Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING

Int. J. Numer. Meth. Engng 2011; 87:149170


Published online 13 September 2010 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/nme.3014
On micro-to-macro transitions for multi-scale analysis
of non-linear heterogeneous materials: unied variational
basis and nite element implementation
D. Peri c
1, ,
, E. A. de Souza Neto
1
, R. A. Feijo
2
, M. Partovi
1
and A. J. Carneiro Molina
1
1
Civil and Computational Engineering Centre, School of Engineering, Swansea University, Singleton Park,
Swansea SA2 8PP, U.K.
2
Laboratrio Nacional de Computao Cientca (LNCC/MCT), Av. Getlio Vargas, 333, Quitandinha,
PetrpolisRio de Janeiro, CEP 25651-075, Brazil
SUMMARY
This work describes a homogenization-based multi-scale procedure required for the computation of the
material response of non-linear microstructures undergoing small strains. Such procedures are important
for computer modelling of heterogeneous materials when the length-scale of heterogeneities is small
compared to the dimensions of the body. The described multi-scale procedure relies on a unied variational
basis which, apart from the continuum-based variational formulation at both micro- and macroscales of
the problem, also includes the variational formulation governing micro-to-macro transitions. This unied
variational basis leads naturally to a generic nite element-based framework for homogenization-based
multi-scale analysis of heterogenous solids. In addition, the unied variational formulation provides clear
axiomatic basis and hierarchy related to the choice of boundary conditions at the microscale. Classical
kinematical constraints are considered over the representative volume element: (i) Taylor, (ii) linear
boundary displacements, (iii) periodic boundary displacement uctuations and (iv) minimal constraint,
also known as uniform boundary tractions. In this context the Hill-Mandel averaging requirement, which
links microscopic and macroscopic stress power, plays a fundamental role in dening the microscopic
forces compatible with the assumed kinematics. Numerical examples of both microscale and two-scale
nite element simulations of elasto-plastic material with microcavities are presented to illustrate the main
features and scope of the described computational strategy. Copyright 2010 John Wiley & Sons, Ltd.
Received 4 February 2010; Revised 2 July 2010; Accepted 13 July 2010
KEY WORDS: multi-scale analysis; homogenization; nite element; micro-to-macro transitions
1. INTRODUCTION
The ever increasing requirements in high-performance applications have provided a constant stim-
ulus for the design of new materials. Often, this has been achieved by appropriately manipulating the
material microstructure, for instance, by adding certain constituents to a matrix phase, thus tailoring
the overall material properties to specic applications. The added phase is typically of a scale much
smaller than the overall structural size, hence making the direct modelling of the material behaviour
impractical. In many situations scales remain tightly coupled and the traditional phenomenological
approach for constitutive description does not provide a sufciently general predictive modelling

Correspondence to: D. Peri c, Civil and Computational Engineering Centre, School of Engineering, Swansea Univer-
sity, Singleton Park, Swansea SA2 8PP, U.K.

E-mail: d.peric@swansea.ac.uk
Copyright 2010 John Wiley & Sons, Ltd.
150 D. PERI

C ET AL.
capability. Therefore a means of continuous interchange of information between scales is needed
if better predictive modelling of material behaviour is to be attempted.
As the basic principles for the micromacro modelling of heterogeneous materials were intro-
duced (see [13]), this technique has proved to be a very effective way to deal with arbitrary
physically non-linear and time-dependent material behaviour at microlevel. During the last decade
various approaches and techniques for the micromacro modelling and simulation of heteroge-
neous materials have been proposed. Among these we highlight the contributions by Oden and
co-workers [4, 5], Ghosh et al. [6], Suquet and co-workers [7, 8], Fish and co-workers [911],
Smit et al. [12], Feyel and Chaboche [13], Miehe and co-workers [1416], Pellegrino et al. [17],
Kouznetsova et al. [18, 19], Ladevze et al. [20, 21], Terada and co-workers [2224], Markovi c
and Ibrahimbegovi c [25], Zohdi and Wriggers [26], Belytschko et al. [27, 28] and Hund and
Ramm [29].
The present article discusses issues related to a computational strategy for homogenization
of microstructures with non-linear material behaviour undergoing small strains. As the aim is to
provide the basic ingredients of the computational strategy allowing for concurrent simulation at
different scales of the model, a simple model is considered comprising two scales arising, for
instance, in the modelling of heterogeneous composite materials.
The aims of this article are two fold: rst, a unied variational basis is outlined, which, apart
from the continuum-based variational formulation at both micro- and macroscales of the problem,
also includes the variational formulation governing micro-to-macro transitions. The unied varia-
tional formulation provides clear axiomatic basis and hierarchy related to the choice of boundary
conditions at the microscale. In this way, issues and ambiguities are avoided that often arise when
the microscale boundary conditions are imposed in an ad hoc manner.
This unied variational basis leads naturally to a generic nite element-based framework for
homogenization-based multi-scale analysis of heterogenous solids. Hence, as the second impor-
tant aim of this article, the nite element implementation arising from the multi-scale varia-
tional framework is developed. The attention is restricted to deformation-driven microstructures,
which have been proven to provide a convenient computational format [14, 16, 30]. Four types of
boundary conditions are imposed over the Representative Volume Element (RVE): (i) Taylor model,
(ii) linear displacements, (iii) periodic displacement uctuations with antiperiodic tractions and
(iv) uniform boundary tractions. These boundary conditions satisfy the fundamental Hill-Mandel
averaging principle, which equates microscopic and macroscopic stress power [3134]. The
resulting computational strategy is characterized by the NewtonRaphson solution of the discrete
boundary value problem, and incorporates the appropriate exact tangent operators.
Numerical examples of both microscale and two-scale nite element simulations are presented
to illustrate the scope and the benets of the described computational strategy.
2. MULTI-SCALE CONSTITUTIVE THEORY: VARIATIONAL FRAMEWORK
This section summarizes the (kinematical) variational basis of the family of homogenization-
based multi-scale constitutive theories addressed in this paper. The variational setting summarized
here is described in further detail by de Souza Neto and Feijo [35]. It provides a structured
axiomatic framework whereby a clear distinction is made between the basic assumptions and their
consequences to the resulting theory. As often seen in the literature, these are usually blurred and
difcult to grasp when the theory is presented on the basis of ad hoc arguments.
The starting point of these theories is the assumption that any material point x of the (macro-
scopic) continuum is associated to a local RVE whose domain

, with boundary *

, (refer to
Figure 1), has a characteristic length, l

, much smaller than the characteristic length, l, of the


macro-continuum. The domain

of the RVE is assumed to consist in general of a solid part,


s

,
and a void part,
v

=
s

. (1)
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 151
macroscopic
continuum
(macro-scale)
x
representative
volume element
(micro-scale)
y

l
<< l

s
U

Figure 1. Macro-continuum with a locally attached microstructure.


For simplicity, in what follows, we shall consider only RVEs whose void part does not intersect
the RVE boundary.
Having dened the above, the entire multi-scale theory can be derived in an axiomatic manner
from ve basic assumptions: (i) a volume averaging relation linking the macroscopic and micro-
scopic strains; (ii) a simple constraint on the possible functional sets of kinematically admissible
displacement uctuations of the RVE; (iii) the Principle of Virtual Work establishing the equilibrium
for the RVE; (iv) a volume averaging relation linking the macro- and microscopic stress tensors
and (v) the Hill-Mandel Principle of Macrohomogeneity that establishes the energy consistency
between the macro- and microscales. These are described in the following.
2.1. Strain averaging
The rst basic assumption is that at any instant t , the strain tensor at an arbitrary point x of
the macro-continuum is dened as the volume average of the microscopic strain tensor eld, e

,
dened over

:
e(t ) =
1
V

(y, t ) dV, (2)


where y is the local coordinate of the RVE, V

=||

|| is the volume of the RVE and


e

=
s
u

, (3)
where
s
u

denotes the symmetric gradient of the microscopic displacement eld u

of the RVE.
2.2. Minimum and actual RVE kinematical constraint
By replacing (3) into (2) and making use of Greens theorem, it can easily be established that the
averaging relation (2) is equivalent to the following constraint on the displacement eld of the
RVE [35]:
_
*

s
ndA=V

e, (4)
where n denotes the outward unit normal eld on *

and
a
s
b
1
2
(ab+ba), (5)
for any vectors a and b. Expression (4) remains valid in the presence of voids, i.e. when
v

=.
From the above considerations it is clear that the essential consequence of assumption (2) is
the constraint (4) on the possible displacement elds of the RVE. Within the virtual work-based
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
152 D. PERI

C ET AL.
variational framework used here to cast this family of multi-scale models, this constraint is promptly
incorporated by simply requiring the (as yet not dened) set K

of kinematically admissible RVE


displacement elds to be a subset of the minimally constrained set of kinematically admissible
microscopic displacements, denoted K

:
K

_
v, sufciently regular

_
*

v
s
ndA=V

e
_
(6)
with sufciently regular meaning that the relevant functions have the sufcient degree of regularity
so that all operations in which they are involved make sense.
Alternatively, by splitting u

, without loss of generality, into a sum


u

(y, t ) =e(t ) y+ u

(y, t ) (7)
of a displacement due to a homogeneous strain eld, e(t ) y, and a displacement uctuation eld,
u

, the above constraint is made equivalent to requiring that the functional set

K

of kinematically
admissible displacement uctuations of the RVE be a subset of the minimally constrained space
of kinematically admissible displacement uctuations,

K

_
v, sufciently regular

_
*

v
s
ndA=0
_
. (8)
At this point, a further fundamental assumption is introduced: The (as yet not dened) set

K

is required to be a subspace of

K

. For application within a virtual work-based variational setting,


it is worth noting that in this case the associated space of virtual kinematically admissible of RVE
displacements, dened as
V

{g=v
1
v
2
|v
1
, v
2
K

}, (9)
trivially coincides with the space of kinematically admissible displacement uctuations:
V

=

K

. (10)
Further, the same arguments applied to the rate form
u

= ey+

u

(11)
of the additive split (7) establish that any kinematically admissible uctuation velocity,

u

, satises

=

K

. (12)
Following the split (7) the microscopic strain (3) can be expressed as the sum
e

(y, t ) =e(t )+
s
u

(y, t ) (13)
of a homogeneous strain eld (coinciding with the macroscopic, average strain) and a eld
s
u

that represents a uctuation about the average strain.


2.3. Equilibrium of the RVE
The next basic axiom of the theory establishes that the RVE must be in equilibrium at any instant
of its deformation history. Assuming that the RVE is subjected in general to a body force eld
b=b(y, t ) and an external traction eld t
e
=t
e
(y, t ) exerted upon the RVE across its external
boundary *

, the Principle of Virtual Work establishes that the RVE is in equilibrium if and only
if the variational equation
_

(y, t ) :
s
gdV
_

b(y, t )gdV
_
*

t
e
(y, t )gdA=0
gV

(14)
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 153
holds at each t , where V

is an appropriate space of virtual displacements of the RVE, subjected


to the general constraints set out in (8,10). In establishing the RVE equilibrium, we have assumed
for simplicity that possible non-zero tractions exerted upon the solid part of the RVE across its
solid/void interfaces (which could arise, for instance, in the presence of a pressurized uid in the
voids) are absent. These can be incorporated into the theory, however, in a straightforward manner.
2.4. Macroscopic stress
Similar to the macroscopic strain denition (3), the macroscopic stress tensor, r, is taken as the
volume average of the microscopic stress eld, r

, over the RVE:


r(t )
1
V

(y, t ) dV. (15)


2.5. The Hill-Mandel principle of macro-homogeneity
Another important axiom underlying models of the present type is the Hill-Mandel Principle of
Macro-homogeneity [31, 36] which requires the macroscopic stress power to equal the volume
average of the microscopic stress power for any kinematically admissible motion of the RVE. This
is expressed by the equation
r: e=
1
V

: e

dV (16)
that must hold for any kinematically admissible microscopic strain rate eld, e

. By combining
(16) with (14) and taking (1012) into account together with the fact that V

is a vector space, we
can establish after straightforward manipulations that (16) is equivalent to the following variational
equation:
_
*

t
e
gdA=0,
_

bgdV =0 gV

(17)
in terms of the RVE boundary traction and body force elds denoted, respectively, t
e
and b. That
is, the virtual work of the RVE body force and surface traction elds vanishthey are the reaction
forces associated to the imposed kinematical constraints embedded in the choice of V

.
2.6. The RVE equilibrium problem
As a consequence of the above, the variational equilibrium statement (14) for the RVE is reduced to
_

:
s
gdV =0 gV

. (18)
Further, we assume that at any time t the stress at each point y of the RVE is delivered by a
generic constitutive functional S
y
of the strain history e
t

(y) at that point up to time t :


r(y, t ) =S
y
(e
t

(y)). (19)
This constitutive assumption, together with the equilibrium equation (18) leads to the denition
of the RVE equilibrium problem which consists in nding, for a given macroscopic strain e (a
function of time), a displacement uctuation function u

such that
_

S
y
{[e+
s
u

(y)]
t
}:
s
gdV =0 gV

. (20)
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
154 D. PERI

C ET AL.
2.7. Characterization of the multi-scale constitutive model
The general multi-scale constitutive model in the present context is dened as follows. For a given
macroscopic strain history, we must rst solve the RVE equilibrium problem dened by (20). With
the solution u

at hand, the macroscopic stress tensor is determined according to the averaging


relation (15), i.e. we have
r(t ) =S(e
t
)
1
V

S
y
{[e+
s
u

]
t
}dV, (21)
where S denotes the resulting (homogenized) macroscopic constitutive response functional.
2.7.1. The choice of kinematical constraints. The characterization of a multi-scale model of the
present type is completed with the choice of a suitable space of kinematically admissible displace-
ment uctuations, V

, so as to make the RVE equilibrium problem (18) well-posed. In


general, different choices lead to different macroscopic constitutive response functionals. The
following choices are frequently employed:
(i) The Taylor model, or rule of mixtures [37] :
V

=
Taylor
V

{0}. (22)
The strain in this case is uniform over the RVE:
e

(y) =e y

. (23)
The (generally non-zero) RVE body force and boundary surface traction elds in this case
are reactions to the imposed uniform strain eld.
(ii) Linear boundary displacements (or zero boundary uctuations) model:
V

=
lin
V

{v, sufciently regular | v(y) =0 y*

}. (24)
The displacements of the boundary of the RVE for this class of models are fully prescribed
as
u

(y) =ey y*

. (25)
The boundary surface traction eld of the RVE is this case is a reaction to the prescribed
boundary displacements of the RVE. The reactive body force eld in this case is
b(y) =0 y

. (26)
(iii) Periodic boundary uctuations. This assumption is typically associated with the description
of media with periodic microstructure. The macrostructure in this case is generated by
the periodic repetition of the RVE [8]. For simplicity, we will focus the description on
two-dimensional problems and we shall follow the notation adopted by Miehe et al. [14].
Consider, for example, the square or hexagonal RVEs, as illustrated in Figure 2. In this
case, each pair i of sides consists of equally sized subsets

+
i
and

i
of *

, with respective unit normals


n
+
i
and n

i
,
such that
n

i
=n
+
i
. (27)
A one-to-one correspondence exists between the points of
+
i
and

i
. That is, each point
y
+

+
i
has a corresponding pair y

i
.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 155
Figure 2. RVE geometries for periodic media. Square and hexagonal cells.
The key kinematical constraint for this class of models is that the displacement uctuation
must be periodic on the boundary of the RVE. That is, for each pair {y
+
, y

} of boundary
material points we have
u

(y
+
, t ) = u

(y

, t ). (28)
Accordingly, the space V

is dened as
V

=
per
V

{ u

, suff. reg. u

(y
+
, t ) = u

(y

, t ) pairs {y
+
, y

}}. (29)
The variational relations (17) in this case imply anti-periodic boundary surface tractions:
t(y
+
, t ) =t(y

, t ) pairs {y
+
, y

}, (30)
and zero body force eld b=0.
(iv) The minimally constrained (or uniform boundary traction) model:
V

=
uni
V

. (31)
It can be shown [35] that the distribution of stress vector on the RVE boundary that satises
(17)
1
is given by
r

(y, t ) n(y) =r(t ) n(y) y*

. (32)
The reactive body force eld, in turn, is zero. As for the Taylor model and linear boundary
displacements assumption, there are no restrictions on the geometry of the RVE in the
present case.
Remark 1
The use of different denitions of V

for a given RVE produces, in general, different estimates


of the corresponding macroscopic constitutive response. The general rule is as follows: First note
that upon inspection of denitions (22), (24), (29) and (31) it follows that
Taylor
V

lin
V

per
V

uni
V

. (33)
That is, the Taylor model gives the stiffest (most kinematically constrained) solution to the
microscopic equilibrium problem, followed in order of decreasing stiffness, by the linear boundary
displacement, the periodic displacement uctuation and the uniform boundary traction model.
The uniform traction model produces the most compliant (least kinematically constrained)
solution.
3. NUMERICAL APPROXIMATION
This section provides a brief description of the computational implementation of multi-scale
constitutive theories of the above type within a non-linear nite element framework.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
156 D. PERI

C ET AL.
3.1. The incremental equilibrium problem
At the outset, we shall assume the constitutive behaviour at the RVE level to be described by
conventional internal variable-based dissipative constitutive theories, whereby the stress tensor is
obtained by integrating a set of ordinary differential equations in time (or pseudo-time) for the given
strain tensor history. Elasto-plasticity and visco-plasticity are classical, widely used examples of
such specializations of (19). In these cases, numerical approximations to the initial value problem
dened by the constitutive equations of the model are usually obtained by Euler-type difference
schemes. For a typical time (or pseudo-time) interval [t
n
, t
n+1
], with known set a
n
of internal
variables at t
n
, the stress r
n+1

at t
n+1
is a (generally implicit) function of the (prescribed) strain
e
n+1

at t
n+1
(refer, for instance, to [3841] for a detailed account of procedures of this kind in
the context of plasticity and visco-plasticity). This can be symbolically represented as
r
n+1

= r
y
(e
n+1

; a
n
), (34)
where r
y
denotes the integration algorithm-related implicit incremental constitutive function at the
point of interest, y.
The above leads to the denition of an incremental version of the homogenized constitutive
function dened in (21), obtained by replacing S
y
with its time-discrete counterpart r
y
:
r
n+1
= r(e
n+1
; a
n
)
1
V

r
y
(e
n+1
+
s
u
n+1

; a
n
) dV, (35)
where a
n
denotes the eld of internal variable sets over
s

at time t
n
and u
n+1

is the displacement
uctuation eld of the RVE at t
n+1
the solution to the time-discrete version of equilibrium
problem (20):
_

r
y
(e
n+1
+
s
u
n+1

; a
n
) :
s
gdV =0 gV

. (36)
3.2. The incremental constitutive function: homogenized constitutive tangent
3.2.1. The incremental constitutive function. Consider rst the simplest multi-scale constitutive
modelthe Taylor model. In this case, we have
V

={0} u
n+1
=e
n+1
y u

n+1
=0, (37)
and the homogenized constitutive function can be written as
Taylor
r(e
n+1
, t ; a
n
)
1
V

(e
n+1
, t ; a
n
) dV, (38)
where a
n
denotes the eld of variables a
n
over the entire domain
s

. That is, at each point y


s

,
a
n
= a
n
(y). (39)
In the general case, the homogenized incremental constitutive function for the stress is dened
implicitly through the incremental microscopic equilibrium equation (36). The stress r
n+1
is
obtained by rst solving (36) and then, with u

|
n+1
at hand, computing
r
n+1
=
1
V

(e
n+1
+
s
u

|
n+1
, t ; a
n
) dV. (40)
That is, the incremental macroscopic stress constitutive function is dened as
hom
r(e
n+1
, t ; a
n
)
1
V

(e
n+1
+
s
u

|
n+1
, t ; a
n
) dV, (41)
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 157
where u

|
n+1
is itself a function solely of e
n+1
(for given t and a
n
) dened as the solution
of (36). Note that, under the Taylor assumption (37), denition (41) recovers the Taylor model
incremental constitutive function (38).
3.2.2. Homogenized constitutive tangent. Let
e

=e
n+1
+e (42)
be the perturbed macroscopic strain, where is a scalar and e denotes a generic incremental strain
tensor. The incremental tangent, denoted
hom
D is a fourth-order tensor that expresses the tangential
relationship between the macroscopic stress and macroscopic strain tensor at t
n+1
, consistently
with the homogenized incremental constitutive function (41). That is, for any macroscopic strain
direction e, we have
hom
r(e

, t ; a
n
) =
hom
r(e
n+1
, t ; a
n
)+
hom
D: e+o(), (43)
where
hom
D: e is the directional derivative of the incremental constitutive function
hom
r in the
direction e:
hom
D: e
d
d

=0
hom
r(e
n+1
+e, t ; a
n
). (44)
The operator
hom
D is simply
hom
D
*
*e

n+1
hom
r(e, t ; a
n
). (45)
Taylor model. For the incremental version of the Taylor model, we obtain, by differentiating (38),
the following homogenized tangent operator:
hom
D=
Taylor
D
1
V

dV, (46)
where
D

=
*
*e

|
n+1
r

(e

, t ; a
n
) (47)
is the tangent operator consistent with the microscopic incremental constitutive law. That is, for the
Taylor model, the incremental homogenized tangent tensor is the volume average of the microscopic
incremental constitutive tangent tensor.
The general case. To obtain a canonical expression for the general case, let us rst consider the
perturbed microscopic displacement uctuation
u

(y) u

n+1
(y)+ u

(y). (48)
The tangential relation between e and u

is obtained from the linearization of the incremental


equilibrium problem dened by (36). Given e, nd the eld u

that solves the linear


variational equation
_

s
g: D

:
s
u

dV =
_
_

s
g: D

dV
_
: e gV

. (49)
It can be shown that the following compact canonical formula for the general homogenized
incremental constitutive tangent operator, can be obtained:
hom
D=
Taylor
D+

D. (50)
Here, only the contribution

D depends on the choice of space V

.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
158 D. PERI

C ET AL.
3.3. Finite element discretization and solution
We now focus on the nite element solution of the time-discrete equilibrium problem (36)a
crucial step in the denition of the approximate homogenized constitutive functional. Following
a standard notation, the nite element approximation to problem (36) for a given discretization h
consists in determining the unknown vector u
n+1

V
h

of global nodal displacement uctuations


such that
G
h
( u
n+1

)
_
_

s,h

B
T

y
(
n+1
+B u
n+1

) dV
_
=0 V
h

, (51)
where
s,h

denotes the discretized RVE domain, B the global straindisplacement matrix (or
discrete symmetric gradient operator),
n+1
is the xed (given) array of macroscopic engineering
strains at t
n+1
,
y
(with upright ) is the functional that delivers the nite element array of stress
components, denotes global vectors of nodal virtual displacements of the RVE and V
h

is the
nite-dimensional space of virtual nodal displacement vectors associated with the nite element
discretization h of the domain
s

.
The solution to the (generally non-linear) problem (51) can be efciently undertaken by the
NewtonRaphson iterative scheme, whose typical iteration (k) consists in solving the linearized
form,
_
F
(k1)
+K
(k1)
u
(k)

_
=0 V
h

, (52)
for the unknown iterative nodal displacement uctuations vector, u
(k)

V
h

where
F
(k1)

s,h

B
T

y
(
n+1
+B u
(k1)

) dV, (53)
and
K
(k1)

s,h

B
T
D
(k1)
BdV (54)
is the tangent stiffness matrix of the RVE with
D
(k1)

d
y
d

=
n+1
+B u
(k1)

(55)
denoting the consistent constitutive tangent matrix eld over the RVE domain. In the above the
bracketed superscript denotes the Newton iteration number and the time station superscript n+1
has been dropped whenever convenient for ease of notation. With the solution u
(k)

at hand, the new


guess u
(k)

for the displacement uctuation at t


n+1
is obtained according to the NewtonRaphson
update formula:
u
(k)

= u
(k1)

+ u
(k)

. (56)
4. FINITE ELEMENT IMPLEMENTATION
Trivially, for the Taylor model, the choice of space of admissible uctuations already denes
the solution to the equilibrium problem (52) so that no numerical procedures are required in
this case, with the macro-stress being computed simply as the volume-fraction weighted average
of the microstresses resulting from the application of the macro-strain history to the different
phases of the RVE. Under the assumption of linear boundary displacements, the solution of
problem (52) follows the conventional route of general linear solid mechanics problemshere
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 159
y
y
+
-
y
-
y
+
i
-
i
+
i
+
i
-
Figure 3. Discretized RVEs for periodic media.
with the degrees of freedom (uctuations) of the boundary fully prescribed as zero. Hence, the
nite element implementation of this class of multi-scale models requires no further consideration.
For the periodic boundary condition and minimally constrained models, however, the kinematic
boundary conditions of the RVE are non-conventional. The main difference lies in the nite element-
generated nite dimensional spaces of admissible uctuations and virtual displacements whose
constraints, here, are not simply described in terms of either fully constrained or completely free
nodal degrees of freedom. Thus, some ambiguities exist in the literature regarding the imposition
of the non-conventional boundary conditions, which are often prescribed in an ad hoc manner.
It is demonstrated below that nite element implementation of the periodic boundary condition
and minimally constrained models follows naturally from the unied variational framework for
multi-scale analysis described in Section 2.
4.1. Periodic boundary uctuations model
For the periodic boundary displacement uctuations model, the RVE geometry must comply
with the constraints set out in item (iii) listed in Section 2.7.1. In this case, it is convenient to
assume,

further, that each boundary node i


+
, with coordinates y
+
i
, has a pair i

, with coordinates
y

i
, as schematically illustrated in Figure 3. Under this assumption, the space V
h

of discretized
kinematically admissible nodal displacement uctuation vectors (with periodicity on the boundary)
can then be dened as
V
h

=
_

_
v=
_
_
_
v
i
v
+
v

_
|v
+
=v

_
, (57)
where v
i
, v
+
and v

denote the vectors containing, respectively, the degrees of freedom of the


RVE interior and the portions
+
and

of the RVE boundary. Here we adopt the direct approach


suggested by Michel et al. [8] whereby the periodicity constraint is enforced exactly in the
discretized space of uctuations and virtual displacements. This is at variance with the treatment
adopted by Miehe and Koch [16] who used a Lagrange multiplier method to enforce the discrete
space constraint.
By splitting F, K, u

and in the same fashion as v in the above and taking into account
denition (57) as well as the fact that both and u
(k)

belong to space V
h

, the linearized
equation (52) takes the form
_

_
_
_
_
F
i
F
+
F

_
(k1)
+
_
_
_
_
k
ii
k
i +
k
i
k
+i
k
++
k
+
k
i
k
+
k

_
(k1)
_
_
_
u
i
u
+
u
+
_

_
(k)
_

_
_
_

+
_

_
=0
i
,
+
. (58)

This assumption is not necessary, but simplies considerably the nite element implementation of the periodic
uctuations model.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
160 D. PERI

C ET AL.
Straightforward manipulations, considering the repetition of u
+
and
+
in the relevant vectors of
nodal degrees of freedom, reduce the linearized discrete equilibrium equation (58) to the following
form:
_

_
_
F
i
F
+
+F

_
(k1)
+
_
_
k
ii
k
i +
+k
i
k
+i
+k
i
k
++
+k
+
+k
+
+k

_
_
(k1) _
u
i
u
+
_
(k)
_

+
_
=0

i
,
+
, (59)
which, nally, in view of the arbitrariness of
i
and
+
, yields the linear system of algebraic
equations for the unknown vectors u
(k)
i
and u
(k)
+
,
_
k
ii
k
i +
+k
i
k
+i
+k
i
k
++
+k
+
+k
+
+k

_
(k1)
_
u
i
u
+
_
(k)
=
_
F
i
F
+
+F

_
(k1)
.
(60)
4.2. Minimally constrained model
A procedure completely analogous to the one above is followed to obtain the nal NewtonRaphson
set of algebraic nite element equations under the assumption of minimally constrained kinematics
(or uniform RVE boundary traction). We then start by dening the discrete counterpart of the
minimally constrained space of uctuations and virtual displacements (8, 31):
V
h


_
v=
_
v
i
v
b
_

_
*
h

N
b
v
b

s
ndA=0
_
, (61)
where v
b
is the vector containing the boundary degrees of freedom and N
b
is the global interpolation
matrix associated solely with the boundary nodes of the discretized RVE.
It can be easily established that the integral constraint on v
b
can be equivalently written in
matrix form as
Cv
b
=0, (62)
where C is the constraint matrix . For an RVE mesh with k interior nodes and m boundary nodes,
in the two-dimensional case v
b
is a vector of dimension 2m and C is the 32m matrix given by
C=
_
_
_
_
_
_
_
_
_
_
_
_
_
_
*
h

N
k+1
n
1
dA 0
_
*
h

N
k+m
n
1
dA 0
0
_
*
h

N
k+1
n
2
dA 0
_
*
h

N
k+m
n
2
pacedA
_
*
h

N
k+1
n
2
dA
_
*
h

N
k+1
n
1
dA
_
*
h

N
m
n
2
dA
_
*
h

N
m
n
1
dA
_

_
, (63)
where n
1
and n
2
denote the components of the outward unit normal eld along the global
orthonormal basis {e
1
, e
2
} and N
j
, j =1, , m, are the global shape functions associated with the
boundary nodes. In this case, Equation (62) poses three linear constraints upon the total number
of 2m boundary degrees of freedom of the discrete RVE. For three-dimensional RVEs, v
b
is of
dimension 3m and matrix C has dimension 63m.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 161
In practice, rather than using global shape functions, matrix C is obtained by assembling
elemental matrices which in two dimensions, for an element e with p nodes on the intersection

(e)
between the boundary of the element and the boundary of the RVE, read
C
(e)
=
_
_
_
_
_
_
_
_
_
_

(e)
N
(e)
1
n
1
dA 0
_

(e)
N
(e)
p
n
1
dA 0
0
_

(e)
N
(e)
1
n
2
dA 0
_

(e)
N
(e)
p
n
2
dA
_

(e)
N
(e)
1
n
2
dA
_

(e)
N
(e)
1
n
1
dA
_

(e)
N
(e)
p
n
2
dA
_

(e)
N
(e)
p
n
1
dA
_

_
, (64)
where we have assumed that the nodes of element e lying on
(e)
are locally numbered 1 to p
and N
(e)
j
, j =1, . . . , p, are the associated local shape functions. For example, a conventional eight-
noded bilinear quadrilateral element (of the type employed in Section 5), having a single straight
side of length l
(e)
with n=e
1
and three equally spaced nodes intersecting the RVE boundary, has
C
(e)
=l
(e)
_
_
_
_
1
6
0
2
3
0
1
6
0
0 0 0 0 0 0
0
1
6
0
2
3
0
1
6
_

_
. (65)
In order to handle constraint (62) upon the discrete space of uctuations and virtual displacements
it is convenient to split v
b
as
v
b
=
_
_
_
v
f
v
d
v
p
_

_
, (66)
where the subscripts f , d and p stand, respectively, for free, dependent and prescribed degrees
of freedom on the boundary of the discrete RVE. Accordingly, the global constraint matrix is
partitioned as
C=[C
f
C
d
C
p
], (67)
so that the constraint equation (62) reads as
[C
f
C
d
C
p
]
_
_
_
v
f
v
d
v
p
_

_
=0. (68)
Prescribed degrees of freedom are needed here to remove rigid body displacements of the RVE and
make the corresponding discrete equilibrium problem (51) well-posed. Trivially, we then prescribe
v
p
=0, (69)
where, in two and three dimensions, v
p
contains, respectively, three and six suitably chosen degrees
of freedom. The constraint equation is now reduced to
_
C
f
C
d
_
_
v
f
v
d
_
=0. (70)
In two dimensions, the above represents three scalar equations involving 2m3 variables, whereas
in the three-dimensional case, we have six scalar equations and 3m6 variables. Hence, the
number of dependent variablesthe dimension of v
d
and of the square sub-matrix C
d
is 3 and
6 for the two- and three-dimensional cases, respectively. The total number of free variablesthe
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
162 D. PERI

C ET AL.
dimension of v
f
and number of columns of C
f
is 2m6 and 3m12, respectively, in two and
three dimensions. Finally, following a trivial manipulation of (70), v
d
can be expressed in terms
of v
f
as
v
d
=Rv
f
, (71)
where
RC
1
d
C
f
. (72)
Note that the dependent degrees of freedom (corresponding to v
d
) must be chosen such that C
d
is
invertible.
With the above considerations at hand, we can re-dene the discrete space (61) of uctuations
and virtual displacements of the RVE as
V
h

_
v=
_
_
_
v
i
v
f
v
d
_

_
|v
d
=Rv
f
_

_
, (73)
which, for convenience, contains now only the non-prescribed degrees of freedom.
The particularization of the linearized nite element equation (52) for the present case is obtained,
analogously to (58), by splitting the corresponding vectors and tangential stiffness matrix according
to the above partitioning and taking (73) into account. This gives
_

_
_
_
_
F
i
F
f
F
d
_

_
(k1)
+
_
_
_
_
k
ii
k
if
k
id
k

k
ff
k
fd
k
di
k
df
k
dd
_

_
(k1)
_
_
_
u
i
u
f
R u
f
_

_
(k)
_

_
_
_
_

f
R
f
_

_
=0
i
,
f
, (74)
which, after straightforward matrix manipulations taking into account the arbitrariness of
i
and

f
, is reduced to the nal form
_
k
ii
k
if
+k
id
R
k

+R
T
k
di
k
ff
+k
fd
R+R
T
k
df
+R
T
k
dd
R
_
(k1)
_
u
i
u
f
_
(k)
=
_
F
i
F
f
+R
T
F
d
_
(k1)
. (75)
5. NUMERICAL EXAMPLES
In this section numerical examples are presented to illustrate the scope and benets of the described
computational strategy. The rst set of numerical examples focuses on microstructure simulations
and discusses some important issues regarding numerical analysis at the microlevel, such as the
effect of boundary conditions, topology and distribution of heterogeneities. The second numerical
example considers a full two-scale simulation of a boundary value problem and incorporates all
the computational ingredients described in this paper. This example also includes a comparison
with a detailed single-scale analysis. The standard computational algorithms for elasto-plasticity
and viscoplasticity are employed that underly the multi-scale approach, and we refer the reader to
[40, 4246] for detailed descriptions of the computational schemes behind the present simulations.
It should be emphasized that in order to ensure the robustness and efciency of the multi-scale
simulations the recently proposed sub-stepping procedure is utilized [47].
5.1. Effect of cavity distribution on homogenized properties
5.1.1. Problem specications. A square cell is considered representing an RVE at the microlevel.
The cell is composed of an elasto-plastic material with heterogeneities in the form of cavities.
Two models are considered: (i) a regular cell with a single circular void embedded in a soft matrix
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 163
Figure 4. Regular cavity model.
depicted in Figure 4 and (ii) randomly generated distribution of voids surrounded by the soft matrix
given in Figure 6. For both models the void volume fraction of the cell is 15%.
Two types of nite elements are employed: linear three-noded triangle element and eight-noded
quadrilateral element. The matrix in all models is assumed to be composed of a von Mises elasto-
plastic material with linear strain hardening. The material properties assigned are: Youngs modulus
E =70GPa, Poissons ratio =0.2, the initial yield stress
Y
0
=0.243GPa and the strain hardening
modulus H =0.2GPa.
5.1.2. Analysis approach. All simulations in this section have been performed by employing the
computational homogenization under the plane stress assumption in small strain regime. The
average stress is obtained by imposing the macro-strain over the RVE and subsequently solving
the microscopic initial boundary value problem for the boundary condition assumed. The generic
imposed macro-strain tensor is expressed by
[
11
,
22
, 2
12
] =[0.001, 0.001, 0.0034].
The loading programme considered is proportional in the sense that the nal strain at the end of
each step is dened by multiplying the above strain array by the relevant load factor. The analysis
is performed under three different boundary conditions considered in this paper: (i) linear boundary
displacements, (ii) periodic boundary displacement uctuations and (iii) uniform boundary traction.
5.1.3. Study of the regular cavity model. A mesh of 350 8-noded isoparametric quadrilateral
elements (see Figure 4) is employed in this simulation. The mesh contains a total number of 1158
nodes.
Figures 5(a) and (b) show, respectively, the deformed mesh and the equivalent plastic strain
distribution for the linear displacement boundary condition. This plastic zone is clearly positioned
along the diagonal side of the unit cell in the direction of the imposed shear. The corresponding
results for the periodic boundary condition and uniform boundary traction assumption are given in
Figures 5 (c, d) and Figures 5 (e, f), respectively. From Figure 5, it can be seen that the plastic zone
has a distinctively different pattern under different boundary conditions. To ease visualization, all
deformed meshes of Figure 5 have been plotted with exaggerated displacements.
The overall stressstrain response is presented in terms of the effective homogenized stress and
the Euclidean norm of the average strain, given, respectively as

eff
=
_
3

J
2
, =
_

2
11
+
2
22
+2
2
12
.
Figure 8 shows the resulting average stressstrain curves for this model. The obtained results
indicate that under linear displacement boundary condition the overall response of the regular cavity
model shows signicantly stiffer homogenized behaviour than that predicted under the periodic
boundary condition. On the other hand, the results obtained for the uniform boundary traction
assumption show the softest behaviour.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
164 D. PERI

C ET AL.
(a) (b)
(c) (d)
(e) (f )
Figure 5. Regular cavity model under linear displacement boundary condition (a)(b), periodic condition
(c)(d) and uniform traction boundary condition (e)(f). (a), (c) and (e) represent the deformed mesh,
whereas (b), (d) and (f) are corresponding effective plastic strain contour plots.
Figure 6. RVE with randomly generated voids.
5.1.4. The RVE with randomly generated voids. The analysis carried out here is identical to the
one described above except for the RVE which now contains a randomly generated distribution of
void placements and sizes (see Figure 6). A mesh of standard three-node linear triangular elements
is employed in this simulation.
Figure 7 shows the equivalent plastic strain distribution for all the prescribed boundary condi-
tions. The occurrence of localized bands with signicant plastic straining can be observed on all
contour plots. Signicantly, unlike the case of the single cavity model all boundary conditions give
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 165
Figure 7. Effective plastic strain contours for the unit cell with randomly generated
voids for different boundary conditions.
a similar distribution of the plastic strain indicating the convergence of the results at the microlevel
with the increase of the statistical sample of heterogeneities.
Figure 8 shows the average stressstrain curves for this model. It can be observed that the
microcell with randomly generated void distribution results in the stressstrain behaviour that
shows a small difference between the three different boundary conditions imposed at the microlevel.
This clearly indicates the convergence of the average properties with the increase of the statistical
sample representing the heterogeneities at the microlevel.
5.2. Two-scale analysis of stretching of an elasto-plastic perforated plate
In this section the analysis of the stretching of a perforated plate is performed. The perforated plate
problem is often used as a benchmarking example in computational plasticity. Here we consider
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
166 D. PERI

C ET AL.
Regular cavity. Linear
Regular cavity. Periodic
Regular cavity. Uniform traction
Random distribution. Linear
Random distribution. Uniform traction
Random distribution. Periodic
Figure 8. Effective stressstrain norm curves. Regular (single) cavity model and random cavity distribution
model under different RVE boundary conditions.
R=
Figure 9. Plane-stress strip with a circular hole. Geometry and boundary conditions.
the plate composed of an elasto-plastic material and containing regularly distributed voids. The
plate has width 10 mm, length 18 mm and uniform thickness of 1 mm (see Figure 9). For obvious
symmetry reasons only one-quarter of the specimen is considered (see Figure 9). The simulation is
performed by imposing uniform displacement along the upper boundary. The elasto-plastic material
properties are identical to those used in the RVE simulations described above. Two simulations are
performed: (i) a coupled two-scale analysis using a relatively coarse macroscopic nite element
mesh with the constitutive response at each Gauss quadrature point dened by the computational
homogenization of an RVE consisting of a single cavity model (as discussed above) embedded
in an elasto-plastic matrix and (ii) a single-scale detailed analysis of the plate where the regular
distribution of cavities is discretized directly at the macroscale.
We remark that this example is presented merely to illustrate the potential applicability of
coupled two-scale analysis. It should be pointed out that we are not looking into the convergence
properties of such solutions, which would require a much more detailed analysis.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 167
Figure 10. Single-scale analysis of an elasto-plastic perforated plate: (a) nite element mesh and
(b) distribution of equivalent plastic strain.
Figure 11. FE meshes at macro- and microlevels for multi-scale analysis.
5.2.1. Single-scale analysis. Single-scale analysis is used for comparative purposes and is
performed on a detailed nite element mesh of the problem given in Figure 10(a). The mesh is
composed of 11 216 4-node quadrilateral elements and 12 147 nodes. Figure 10(b) illustrates the
distribution of an equivalent plastic strain at the latter stages of the simulation.
5.2.2. Two-scale analysis. For the multi-scale analysis the perforated plate is dened as a homo-
geneous structure at the macrolevel, whereas at the microlevel an RVE is dened with side length

equal to 1 and a centred single void giving a void volume fraction of 50%. Linear three-noded
triangle elements are employed at both macro- and microlevels (see Figure 11). The mesh at the
macrolevel is composed of 25 elements and 21 nodes, whereas at the microlevel the FE mesh is
composed of 603 elements and 352 nodes.
The multi-scale analysis has been performed under the three types of RVE boundary constraints
considered in this paper. In addition, an analysis using the Taylor assumption (the uniform RVE
strain or rule of mixtures) is carried out for comparison. Figure 12 plots the reaction force against
the prescribed top edge displacement for the four RVE assumptions considered. The different

The actual size of the RVE is immaterial here since size effects are not accounted for within the present approach.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
168 D. PERI

C ET AL.
Regular cavity. Linear
Single scale
Taylor assumption
Multi-scale. Linear
Multi-scale. Uniform traction
Multi-scale. Periodic
Single scale
Taylor assumption
Multi-scale. Linear
Multi-scale. Uniform traction
Multi-scale. Periodic
E
d
g
e

r
e
a
c
t
i
o
n

[
k
N
]
Prescribed edge displacement [mm]
Figure 12. Perforated plate. Reactiondisplacement curve.
assumptions produce markedly different forcedisplacement diagrams. As expected, the results
obtained for the Taylor model show substantially stiffer behaviour than all the others. The resulting
overall behaviour for the periodic boundary condition shows very good correspondence with the
results obtained by the detailed single-scale analysis. This is expected since the actual heterogeneity
is periodic in the macrostructure. Finally, uniform boundary traction assumption generates the
softest response.
6. CONCLUSIONS
A compact and efcient computational framework has been developed for the homogenization-
based multi-scale nite element analysis of solids. The computational framework relies on an
elegant variational formulation, which, in a natural way, introduces a hierarchy of boundary condi-
tions at the microscale, and allows for direct treatment of micro-to-macro transitions.
Details of the novel implementation procedure for three different types of RVE kinematical
constraints have been described, including the corresponding tangent constitutive operators. The
implementation procedure is based on an algebraic factorization and, importantly, does not involve
any articial parameters. As a result an efcient and robust overall scheme for multi-scale analysis
of heterogeneous materials has been developed.
The presented numerical tests conrm the successful implementation of the computational
procedure and efcient solution of the discrete multi-scale problem. The examples clearly illustrate
the bounding properties related to the choice of the microscale boundary conditions, which emanate
from the described variational structure. The signicance of the statistical sample representing the
heterogenities at the microlevel is also illustrated.
The ongoing research is concerned with the analysis of more general non-linear material
behaviour at the microscale at both small and nite strains (see [48] for representative application
to bio-materials) and design and optimization of microstructures [4951].
REFERENCES
1. Sanchez-Palencia E. Comportement local et macroscopique dun type de milieux physique htrognes.
International Journal of Engineering Science 1974; 12:231251.
2. Suquet PM. Local and global aspects in the mathematical theory of plasticity. In Plasticity Today: Modelling
Methods and Applications, Sawczuk A (ed.). Elsevier Applied Science Publishers: Amsterdam, 1985.
3. Suquet PM. Elements of homogenization for inelastic solid mechanics. In Homogenization Techniques for
Composite Media, Sanchez-Palencia E, Zaoui A (eds). Springer: Berlin, 1987.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
ON MICRO-TO-MACRO TRANSITIONS FOR MULTI-SCALE ANALYSIS 169
4. Oden JT, Vemaganti K, Moes N. Hierarchical modelling of heterogenous solids. Computer Methods in Applied
Mechanics and Engineering 1999; 172:225.
5. Zohdi TI, Oden JT, Rodin GJ. Hierarchical modeling of heterogeneous bodies. Computer Methods in Applied
Mechanics and Engineering 1996; 138:273298.
6. Ghosh S, Lee K, Murthy S. Two scale analysis of heterogeneous elastic-plastic materials with asymptotic
homogenization and Voronoi cell nite element method. Computer Methods in Applied Mechanics and Engineering
1996; 132:63116.
7. Moulinec H, Suquet P. A numerical method for computing the overall response of nonlinear composites with
complex microstructure. Computer Methods in Applied Mechanics and Engineering 1998; 157:6994.
8. Michel JC, Moulinec H, Suquet P. Effective properties of composite materials with periodic microstructure: a
computational approach. Computer Methods in Applied Mechanics and Engineering 1999; 172:109143.
9. Fish J, Shek K, Pandheeradi M, Shephard MS. Computational plasticity for composite structures based on
mathematical homogenization: theory and practice. Computer Methods in Applied Mechanics and Engineering
1997; 148:5373.
10. Fish J, Zheng Y. Multiscale enrichment based on partition of unity. International Journal for Numerical Methods
in Engineering 2005; 62:13411359.
11. Yuan Z, Fish J. Toward realization of computational homogenization in practice. International Journal for
Numerical Methods in Engineering 2008; 73:361380.
12. Smit RJM, Brekelmans WAM, Meijer HEH. Prediction of the mechanical behaviour of nonlinear heterogeneous
systems by multi-level nite element modeling. Computer Methods in Applied Mechanics and Engineering 1998;
155:181192.
13. Feyel F, Chaboche J-L. FE
2
multiscale approach for modelling the elastoviscoplastic behaviour of long bre
SiC/Ti composite materials. Computer Methods in Applied Mechanics and Engineering 2000; 183:309330.
14. Miehe C, Schotte J, Schrder J. Computational micro-macro transitions and overall tangent moduli in the analysis
of polycrystals at large strains. Computational Materials Science 1999; 16:372382.
15. Miehe C, Schrder J, Schotte J. Computational homogenization analysis in nite plasticity. Simulation of texture
development in polycrystalline materials. Computer Methods in Applied Mechanics and Engineering 1999;
171:387418.
16. Miehe C, Koch A. Computational micro-macro transitions of discretized microstructures undergoing small strains.
Archives of Applied Mechanics 2002; 72:300317.
17. Pellegrino C, Galvanetto U, Schreer BA. Numerical homogenization of periodic composite materials with non-
linear material components. International Journal for Numerical Methods in Engineering 1999; 46:16091637.
18. Kouznetsova VG, Brekelmans WAM, Baaijens FPT. An approach to micro-macro modelling of heterogeneous
materials. Computational Mechanics 2001; 27:3748.
19. Kouznetsova VG. Computational homogenization for multiscale analysis of multi-phase materials. Ph.D. Thesis,
TU University Eindhoven, Eindhoven, 2002.
20. Ladevze P, Loiseau O, Dureisseix D. A micro-macro and parallel computational strategy for highly heterogeneous
structures. International Journal for Numerical Methods in Engineering 2001; 52:121138.
21. Ladevze P. Multiscale modelling and computational strategies for composites. International Journal for Numerical
Methods in Engineering 2004; 60:233253.
22. Terada K, Kikuchi N. A class of general algorithms for multi-scale analyses of heterogeneous media. Computer
Methods in Applied Mechanics and Engineering 2001; 190:54275464.
23. Matsui K, Terada K, Yuge K. Two-scale nite element analysis of heterogeneous solids with periodic
microstructures. Computers and Structures 2004; 82:593606.
24. Watanabe I, Terada K, de Souza Neto EA, Peric D. Characterization of macroscopic tensile strength of
polycrystalline metals with two-scale nite element analysis. Journal of the Mechanics and Physics of Solids
2008; 56:11051125.
25. Markovi c D, Ibrahimbegovi c A. On micro-macro interface conditions for micro-scale based FEM for inelastic
behaviour of heterogeneous materials. Computer Methods in Applied Mechanics and Engineering 2004; 193:5503
5523.
26. Zohdi TI, Wriggers P. Introduction to Computational Micromechanics. Springer: Berlin, 2005.
27. Belytschko T, Loehnert S, Song JH. Multiscale aggregating discontinuities: a method for circumventing loss of
material stability. International Journal for Numerical Methods in Engineering 2008; 73:869894.
28. Belytschko T, Song JH. Coarse-graining of multiscale crack propagation. International Journal for Numerical
Methods in Engineering 2010; 81:537563.
29. Hund A, Ramm E. Locality constraints within multiscale model for non-linear material behaviour. International
Journal for Numerical Methods in Engineering 2007; 70:16131632.
30. Swan CC. Techniques for stress- and strain-controlled homogenization of inelastic periodic composites. Computer
Methods in Applied Mechanics and Engineering 1994; 117:249267.
31. Mandel J. Plasticit Classique et Viscoplasticit. CISM Courses and Lectures No. 97. Springer: Udine, Italy,
1971.
32. Hill R. On constitutive macro-variables for heterogeneous solids at nite strain. Proceedings of the Royal Society
of London, Series A 1972; 326:131147.
33. Willis JR. Variational and related methods for the overall properties of composites. Advances in Applied Mechanics
1981; 21:178.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme
170 D. PERI

C ET AL.
34. Nemat-Nasser S, Hori M. Micromechanics: Overall Properties of Heterogeneous Materials. North-Holland:
Amsterdam, 1993.
35. de Souza Neto EA, Feijo RA. Variational foundations of multi-scale constitutive models of solid: small and
large strain kinematical formulation. National Laboratory for Scientic Computing (LNCC/MCT), Brazil, Internal
Research & Development Report No. 16/2006, 2006.
36. Hill R. A self-consistent mechanics of composite materials. Journal of the Mechanics and Physics of Solids
1965; 13(4):213222.
37. Taylor GI. Plastic strains in metals. Journal Institute of Metals 1938; 62:307324.
38. Simo JC, Hughes TJR. Computational inelasticity. Springer: New York, 1998.
39. Belytschko T, Liu WK, Moran B. Nonlinear Finite Elements for Continua and Structures. Wiley: New York,
2000.
40. de Souza Neto EA, Peri c D, Owen DRJ. Computational Methods for Plasticity: Theory and Applications. Wiley:
Chichester, 2008.
41. Wriggers P. Nonlinear Finite Element Methods. Springer: Berlin, 2008.
42. de Souza Neto EA, Peri c D. A computational framework for a class of fully coupled models for elastoplastic
damage at nite strains with reference to the linearization aspects. Computer Methods in Applied Mechanics and
Engineering 1996; 130:179193.
43. de Souza Neto EA, Peri c D, Owen DRJ. Continuum modelling and numerical simulation of material damage at
nite strains. Archives of Computational Methods in Engineering 1998; 5:311384.
44. Dutko M, Peric D, Owen DRJ. Universal anisotropic yield criterion based on superquadric functional
representation1: algorithmic issues and accuracy analysis. Computer Methods in Applied Mechanics and
Engineering 1993; 109:7393.
45. Peri c D. On a class of constitutive equations in viscoplasticity: formulation and computational issues. International
Journal for Numerical Methods in Engineering 1993; 36:13651393.
46. Peri c D, de Souza Neto EA. A new computational model for Tresca plasticity at nite strains with an optimal
parametrization in the principal space. Computer Methods in Applied Mechanics and Engineering 1999; 171:
463489.
47. Somer DD, de Souza Neto EA, Dettmer WG, Peri c D. A sub-stepping scheme for multi-scale analysis of solids.
Computer Methods in Applied Mechanics and Engineering 2009; 198:10061016.
48. Speirs DCD, de Souza Neto EA, Peri c D. An approach to the mechanical constitutive modelling of arterial tissue
based on homogenization and optimization. Journal of Biomechanics 2008; 41:26732680.
49. Giusti SM, Novotny AA, de Souza Neto EA, Feijo RA. Sensitivity of the macroscopic elasticity tensor to
topological microstructural changes. Journal of the Mechanics and Physics of Solids 2009; 57:555570.
50. Giusti SM, Novotny AA, de Souza Neto EA, Feijo RA. Sensitivity of the macroscopic thermal conductivity
tensor to topological microstructural changes. Computer Methods in Applied Mechanics and Engineering 2009;
198:727739.
51. Giusti SM, Novotny AA, de Souza Neto EA. Sensitivity of the macroscopic response of elastic microstructures
to the insertion of inclusions. Proceedings of the Royal Society A 2010; 466:17031723.
Copyright 2010 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2011; 87:149170
DOI: 10.1002/nme

You might also like