011. A variational formulation for
the incremental homogenization of elasto-plastic composites. J. Mech. Phys.
Solids 59, 2455–2475.
Appendix E. ‘‘Consistent’’ linearization of the homogenized Brassart, L., Stainier, L., Doghri, I., Delannay, L., 2012. Homogenization of elasto-
stress (visco) plastic composites based on an incremental variational principle. Int. J.
Plast. 36, 86–112.
Byström, J., 2009. Optimal design of a long and slender compressive strut. Int. J.
The ‘‘consistent’’ linearization of the homogenized stress can be Multiphys. 3, 235–257.
rewritten using Eqs. (57) and (C.1) for the homogenized material, Carrere, N., Valle, R., Bretheau, T., Chaboche, J.L., 2004. Multiscale analysis of the
transverse properties of ti-based matrix composites reinforced by sic fibres:
leading to from the grain scale to the macroscopic scale. Int. J. Plast. 20, 783–810.
  Chaboche, J., Kanouté, P., Roos, A., 2005. On the capabilities of mean-field
@D
 ¼ v I drI þ v 0 dr0 ¼ v I C alg
dr I : deI þ v 0 C algD
0 r^0  approaches for the description of plasticity in metal matrix composites. Int. J.
@ e0 Plast. 21, 1409–1434.
@D Christman, T., Needleman, A., Nutt, S., Suresh, S., 1989. On microstructural evolution
: de0  v 0 r ^0 ~
dp and micromechanical modelling of deformation of a whisker-reinforced metal-
@p ~ matrix composite. Mater. Sci. Eng.: A 107, 49–61 (Proceedings of the
   
@ eI @ eI @D Symposium on Interfacial Phenomena in Composites: Processing
¼ v I C alg
I : : de
 þ dp~ þ v 0 C algD
0  r
^ 0  Characterization and Mechanical Properties).
@ e @p~ @ e0 Coenen, E., Kouznetsova, V., Geers, M., 2011. Enabling microstructure-based
 
@ e0 @ e0 @D damage and localization analyses and upscaling. Model. Simul. Mater. Sci.
: : de þ dp~  v 0r ^0 ~;
dp ðE:1Þ Eng. 19, 074008.
@ e @p~ ~
@p Coenen, E., Kouznetsova, V., Geers, M., 2012. Novel boundary conditions for strain
localization analyses in microstructural volume elements. Int. J. Numer.
with, see Appendix D, Methods Eng. 90, 1–21.
  Corbin, S., Wilkinson, D., 1994. Influence of matrix strength and damage
@ eI @F @ e0 1 @ eI
¼ J 1 : ; ¼ I  v1 ; ðE:2Þ accumulation on the mechanical response of a particulate metal matrix
@ e @ e @ e v0 @ e composite. Acta Metall. Mater. 42, 1329–1335.
Dascalu, C., 2009. A two-scale damage model with material length. CR Méc. 337,
@ eI @F @ e0 v I @ eI
¼ J 1 : ; and ¼ : ðE:3Þ 645–652.
@p~ ~
@p @p~ v 0 @ p~ De Borst, R., 1991. Simulation of strain localization: a reappraisal of the Cosserat
continuum. Eng. Comput. 8, 317–332.
Thus, the consistent tangent operators read De Borst, R., Sluys, L., Mühlhaus, H.B., Pamin, J., 1993. Fundamental issues in finite
  element analyses of localization of deformation. Eng. Comput. 10, 99–121.
alg @ eI @D @ e0 Doghri, I., 1995. Numerical implementation and analysis of a class of metal
C alg
nþ1 ¼ v I C I : þ v 0 C algD
0  r
^ 0  : ; ðE:4Þ
@ e @ e0 @ e plasticity models coupled with ductile damage. Int. J. Numer. Methods Eng. 38,
3403–3431.
  Doghri, I., 2000. Mechanics of Deformable Solids- Linear, Nonlinear, Analytical and
@ eI @D @ e0 @D Computational Aspects. Springer-Verlag, Berlin.
C p~ ¼ v I C alg
I : ~ þ v 0 C algD
dp 0  r
^ 0  :  v0r
^0 : ðE:5Þ
@p~ @ e0 @p~ ~
@p Doghri, I., Adam, L., Bilger, N., 2010. Mean-field homogenization of elasto-
viscoplastic composites based on a general incrementally affine linearization
method. Int. J. Plast. 26, 219–238.
Doghri, I., Brassart, L., Adam, L., Gérard, J.S., 2011. A second-moment incremental
Appendix F. Eshelby tensor and its derivative formulation for the mean-field homogenization of elasto-plastic composites.
Int. J. Plast. 27, 352–371.
The derivative of the Eshelby tensor can be written as Doghri, I., Ouaar, A., 2003. Homogenization of two-phase elasto-plastic composite
materials and structures: study of tangent operators, cyclic plasticity and
 
@S @S @m @j @ m @ ls numerical algorithms. Int. J. Solids Struct. 40, 1681–1712.
¼  þ : ðF:1Þ Doghri, I., Tinel, L., 2005. Micromechanical modeling and computation of elasto-
@ Der @ m @ j @ Der @ ls @ Der plastic materials reinforced with distributed-orientation fibers. Int. J. Plast. 21,
1919–1940.
If damage is not considered, we have Engelen, R.A.B., Geers, M., Baaijens, F., 2003. Nonlocal implicit gradient-enhanced
elasto-plasticity for the modelling of softening behaviour. Int. J. Plast. 19, 403–
@j 433.
¼ 0; ðF:2Þ
@ Der Eshelby, J.D., 1957. The determination of the elastic field of an ellipsoidal inclusion,
and related problems. Proc. R. Soc. London Ser. A, Math. Phys. Sci. 241, 376–396.
therefore, Geers, M., 1997. Experimental Analysis and Computational Modelling of Damage
and Fracture. Ph.D. Thesis. University of Technology, Eindhoven (Netherlands).
@S @S @ m @ ls Geers, M., de Borst, R., Brekelmans, W., Peerlings, R., 1999. Validation and internal
¼  : ðF:3Þ
@ Der @ m @ ls @ Der length scale determination for a gradient damage model: application to short
glass-fibre-reinforced polypropylene. Int. J. Solids Struct. 36, 2557–2583.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022
18 L. Wu et al. / International Journal of Solids and Structures xxx (2013) xxx–xxx

Geers, M., Kouznetsova, V., Brekelmans, A., 2010. Multi-scale computational Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of
homogenization: trends and challenges. J. Comput. Appl. Math. 234, 2175– materials with misfitting inclusions. Acta Metall. 21, 571–574 (Cited by (since
2182. 1996) 1814).
Geers, M.G.D., de Borst, R., Brekelmans, W.A.M., Peerlings, R.H.J., 1998. Strain-based Moulinec, H., Suquet, P., 2003. Intraphase strain heterogeneity in nonlinear
transient-gradient damage model for failure analyses. Comput. Methods Appl. composites: a computational approach. Eur. J. Mech. A/Solids 22, 751–770.
Mech. Eng. 160, 133–153. Peerlings, R., de Borst, R., Brekelmans, W., Ayyapureddi, S., 1996. Gradient-enhanced
Hill, R., 1965a. Continuum micro-mechanics of elastoplastic polycrystals. J. Mech. damage for quasi-brittle materials. Int. J. Numer. Methods Eng. 39, 3391–3403.
Phys. Solids 13, 89–101. Peerlings, R., de Borst, R., Brekelmans, W., Geers, M., 1998. Gradient-enhanced
Hill, R., 1965b. A self-consistent mechanics of composite materials. J. Mech. Phys. damage modelling of concrete fracture. Mech. Cohes. Frict. Mater. 3, 323–342.
Solids 13, 213–222. Peerlings, R., Geers, M., de Borst, R., Brekelmans, W., 2001. A critical comparison of
Ji, B., Wang, T., 2003. Plastic constitutive behavior of short-fiber/particle reinforced nonlocal and gradient-enhanced softening continua. Int. J. Solids Struct. 38,
composites. Int. J. Plast. 19, 565–581. 7723–7746.
Kanouté, P., Boso, D., Chaboche, J., Schrefler, B., 2009. Multiscale methods for Pettermann, H.E., Plankensteiner, A.F., Böhm, H.J., Rammerstorfer, F.G., 1999. A
composites: a review. Arch. Comput. Methods Eng. 16, 31–75. http://dx.doi.org/ thermo-elasto-plastic constitutive law for inhomogeneous materials based on
10.1007/s11831-008-9028-8. an incremental Mori–Tanaka approach. Comput. Struct. 71, 197–214.
Knockaert, R., Doghri, I., 1999. Nonlocal constitutive models with gradients of Pierard, O., Doghri, I., 2006a. An enhanced affine formulation and the corresponding
internal variables derived from a micro/macro homogenization procedure. numerical algorithms for the mean-field homogenization of elasto-viscoplastic
Comput. Methods Appl. Mech. Eng. 174, 121–136. composites. Int. J. Plast. 22, 131–157.
Kouznetsova, V., Geers, M., Brekelmans, W., 2004. Multi-scale second-order Pierard, O., Doghri, I., 2006b. Study of various estimates of the macroscopic tangent
computational homogenization of multi-phase materials: a nested finite operator in the incremental homogenization of elastoplastic composites. Int. J.
element solution strategy. Comput. Methods Appl. Mech. Eng. 193, 5525– Multiscale Comput. Eng. 4, 521–543.
5550 (Adv. Comput. Plasticity). Pierard, O., LLorca, J., Segurado, J., Doghri, I., 2007. Micromechanics of particle-
Kouznetsova, V., Geers, M.G.D., Brekelmans, W.A.M., 2002. Multi-scale constitutive reinforced elasto-viscoplastic composites: finite element simulations versus
modelling of heterogeneous materials with a gradient-enhanced computational affine homogenization. Int. J. Plast. 23, 1041–1060.
homogenization scheme. Int. J. Numer. Methods Eng. 54, 1235–1260. Ponte Castañeda, P., 1991. The effective mechanical properties of nonlinear
Kröner, E., 1958. Berechnung der elastischen konstanten des vielkristalls aus den isotropic composites. J. Mech. Phys. Solids 39, 45–71.
konstanten des einkristalls. Z. Phys. A Hadron Nucl. 151, 504–518. http:// Ponte Castañeda, P., 1992. A new variational principle and its application to
dx.doi.org/10.1007/BF01337948. nonlinear heterogeneous systems. SIAM J. Appl. Math. 52, 1321–1341.
Lahellec, N., Ponte Castañeda, P., Suquet, P., 2011. Variational estimates for the Ponte Castañeda, P., 1996. Exact second-order estimates for the effective
effective response and field statistics in thermoelastic composites with intra- mechanical properties of nonlinear composite materials. J. Mech. Phys. Solids
phase property fluctuations. Proc. R. Soc. A: Math. Phys. Eng. Sci. 467, 2224– 44, 827–862.
2246. Ponte Castañeda, P., 2002a. Second-order homogenization estimates for nonlinear
Lahellec, N., Suquet, P., 2007a. On the effective behavior of nonlinear inelastic composites incorporating field fluctuations. I: Theory. J. Mech. Phys. Solids 50,
composites. I. Incremental variational principles. J. Mech. Phys. Solids 55, 1932– 737–757.
1963. Ponte Castañeda, P., 2002b. Second-order homogenization estimates for nonlinear
Lahellec, N., Suquet, P., 2007b. On the effective behavior of nonlinear inelastic composites incorporating field fluctuations. II: Applications. J. Mech. Phys.
composites. II. A second-order procedure. J. Mech. Phys. Solids 55, 1964–1992. Solids 50, 759–782.
Lahellec, N., Suquet, P., 2013. Effective response and field statistics in elasto-plastic Segurado, J., Llorca, J., González, C., 2002. On the accuracy of mean-field approaches
and elasto-viscoplastic composites under radial and non-radial loadings. Int. J. to simulate the plastic deformation of composites. Scr. Mater. 46, 525–529.
Plast. 42, 1–30. Simo, J.C., Taylor, R.L., 1985. Consistent tangent operators for rate-independent
Lemaitre, J., 1985. Coupled elasto-plasticity and damage constitutive equations. elastoplasticity. Comput. Methods Appl. Mech. Eng. 48, 101–118.
Comput. Methods Appl. Mech. Eng. 51, 31–49. Suquet, P., 1995. Overall properties of nonlinear composites: a modified secant
Lemaitre, J., Chaboche, J.L., 1991. Mécanique des Matériaux solides. Dunod. moduli theory and its link with Ponte Castañeda’s nonlinear variational
Lemaitre, J., Desmorat, R., 2005. Engineering Damage Mechanics: Ductile, Creep, procedure. CR Acad. Sci. 320, 563–571.
Fatigue and Brittle Failures. Springer-Verlag, Berlin. Svedberg, T., Runesson, K., 1997. A thermodynamically consistent theory of
Lissenden, C., Arnold, S., 1997. Theoretical and experimental considerations in gradient-regularized plasticity coupled to damage. Int. J. Plast. 13, 669–696.
representing macroscale flow/damage surfaces for metal matrix composites. Talbot, D., Willis, J., 1992. Some simple explicit bounds for the overall behaviour of
Int. J. Plast. 13, 327–358. nonlinear composites. Int. J. Solids Struct. 29, 1981–1987.
Liu, X., Hu, G., 2005. A continuum micromechanical theory of overall Talbot, D.R.S., Willis, J.R., 1985. Variational principles for inhomogeneous non-linear
plasticity for particulate composites including particle size effect. Int. J. Plast. media. IMA J. Appl. Math. 35, 39–54.
21, 777–799. Talbot, D.R.S., Willis, J.R., 1987. Bounds and self-consistent estimates for the overall
LLorca, J., González, C., Molina-Aldareguía, J.M., Segurado, J., Seltzer, R., Sket, F., properties of nonlinear composites. IMA J. Appl. Math. 39, 215–240.
Rodríguez, M., Sádaba, S., Muñoz, R., Canal, L.P., 2011. Multiscale modeling of Wieckowski, Z., 2000. Dual finite element methods in homogenization for elastic–
composite materials: a roadmap towards virtual testing. Adv. Mater. 23, 5130– plastic fibrous composite material. Int. J. Plast. 16, 199–221.
5147. Wu, L., Noels, L., Adam, L., Doghri, I., 2012. Multiscale mean-field homogenization
Massart, T., Peerlings, R., Geers, M., 2005. A dissipation-based control method for fiber-reinforced composites with gradient-enhanced damage
method for the multi-scale modelling of quasi-brittle materials. CR Méc. 333, model. Comput. Methods Appl. Mech. Eng. 233–236, 164–179.
521G527. Wu, L., Noels, L., Adam, L., Doghri, I., 2013. Non-local damage-enhanced mfh for
Massart, T., Peerlings, R., Geers, M., 2007. An enhanced multi-scale approach for multiscale simulations of composites. Composite Materials and Joining
masonry wall computations. Int. J. Numer. Methods Eng. 69, 1022–1059. Technologies for Composites, vol. 7. Springer (chapter 13).
Masson, R., Bornert, M., Suquet, P., Zaoui, A., 2000. An affine formulation for the Wu, L., Noels, L., Adam, L., Doghri, I., (2013). A combined incremental-secant mean-
prediction of the effective properties of nonlinear composites and polycrystals. field homogenization scheme with per-phase residual strains for elasto-plastic
J. Mech. Phys. Solids 48, 1203–1227. composites. Int. J. Plast. in press. http://dx.doi.org/10.1016/j.ijplas.2013.06.006.
Mercier, S., Molinari, A., 2009. Homogenization of elastic viscoplastic heterogeneous Zaoui, A., Masson, R., 2002. Modelling stress-dependent transformation strains of
materials: self-consistent and Mori–Tanaka schemes. Int. J. Plast. 25, 1024– heterogeneous materials. In: Bahei-El-Din, Y.A., Dvorak, G.J., Gladwell, G.M.L.
1048. (Eds.), IUTAM Symposium on Transformation Problems in Composite and Active
Molinari, A., Canova, G., Ahzi, S., 1987. A self consistent approach of the large Materials, Solid Mechanics and Its Applications, vol. 60. Springer, The
deformation polycrystal viscoplasticity. Acta Metall. 35, 2983–2994. Netherlands, pp. 3–15.
Molinari, A., El Houdaigui, F., Tóth, L., 2004. Validation of the tangent formulation Zbib, H.M., Aifantis, E.C., 1989. A gradient-dependent flow theory of plasticity:
for the solution of the non-linear eshelby inclusion problem. Int. J. Plast. 20, application to metal and soil instabilities. Appl. Mech. Rev. Trans. ASME 42,
291–307. S295–S304.

Please cite this article in press as: Wu, L., et al. An implicit-gradient-enhanced incremental-secant mean-field homogenization scheme for elasto-plastic
composites with damage. Int. J. Solids Struct. (2013), http://dx.doi.org/10.1016/j.ijsolstr.2013.07.022

Das könnte Ihnen auch gefallen