A Short Introduction to Basic Aspects of Continuum Micromechanics - Bohm
A Short Introduction to Basic Aspects of Continuum Micromechanics - Bohm
H.J. Böhm
at the
Institute of Lightweight Design and Structural Biomechanics
Vienna University of Technology
CONTENTS
i
Notes on this Document
The present document is an expanded and modified version of the lecture notes prepared
for the European Advanced Summer Schools “Frontiers of Computational Micromechan-
ics in Industrial and Engineering Materials” held in Galway, Ireland, in July 1998 and in
August 2000. Similar lecture notes were used for graduate courses in a Summer School
held at Ameland, the Netherlands, in October 2000 and during the COMMAS Sum-
mer School held in September 2002 at Stuttgart, Germany. All of the above documents
were based on the micromechanics section of the lecture notes for the course “Composite
Engineering” given regularly at Vienna University of Technology.
The course notes “A Short Introduction to Continuum Micromechanics” (Böhm,2004) for
the CISM Course on Mechanics of Microstructured Materials held in July 2003 in Udine,
Italy, are basically a shortened version of the present report that employs a somewhat
different notation.
Even though this document is focused on the more basic aspects of continuum microme-
chanics, it is continuously updated to reflect current developments as seen by the author.
For information on the research work in the field of micromechanics of materials currently
carried out at the Institut für Leichtbau und Flugzeugbau see the web pages
http://ilfb.tuwien.ac.at/ilfb/ilfb ra3e.html
ii
List of Variables
h i Volume average
iii
α: Thermal expansion coefficient tensor (rank 2; vector in Nye notation)
I: Identity tensor (rank 4; matrix in Nye notation)
S: Eshelby tensor (rank 4; matrix in Nye notation)
t: Traction vector
ta : Vector of applied tractions
nΓ : Surface normal vector
E: Young’s modulus
G: Shear modulus
K: Bulk modulus
ν: Poisson’s ratio
KT : Transverse bulk modulus
T : Temperature
σ: Stress component
ε: Strain component
γ: Shear angle
α: Coefficient of thermal expansion
Constituents (or “phases”, i.e. matrix, inclusions, fibers, . . .) are denoted by superscripts
in brackets, e.g. (p) . Axial and transverse properties of transversely isotropic materials
are marked by subscripts A and T , respectively, and effective properties are denoted by
a superscript asterisk ∗ .
Nye notation (also known as Voigt–Mandel notation) is used in chapters 1 to 3 of these
notes, i.e. the stress and strain tensors are formally denoted as pseudo-vectors with the
components
where γij = 2εij are the shear angles1) . Tensors of rank 4 such as elasticity and com-
pliance tensors are denoted in the form of matrices. It should be noted that the above
definitions of the stress and strain tensors influence not only the coefficients of the elastic
tensors, but also those of the Eshelby tensor and the concentration tensors.
1)
This notation allows strain energy densities to be obtained as U = 21 σε. Furthermore,
the stress and strain transformation tensors follow the relationship TΘ Θ −1 T
ε = [(Tσ ) ] .
iv
1. INTRODUCTION
In the present document some basic issues and some important approaches in the field
of continuum micromechanics of materials are discussed. The main emphasis lies on
application related (or “engineering”) aspects, and neither a comprehensive theoretical
treatment nor a review of the pertinent literature are attempted. For more formal treat-
ments of many of the concepts and methods discussed in the following see e.g. (Mura,1987;
Aboudi,1991; Nemat–Nasser/Hori,1993; Suquet,1997; Markov,2000; Bornert et al.,2001;
Torquato,2001; Milton,2002); Short overviews were given e.g. by Hashin(1983) and Za-
oui (2002). Discussions of the history of the development of the field can be found in
(Markov,2000; Zaoui,2002).
Due to the author’s research interests, more room is given to the thermomechanical
behavior of two-phase materials showing a matrix–inclusion topology and especially to
metal matrix composites (MMCs), than to materials with other phase arrangements. Ex-
tending the methods presented here to other types of inhomogeneous materials, however,
in general does not cause principal difficulties.
Many industrial and engineering materials as well as the majority of “natural” materi-
als are inhomogeneous, i.e. they consist of dissimilar constituents (or “phases”) that are
distinguishable at some (small) length scale. Each constituent shows different material
properties and/or material orientations and may itself be inhomogeneous at some smaller
length scale(s). Typical examples of inhomogeneous materials are composites, polycrys-
talline materials, porous and cellular materials, functionally graded materials, wood, and
bone.
An important aim of theoretical studies of multiphase materials lies in homogenization,
i.e. in deducing their overall (“effective” or “apparent”) behavior (e.g. stiffness, thermal
expansion and strength properties, heat conduction and related transport properties,
electrical and magnetic properties, electromechanical properties, . . .) from the corre-
sponding material behavior of the constituents (and of the interfaces between them) and
from the geometrical arrangement of the phases. In what follows, the main focus will
lie on describing the thermomechanical behavior of inhomogeneous two-phase materials
by methods of continuum mechanics. It should be noted, however, that there is a large
body of literature applying analogous or related continuum approaches to other physi-
cal properties of inhomogeneous materials, see e.g. (Hashin,1983; Markov/Preziosi,2000;
Milton,2001).
In micromechanical approaches the stress and strain fields in an inhomogeneous material
are split into contributions corresponding to different length scales. It is assumed that
these length scales are sufficiently different so that for each pair of them
• fluctuations of the stress and strain fields at the smaller length scale (microfields,
“fast variables”) influence the macroscopic behavior at the larger length scale only
via their volume averages and
–1–
• gradients of the stress and strain fields as well as compositional gradients at the
larger length scale (macrofields, “slow variables”) are not significant at the smaller
length scale, where these fields appear to be locally constant and can be described
in terms of uniform “applied” or “far field” stresses and strains.
Formally, this splitting of the strain and stress fields into slow and fast contributions can
be written as
ε(x) = hεi + ε0 (x) and σ(x) = hσi + σ 0 (x) , (1)
where hεi and hσi are the (slow) macroscopic fields, whereas ε0 and σ 0 stand for the
microscopic fluctuations.
For example, in layered composites (“laminates”) there are at least three length scales
that can be described by continuum mechanics, which are often called the macro scale
(component or sample), the mesoscale (individual plies), and the microscale (fiber–matrix
structure of the plies)2) . Smaller length scales in such material systems may or may not be
amenable to continuum mechanical descriptions; for an overview of methods applicable
below the continuum range see e.g. (Raabe,1998). From the point of view of a given
length scale the material behavior at any smaller length scale can be described by that of
an equivalent homogenized continuum provided the separation between the two length
scales is sufficiently large and the above conditions are met.
If the above relations between “fast” and “slow” variables are not fulfilled to a sufficient
degree (e.g. due to insufficiently separated length scales, near free surfaces of inhomo-
geneous materials, at macroscopic interfaces adjoined by at least one inhomogeneous
material, or in the presence of marked compositional or load gradients, the latter be-
ing e.g. present in inhomogeneous beams under bending loads), special homogenization
methods must be used, such as second order schemes explicitly accounting for deforma-
tion gradients (Kouznetsova et al.,2002), and the homogenized continuum is nonlocal,
see e.g. (Feyel,2003).
An important classification scheme for inhomogeneous materials is based on the mi-
croscopic phase topology. In matrix–inclusion arrangements (as found in “particulate
materials” such as most composite materials, porous materials, and closed cell foams 3) )
only the matrix shows a connected topology and the constituents play clearly distinct
roles. In interwoven phase arrangements (as found e.g. in functionally graded materials,
in open cell foams, or in woven and braided tissues) and in many polycrystals (“granular
materials”), in contrast, the phases cannot be distinguished on the basis of topology.
2)
This nomenclature is far from universal, the naming of the length scales of inhomogeneous
materials being notoriously inconsistent in the literature.
3)
With respect to their thermomechanical behavior, porous and cellular materials can
usually be treated as inhomogeneous materials in which one constituent has vanishing stiffness,
thermal expansion, conductivity etc.
–2–
1.2 Homogenization and Localization
For any region of an inhomogeneous material the microscopic strain and stress fields, ε(x)
and σ(x), and the corresponding macroscopic responses, hεi and hσi, can be formally
linked by localization (or projection) relations of the type
ε(x) = A(x)hεi
σ(x) = B(x)hσi . (2)
If, in addition, the region is sufficiently large and contains no significant macroscopic
gradients of stress, strain or composition, homogenization relations can be defined as
1 1
Z Z
hεi = ε(x) dΩ = (u(x) ⊗ nΓ + nΓ ⊗ u(x)) dΓ
Ω s Ωs 2Ωs Γs
1 1
Z Z
hσi = σ(x) dΩ = t(x) ⊗ x dΓ , (3)
Ω s Ωs Ω s Γs
where Ωs and Γs stand for the volume and the surface of the region under consideration,
u(x) is the deformation vector, t(x) = σ(x)nΓ are the surface traction vectors, nΓ are
surface normal vectors and ⊗ is the dyadic product of vectors defined as a⊗b=[a i bj ].
A(x) and B(x) are called mechanical strain and stress concentration tensors (or influence
functions; Hill,1963), respectively. The surface integral formulation for hεi given above
holds only for perfect interfaces between the constituents (otherwise correction terms
involving the displacement jumps across failed interfaces must be added, see e.g. (Nemat–
Nasser/Hori,1993)), in which case the mean strains and stresses in a control volume, hεi
and hσi, are fully determined by the surface displacements and tractions. In the absence
of body forces the microstresses σ(x) are self-equilibrated (but typically not zero). In
the above form, eqn.(2) applies only to linear elastic behavior, but it can be easily
modified to cover thermoelastic behavior, compare eqn.(16), and the formal extension
to the nonlinear range (e.g. to elastoplastic materials described by secant or incremental
plasticity) is rather straightforward, see eqn.(41). For a discussion of averaging at finite
deformations see e.g. (Nemat–Nasser,1999).
Note that eqns.(1) and (3) immediately imply that volume averages of fluctuations vanish
for sufficiently large integration volumes vanish,
1 1
Z Z
0
ε (x) dΩ = 0 = σ 0 (x) dΩ , (4)
Ω s Ωs Ω s Ωs
4)
Whereas integrals over products of slow and fast variables vanish, integrals over products
of fluctuating components do not vanish in general, compare section 2.4.
–3–
can be shown to hold for general statically admissible stress fields σ ∗ and kinematically
admissible strain fields ε∗ , see (Hill,1967). Equation (5) is known as Hill’s macrohomo-
geneity condition or the Mandel–Hill condition, compare (Bornert,2001; Zaoui,2001). For
the special cases of homogeneous stress and strain boundary conditions (for which the
above relation can be shown to hold by transforming the volume integrals into surface
integrals, to which the fluctuations σ 0 and ε0 give zero contributions) it is referred to
as Hill’s lemma. Such conditions imply that the volume averaged strain energy density
of an inhomogeneous material can be obtained from the volume averages of the stresses
and strains, provided the micro- and macroscales are sufficiently different. Accordingly,
homogenization can be interpreted as finding a homogeneous comparison material that
is energetically equivalent to a given microstructured material.
The microgeometries of real inhomogeneous materials are at least to some extent random
and, in nearly all cases of practically relevance, they are highly complex. Accordingly,
exact expressions for A(x), B(x), ε(x), σ(x) etc. cannot realistically be provided and
approximations have to be introduced. Typically, these approximations are based on the
ergodic hypothesis, the heterogeneous material being assumed to be statistically homo-
geneous. Thus, sufficiently large subvolumes (mesodomains, homogenization volumes)
selected randomly within the sample are taken to give rise to the same averaged mate-
rial properties which, in turn, correspond to the material’s overall or effective material
properties5) .
Ideally, the homogenization volume should be chosen to be a proper reference volume
element (RVE), i.e., a subvolume of Ωs that is statistically representative of the micro-
geometry of the material6) . Such a volume element must sufficiently large to allow a
meaningful sampling of the microfields and sufficiently small for the influence of macro-
scopic gradients to be negligible and for an analysis of the microfields to be possible7) .
The volume averages of stresses, strains, etc. over a RVE are equal to the correspond-
ing ensemble averages over many RVEs. For discussions of homogenization volumes
and RVEs see e.g. (Hashin,1983; Nemat–Nasser/Hori,1993; Drugan/Willis,1996; Ostoja–
Starzewski,1998; Zaoui,2001; Bornert et al.,2001; Stroeven et al.,2002; Kanit et al.,2003).
In practice it may be rather difficult to study (or, in fact, identify) proper RVEs of actual
5)
Note that some inhomogeneous materials, such as graded materials, are not statistically
homogeneous and, consequently, may require nonstandard treatment. For such materials it is
not possible to define effective material properties in the sense of eqn.(6) and reference volume
elements may not exist.
6)
Alternatively a RVE may be defined as a volume that shows the same overall material
properties irrespective of the boundary conditions applied (Sab,1992; Huet,1999). The latter
definition of an RVE, however, gives rise to a dependence of the size of the RVE on the physical
property considered and on the pertinent phase contrast (Kanit et al.,2003).
7)
This requirement was symbolically denoted as MICROMESOMACRO by Hashin
(1983), where MICRO and MACRO have their “usual” meanings and MESO stands for the
length scale of the homogenization volume. As noted by Nemat–Nasser (1999) it is the dimension
relative to the microstructure essential for a given problem that is important for the size of an
RVE.
–4–
inhomogeneous materials and, accordingly, approximations to them are employed e.g. in
investigations based on analyzing discrete microgeometries.
–5–
1.3 Overall Behavior, Material Symmetries
The homogenized strain and stress fields of an elastic inhomogeneous material as obtained
by eqn.(3), hεi and hσi, can be linked by effective elastic tensors E∗ and C∗ as
hσi = E∗ hεi
hεi = C∗ hσi , (6)
1
Z
∗
E = E(x)A(x) dΩ
Ω s Ωs
1
Z
∗
C = C(x)B(x) dΩ . (7)
Ω s Ωs
Two independent parameters are sufficient for describing such an overall linear elastic
behavior (e.g. the effective Young’s modulus E ∗ , the effective Poisson number ν ∗ , the
8)
Sometimes, the term “effective material properties” is employed to explicitly imply the
response of a proper RVE or of an infinitely large sample, whereas the term “apparent” is used
for smaller volume elements (Huet,1990).
–6–
effective shear modulus G∗ , the effective bulk modulus K ∗ =E ∗ /3(1−2ν ∗ ), or the effective
Lamé constants) and one is required for the effective thermal expansion behavior in the
linear range (the effective coefficient of thermal expansion α∗ = α11 ).
The effective elasticity and thermal expansion tensors for statistically transversely
isotropic materials have the structure
E11 E12 E12 0 0 0 α11
E12 E22 E23 0 0 0 α22
E E23 E22 0 0 0 α
E∗ = 12 α∗ = 22 , (9)
0 0 0 E44 0 0 0
0 0 0 0 E44 0 0
0 0 0 0 0 E66 = 21 (E22 − E23 ) 0
where 1 is the axial direction and 2–3 is the transverse plane of isotropy. Generally, the
thermoelastic behavior of transversely isotropic materials is described by five indepen-
dent elastic constants and two independent coefficients of thermal expansion. Appropri-
ate elastic parameters in this context are e.g. the axial and transverse effective Young’s
∗ ∗
moduli, EA and ET , the axial and transverse effective shear moduli, G∗A and G∗T , the
∗ ∗
axial and transverse effective Poisson numbers, νA and νT , as well as the effective trans-
∗ ∗ ∗ ∗ ∗ ∗2
verse bulk modulus KT =EA /2[(1 − νT )(EA /ET ) − 2νA ]. The transverse (“in-plane”)
properties are related via G∗T =ET ∗
/2(1 + νT ∗
), but for general transversely isotropic ma-
∗
terials there is no linkage between the axial properties EA , G∗A and νA ∗
beyond the above
∗
definition of KT . Both an axial and a transverse effective coefficient of thermal expan-
sion, α∗A = α11 and α∗T = α22 , are required. For the special case of materials reinforced
by aligned continuous fibers Hill’s (1964) relations
(f)
4(νA − ν (m) )2
∗ (f) (m) ξ 1−ξ 1
EA = ξEA + (1 − ξ)E + (f) (m) (f)
+ (m) − ∗
(1/KT − 1/KT )2 KT KT KT
(f)
νA − ν (m)
∗ (f) (m) ξ 1−ξ 1
νA = ξνA + (1 − ξ)ν + (f) (m) (f)
+ (m) − ∗ (10)
1/KT − 1/KT KT KT KT
∗ ∗ ∗
allow the effective moduli EA and νA to be expressed by KT , some constituent properties,
and the fiber volume fraction ξ. Equations (10) can be used to reduce the number of in-
dependent effective elastic parameters required for unidirectional continuously reinforced
∗
composites to three (usually KT , G∗A and G∗T ).
It is worth noting that the overall material symmetries of inhomogeneous materials and
their effect on various physical properties can be treated in full analogy to the symmetries
of crystals as discussed e.g. in (Nye,1957). The influence of the overall symmetry of the
phase arrangement on the overall mechanical behavior of inhomogeneous materials can be
marked9) , especially on the post-yield and other nonlinear responses to mechanical loads,
compare e.g. (Nakamura/Suresh,1993; Weissenbek,1994). Accordingly, it is important to
aim at closely following the symmetry of the actual material in any modeling effort.
9)
Overall properties described by lower order tensors, e.g. the thermal expansion response,
are much less sensitive to symmetry effects, compare (Nye,1957).
–7–
1.4 Basic Modeling Strategies in Continuum Micromechanics of Materials
10)
It should be noted that the overall thermomechanical behavior of homogenized mate-
rials is typically richer than that of the constituents, i.e. the effects of the interaction of the
constituents in many cases cannot be satisfactorily described by simply adapting material pa-
rameters and not changing the functional relationships in the constitutive laws (for example,
two constituents showing viscoplastic power law behavior with different exponents in general
do not give rise to a power-law type overall response). Accordingly, using predicted uniaxial
stress–strain curves for “calibrating” the free parameters of macroscopic constitutive models
may not result in satisfactory descriptions for the multiaxial material response.
–8–
(limited) statistical information11) :
• Mean Field Approaches (MFAs) and related methods (see chapter 2): The mi-
crofields within each constituent are approximated by their phase averages hεi (p)
and hσi(p) , i.e. piecewise (phasewise) uniform stress and strain fields are employed.
Such descriptions typically use information on the microscopic topology, the in-
clusion shape and orientation, and, to some extent, on the statistics of the phase
distribution. The localization relations then take the form
1
Z X
(p)
hεi = (p) ε(x) dΩ with hεi = V (p) hεi(p)
Ω Ω(p) p
1
Z X
hσi(p) = σ(x) dΩ with hσi = V (p) hσi(p) (12)
Ω(p) Ω(p) p
where (p) denotes a given phase of the material, Ω(p) is the corresponding phase
volume, and V (p) =Ω(p) / k Ω(k) is the volume fraction of the phase. Note that
P
in contrast to eqn.(2), for MFAs the phase concentration tensors Ā and B̄ are
not functions of the spatial coordinates within the reference volume. Mean field
approaches tend to be formulated in terms of the phase concentration tensors, they
pose relatively low computational requirements, they have been highly successful
in describing the thermoelastic response of inhomogeneous materials, and their use
for modeling nonlinear inhomogeneous materials is a subject of active research.
• Variational Bounding Methods (see chapter 3): Variational principles are used
to obtain upper and (in many cases) lower bounds on the overall elastic tensors,
elastic moduli, secant moduli, and other physical properties of inhomogeneous ma-
terials. Many analytical bounds are obtained on the basis of phase-wise constant
stress (polarization) fields. Bounds — aside from their intrinsic interest — are im-
portant tools for assessing other models of inhomogeneous materials. In addition,
in many cases one of the bounds provides good estimates for the physical property
under consideration, even if the bounds are rather slack (Torquato,1991). Many
bounding methods are closely related to MFAs.
The second group of approximations is based on studying discrete microstructures and
includes
11)
General phase arrangements can be described by n-point correlation functions that give
the probability that a some geometrical configuration of n points lies within prescribed phases.
For statistically isotropic materials 2-point correlations correspond to the volume fractions.
–9–
• Periodic Microfield Approaches (PMAs) or Unit Cell Methods (see chapter 5):
The real inhomogeneous material is approximated by an infinitely extended model
material with a periodic phase arrangement12) . The corresponding periodic mi-
crofields are usually evaluated by analyzing unit cells (which may describe mi-
crogeometries ranging from rather simplistic to highly complex) via analytical or
numerical methods. Unit cell methods are typically used for performing materials
characterization of inhomogeneous materials in the nonlinear range, but they can
also be employed as micromechanically based constitutive models. The high res-
olution of the microfields provided by PMAs can be very useful for studying the
initiation of damage at the microscale. However, because they inherently give rise
to periodic configurations of damage, PMAs are not well suited for investigating
phenomena such as the interaction of the microstructure with macroscopic cracks.
Periodic microfield approaches can give detailed information on the local stress
and strain fields within a given unit cell13) , but they tend to be computationally
expensive.
• Embedded Cell Approaches (ECAs; see chapter 6): The real inhomogeneous ma-
terial is approximated by a model material consisting of a “core” containing a
discrete phase arrangement that is embedded within some outer region to which
far field loads or displacements are applied. The material properties of this outer
region may be described by some macroscopic constitutive law, they can be deter-
mined self-consistently or quasi-self-consistently from the behavior of the core, or
the embedding region may take the form of a coarse description and/or discretiza-
tion of the phase arrangement. ECAs can be used for materials characterization,
and they are usually the best choice for studying regions of special interest, such
as crack tips and their surroundings, in inhomogeneous materials. Like PMAs,
embedded cell approaches can resolve local stress and strain fields in the core
region at high detail, but tend to be computationally expensive.
• For small inhomogeneous samples (i.e. samples the size of which exceeds the mi-
croscale by not more than, say, an order of magnitude, the full microstructure can
be modeled and studied, see e.g. (Guidoum,1994; Papka/Kyriakides,1994; Silber-
schmidt/Werner,2001). In such cases boundary effects (and thus the boundary
12)
Whereas mean field methods make the implicit assumption that there is no long range
order in the inhomogeneous material, periodic phase arrangements imply the existence of just
such an order.
13)
Due to limitations in available computing power the phase arrangements studied in pe-
riodic microfield and embedded cell approaches in practice are typically too restricted in com-
plexity to be representative volume elements in the strict sense.
– 10 –
conditions applied to the sample) typically play a prominent role in its behavior.
Note that no scale transition is involved in problems of this type.
Figure 1 shows a sketch of a microstructure as well as PMA, ECA and windowing ap-
proximations applied to it.
Further descriptions of inhomogeneous materials, such as rules of mixtures (isostrain
and isostress models) and semiempirical formulae such as the Halpin–Tsai equations
(Halpin/Kardos,1976), are not discussed here due to their typically rather weak physical
background and their rather limited predictive capabilities. For brevity, also a number
of models with solid physical backgrounds, such as expressions for self-similar compos-
ite sphere assemblages (CSA, Hashin,1962) and composite cylinder assemblages (CCA,
Hashin/Rosen,1964), information theoretical approaches (Kreher,1990), and the semiem-
pirical methods proposed for describing the elastoplastic behavior of composites rein-
forced by aligned continuous fibers (Dvorak/Bahei-el-Din,1987), are not covered within
the present discussions.
– 11 –
For studying materials that are inhomogeneous at a number of (sufficiently widely spread)
length scales (e.g. materials in which well defined clusters of inclusions are present),
hierarchical procedures that use homogenization at more than one level are a natural
extension of the above concepts. Such multi-scale models will be the subject of a short
discussion in chapter 8.
– 12 –
2. MEAN FIELD APPROACHES
For brevity the following treatment concentrates mainly on Mori–Tanaka (effective field)
methods, which may be viewed as the simplest mean field approaches for inhomogeneous
materials that encompass the full physical range of phase volume fractions14) . Only
two-phase materials are covered explicitly and perfect bonding between constituents is
assumed. Unless specifically stated otherwise, the material behavior of both inclusion and
matrix is taken to be linear (thermo)elastic. It should be noted that there is an extensive
body of literature covering mean field approaches (most of which can be classified as
effective field, self-consistent, and differential theories) as well as related methods.
For linear thermoelastic materials, depending on the boundary conditions the overall
stress–strain relations can be denoted in the form
hσi = E∗ hεi + e∗ ∆T
hεi = C∗ hσi + α∗ ∆T (13)
where e∗ = −E∗ α∗ is the overall specific thermal stress tensor (i.e. the overall stress
response of the fully constrained composite to a purely thermal unit load) and ∆T is
the (spatially uniform) temperature difference with respect to some stress-free reference
temperature. The constituents, here a matrix (m) and inclusions (i) , are also taken to
behave thermoelastically, i.e.
Here ξ=V (i) =Ω(i) /(Ω(i) + Ω(m) ) stands for the volume fraction of the inclusions and
εa =C∗ σa is an applied strain or the elastic strain response of the composite to a purely
mechanical applied load σa . In the isothermal case, the equalities hεi=εa and hσi=σa
14)
Because Eshelby and Mori–Tanaka methods are specifically suited for matrix–inclusion-
type microtopologies, the expression “composite” is often used in the present chapter instead of
the more general designation “inhomogeneous material”.
– 13 –
correspond to homogeneous strain and homogeneous stress boundary conditions, respec-
tively15) , perfectly bonded interfaces being also assumed in the former case.
The phase averaged strains and stresses can be related to the overall strains, stresses,
and temperature changes by the elastic and thermal phase strain and stress concentration
tensors Ā, ā, B̄, and b̄ (Hill,1963), respectively, which for thermoelastic inhomogeneous
materials are defined by the expressions
(compare eqn.(11) for the isothermal elastic case). By using eqns.(15) and (16), the
phase strain and stress concentration tensors (or strain and stress concentration tensors
for short) can be shown to fulfill the relations
From eqn.(7) the effective elasticity and compliance tensors of the composite can be
obtained from the properties of the phases and from the concentration tensors as16)
The effective specific thermal stress tensor and the effective thermal expansion coefficient
tensor can be written as
respectively.
The stress and strain concentration tensors for a given phase are linked to each other by
expressions such as
Ā(m) = C(m) B̄(m) E(i) [I + (1 − ξ)(C(m) − C(i) )B̄(m) E(i) ]−1 = C(m) B̄(m) E∗
B̄(m) = E(m) Ā(m) C(i) [I + (1 − ξ)(E(m) − E(i) )Ā(m) C(i) ]−1 = E(m) Ā(m) C∗ . (20)
15)
For more general boundary conditions it can be shown by using the Gauss theorem
that the average stresses in a volume element of an inhomogeneous material with perfect phase
boundaries are fully determined by the tractions on its boundaries and the average strains are
determined by the displacements at the boundaries, compare eqn.(3).
16)
It is interesting to note that, because E∗ and its inverse, C∗ , depend not only on the
properties of the constituents, but also on the concentration tensors and thus on the microscopic
phase arrangement, it may be argued that overall properties of inhomogeneous materials, e.g.
their effective Young’s modulus, are not material parameters in the strict sense.
– 14 –
In addition, relations can be derived that connect the thermal strain and stress concen-
tration tensors of a given phase with the elastic strain and stress concentration tensors,
respectively (see e.g. Benveniste/Dvorak,1990; Benveniste et al.,1991), typical expres-
sions being17)
From eqns.(18) to (21) it is evident that the knowledge of one elastic phase concentration
tensor is sufficient for describing the full thermoelastic behavior of a two-phase inhomo-
geneous material within the mean field framework18) . There are a number of additional
relations that allow e.g. the phase concentration tensors to be obtained from the overall
and phase elastic tensors, see e.g. (Böhm,1998).
The macrohomogeneity condition and Hill’s lemma, eqn.(5), can be denoted for elastic
inhomogeneous materials as
X 1 Z T T
T
hσ εi = (p)
εT (x) E(p) ε(x) dΩ = hεi hσi = hεi E∗ hεi . (22)
p
Ω Ω(p)
For detailed discussions of Hill’s lemma for the case of periodicity boundary conditions
see e.g. (Suquet,1987; Michel et al.,1999).
εc = Sεt , (23)
where S is called the (interior point) Eshelby tensor. For eqn.(23) to hold, εt may be any
kind of eigenstrain which is uniform over the inclusion (e.g. a thermal strain, or a strain
17)
Relations for the inclusion concentration tensors corresponding to eqns.(20) and (21) take
analogous forms.
18)
Similarly, n−1 elastic phase concentration tensors must be known for describing the
overall thermoelastic behavior of an n-phase material.
19)
This “Eshelby property” is limited to ellipsoidal inclusions (Lubarda/Markenscoff,1998).
– 15 –
due to some phase transformation which involves no changes in the elastic constants of
the inclusion).
For mean field descriptions of dilute matrix–inclusion composites, of course, it is the
stress and strain fields in inhomogeneous inclusions (“inhomogeneities”) embedded in a
matrix that are of interest. Such cases can be handled on the basis of Eshelby’s theory for
homogeneous inclusions, eqn.(23), by introducing the concept of equivalent homogeneous
inclusions. This strategy involves replacing an actual perfectly bonded inhomogeneous
inclusion (which has different material properties than the matrix) subjected to a given
unconstrained eigenstrain εt with a (fictitious) “equivalent” homogeneous inclusion on
which an appropriate (fictitious) “equivalent” eigenstrain ετ is made to act. This equiv-
alent eigenstrain must be chosen in such a way that the same stress and strain fields,
hσi(i) and hεi(i) =εc , are obtained in the constrained state for the actual inhomogeneous
inclusion and the equivalent homogeneous inclusion. Following (Withers et al.,1989) this
process can be visualized as consisting of “cutting and welding” operations as shown in
fig. 2.
The concept of the equivalent homogeneous inclusion can be extended to cases where
a uniform mechanical strain εa or external stress σa is applied to a perfectly bonded
– 16 –
inhomogeneous elastic inclusion in an infinite matrix. Here, the strain in the inclusion,
hεi(i) , will be a superposition of the applied strain and of an additional term εc due to the
constraint effects of the surrounding matrix. A fair number of different expressions for
concentration tensors obtained by such procedures have been reported in the literature,
among the simplest being those proposed by Hill (1965) and elaborated by Benveniste
(1987). For deriving them, the condition of equal stresses and strains in the actual
inclusion (elasticity tensor E(i) ) and the equivalent inclusion (elasticity tensor E(m) ) under
an applied far field strain εa is denoted as
where eqn.(23) is used to describe the constrained strain of the equivalent homogeneous
inclusion. From these expressions and from eqn.(16) the inclusion strain concentration
tensor can be obtained as
(i)
Ādil = [I + SC(m) (E(i) − E(m) )]−1 (25)
Equations (25) and (26) were derived under the assumption that the inclusions are di-
lutely dispersed in the matrix and thus do not “feel” any effects due to their neighbors
(i.e. they are loaded by the unperturbed applied stress σa or applied strain εa , the so
called dilute case). Accordingly, they are independent of the inclusion volume fraction ξ.
The stress and strain fields outside a transformed homogeneous or inhomogeneous inclu-
sion in an infinite matrix are not uniform on the microscale20) (Eshelby,1959). Within
the framework of mean field theories, which aim to link the average fields in matrix and
inclusions with the overall response of inhomogeneous materials, however, it is only the
average matrix stresses and strains that are of interest. For dilute composites, such ex-
pressions follow directly by combining eqns.(25) and (26) with eqn.(17). Estimates for
the overall elastic and thermal expansion tensors can, of course, be obtained in a straight-
forward way from the concentration tensors by using eqns.(19) and (21). It must be kept
in mind, however, that the dilute expressions are strictly valid only for vanishingly small
inclusion volume fractions and should be used only for ξ 0.1.
The Eshelby tensor S depends only on the material properties of the matrix and on
the aspect ratio a of the inclusions, i.e. expressions for the Eshelby tensor of ellipsoidal
20)
The fields outside a single inclusion can be described via the (exterior point) Eshelby
tensor, see e.g. (Ju/Sun,1999). From the (constant) interior point fields and the (position)
dependent exterior point fields the stress and strain jumps at the interface between inclusion
and matrix can be evaluated.
– 17 –
inclusions are independent of the material symmetry and properties of the inclusions. Ex-
pressions for the Eshelby tensor of spheroidal inclusions in an isotropic matrix are given
e.g. in (Pedersen,1983; Tandon/Weng,1984; Mura,1987; Clyne/Withers,1993) 21) , the for-
mulae being very simple for continuous fibers (a → ∞), spherical inclusions (a = 1) and
thin circular discs (a → 0). Analytical expressions for the Eshelby tensor can also be
given for spheroidal inclusions in a matrix of transversely isotropic (Withers,1989) or
cubic material symmetry (Mura,1987), provided the material axes of the matrix con-
stituents are aligned with the orientations of nonspherical inclusions. In cases where
no analytical solutions are available, the Eshelby tensor can be evaluated numerically,
compare (Gavazzi/Lagoudas,1990).
Mori–Tanaka-Type Estimates
One way for achieving this consists of approximating the stresses acting on an inclusion,
which may be viewed as the perturbation stresses caused by the presence of other inclu-
sions (“image stresses”, “background stresses”, “mean field stresses”, “effective stresses”)
superimposed on the applied far field stress, by an appropriate average matrix stress
hσi(m) . The idea of combining this concept of an average matrix stress with Eshelby-type
equivalent inclusion approaches goes back to Brown and Stobbs (1971) as well as Mori
and Tanaka (1973). Effective field theories of this type are generically called Mori–Tanaka
methods or “Equivalent Inclusion — Average Stress” (EIAS) approaches.
It was pointed out by Benveniste (1987) that in the isothermal case the central assump-
tion involved in Mori–Tanaka approaches can be denoted as
21)
Note that instead of evaluating the Eshelby tensor for a given configuration, the so
called mean polarization factor tensor Q=E(m) (I − S) may be evaluated instead, see e.g. (Ponte
Castañeda,1996).
– 18 –
These expressions can be viewed as modifications of eqn.(11) in which the applied strain
and stress, εa and σa , are replaced by the (unknown) matrix strain or stress, hεi(m) and
hσi(m) , respectively, according to the EIAS strategy.
Based on eqn.(27) very simple and straightforward Mori–Tanaka-type expressions for the
matrix strain and stress concentrations tensors were proposed by (Benveniste,1987) as
(m) (i)
ĀM = [(1 − ξ)I + ξ Ādil ]−1
(m) (i)
B̄M = [(1 − ξ)I + ξ B̄dil ]−1 , (28)
the corresponding relations for the inclusion strain and stress concentration tensors fol-
lowing from eqns.(27) as
(i) (i) (i)
ĀM = Ādil [(1 − ξ)I + ξ Ādil]−1
(i) (i) (i)
B̄M = B̄dil [(1 − ξ)I + ξ B̄dil ]−1 , (29)
Equations (28) and (29) may be evaluated with any strain and stress concentration
tensors pertaining to dilute inclusions in a matrix. If, for example, eqns.(25) and (26)
are employed, the Mori–Tanaka matrix strain and stress concentration tensors take the
form
(m)
ĀM = {(1 − ξ)I + ξ[I + SC(m) (E(i) − E(m) )]−1 }−1
(m)
B̄M = {(1 − ξ)I + ξE(i) [I + SC(m) (E(i) − E(m) )]−1 C(m) }−1 , (30)
which can be easily modified for porous materials by letting E(i) → 0 to give
ξ −1
E∗M,por = E(m) I +
(I − S)−1 . (32)
1−ξ
Equation (32) should not, however, be used for void volume fractions that are much in
excess of, say, ξ = 0.2522) .
22)
Mori–Tanaka expressions for two-phase materials are based on the assumption that the
shape of the inhomogeneities can be described by ellipsoids of a given aspect ratio throughout
the deformation history. In the case of porous materials with high void volume fractions, e.g.
cellular materials and foams, deformation at the microscale takes place mainly by bending and
buckling of the matrix “walls” between voids, see e.g. (Gibson/Ashby,1988). Such “large strain”
effects cannot be properly described by Mori–Tanaka approaches, which, consequently, tend to
markedly overestimate the effective stiffness of cellular materials.
– 19 –
FIG.3: Sketch of ellipsoidal inclusions in an aligned ellipsoidally distributed spatial
arrangement as used implicitly in Mori–Tanaka-type approaches (a=2.0).
– 20 –
Self-Consistent Estimates
Another group of estimates for the overall thermomechanical behavior of inhomogeneous
materials are self-consistent schemes (or effective medium theories), in which an inclusion
or some phase arrangement is embedded in the effective material (the properties of which,
of course, are not known a priori)24) . Figure 4 shows a schematic comparison of the
material and loading configurations underlying Eshelby, Mori–Tanaka, and self-consistent
methods.
i
i
i i
m
m m eff eff
σa σ(m)
tot
σa σa
dilute MTM CSCS GSCS
FIG.4: Schematic comparison of the Eshelby method for dilute composites, Mori–
Tanaka approaches, and classical as well as generalized self-consistent schemes.
If the kernel consists of an inclusion only, classical (or two-phase) self-consistent schemes
are obtained (CSCS, see e.g. Hill,1965), which formally can be described by relations
such as
E∗SCS = E(m) + ξ[E(i) − E(m) ]Āi,∗
dil . (33)
Because Āi,∗
dil , the strain concentration tensor of an inclusion embedded in the effective
medium in analogy to eqn.(25), is a function of E∗SCS , this expression is implicit and
E∗SCS has to be solved for by self-consistent iteration using a scheme of the type
The Eshelby tensor used in this expression, Sn , has to be recomputed for each iteration.
Generally, classical self-consistent schemes are best suited for describing the overall elastic
properties of two-phase materials that do not show a matrix–inclusion microtopology
in at least part of their range of volume fractions (e.g. polycrystals, for which multi-
phase expressions analogous to eqn.(33) are often used, and many functionally graded
materials)25) , but they have also been employed for studying particle reinforced MMCs in
which the grain size of the matrix and the size of the particulates are similar (Corvasce et
24)
Within the classification of micromechanical methods given in section 1.4, self-consistent
mean field methods can also be viewed as as analytical ECAs.
25)
Classical self consistent schemes can be shown to correspond to perfecly disordered ma-
terials (Kröner,1978) or self-similar hierarchical materials (Torquato,2001). There is also a
realization in terms of a single-scale dispersion (Torquato/Hyun,2001).
– 21 –
al.,1991). When microstructures described by the CSCS show a geometrical anisotropy
of the transversely isotropic type (e.g. in terms of the free lengths in the phases in a
metallographical sections), this anisotropy determines the “inclusion” aspect ratio to be
used.
For porous materials classical self-consistent schemes predict a breakdown of stiffness
due to the percolation of the pores at ξ = 21 for spherical voids and ξ = 13 for aligned
cylindrical voids (Torquato,2001).
with the initial conditions E∗D =E(m) and C∗D =C(m) , respectively, at ξ=0. In analogy to
eqn.(33) Āi,∗ i,∗ ∗
dil and B̄dil depend on ED . Differential schemes describe aligned matrix–
inclusion microgeometries with markedly polydisperse distributions of inclusion sizes
(corresponding to the repeated homogenization steps) and, accordingly, they have not
been used extensively for studying composite materials.
Recently second order estimates for the thermoelastic properties of two-phase materials
were developed by Torquato (1998), which correspond to the same phase arrangements
and use the same three-point microstructural parameters η and ζ as discussed for three-
point bounds in chapter 3.2. These estimates lie between the corresponding bounds and
give excellent analytical predictions for the overall thermoelastic response of inhomoge-
neous materials that show the appropriate microstructures and are free of flaws.
where τ (x) is known as the polarization tensor and E0 is the elastic tensor of the compar-
ison material. Equation (36) can be interpreted as using the polarizations for describing
the “fast” contributions to the actual stress field, σ(x), while the “slow” contributions
are covered by the behavior of the reference material. Phase-wise constant polarization
tensors
τ (p) = (E(p) − E0 )hεi(p) (37)
– 22 –
can be obtained by phase averaging over phase (p).
Equation (37) can be used as the starting point for deriving general mean field strain
concentration tensors of the type
(p) −1
ĀHS = (L0 + E(p) )−1 ξ(L0 + E(i) )−1 + (1 − ξ)(L0 + E(m) )−1
. (38)
This expression leads to estimates of the overall elastic stiffness of microstructured two-
phase materials in the form
−1
E∗HS = E(m) + ξ(L0 − E(m) ) I + (1 − ξ)(L0 + E(m) )−1 (E(i) − E(m) )
, (39)
Here the Eshelby tensor S0 must be evaluated with respect to the comparison mate-
rial. L0 can be shown to depend on the two-point statistics of the microgeometrical
arrangement, see (Bornert,2001).
Standard mean field models can be obtained as special cases of Hashin–Shtrikman meth-
ods by appropriate choices of the comparison material. For example, using the matrix as
the comparison material results in Mori–Tanaka methods, whereas the selection of the
effective material as comparison material leads to classical self-consistent schemes.
The above Hashin–Shtrikman formalism can be extended in a number of ways, one of
which consists in not only studying inhomogeneities but more complex geometrical enti-
ties, called patterns or motifs, embedded in a matrix, see (Bornert,1996; Bornert,2001).
Typically the stress and strain fields in such patterns are inhomogeneous even in the di-
lute case and numerical methods have to be resorted to in evaluating the corresponding
overall constraint tensors. A simple example of such a pattern is a spherical particle or
cylindrical fiber surrounded by a layer of matrix at an appropriate volume fraction, which,
when embedded in the effective material, allows to recover the three-phase or generalized
self-consistent scheme (GSCS) of (Christensen/Lo,1979; Christensen/Lo,1986), compare
fig. 4. GSCS-expressions for the overall elastic moduli are obtained by considering the
differential equations describing the elastic response of such three-phase regions under ap-
propriate boundary and loading conditions26) . The GSCS in its original version is limited
to inclusions of spherical or cylindrical shape, where it describes microgeometries of the
CSA and CCA types, respectively, and gives rise to third order equations for the shear
modulus G and the transverse shear modulus GT , respectively (for the other moduli,
CSA and CCA results are available). Extensions of the GSCS to multi-layered spheri-
cal inclusions (Hervé/Zaoui,1993) and to aligned ellipsoidal inclusions (Huang/Hu,1995),
can be found in the literature.
26)
Even though the GSCS is not a mean field method in the strict sense, pertinent phase
concentration tensors can easily be obtained from the overall elastic tensors, so that mean field
formalisms may be used to generate e.g. effective coefficients of thermal expansion.
– 23 –
Generalized self-consistent schemes give excellent results for inhomogeneous materials
with matrix–inclusion topologies and are accordingly highly suited for obtaining esti-
mates for the thermoelastic moduli of composite materials reinforced by (approximately)
spherical particles or aligned continuous fibers. Because self-consistent schemes are by
definition implicit methods, their computational requirements are in general higher than
those of Mori–Tanaka-type approaches. Like Mori–Tanaka approaches, they can form
the basis for describing the behavior of nonlinear inhomogeneous materials.
Descriptions for viscoelastic microstructured materials are closely related to those for
elastic composites. Relaxation moduli and creep compliances can be obtained by ap-
plying micromechanical methods in the Laplace transform domain, where the problems
are analogous to elastic ones for the same microgeometries. Standard micromechanical
methods can be used in conjunction with complex moduli for steady state vibration anal-
yses of viscoelastic composites. For correspondence principles between descriptions for
elastic and viscoelastic inhomogeneous materials see e.g. (Hashin,1983). The extension
of mean field estimates and bounding methods to elastoplastic, viscoelastoplastic, and
damaging inhomogeneous materials, however, has proven to be challenging.
The aim of mean field models of microstructured materials with at least one elastoplastic
constituent consists in providing accurate estimates for the material response for any
load state and load history at reasonable computational cost. The main difficulties in
attaining this goal lie in the typically strong intra-phase fluctuations of the stress and
strain fields in elastoplastic inhomogeneous materials and in the hereditary nature of
plasticity. The responses of the constituents can thus vary markedly at the microscale,
with each point following its own trajectory in stress space. Accordingly, a 2-phase
elastoplastic composite effectively behaves as a multiphase material and phase averages
are less useful descriptors than in the linear elastic case.
Mean field models of elastoplastic inhomogeneous materials27) typically are based on
solving sequences or sets of linearized problems in terms of linear inhomogeneous ref-
erence media. Accordingly, choices have to be made with respect to the linearization
procedure, the linear homogenization model, and the phase-wise equivalent stresses and
equivalent strains to be used in evaluating the elastoplastic constituent material behavior,
see (Zaoui,2001).
In the literature several lines of development of mean field and Hashin–Shtrikman-type
approaches for elastoplastic inhomogeneous materials can be found. Historically, the
most important of them have been secant plasticity concepts based on deformation
theory, see e.g. (Tandon/Weng,1988; Dunn/Ledbetter,1997), and incremental plasticity
27)
The present chapter is restricted to continuum mechanical descriptions of thermoelasto-
plastic composites. In a separate line of development, phase averaged stress fields obtained
by Eshelby-type methods have been used in dislocation-based descriptions of nonlinear ma-
trix behavior, e.g. for studying the response of MMCs to temperature excursions, see e.g.
(Taya/Mori,1987).
– 24 –
models, compare e.g. (Hill,1965; Lagoudas et al.,1991; Karayaka/Sehitoglu,1993; Pet-
termann,1997). In addition, “tangent” concepts (Molinari et al.,1987) and, recently, the
affine formulation (Masson et al.,2000), which treats the linearized constitutive equations
as thermoelastic ones, have been proposed. A recent overview of mean field methods for
elastoplastic composites can be found in (Ponte Castañeda/Suquet,1998).
Secant models aim at directly arriving at solutions in terms of the overall response for
a given load state. They are based on the deformation theory of plasticity and can be
formulated in terms of potentials, which allows for a concise mathematical presentation
(Bornert/Suquet,2001). Secant models are limited to monotonic loading and radial (or
approximately radial) trajectories of the constituents in stress space during loading 28) ,
which precludes their use as micromechanically based constitutive models in multi-scale
analyses. Advanced secant formulations, however, are highly suitable for materials char-
acterization, where they give excellent results.
Incremental plasticity (and affine) approaches are not subject to limitations with respect
to loadint paths. However, especially for matrix dominated deformation modes they tend
to markedly overestimate the overall strain hardening in the post-yield regime, compare
e.g. (Gilormini,1995; Suquet,1997). Recent algorithmic improvements are reported to
markedly reduce this weakness, compare (Doghri/Ouaar,2003).
A problem common to all mean field approaches for elastoplastic composites lies in han-
dling the intraphase variations of the stress and strain fields within the assumption of
phasewise uniformity29) . Improvements in this respect have been obtained by evaluat-
ing the phase averages of the von Mises equivalent stresses from energy considerations
(Qiu/Weng,1992; Hu,1996) or by using statistically based theories (Buryachenko,1996;
Ju/Sun,2001). Such algorithms, however, have been mostly limited to secant plasticity
(deformation theory) approaches, leading to “modified secant models”, see e.g. (Ponte
Castañeda/Suquet,1998). Excellent agreement has been reported between the latter
method and multi-particle unit cell models in a materials characterization context (Se-
gurado et al.,2002).
In the following a short discussion is given of the Incremental Mori–Tanaka (IMT) method
developed by Pettermann (1997). Incremental mean field methods can be formulated in
terms of phase averaged strain and stress rate tensors, dhεi(p) and dhσi(p) , which can be
28)
The condition of radial loading paths in stress space at the constituent level is generally
not fulfilled by a considerable margin for elastoplastic inhomogeneous materials, even for overall
loading paths that are are perfectly radial, see e.g. (Pettermann,1997). This effect is due to
changes in the accommodation of the phase stresses and strains in inhomogeneous materials
upon yielding of a constituent.
29)
Note that, in general, the phase average of the equivalent stress in a given phase is greater
than the equivalent stress computed from the phase averages of the stress components, which
leads to a tendency of simple mean field methods to overestimate overall yield and flow stresses.
As an extreme case, evaluating the von Mises equivalent stresses from the phase averaged stress
components in such approaches leads to predictions that materials with spherical reinforcements
will not yield under overall hydrostatic loads and homogeneous temperature loads.
– 25 –
expressed in analogy to eqn.(16) as
(p) (p)
dhεi(p) = Āt dhεi + āt dT
(p) (p)
dhσi(p) = B̄t dhσi + b̄t dT , (41)
where dhεi stands for the far field mechanical strain rate tensor, dhσi for the far field
(p) (p)
(applied) stress rate tensor, and dT for a spatially uniform temperature rate. Āt , āt ,
(p) (p)
B̄t , and b̄t are instantaneous strain and stress concentration tensors, respectively. Us-
ing the assumption that the inclusions show elastic and the matrix elastoplastic material
behavior30) , the overall instantaneous (tangent) stiffness tensor can be written in terms
of the phase properties and the instantaneous concentration tensors as
(m) (m)
E∗t = E(i) + (1 − ξ)[Et − E(i) ]Āt
(m) (m)
= [C(i) + (1 − ξ)[Ct − C(i) ]B̄t ]−1 , (42)
in analogy to eqn.(18). Note that eqn.(42) is formulated such that dhσi = E ∗t dhεi, i.e.
E∗t is a “continuum” tangent operator, compare (Doghri/Ouaar,2003).
Using the Mori–Tanaka formalism of Benveniste (1987), compare eqn.(30), the instanta-
neous matrix concentration tensors take the form
(m) (m) (m)
Āt = {(1 − ξ)I + ξ[I + St Ct (E(i) − Et )]−1 }−1
(m) (m) (m) (m)
B̄t = {(1 − ξ)I + ξE(i) [I + St Ct (E(i) − Et )]−1 Ct }−1 . (43)
The instantaneous Eshelby tensor St depends on the current state of the matrix material
and in general has to be evaluated numerically due to the anisotropic structure of the
(m)
instantaneous tangent stiffness tensor Et . Expressions for the instantaneous coefficients
of thermal expansion and for the instantaneous thermal concentration tensors can be
derived in analogy to the corresponding thermoelastic relations, e.g. eqns.(19) and (21).
Due to their combination of usually reasonable accuracy, flexibility in terms of inclusion
geometries, and relatively low computational requirements, incremental Mori–Tanaka
methods can be useful as micromechanically based constitutive laws within Finite El-
ement codes for analyzing components and structures made of elastoplastic composite
materials such as metal matrix composites (MMCs). In view of their limitations as dis-
cussed above, however, they should not be used at overall plastic strains exceeding a few
percent.
Formulations of eqns.(41) to (43) that are directly suitable for implementation as mi-
cromechanically based constitutive models at the integration point level within Finite
Element codes can be obtained by replacing rates such as dhσi(p) with finite increments
such as ∆hσi(p) . It is worth noting that in the resulting incremental Mori–Tanaka meth-
ods no assumptions on the overall yield locus and the overall flow potential are made,
30)
Analogous expressions can be derived for elastoplastic inclusions in an elastic matrix or,
in general, for composites containing any required number of elastoplastic phases.
– 26 –
the effective material behavior being entirely determined by the incremental mean field
equations and the constitutive behavior of the phases. However, mapping of the stresses
onto the yield surface cannot be handled at the level of the homogenized material and
the radial return mapping algorithm has to be applied to the matrix at the microscale in-
stead. Accordingly, the constitutive equations describing the overall behavior cannot be
integrated directly (as is the case for homogeneous elastoplastic materials), and iterative
algorithms are required. For example, Pettermann (1997) used an implicit Euler scheme
in an implementation of an incremental Mori–Tanaka method as a user supplied material
routine (UMAT) for the commercial Finite Element code ABAQUS (Hibbit, Karlsson &
Sorensen, Inc., Pawtucket, RI, 1997). Algorithms of this type can also handle thermal
expansion effects31) and temperature dependent material parameters.
Incremental Mori–Tanaka-type approaches have also been employed to obtain estimates
for the nonlinear response of inhomogeneous materials due to microscopic damage or to
combinations of damage and plasticity, see e.g. (Tohgo/Chou,1996; Guo et al.,1997).
Self-consistent schemes can also be combined with secant or incremental approaches
to describe elastoplastic inhomogeneous materials, see e.g. (Hill,1965; Hutchinson,1970;
Berveiller/Zaoui,1979; Bornert,1996), that may also be subject to microscopic damage
(Estevez et al.,1999; González/LLorca,2000). Such procedures (as well as the weaknesses
and strengths of the approaches) are closely related to those discussed above for the
Mori–Tanaka method32) . Analogous statements hold for Hashin–Shtrikman estimates
(and bounds, compare section 3.3) in general, which have e.g. been developed to account
for geometrical nonlinearities due to the evolution of of pore shapes and orientations in
porous materials, compare (Kailasam et al.,2000).
As an alternative to directly extending mean field theories into secant or incremen-
tal plasticity, they can also be combined with Dvorak’s Transformation Field Analysis
(TFA; Dvorak,1992) for obtaining descriptions of the overall behavior of inhomogeneous
materials in the elastoplastic range, compare e.g. (Plankensteiner,2000). The TFA es-
sentially evaluates elastoplastic stress and strain distribution on the basis of the elastic
accommodation of the stresses and strains. Such approaches are in principle suitable
for use as micromechanically based constitutive models in FE codes, are very attractive
in terms of computational requirements, but tend to strongly overestimate the overall
strain hardening of elastoplastic inhomogeneous materials because they essentially use
elastic accommodation of microstresses and strains throughout the loading history.
Finally, it is worth mentioning that even at dilute volume fractions the “Eshelby prop-
31)
It should be noted that elastoplastic inhomogeneous materials such as metal matrix com-
posites typically show a hysteretic thermal expansion response, i.e. the “coefficients of thermal
expansion” are not material properties in the strict sense. This dependence of the thermal
expansion behavior on the instantaneous stiffness requires special treatment within the IMT
framework, compare (Pettermann,1997).
32)
In analogy to the elastic case, Mori–Tanaka based methods are more suitable for describ-
ing elastoplastic matrix–inclusion composites, whereas two-phase self-consistent schemes are
more suited for modeling inhomogeneous materials with topologically indistinguishable phases,
e.g. polycrystals or interwoven materials.
– 27 –
erty” (i.e. uniformity) of the stress and strain fields within the inclusions is lost. Semi-
analytical deformation plasticity solutions for the case of single spheroidal inclusions
embedded in an elastoplastic matrix were given in (Lee/Mear,1999).
The overall symmetry of composites reinforced by nonaligned short fibers in many cases
is isotropic (random fiber orientations) or transversely isotropic (planar random fiber
orientations, fiber arrangements with axisymmetric orientation distributions). However,
processing conditions can give rise to a wide range of fiber orientation distributions and,
consequently, lower overall symmetries, compare e.g. (Allen/Lee,1990).
The statistics of the microgeometries of nonaligned matrix–inclusion composites can be
described on the basis of orientation distribution functions (ODFs) and aspect ratio or
length distribution funcions (LDFs) of the reinforcements. Analytical models of such ma-
terials typically make use of configurational averaging procedures, which may encompass
aspect ratio averaging in addition to orientational averaging. Orientational averaging
of tensor valued variables can be done by direct (numerical) integration, see e.g. (Pet-
termann et al.,1987), or on the basis of expansions of the ODF in terms of generalized
spherical harmonics (Viglin expansions), compare e.g. (Advani/Tucker,1987). The latter
approach can be formulated in a number of ways, e.g. via texture coefficients or texture
matrices, compare (Siegmund et al.,2004). For a discussion of a number of issues relevant
to configurational averaging see Eduljee/McCullough (1993).
Mori–Tanaka methods have been adapted to describing the elastic behavior of rather
general microstructures consisting of nonaligned inclusions embedded in a matrix. A
typical starting point for such “extended Mori–Tanaka” algorithms are dilute stress con-
centration tensors for reinforcement that have some given orientation with respect to the
(i),Θ
global coordinate system, B̄dil , which can be expressed as
(i),Θ −1 Θ −1
B̄dil = TΘ (m)
(I − S)(C(i) − C(m) )
σ I+E Tσ . (44)
on the basis of eqn.(26). Here TΘ σ stands for the stress transformation tensor from the
local (reinforcement based) to the global coordinate system. A phase averaged dilute
fiber stress concentration tensor can then be obtained by orientational averaging as
Z
(i)
D E
(i),Θ (i),Θ
Bdil = ρ (Θ)B̄dil dΘ = ρ B̄dil (45)
Θ
where Θ stands for the orientation angle and the orientation distribution function ρ(Θ)
is assumed to be normalized to give hρi = 1. The core statement of the Mori–Tanaka
approach according to Benveniste (1987), eq.(27), can then be written in the form
– 28 –
The phase averaged Mori–Tanaka concentration tensors can then be expressed as
(m) (i)
BM = [(1 − ξ)I + ξBdil ]−1
(i) (i) (i)
BM = Bdil [(1 − ξ)I + ξBdil ]−1 (47)
– 29 –
with randomly oriented phases of the matrix–inclusion and interwoven types. Another
mean field approach for such materials, the Kuster–Toksőz model, essentially is a dilute
description applicable to matrix–inclusion topologies and tends to give unphysical results
at high inclusion volume fractions. In addition, the Hashin–Shtrikman approach of Ponte
Castañeda and Willis (1995) can be adapted to randomly oriented reinforcements 35) . For
recent discussions on the relationships between some of the above approaches see e.g.
(Berryman/Berge,1996; Hu/Weng,2000). Due to the overall isotropic behavior of com-
posites reinforced by randomly oriented fibers or platelets, their elastic behavior must
comply with the “standard” Hashin–Shtrikman bounds (Hashin/Shtrikman,1963), see
also table 3 in section 5.4.
In addition to providing estimates for the overall thermoelastic behavior of compos-
ites containing nonaligned reinforcements, Mori–Tanaka theories can also be used for
describing the dependence of the microscopic stress and strain tensors in individual re-
inforcements in such materials on the latter’s orientation. For this purpose eqn.(27) is
rewritten as
(i),Θ (i),Θ (m)
B̄M = B̄dil BM , (48)
(i),Θ (m)
where B̄M and BM are defined by eqns. (44) and (47). Results obtained with the
above relation are in good agreement with numerical predictions for moderate volume
fractions and elastic contrasts, compare fig. 5.
Besides mean field and unit cell models (compare section 5.4) a number of other meth-
ods have been proposed for studying composites with nonaligned reinforcements. Most
of them are based on the assumption that the contribution of a given fiber to the overall
stiffness and strength depends solely on its orientation with respect to the applied load
and on its length, interactions between neighboring fibers being neglected. The paper
physics approach (Cox,1952) and the Fukuda–Kawata theory (Fukuda/Kawata,1974)
are based on summing up stiffness contributions of fibers crossing an arbitrary nor-
mal section on the basis of fiber orientation and length distribution functions, com-
pare also (Jayaraman/Kortschot,1996). Such theories use shear lag models (Cox,1952;
Fukuda/Chou,1982) or improvements thereof (Fukuda/Kawata,1974) to describe the be-
havior of a single fiber embedded in matrix, the results being typically given as modified
rules of mixtures with fiber direction and fiber length corrections. Laminate analogy ap-
proaches, see e.g. (Fu/Lauke,1998), approximate nonaligned reinforcement arrangements
by a stack of layers each of which handles one fiber orientation and, where appropriate,
one fiber length. The latter methods, in turn, are closely related to models that are
based on orientational averaging over the elastic tensors corresponding to different fiber
orientations, see e.g. (Siegmund et al.,2004).
35)
Mori–Tanaka methods, which intrinsically describe arrangements of aligned ellipsoids
(compare section 2.3) when used with the Wu tensor generally provide symmetric overall elastic
tensors for materials with randomly oriented reinforcements.
– 30 –
4
σ1 (MPa)
2
1
σ1 - MTM
0 σax - MTM
σ1 - UC
-1
0 10 20 30 40 50 60 70 80 90
Θ(°)
FIG.5: Mori–Tanaka and unit cell predictions for the angular dependence of the max-
imum principal stresses (solid line) in the randomly oriented fibers (aspect ratio 5) of
an SiC/Al MMC subjected to uniaxial tension. In addition, the variation of the axial
stresses is given (dashed line; note that the axial stresses in fibers oriented at angles to the
loading direction in excess of Θ≈65◦ .) Numerical results are presented in terms of mean
values (solid circles) and standard deviations within individual fibers; from (Duschlbauer
et al.,2003).
There are strong conceptual similarities between inclusion problems describing the elas-
tostatic behavior of and diffusive transport processes in inhomogeneous materials, see
e.g. (Hashin,1983). Specifically, in diffusion-type problems (such as heat conduction,
electrical conduction and diffusion of moisture)36) the effects of dilute ellipsoidal inho-
mogeneous inclusions embedded in a matrix can be described in analogy to eqns.(23) to
(26) via “diffusion Eshelby tensors”37) , the Eshelby property taking the form of constant
fluxes (such as temperature gradients) within the inclusion. Similar descriptions can also
be devised for a number of coupled problems, such as the electromechanical behavior of
inhomogeneous materials with at least one piezoelectric constituent.
For nondilute materials direct analoga of the Mori–Tanaka and self-consistent ap-
proaches introduced in chapters 2.3 and 2.7 can be derived, see e.g. (Hatta/Taya,1986;
36)
Note that the differential equations describing antiplane shear in elastic solids, d’Arcy
creep flow in porous media, and equilibrium properties such as overall dielectric constants and
magnetic permeabilities are also of the Poisson type and thus mathematically equivalent to
diffusive transport problems.
37)
In contrast to the “mechanical Eshelby tensor”, which (like the elasticity tensor) is of rank
4, the “diffusion Eshelby tensor” is of rank 2 (like the conductivity tensor), and, for spheroidal
inclusions, it depends only on their aspect ratios. Explicit expressions for the elements of
“diffusion Eshelby tensors” are given e.g. in (Hatta/Taya,1986).
– 31 –
Miloh/Benveniste,1988; Dunn/Taya,1993; Chen,1997). Similar analogies exist for varia-
tional bounding approaches.
Extensive discussions of diffusion-type problems in inhomogeneous media can be found
e.g. in (Markov/Preziosi,2000; Torquato,2001; Milton,2002; Duschlbauer,2003).
– 32 –
3. VARIATIONAL BOUNDING METHODS
Whereas mean field methods, unit cell approaches and embedding strategies can typically
be used for both homogenization and localization tasks, bounding methods are restricted
to homogenization, i.e. only the overall response of inhomogeneous materials is bounded.
Discussions in this chapter again are restricted to materials consisting of two perfectly
bonded constituents.
Rigorous bounds for the overall elastic properties of inhomogeneous materials are typi-
cally obtained from appropriate variational (minimum energy) principles. In the follow-
ing, only outlines of bounding methods are given; formal treatments can be found e.g. in
(Nemat–Nasser/Hori,1993; Bornert,1996; Ponte Castañeda/Suquet,1998; Markov,2000;
Bornert et al.,2001; Milton,2002).
Hill Bounds
Classical expressions for the minimum potential energy and the minimum complementary
energy in combination with uniform stress and strain trial functions lead to the simplest
variational bounding expressions, the upper bounds of Voigt (1889) and the lower bounds
of Reuss (1929). In their tensorial form (Hill,1952) they are known as the Hill bounds
and can be written as
hX i−1 X
(p) (p)
V C ≤ E∗ ≤ V (p) E(p) . (49)
(p) (p)
These bounds, while universal and very simple, do not contain any information on the
microgeometry of an inhomogeneous material beyond the phase volume fractions, and
are typically too slack to be of much practical use38) , but in contrast to the Hashin–
Shtrikman and higher order bounds they also hold for control volumes that are too small
to be proper RVEs.
Hashin–Shtrikman-Type Bounds
Much tighter bounds can be obtained from a variational formulation due to Hashin and
Shtrikman (1962). It is formulated in terms of a homogeneous reference material and
of polarization fields τ (x) that describe the difference between the microscopic stress
fields in the inhomogeneous material and a homogeneous reference material, compare
eqn.(36). In combination with the macrohomogeneity condition (compare section 2.1)
38)
The bounds on the Young’s and shear moduli obtained from eqn.(49) are equivalent to
Voigt and Reuss expressions in terms of the corresponding phase moduli only if the constituents
have equal Poisson’s ratios. Due to the homogeneous stress and strain assumptions used for
obtaining the Hill bounds, the corresponding phase strain and stress concentration tensors are
(p) (p)
ĀV =I and B̄R =I, respectively.
– 33 –
polarization fields allow the complementary energy of the composite to be formulated
in such a way that the “fast” and “slow” contributions are separated, giving rise to the
Hashin–Shtrikman variational principle.
The highly complex position dependent polarizations τ (x) are approximated by phase-
wise constant polarizations τ (p) , compare eqn.(37), which can be optimized with respect to
the Hashin–Shtrikman variational principle. The use of the softer constituent as reference
material then leads to lower bounds and that of the stiffer constituent to upper bounds on
the overall elastic behavior of two-phase materials. The Hashin–Shtrikman bounds proper
(Hashin/Shtrikman,1963) apply to inhomogeneous materials with (statistically) isotropic
overall symmetry; they provide bounds on the effective bulk modulus K ∗ and the effective
shear moduli G∗ . For obtaining bounds on the effective Young’s modulus E ∗ from the
above results see (Hashin,1983), and for bounding expressions on the Poisson numbers ν ∗
see (Zimmerman,1992)39) . These bounds apply to any combination of phase properties
with well defined softer and stiffer constituents, i.e. (K (i) − K (m) )(G(i) − G(m) ) < 0. The
Hashin–Shtrikman formalism also allows to recover the Voigt and Reuss bounds when
the reference material is chosen to be much stiffer or much softer than either of the
constituents.
Another set of Hashin–Shtrikman bounds pertains to (statistically) transverse isotropic
composites reinforced by aligned continuous fibers40) . For a discussion of evaluating these
bounds for composites reinforced by transversely isotropic fibers see (Hashin,1983).
Further developments led to the Walpole (1966) and Willis (1977) bounds, which pertain
to thermoelastically anisotropic overall behavior due to anisotropic constituents, statis-
tically anisotropic phase arrangements, and/or nonspherical reinforcements, a typical
example being the aligned ellipsoidal microgeometry sketched in fig.3. These bounds
include the Hashin–Shtrikman bounds discussed above as special cases and they can
also handle situations where a “softest” and “stiffest” constituent cannot be identified
unequivocally so that a fictive reference material must be used.
Hashin–Shtrikman-type bounds are not restricted to materials with matrix–inclusion
topologies, but hold for any phase arrangement of the appropriate symmetry and phase
volume fractions. In addition, it is worth noting that Hashin–Shtrikman-type bounds
are sharp, i.e. they are the tightest bounds that can be given for the type of geometrical
information used (volume fraction and overall symmetry, i.e. two point correlations).
From the practical point of view it is of interest that for two-phase materials Mori–
Tanaka estimates correspond to one of the Walpole bounds, compare e.g. (Weng,1990),
while the other bound can be obtained after a “color inversion” (i.e. exchanging the
39)
Whereas engineers tend to describe elastic material behavior in terms of Young’s moduli
and Poisson’s ratios (which can be measured experimentally in a relatively straightforward way),
many homogenization expressions are best formulated in terms of bulk and shear moduli, which
directly describe the physically relevant hydrostatic and deviatoric responses of materials.
40)
These bounds apply to transversely isotropic two-phase materials in which the phases
are continuous in the “axial” direction.
– 34 –
roles of inclusions and matrix)41) . Accordingly, the bounds can be evaluated for fairly
general aligned phase geometries by simple matrix algebra such as eqn.(31). With the
exception of random fiber composites, which follow the “standard” Hashin–Shtrikman
bounds for statistically isotropic materials, however, no Hashin–Shtrikman-type bounds
are available for matrix–inclusion materials with nonaligned reinforcements. In the case
of porous materials (where the pores may be viewed as an ideally compliant phase) only
upper Hashin–Shtrikman bounds can be obtained, the lower bounds for the elastic moduli
being zero.
Hashin–Shtrikman-type variational formulations can also be employed to generate
bounds for more general phase arrangements. Evaluating the polarizations for “com-
posite regions” consisting of inclusions embedded in matrix gives rise to Hervé–Stolz–
Zaoui bounds (Hervé et al.,1991). When complex phase patterns are to be considered
(Bornert,1996; Bornert et al.,1996) numerical methods must be employed for evaluating
the polarization fields.
When more complex trial functions are used in variational bounding, their optimization
requires statistical information on the phase arrangement in the form of n-point corre-
lation functions42) . This way improved bounds can be generated that are significantly
tighter than Hashin–Shtrikman-type expressions (which in this context can be classified
as two-point bounds).
Three-point bounds for statistically isotropic two-phase materials can be formulated in
such a way that the information on the phase arrangement statistics is contained in two
41)
Put more precisely, the lower bounds for two-phase materials are obtained from Mori–
Tanaka methods when the more compliant material is used as matrix phase, and the upper
bounds when it is used as inclusion phase. For matrix–inclusion materials with more than two
constituents, Mori–Tanaka expressions are bounds if the matrix is the stiffest or the softest
phase, and otherwise fall between the appropriate bounds (Weng,1990). Note, however, that
extended Mori–Tanaka methods for nonaligned composites as discussed in section 2.6 in general
do not coincide with bounds.
42)
For discussions on statistical descriptions of phase arrangements see e.g. (Torquato,1991;
Ghosh et al.,1997; Pyrz/Bochenek,1998; Torquato,1998; Torquato,2001).
– 35 –
three-point microstructural parameters, η(ξ) and ζ(ξ). The evaluation of these param-
eters for a given microgeometry is a considerable task. However, analytical expressions
or tabulated data in terms of the inclusion volume fraction ξ are available for a number
of generic microgeometries of practical importance, among them statistically homoge-
neous isotropic materials containing identical, bidisperse and polydisperse impenetra-
ble (“hard”) spheres (describing matrix–inclusion composites) as well as monodisperse
interpenetrating spheres (“boolean models” that can describe many interwoven phase
arrangements), and statistically homogeneous transversely isotropic materials reinforced
by impenetrable or interpenetrating aligned cylinders.
References to some three-point bounds and to a number of expressions for η and ζ
applicable to some two-phase composites can be found in section 3.4, where results from
mean field and bounding approaches are compared. For a review of higher order bounds
for elastic (as well as other) properties of inhomogeneous materials see (Torquato,1991;
Torquato,2001).
Analoga to the Hill bounds for nonlinear inhomogeneous materials were introduced by
Bishop and Hill (1951). For polycrystals the nonlinear equivalents to Voigt and Reuss
expressions are usually referred to as Taylor and Sachs bounds, respectively.
In analogy to mean field estimates for elastoplastic material behavior, nonlinear bounds
are typically obtained by evaluating a sequence of linear bounds. Talbot and Willis
(1985) extended the Hashin–Shtrikman variational principles to obtain one-sided bounds
(i.e. upper or lower bounds, depending on the material combination) on the nonlinear
mechanical behavior of inhomogeneous materials.
An important development was the derivation by Ponte Castañeda (1992) of a variational
principle that allows upper bounds on the effective nonlinear behavior of inhomogeneous
materials to be generated on the basis of upper bounds for the elastic response43) by
using a series of inhomogeneous reference materials, the properties of which have to be
obtained by optimization procedures for each strain level. Essentially, the variational
principle guarantees the best choice for the comparison material at a given load. The
Ponte Castañeda bounds are closely related to mean field approaches using improved
secant plasticity methods (compare the remarks in section 2.4). For higher order bounds
on the nonlinear response of inhomogeneous materials see also (Talbot/Willis,1998).
The study of bounds as well as the development of improved estimates (e.g. of the self-
consistent type) for the overall mechanical behavior of nonlinear inhomogeneous materials
have been active fields of research during the last decade, see the recent reviews by
Suquet(1997), Ponte Castañeda and Suquet (1998), and Willis (2000).
43)
The Ponte Castañeda bounds are rigorous for nonlinear elastic inhomogeneous materials
and, on the basis of deformation theory, are very good approximations for materials with at
least one elastoplastic constituent. Applying the Ponte Castañeda variational procedure to
elastic lower bounds does not necessarily lead to a lower bound for the inelastic behavior.
– 36 –
3.4 Some Comparisons of Mean Field and Bounding Predictions
In order to give some idea of the type of predictions obtained by different mean field
(and related) approaches and by bounding methods for the thermomechanical responses
of inhomogeneous thermoelastic materials, results for the overall elastic moduli and coef-
ficients of thermal expansion are presented as functions of the inclusion volume fraction
ξ in this section. All comparisons are based on E-glass particles or fibers embedded in
an epoxy matrix. The elastic contrast of these constituents is approximately 21 and the
thermal expansion contrast approximately 0.14, see table 1.
E ν α
−6]
[GPa] [] [1/K ×10
matrix 3.5 0.35 36.0
reinforcements 74.0 0.2 4.9
TABLE 1: Constituent material parameters for epoxy matrix and E-glass reinforce-
ments used in generating figs. 6 to 12.
Figures 6 and 7 show predictions for the overall Young’s and shear moduli of a particle
reinforced glass/epoxy composite based on the above constituent material parameters.
The Hill bounds can be seen to be very slack. The Mori–Tanaka results (MTM) —
as mentioned before — coincide with the lower Hashin–Shtrikman bounds (H/S LB),
whereas the classical self-consistent scheme (CSCS) shows a typical behavior in that it
is close to one Hashin–Shtrikman bound at low volume fractions, approaches the other
at high volume fractions, and displays a transition behavior in the form of a sigmoid
curve inbetween. The three-point bounds (3PLB and 3PUB) given in figs. 6 and 7 apply
to monodisperse impenetrable spherical inclusions, are based on formalisms developed
by Beran/Molyneux (1966), Milton (1981) and Phan–Thien/Milton (1983), and use ex-
pressions for the statistical parameters η and ζ reported in (Miller/Torquato,1990) and
(Torquato et al.,1987), respectively. Like Torquato’s second order estimates (3PE) these
expressions are available for inclusion volume fractions up to ξ=0.6. Interestingly, for
the case considered here, the results from the generalized self-consistent scheme (GSCS),
which as usual predicts a slightly stiffer behavior than the Mori–Tanaka method (tak-
ing the “standard” case of stiff inclusions in a compliant matrix), are very close to the
lower three-point bounds even though microgeometries corresponding to the GSCS are
not monodisperse in the inclusion sizes.
Predictions for the effective coefficients of thermal expansion of statistically isotropic
glass-epoxy composites are given in fig. 8, the bounds being obtained from a relation
proposed by Levin (1967) and Schapery (1968) that can be evaluated with any bounds
or estimates for the effective bulk modulus. Here the Hashin–Shtrikman bounds and
the three-point bounds are used as basis for bounding the overall thermal expansion
– 37 –
EFFECTIVE YOUNGs MODULUS [GPa]
CSCS
3PE (mono/h)
GSCS
3PUB (mono/h)
3PLB (mono/h)
H/S UB
H/S LB; MTM
40.0
Hill/Voigt UB
Hill/Reuss LB
0.0
CSCS
3PE (mono/h)
GSCS
3PUB (mono/h)
3PLB (mono/h)
H/S UB
H/S LB; MTM
20.0
Hill/Voigt UB
Hill/Reuss LB
0.0
– 38 –
40.0
CSCS
3PE (mono/h)
GSCS
EFFECTIVE CTE [1/K x 10 −6 ]
3PLB (mono/h)
3PUB (mono/h)
H/S LB
H/S UB; MTM
20.0
0.0
44)
Note that upper bounds for the bulk modulus give rise to lower bounds for the CTEs
and vice versa.
45)
In the case of the axial Young’s moduli the different estimates and bounds are essentially
indistinguishable from each other (and from the rules of mixture result) for the material param-
eters and scaling used in fig. 9 and are thus not displayed. It is, however, worth noting that the
effective axial Young’s modulus of a composite reinforced by aligned continuous fibers is always
equal to or larger than the rules of mixture result.
– 39 –
EFFECTIVE YOUNGs MODULUS [GPa]
CSCS
3PE (mono/h)
GSCS
3PUB (mono/h)
3PLB (mono/h)
H/S UB
H/S LB; MTM
40.0
Hill/Voigt UB
Hill/Reuss LB
0.0
CSCS
3PE (mono/h)
GSCS
3PUB (mono/h)
3PLB (mono/h)
20.0
– 40 –
EFFECTIVE SHEAR MODULUS [GPa]
30.0
CSCS
3PUB (mono/h)
GSCS
3PUB (mono/h)
3PLB (mono/h)
20.0
H/S UB
H/S LB; MTM
Hill/Voigt UB
Hill/Reuss LB
10.0
0.0
40.0
CSCS tr.
EFFECTIVE CTE [1/K x 10
CSCS ax.
3PE (mono/h) tr.
3PE (mono/h) ax.
GSCS tr.
GSCS ax.
MTM tr.
MTM ax.
20.0
0.0
– 41 –
Nearly the same axial CTEs (“ax”) are obtained for all models considered. No bounds
for the CTEs are given in fig. 12 because there is no equivalent to the Levin formula for
overall transversely isotropic materials.
In figs. 6 to 12, the classical self-consistent scheme is not in agreement with the three-
point bounds shown, which explicitly correspond to matrix–inclusion topologies. Con-
siderably better agreement with the CSCS is found by using three-point parameters of
the interpenetrating sphere or cylinder type (which can also describe cases where both
phases percolate, but are not as symmetrical with respect to the constituents as the
CSCS). From a practical point of view it is worth noting that despite their sophistication
improved bounds (and Torquato’s second order estimates) may give overly optimistic pre-
dictions for the overall moduli because they describe ideal two-phase materials, whereas
in actual materials it is practically impossible to avoid flaws such as porosity.
Before closing this chapter it should be mentioned that, even though the behaviors shown
in figs. 6 to 12 are rather straightforward, somewhat more “interesting” responses may be
obtained when one or both of the constituents are transversely isotropic with a high stiff-
ness contrast between the axial and transverse directions (note that e.g. carbon fibers
typically exhibit transverse Young’s moduli that are much smaller than the Young’s
moduli of, say, metallic matrices, whereas the opposite holds true for the axial Young’s
moduli). Such effects may be especially marked in the nonlinear range. Also, the angu-
lar dependences of stiffnesses and CTEs in fiber reinforced materials can be quite rich,
compare e.g. (Pettermann et al.,1997).
– 42 –
4. GENERAL REMARKS ON APPROACHES BASED ON DISCRETE
MICROSTRUCTURES
46)
As understood here, statistically reconstructed microstructures are equivalent to real
ones in a statistical sense, i.e. they show equal or very similar phase distribution functions.
Accordingly, each such model geometry may be viewed as being a member of a family of mi-
crogeometries, all of which are (approximate) realizations of of some statistical process. Results
from sets of such simulations are properly evaluated in terms of ensemble averages. Algorithms
for reconstructing matrix–inclusion and more general microgeometries that approximate pre-
defined statistical descriptors are discussed e.g. in (Rintoul/Torquato,1997), (Torquato,1998),
(Roberts/Garboczi,1999), (Zeman/Šejnoha,2001) and (Bochenek/Pyrz,2002). For in-depth dis-
cussions of many of the underlying issues, such as statistical descriptions of inhomogeneous
materials, see (Torquato,2001).
47)
For uniform boundary conditions it can be shown, in fact, that the overall elastic behavior
of matrix–inclusion type composites can be bounded by approximating the actual shape of
particles by inner and outer envelopes of “smooth” (e.g. ellipsoidal) shape. This is known as
the Hill modification theorem, see (Hill,1963; Huet et al.,1991).
– 43 –
from metallographic sections, serial sections, or tomographic data, see e.g. (Fischmeis-
ter/Karlsson,1977; Terada/Kikuchi,1996; Li et al.,1999), the resulting descriptions being
termed “real structure” or “digital image based” (DIB) models.
Unless materials with simple periodic phase arrangements are considered, for both mod-
eling strategies the question immediately arises of how complex (and thus “large”) the
model geometry must be in order to adequately capture the physical behavior of the
material to be studied. For the case of elastic statistically isotropic composites with
matrix–inclusion topology and sphere-like particles, Drugan and Willis (1996) estimated
that for approximating the overall moduli with errors of less than 5% or less than 1%,
respectively, volume elements with sizes of approximately two or five inclusion diame-
ters are sufficient for any volume fraction48) ; compare also (Drugan,2000). Alternatively,
adequate sizes of model geometries for elastic studies may be estimated on the basis of ex-
perimentally obtained correlation lengths of the phase arrangement (Bulsara et al.,1999),
by posing the requirement of having at least 2 statistically independent inhomogeneities
in the cell (Zeman/Šejnoha,2001), or by using windowing approaches (compare chapter
7) to bound the overall response from above and below. In addition, the adequacy of
the size of a volume element may be judged on the basis of deviations from the required
symmetry of the overall response. At present, however, there appears to be no reliable
method for directly assessing the representativeness of the microscopic stress and strain
distributions obtained from a given volume element.
For nonlinear matrix behavior a number of numerical studies (Zohdi,1999; Jiang et
al.,2001; Böhm/Han,2001) have indicated that substantially larger volume elements are
necessary for satisfactorily approximating the required overall symmetry and for obtain-
ing good agreement between the responses of (nominally) statistically equivalent phase
arrangements, especially at elevated overall inelastic strains. The reason for this behavior
lies in the marked inhomogeneity of the strain fields typically found in these regimes, with
contiguous regions of elevated plastic strains or of microscopic damage tending to appear.
The latter are considerably larger than individual inclusions and effectively introduce a
new length scale into the problem. Accordingly, the size of satisfactory multi-inclusion
unit cells depends markedly on the phase material behavior49) .
The majority of published micromechanical analyses of discrete microstructures have
employed standard numerical engineering methods for resolving the microfields, stud-
ies using Finite Difference (FD) algorithms (Adams/Doner,1967), spring lattice models
(see e.g. Ostoja–Starzewski,2002), the Boundary Element Method (BEM; see e.g. Achen-
bach/Zhu,1989), the Finite Element Method (FEM), techniques using Fast Fourier Trans-
forms (FFT; see e.g. Moulinec/Suquet,1994) and Discrete Fourier Transforms (DFT;
48)
Note that a volume element fulfilling a condition of this type is not necessarily a RVE in
the sense discussed in chapter 1.2.
49)
For elastoplastic matrices, weaker strain hardening tends to translate into unit cells
requiring a higher number of inclusions, and very large model geometries are necessary when
the matrix shows softening e.g. due to damage (Zohdi/Wriggers,2001).
– 44 –
Müller,1996)50) , as well as FE-based discrete dislocation models (Cleveringa et al.,1997)
having been reported. A number of more specialized approaches are discussed in connec-
tion with periodic microfield analyses, see section 5.1. Generally speaking, spring lattice
models tend to have advantages in handling pure traction boundary conditions and in
modeling the progress of microcracks due to local (brittle) failure. Boundary elements
tend to be at their best in studying geometrically complex linear elastic problems. For
all the above methods the characteristic length of the discretization (“mesh size”) must,
of course, be considerably smaller than the microscale of a given problem in order to
obtain spatially well resolved results.
At present, the FEM is the most commonly used numerical scheme for evaluating discrete
microgeometries, especially in the nonlinear range, where its flexibility and capability of
supporting a wide range of constitutive models for the constituents and for the interfaces
between them are especially appreciated51) . An additional asset of the FEM in the
context of continuum micromechanics is its ability to handle discontinuities in the stress
and strain components (which typically occur at interfaces between different constituents)
in a natural way via appropriately placed element boundaries.
Applications of the FEM to micromechanical studies tend to fall into four main groups,
compare fig. 13. In most published works the phase arrangements are discretized by
an often high number of “standard” continuum elements, the mesh being designed in
such a way that element boundaries (and, where required, special interface elements)
are positioned at all interfaces between constituents. Such an approach has the ad-
vantage that in principle any microgeometry can be handled and that readily available
commercial FE packages may be used. However, the actual modeling of complex phase
configurations in many cases requires sophisticated and/or specialized preprocessors for
generating the mesh, a task that has been tricky to automatize. The resulting stiffness
matrices may show unfavorable conditioning due to suboptimal element shapes and satis-
factory resolution of the microfields at local “hot spots” (e.g. between closely neighboring
reinforcements) can lead to very large models indeed.
Alternatively, a smaller number of special hybrid elements may be used, which are specif-
ically formulated to model the deformation, stress, and strain fields in an inhomogeneous
region consisting of a single inclusion or void together with the surrounding matrix on
the basis of some appropriate analytical theory, see e.g. (Accorsi,1988). The most highly
developed approach of this type at present is the Voronoi Finite Element Method (Ghosh
et al.,1996), in which the mesh for the hybrid elements is obtained by Voronoi tesselations
based on the positions of the reinforcements. Large planar multi-inclusion arrangements
can be analyzed this way using a limited number of (albeit rather complex) elements,
50)
FFT and DFT based methods are limited to periodic configurations, for which they tend
to be highly efficient, compare e.g. (Michel et al.,1999). Due to their use of a fixed regular grid
they are well suited for real structure models and for studying the evolution of microstructures,
see e.g. (Dreyer et al.,1999).
51)
Constitutive models for constituents used in FEM based micromechanics have included
a wide range of elastoplastic, viscoelastoplastic and continuum damage mechanics type descrip-
tions as well as crystal plasticity models (McHugh et al.,1993; Bruzzi et al.,2001).
– 45 –
a) b) c) d)
and good accuracy as well as significant gains in efficiency have been claimed. Com-
putational strategies of this type are, of course, specifically tailored for inhomogeneous
materials with matrix–inclusion topologies.
Especially when the phase arrangements to be studied are based on digital images of
actual microgeometries, a third approach for discretizing microgeometries of interest. It
consists of using a regular square or hexahedral mesh that has the same resolution as the
digital image, each element being assigned to one of the constituents by operations such
as thresholding of the grey values of the corresponding pixel or voxel, respectively. Such
meshes have the advantage of allowing a straightforward automatic model generation
from appropriate experimental data (metallographic sections, tomographic scans) and
of avoiding ambiguities in smoothing the digital data (which are generally present if a
“standard” FE mesh is employed to discretize experimental data of this type). Obviously
“voxel element” strategies lead to ragged phase boundaries, which may give rise to some
oscillatory behavior of the solutions (Niebur et al.,1999), can lead to very high local stress
maxima (Terada et al.,1997), and may degrade accuracy. They are, however, claimed
not to cause unacceptably large errors even for relatively coarse discretizations, at least
in the linear elastic range (Guldberg et al.,1998). Good spatial resolution, i.e. a high
number of pixels or voxels, of course, is very beneficial in such models.
A fourth approach also uses regular FE meshes, but assigns phase properties at the inte-
gration point level of standard elements (“multiphase elements”, see e.g. Schmauder et
al.,1996; Quilici/Cailletaud,1999). Essentially, this amounts to trading off ragged bound-
– 46 –
aries at element edges for smeared-out (and typically degraded) microfields within those
elements that contain a phase boundary, because stress or strain discontinuities within
elements cannot be adequately handled by standard FE shape functions. With respect
to the element stiffnesses the latter concern can be much reduced by overintegrating ele-
ments containing phase boundaries, which leads to good approximations of integrals in-
volving non-smooth displacements by numerical quadrature, see (Zohdi/Wriggers,2001).
The resulting stress and strain distributions, however, remain smeared-out approxima-
tions in elements that contain phase boundaries, the rate of convergence of the microfields
being accordingly slow. This disadvantage can be eliminated by using “extended” FEM
methods, in which the mesh does not have to conform to the phase boundaries, the latter
being accounted for by appropriately enriching the Finite Element approximations (Moës
et al.,2003).
A fairly recent development for studying microgeometries involving a large number of
inclusions (tens to thousands) by Finite Element methods involves programs specially
geared towards solving micromechanical problems. Such codes may be based on matrix-
free iterative solvers such as Conjugate Gradient (CG) methods, analytical solutions
for the microfields (e.g. constant strain approximations corresponding to the Hill upper
bounds, eqn.(49)), being used as starting solutions to speed convergence52) . For studies
involving such codes see e.g. (Gusev et al.,2000; Zohdi/Wriggers,2001). Fast solvers of
this type open the possibility of solving inverse problems, e.g. for finding optimal particle
shapes for given load cases and damage modes, compare (Zohdi,2003).
52)
Essentially, in such a scheme the initial guess gives a good estimate of long wavelength
contributions to the solution, and the CG iterations take care of short wave contributions.
– 47 –
5. PERIODIC MICROFIELD APPROACHES
Periodic Microfield Approaches (PMAs)53) aim to describe the macroscopic and mi-
croscopic behavior of inhomogeneous materials by studying model materials that have
periodic microstructures.
The first two sections of this chapter cover some basic concepts of unit cell based PMAs.
Following this, applications to continuously reinforced composites, discontinuously rein-
forced composites, and some other inhomogeneous materials are discussed.
Periodic microfield approaches analyze the behavior of infinite (one, two- or three-
dimensional) periodic arrangements of constituents making up a given inhomogeneous
material under the action of far field mechanical loads or uniform temperature fields54) .
The most common approach for studying the stress and strain fields in such periodic
configurations is based on describing the microgeometry by a periodically repeating unit
cell to which the investigations may be limited without loss of information or generality,
at least for static analyses55) . The literature on such methods is fairly extensive and well
developed mathematical theories are available on scale transitions in periodic structures,
compare (Michel et al.,2001).
A wide variety of unit cells have been employed in published PMA studies, ranging from
geometries used to describe simple periodic arrays of inclusions to highly complex phase
arrangements, such as multi-inclusion cells (“supercells”). For simple periodic phase
arrangements it may also be possible to find analytical solutions via series expansions
that make explicit use of the periodicity (see e.g. Sangani/Lu,1987) or via appropriate
potential methods (Wang et al.,2000).
53)
Despite the name “periodic microfield approaches” used here, these methods are not
restricted to micromechanical applications, but can be employed for a wide range of problems
involving two appropriate length scales.
54)
Standard PMAs cannot handle free boundaries or macroscopic gradients in mechanical
loads, temperatures or composition in any direction in which the fields are periodic. If such gra-
dients or free boundaries are to be studied, however, in many cases layer models can be generated
that are nonperiodic in one direction and periodic in the other(s), see e.g. (Wittig/Allen,1994;
Reiter et al.,1997; Weissenbek et al.,1997).
55)
Periodic phase arrangements, especially simply periodic microstructures, typically are
not well suited for dynamic analyses, because they act as filters that exclude all waves with
frequencies that do not fall within certain bands, see e.g. (Suzuki/Yu,1998). In addition, due to
the boundary conditions required for obtaining periodicity, unit cell analyses can only handle
wavelengths that are smaller than or equal to the appropriate cell dimension, and mechanical
waves are “locked in” within the cell instead of being allowed to pass through and continue into
the far field. For the same reason, unit cell based methods typically can resolve only subsets
of the buckling modes and buckling loads in stability analyses (depending on the geometry
of the cell), so that considerable care is required in using them for such purposes, compare
(Vonach,2001).
– 48 –
Even though most PMA studies in the literature have used standard numerical engineer-
ing methods as described in chapter 4, some more specialized approaches for evaluating
microscopic stress and strain fields are worth mentioning. One of them, known as the
Method of Cells (Aboudi,1989; Aboudi,1991), discretizes unit cells that correspond to
square arrangements of square fibers into four subcells, within each of which displace-
ments are approximated by low-order polynomials. Traction and displacement continuity
conditions at the faces of the subcells are imposed in an average sense and analytical
and/or semi-analytical approximations to the deformation fields are obtained in the elas-
tic and inelastic ranges. While using highly idealized microstructures and providing only
limited information on the microscopic stress and strain fields, the resulting models pose
relatively low computational requirements but have limited capabilities for handling axial
shear. The method of cells has been used as a constitutive model for analyses of struc-
tures made of continuously reinforced composites, see e.g. (Arenburg/Reddy,1991). A
development, the Generalized Method of Cells (Aboudi,1996), allows finer discretizations
of unit cells for fiber and particle reinforced composites, reinforcement and matrix being
essentially split into a number of “subregions” of rectangular or hexahedral shape; for
some comparisons with microfields obtained by Finite Element based unit cells see e.g.
(Iyer et al.,2000; Pahr/Arnold,2002). Recently, further improvements of the Method of
Cells have been reported, see (Aboudi et al.,2003).
An alternative approach, Dovrak’s (1992) Transformation Field Analysis, allows the pre-
diction of the nonlinear response of inhomogeneous materials based on either mean field
descriptions (compare the remarks in section 2.4) or on periodic microfield models for
the elastic behavior. Provided sufficiently fine discretizations of the phases are em-
ployed the use of essentially elastic accommodation has been found to be acceptable,
and high computational efficiency is claimed for the method (Dvorak et al.,1994). A
further group of solution strategies reported in the literature (Axelsen and Pyrz,1995;
Fond et al.,2001; Schjødt–Thomsen and Pyrz,2004) use numerically evaluated equivalent
inclusion approaches that account for interacting inhomogeneities and may be based e.g.
on the work of Moschovidis and Mura (1975).
In typical periodic microfield approaches strains and stresses are split into constant
macroscopic contributions (“slow variables”), hεi and hσi, and periodically varying mi-
croscopic fluctuations (“fast variables”), ε0 (z) and σ 0 (z), in analogy to eqn.(1). The
position vectors, however, are denoted as z to indicate that the unit cell “lives” on the
microscale. The volume integrals used to obtain averages, eqns.(3) and (4), must, of
course, be solved over the volume of the unit cell, ΩUC . Formal derivations of the above
relationships for periodically varying microstrains and microstresses show that the work
done by the fluctuating stress and strain contributions vanishes, compare (Michel et
al.,2001).
Evidently in periodic microfield approaches each unit of periodicity (unit cell) con-
tributes the same displacement increment ∆u and the macroscopic displacements vary
(multi)linearly. An idealized depiction of such a situation is presented in fig. 14, which
shows the variations of the strains εs (s) = hεs i + ε0s (s) and of the corresponding displace-
ments us (s) = ūs s + u0s (s) along some section line s in a hypothetical periodic two-phase
– 49 –
ε, u
<εs >
ε’s
u’s −
us s
∆us
A B A B A B A
s
FIG.14: Schematic depiction of the variation of the strains εs (s) and the displacements
us (s) along a section line (coordinate direction s) in some hypothetical inhomogeneous
material consisting of constituents A and B.
material consisting of constituents A and B. Here ūs is defined as ūs = ∆us /U , where U
stands unit of periodicity in direction s and ∆us for the corresponding displacement in-
crement (note that ∆us =hεs iU for linear displacement–strain relations). The periodicity
of the strains and of the displacements is immediately apparent. In addition, symmetry
points of εs (s) and us (s) are indicated by small circles.
Boundary Conditions
The proper use of unit cell based methods requires that the cells together with the bound-
ary conditions (B.C.s) prescribed on them generate valid tilings both for the undeformed
geometry and for all deformed states pertinent to the investigation (i.e. gaps and over-
laps between neighboring unit cells as well as unphysical constraints on the deformations
must not be allowed). In order to achieve this, the boundary conditions for the unit cells
must be specified in such a way that all deformation modes appropriate for the load cases
to be studied can be attained. The three major types of boundary conditions used in
periodic microfield analyses of the mechanical behavior of periodic model materials are
periodic, symmetry, and antisymmetry (or point symmetry) B.C.s56) .
56)
In addition, free surface boundary conditions have been used for layer type models. For
more formal treatments of boundary conditions for unit cells than given here see e.g. (An-
thoine,1995; Michel et al.,1999)
– 50 –
C
B
A F
E
D
H
G
I J K
Generally, for a given periodic phase arrangement unit cells are nonunique, the range of
possible shapes being especially wide when point or mirror symmetries are present in the
microgeometry. As an example, fig. 15 depicts a (planar) periodic hexagonal array of
circular inclusions (fibers with out-of-plane orientation) and some of the unit cells that
can be used to study aspects of the mechanical behavior of this arrangement. There are
considerable difference in the sizes and capabilities of the unit cells shown.
In fig. 16 schematic sketches of the three most important types of boundary conditions,
periodicity, symmetry and antisymmetry B.C.s, are given for two-dimensional unit cells.
The sides and corner points of the cells are annotated; an extension of this nomenclature
to three dimensions is trivial.
The most general boundary conditions for unit cells are periodicity (“toroidal”, “cyclic”)
B.C.s, which can handle any possible deformation state of the cell and, consequently, of
the inhomogeneous material to be modeled. In fig. 15 cells A to F belong to this group.
Because such unit cells tile the computational space by translation, neighboring cells
(and, consequently, opposite faces of a given cell) must fit into each other like parts of
a jigsaw puzzle in both undeformed and deformed states. For the case of quadrilateral
two-dimensional unit cells this can be achieved by pairing opposite faces and linking
the corresponding degrees of freedom from each pair of faces. Using the conventions
for naming the vertices and faces shown in fig. 16 (left), the resulting equations for the
– 51 –
u’
u’ NW
N
NE
u’ NE
u’NW
NE
NW
NW
u’
u’ NE
N
N
NW
NW U
U
E
E
W
u’
P
u’ SE
W
SE
P
z2
SW u’
SW
L
u’
L
SE
SE
SW
S S S
z1
FIG.16: Sketch of periodicity (left), symmetry (center), and antisymmetry (right)
boundary conditions as used with two-dimensional unit cells.
u0N (z̃1 ) = u0S (z̃1 ) + u0NW u0E (z̃2 ) = u0W (z̃2 ) + u0SE u0NE = u0NW + u0SE . (50)
Here z̃1 and z̃2 denote corresponding positions on the N and S faces and on the E and W
faces of the unit cell, respectively. Equation (50) slaves the microscopic displacements of
nodes on faces N and E to those of corresponding nodes on the “master faces” S and W,
with the nodes SE and NW acting as “master nodes”. The undeformed geometry must
fulfill the compatibility conditions
z̃E − z̃W = zSE − zSW = lEW and z̃N − z̃S = zNW − zSW = lNS (51)
for pairs of nodes z̃E and z̃W as well as z̃N and z̃S , respectively, i.e. it must be possible
to translate a master face into its corresponding slave face. In addition, for each master–
slave pair of faces the phase arrangements and the FE discretizations must be compatible,
compare (Swan,1994). Beyond these conditions there is wide latitude in choosing the
shapes and the microgeometries of unit cells that use periodicity boundary conditions 58) .
Periodic B.C.s generally are the least restrictive option for multi-inclusion unit cell models
using phase arrangements obtained by statistically based algorithms (“supercells”) or by
experimental techniques (real structure cells).
57)
In principle, all variables (i.e. for mechanical analyses the displacements, strains and
stresses) must be linked by appropriate periodicity conditions. When a displacement based
FE code is used, however, such conditions can be specified explicitly only for the displacement
components (including, where appropriate, rotation D.O.F.s), and the periodicity of the stresses
and strains is fulfilled approximately. For periodicity and antisymmetry boundary conditions,
the main error is due to the fact that in typical implementations the nodal stresses and strains
are not averaged across cell boundaries, even though they ought to be. Note also that when
stresses are periodic, unit cell boundary tractions are antiperiodic.
58)
Conditions analogous to eqn.(50) can be specified for any appropriately chosen regular
space-filling two-dimensional cell having an even number of sides, i.e. squares, rectangles and
hexagons (compare cell F in fig. 15), and for three-dimensional cells with an even number of faces,
i.e. cubes, hexahedra, rhombic dodecahedra (bcc symmetry) and regular tetrakaidecahedra (fcc
symmetry). More complicated conditions have to be used for unit cells of less regular shape,
compare e.g. (Cruz/Patera,1995; Estrin et al.,1999; Xia et al.,2003), or more complex periodicity.
– 52 –
In practice FE-based unit cell studies using periodicity boundary conditions can be rather
expensive in terms of computing time and memory requirements, because the multi-point
constraints required for implementing eqn.(50) or its equivalents tend to degrade the band
structure of the system matrix, especially in three-dimensional problems. In addition it
is worth noting that — especially for simple phase arrangements — considerable care
may be required to prevent over- and underconstraining due to inappropriate selection
of regions with periodicity boundary conditions, compare (Pettermann/Suresh,2000).
For rectangular and hexahedral unit cells in which the faces of the cell coincide with
symmetry planes of the phase arrangement and for which this property is retained for all
deformed states that are to be studied, periodicity B.C.s simplify to symmetry boundary
conditions. For the arrangement shown in fig. 16 (center) these B.C.s take the form
where u0 and v 0 are the components of the microscopic displacement vector u0 . The
roles of slaves and masters are the same as in eqn.(50). Evidently, eqn.(52) enforces
the condition that pairs of master and slave faces must stay parallel throughout the
deformation history. Note that for symmetry B.C.s there are no requirements with
respect to compatibility of phase arrangements or meshes at paired faces.
Symmetry boundary conditions are fairly easy to use and tend to give rise to small unit
cells for simply periodic phase arrangements, but the load cases that can be handled
are limited to uniform thermal loads, mechanical loads that act in directions normal to
one or more pairs of faces, and combinations of the above59) . In fig. 15, unit cell G uses
boundary conditions of this type60) . A modified version of symmetry boundary conditions
can be used to model the bending of layer models, see e.g. (Weissenbek et al.,1997).
Symmetry boundary conditions are typically very useful for describing relatively simple
microgeometries, but tend to be somewhat restrictive for modeling phase arrangements
generated by statistically based algorithms or obtained from micrographs.
Antisymmetry boundary conditions are even more limited in terms of the microgeometries
that they can handle, because they require the presence of centers of point symmetry
(“pivot points”). In contrast to symmetry boundary conditions, however, unit cells
employing them on all faces are subject to few restrictions to the load cases that can be
handled. Among the unit cells shown in fig. 15, cells H and I use point symmetry B.C.s
59)
Because these load cases include uniaxial loading normal to the faces, loading by ex-
tensional shear (obtained e.g. in the two-dimensional case by applying normal stresses σ a to
the vertical and −σa to the horizontal faces), hydrostatic loading, and (hygro)thermal loading,
symmetry B.C.s in many cases are sufficient for materials characterization.
60)
Whereas for a given periodic phase arrangement there are infinitely many equivalent
unit cells with periodicity B.C.s, the number of “smallest” unit cells with symmetry B.C.s is
always limited. It is also worth noting that for an n-dimensional problem 2 n unit cells with
symmetry B.C.s can be assembled (by mirroring operations) into a unit cell that is suitable for
the application of periodicity B.C.s.
– 53 –
on all faces and can handle any in-plane load61) . Alternatively, antisymmetry B.C.s can
be combined with symmetry B.C.s to obtain very small unit cells that are restricted
to loads acting normal to the symmetry faces, compare unit cells J and K in fig. 15.
The right hand sketch in fig. 16 shows such a “reduced” unit cell, the antisymmetry
boundary conditions being applied on the E-side where a pivot point P is present. For
this configuration the boundary conditions
must be fulfilled, where u0U (s̃) and u0L (−s̃) are the deformation vectors of pairs of points
U and L that are positioned symmetrically on face E (where a local coordinate system
s̃ centered on the pivot P is defined in analogy to fig. 24) and deform symmetrically
with respect to the pivot P. The undeformed geometry of such a face also must be
antisymmetric with respect to the pivot point P, i.e.
and the phase arrangements as well as the discretizations on both halves of the face
must be compatible. Three-dimensional unit cells employing combinations of symmetry
and point symmetry B.C.s can e.g. be used to advantage for studying cubic arrays of
particles, see (Weissenbek et al.,1994).
Within the nomenclature shown in fig. 16 the displacements at the master nodes SE and
NW, u0SE and u0NW , control the overall deformation of the unit cell in 1- and 2-directions
and can be used to evaluate the macroscopic strain of the corresponding periodic model
material, compare eqn.(63). With respect to models using the FEM it is worth noting
that all of the above types of boundary conditions can be handled by any code that
provides linear multipoint constraints between degrees of freedom and thus allows the
linking of three or more D.O.F.s by linear equations.
Obviously, the description of real materials, which in general are not periodic, by peri-
odic model materials entails some geometrical approximations. These take the form of
periodicity constraints on computer generated generic phase arrangements or of appro-
priate modifications in the case of real structure microgeometries62) , compare fig. 1. The
effects of such approximations in the vicinity of the cell surfaces, of course, diminish in
importance with growing size of the model.
61)
Triangular unit cells similar to I in fig. 15 were used e.g. by Teply and Dvorak (1988)
to study the transverse mechanical behavior of hexagonal arrays of fibers. Analogous cells for
perturbed square arrangements of inclusions were introduced in (Marketz/Fischer,1994). Note
that the periodic cells D, E and F also show point symmetry on their faces.
62)
Even though “cutting off” a microgeometry and using symmetry B.C.s for these bound-
aries at first sight does not appear to require modifications, the introduction of a symmetry line
or plane, in fact, amounts to a marked local constraint on or a modification of the geometry. In
the case of sufficiently large multi-inclusion cells it is also possible to obtain useful results when
applying periodicity boundary conditions to nonperiodic geometries, see (Terada et al.,2000).
– 54 –
Equations (50) to (54) refer to the usual case where “standard” (local) material behavior
is used on the macroscale and the scale transition is handled via homogenized stress and
strain fields. Periodic boundary conditions that are conceptually similar to eqn.(50) can
be devised for cases where gradient theories are employed on the macroscale and higher
order stresses as well as strain gradients figure in coupling the length scales (Geers et
al.,2001).
Linking Macroscale and Microscale
The primary practical challenge in using periodic microfield approaches for modeling
inhomogeneous materials lies in choosing and generating suitable unit cells that — in
combination with appropriate boundary conditions — allow a realistic representation
of the actual microgeometries within available computational resources. The unit cells
must then be subjected to appropriate macroscopic stresses, strains and temperature
excursions. Whereas loads of the latter type generally do not pose major difficulties,
applying far field stresses or strains may be difficult (note that the variations of the
stresses along the faces of a unit cell in general are not known a priori, so that it is not
possible to prescribe appropriate boundary tractions via distributed loads63) ).
The most versatile and elegant strategy for linking the macroscale and microscale in unit
cell analyses is based on a mathematical framework known as asymptotic homogenization,
asymptotic expansion homogenization, or homogenization theory, see e.g. (Suquet,1987).
Instead of formulating the unit cell problem solely on the microscale, macroscopic and
microscopic coordinates, Z and z, respectively, are explicitly introduced. They are related
by the expression
zi = Zi / , (55)
where = `/L 1 is a scaling parameter, L and ` being characteristic lengths of the
macro- and microscales64) . The displacement field in the unit cell can then be represented
by an asymptotic expansion of the type
(0) (1) (2)
ui (Z, z, ) = ui (Z) + ui (Z, z) + 2 ui (Z, z) + H.O.T. , (56)
(0) (1)
where the ui are the effective or macroscopic displacements and ui stands for the 1st
order displacement perturbations due to the microstructure, which are assumed to vary
periodically65) . Using the chain rule, i.e.
∂ Z ∂ 1 ∂
f (Z, z = ) → f+ f , (57)
∂x ∂Z ∂z
63)
Prescribing homogeneous strains or tractions at the boundaries essentially corresponds
to the “windowing” approaches discussed in chapter 7.
64)
Equation (55) may be viewed as “stretching” the microscale so it becomes comparable
to the macroscale, f (x) → f (Z, Z/) = f (Z, z).
65)
The nomenclature used in eqns.(55) to (61) follows typical usage in asymptotic homog-
enization. It is more general than but can be directly compared with the one used in eqns.(1)
to (54), where only microscopic coordinates are employed. Zero order terms can generally be
identified with macroscopic quantities and first order terms generally correspond to periodically
varying microscopic quantities.
– 55 –
the strains can be related to the displacements (in the case of small strains) as
1 n ∂ (0) ∂ (0) ∂ (1) ∂ (1) o
εij (Z, z, ) = u + u + u + u
2 ∂Zj i ∂Zi j ∂zj i ∂zi j
n ∂ (1) ∂ (1) ∂ (2) ∂ (2) o
+ u + u + u + u + H.O.T.
2 ∂Zj i ∂Zi j ∂zj i ∂zi j
(1) (2)
= εij (Z, z) + εij (Z, z) + H.O.T. , (58)
(0) (0)
where terms of the type εij = 1 ∂z∂ j ui are deleted due to the underlying assumption
that the variations of slow variables are negligible at the microscale. The stresses are
expanded in analogy to the strains so that the equilibrium condition can be written as
∂ 1 ∂ h (1) i
(2)
+ σij (Z, z) + σij (Z, z) + fi (Z) = 0 , (59)
∂Zj ∂zj
the fi being macroscopic body forces. Working out eqn.(59) and sorting the contributions
by order of gives rise to a hierarchical system of partial differential equations.
∂ (1)
σ =0 (order −1 )
∂zj ij
∂ (1) ∂ (2)
σij + σ + fi = 0 (order 0 ) , (60)
∂Zj ∂zj ij
the first of which gives rise to a boundary value problem at the unit cell level that is
referred to as the “micro equation”. By making a specific ansatz for the strains at unit
cell level and by volume averaging over the second equation in the system (60), the
“macro equation”, the microscopic and macroscopic fields can be linked such that the
homogenized elasticity tensor is obtained as
1 ∂ mn
Z
H
Eijmn = Eijkl δkm δln + χ dΩ , (61)
ΩUC ΩUC ∂zl k
Here ΩUC is the volume of the unit cell, Eijkl (z) is the phase-level elasticity tensor, Iij is
a unit tensor, and the “characteristic function” χmnk (z) describes the deformation modes
66)
of the unit cell . Analogous expressions can be derived for tangent modulus tensors in
nonlinear analyses, compare e.g. (Ghosh et al.,1996).
The above relations can be used as the basis of Finite Element algorithms that solve for
the characteristic function χmn
k . For detailed discussions of asymptotic homogenization
methods within the framework of FEM-based micromechanics see e.g. (Jansson,1992;
Ghosh et al.,1996; Hassani/Hinton,1999; Chung et al.,2001).
66)
Note that even though eqns.(60) and (61) are derived from an explicit two-scale formu-
lation neither of them contains the scale parameter , see the discussion in (Chung et al.,2001).
– 56 –
Asymptotic homogenization allows to directly couple FE analyses on the macro- and
microscales, compare e.g. (Terada et al.,2003), an approach that has been used in a
number of multi-scale analyses (compare section 8) and is sometimes referred to as FE 2
method (Feyel/Chaboche,2000). A treatment of homogenization in the vicinity of macro-
scopic boundaries can be found in (Schrefler et al.,1997), an asymptotic homogenization
procedure for elastic composites that uses standard elements within a commercial FE
package was proposed by Banks–Sills et al.(1997). Asymptotic homogenization schemes
have also been involved for problems involving higher order stresses and strain gradients
(Kouznetsova et al.,2002)67).
When asymptotic homogenization is not used, it is good practice to apply far field stresses
and strains to a given unit cell via concentrated nodal forces or prescribed displacements
at the master nodes and/or pivots, an approach termed the method of macroscopic
degrees of freedom by Michel et al.(1999). Employing the divergence theorem and using
the nomenclature and configuration of fig. 16 (left), the forces to be applied to the master
nodes 2 and 4, P(2) and P(4), of a two-dimensional unit cell with periodicity boundary
conditions can be shown to be given by the surface integrals
Z Z
P(2) = t0a (z) dΓ P(4) = t0a (z) dΓ , (62)
ΓE ΓN
compare (Smit et al.,1988). Here t0a (z) = σa nΓ (z) stands for the surface traction vector
corresponding to the applied stress field68) at some given point z on the cell’s surface
ΓUC , with nΓ (z) being the local surface normal. Equation (62) can be generalized to
require that each master node is loaded by a force corresponding to the surface integral
of the applied traction vectors over the face slaved to it via an equivalent of eqn.(50).
An analogous procedure holds for three-dimensional cases, and symmetry as well as
antisymmetry boundary conditions as described by eqns.(52) and (53) can be handled
similarly. For geometrically nonlinear analyses eqn.(62) must be applied to the current
configuration.
For applying far field strain boundary conditions within the method of macroscopic de-
grees of freedom, displacements are prescribed to the master nodes. These nodal displace-
ments must be obtained from the macroscopic strains via the appropriate displacement–
strain relations. For example, following the notation of eqns.(50) to (53) the displace-
ments to be prescribed to the master nodes 2 and 4 of the unit cells shown in fig. 16 can
be evaluated as
67)
Such methods are useful for problems in with the length scales are not well spread; in
them the “unit cells” do not necessarily remain periodic during the deformation process.
68)
Note that the t0a (z) are not identical with the actual local values of the tractions, t0 (z)
at the cell boundaries, but are equal to them over the cell face in an integral sense.
– 57 –
for the case of linear displacement–strain relations69) , the reference lengths of the rectan-
gular unit cells being defined in eqn.(51). Strain controlled analyses are typically much
easier to carry out by the method of macroscopic degrees of freedom than are stress
controlled ones70) .
In general, the overall stress and strain tensors within a unit cell can be evaluated by
volume averaging71) (e.g. using some numerical integration scheme) or by using the equiv-
alent surface integrals given in eqn.(3), i.e.
1 1
Z Z
0
hσi = σ (z) dΩ = t0 ⊗ z dΓ
ΩUC ΩUC ΩUC ΓUC
1 1
Z Z
0
hεi = ε (z) dΩ = (u ⊗ nΓ + nΓ ⊗ u) dΓ . (64)
ΩUC ΩUC 2ΩUC ΓUC
In the case of rectangular or hexahedral unit cells that are aligned with the coordinate
axes, averaged engineering stress and strain components can, of course, be evaluated by
dividing the applied or reaction forces at the master nodes by the appropriate surface
areas and by dividing the displacements at the master nodes by the appropriate cell
lengths, respectively.
For evaluating phase averaged quantities from unit cell analyses, it is usually possible to
use direct volume integration of eqn.(12) or similar expressions72) . In many FE codes
this can be done by approximate numerical quadrature according to
N
1 1 X
Z
hf i = f (z)dΩ ≈ fl Ωl , (65)
Ω Ω Ω
l=1
where fl and Ωl are the function value and the integration weight (in terms of an inte-
gration point volume), respectively, associated with the l-th integration point within a
69) 0
One of the two displacement components, vSE and u0NW , is an integration constant
describing solid body rotations of the unit cell and, accordingly, can be chosen at will. The
displacements at node NE are fully determined by those of the master nodes, compare eqn.(50).
Analogous relations hold for three-dimensional cases.
70)
Even though the loads acting on the master nodes obtained from eqn.(62) are in principle
equilibrated, in stress controlled analyses solid body rotations may be induced through small
numerical errors. When these solid body rotations are suppressed by deactivating degrees of
freedom beyond those required for enforcing periodicity, the quality of the solutions may be
degraded or transformations of the averaged stresses and strains may be required.
71)
For large strain analysis the consistent procedure is to first average the deformation
gradients and then evaluate the strains from these averages.
72)
It is of some practical interest that volume averaged and phase averaged microfields
obtained from unit cells must fulfill all relations given in section 2.1 and can thus be conveniently
used to check the consistency of a given model by inserting them into eqns.(15). Note, however,
that at finite deformations appropriate stress and strain measures must be used for this purpose
(Nemat–Nasser,1999).
– 58 –
given integration volume Ω that contains N integration points73) , When effective values
or phase averages are to be generated of variables that are nonlinear functions of the
stress or strain components (e.g. equivalent stresses, equivalent strains, stress triaxiali-
ties, stress principals), the variable must be evaluated first (e.g. at the integration points)
and volume averaging of the results performed subsequently, compare also the remarks in
section 2.4. Besides generating overall and phase averages, microscopic variables can also
be evaluated in terms of higher statistical moments and of distribution functions (“stress
spectra”, see e.g. Bornert et al.,1994; Böhm/Rammerstorfer,1995; Böhm/Han,2001).
– 59 –
PH0 PS0
CH1 MS5
RH2 CS7
CH3 CS8
FIG.17: Eight periodic fiber arrangements of fiber volume fraction ξ=0.475 for modeling
continuously fiber reinforced composites (Böhm/Rammerstorfer,1995). Note that the
minimum fiber distances are the same for all configurations except the simple periodic
arrangements PH0 and PS0.
– 60 –
∗ ∗ ∗ ∗
EA ET νA νT α∗A α∗T
−1 −6]
[GPa] [GPa] [] [] [K ×10 [K ×10−6]
−1
77)
Note that the constituents’ material properties underlying table 2 have an elastic contrast
of less than 3. In general the elastic stiffnesses obtained from square-type arrangements may to
violate the Hashin–Shtrikman bounds and usually lie outside the three-point bounds.
– 61 –
FIG.18: Multi-fiber unit cell for a continuously reinforced MMC (ξ=0.453) using 60
quasi-randomly positioned fibers and symmetry B.C.s (Nakamura/Suresh,1993)
CURVE
Date:
6.00E+01
06/09/98
Time:
15:57:39
X,max:
PS/00 4.00E-03
5.00E+01
0.00E+00
APPLIED STRESS [MPa]
PS/45
3.00E+01
Matrix
2.00E+01
1.00E+01
DN/00
PS0/45
PS0/00
PH0/90
PH0/00
0.00E+00
Al99.9 MATRIX
ALTEX FIBER
– 62 –
EPS.EFF.PLASTIC
2.25E-02
1.75E-02
1.25E-02
7.50E-03
2.50E-03
loading, however, tends to be very sensitive to the fiber arrangement. As to simply pe-
riodic models, the behavior of hexagonal arrangements typically is sandwiched between
the stiff (in 0◦ –direction) and the compliant (in 45◦ –direction) responses of periodic
square arrangements in both the elastic and elastoplastic ranges, see fig. 19. Multi-fiber
unit cells that approach statistical transverse isotropy such as fig. 18 tend to show no-
ticeably stronger strain hardening than periodic hexagonal arrangements, compare also
(Moulinec/Suquet,1997). Generally, it may be noted that the macroscopic yield sur-
faces of uniaxially reinforced MMCs are not of the Hill type, but rather follow bimodal
descriptions (Dvorak/Bahei-el-Din,1987).
The distributions of microstresses and microstrains in fibers and matrix typically depend
markedly on the fiber arrangement, especially under thermal and transverse mechanical
loading78) . In the plastic regime, the microscopic distributions of equivalent stresses
78)
Irrespective of the types of load applied, for technologically relevant inclusion volume
fractions Eshelby’s (1957) result of uniform stresses and strains in dilute inclusions holds neither
in the elastic nor the inelastic regimes for nondilute composites.
– 63 –
and equivalent plastic strains79) , of hydrostatic stresses, and stress triaxialities tend
to be markedly inhomogeneous80) , see e.g. fig. 20. This type of behavior typically is
most marked for applied transverse shear loads (and for elastic–ideally plastic matrix
behavior). As a consequence of the inhomogeneity of the microfields, there tend to be
strong constrained plasticity effects in continuously reinforced MMCs and the onset of
damage in the matrix, of fracture of the fibers, and of interfacial decohesion at the fiber–
matrix interfaces show a strong dependence on the fiber arrangement. Marked local
irregularities in the microgeometry, such as clusters of closely approaching fibers and/or
“matrix islands”, as present e.g. in arrangement RH2 or in the multi-fiber unit cell shown
in fig. 18, appear to be especially critical in this respect.
Finally, it should be noted that there are other classes of materials reinforced by continu-
ous fibers that can be studied by unit cell methods, among them composites using woven,
knitted, and braided reinforcements. For such materials the thermomechanical behavior
of different weaves can be described by mesoscopic unit cells that model tows containing
many individual fibers which are embedded in a matrix, see e.g. (Chang/Kikuchi,1993;
Dasgupta et al.,1996). In addition, it is possible to investigate local stress and strain
fields in certain cross-ply laminates via unit cell methods, compare e.g. (Lerch et al.,1991;
Ismar/Schröter,2000).
– 64 –
simple conceptually and, when employing periodicity B.C.s, can in principle be used to
model the full thermomechanical response of aligned short fiber reinforced composites.
FIG.21: Sketch of fiber positions in non-staggered hexagonal (left) and staggered square
(right) periodic arrangements of aligned short fibers.
When limitations are accepted with respect to the load cases that can be studied, con-
siderably smaller unit cells that employ symmetry boundary conditions may be used
for staggered and non-staggered fiber arrangements, see e.g. (Levy/Papazian,1991) and
compare fig. 22. Analogous microgeometries based on periodic hexagonal arrangements
of non-staggered (Järvstråt,1992) or staggered (Tucker/Liang,1999) fibers as well as
ellipsoidal unit cells aimed at describing composite ellipsoid assembly microstructures
(Järvstråt,1993) were also proposed82) .
FIG.22: Three-dimensional unit cells for short fibers employing symmetry boundary
conditions used for modeling non-staggered (top) and staggered (bottom) square ar-
rangements of fibers (Weissenbek,1994).
Staggered and non-staggered unit cells of the above types, however, are rather restrictive
in terms of fiber arrangements and shapes, a difficulty that can be resolved by using much
larger multi-fiber unit cells that contain randomly positioned aligned fibers of different
lengths see e.g. (Ingber/Papathanasiou,1997; Gusev,2001).
82)
For a comparison between unit cell and analytical predictions for the overall elastic
properties for short fiber reinforced composites see e.g. (Tucker/Liang,1999).
– 65 –
For many materials characterization studies, a more economical alternative to the above
three-dimensional unit cells are axisymmetric models describing the axial behavior of
non-staggered or staggered arrays of aligned cylindrical short fibers in an approximate
way. The basic idea behind these descriptions is to replace unit cells for square or
hexagonal arrangements by circular composite cylinders of equivalent cross sectional
area (and volume fraction) as sketched in fig. 23.
FIG.23: Periodic arrays of aligned non-staggered (top) and staggered (bottom) short
fibers and corresponding axisymmetric cells (left: cross sections in transverse plane; right:
sections parallel to fibers).
The resulting axisymmetric cells are not unit cells in the strict sense, because they overlap
and are not space filling83) . In addition, they do not have the same transverse fiber
spacing as the corresponding three-dimensional arrangements and they are restricted to
handling load cases that lead to axisymmetric deformations (i.e. they cannot be used to
study the response to transverse uniaxial or shear loading). They have the advantage,
however, of significantly reduced computational requirements.
Axisymmetric cell models employ symmetry boundary conditions for the top and bottom
faces of the cells, and the B.C.s at the cylinder surface are chosen to maintain the
same cross sectional area along the axial direction for an aggregate of cells. In the case
of non-staggered fibers this can be easily done by specifying symmetry-type boundary
conditions, whereas for staggered arrangements a pair of cells with opposite fiber positions
is considered, for which the total cross sectional area is required to be independent of
83)
Analogous axisymmetric cells may also be used for studying aspects of the behavior of
continuously reinforced composites.
– 66 –
N NE
N NE
~s
NW U NW U
E
P E
W W P
z z
L ~
−s
SW SW L
S SE S SE
r r
UNDEFORMED DEFORMED
FIG.24: Axisymmetric cell for staggered arrangement of short fibers: undeformed and
deformed shape.
the axial coordinate84) . Using the nomenclature of fig. 24 and the notation of eqns.(50)
to (53), nonlinear relations are obtained for the radial displacements, u0 , together with
linear constraints for the axial displacements, v 0 ,
Here U and L are nodes on the E-side of the unit cell that are positioned symmetrically
with respect to the pivot point P. Nonlinear constraints of this type may not be available
or can be cumbersome to use in standard FE codes85) . However, for most applications
the B.C.s for the radial displacements in eqn.(66) can be linearized without major loss
in accuracy, so that antisymmetry-type boundary conditions analogous to eqn.(53) are
obtained for the radial surface, i.e.
For the last decade, axisymmetric cell models have been the workhorses of PMA studies
of short fiber reinforced composites, see e.g. (Povirk et al.,1992; Tvergaard,1994), and
they have been extended to cover composites containing aligned inclusions of different
sizes, shapes and aspect ratios (essentially by coupling two different cells instead of two
equal ones as shown in fig. 24), compare (Böhm et al.,1993; Weissenbek,1994; Tver-
gaard,2004). Typically, descriptions using staggered arrangements allow a wider range
of microgeometries to be covered and give somewhat more realistic descriptions of actual
composites.
84)
As originally proposed by Tvergaard (1990), the nonlinear displacement boundary con-
ditions in eqn.(66) were combined with antisymmetry traction B.C.s for use with a hybrid FE
formulation.
85)
For an example of the use of the nonlinear BCs described in eqn.(66) with a commercial
FE code see e.g. (Ishikawa et al.,2000), where a cell of truncated cone shape is employed to
describe bcc arrangements of particles.
– 67 –
Composites Reinforced by Nonaligned Short Fibers
– 68 –
FIG.25: Unit cell for a short fiber reinforced MMC (ξ=0.15 nominal) containing 15
cylindrical short fibers of aspect ratio 5 in a quasi-random arrangement suitable for
using periodicity B.C.s (Böhm et al.,2002).
of the phase arrangement and for use in the elastoplastic range unit cells containing a
considerably higher number of fibers are certainly desirable.
E∗ ν∗
[GPa] []
SiC particles 450.0 0.17
Al2618-T4 matrix 70.0 0.30
HS/lo 87.6 0.246
HS/hi 106.1 0.305
MTM 89.8 0.285
CSCS 91.2 0.284
KTM 90.3 0.285
MFUC/sph 89.4 0.285
MFUC/cyl 90.0 0.284
– 69 –
Composites Reinforced by Particles
Materials reinforced by statistically uniformly distributed particles show isotropic over-
all behavior, but there exists no simple periodic arrangement of particles that is both
inherently elastically isotropic and space filling86) . In actual materials, which often show
rather irregular particle shapes, anisotropies in the microgeometries and in the overall
response may be introduced by processing effects, e.g. extrusion textures. Accordingly,
unit cell studies of particle reinforced composites that use generic phase arrangements
are subject to nontrivial tradeoffs between keeping computational requirements at man-
ageable levels (which favors simple particle shapes combined with two-dimensional or
simple three-dimensional microgeometries) and obtaining sufficient level of realism for
a given purpose (which often leads to requirements for unit cells containing a consider-
able number of particles of complex shape at random positions in three dimensions). In
many respects periodic microfield analyses of particle reinforced composites are subject
to similar constraints and use analogous approaches as work on short fiber reinforced
composites.
1-part.cell
1-particle 2-particle
cell cell
Reference
2-particle cell Volume
2-particle cell 1-particle cell
FIG.26: Simple cubic, face centered cubic, and body centered cubic arrangements of
spheres with some simple unit cells (Weissenbek et al.,1994)
Most three-dimensional unit cell studies of generic microgeometries for particle rein-
forced composites have been based on simple cubic (sc), face centered cubic (fcc) and
body centered cubic (bcc) arrangements of spherical, cylindrical and cube-shaped inclu-
sions, see e.g. (Hom/McMeeking,1991). By invoking the symmetries of these arrange-
ments and using symmetry as well as antisymmetry boundary conditions, relatively sim-
ple unit cells for materials characterization can be obtained87) , see e.g. (Weissenbek et
al.,1994; Sanders/Gibson,2003) and compare figs. 26 and 27. In addition, work employ-
ing hexagonal or tetrakaidecahedral inclusion configurations has been reported, see e.g.
(Rodin,1993). Of the above arrangements, simple cubic models are the easiest to handle,
86)
Even though pentagonal dodecahedra and icosahedra have the appropriate symmetry
properties (Christensen,1987), they are not space filling.
87)
Again, larger unit cells with periodicity rather than symmetry boundary conditions
are required for unrestricted modeling of the full thermomechanical response of cubic phase
arrangements.
– 70 –
FIG.27: Some unit cells for particle reinforced composites using cubic arrangements
(Weissenbek et al.,1994)
but they show a marked anisotropy. The overall elastic moduli obtained from cubic ar-
rangements of particles do not necessarily fulfill the Hashin–Shtrikman bounds and they
typically lie outside the three-point bounds, compare table 4.
Recently, three-dimensional studies based on more complex phase arrangements have
appeared in the literature. Finite Element methods were used in combination with unit
cells containing up to 64 statistically distributed particles to describe the overall behavior
of elastic particle reinforced composites (Gusev,1997). For studying elastoplastic parti-
cle reinforced MMCs and related materials, hexahedral unit cells containing up to 10
particles in a perturbed cubic configuration (Watt et al.,1966) as well as cube shaped
cells incorporating 15 to 20 spherical inclusions in quasi-random arrangements (Böhm et
al.,1999, Böhm/Han,2001), compare fig. 28, were proposed.
Three-dimensional simulations involving high numbers of particles for investigating com-
posites with statistically uniform phase arrangements have been reported for models of
elastic composites (Michel et al.,1999) and for studying brittle matrix composites that
develop damage (Zohdi/Wriggers,2001). In addition composites with nonuniformly ar-
ranged reinforcements were investigated by unit cells containing 7 clusters of 7 particles
each (Segurado et al.,2003).
In analogy to short fiber reinforced materials, axisymmetric cell models with staggered or
non-staggered inclusions (compare fig. 21) can be used for materials characterization of
particle reinforced composites. Such approaches have been a mainstay of PMA modeling
of these materials, see e.g. (Bao et al.,1991; LLorca,1996). It is interesting to note
that by appropriate choices of the cells’ aspect ratios axisymmetric cell models can be
generated that in many ways correspond to simple, face centered, and body centered
cubic arrangements (Weissenbek,1994).
Due to their relatively low computational requirements, planar unit cell models of parti-
cle reinforced materials are often found in the literature. Generally, plane stress models
(which actually describe thin “reinforced sheets”) show a more compliant and plane strain
models (which correspond to reinforcement by aligned continuous fibers rather than par-
ticles) show a stiffer overall response than three-dimensional descriptions, compare table
– 71 –
FIG.28: Unit cell for a particle reinforced MMC (ξ=0.2 nominal) containing 20
spherical particles in a quasi-random arrangement suitable for using periodicity B.C.s
(Böhm/Han,2001)
4. With respect to the elastoplastic overall behavior, plane stress analyses may be prefer-
able to plane strain analyses, see e.g. (Weissenbek,1994), but no two-dimensional model
reliably gives satisfactory results in terms of the predicted microstress and microstrain
distributions88) , see table 4 and (Böhm/Han,2001). Axisymmetric cell models typically
provide considerably better results than planar ones. Accordingly, extreme care on the
part of the analyst is required in using planar unit cell models for particle reinforced com-
posites and quantitative agreement with experiments cannot be expected. Evidently,
three-dimensional (or axisymmetric) analyses should be used for discontinuously rein-
forced materials wherever possible.
In table 4 bounds, MFA results, and PMA predictions for the overall thermoelastic
moduli of a particle reinforced SiC/Al MMC (elastic contrast ≈ 6) are compared, the
88)
Plane stress configurations tend to predict much higher levels of equivalent plastic strains
in the matrix and much weaker overall hardening than do plane strain and generalized plane
strain models using the same phase geometry. For a given particle volume fraction, results from
three-dimensional arrangements typically lie between those of the above planar models. Evi-
dently, the configurations of regions of concentrated strains that underlie this behavior depend
strongly on the geometrical constraints, compare also (Iung/Grange,1995; Gänser et al.,1998;
Böhm et al.,1999, Shen/Lissenden,2002). For continuously reinforced composites subjected to
transverse loading the equivalent plastic strains tend to concentrate in bands oriented at 45 ◦ to
the loading directions for, whereas the patterns appear to be qualitatively different in particle
reinforced materials.
– 72 –
E∗ E ∗ [100] E ∗ [110] ν∗ α∗
[GPa] [GPa] [GPa] [] [K−1 ×10−6]
SiC particles 429.0 — — 0.17 4.30
Al99.9 matrix 67.2 — — 0.35 23.0
HS/lo 90.8 — — 0.286 16.8
HS/hi 114.6 — — 0.340 18.6
3PBm/lo 91.4 — — 0.323 18.5
3PBm/hi 93.9 — — 0.328 18.6
MTM 90.8 — — 0.329 18.6
GSCS 91.5 — — 0.327 18.6
2OEm 91.8 — — 0.327 18.6
sc — 96.4 90.5 — 18.7
fcc — 89.0 91.7 — 18.6
bcc — 90.0 90.6 — 18.6
axi/sc 95.4 — — 0.318 18.6/18.6
axi/fcc 88.1 — — 0.342 18.1/19.0
axi/bcc 87.9 — — 0.352 18.4/18.8
3/D MPUC 92.4 — — 0.326
2/D PST MPUC 85.5 — — 0.334
2/D PSE MPUC 98.7 — — 0.500
loading directions for the cubic arrangements being identified by Miller indices. It is
interesting that — for the loading directions considered here, which do not span the
full range of possible responses of the configurations — none of the results from the
cubic arrangements falls within the three-point bounds for identical spherical inclusions.
The overall anisotropy of the simple cubic arrangement is marked, whereas the body
centered and face centered arrays deviate much less from overall isotropy. The predictions
of the axisymmetric analyses can be seen to be of comparable quality to those of the
corresponding cubic arrays. The results given for the three-dimensional multi-particle
models are ensemble averages89) over a number of unit cells and loading directions, and
they show very good agreement with the three-point bounds and second order estimates,
89)
When volume elements are used that are smaller than proper RVEs ensemble averaging
over a number of such results can be used to estimate the effective material properties, compare
e.g. (Kanit et al.,2003).
– 73 –
whereas there are marked differences to the plane stress and plane strain models90) .
Predicted overall moduli evaluated from individual multi-particle cells differ by about 2%,
and due to the use of a periodic pseudo-random rather than a statistically homogeneous
particle distribution, the Drugan–Willis estimates for the errors in the overall moduli are
exceeded somewhat.
E∗ ν∗ G∗ K∗ α∗
[GPa] [] [GPa] [GPa] [K−1 ×10−6]
glass particles 73.1 0.18 30.1 38.1 4.90
expoxy matrix 3.16 0.35 1.17 3.51 40.0
HS/lo 6.80 0.024 2.59 6.11 12.1
HS/hi 20.9 0.404 8.53 12.7 23.5
3PB(m)/lo 7.05 0.236 2.69 6.16 20.1
3PB(m)/hi 10.0 0.334 3.95 7.22 23.6
3PB(p)/lo 7.17 0.195 2.74 6.26 17.2
3PB(p)/hi 12.1 0.356 4.79 8.56 23.0
MTM 6.80 0.315 2.59 6.11 23.5
GSCS 7.23 0.303 2.77 6.11 23.5
2OE(m) 7.29 0.304 2.80 6.22 23.2
2OE(p) 7.52 0.304 2.88 6.39 22.6
3/D MPUC(m) 7.95 0.291 3.11 6.34 22.8
3/D MPUC(p) 8.06 0.288 3.11 6.36 22.7
Table 5 lists predictions for the overall behavior of a particle reinforced glass-epoxy
composite (elastic contrast ≈ 23) at much higher volume fraction, ξ=0.389, where the
three-point bounds are no longer particularly sharp. Both the three-point bounds and
Torquato’s second order estimates show some sensitivity to the size distribution of the
spherical reinforcements, monodisperse (m) and polydisperse (p) cases being compared.
Again there is very good agreement between the analytical results and the predictions
of multi-particle unit cells, which used 80 monodispersely or polydispersely sized quasi-
randomly positioned spherical particles. Both cells were generated and evaluated with
the FE-based program PALMYRA. On closer inspection the moduli listed for the unit
cells can be seen to deviate somewhat from the relationships pertaining to isotropic
materials; this due to the fact that averaged moduli obtained from slightly anisotropic
cells are given.
90)
Note that the comparisons given in tables 2 to 5 are strictly applicable only for the
material parameters used there, but do show typical trends.
– 74 –
Like fiber reinforced MMCs, particle reinforced composites typically display highly in-
homogeneous distributions of the microstresses and microstrains in the nonlinear range,
compare fig. 29 (which shows the predicted equivalent accumulated plastic strains inside
a multi-particle model of an MMC). The shapes of the regions of elevated plastic strains,
however, may differ considerably between the two groups of composites. If a constituent
is driven into strain softening behavior,shear bands typically appear. These effects, which
tend to be strongest under overall shear loading, give rise to microscopic “structures”
that can be considerably larger than individual particles, leading to longer ranged inter-
actions between inclusions. These, in turn, are thought to underlie the need for larger
control volumes for studying composites with nonlinear matrix behavior mentioned in
chapter 4.
5.0000E-02
4.0000E-02
3.0000E-02
2.0000E-02
1.0000E-02
FIG.29: Predictions for the equivalent plastic strains in a particle reinforced MMC
(ξ=0.2 nominally) subjected to uniaxial tensile loading obtained by a unit cell with 20
spherical particles in a quasi-random arrangement (Böhm/Han,2001)
If the inclusions are highly irregular in shape or if their volume fraction markedly exceeds
0.5 (as is typically the case for cermets such as WC/Co), generic unit cell models employ-
ing relatively regular particle shapes may not result in very satisfactory microgeometries.
For such materials a typical approach consists of basing PMA models on a “real struc-
ture” obtained (maybe with some minor modifications) from metallographic sections,
compare e.g. (Fischmeister/Karlsson,1977). Due to the nature of the underlying experi-
mental data, models of this type often take the form of planar analyses, the limitations of
– 75 –
which have been discussed above. This difficulty can be overcome e.g. by voxel-based dis-
cretization schemes using digitized serial sections, see (Terada/Kikuchi,1996). Alterna-
tively, more or less irregular particles identified by such procedures may be approximated
by equivalent ellipses or ellipsoids, compare (Li et al.,1999).
In this section some additional applications of periodic microfield approaches are pre-
sented, which were selected mainly in view of discussing modeling concepts. The reader
should be aware that it touches only a very small fraction of the research activities in
which unit cell methods have been applied to the study of the thermomechanical prop-
erties of materials.
91)
HCT and related models of matrix–inclusion topologies are, however, typically restricted
to low inclusion volume fractions and tend to be much more regular than actual materials.
Note that an alternative modeling strategy for layer and cluster structured HSSs is discussed in
chapter 8.
92)
In contrast to the other inhomogeneous materials discussed in these course notes, FGMs
and layer-structured HSSs are not statistically homogeneous.
– 76 –
FIG.30: HCT unit cell models for a layer structured high-speed steel containing
equiaxed and elongated carbides (left; Plankensteiner et al.,1997) and for a function-
ally graded material (right).
– 77 –
tesselations may have to be modified by eliminating small boundary faces and/or edges
that are rarely found in actual grain geometries). In addition, the use of crystal plasticity
formulations for describing the material behavior of the individual grains makes for high
demands on computational resources, especially for three-dimensional analyses, see e.g.
(Quilici/Cailletaud,1999).
Methods of this type have, for example, successfully described the evolution of texture
in metals and are also being employed for studying dislocation structures which play an
important role in the hardening response. For a general discussion of the issues involved
in micromechanical models in crystal plasticity and related fields see e.g the overview by
Dawson (2000).
PMA Models for Porous and Cellular Materials, Cancellous Bone, Wood
Elastoplastic porous materials have been the subject of a considerable number of PMA
studies due to their relevance to ductile damage and failure of metals94) . Generally,
modeling concepts are closely related to those employed for particle reinforced compos-
ites, compare section 2.4, the main difference being that the shapes of the voids may
evolve significantly through the loading history95) . Accordingly, axisymmetric cells, see
e.g. (Koplik/Needleman,1988; Pardoen/Hutchinson,2000; Gărăjeu et al.,2000) and three-
dimensional unit cells based on cubic arrangements of voids, compare e.g. (McMeek-
ing/Hom,1990; Segurado et al.,2002), have been used in the majority of published unit
cell studies of porous materials. Recently, studies based on multi-void cells using geom-
etry data from serial sectioning have been reported (Shan/Gokhale,2001).
In cellular materials, such as foams and cancellous bone, the volume fraction of the solid
phase is low (often amounting to no more than a few percent) and the void phase may
94)
Most constitutive models describing ductile damage in metals are based on microme-
chanical considerations, see e.g. (Rice/Tracey,1969; Gurson,1977; Tvergaard/Needleman,1984;
Gologanu et al.,1997; Kailasam et al.,2000).
95)
Additional complications are introduced by void size effects (Tvergaard,1996) and void
coalescence (Faleskog/Shih,1997). It may be noted that the evolution of the shapes of initially
spheroidal voids under non-hydrostatic loads has also been the subject of intensive studies
by mean field type methods, see e.g. (Kailasam/Ponte Castañeda,1998; Kailasam et al.,2000).
Analyses of this type are typically based on the assumption that initially spherical pores will
stay ellipsoids throughout the deformation history, which axisymmetric cell analyses Gărăjeu
et al. (2000) have shown to be an excellent approximation for axisymmetric tensile load cases.
For compressive loading, however, initially spherical pores may evolve into markedly different
shapes (Segurado et al.,2002).
– 78 –
be topologically connected (open cell foams) or unconnected (closed cell foams). Such
materials often display a very small linear range, and at higher strains gross shape changes
of the cells typically take place. This behavior is especially pronounced for compressive
loading, where bending, elastic buckling, plastic buckling, and brittle failure of cell walls
or struts can play a major role. Periodic microfield methods are well suited to studying
the thermomechanical behavior of cellular materials96) , special provision for handling
large deformations of and contact between cell walls being required in many cases. In
addition, care must be taken that boundaries for which symmetry B.C.s are specified
do not coincide with cell walls that can be expected to bend or buckle, and models in
general must be sufficiently large so that appropriate nontrivial deformation patterns can
develop97) .
The geometrically most simple cellular materials are regular honeycombs, which can be
modeled by planar hexagonal cell models. Despite their geometrical simplicity regular
honeycombs show quite complex deformation patterns, the resolution of which requires
sufficiently large unit cells, compare (Daxner et al.,2002), and are subject to marked lo-
calization effects under compressive “in-plane” loading (Papka/Kyriakides,1994). Some-
what less ordered two-dimensional arrangements have been used for studying the crush-
ing behavior of soft woods, see e.g. (Holmberg et al.,1999), and highly irregular planar
arrangements, compare fig. 31, can be used to study many aspects of the geometry de-
pendence of the mechanical response of cellular materials.
For three-dimensional studies of closed cell foams, various generic microstructures based
on cubic arrangements, see e.g. (Hollister et al.,1991), truncated cubes plus small cubes
(Santosa/Wierzbicki,1998), rhombic dodecahedra, and regular tetrakaidecahedra (Gren-
estedt,1998; Simone/Gibson,1998; Ableidinger,2000), compare fig. 32, have been used. In
models of this type cell walls are typically described as thin shells. Analogous regular and
irregular microgeometries have formed the basis for analytical and numerical beam-type
models for open cell foams (Zhu et al.,1997; Shulmeister et al.,1998; Vajjhala et al.,2000).
A recent development have been unit cell models employing voxel-type discretizations 98)
of actual foams, see e.g. (Maire et al.,2000; Roberts/Garboczi,2001).
A further group of cellular materials amenable to PMA modeling are syntactic foams
(i.e. hollow spheres embedded in a solid matrix) and hollow sphere foams (in which the
spaces between the spheres are “empty”), for which axisymmetric or three-dimensional
96)
The widely used results of Gibson and Ashby (1988), in which thermomechanical moduli
and other overall physical properties of cellular materials are expressed in terms of the relative
density, were derived by analytically studying specific arrangements of beams (for open cell
foams) and plates (for closed cell foams). The majority of these geometries are proper unit cells
that describe periodic phase arrangements.
97)
For perfectly regular structures such as hexagonal honeycombs the minimum size of a
unit cell for capturing bifurcation effects can be obtained from extended homogenization theory
(Saiki et al.,2002).
98)
For voxel-based models of open-cell and closed-cell foams Roberts and Garbozci (2001)
estimate the “systematic discretization error” to be of the order of 10% in terms of the overall
moduli.
– 79 –
Y
Z X
FIG.31: Planar periodic unit cell for studying irregular cellular materials (Daxner et
al.,2002).
cell models describing cubic arrangements of spheres can be used, see e.g. (Rammerstor-
fer/Böhm,2000; Sanders/Gibson,2003).
The effects of details of the microgeometries of cellular materials (e.g. thickness distri-
butions and geometrical imperfections or flaws of cell walls or struts), which can con-
siderably influence the overall behavior (a typical example being the very small elastic
– 80 –
ranges of metallic foams), have been an active field of research for the last years, see
e.g. (Grenestedt,1998; Daxner,2002). Unit cell analyses of cellular materials with re-
alistic microgeometries tend to be rather complex and numerically demanding due to
these materials’ tendency to deform by local mechanisms and instabilities99) , but analyt-
ical solutions have been reported for some simply periodic phase arrangements, see e.g.
(Warren/Kraynik,1991).
A special type of cellular material, cancellous (or spongy) bone, has attracted consid-
erable research interest for the last twenty years. Cancellous bone shows a wide range
of microstructures, which can be idealized as beam or beam–plate configurations (Gib-
son,1985), and the solid phase again is an inhomogeneous material at a lower length
scale. In studying the mechanical behavior of cancellous bone, large three-dimensional
unit cell models based on tomographic scans of actual samples and using voxel-based
discretization schemes have become fairly widely used, see e.g. (Hollister et al.,1994; van
Rietbergen et al.,1999).
Periodic microfield methods analogous to those discussed in sections 5.1 to 5.5 can be
used to study linear diffusion-type problems of the types mentioned in section 2.7, Laplace
solvers (typically for use with heat conduction problems) being available with many FE
packages. Specialized solvers are, however, required in some cases, e.g. when studying
the frequency dependent dielectric properties of composites via complex potentials, see
e.g. (Krakovsky/Myroshynchenko,2002).
There are, however, intrinsic difficulties in employing periodic microfields for transport
problems where nonlinear conductivity/resistivity behavior of the constituents is present.
Whereas in solid mechanics, material nonlinearities are typically formulated terms of
the “generalized intensity” or “generalized flux” fields, i.e. the microscopic strains and
stresses, nonlinear resisitivities or diffusivities in transport problems typically depend
on the direct variable, e.g. the temperature in heat conduction. As is evident from
fig. 14, in PMAs the generalized intensities and fluxes are periodic, whereas the direct
variables consist of fluctuating plus linear contributions, so that they accumulate from
cell to cell. As a consequence, in solid mechanics problems the position dependence of
material properties is periodic also in the presence of nonlinearities, but this is not the
case in transport problems. Accordingly, even though seemingly workable solutions, e.g.
in terms of temperature fields, can be obtained from unit cells in which temperature
dependent material properties are specified, the results do not correspond to proper
periodic solutions. Periodic microfield approaches should not be used in such cases,
embedded cell models being more suitable for studying the problems.
99)
Interestingly, instability phenomena in the cell walls of closed cell foams may even occur
under uniaxial tensile macroscopic loading, see (Ableidinger,2000).
– 81 –
6. EMBEDDED CELL APPROACHES
Like periodic microfield methods, Embedded Cell Approaches (ECAs) aim at predicting
the microfields in inhomogeneous materials at high spatial resolution. For this purpose
they use models consisting of a core (“local heterogeneous region”), which can range
from relatively simple configurations to highly detailed phase arrangements, embedded
in an outer region that serves mainly for introducing loads into the core, compare fig.
33. This strategy avoids some of the drawbacks of PMAs, especially the requirement
that the geometry and all microfields must be strictly periodic100) . Of course, a suitable
description of the outer region must be chosen, so that errors in the accommodation
of stresses and strains are avoided, and some care is required with respect to spurious
boundary layers which may occur at the “interfaces” between the core and the surround-
ing material101) . Although some analytical methods such as classical and generalized
self-consistent schemes, see section 2.3, may be viewed as embedding schemes, most
ECAs with more complex core microgeometries have tended to use numerical engineer-
ing methods.
CORE
(discrete phase arrangement)
INTERFACE between
core and embedding region
EMBEDDING
REGION
FIG.33: Schematic depiction of the arrangement of core, embedding region, and inter-
face in an embedded cell approach.
100)
Note that ECAs can be used without intrinsic restrictions at or in the vicinity of surfaces
and interfaces, and they can (at least in principle) handle gradients in composition and loads.
101)
Note that these interfaces are solely a consequence of the modeling approach and do not
have any physical background. The resulting boundary layers typically have a thickness of, say,
an inclusion diameter for elastic materials, but they may be longer ranged for nonlinear material
behavior.
– 82 –
Three basic types of embedding approaches can be found in the literature. One of
them uses discrete phase arrangements in both the core region and in the surrounding
material, the latter, however, being discretized by a much coarser FE mesh, see e.g.
(Sautter et al.,1993). Mesh superposition techniques, which use a coarse mesh over the
whole model plus a geometrically independent much finer mesh in regions of interest
(Takano et al.,2001) are conceptually similar to the above modeling strategy. Models of
the above types are not subject to boundary layers between core and outer regions.
In the second group of embedding methods the behavior of the outer regions is described
via appropriate smeared-out constitutive models. In the simplest case these take the
form of semiempirical or micromechanically based constitutive laws that are prescribed
a priori for the embedding zone and which must be chosen to correspond closely to the
overall behavior of the core102) . This way, conceptually simple models are obtained that
are very well suited for studying local phenomena such as the stress and strain distribu-
tions in the vicinity of crack tips or at macroscopic interfaces in composites (Chimani
et al.,1997), the growth of cracks in inhomogeneous materials (Wulf et al.,1996; Ablei-
dinger,2000), or damage due to the processing of composites (Monaghan/Brazil,1998).
A closely related approach uses appropriate coupling conditions (which may be imple-
mented via the boundary conditions of the micromodel) to link a macromodel employing
smeared-out material data with a micromodel using discrete regions for the constituents.
For a applications of such submodeling techniques see e.g. (Heness et al.,1999; Váradi et
al.,1999).
The third type of embedding scheme uses the homogenized thermomechanical response
of the core to determine the effective behavior of the surrounding medium, giving rise to
models of the self-consistent type, which are mainly employed for materials characteriza-
tion. The use of such approaches is, of course, predicated on the availability of suitable
parameterizable constitutive laws for the embedding material that can follow the core’s
instantaneous homogenized behavior with sufficient accuracy for all load cases and for
any loading history. This requirement typically can be fulfilled easily in the linear range
(see e.g. Chen et al.,1994), but leads to considerable difficulties when at least one of
the constituents shows elastoplastic or viscoplastic material behavior103) . Accordingly,
approximations have to be used (the consequences of which may be difficult to assess
102)
Note that by restricting damage effects to the core of the model and prescribing the
behavior of the damage-free composite to the embedding material, the volume fractions of
damaged regions can be varied at will in ECA models, compare e.g. (Bao,1992).
103)
Typically, the effective yielding behavior of the (inhomogeneous) core shows some de-
pendence on the first stress invariant, and, for low plastic strains, the homogenized response of
the core tends to be strongly influenced by the fractions of the elastoplastic constituents that
have actually yielded. In addition, in many cases anisotropies of the yielding and hardening
responses are introduced by the phase geometry (e.g. aligned fibers) and by the phase arrange-
ment of the core. At present, there appears to be no constitutive law that, on the one hand, can
fully account for these phenomena (compare e.g. Dvorak,1991) and, on the other hand, has the
capability of being adapted to the instantaneous response of the core by adjusting appropriate
free parameters.
– 83 –
in view of the nonlinearity and path dependence of elastoplastic material behavior) 104) ,
and models of this type are best termed “quasi-self-consistent schemes”. Such approaches
are discussed e.g. in (Bornert et al.,1994; Dong/Schmauder,1996). When a multi-particle
unit cell is used as the core in a quasi-self-consistent embedding scheme, the relaxation
of the periodicity constraints tends to make the overall responses of the embedded con-
figuration softer than that of the periodic arrangement (Bruzzi et al.,2001)105).
For materials characterization embedded cells can be loaded by homogeneous stresses
or strains applied to the outer boundaries. If crack problems (and related questions)
are to be studied, however, it is often preferable to impose displacement boundary con-
ditions corresponding to the far field behavior of suitable analytical solutions (e.g. for
the displacement field around a crack tip). Embedded cell models also allow complete
samples to be considered in “simulated experiments”, see e.g. fig. 34, which shows an
embedding model consisting of one half of a CT specimen made of an open cell foam
(Ableidinger,2000). In it, the embedding region is described as a homogenized elastic
continuum, in the core region a highly idealized beam model is used for the open cell
foam, and the crack is assumed to run along the symmetry plane.
3 1
FIG.34: Embedded cell model of a CT specimen for studying crack progress in an open
cell metallic foam (Ableidinger,2000).
104)
Note that in case a uniaxial stress–strain diagram is adapted to the core’s behavior and
multiaxial stresses are handled via standard metal plasticity relations (e.g. using von Mises or
Hill-type plasticity), such approximations are implicitly invoked.
105)
In the study mentioned above, the difference between PMA and ECA results is relatively
small. Essentially identical responses can be expected from the two types of model if the phase
arrangement is a proper RVE and the embedding material can fully describe the homogenized
behavior of the core.
– 84 –
Core and embedding region may be planar, axisymmetric or fully three-dimensional, and
symmetries present in the geometries can, as usual, be employed to reduce the size of the
model, compare (Peters,1997). Effective and phase averaged stresses and strains from
embedded cell analyses are best evaluated from eqn.(64) or its equivalents, and in the
case of geometrically complex phase arrangements it is good practice to use only the
central regions of the core for this purpose in order to avoid possible boundary layers.
– 85 –
7. WINDOWING APPROACHES
Windowing approaches are based on placing mesoscopic test windows at random positions
in an inhomogeneous material, compare fig. 35, and subjecting the resulting samples of
the microstructure to homogeneous stress (uniform traction, “uniform static”, Neumann,
natural) and strain (uniform displacement, “uniform kinematic”, Dirichlet, essential)
boundary conditions. Provided sufficiently large and “nonpathological” windows are
chosen, predictions of this type give rise to lower and upper estimates, respectively,
compare e.g. (Nemat–Nasser/Hori,1993), and ensembles of such results provide lower
106)
For all methods studying discrete microstructures the influence of boundary perturba-
tions decreases with growing size of the volume element and vanishes for a proper RVE, compare
the study on different square arrangements of fibers subjected to homogeneous stress and strain
boundary conditions given in (Michel,2001).
107)
Bounds of this type were originally developed to obtain rigorous results for the apparent
elastic properties of inhomogeneous materials on the basis of volume elements that are known
to be too small to be proper RVEs (Huet,1990).
– 86 –
and upper bounds on the overall behavior of the material108) . By definition, for proper
reference volume elements such lower and upper estimates and bounds on the overall
elastic properties must coincide (Hill,1963).
By using a series of windows of increasing size a hierarchy of bounds can be generated
for a given microstructure, which allows to assess the dependence of the predicted overall
moduli on the size of the windows109) . Alternatively, one of the bounds may used as
estimate for the overall response of an inhomogeneous material110) , see e.g. (Zohdi,1999).
The concept of hierarchical bounds obtained from samples that are smaller than RVEs has
been verified by planar (Amieur,1994) and three-dimensional (Guidoum,1994) modeling
schemes and by experimental studies (Huet,1999).
Recently windowing approaches have been extended into the nonlinear range (Jiang et
al.,2001), where the bounding properties of the two types of boundary conditions can be
proven within the context of deformation theory. For relatively small windows (having
a ratio between inclusion diameter and window size of, say, 5) natural and essential
uniform boundary conditions can give rise to marked differences in the distributions of
the microfields, especially the plastic strains111) . When larger windows are used, however,
increasingly similar microfields are obtained which allows to assess the dependence of the
predicted overall (instantaneous or secant) moduli on the size of the windows.
A conceptually different type of windowing models was reported by Soppa et al.(2003),
who subjected rectangular test windows to experimentally determined boundary dis-
placements. The main interest in such approaches lies in correlating microstrain fields
obtained by experiments and by modelling.
108)
It was shown by Suquet (1987) that elastic tensors obtained with periodicity bound-
ary conditions always lie between results generated with uniform traction and uniform strain
boundary conditions; in (Jiang et al.,2002) a similar result is reported for deformation plasticity.
109)
In analogy, windows of increasing size subjected to periodicity or symmetry boundary
conditions give rise to hierarchies of estimates, see e.g. (Sautter et al.,1993; Terada et al.,2000).
110)
For materials with matrix–inclusion topology, the lower bounds (and thus homogeneous
stress boundary conditions) typically give good estimates in case the reinforcements are stiffer
than the matrix, whereas the upper bounds (and thus homogeneous strain boundary conditions)
are better suited when the matrix is stiffer than the reinforcements.
111)
The application of the above boundary conditions at the surfaces of a window gives rise
to a boundary layer in which the microfields are perturbed; its width depends on the material
behavior.
– 87 –
8. MULTI-SCALE MODELS
scale scale
transition transition
#1 #2
112)
Describing the material behavior at lower length scales by a homogenized model implies
that characteristic lengths differ by, say, two orders of magnitude or more, so that valid volumes
can be defined for homogenization. It is not technically correct to employ hierarchical approaches
within “bands” of more or less continuous distributions of length scales, as can be found e.g. in
some metallic foams with highly disperse cell sizes.
– 88 –
above methods, i.e. mean field, unit cell, and embedding approaches, may be used as
“building blocks” at any level within hierarchical schemes113) . Such multi-scale model-
ing strategies have the additional advantage of allowing the behavior of the constituents
at all lower length scales to be assessed via the corresponding localization relations 114) .
Because local stresses and strains at the lower length scales can differ markedly from
the applied macroscopic conditions in both magnitude and orientation, it is advisable
that the homogenized descriptions of the thermomechanical responses at all length scales
(with the possible exception of the top one) take the form of proper constitutive models
that are capable of handling any loading conditions and any loading history. In most
cases, this requirement can readily be fulfilled by micromechanically based constitutive
models that employ mean field methods or periodic microfield approaches.
Among the continuum mechanical multi-scale descriptions of the thermomechanical be-
havior of inhomogeneous materials reported in the literature, some combine mean field
methods at the higher length scale with mean field (Hu et al.,1998; Tszeng,1998) or
periodic microfield approaches (González/LLorca,2000) at the lower length scale. The
most common strategy for multi-scale modeling, however, uses Finite Element based unit
cell or embedding methods at the topmost length scale, which implies that the homoge-
nized material models describing the lower level(s) of the hierarchy must take the form
of micromechanically based constitutive laws that can be evaluated at each integration
point.
This requirement typically does not give rise to problematic computational workloads in
the elastic range, where the superposition principle can be used to obtain the full ho-
mogenized elasticity and thermal expansion tensors from a limited number of analyses of
an appropriate unit cell115) . For simulating the thermomechanical response of nonlinear
inhomogeneous materials, however, essentially a full micromechanical submodel has to
be maintained and solved at each integration point in order to account for the history
dependence of the local responses116) . Even though within such a framework the use of
113)
Micromechanical methods that by construction describe inhomogeneous materials with
a wide range of length scales, e.g. differential schemes, are, however, of limited suitability for
multi-scale analyses.
114)
It should be noted, however, that generating predictions for all relevant fields at two or
more length scales can give rise to very large amounts of data, especially if models based on
numerical discretizing methods are used at one or more of the length scales.
115)
In the elastic range domain decomposition techniques can be used to formulate the prob-
lems at the lower length scales in such a way that they are well suited for parallel process-
ing, allowing the development of computationally highly efficient multi-scale procedures, see
(Oden/Zohdi,1997).
116)
Some approaches have been reported that aim at partially “decoupling” analyses at the
higher and lower length scales by parameterizing results from unit cell analyses and using the
stored data together with some interpolation scheme as an “approximate constitutive model”,
see e.g. (Terada/Kikuchi,1996; Gänser,1998; Schrefler et al.,1999; Ghosh et al.,2001). The main
difficulty in such modeling strategies lies in handling load history and load path dependences
to account for elastoplastic constituents or microscopic damage to constituents. Appropriate
representation of general multiaxial stress and/or strain states also tends to be a challenge.
– 89 –
Eps_eff,p_(m)
3.5000E-03
3.0000E-03
2.5000E-03
2.0000E-03
1.5000E-03
FIG.37: Phase averaged microscopic equivalent plastic strains in the matrix within the
inclusion-poor regions of a cluster-structured high speed steel under mechanical loading
as predicted by a mesoscopic unit cell model combined with an incremental Mori–Tanaka
model at the microscale (Plankensteiner,2000).
sophisticated unit cell based models at the lower length scales usually is a very expensive
proposition in terms of computational requirements, work of this type was reported e.g.
by (Wu et al.,1989), and asymptotic homogenization has been used successfully in this
context for two-dimensional analyses (Ghosh et al.,1996). Lower (but by no means negli-
gible) computational costs can be achieved by using constitutive models based on mean
field approaches, e.g. incremental Mori–Tanaka methods or corresponding versions of the
Transformation Field Analysis, compare section 2.5117) . It is also possible to handle the
“lowest” scale transition by a combination of a mean field or TFA-type homogeniza-
tion procedure and unit-cell based localization algorithms, the latter being used only in
regions of special interest (Fish et al.,1997).
Figure 37 shows a result obtained by applying a relatively straightforward multi-scale
model to the materials characterization of a cluster-structured high speed steel. At the
mesoscale, a unit cell approach is used in which particle-rich clusters and inclusion-poor
regions are described as separate phases. Each of the latter is treated as a particle rein-
forced MMC with appropriate inclusion volume fraction at the microscale, an incremental
Mori–Tanaka model being employed for homogenization. Such an approach not only pre-
117)
Multi-scale approaches using micromechanically based constitutive laws may lead to spu-
rious stress and strain fields close to interior boundaries at the higher length scale, e.g. between
reinforced and monolithic regions in selectively reinforced components, because in such regions
the conditions for standard homogenization methods may not be fulfilled locally, compare e.g.
(Chimani et al.,1997).
– 90 –
dicts macroscopic stress–strain relationships, but also allows the distributions of phase
averaged microscopic variables to be evaluated, compare (Plankensteiner et al.,1998) 118).
Alternatively, multiscale approaches may be rather “loosely coupled”, essentially link-
ing together quite different models to obtain an overall result, see e.g. (Onck/van der
Giessen,1997).
In the last few years there has been a surge of research interest in developing and
applying algorithms for the multi-scale and hierarchical continuum modeling of in-
homogeneous materials, see e.g. (Lee/Ghosh,1996; Zohdi et al.,1996; Fish/Shek,2000;
Feyel/Chaboche,2000; Ibrahimbegović/Markovič,2003; Moës et al.,2003). A recent de-
velopment are multi-scale methods in which the computational domain is adaptively
split into regions resolved at appropriate length scale. In an FE-based framework de-
veloped Ghosh et al. (2001) “non-critical” regions are described by continuum elements
with homogenized material properties, and at zones of high macroscopic stress and strain
gradients PMA models are automatically activated at the integration points to monitor
the phase behavior at the microscale. If violations of the homogenization conditions (e.g.
onset of damage, mesoscopic interfaces, free edges) are detected, embedded models of the
fully resolved microstructure are activated at the appropriate positions. This strategy
allows detailed studies of critical regions in inhomogeneous materials, e.g. near free edges.
In an alternative approach, homogenization models employing generalized continua have
been proposed for handling regions close to free surfaces and zones with high local field
gradients (Feyel,2003).
Finally, it should be noted that multi-scale approaches are not limited to using the “stan-
dard” methods of continuum micromechanics as discussed above, see e.g. the discussions
on metal plasticity in (Tadmor et al.,2000). Especially the capability of the Finite Ele-
ment method of handling highly complex constitutive descriptions has been used to build
multi-scale descriptions that employ, for example, material models based on atomistics
(Shenoy et al.,1998). Evidently, the distinction between micromechanical studies using
sophisticated constitutive models for the phases on the one hand and multi-scale model-
ing on the other hand as used here is to some extent a matter of definition (and personal
preferences).
118)
When assessing results from multi-scale analyses using mean field methods at the smallest
length scale, such as fig. 37, it must be kept in mind that the maximum stresses and strains
shown are the maxima of the local phase averaged values and not maximum microscopic stresses
as obtained by “standard” micromechanical unit cell studies.
– 91 –
9. CLOSING REMARKS
119)
For this special case unit cell or embedding models employing anisotropic or crystal
plasticity models for the phases can be used which, however, can become very large when the
statistics of grain orientation are accounted for.
– 92 –
the constitutive models of the constituents120) . Absolute length scales can be provided
“explicitly” via discrete dislocation models (Cleveringa et al.,1997), via gradient or non-
local constitutive laws see e.g. (Tomita et al.,2000; Niordson/Tvergaard,2002), and dam-
age models, or “implicitly”, e.g. by adjusting the phase material parameters to account
for grain sizes via the Hall–Petch effect. Analysts should also be aware that absolute
length scales may be introduced inadvertently into a model by mesh dependence effects
of discretizing numerical methods, a “classical” example being strain localization due to
softening121) in the material behavior of a constituent, compare also the discussions in
the Appendix. When multi-scale models are used special care may be necessary to avoid
introducing inconsistent length scales at different modeling levels.
In addition it is worth noting that usually the macroscopic response of inhomogeneous
materials is much less sensitive to the phase arrangement (and to modeling approx-
imations) than are the distributions of the microfields. Consequently, whereas good
agreement in the overall behavior of a given model with “benchmark” theoretical results
or experimental data typically indicates that the phase averages of stresses and strains
are described satisfactorily122) , it does not necessarily imply that the higher statistical
moments of the stress and strain distributions have been captured correctly.
It is important to be aware that work in the field of micromechanics of materials invariably
involves finding viable compromises in terms of the complexity of the models, which, on
the one hand, have to be able to account (at least approximatively) for the physical
phenomena relevant to the given problem, and, on the other hand, must be sufficiently
simple to allow solutions to be obtained within the relevant constraints of time, cost, and
computational resources. Obviously, actual problems cannot be solved without recourse
to various approximations and tradeoffs — the important point is to be aware of them.
It is worth keeping in mind that there is no such thing as a “best micromechanical
approach” for all applications and that models, while indispensable in understanding
the behavior of inhomogeneous materials, are “just” models and do not reflect the full
complexities of real materials.
120)
One important exception are models for studying macroscopic cracks (e.g. via embedded
cells), where the crack length does introduce an absolute length scale.
121)
Note that here the expression “localization” is used to describe a physical phenomenon
in materials and not in its “mathematical” meaning as introduced in section 1.2.
122)
In the elastoplastic range heterogeneities of the stress and strain fields within the phases
also play a considerable role, which can be approximately described by the second statistical
moments of the microfields.
– 93 –
APPENDIX: SOME ASPECTS OF MODELING DAMAGE AT THE
MICROLEVEL
Modeling efforts aimed at studying the damage and failure behavior of inhomogeneous
materials may take the form of microscopic or macroscopic approaches. Methods be-
longing to the latter group often are based to some extent on micromechanical con-
cepts and encompass, among others, the “classical” failure surfaces for continuously
reinforced laminae such as the Tsai–Wu criterion (Tsai/Wu,1971) and macroscopic con-
tinuum damage models of various degrees of complexity, see e.g. (Chaboche et al.,1998;
Voyiadjis/Kattan,1999). Also on the macroscopic level the Considére criterion may be
used in combination with micromechanical analyses to check for (macroscopic) plastic
instability when studying the ductility of inhomogeneous elastoplastic materials, see e.g.
(González/LLorca,1996). In the following, however, only the modeling of damage on the
microscale in the context of (mainly Finite Element based) micromechanical methods of
the types discussed in chapters 4 to 6 will be considered.
General Remarks
When damage and failure of inhomogeneous materials are to be studied at the microscale,
it is physically intuitive to represent them as acting at the level of the constituents and of
the interfaces between them, where appropriate damage and fracture models have to be
provided. All of the resulting microscopic failure mechanisms contribute to the energy ab-
sorption during crack growth. Taking MMCs as an example, this concept translates into
viewing damage and failure as involving a mutual interaction and competition between
ductile damage of the matrix, brittle cleavage of the reinforcements, and decohesion of
the interfaces. The dominant microscopic damage process is decided by the individual
constituents’ strength parameters, by the microgeometry and by the loading conditions,
compare (Needleman et al.,1993).
If only information on the initiation of damage is required it may be sufficient to eval-
uate appropriate “damage relevant fields” (e.g. the maximum principal stress in brittle
materials, a ductile damage indicator as described by eqn.(69) for metals, or the traction
vectors at interfaces) without accounting for the loss of stiffness that accompanies the
evolution of damage. Alternatively, it may be of interest to study microstructures con-
taining flaws (e.g. fiber breaks, matrix flaws) in order to assess the resulting microfields,
see e.g. (Tsotsis/Weitsman,1990), where J -integrals were evaluated for specific matrix
crack configurations.
Constituent damage at the microscale can be described by standard continuum damage
mechanics (CDM) approaches, the implication being that the characteristic length scale
of the damage (pores, microcracks, etc.) is sufficiently smaller than the characteristic
length scale of phase arrangement to be studied (essentially such models amount to multi-
scale methods as discussed in section 8). CDM methods can account for the stiffness
degradation due to the evolution of damage, which tends to give rise to localization
and may lead to the formation of proper shear bands within the microscopic phase
arrangement. Alternatively, the effects of ductile damage may be introduced via ductile
– 94 –
damage indicator triggered (DDIT) techniques, in which the stiffness of an element is
drastically reduced or the element removed when a certain value of a ductile damage
indicator is reached.
A numerical difficulty typically encountered with CDM and DDIT models lies in their
a tendency towards a marked mesh dependence of the solutions123) (typically the char-
acteristic size of the localized region is comparable to the dimensions of an element,
regardless of the actual degree of mesh refinement). This behavior can be avoided by
using continuum damage theories that are enriched by higher order gradients, see e.g.
(Geers et al.,2001), by introducing nonlocal averaging of an appropriate internal vari-
able, compare (Jirásek/Rolshoven,2002), or by employing time dependent formulations
involving rate effects (viscoelasticity, viscoplasticity), compare (Needleman,1987). Alter-
natively, elements of an appropriate constant size (“cell elements”, Xia/Shi,1995) may
be employed. All of the above procedures explicitly introduce an absolute length scale
into the problem124) . A rather ad-hoc alternative consists in selecting a mesh size that
leads to physically reasonable sizes of localization features.
Macrocracks as well as microcracks that have characteristic lengths comparable to the
microscale of a given phase arrangement must be resolved explicitly in micromechanical
studies (and cannot be smeared out as “damage”). For describing them “standard” FE
models of cracks may be used; for an overview covering some of the relevant methods
see e.g. (Liebowitz et al.,1995). When crack growth is to be considered, local criteria for
the initiation and/or growth of cracks (J -integral, CTOD or CTOA criteria, . . .) as well
as appropriate algorithms for geometrically modeling the generation and/or extension of
cracks must be provided.
Numerical strategies for simulating the progress of cracks typically fall into three groups,
the first of which emphasizes the concept that the nucleation and growth of cracks cor-
respond to the creation of new free surface, leading to “surface oriented” discrete crack
models in which crack growth involves breaking the connectivity between neighboring
elements, e.g. by node release techniques (Bakuckas et al.,1993; Narasimhan,1994) or by
cohesive surface elements, compare (Elices et al.,2002). Evidently, in such approaches
the growth of cracks is restricted to element boundaries at which appropriate provisions
for crack opening have been made, so that mesh dependences of the predicted crack
path tends to be a problem. In cases where the crack path is not known a priori such
mesh dependences can be counteracted either by using fine and irregular discretizations
or by introducing remeshing schemes that locally adapt the element geometry for each
crack growth increment, compare e.g. (Fish et al.,1999). Both of these remedies tend to
lead to large and complex models, so that in micromechanics of materials crack opening
123)
Within classical continuum theory strain localization leads to ill-posed boundary value
problems, i.e. the governing equations lose their ellipticity or hyperbolicity, giving rise to the
mesh sensitivity of the solutions, see e.g. (Baaser/Gross,2002). In order to provide an objective
description of localization, the problem has to be regularized.
124)
A precondition for the proper regularizaton by higher order gradients or by nonlocal
averaging is that the discretization must be finer than than this absolute length scale.
– 95 –
approaches have been reported mainly for symmetrical phase arrangements where the
crack paths are known to follow some plane of symmetry.
The mesh dependence of the crack paths predicted by surface oriented approaches can
be largely mitigated by using specially formulated solid elements, known as embedded
discontinuity elements. These, on the one hand, allow discontinuities to develop at any
position and in any direction within each element and on the other hand, provide for
continuous crack paths, see e.g. (Jirásek,2000; Alfaiate et al.,2002). Embedded discon-
tinuity elements at present are in a developmental stage, are typically not available in
standard FE packages, and up till now do not seem to have been used in the context of
continuum micromechanics.
The third group of approaches for modeling the progress of cracks is based on the idea
of a fracture process zone within which the constitutive parameters of the material are
degraded. This results in “volume oriented” smeared-out models of cracks (“fictitious
crack models”), see e.g. (Weihe/Kröplin,1995). Many such models are closely related to
damage models.
All three groups of formulations involve some characteristic length scale, so that an
absolute length scale is introduced into the micromechanical model, either explicitly
via the formulation or implicitly via the discretization (i.e. via the element sizes along
the crack path). Of course, appropriate numerical schemes are required for handling
the overall softening and the local stress redistribution caused by crack propagation or
element elimination.
In composite materials with matrix–inclusion topology, damage may either occur in the
reinforcements, in the matrix, or within the interface between them. Similarly, in poly-
crystals intergranular or intragranular failure may be observed. The dominant micro-
scopic damage process is decided by the strength parameters of the constituents and
interfaces, by the microgeometry, and by the loading conditions.
– 96 –
(Monte-Carlo-type procedure) is provided which decides if a given reinforcement will
fail at a given load level on the basis of appropriate fracture probabilities, compare
e.g. (Pandorf,2001; Eckschlager,2002). Within micromechanical analyses such fracture
probabilities can be evaluated from the predicted distributions of the microstresses by
Weibull-type weakest link models125) that were originally developed for describing the
failure statistics of bulk brittle materials. A simple Weibull expression for evaluating the
failure probability Pn of a given particle n may take the form (Eckschlager,2002)
( )
1 σ (x) m
Z
1
Pn = 1 − exp − dΩ , (68)
V0 Ωn :σ1 >0 σ0
where σ1 (x) is the microscopic field of maximum principal stresses in the inclusion and
Ω:σ1 >0 is that part of the particle volume in which the maximum principal stress is
tensile. Equation (68) involves three material parameters, viz. σ0 , which stands for
the particle’s average strength, m, which is known as the Weibull modulus, and V0 , a
reference volume. Note that expressions such as eqn.(68) assign a failure probability to
a particle or inclusion as a whole126) , which fits well with the concept of “instantaneous
cleavage” of particles.
In models of discontinuously reinforced composites that use axisymmetric cells, a natural
choice for the position of a predefined crack is a symmetry plane that is oriented normal
to the axis of rotation (which coincides with the loading direction) and runs through the
center of the inclusion, see e.g. (Bao,1992; Davis et al.,1998; González/LLorca,2000). For
more general arrangements of particles or short fibers the most realistic approach would
be to consider the stress state of each particle in selecting the crack plane127) , so that
perturbations due to (possibly failed) neighboring particles can be accounted for. Due
to their high computational requirements up till now such algorithms have only been
reported for planar arrangements of ellipsoids reinforcements, the crack being taken to
pass through the position of maximum principal stress in the particle and to be oriented
normally to the local maximum principal stress at this point (Ghosh et al.,2001).
For three-dimensional particle arrangements a simpler concept for approximating the
orientations and positions of cracks within particles is based on the assumption that
125)
Whereas there is agreement that the weakest-link behavior of long fibers is mainly due
to surface defects, the physical interpretation is not as clear for Weibull models of particles
embedded in a ductile matrix. Weibull models based on eqn.(68) or similar expressions, however,
have a number of advantages for the latter class of applications, compare (Wallin et al.,1987),
not the least of which is their capability for describing particle size effects. It should be noted,
however, that simple Weibull-type models were reported to give predictions for the dependence
of the number of fracture events per volume on particle size and volume fraction that do not
agree with acoustic emission data from ductile matrix composites (Mummery et al.,1993).
126)
As a consequence, such approaches are not well suited for providing information on the
position of crack initiation within the particle.
127)
For small monocrystalline particulates the crystallographic orientations can also be ex-
pected to play a role in determining the crack plane, thus introducing a statistical element into
the problem, compare e.g. the experimental results in (Mawsouf,2000).
– 97 –
they passes through the particle center and is oriented normal to the average maximum
principal stress in the inclusion or to the macroscopic maximum principal stress. Such
relatively simple models are adequate for qualitative studies of the fracture of spherical
particles (Eckschlager et al.,2001), but local stress concentrations must be considered for
particulates of irregular shape and multiple cracks must be provided for in modeling the
fracture of elongated particles.
An alternative to the above approach of instantaneous particle cracking consists in using
cohesive surface models for opening predefined cracks in reinforcements, see e.g. (Finot
et al.,1994; Tvergaard,1994). Such models are capable of resolving the growth sequence
of a predefined crack (rather than “popping” it open instantaneously).
where εeq,p stands for the accumulated equivalent plastic strain, σm for the mean (hy-
drostatic) stress, σeq for the equivalent (von Mises) stress, and εf for the uniaxial failure
strain of the material. This ductile damage indicator has been used in continuum mi-
cromechanics in the sense of a “damage relevant field” e.g. in (Gunawardena et al.,1993;
Böhm/Rammerstorfer,1996).
For micromechanical studies that employ “progressively cavitating” Gurson-type models
for describing ductile failure of the matrix of MMCs see e.g. (Needleman et al.,1993,
Geni/Kikuchi,1999); Gurson models were also incorporated into Voronoi cell elements
(Hu et al.,2004). Alternatvely, the ductile damage model of Rousselier (1987) can be
used in micromechanical models (Drabek/Böhm,2004). The use of DDIT element elimi-
nation schemes was reported in the context of continuum micromechanics of inhomoge-
neous materials e.g. in (Schmauder et al.,1996; Berns et al.,1998). Other criteria used
128)
In contrast to the damage parameters used in CDM, which allow the degree of damage
to be followed throughout the failure process, only two values of the damage parameter D id
defined in eqn.(69) have a direct physical interpretation: 0 indicates no damage and 1 ductile
failure. Accordingly, in contrast to CDM models, DDIT approaches cannot follow the degrada-
tion of stiffness as damage develops. The main advantage of the ductile damage indicator lies in
requiring only one material parameter, εf , which can be evaluated experimentally in a relatively
straightforward way, compare the discussion in (Fischer et al.,1995).
– 98 –
for controlling element elimination in ductile matrices have included strain energy crite-
ria (Mahishi/Adams,1982), the so called s-criterion (Spiegler et al.,1990), plastic strain
criteria (Adams,1974), and stress triaxiality criteria (Steinkopff et al.,1995). Wulf et
al. (1996) reported comparisons between predictions obtained by planar embedded cell
analyses using element elimination models controlled by stress triaxiality, plastic strain,
and Rice–Tracey criteria, the latter giving the best agreement with experimental re-
sults for that special case. The use of nonlocal versions of damage models and DDIT
schemes, which allow the mesh dependence of the solutions to be controlled, is a recent
development, see e.g. (Böhm et al.,2004).
Approaches for modeling matrix failure that do not fall into the above group were re-
ported e.g. by Finot et al. (1994), who employed a cohesive surface model to simulate
crack propagation along a symmetry plane from the reinforcement into the matrix of short
fiber reinforced MMCs, and by Biner (1996), who studied creep damage to the matrix
of an MMC using Tvergaard’s (1984) grain boundary cavitation model. Mahishi (1986)
combined a node release technique with an energy release criterion to study cracks in poly-
mer matrices129) . Composites with brittle elastic matrix behavior can be investigated via
fracture mechanics based methods see e.g. (Ismar/Reinert,1997), or Weibull-type criteria
may be employed (Brockmüller et al.,1995).
Finally, it may be mentioned that in the case of cyclic loading shakedown theorems
may be used on the microscale to assess if elastic shakedown may take place in an
inhomogeneous material or if low cycle fatigue will act on the microscale, see (Böhm,1993;
Schwabe,2000).
129)
Because crack paths in inhomogeneous materials may be rather complicated, appropriate
criteria for determining the crack direction have to be employed and remeshing typically is
necessary when node release techniques are used, compare (Vejen/Pyrz,2002).
– 99 –
gether in a “shrink fit” by the thermal stresses caused by cooling down from the processing
temperature, see e.g. (Gunawardena et al.,1993; Sherwood/Quimby,1995). As an exten-
sion of this type of description, Coulomb-type models, i.e. a linkage between normal and
shear stresses analogous to that employed for dry friction, may be used for the interface
(Benabou et al.,2002).
Other interfacial decohesion models reported in the literature include strain softening in-
terfacial springs (Vosbeek et al.,1993; Li/Wisnom,1994), a node release technique based
on fracture mechanics solutions for bimaterial criterion (Moshev/Kozhevnikova,2002).
Interphases, i.e. interfacial layers of finite thickness, were studied by various ficti-
tious crack techniques, see e.g. (Ismar/Schmitt,1991; McHugh/Connolly,1994; Walter
et al.,1997) and thermomechanically based continuum damage models (Benabou et
al.,2002).
A rather different approach has been developed specifically for accounting for the stiff-
ness loss due to debonding within mean field models. It is based on describing debonded
reinforcements by perfectly bonded (transversally isotropic) “fictitious inclusions” of ap-
propriately reduced stiffness. Within such a framework, the fraction of debonded rein-
forcements may be controlled via Weibull-type models, see e.g. (Zhao/Weng,1995; Sun
et al.,2003).
– 100 –
REFERENCES
Ableidinger A.: Some Aspects of the Fracture Behavior of Metal Foams. Diploma Thesis,
Vienna University of Technology, Vienna, Austria, 2000.
Aboudi J.: Micromechanical Analysis of Composites by the Method of Cells;
Appl.Mech.Rev. 42, 193–221, 1989.
Aboudi J.: Mechanics of Composite Materials. Elsevier, Amsterdam, The Netherlands,
1991.
Aboudi J.: Micromechanical Analysis of Composites by the Method of Cells — Update;
Appl.Mech.Rev. 49, S83–S91, 1996.
Aboudi J., Pindera M.J., Arnold S.M.: Higher-Order Theory for Periodic Multiphase
Materials with Inelastic Phases; Int.J.Plast. 19, 805–847, 2003.
Accorsi M.L.: A Method for Modeling Microstructural Material Discontinuities in a Fi-
nite Element Analysis; Int.J.Num.Meth.Engng. 26, 2187–2197, 1988.
Achenbach J.D., Zhu H.: Effect of Interfacial Zone on Mechanical Behavior and Failure
of Fiber-Reinforced Composites; J.Mech.Phys.Sol. 37, 381–393, 1989.
Adams D.F.: A Micromechanical Analysis of Crack Propagation in an Elastoplastic Com-
posite Material; Fibre Sci.Technol. 7, 237–256, 1974.
Adams D.F., Crane D.A.: Finite Element Micromechanical Analysis of a Unidirectional
Composite Including Longitudinal Shear Loading; Comput.Struct. 18, 1153–1165, 1984.
Adams D.F., Doner D.R.: Transverse Normal Loading of a Uni-Directional Composite;
J.Compos.Mater. 1, 152–164, 1967.
Advani S.G., Tucker C.L.: The Use of Tensors to Describe and Predict Fiber Orientation
in Short Fiber Composites; J.Rheol. 31, 751–784, 1987.
Alfaiate J., Wells G.N., Sluys L.J.: On the Use of Embedded Discontinuity Elements with
Crack Path Continuity for Mode-I and Mixed Mode Fracture; Engng.Fract.Mech. 69,
661–686, 2002.
Allen D.H., Lee J.W.: The Effective Thermoelastic Properties of Whisker-Reinforced
Composites as Functions of Material Forming Parameters; in “Micromechanics and In-
homogeneity” (Eds. G.J.Weng, M.Taya, H.Abé), pp. 17–40; Springer–Verlag, New York,
NY, 1990.
Amieur M.: Etude numérique et expérimentale des effets d’échelle et de conditions aux
limites sur des éprouvettes de béton n’ayant pas le volume représentatif. Doctoral Thesis,
Ecole Polytécnique Fédérale de Lausanne, Lausanne, Switzerland, 1994.
Anthoine A.: Derivation of the In-Plane Elastic Characteristics of Masonry Through
Homogenization Theory; Int.J.Sol.Struct. 32, 137–163, 1995.
Antretter T.: Micromechanical Modeling of High Speed Steel. VDI–Verlag (Reihe 18,
Nr.232), Düsseldorf, Germany, 1998.
Arenburg R.T., Reddy J.N.: Analysis of Metal-Matrix Composite Structures — I. Mi-
cromechanics Constitutive Theory; Comput.Struct. 40, 1357–1368, 1991.
Axelsen M.S., Pyrz R.: Correlation Between Fracture Toughness and the Microstruc-
ture Morphology in Transversely Loaded Unidirectional Composites; in “Microstructure–
Property Interactions in Composite Materials” (Ed. R.Pyrz), pp. 15–26; Kluwer Aca-
demic Publishers, Dordrecht, The Netherlands, 1995.
– 101 –
Baaser H., Gross D.: On the Limitations of CDM Approaches to Ductile Fracture Prob-
lems; in “Proc. Fifth World Congress on Computational Mechanics (WCCM V)” (Eds.:
H.A.Mang, F.G.Rammerstorfer, J.Eberhardsteiner), paper # 81462; Vienna University
of Technology, Vienna, 2002.
Bakuckas J.G., Tan T.M., Lau A.C.W., Awerbuch J.: A Numerical Model for Pre-
dicting Crack Path and Modes of Damage in Unidirectional Metal Matrix Composites;
J.Reinf.Plast.Compos. 12, 341–358, 1993.
Banerjee P.K., Henry D.P.: Elastic Analysis of Three-Dimensional Solids with Fiber
Inclusions by BEM; Int.J.Sol.Struct. 29, 2423–2440, 1992.
Banks–Sills L., Leiderman V., Fang D.: On the Effect of Particle Shape and Orientation
on Elastic Properties of Metal Matrix Composites; Composites 28B, 456–481, 1997.
Bao G.: Damage Due to Fracture of Brittle Reinforcements in a Ductile Matrix;
Acta.metall.mater. 40, 2547–2555, 1992.
Bao G., Hutchinson J.W., McMeeking R.M.: Particle Reinforcement of Ductile Matrices
Against Plastic Flow and Creep; Acta metall.mater. 39, 1871–1882, 1991.
Benabou L., Benseddiq, Naı̈t–Abdelaziz M.: Comparative Analysis of Damage at Inter-
faces of Composites; Composites 33B, 215–224, 2002.
Benveniste Y.: A New Approach to the Application of Mori–Tanaka’s Theory in Com-
posite Materials; Mech.Mater. 6, 147–157, 1987.
Benveniste Y.: Some Remarks on Three Micromechanical Models in Composite Media;
J.Appl.Mech. 57, 474–476, 1990.
Benveniste Y., Dvorak G.J.: On a Correspondence Between Mechanical and Ther-
mal Effects in Two-Phase Composites; in “Micromechanics and Inhomogeneity” (Eds.
G.J.Weng, M.Taya, H.Abé), pp. 65–82; Springer–Verlag, New York, NY, 1990.
Benveniste Y., Dvorak G.J., Chen T.: On Diagonal and Elastic Symmetry of the Approx-
imate Effective Stiffness Tensor of Heterogeneous Media; J.Mech.Phys.Sol. 39, 927–946,
1991.
Beran M.J., Molyneux J.: Use of Classical Variational Principles to Determine Bounds
for the Effective Bulk Modulus in Heterogeneous Media; Quart.Appl.Math. 24, 107–118,
1966.
Berns H., Melander A., Weichert D., Asnafi N., Broeckmann C., Gross–Weege A.: A
New Material for Cold Forging Tools; Comput.Mater.Sci. 11, 166–188, 1998.
Berryman J.G.: Long-Wavelength Propagation in Composite Elastic Media, II. Ellip-
soidal Inclusions; J.Acoust.Soc.Amer. 68, 1820–1831, 1980.
Berryman J.G., Berge P.A.: Critique of Two Explicit Schemes for Estimating Elastic
Properties of Multiphase Composites; Mech.Mater. 22, 149–164, 1996.
Berveiller M., Zaoui A.: An Extension of the Self-Consistent Scheme to Plastically Flow-
ing Polycrystals; J.Mech.Phys.Sol. 26, 325–344, 1979.
Bhattacharyya A., Weng G.J.: The Elastoplastic Behavior of a Class of Two-Phase
Composites Containing Rigid Inclusions; Appl.Mech.Rev. 47, S45–S65, 1994.
Biner S.B.: A Finite Element Method Analysis of the Role of Interface Behavior in the
Creep Rupture Characteristics of a Discontinuously Reinforced Composite with Sliding
Grain Boundaries; Mater.Sci.Engng. A208, 239–248, 1996.
Bisegna P., Luciano R.: Bounds on the Off-Diagonal Coefficients of the Homogenized
Constitutive Tensor of a Composite Material; Mech.Res.Comm. 23, 239–246, 1996.
Bishop J.F.W., Hill R.: A Theory of the Plastic Distortion of a Polycrystalline Aggregate
Under Combined Stress; Phil.Mag. 42, 414–427, 1951.
– 102 –
Bochenek B., Pyrz R.: Reconstruction Methodology for Planar and Spatial Random Mi-
crostructures; in “New Challenges in Mesomechanics” (Eds. R.Pyrz, J.Schjødt–Thomsen,
J.C.Rauhe, T.Thomsen, L.R.Jensen), pp. 565–572; Aalborg University, Aalborg, Den-
mark, 2002.
Böhm H.J.: Numerical Investigation of Microplasticity Effects in Unidirectional Longfiber
Reinforced Metal Matrix Composites; Modell.Simul.Mater.Sci.Engng. 1, 649–671, 1993.
Böhm H.J.: Notes on Some Mean Field Approaches for Composites. CDL–FMD Report
1–1998, Vienna University of Technology, Vienna, Austria, 1998.
Böhm H.J.: A Short Introduction to Continuum Micromechanics; in “Mechanics of Mi-
crostructured Materials” (Ed. H.J.Böhm), pp. 1–40; Springer Verlag, Vienna, 2004 (in
print).
Böhm H.J., Han W.: Comparisons Between Three-Dimensional and Two-Dimensional
Multi-Particle Unit Cell Models for Particle Reinforced MMCs; Modell.Simul.Mater.Sci.
Engng. 9, 47–65, 2001.
Böhm H.J., Rammerstorfer F.G.: Fiber Arrangement Effects on the Microscale Stresses
of Continuously Reinforced MMCs; in “Microstructure–Property Interactions in Com-
posite Materials” (Ed. R.Pyrz), pp. 51–62; Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1995.
Böhm H.J., Rammerstorfer F.G.: Influence of the Micro-Arrangement on Matrix and
Fiber Damage in Continuously Reinforced MMCs; in “Micromechanics of Plasticity and
Damage of Multiphase Materials” (Eds. A.Pineau, A.Zaoui), pp. 19–26; Kluwer, Dor-
drecht, The Netherlands, 1996.
Böhm H.J., Eckschlager A., Han W.: Modeling of Phase Arrangement Effects in High
Speed Tool Steels; in “Tool Steels in the Next Century” (Eds. F.Jeglitsch, R.Ebner,
H.Leitner), pp. 147–156; Montanuniversität Leoben, Leoben, Austria, 1999.
Böhm H.J., Eckschlager A., Han W.: Multi-Inclusion Unit Cell Models for Metal Matrix
Composites with Randomly Oriented Discontinuous Reinforcements; Comput.Mater.Sci.
25, 42–53, 2002.
Böhm H.J., Rammerstorfer F.G., Weissenbek E.: Some Simple Models for Microme-
chanical Investigations of Fiber Arrangement Effects in MMCs; Comput.Mater.Sci. 1,
177–194, 1993.
Böhm H.J., Duschlbauer D., Drabek T., Kolednik O., Shan G.X., Chimani C.: Numerical
Studies of Crack Sensitivity under Continuous-Casting Conditions; in “Proceedings of the
VAI Continuous Casting and Hot-Rolling Conference CCR’04”, Paper 6.2;VAI, Linz,
Austria, 2004.
Bornert M.: Morphologie microstructurale et comportement mécanique; charactérisations
expérimentales, approches par bornes et estimations autocohérentes généralisées. Doc-
toral Thesis, Ecole Nationale des Ponts et Chaussées, Paris, France, 1996.
Bornert M.: Homogénéisation des milieux aléatoires: bornes et estimations; in “Ho-
mogénéisation en mécanique des materiaux 1. Matériaux aléatoires élastiques et milieux
périodiques” (Eds. M.Bornert, T.Bretheau, P.Gilormini), pp. 133–221; Editions Hermès,
Paris, 2001.
Bornert M., Suquet P.: Propriétés non linéaires des composites: Approches par les poten-
tiels; in “Homogénéisation en mécanique des matériaux 2. Comportements non linéaires
et problèmes ouverts” (Eds. M.Bornert, T.Bretheau, P.Gilormini), pp. 45–90; Editions
Hermès, Paris, 2001.
Bornert M., Stolz C., Zaoui A.: Morphologically Representative Pattern-Based Bounding
in Elasticity; J.Mech.Phys.Sol. 44, 307–331, 1996.
– 103 –
Bornert M., Bretheau T., Gilormini P. (Eds.): Homogénéisation en mécanique des
matériaux 1. Matériaux aléatoires élastiques et milieux périodiques. Editions Hermès,
Paris, 2001.
Bornert M., Bretheau T., Gilormini P. (Eds.): Homogénéisation en mécanique des
matériaux 2. Comportements non linéaires et problèmes ouverts. Editions Hermès, Paris,
2001.
Bornert M., Hervé E., Stolz C., Zaoui A.: Self Consistent Approaches and Strain Het-
erogeneities in Two-Phase Elastoplastic Materials; Appl.Mech.Rev. 47, S66–S76, 1994.
Brockmüller K.M., Bernhardi O., Maier M.: Determination of Fracture Stress and
Strain of Highly Oriented Oriented Short Fibre-Reinforced Composites Using a Fracture
Mechanics-Based Iterative Finite-Element method; J.Mater.Sci. 30, 481–487, 1995.
Brown L.M., Stobbs W.M.: The Work-Hardening of Copper–Silica. I. A Model Based
on Internal Stresses, with no Plastic Relaxation; Phil.Mag. 23, 1185–1199, 1971.
Bruzzi M.S., McHugh P.E., O’Rourke F., Linder T.: Micromechanical Modelling of the
Static and Cyclic Loading of an Al21244-SiC MMC; Int.J.Plast. 17, 565–599, 2001.
Bulsara V.N., Talreja R., Qu J.: Damage Initiation under Transverse Loading of Unidi-
rectional Composites with Arbitrarily Distributed Fibers; Compos.Sci.Technol. 59, 673–
682, 1999.
Buryachenko V.A.: The Overall Elastoplastic Behavior of Multiphase Materials with
Isotropic Components; Acta Mech. 119, 93–117, 1996.
Chaboche Y.L., Kruch S., Pottier T.: Micromechanics versus Macromechanics: a Com-
bined Approach for Metal Matrix Composite Constitutive Modeling; Eur.J.Mech. A/Solids
17, 885–908, 1998.
Chang W., Kikuchi N.: Analysis of Residual Stresses Developed in the Curing of
Liquid Molding; in “Advanced Computational Methods for Material Modeling” (Eds.
D.J.Benson, R.A.Asaro), pp. 99–113; AMD–Vol.180/PVP–Vol.268, ASME, New York,
NY, 1993.
Chen H.R., Yang Q.S., Williams F.W.: A Self-Consistent Finite Element Approach to
the Inclusion Problem; Comput.Mater.Sci. 2, 301–307, 1994.
Chen T.: Exact Moduli and Bounds of Two-Phase Composites with Coupled Multifield
Linear Responses; J.Mech.Phys.Sol. 45, 385–398, 1997.
Chimani C.M., Mörwald K.: Micromechanical Investigation of the Hot Ductility Behavior
of Steel; ISIJ Int. 39, 1194–1197, 1999.
Chimani C.M., Böhm H.J., Rammerstorfer F.G.: On Stress Singularities at Free Edges
of Bimaterial Junctions — A Micromechanical Study; Scr.mater. 36, 943–947, 1997.
Christensen R.M.: Sufficient Symmetry Conditions for Isotropy of the Elastic Moduli
Tensor; J.Appl.Mech. 54, 772–777, 1987.
Christensen R.M., Lo K.H.: Solutions for Effective Shear Properties Three Phase Sphere
and Cylinder Models; J.Mech.Phys.Sol. 27, 315–330, 1979.
Christensen R.M., Lo K.H.: Erratum to Christensen and Lo, 1979; J.Mech.Phys.Sol. 34,
639, 1986.
Christensen R.M., Schantz H., Shapiro J.: On the Range of Validity of the Mori–Tanaka
Method; J.Mech.Phys.Sol. 40, 69–73, 1992.
Chung P.W., Tamma K.K., Namburu R.R.: Asymptotic Expansion Homogenization for
Heterogeneous Media: Computational Issues and Applications; Composites 32A, 1291–
1301, 2001.
– 104 –
Cleveringa H.H.M., van der Giessen E., Needleman A.: Comparison of Discrete Disloca-
tion and Continuum Plasticity Predictions for a Composite Material; Acta mater. 45,
3163–3179, 1997.
Clyne T.W., Withers P.J.: An Introduction to Metal Matrix Composites. Cambridge
University Press, Cambridge, UK, 1993.
Courage W.M.G., Schreurs P.J.G: Effective Material Parameters for Composites with
Randomly Oriented Short Fibers; Comput.Struct. 44, 1179–1185, 1992.
Cox H.L.: The Elasticity and Strength of Paper and Other Fibrous Materials;
Brit.J.Appl.Phys. 3, 72–79, 1952.
Corvasce F., Lipiński P., Berveiller M.: The Effects of Thermal, Plastic and Elastic
Stress Concentrations on the Overall Behavior of Metal Matrix Composites; in “Inelastic
Deformation of Composite Materials” (Ed. G.J.Dvorak), pp. 389–408; Springer–Verlag,
New York, NY, 1991.
Cruz M.E., Patera A.T.: A Parallel Monte-Carlo Finite Element Procedure for the Anal-
ysis of Multicomponent Random Media; Int.J.Num.Meth.Engng. 38, 1087–1121, 1995.
Dasgupta A., Agarwal R., Bhandarkar S.: Three-Dimensional Modeling of Woven-
Fabric Composites for Effective Thermo-Mechanical and Thermal Properties; Com-
pos.Sci.Technol. 56, 209–223, 1996.
Davis L.C.: Flow Rule for the Plastic Deformation of Particulate Metal Matrix Compos-
ites; Comput.Mater.Sci. 6, 310–318, 1996.
Davis L.C., Andres C., Allison J.E.: Microstructure and Strengthening of Metal Matrix
Composites; Mater.Sci.Engng. A249, 40–45, 1998.
Dawson P.R.: Computational Crystal Plasticity; Int.J.Sol.Struct. 37, 115–130, 2000.
Daxner T.: Multi-Scale Modeling and Simulation of Metallic Foams. Doctoral Thesis,
Vienna University of Technology, Vienna, Austria, 2002.
Daxner T., Böhm H.J., Seitzberger M., Rammerstorfer F.G.: Modeling of Cellular Ma-
terials; in “Handbook of Cellular Metals” (Eds.: H.P.Degischer, B.Kriszt), pp. 245–280;
Wiley–VCH, Weinheim, Germany, 2002.
Doghri I., Ouaar A.: Homogenization of Two-Phase Elasto-Plastic Composite Materials
and Structures; Int.J.Sol.Struct. 40, 1681–1712, 2003.
Dong M., Schmauder S.: Modeling of Metal Matrix Composites by a Self-Consistent
Embedded Cell Model; Acta mater. 44, 2465–2478, 1996.
Drabek T., Böhm H.J.: Simulation of Ductile Damage in Metal Matrix Composites; in
“Proceedings of NUMIFORM 2004” (Eds. S.Ghosh, J.M.Castro, J.K.Lee), pp. 1887–1892;
American Institute of Physics, Melville, NY, 2004.
Dreyer W., Müller W.H., Olschewski J.: An Approximate Analytical 2D-Solution for the
Stresses and Strains in Eigenstrained Cubic Materials; Acta Mech. 136, 171–192, 1999.
Drugan W.J.: Micromechanics-Based Variational Estimates for a Higher-Order Nonlocal
Constitutive Equation and Optimal Choice of Effective Moduli for Elastic Composites;
J.Mech.Phys.Sol. 48, 1359–1387, 2000.
Drugan W.J., Willis J.R.: A Micromechanics-Based Nonlocal Constitutive Equa-
tion and Estimates of Representative Volume Element Size for Elastic Composites;
J.Mech.Phys.Sol. 44, 497–524, 1996.
Dunn M.L., Ledbetter H.: Elastic–Plastic Behavior of Textured Short-Fiber Composites;
Acta mater. 45, 3327–3340, 1997.
Dunn M.L., Taya M.: Micromechanics Predictions of the Effective Electroelastic Moduli
of Piezoelectric Composites; Int.J.Sol.Struct. 30, 161–175, 1993.
– 105 –
Duschlbauer D.: Computational Simulation of the Thermal Conductivity of MMCs under
Consideration of the Inclusion–Matrix Interface. Doctoral Thesis, Vienna University of
Technology, Vienna, Austria, 2003.
Duschlbauer D., Pettermann H.E., Böhm H.J.: Mori–Tanaka Based Evaluation of Inclu-
sion Stresses in Composites with Nonaligned Reinforcements; Scr.mater. 48, 223–228,
2003.
Dvorak G.J.: Plasticity Theories for Fibrous Composite Materials; in “Metal Matrix
Composites: Mechanisms and Properties” (Eds. R.K.Everett, R.J.Arsenault), pp. 1–77;
Academic Press, Boston, MA, 1991.
Dvorak G.J.: Transformation Field Analysis of Inelastic Composite Materials;
Proc.Roy.Soc.London A437, 311–327, 1992.
Dvorak G.J., Bahei–el–Din Y.A.: A Bimodal Plasticity Theory of Fibrous Composite
Materials; Acta Mech. 69, 219–241, 1987.
Dvorak G.J., Bahei–el–Din Y.A., Wafa A.M.: Implementation of the Transformation
Field Analysis for Inelastic Composite Materials; Comput.Mech. 14, 201–228, 1994.
Eckschlager A.: Simulation of Particle Failure in Particle Reinforced Ductile Matrix
Composites. Doctoral Thesis, Vienna University of Technology, Vienna, Austria, 2002.
Eckschlager A., Böhm H.J., Han W.: Modeling of Brittle Particle Failure in Particle
Reinforced Ductile Matrix Composites by 3D Unit Cells; in “Proceedings of the 13th
International Conference on Composite Materials (ICCM–13)” (Ed. Y.Zhang), Paper
No. 1342; Scientific and Technical Documents Publishing House, Beijing, 2001 (CD–
ROM).
Eduljee R.F., McCullough R.L.: Elastic Properties of Composites; in “Materials Sci-
ence and Technology Vol.13: Structure and Properties of Composites” (Eds. R.W.Cahn,
P.Haasen, E.J.Kramer), pp. 381–474; VCH, Weinheim, FRG, 1993.
Elices M., Guinea G.V., Gomez J., Planas J.: The Cohesive Zone Model: Advantages,
Limitations and Challenges; Engng.Fract.Mech. 69, 137–163, 2002.
Eshelby J.D.: The Determination of the Elastic Field of an Ellipsoidal Inclusion and
Related Problems; Proc.Roy.Soc.London A241, 376–396, 1957.
Eshelby J.D.: The Elastic Field Outside an Ellipsoidal Inclusion; Proc.Roy.Soc.London
A252, 561–569, 1959.
Estevez R., Maire E., Franciosi P., Wilkinson D.S.: Effect of Particle Clustering on the
Strengthening versus Damage Rivalry in Particulate Reinforced Elastic Plastic Materials:
A 3-D Analysis from a Self-Consistent Modelling; Eur.J.Mech. A/Solids 18, 785–804,
1999
Estrin Y., Arndt S., Heilmaier M., Bréchet Y.: Deformation Behavior of Particle-
Strengthened Alloys: A Voronoi Mesh Approach; Acta mater. 47, 595–606, 1999.
Faleskog J., Shih C.F.: Micromechanics of Coalescence — I.: Synergistic Effects of
Elasticity, Plastic Yielding and Multi-Size-Scale Voids; J.Mech.Phys.Sol. 45, 21–50,
1997.
Ferrari M.: Asymmetry and the High Concentration Limit of the Mori–Tanaka Effective
Medium Theory; Mech.Mater. 11, 251–256, 1991.
Feyel F.: A Multilevel Finite Element Method (FE2 ) to Describe the Response of Highly
Non-Linear Strucutres Using Generalized Continua; Comput.Meth.Appl.Mech.Engng.
192, 3233–3244, 2003.
Feyel F., Chaboche J.L.: FE2 Multiscale Approach for Modelling the Elastoviscoplastic
Behaviour of Long Fiber SiC/Ti Composite Materials; Comput.Meth.Appl.Mech.Engng.
183, 309–330, 2000.
– 106 –
Finot M.A., Shen Y.L., Needleman A., Suresh S.: Micromechanical Modeling of Re-
inforcement Fracture in Particle-Reinforced Metal-Matrix Composites; Metall.Trans.A
25A, 2403–2420, 1994.
Fischer F.D., Kolednik O., Shan G.X., Rammerstorfer F.G.: A Note on Calibration of
Ductile Failure Damage Indicators; Int.J.Fract. 73, 345–357, 1995.
Fischmeister H., Karlsson B.: Plastizitätseigenschaften grob–zweiphasiger Werkstoffe;
Z.Metallk. 68, 311–327, 1977.
Fish J., Shek K.: Multiscale Analysis of Composite Materials and Structures; Com-
pos.Sci.Technol. 60, 2547–2556, 2000.
Fish J., Shephard M.S., Beall M.W.: Automated Multiscale Fracture Analysis; in “Dis-
cretization Methods in Structural Mechanics” (Eds. H.A.Mang, F.G.Rammerstorfer), pp.
249–256; Kluwer, Dordrecht, The Netherlands, 1999.
Fish J., Shek K., Pandheeradi M., Shephard M.S.: Computational Plasticity for Com-
posite Structures Based on Mathematical Homogenization: Theory and Practice; Com-
put.Meth.Appl.Mech.Engng. 148, 53–73, 1997.
Fond C., Riccardi A., Schirrer R., Montheillet F.: Mechanical Interaction between Spher-
ical Inhomogeneities: An Assessment of a Method Based on the Equivalent Inclusion;
Eur.J.Mech. A/Solids 20, 59–75, 2001.
Fu S.Y., Lauke B.: The Elastic Modulus of Misaligned Short-Fiber-Reinforced Polymers;
Compos.Sci.Technol. 58, 389–400, 1998.
Fukuda H., Chou T.W.: A Probabilistic Theory of the Strength of Short-Fibre Composites
with Variable Fibre Length and Orientation; J.Mater.Sci. 17, 1003–1007, 1982.
Fukuda H., Kawata K.: On Young’s Modulus of Short Fiber Composites; Fibre
Sci.Technol. 7, 207–222, 1974.
Gänser H.P.: Large Strain Behavior of Two-Phase Materials. Doctoral Thesis, Montan-
universität Leoben, Leoben, Austria, 1998.
Gänser H.P., Fischer F.D., Werner E.A.: Large Strain Behaviour of Two-Phase Materials
with Random Inclusions; Comput.Mater.Sci. 11, 221–226, 1998.
Gărăjeu M., Michel J.C., Suquet P.: A Micromechanical Approach of Damage in
Viscoplastic Materials by Evolution in Size, Shape and Distribution of Voids; Com-
put.Meth.Appl.Mech.Engng. 183, 223–246, 2000.
Gavazzi A.C., Lagoudas D.C.: On the Numerical Evaluation of Eshelby’s Tensor and its
Application to Elastoplastic Fibrous Composites; Comput.Mech. 7, 12–19, 1990.
Geers M.G.D., Kouznetsova V., Brekelmans W.A.M.: Micro–Macro Scale Transitions
for Engineering Materials in the Presence of Size Effects; in “Trends in Computational
Structural Mechanics” (Eds. W.A.Wall, K.U.Bletzinger, K.Schweizerhof), pp. 118–127;
CIMNE, Barcelona, Spain, 2001.
Geers M.G.D., Engelen R.A.B., Ubachs R.J.M.: On the Numerical Modelling of Ductile
Damage with an Implicit Gradient-Enhanced Formulation; Rev.Eur.Elem.Fin. 10, 173–
191, 2001.
Geni M., Kikuchi M.: Void Configuration under Constrained Deformation in Ductile
Matrix Materials; Comput.Mater.Sci. 16, 391–403, 1999.
Ghosh S., Moorthy S.: Particle Fracture Simulation in Non-Uniform Microstructures of
Metal-Matrix Composites; Acta mater. 46, 965–982, 1998.
Ghosh S., Lee K.H., Moorthy S.: Two Scale Analysis of Heterogeneous Elastic-Plastic
Materials with Asymptotic Homogenization and Voronoi Cell Finite Element Model;
Comput.Meth.Appl.Mech.Engng. 132, 63–116, 1996.
– 107 –
Ghosh S., Lee K.H., Raghavan P.: A Multi-Level Computational Model for Multi-Scale
Analysis in Composite and Porous Materials; Int.J.Sol.Struct. 38, 2335–2385, 2001.
Ghosh S., Nowak Z., Lee K.H.: Quantitative Characterization and Modeling of Composite
Microstructures by Voronoi Cells; Acta mater. 45, 2215–2234, 1997.
Gibson L.J.: The Mechanical Behavior of Cancellous Bone; J.Biomech. 18, 317–328,
1985.
Gibson L.J., Ashby M.F.: Cellular Solids: Structure and Properties. Pergamon Press,
Oxford, UK, 1988.
Gilormini P.: Insuffisance de l’extension classique du modèle auto-cohérent au comporte-
ment non-linéaire; C.R.Acad.Sci.Paris, série IIb 320, 115–122, 1995.
Gologanu M., Leblond J.P., Perrin G., Devaux J.: Recent Extensions of Gurson’s Model
for Porous Ductile Materials; in “Continuum Micromechanics” (Ed. P.Suquet), pp. 61–
130; Springer–Verlag, Vienna, Austria, 1997.
Gommers B., Verpoest I., van Houtte P.: The Mori–Tanaka Method Applied to Textile
Composite Materials; Acta mater. 46, 2223–2235, 1998.
González C., LLorca J.: Prediction of the Tensile Stress–Strain Curve and Ductility in
SiC/Al Composites; Scr.mater. 35, 91–97, 1996.
González C., LLorca J.: A Self-Consistent Approach to the Elasto-Plastic Behavior of
Two-Phase Materials Including Damage; J.Mech.Phys.Sol. 48, 675–692, 2000.
Grenestedt J.L.: Influence of Wavy Imperfections in Cell Walls on Elastic Stiffness of
Cellular Solids; J.Mech.Phys.Sol. 46, 29–50, 1998.
Guidoum A.: Simulation numérique 3D des comportements des bétons en tant que com-
posites granulaires. Doctoral Thesis, Ecole Polytécnique Fédérale de Lausanne, Lau-
sanne, Switzerland, 1994.
Guldberg R.E., Hollister S.J., Charras G.T.: The Accuracy of Digital Image-Based Finite
Element Models; J.Biomech.Engng. 120, 289–295, 1998.
Gunawardena S.R., Jansson S., Leckie F.A.: Transverse Ductility of Metal Matrix
Composites; in “Failure Mechanisms in High Temperature Composite Materials” (Eds.
K.Haritos, G.Newaz, S.Mall), pp. 23–30; AD–Vol.22/AMD–Vol.122, ASME, New York,
NY, 1991.
Gunawardena S.R., Jansson S., Leckie F.A.: Modeling of Anisotropic Behavior of Weakly
Bonded Fiber Reinforced MMC’s; Acta metall.mater. 41, 3147–3156, 1993.
Guo G., Fitoussi J., Baptiste D., Sicot N., Wolff C.: Modelling of Damage Behavior of
a Short-Fiber Reinforced Composite Structure by the Finite Element Analysis Using a
Micro–Macro Law; Int.J.Dam.Mech. 6, 278–299, 1997.
Gurson A.L.: Continuum Theory of Ductile Rupture by Void Nucleation and
Growth: Part I — Yield Criteria and Flow Rules for Porous Ductile Media;
J.Engng.Mater.Technol. 99, 2–15, 1977.
Gusev A.A.: Representative Volume Element Size for Elastic Composites: A Numerical
Study; J.Mech.Phys.Sol. 45, 1449–1459, 1997.
Gusev A.A.: Numerical Identification of the Potential of Whisker- and Platelet-Filled
Polymers; Macromolecules 34, 3081–3093, 2001.
Gusev A.A., Hine P.J., Ward I.M.: Fiber Packing and Elastic Properties of a Trans-
versely Random Unidirectional Glass/Epoxy Composite; Compos.Sci.Technol. 60, 535–
541, 2000.
Haddadi H., Teodosiu C.: 3D-Analysis of the Effect of Interfacial Debonding on the
Plastic Behaviour of Two-Phase Composites; Comput.Mater.Sci. 16, 315–322, 1999.
– 108 –
Halpin J.C., Kardos J.L.: The Halpin–Tsai Equations: A Review; Polym.Engng.Sci. 16,
344–351, 1976.
Hancock J.W., Mackenzie A.C.: On the Mechanisms of Ductile Failure in High-Strength
Steels Subjected to Multi-Axial Stress-States; J.Mech.Phys.Sol. 24, 147–169, 1976.
Hashin Z.: The Elastic Moduli of Heterogeneous Materials; J.Appl.Mech. 29, 143–150,
1962.
Hashin Z.: Analysis of Composite Materials — A Survey; J.Appl.Mech. 50, 481–505,
1983.
Hashin Z.: The Differential Scheme and its Application to Cracked Materials;
J.Mech.Phys.Sol. 36, 719–733, 1988.
Hashin Z., Rosen B.W.: The Elastic Moduli of Fiber-Reinforced Materials; J.Appl.Mech.
31, 223–232, 1964.
Hashin Z., Shtrikman S.: On Some Variational Principles in Anisotropic and Nonhomo-
geneous Elasticity; J.Mech.Phys.Sol. 10, 335–342, 1962.
Hashin Z., Shtrikman S.: A Variational Approach to the Theory of the Elastic Behavior
of Multiphase Materials; J.Mech.Phys.Sol. 11, 127–140, 1963.
Hassani B., Hinton E.: Homogenization and Structural Topology Optimization. Springer–
Verlag, London, UK, 1999.
Hatta H., Taya M.: Equivalent Inclusion Method for Steady State Heat Conduction in
Composites; Int.J.Engng.Sci. 24, 1159–1172, 1986.
Heness G.L., Ben-Nissan B., Gan L.H., Mai Y.W.: Development of a Finite Element
Micromodel for Metal Matrix Composites; Comput.Mater.Sci. 13, 259–269, 1999.
Hervé E., Zaoui A.: n-Layered Inclusion-Based Micromechanical Modelling; Int.J.Engng.
Sci. 31, 1–10, 1993.
Hervé E., Stolz C., Zaoui A.: A propos de’l assemblage des sphères composites de Hashin;
C.R.Acad.Sci.Paris, série II 313, 857–862, 1991.
Hill R.: The Elastic Behaviour of a Crystalline Aggregate. Proc.Phys.Soc. A65, 349–354,
1952.
Hill R.: Elastic Properties of Reinforced Solids: Some Theoretical Principles;
J.Mech.Phys.Sol. 11, 357–372, 1963.
Hill R.: Theory of Mechanical Properties of Fibre-Strengthened Materials: I. Elastic
Behaviour; J.Mech.Phys.Sol. 12, 199–212, 1964.
Hill R.: A Self Consistent Mechanics of Composite Materials; J.Mech.Phys.Sol. 13,
213–222, 1965.
Hill R.: The Essential Structure of Constitutive Laws for Metal Composites and Poly-
crystals; J.Mech.Phys.Sol. 15, 79–95, 1967.
Hollister S.J., Brennan J.M., Kikuchi N.: A Homogenization Sampling Procedure for
Calculating Trabecular Bone Effective Stiffness and Tissue Level Stress; J.Biomech. 27,
433–444, 1994.
Hollister S.J., Fyhrie D.P., Jepsen K.J., Goldstein S.A.: Application of Homogenization
Theory to the Study of Trabecular Bone Mechanics; J.Biomech. 24, 825–839, 1991.
Holmberg S., Persson K., Petersson H.: Nonlinear Mechanical Behaviour and Analysis
of Wood and Fibre Materials; Comput.Struct. 72, 459–480, 1999.
Hom C.L., McMeeking R.M.: Plastic Flow in Ductile Materials Containing a Cubic
Array of Rigid Spheres; Int.J.Plast. 7, 255–274, 1991.
– 109 –
Hu C., Moorthy S., Ghosh S.: A Voronoi Cell Finite Element for Ductile Damage in
MMCs; in “Proceedings of NUMIFORM 2004” (Eds. S.Ghosh, J.M.Castro, J.K.Lee), pp.
1893–1898; American Institute of Physics, Melville, NY, 2004.
Hu G.K.: A Method of Plasticity for General Aligned Spheroidal Void or Fiber-Reinforced
Composites; Int.J.Plast. 12, 439–449, 1996.
Hu G.K., Weng G.J.: The Connections between the Double-Inclusion Model and the
Ponte Castañeda–Willis, Mori–Tanaka, and Kuster–Toksoz Models; Mech.Mater. 32,
495–503, 2000.
Hu G.K., Weng G.J.: Some Reflections on the Mori–Tanaka and Ponte Castañeda–Willis
Methods with Randomly Oriented Ellipsoidal Inclusions; Acta Mech. 140, 31–40, 2000.
Hu G.K., Guo G., Baptiste D.: A Micromechanical Model of Influence of Particle Frac-
ture and Particle Cluster on Mechanical Properties of Metal Matrix Composites; Com-
put.Mater.Sci. 9, 420–430, 1998.
Huang J.H.: Some Closed-Form Solutions for Effective Moduli of Composites Containing
Randomly Oriented Short Fibers; Mater.Sci.Engng. A315, 11–20, 2001.
Huang Y., Hu K.X.: A Generalized Self-Consistent Mechanics Method for Solids Con-
taining Elliptical Inclusions; J.Appl.Mech. 62, 566–572, 1995.
Huet C.: Application of Variational Concepts to Sizes Effects in Elastic Heterogeneous
Bodies; J.Mech.Phys.Sol. 38, 813–841, 1990.
Huet C.: Coupled Size and Boundary-Condition Effects in Viscoelastic Heterogeneous
and Composite Bodies; Mech.Mater. 31, 787–829, 1999.
Huet C., Navi P., Roelfstra P.E.: A Homogenization Technique Based on Hill’s Modifi-
cation Theorem; in “Continuum Models and Discrete Systems” (Ed. G.A.Maugin), pp.
135–143; Longman, Harlow, UK, 1991.
Hutchinson J.W.: Elastic-Plastic Behavior of Polycrystalline Metals and Composites;
Proc.Roy.Soc.London A319, 247–272, 1970.
Hutchinson J.W.: Plasticity on the Micron Scale; Int.J.Sol.Struct. 37, 225–238, 2000.
Ibrahimbegović A., Markovič D.: Strong Coupling Methods in Multi-Phase and
Multi-Scale Modeling of Inelastic Behavior of Heterogeneous Structures; Com-
put.Meth.Appl.Mech.Engng. 192, 3089–3107, 2003.
Ingber M.S., Papathanasiou T.D.: A Parallel-Supercomputing Investigation of the Stiff-
ness of Aligned Short-Fiber-Reinforced Composites Using the Boundary Element Method;
Int.J.Num.Meth..Engng. 40, 3477–3491, 1997.
Ishikawa N., Parks D., Socrate S., Kurihara M.: Micromechanical Modeling of Ferrite–
Pearlite Steels Using Finite Element Unit Cell Models; ISIJ Int. 40, 1170–1179, 2000.
Ismar H., Reinert U.: Modelling and Simulation of the Macromechanical Nonlinear Be-
havior of Fibre-Reinforced Ceramics on the Basis of a Micromechanical–Statistical Ma-
terial Description; Acta Mech. 120, 47–60, 1997.
Ismar H., Schmitt U.: Simulation von Einspiel- und Rißbildungsvorgängen in Faserver-
bundwerkstoffen mit metallischer Matrix; Arch.Appl.Mech. 61, 18–29, 1991.
Ismar H., Schröter F.: Three-Dimensional Finite Element Analysis of the Mechanical
Behavior of Cross Ply-Reinforced Aluminum; Mech.Mater. 32, 329–338, 2000.
Iung T., Grange M.: Mechanical Behavior of Two-Phase Materials Investigated by the
Finite Element Method: Necessity of Three-Dimensional Modeling; Mater.Sci.Engng.
A201, L8–L11, 1995.
Iyer S.K., Lissenden C.J., Arnold S.M.: Local and Overall Flow in Composites Predicted
by Micromechanics; Composites 31B, 327–343, 2000.
– 110 –
Jansson S.: Homogenized Nonlinear Constitutive Properties and Local Stress Concentra-
tions for Composites with Periodic Internal Structure; Int.J.Sol.Struct. 29, 2181–2200,
1992.
Järvstråt N.: Homogenization and the Mechanical Behavior of Metal Composites; in
“Residual Stresses – III” (Eds. H.Fujiwara, T.Abe, K.Tanaka), Vol.1, pp. 70–75; Elsevier,
London, UK, 1992.
Järvstråt N.: An Ellipsoidal Unit Cell for the Calculation of Micro-Stresses in Short
Fibre Composites; Comput.Mater.Sci. 1, 203–212;
Jayaraman K., Kortschot M.T.: Correction to the Fukuda–Kawata Young’s Modulus The-
ory and the Fukuda–Chou Strength Theory for Short Fibre-Reinforced Composite Mate-
rials; J.Mater.Sci. 31, 2059–2064, 1996.
Jiang M., Ostoja–Starzewski M., Jasiuk I.: Scale-Dependent Bounds on Effective Elasto-
plastic Response of Random Composites; J.Mech.Phys.Sol. 49, 655–673, 2001.
Jiang M., Ostoja–Starzewski M., Jasiuk I.: Apparent Elastic and Elastoplastic Behavior
of Periodic Composites; Int.J.Sol.Struct. 39, 199–212, 2002.
Jirásek M.: Comparative Study on Finite Elements with Embedded Discontinuities; Com-
put.Meth.Appl.Mech.Engng. 188, 307–330, 2000.
Jirásek M., Rolshoven S.: Comparison of Integral-Type Nonlocal Plasticity Models for
Strain-Softening Materials; Int.J.Engng.Sci. 41, 1553–1602, 2003.
Johannesson B., Pedersen O.B.: Analytical Determination of the Average Eshelby Tensor
for Transversely Isotropic Fiber Orientation Distributions; Acta mater. 46(9), 3165–
3173, 1998.
Ju J.W., Sun L.Z.: A Novel Formulation for the Exterior Point Eshelby’s Tensor of an
Ellipsoidal Inclusion; J.Appl.Mech. 66, 570–574, 1999.
Ju J.W., Sun L.Z.: Effective Elastoplastic Behavior of Metal Matrix Composites Con-
taining Randomly Located Aligned Spheroidal Inhomogeneities. Part I: Micromechanics-
Based Formulation; Int.J.Sol.Struct. 38, 183–201, 2001.
Kailasam M., Ponte Castañeda P.: A General Constitutive Theory for a Linear and
Nonlinear Particulate Media with Microstructure Evolution; J.Mech.Phys.Sol. 46, 427–
465, 1998.
Kailasam M., Aravas N., Ponte Castañeda P.: Porous Metals with Developing Anisotropy:
Constitutive Models, Computational Issues and Applications to Deformation Processing;
Comput.Model.Engng.Sci. 1, 105–118, 2000.
Kanit T., Forest S., Gallier I., Mounoury V., Jeulin D.: Determination of the Size of
the Representative Volume Element for Random Composites: Statistical and Numerical
Approach; Int.J.Sol.Struct. 40, 3647–3679, 2003.
Karayaka M., Sehitoglu H.: Thermomechanical Deformation Modeling of Al2xxx-
T4/SiCp Composites; Acta metall.mater. 41, 175–189, 1993.
Kouznetsova V.G., Geers M.G.D., Brekelmans W.A.M.: Multi-Scale Constitutive Mod-
eling of Heterogeneous Materials with a Gradient-Enhanced Computational Homogeniza-
tion Scheme; Int.J.Num.Meth.Engng. 54, 1235–1260, 2002.
Krakovsky I., Myroshnychenko V.: Modelling Dielectric Properties of Composites by
Finite Element Method; J.Appl.Phys. 92, 6743–6748, 2002.
Kreher W.S.: Residual Stresses and Stored Elastic Energy of Composites and Polycrys-
tals; J.Mech.Phys.Sol. 38, 115–128, 1990.
Kuster G.T., Toksőz M.N.: Velocity and Attenuation of Seismic Waves in Two-Phase
Media: I. Theoretical Formulation; Geophysics 39, 587–606, 1974.
– 111 –
Lagoudas D.C., Gavazzi A.C., Nigam H.: Elastoplastic Behavior of Metal Matrix Com-
posites Based on Incremental Plasticity and the Mori–Tanaka Averaging Scheme; Com-
put.Mech. 8, 193–203, 1991.
Lee B.J., Mear M.E.: Stress Concentration Induced by an Elastic Spheroidal Particle in
a Plastically Deforming Solid; J.Mech.Phys.Sol. 47, 1301–1336, 1999.
Lee H.K., Simunovic S.: Modeling of Progressive Damage in Aligned and Randomly Ori-
ented Discontinuous Fiber Polymer Matrix Composites; Composites 31B, 77–86, 2000.
Lee K.H., Ghosh S.: Small Deformation Multi-Scale Analysis of Heterogeneous Mate-
rials with the Voronoi Cell Finite Element Model and Homogenization Theory; Com-
put.Mater.Sci. 7, 131–146, 1996.
Lerch B.A., Melis M.E., Tong M.: Experimental and Analytical Analysis of the Stress–
Strain Behavior in a [90/0]2S SiC/Ti-15-3 Laminate. NASA Technical Memorandum
NASA–TM–104470, 1991.
Levin V.M.: On the Coefficients of Thermal Expansion of Heterogeneous Materials;
Mech.Sol. 2, 58–61, 1967.
Levy A., Papazian J.M.: Elastoplastic Finite Element Analysis of Short-Fiber-Reinforced
SiC/Al Composites: Effects of Thermal Treatment; Acta metall.mater. 39, 2255–2266,
1991.
Li D.S., Wisnom M.R.: Unidirectional Tensile Stress–Strain Response of BP-SiC Fiber
Reinforced Ti-6Al-4V; J.Compos.Technol.Res. 16, 225–233, 1994.
Li M., Ghosh S., Richmond O., Weiland H., Rouns T.N.: Three Dimensional Character-
ization and Modeling of Particle Reinforced Metal Matrix Composites, Part II: Damage
Characterization; Mater.Sci.Engng. A266, 221–240, 1999.
Liebowitz H., Sandhu J.S., Lee J.D., Menandro F.C.M.: Computational Fracture Me-
chanics: Research and Application; Engng.Fract.Mech. 50, 653–670, 1995.
LLorca J.: A Numerical Analysis of the Damage Mechanisms in Metal-Matrix Composites
under Cyclic Deformation; Comput.Mater.Sci. 7, 118–122, 1996.
Lubarda V.A., Markenscoff X.: On the Absence of Eshelby Property for Non-Ellipsoidal
Inclusions; Int.J.Sol.Struct. 35, 3405–3411, 1998.
Lusti H.R., Hine P.J., Gusev A.A.: Direct Numerical Predictions for the Elastic and
Thermoelastic Properties of Short Fibre Composites; Compos.Sci.Technol. 62, 1927–
1934, 2002.
Mahishi J.M.: An Integrated Micromechanical and Macromechanical Approach to Frac-
ture Behavior of Fiber-Reinforced Composites; Engng.Fract.Mech. 25, 197–228, 1986.
Mahishi J.M., Adams D.F.: Micromechanical Predictions of Crack Initiation, Propaga-
tion and Crack Growth Resistance in Boron/Aluminum Composites; J.Compos.Mater.
16, 457–469, 1982.
Maire E., Wattebled F., Buffière J.Y., Peix G.: Deformation of a Metallic Foam Studied
by X-Ray Computed Tomography and Finite Element Calculations; in “Metal Matrix
Composites and Metallic Foams” (Eds. T.W.Clyne, F.Simancik), pp. 68–73; Wiley–VCH,
Weinheim, Germany, 2000.
Marketz F., Fischer F.D.: Micromechanical Modelling of Stress-Assisted Martensitic
Transformation; Modell.Simul.Mater.Sci.Engng. 2, 1017–1046, 1994.
Markov K.: Elementary Micromechanics of Heterogeneous Media; in “Heterogeneous Me-
dia: Micromechanics Modeling Methods and Simulations” (Eds. K.Markov, L.Preziosi),
pp. 1–162; Birkhäuser, Boston, MA, 2000.
– 112 –
Markov K., Preziosi L. (Eds.): Heterogeneous Media: Micromechanics Modeling Methods
and Simulations. Birkhäuser, Boston, MA, 2000.
Masson R., Bornert M., Suquet P., Zaoui A.: An Affine Formulation for the Prediction
of the Effective Properties of Nonlinear Composites and Polycrystals; J.Mech.Phys.Sol.
48, 1203–1227, 2000.
Mawsouf N.M.: A Micromechanical Mechanism of Fracture Initiation in Discontinuously
Reinforced Metal Matrix Composite; Mater.Charact. 44, 321–327, 2000.
McDowell D.L.: Modeling and Experiments in Plasticity; Int.J.Sol.Struct. 37, 293–310,
2000.
McHugh P.E., Connolly P.: Modelling the Thermo-Mechanical Behavior of an Al Alloy-
SiCp Composite. Effects of Particle Shape and Microscale Failure; Comput.Mater.Sci.
3, 199–206, 1994.
McHugh P.E., Mohrmann R.: Modelling of Creep in a Ni Base Superalloy using a Single
Crystal Plasticity Model; Comput.Mater.Sci. 9, 134–140, 1997.
McHugh P.E., Asaro R.J., Shih C.F.: Computational Modeling of Metal-Matrix Compos-
ite Materials, Parts I to IV; Acta metall.mater. 41, 1461–1510, 1993.
McLaughlin R.: A Study of the Differential Scheme for Composite Materials;
Int.J.Engng.Sci. 15, 237–244, 1977.
McMeeking R., Hom C.L.: Finite Element Analysis of Void Growth in Elastic-Plastic
Materials; Int.J.Fract. 42, 1–19, 1990.
Michel J.C.: Théorie des modules effectifs. Approximations de Voigt et de Reuss; in
“Homogénéisation en mécanique des materiaux 1. Matériaux aléatoires élastiques et
milieux périodiques” (Eds. M.Bornert, T.Bretheau, P.Gilormini), pp. 41–56; Editions
Hermès, Paris, 2001.
Michel J.C., Suquet P.: An Analytical and Numerical Study of the Overall Behaviour of
Metal-Matrix Composites; Modell.Simul.Mater.Sci.Engng. 2, 637–658, 1994.
Michel J.C., Moulinec H., Suquet P.: Effective Properties of Composite Materials with
Periodic Microstructure: A Computational Approach; Comput.Meth.Appl.Mech.Engng.
172, 109–143, 1999.
Michel J.C., Moulinec H., Suquet P.: Composites à microstructure périodique; in “Ho-
mogénéisation en mécanique des materiaux 1. Matériaux aléatoires élastiques et milieux
périodiques” (Eds. M.Bornert, T.Bretheau, P.Gilormini), pp. 57–94; Editions Hermès,
Paris, 2001.
Miehe C., Schröder J., Schotte J.: Computational Homogenization Analysis in Fi-
nite Plasticity. Simulation of Texture Development in Polycrystalline Materials; Com-
put.Meth.Appl.Mech.Engng. 171, 387–418, 1999.
Miller C.A., Torquato S.: Effective Conductivity of Hard Sphere Suspensions;
J.Appl.Phys. 68, 5486–5493, 1990.
Miloh T., Benveniste Y.: A Generalized Self-Consistent Method for the Effective Conduc-
tivity of Composites with Ellipsoidal Inclusions and Cracked Bodies; J.Appl.Phys. 63,
789–7796, 1988.
Milton G.W.: Bounds on the Electromagnetic, Elastic, and Other Properties of Two-
Component Composites; Phys.Rev.Lett. 46, 542–545, 1981.
Milton G.W.: The Theory of Composites. Cambridge University Press, Cambridge, UK,
2002.
Mlekusch B.: Thermoelastic Properties of Short-Fibre-Reinforced Thermoplastics; Com-
pos.Sci.Technol. 59, 911–923, 1999.
– 113 –
Moës N., Cloirec M., Cartraud P., Remacle J.F.: A Computational Approach to Handle
Complex Microstructure Geometries; Comput.Meth.Appl.Mech.Engng. 192, 3163–3177,
2003.
Molinari A., Canova G.R., Ahzi S.: A Self-Consistent Approach for Large Deformation
Viscoplasticity; Acta metall. 35, 2983–2984, 1987.
Monaghan J., Brazil D.: Modeling the Sub-Subsurface Damage Associated with the Ma-
chining of a Particle Reinforced MMC; Comput.Mater.Sci. 9, 99–107, 1997.
Mori T., Tanaka K.: Average Stress in the Matrix and Average Elastic Energy of Mate-
rials with Misfitting Inclusions; Acta metall. 21, 571–574, 1973.
Moschovidis Z.A., Mura T.: Two Ellipsoidal Inhomogeneities by the Equivalent Inclusion
Method; J.Appl.Mech. 42, 847–852, 1975.
Moshev V.V., Kozhevnikova L.L.: Structural Cell of Particulate Elastomeric Composites
under Extension and Compression; Int.J.Sol.Struct. 39, 449–465, 2002.
Moulinec H., Suquet P.: A Fast Numerical Method for Computing the Linear and Non-
linear Mechanical Properties of Composites; C.R.Acad.Sci.Paris, série II 318, 1417–1423,
1994.
Moulinec H., Suquet P.: A Numerical Method for Computing the Overall Response of
Nonlinear Composites with Complex Microstructure. Comput.Meth.Appl.Mech.Engng.
157, 69–94, 1997.
Müller W.H.: Mathematical versus Experimental Stress Analysis of Inhomogeneities in
Solids; J.Phys.IV 6, C1-139–C1-148, 1996.
Mummery P., Derby B., Scruby C.B.: Acoustic Emission from Particulate-Reinforced
Metal Matrix Composites; Act metall.mater. 41, 1431–1445, 1993.
Mura T.: Micromechanics of Defects in Solids. Martinus Nijhoff, Dordrecht, The Nether-
lands, 1987.
Nakamura T., Suresh S.: Effects of Thermal Residual Stresses and Fiber Packing on
Deformation of Metal–Matrix Composites; Acta metall.mater. 41, 1665–1681, 1993.
Narasimhan R.: A Numerical Study of Fracture Initiation in a Ductile Population of 2nd
Phase Particles. I. Static Loading; Engng.Fract.Mech. 47, 919–934, 1994.
Needleman A.: A Continuum Model for Void Nucleation by Inclusion Debonding;
J.Appl.Mech. 54, 525–531, 1987.
Needleman A.: Material Rate Dependence and Mesh Sensitivity in Localization Problems;
Comput.Meth.Appl.Mech.Engng. 67, 68–85, 1987.
Needleman A., Nutt S.R., Suresh S., Tvergaard V.: Matrix, Reinforcement, and
Interfacial Failure: in “Fundamentals of Metal Matrix Composites” (Eds. S.Suresh,
A.Mortensen, A.Needleman), pp. 233–250; Butterworth–Heinemann, Boston, MA, 1993.
Nemat–Nasser S.: Averaging Theorems in Finite Deformation Plasticity;, Mech. mater.
31, 493–523, 1999.
Nemat–Nasser S., Hori M.: Micromechanics: Overall Properties of Heterogeneous Solids.
North–Holland, Amsterdam, The Netherlands, 1993.
Niebur G.L., Yuen J.C., Hsia A.C., Keaveny T.M. Convergence Behavior of High-
Resolution Finite Element Models of Trabecular Bone; J.Biomech.Engng. 121, 629–635,
1999.
Niordson C.F., Tvergaard V.: Nonlocal Plasticity Effects on Fibre Debonding in a
Whisker-Reinforced Metal; Eur.J.Mech. A/Solids 21, 239–248, 2002.
Norris A.N.: A Differential Scheme for the Effective Moduli of Composites; Mech.Mater.
4, 1–16, 1985.
– 114 –
Nutt S.R., Needleman A.: Void Nucleation at Fiber Ends in Al–SiC Composites;
Scr.metall. 21, 705–710, 1987.
Nye J.F.: Physical Properties of Crystals, Their Representation by Tensors and Matrices.
Clarendon, Oxford, UK, 1957.
Oden T.J., Zohdi T.I.: Analysis and Adaptive Modeling of Highly Heterogeneous Elastic
Structures; Comput.Meth.Appl.Mech.Engng. 148, 367–391, 1997.
Onck P., van der Giessen E.: Microstructurally-Based Modelling of Intergranular Creep
Fracture Using Grain Elements; Mech.Mater. 26, 109–126, 1997.
Ostoja–Starzewski M.: Random Field Models of Heterogeneous Materials; Int.J.Sol.Struct.
35, 2429–2455, 1998.
Ostoja–Starzewski M.: Lattice Models in Micromechanics; Appl.Mech.Rev. 55, 35–60,
2002.
Pahr D.H., Arnold S.M.: The Applicability of the Generalized Method of Cells for Ana-
lyzing Discontinuously Reinforced Composites. Composites 33B, 153–170, 2002.
Pandorf R.: Ein Beitrag zur FE-Simulation des Kriechens partikelverstärkter metallis-
cher Werkstoffe. VDI–Verlag (Reihe 5, Nr.585), Düsseldorf, FRG, 2000.
Papka S.D., Kyriakides S.: In-Plane Compressive Response and Crushing of Honeycomb;
J.Mech.Phys.Sol. 42, 1499–1532, 1994.
Pardoen T., Hutchinson J.W.: An Extended Model for Void Growth Coalescence;
J.Mech.Phys.Sol. 48, 2467–2512, 2000.
Pedersen O.B.: Thermoelasticity and Plasticity of Composites — I. Mean Field Theory;
Acta metall. 31, 1795–1808, 1983.
Pedersen O.B., Withers P.J.: Iterative Estimates of Internal Stresses in Short-Fibre Metal
Matrix Composites; Phil.Mag. A65, 1217–1233, 1992.
Peters P.W.M.: Modelling of Matrix Cracking in Fiber Reinforced Ceramics; Key En-
gng.Mater. 127–131, 1093–1100, 1997.
Pettermann H.E.: Derivation and Finite Element Implementation of Constitutive Mate-
rial Laws for Multiphase Composites Based on Mori–Tanaka Approaches. VDI–Verlag
(Reihe 18, Nr.217), Düsseldorf, Germany, 1997.
Pettermann H.E., Suresh S.: A Comprehensive Unit Cell Model — A Study of Coupled
Effects in Piezoelectric 1–3 Composites; Int.J.Sol.Struct. 37, 5447–5464, 2000.
Pettermann H.E., Böhm H.J., Rammerstorfer F.G.: Some Direction Dependent Proper-
ties of Matrix–Inclusion Type Composites with Given Reinforcement Orientation Distri-
butions; Composites 28B, 253–265, 1997.
Phan–Thien N., Milton G.W.: New Third-Order Bounds on the Effective Moduli of N -
Phase Composites; Quart.Appl.Math. 41, 59–74, 1983.
Plankensteiner A.F.: Multiscale Treatment of Heterogeneous Nonlinear Solids and Struc-
tures. VDI–Verlag (Reihe 18, Nr.248), Düsseldorf, Germany, 2000.
Plankensteiner A.F., Böhm H.J., Rammerstorfer F.G., Pettermann H.E.: Multiscale
Modeling of Highly Heterogeneous Particulate MMCs; J.Physique IV 8, Pr-301–Pr-308,
1998.
Plankensteiner A.F., Böhm H.J., Rammerstorfer F.G., Buryachenko V.A., Hackl G.:
Modeling of Layer-Structured High Speed Steel; Acta mater. 45, 1875–1887, 1997.
Ponte Castañeda P.: New Variational Principles in Plasticity and their Application to
Composite Materials; J.Mech.Phys.Sol. 40, 1757–1788, 1992.
– 115 –
Ponte Castañeda P.: A Second-Order Theory for Non-Linear Composite Materials; C.R.
Acad.Sci.Paris, série IIb 322, 3–10, 1996.
Ponte Castañeda P., Suquet P.: Nonlinear Composites; in “Advances in Applied Mechan-
ics 34” (Eds. E.van der Giessen, T.Y.Wu), pp. 171–302; Academic Press, New York, NY,
1998.
Ponte Castañeda P., Willis J.R.: The Effect of Spatial Distribution on the Effective
Behavior of Composite Materials and Cracked Media; J.Mech.Phys.Sol. 43, 1919–1951,
1995.
Povirk G.L., Nutt S.R., Needleman A.: Analysis of Creep in Thermally Cycled Al/SiC
Composites; Scr.metall.mater. 26, 461–66, 1992.
Pyrz R., Bochenek B.: Topological Disorder of Microstructure and its Relation to the
Stress Field; Int.J.Sol.Struct. 35, 2413–2427, 1998.
Qiu Y.P., Weng G.J.: A Theory of Plasticity for Porous Materials and Particle-
Reinforced Composites; J.Appl.Mech. 59, 261–268, 1992.
Quilici S., Cailletaud G.: FE Simulation of Macro-, Meso- and Microscales in Polycrys-
talline Plasticity; Comput.Mater.Sci. 16, 383–390, 1999.
Raabe D.: Computational Materials Science. Wiley–VCH, Weinheim, Germany, 1998.
Rammerstorfer F.G., Böhm H.J.: Finite Element Methods in Micromechanics of Com-
posites and Foam Materials; in “Computational Mechanics for the Twenty First Century”
(Ed. B.H.V.Topping), pp. 145–164; Saxe–Coburg Publications, Edinburgh, UK, 2000.
Reiter T.J., Dvorak G.J., Tvergaard V.: Micromechanical Models for Graded Composite
Materials; J.Mech.Phys.Sol. 45, 1281–1302, 1997.
Reuss A.: Berechnung der Fließgrenze von Mischkristallen auf Grund der Plastizitätsbe-
dingung für Einkristalle; ZAMM 9, 49–58, 1929.
Rice J.R., Tracey D.M.: On the Ductile Enlargement of Voids in Triaxial Stress Fields;
J.Mech.Phys.Sol. 17, 201–217, 1969.
Rintoul M., Torquato S.: Reconstruction of the Structure of Dispersions; J.Colloid In-
terf.Sci. 186, 467–476, 1997.
Roberts A.P., Garboczi E.J.: Elastic Properties of a Tungsten–Silver Composite by Re-
construction and Computation; J.Mech.Phys.Sol. 47, 2029–2055, 1999.
Roberts A.P., Garboczi E.J.: Elastic Moduli of Model Random Three-Dimensional
Closed-Cell Cellular Solids; Acta mater. 49, 189–197, 2001.
Rodin G.J.: The Overall Elastic Response of Materials Containing Spherical Inhomo-
geneities; Int.J.Sol.Struct. 30, 1849–1863, 1993.
Rousselier G.: Ductile Fracture Models and their Potential in Local Approach of Fracture;
Nucl.Engng.Design 105, 97–111, 1987.
Sab K.: On the Homogenization and the Simulation of Random Materials; Eur.J.Mech.
A/Solids 11, 585–607, 1992.
Saiki I., Terada K., Ikeda K., Hori M.: Appropriate Number of Unit Cells in a Repre-
sentative Volume Element for Micro-Structural Bifurcation Encountered in a Multi-Scale
Modeling; Comput.Meth.Appl.Mech.Egng. 191, 2561–2585, 2002.
Sanders W., Gibson L.J.: Mechanics of Hollow Sphere Foams; Mater.Sci.Engng. A347,
70–85, 2003.
Sangani A., Lu W.: Elastic Coefficients of Composites Containing Spherical Inclusions
in a Periodic Array; J.Mech.Phys.Sol. 35, 1–21, 1987.
– 116 –
Santosa S.P., Wierzbicki T.: On the Modeling of Crush Behavior of a Closed-Cell Alu-
minum Foam Structure; J.Mech.Phys.Sol. 46, 645–669, 1998.
Sautter M., Dietrich C., Poech M.H., Schmauder S., Fischmeister H.F.: Finite Element
Modelling of a Transverse-Loaded Fibre Composite: Effects of Section Size and Net Den-
sity; Comput.Mater.Sci. 1, 225–233, 1993.
Schapery R.A.: Thermal Expansion Coefficients of Composite Materials Based on Energy
Principles; J.Compos.Mater. 2, 380–404, 1968.
Schjødt–Thomsen J., Pyrz R.: Influence of Statistical Cell Description on the Local Strain
and Overall Properties of Cellular Materials; in “Proceedings of NUMIFORM 2004”
(Eds. S.Ghosh, J.M.Castro, J.K.Lee), pp. 1630–1635; American Institute of Physics,
Melville, NY, 2004.
Schmauder S., Wulf J., Steinkopff T., Fischmeister H.: Micromechanics of Plasticity
and Damage in an Al/SiC Metal Matrix Composite; in “Micromechanics of Plasticity
and Damage of Multiphase Materials” (Eds. A.Pineau, A.Zaoui), pp. 255–262; Kluwer,
Dordrecht, The Netherlands, 1996.
Schrefler B.A., Lefik M., Galvanetto U.: Correctors in a Beam Model for Unidirectional
Composites; Compos.Mater.Struct. 4, 159–190, 1997.
Schrefler B.A., Galvanetto U., Pellegrino C., Ohmenhäuser F.: Global Non-Linear Be-
havior of Periodic Composite Materials; in “Discretization Methods in Structural Me-
chanics” (Eds. H.A.Mang, F.G.Rammerstorfer), pp. 265–272; Kluwer, Dordrecht, The
Netherlands, 1999.
Schwabe F.: Einspieluntersuchungen von Verbundwerkstoffen mit periodischer Mikrostruk-
tur. Doctoral Thesis, Rheinisch-Westfälische Technische Hochschule, Aachen, Germany,
2000.
Segurado J., LLorca J., González C.: On the Accuracy of Mean-Field Approaches to
Simulate the Plastic Deformation of Composites; Scr.Mater. 46, 525–529, 2002.
Segurado J., González C., LLorca J.: A Numerical Investigation of the Effect of Particle
Clustering on the Mechanical Properties of Composites; Acta mater. 51, 2355–2369,
2003.
Segurado J., Parteder E., Plankensteiner A.F., Böhm H.J.: Micromechanical Studies of
the Densification of Porous Molybdenum; Mater.Sci.Engng. A333, 270–278, 2002.
Šejnoha M., Zeman J.: Overall Viscoelastic Response of Random Fibrous Com-
posites with Statistically Quasi Uniform Distribution of Reinforcements; Com-
put.Meth.Appl.Mech.Engng. 191, 5027–5044, 2002.
Shan Z., Gokhale A.M.: Micromechanics of Complex Three-Dimensional Microstruc-
tures; Acta mater. 49, 2001–2015, 2001.
Shen H., Lissenden C.J.: 3D Finite Element Analysis of Particle-Reinforced Aluminum;
Mater.Sci.Engng. A338, 271–281, 1997.
Shenoy V.B., Miller R., Tadmor E.B., Phillips R., Ortiz M.: Quasicontinuum Models of
Interfacial Structures and Deformation; Phys.Rev.Lett. 80, 742–745, 1998.
Shephard M.S., Beall M.W., Garimella R., Wentorf R.: Automatic Construction of 3/D
Models in Multiple Scale Analysis; Comput.Mech. 17, 196–207, 1995.
Sherwood J.A., Quimby H.M.: Micromechanical Modeling of Damage Growth in Tita-
nium Based Metal-Matrix Composites; Comput.Struct. 56, 505–54, 1995.
Shulmeister V., van der Burg M.W.D., van der Giessen E., Marissen R.: A Numerical
Study of Large Deformations of Low-Density Elastomeric Open-Cell Foams; Mech.Mater.
30, 125–140, 1998.
– 117 –
Siegmund T., Werner E., Fischer F.D.: On the Thermomechanical Deformation Behavior
of Duplex-Type Materials; J.Mech.Phys.Sol. 43, 495–552, 1995.
Siegmund T., Cipra R., Liakus J., Wang B., LaForest M., Fatz A.: Processing–
Microstructure–Property Relationships in Short Fiber Reinforced Carbon–Carbon Com-
posite System; in “Mechanics of Microstructured Materials” (Ed. H.J.Böhm), pp. 235–
258; Springer Verlag, Vienna, 2004 (in print).
Silberschmidt V.V., Werner E.A.: Analyses of Thermal Stresses’ Evolution in Ferritic–
Austenitic Duplex Steels; in “Thermal Stresses 2001” (Eds. Y.Tanigawa, R.B.Hetnarski,
N.Noda), pp. 327–330; Osaka Prefecture University, Osaka, Japan, 2001.
Silnutzer N.: Effective Constants of Statistically Homogeneous Materials. Doctoral The-
sis, University of Pennsylvania, Philadelphia, PA, 1972.
Simone A.E., Gibson L.J.: Deformation Characteristics of Metal Foams; Acta mater.
46, 2139–2150, 1998.
Smit R.J.M., Brekelmans W.A.M., Meijer H.E.H.: Prediction of the Mechanical Be-
havior of Non-Linear Heterogeneous Systems by Multi-Level Finite Element Modeling;
Comput.Meth.Appl.Mech.Engng. 155, 181–192, 1998.
Soppa E., Schmauder S., Fischer G., Brollo J., Weber U.: Deformation and Damage in
Al/Al−2O3 ; Comput.Mater.Sci. 28, 574–586, 2003.
Sørensen N.J.: A Planar Type Analysis for the Elastic-Plastic Behaviour of Continuous
Fibre-Reinforced Metal-Matrix Composites Under Longitudinal Shearing and Combined
Loading. DCAMM Report No.418, Technical University of Denmark, Lyngby, 1991.
Sørensen N.J., Suresh S., Tvergaard V., Needleman A.: Effects of Reinforcement Ori-
entation on the Tensile Response of Metal Matrix Composites; Mater.Sci.Engng. A197,
1–10, 1995.
Spiegler R., Schmauder S., Fischmeister H.F.: Simulation of Discrete Void Formation in
a WC–Co Microstructure; in “Finite Elements in Engineering Applications 1990”, pp.
21–42;INTES GmbH, Stuttgart, Germany, 1990.
Steinkopff T., Sautter M., Wulf J.: Mehrphasige Finite Elemente in der Verformungs-
und Versagensanalyse grob mehrphasiger Werkstoffe; Arch.Appl.Mech. 65, 496–506,
1995.
Stroeven M., Askes H., Sluys L.J.: A Numerical Approach to Determine Representative
Volumes for Granular Materials; in “Proc. Fifth World Congress on Computational Me-
chanics (WCCM V)” (Eds.: H.A.Mang, F.G.Rammerstorfer, J.Eberhardsteiner), paper
# 80395; Vienna University of Technology, Vienna, 2002.
Sun L.Z., Ju J.W., Liu H.T.: Elastoplastic Modeling of Metal Matrix Composites with
Evolutionary Particle Debonding; Mech.Mater. 35, 559–569, 2003.
Suquet P.M.: Elements of Homogenization for Inelastic Solid Mechanics; in “Homoge-
nization Techniques in Composite Media” (Eds. E.Sanchez–Palencia, A.Zaoui), pp. 194–
278; Springer–Verlag, Berlin, Germany, 1987.
Suquet P.M.: Effective Properties of Nonlinear Composites; in “Continuum Microme-
chanics” (Ed. P.Suquet), pp. 197–264; Springer–Verlag, Vienna, Austria, 1997.
Suzuki T., Yu P.K.L.: Complex Elastic Wave Band Structures in Three-Dimensional
Periodic Elastic Media; J.Mech.Phys.Sol. 46, 115–138, 1998.
Swan C.C.: Techniques for Stress- and Strain-Controlled Homogenization of Inelastic
Periodic Composites; Comput.Meth.Appl.Mech.Engng. 117, 249–267, 1994.
Tadmor E.B., Phillips R., Ortiz M.: Hierarchical Modeling in the Mechanics of Materials;
Int.J.Sol.Struct. 37, 379–390, 2000.
– 118 –
Takano N., Zako M., Okazaki T.: Efficient Modeling of Microscopic Heterogeneity and Lo-
cal Crack in Composite Materials by Finite Element Mesh Superposition Method; JSME
Int.J. Srs.A 44, 602–609, 2001.
Talbot D.R.S., Willis J.R.: Variational Principles for Inhomogeneous Non-Linear Media;
J.Appl.Math. 35, 39–54, 1985.
Talbot D.R.S., Willis J.R.: Bounds of Third Order for the Overall Response of Nonlinear
Composites; J.Mech.Phys.Sol. 45, 87–111, 1997.
Tandon G.P., Weng G.J.: The Effect of Aspect Ratio of Inclusions on the Elastic Prop-
erties of Unidirectionally Aligned Composites; Polym.Compos. 5, 327–333, 1984.
Tandon G.P., Weng G.J.: A Theory of Particle-Reinforced Plasticity; J.Appl.Mech. 55,
126–135, 1988.
Taya M., Mori T.: Dislocations Punched Out Around a Short Fibre in MMC Subjected
to Uniform Temperature Change; Acta metall. 35, 155–162, 1987.
Taya M., Armstrong W.D., Dunn M.L., Mori T.: Analytical Study on Dimensional
Changes in Thermally Cycled Metal Matrix Composites; Mater.Sci.Engng. A143, 143–
154, 1991.
Teply J.L., Dvorak G.J.: Bounds on Overall Instantaneous Properties of Elastic-Plastic
Composites; J.Mech.Phys.Sol. 36, 29–58, 1988.
Terada K., Kikuchi N.: Microstructural Design of Composites Using the Homogenization
Method and Digital Images; Mater.Sci.Res.Int. 2, 65–72, 1996.
Terada K., Kikuchi N.: Nonlinear Homogenization Method for Practical Applications; in
“Computational Methods in Micromechanics” (Eds. S.Ghosh, M.Ostoja–Starzewski), pp.
1–16; AMD–Vol.212/MD–Vol.62, ASME, New York, NY, 1996.
Terada K., Miura T., Kikuchi N.: Digital Image-Based Modeling Applied to the Homog-
enization Analysis of Composite Materials; Comput.Mech. 20, 331–346, 1997.
Terada K., Hori M., Kyoya T., Kikuchi N.: Simulation of the Multi-Scale Convergence
in Computational Homogenization Approach; Int.J.Sol.Struct. 37, 2285–2311, 2000.
Terada K., Saiki I., Matsui K., Yamakawa Y.: Two-Scale Kinematics and Linearization
for Simultaneous Two-Scale Analysis of Periodic Heterogeneous Solids at Finite Strain;
Comput.Meth.Appl.Mech.Engng. 192, 3531–3563, 2003.
Tohgo K., Chou T.W.: Incremental Theory of Particulate-Reinforced Composites Includ-
ing Debonding Damage; JSME Int.J. Srs.A 39, 389–397, 1996.
Tomita Y., Higa Y., Fujimoto T.: Modeling and Estimation of Deformation Behavior of
Particle Reinforced Metal-Matrix Composite; Int.J.Mech.Sci. 42, 2249–2260, 2000.
Torquato S.: Random Heterogeneous Media: Microstructure and Improved Bounds on
Effective Properties; Appl.Mech.Rev. 44, 37–75, 1991.
Torquato S.: Morphology and Effective Properties of Disordered Heterogeneous Media;
Int.J.Sol.Struct. 35, 2385–2406, 1998.
Torquato S.: Effective Stiffness Tensor of Composite Media: II. Applications to Isotropic
Dispersions; J.Mech.Phys.Sol. 46, 1411–1440, 1998.
Torquato S.: Random Heterogeneous Media. Springer–Verlag, New York, NY, 2001.
Torquato S., Hyun S.: Effective Medium Approximation for Composite Media: Realizable
Single-Scale Dispersions; J.Appl.Phys. 89, 1725–1729, 2001.
Torquato S., Lado F.: Improved Bounds on the Effective Moduli of Random Arrays of
Cylinders; J.Appl.Mech. 59, 1–6, 1992.
– 119 –
Torquato S., Lado F., Smith P.A.: Bulk Properties of Two-Phase Disordered Media. IV.
Mechanical Properties of Suspensions of Penetrable Spheres at Nondilute Concentrations;
J.Chem.Phys. 86, 6388–6392, 1987.
Tsai S.W., Wu E.M.: A General Theory of Strength for Anisotropic Materials;
J.Compos.Mater. 5, 38–80, 1971.
Tsotsis T.K., Weitsman Y.: Energy Release Rates for Cracks Caused by Moisture Ab-
sorption in Graphite/Epoxy Composites; J.Compos.Mater. 24, 483–496, 1990.
Tszeng T.C.: The Effects of Particle Clustering on the Mechanical Behavior of Particle
Reinforced Composites; Composites 29B, 299–308, 1998.
Tucker C.L., Liang E.: Stiffness Predictions for Unidirectional Short-Fiber Composites:
Review and Evaluation; Compos.Sci.Technol. 59, 655–671, 1999.
Tvergaard V.: Constitutive Relations for Creep in Polycrystals with Boundary Cavita-
tion; Acta metall. 32, 1977–1990, 1984.
Tvergaard V.: Analysis of Tensile Properties for a Whisker-Reinforced Metal-Matrix
Composite; Acta metall.mater. 38, 185–194, 1990.
Tvergaard V.: Fibre Debonding and Breakage in a Whisker-Reinforced Metal.
Mater.Sci.Engng. A190, 215–222, 1994.
Tvergaard V.: Effect of Void Size Difference on Growth and Cavitation Instabilities;
J.Mech.Phys.Sol. 44, 1237–1253, 1996.
Tvergaard V.: Breakage and Debonding of Short Brittle Fibres Among Particulates in a
Metal Matrix; Mater.Sci.Engng. A369, 192–200, 2004.
Tvergaard V., Needleman A.: Analysis of the Cup-Cone Fracture in a Round Tensile
Bar; Acta metall. 32, 157–169, 1984.
Vajjhala S., Kraynik A.M., Gibson L.J.: A Cellular Solid Model for Modulus Reduction
due to Resorption of Trabeculae in Bone; J.Biomech.Engng. 122, 511–515, 2000.
van Rietbergen B., Müller R., Ulrich D., Rüegsegger P., Huiskes R.: Tissue Stresses
and Strain in Trabeculae of a Canine Proximal Femur Can be Quantified from Computer
Reconstructions; J.Biomech. 32, 165–173, 1999.
Váradi K., Néder Z., Friedrich K., Flöck J.: Finite-Element Analysis of a Polymer Com-
posite Subjected to Ball Indentation; Compos.Sci.Technol. 59, 271–281, 1999.
Vejen N., Pyrz R.: Transverse Crack Growth in Glass/Epoxy Composites with Exactly
Positioned Long Fibers. Part II: Numerical; Composites 33B, 279–290, 2002.
Voigt W.: Über die Beziehung zwischen den beiden Elasticitäts-Constanten isotroper
Körper; Ann.Phys. 38, 573–587, 1889.
Vonach W.K.: A General Solution to the Wrinkling Problem of Sandwiches. VDI–Verlag
(Reihe 18, Nr.268), Düsseldorf, Germany, 2001.
Vosbeek P.H.J., Meurs P.F.M., Schreurs P.J.G.: Micro-Mechanical Constitutive Models
for Composite Materials with Interfaces; in “Proc.1st Forum of Young European Re-
searchers”, pp. 223–228; Université de Liége, Liége, Belgium, 1993.
Voyiadjis G.Z., Kattan P.I.: Advances in Damage Mechanics: Metals and Metal Matrix
Composites. Elsevier, Amsterdam, 1999.
Wakashima K., Tsukamoto H., Choi B.H.: Elastic and Thermoelastic Properties of Metal
Matrix Composites with Discontinuous Fibers or Particles: Theoretical Guidelines to-
wards Materials Tailoring; in “The Korea–Japan Metals Symposium on Composite Ma-
terials”, pp. 102–115; The Korean Institute of Metals, Seoul, Korea, 1988.
– 120 –
Wallin K., Saario T., Törrönen K.: Fracture of Brittle Particles in a Ductile Matrix;
Int.J.Fract. 32, 201–209, 1987.
Walpole L.J.: On Bounds for the Overall Elastic Moduli of Inhomogeneous Systems —
II; J.Mech.Phys.Sol. 14, 289–301, 1966.
Walter M.E., Ravichandran G., Ortiz M.: Computational Modeling of Damage Evolution
in Unidirectional Fiber Reinforced Ceramic Matrix Composites; Comput.Mech. 20, 192–
198, 1997.
Wang J., Andreasen J.H., Karihaloo B.L.: The Solution of an Inhomogeneity in a Finite
Plane Region and its Application to Composite Materials; Compos.Sci.Technol. 60, 75–
82, 2000.
Warren W.E., Kraynik A.M.: The Nonlinear Elastic Properties of Open-Cell Foams;
J.Appl.Mech. 58, 376–381, 1991.
Watt D.F., Xu X.Q., Lloyd D.J.: Effects of Particle Morphology and Spacing on the
Strain Fields in a Plastically Deforming Matrix; Acta mater. 44, 789–799, 1996.
Weihe S., Kröplin B.H.: The Fictitious Crack Concept in the Mechanics of Composites, in
“Computational Plasticity: Fundamentals and Applications” (Eds. D.R.Owen, E.Oñate,
E.Hinton), pp. 1215–1226; Pineridge Press, Swansea, 1995.
Weissenbek E.: Finite Element Modelling of Discontinuously Reinforced Metal Matrix
Composites. VDI–Verlag (Reihe 18, Nr.164), Düsseldorf, Germany, 1994.
Weissenbek E., Böhm H.J., Rammerstorfer F.G.: Micromechanical Investigations of Ar-
rangement Effects in Particle Reinforced Metal Matrix Composites; Comput.Mater.Sci.
3, 263–278, 1994.
Weissenbek E., Pettermann H.E., Suresh S.: Numerical Simulation of Plastic Deforma-
tion in Compositionally Graded Metal–Ceramic Structures; Acta mater. 45, 3401–3417,
1997.
Weng G.J.: The Theoretical Connection Between Mori–Tanaka Theory and the Hashin–
Shtrikman–Walpole Bounds; Int.J.Engng.Sci. 28, 1111–1120, 1990.
Willis J.R.: Bounds and Self-Consistent Estimates for the Overall Properties of
Anisotropic Composites; J.Mech.Phys.Sol. 25, 185–202, 1977.
Willis J.R.: The Overall Response of Nonlinear Composite Media; Eur.J.Mech. A/Solids
19, S165–S184, 2000.
Withers P.J.: The Determination of the Elastic Field of an Ellipsoidal Inclusion in a
Transversely Isotropic Medium, and its Relevance to Composite Materials; Phil.Mag.
A59, 759–781, 1989.
Withers P.J., Stobbs W.M., Pedersen O.B.: The Application of the Eshelby Method of
Internal Stress Determination to Short Fibre Metal Matrix Composites; Acta metall. 37,
3061–3084, 1989.
Wittig L.A., Allen D.H.: Modeling the Effect of Oxidation on Damage in SiC/Ti-15-3
Metal Matrix Composites; J.Engng.Mater.Technol. 116, 421–427, 1994.
Wu J.F., Shephard M.S., Dvorak G.J., Bahei–el–Din Y.A.: A Material Model for the
Finite Element Analysis of Metal Matrix Composites; Compos.Sci.Technol. 35, 347–
366, 1989.
Wu T.T: The Effect of Inclusion Shape on the Elastic Moduli of a Two-Phase Material;
Int.J.Sol.Struct. 2, 1–8, 1966.
Wulf J., Steinkopff T., Fischmeister H.: FE-Simulation of Crack Paths in the Real Mi-
crostructure of an Al(6061)/SiC Composite; Acta mater. 44, 1765–1779, 1996.
– 121 –
Xia L., Shih C.F.: Ductile Crack Growth — II. Void Nucleation and Geometry Effects
on Macroscopic Fracture Behavior; J.Mech.Phys.Sol. 43, 1953–1981, 1995.
Xia Z., Zhang Y., Ellyin F.: A Unified Periodical Boundary Conditions for Representative
Volume Elements of Composites and Applications; Int.J.Sol.Struct. 40, 1907–1921, 2003.
Zaoui A.: Changement d’échelle: motivation et méthodologie; in “Homogénéisation en
mécanique des materiaux 1. Matériaux aléatoires élastiques et milieux périodiques” (Eds.
M.Bornert, T.Bretheau, P.Gilormini), pp. 19–40; Editions Hermès, Paris, 2001.
Zaoui A.: Plasticité: Approches en champ moyen; in “Homogénéisation en mécanique
des materiaux 2. Comportements non linéaires et problèmes ouverts” (Eds. M.Bornert,
T.Bretheau, P.Gilormini), pp. 17–44; Editions Hermès, Paris, 2001.
Zaoui A.: Continuum Micromechanics: Survey; J.Engng.Mech. 128, 808–816, 2002.
Zeman J., Šejnoha M.: Numerical Evaluation of Effective Elastic Properties of Graphite
Fiber Tow Impregnated by Polymer Matrix; J.Mech.Phys.Sol. 49, 69–90, 2001.
Zhao Y.H., Weng G.J.: A Theory of Inclusion Debonding and Its Influence on the Stress–
Strain Relations of a Ductile Matrix Composite; Int.J.Dam.Mech. 4, 196–211, 1995.
Zhong X.A., Knauss W.G.: Effects of Particle Interaction and Size Variation on Damage
Evolution in Filled Elastomers; Mech.Compos.Mater.Struct. 7, 35–53, 2000.
Zhou S.J., Curtin W.A.: Failure of Fiber Composites: A Lattice Green Function Ap-
proach; Acta metall.mater. 43, 3093–3104, 1995.
Zhu H.X., Mills N.J., Knott J.F.: Analysis of the High Strain Compression of Open-Cell
Foams; J.Mech.Phys.Sol. 45, 1875–1904, 1997.
Zimmerman R.W.: Hashin–Shtrikman Bounds on the Poisson Ratio of a Composite
Material; Mech.Res.Comm. 19, 563–569, 1992.
Zohdi T.I.: A Model for Simulating the Deterioration of Structural-Scale Material Re-
sponses of Microheterogeneous Solids; in “Euromech Colloqium 402 — Micromechanics
of Fracture Processes” (Eds. D.Gross, F.D.Fischer, E.van der Giessen), pp. 91–92; TU
Darmstadt, Darmstadt, Germany, 1999
Zohdi T.I.: Constrained Inverse Formulations in Random Material Design; Com-
put.Meth.Appl.Mech.Engng. 192, 3179–3194, 2003.
Zohdi T.I., Wriggers P.: A Model for Simulating the Deterioration of Structural-Scale Ma-
terial Responses of Microheterogeneous Solids; Comput.Meth.Appl.Mech.Engng. 190,
2803–2823, 2001.
Zohdi T.I., Oden J.T., Rodin G.J.: Hierarchical Modeling of Heterogeneous Bodies; Com-
put.Meth.Appl.Mech.Engng. 138, 273–289, 1996.
– 122 –