You are on page 1of 11

Mechanics of Materials 38 (2006) 873–883

www.elsevier.com/locate/mechmat

Computational micromechanics of composites:


The effect of particle spatial distribution
Javier Segurado, Javier LLorca *

Department of Materials Science, Polytechnic University of Madrid, E.T.S. de Ingenieros de Caminos, 28040 Madrid, Spain

Received 10 December 2004; received in revised form 15 February 2005

Abstract

The effect of the reinforcement spatial distribution on the tensile deformation was investigated in a sphere-reinforced
ductile-matrix composite through the finite element simulation of a representative volume element of the material. Com-
posites with four different random and isotropic particle spatial distributions were considered: a homogeneous one and
three heterogeneous ones, where the particles were clustered in spherical regions with increasing local reinforcement vol-
ume fractions. The inhomogeneities in the particle spatial distribution had a negligible influence on the effective composite
properties in the elastic and plastic regimes. The average stresses borne by the particles were, however, higher in the clus-
tered microstructures and marked stress concentrations were found between particles very close to each other along the
loading direction. This affected the damage: the probability of fracture of the spherical reinforcements in the clustered
materials was 3–6 times higher than in the homogeneous one, and decohesion at the particle/matrix interface (which
was simulated by interface elements around the particle surfaces) occurred earlier and grew faster in the inhomogeneous
composites, leading to a significant reduction in the hardening rate and ductility. These simulations showed that damage
(either by particle or interface decohesion) was preferentially nucleated within the clusters between particles very close to
each other and oriented along the loading axis within a conic angle of ±45°, in excellent agreement with the experimental
observations reported in the literature.
Ó 2005 Elsevier Ltd. All rights reserved.

Keywords: Micromechanical modeling; Particle-reinforced composites; Clustering; Damage; Interface decohesion

1. Introduction strategies may be used. On the one hand, homogeni-


zation techniques are applied to finding the appro-
Micromechanics of composites is aimed at deter- priate homogenized (or averaged) equations which
mining the macroscopic (or effective) composite describe the mechanical deformation of a represen-
properties through models which incorporate tative volume element (RVE) of the composite
the microstructural details. Two different analysis which is much larger than the heterogeneities in the
microstructure. Various averaging theorems—
*
Corresponding author. Tel.: +34 913365375; fax: +34
which normally assume a random and homogeneous
915437845. reinforcement distribution—provide the volume-
E-mail address: jllorca@mater.upm.es (J. LLorca). averaged stress and strain fields in each phase

0167-6636/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechmat.2005.06.026
874 J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883

(Nemat-Nasser and Hori, 1999; Torquato, 2001). 1997; Michel et al., 1999; Segurado and LLorca,
On the other hand, periodic microfield approaches 2002) and elasto-plastic deformation (Böhm and
presume that the reinforcements are placed at pre- Han, 2001; Segurado et al., 2003; González et al.,
scribed (regular) locations in space, and determine 2004), and more recent results have explored the
the macroscopic properties from the numerical or development of damage by matrix void growth
analytical simulation of the mechanical response (LLorca and Segurado, 2004) and particle fracture
of a unit cell with symmetric (or antisymmetric) (Böhm et al., 2004) in RVEs with homogeneous par-
boundary conditions which respect the microstruc- ticle distributions. In this paper, we extend the pre-
tural symmetries (Böhm, 2004). vious results obtained with this technique to study
Both approaches have advantages and limita- systematically the effect of inhomogeneous particle
tions. Homogenization techniques provide analytical distributions on the elastic, elasto-plastic and
or semi-analytical constitutive equations in linear elasto-plastic deformation with damage (particle
and non-linear regimes, which can be readily used fracture or interface decohesion) of a model
as input for the composite behavior in structural composite made up of a random and isotropic dis-
analysis codes. However, they are not well suited to persion of ceramic spheres embedded in an elasto-
account for the effect of damage because damage is plastic continuous matrix.
triggered by the extremal values and they supply
information only on the average ones. In addition, 2. Composite microstructure and finite element model
damage leads to the rapid localization of the defor-
mation, and the influence of this phenomenon on The RVE for the numerical simulations was a
the macroscopic properties cannot be captured easily cubic cell containing a random and isotropic disper-
with techniques which relay in averaging theorems. sion of spherical particles. The average sphere
Microfield approaches can obtain the complex stress volume fraction in the RVE, n, was 15% in all the
and strain fields generated in the matrix and rein- simulations presented in this paper. RVEs with four
forcements upon deformation, and so they overcome different kinds of random microstructures were cre-
the first limitation of homogenization techniques ated. In the first one, 30 particles generated with the
to model damage. Nevertheless, the periodicity random sequential adsorption algorithm (Segurado
imposed by the model assumptions means that dam- and LLorca, 2002; Rintoul and Torquato, 1997)
age nucleates and grows simultaneously throughout were homogeneously distributed throughout the
the microstructure, in opposition to the experimental RVE. The other three microstructures, with increas-
data which show that damage is initially confined in ing inhomogeneity levels, were formed by a random
small regions of the material and spreads progres- and homogeneous dispersion of spherical regions
sively with deformation (LLorca and González, (clusters). The spherical particles were concentrated
1998). In fact, it is well established experimentally around the center of each cluster and the inhomoge-
that damage nucleation in polymer- and metal– neity of the microstructures was defined by the
matrix composites occurs in regions of the micro- sphere volume fraction in the cluster regions,
structure that contain high local volume fraction of ncl = 20%, 30%, and 40%. Each cluster was made
reinforcements (Cantwell and Roulin-Moloney, up of seven spheres and there were seven clusters
1989; Lewandowski et al., 1989; Poza and LLorca, in each RVE, a total of 49 particles. It should be
1996). Thus, the accurate simulation of deformation noted that the RVE microstructure was periodic
and damage in composites requires new simulation along the three cube axes to apply periodic bound-
techniques which can include inhomogeneous rein- ary conditions to the cube faces, because the effec-
forcement distributions. tive behavior derived under these conditions for a
The limitations of the classic micromechanical finite RVE is always closer to that of an infinite
model for composites have been overcome by a RVE than those obtained under imposed forces or
new simulation approach based on the finite ele- displacements. Two inhomogeneous particle disper-
ment simulation of a three-dimensional RVE con- sions corresponding to ncl = 20% and 40% are
taining several dozens of reinforcements randomly shown in Fig. 1(a) and (b), respectively. Highly
distributed. Various studies have demonstrated the inhomogeneous distributions present particle-rich
capabilities of this method to capture with great and particle-free regions in the microstructure
accuracy the effective properties of particle-rein- (Fig. 1(b)) as opposed to the homogeneous
forced composites subjected to elastic (Gusev, distributions.
J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883 875

hourglass control, which exhibit minimal volumetric


locking during plastic straining. Approximately 500
elements were used to represent each sphere, ensur-
ing that the sphere volume fraction in the discretized
models was within 0.1% of the theoretical value. In
addition, an adaptative meshing technique was used
to reduce the element size in the matrix within the
cluster zone, coarsening the mesh in the rest of the
cell. Special care was taken to minimize the number
of distorted elements to a few (<10) in each model
and to include at least two tetrahedra through the
matrix thickness between neighbor particles closely
packed in the clusters. A typical model of the clus-
tered microstructures had 100,000 elements and
150,000 nodes and it was checked that the discreti-
zation of the cubic cells was fine enough to give
accurate results from the numerical viewpoint
(Segurado et al., 2003).
The RVE mechanical response in uniaxial ten-
sion along one of the cube axes was computed by
the finite element method using periodic boundary
conditions on the cube faces. Simulations were
carried out with Abaqus/Standard (Abaqus, 2004)
within the framework of the small displacements
theory. The spherical particles behaved as elastic,
isotropic solids characterized by their elastic modu-
lus Ep = 400 GPa and PoissonÕs ratio mp = 0.20. The
matrix was modeled as an isotropically hardening
elastic plastic solid following the incremental (J2)
theory of plasticity, and the total matrix strain
was given by the sum of the elastic and plastic strain
components. The matrix elastic constants were
Em = 70 GPa and mm = 0.30, and the isotropic
matrix hardening during plastic deformation was
given by the expression
nm
Fig. 1. Lateral view of the particle spatial distribution in the req p
m ¼ Am ½m  ð1Þ
cubic RVEs. Inhomogeneous distribution with (a) ncl = 0.20 and
(b) ncl = 0.40. Each RVE contains seven clusters with seven where req
m is the Von Mises equivalent stress and pm
spheres and one of the clusters has been shaded for clarity. Notice stands for accumulated plastic strain. The constants
that the microstructure is periodic and that the particles Am = 400 MPa and nm = 0.15 are typical of an Al
intersecting a cube face also appear at the opposite face. alloy matrix reinforced with ceramic particles. More
details of the microstructure generation, the finite
Cubic cells for the finite element simulations were element discretization and simulation, and the
created from these sphere distributions: the particles boundary conditions can be found in Segurado
intersecting the cube faces were split into an appro- and LLorca (2002) and Segurado et al. (2002).
priate number of parts and copied on the opposite
sides of the cube. Three faces of the cube were 3. Results
meshed with quadratic triangles, and the meshes
were copied on the opposite sides. The model 3.1. Elastic behavior
volume (matrix and spheres) was meshed using
modified 10-node tetrahedra (C3D10M in Abaqus, Four different cubic RVEs were generated and
2004) with integration at four Gauss points and analyzed for the homogeneous microstructure
876 J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883

1.05 ness tensor of the cluster region Lcl was obtained


3D numerical simulation using one of the homogenization methods (Mori
of RVE and Tanaka, 1973; Benveniste, 1987; self-consistent,
Self-consistent method
1.04 Christensen and Lo, 1979; and TorquatoÕs third
Third order approximation
Mori-Tanaka method order approximation, Torquato, 1998) assuming
that the cluster region behaves as a composite rein-
1.03
forced with a volume fraction ncl of spheres. The
E / Ehom

inhomogeneous composite was represented by a ran-


1.02
dom and homogeneous distribution of clustered
regions occupying a volume fraction n/ncl and the
elastic stiffness tensor was computed using again
1.01 the corresponding homogenization method.
While the results of the double-homogenization
technique based on Mori–TanakaÕs model did not
1 show any effect of clustering on the elastic properties
(as is proved elsewhere, Segurado and LLorca,
0.1 0.2 0.3 0.4 0.5 0.6 0.7 2005a), those based on TorquatoÕs model predicted
ξcl the same influence of clustering on the elastic modu-
lus as the numerical simulations, although the max-
Fig. 2. Average elastic modulus of composites, E, with different
particle spatial distributions (normalized by the modulus of the
imum was attained at higher levels of inhomogeneity
homogeneous particle distribution, Ehom) as a function of the in the microstructure (ncl  55%). The self-consistent
degree of clustering represented by the local particle volume approximation also led to a maximum for the same
fraction in the clusters, ncl. The average results from the degree of inhomogeneity but its value was far higher.
numerical simulations of the 3D RVEs are plotted together with This is very likely a consequence of the self-consis-
those obtained from the self-consistent method, TorquatoÕs third
order approximation and the Mori–Tanaka method.
tent method, which provides very stiff predictions
for the effective properties when the sphere volume
fraction is very high, overestimating Lcl, and hence
(ncl = n = 0.15) and another four for each of the the effect of clustering on the composite response.
three inhomogeneous microstructures with ncl = On the contrary, TorquatoÕs third order approxima-
0.20, 0.30 and 0.40. They were subjected to uniaxial tion is known to be the most accurate homogeniza-
tension along the three perpendicular cube axes, tion technique to predict the elastic properties of
leading to 12 values of the elastic modulus for each this class of composites up to volume fractions
composite. The average value of the elastic modulus of 50% (Segurado and LLorca, 2002). The origin
E (normalized by the average one of the composite of the differences between the numerical simulation
with a statistically homogeneous sphere distribution, and the homogenization models in the clustering
Ehom) is plotted in Fig. 2 as a function of the degree level which led to the maximum stiffness is not
of clustering represented by ncl. Drugan and Willis known, but it may be related to the small number
(1996) showed that a RVE containing 49 spheres of spheres (only 7) in each cluster of the RVE.
was large enough to simulate the macroscopic com- Although the distance between the spheres in the
posite behavior in the elastic regime and the error cluster is minimum (Segurado et al., 2003), the local
bars in the figure support this conclusion as the dis- volume fraction ncl is limited to 40% due to the num-
persion in the elastic modulus among different RVEs ber spheres. Higher local volume fractions (close to
of the same microstructure was negligible. These the theoretical limit of 64%) could be attained by
results indicate that the influence of the inhomoge- increasing the number of spheres within the cluster
neous particle distribution on the elastic properties while the average minimum distance was held
was negligible (the maximum difference was below constant, and this could modify the position of the
1%) and that a maximum was attained when the maximum for the numerical simulations in Fig. 2.
local volume fraction in the cluster was approxi-
mately 30%. The results obtained using various 3.2. Elasto-plastic behavior
homogenization methods are also plotted for com-
parison. They were computed using a double The simulations presented in the previous para-
homogenization technique. Firstly, the elastic stiff- graphs were extended into the elasto-plastic regime.
J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883 877

500

particles
400
σ eq, σpeq (MPa)

300

composite
200
ξcl = 0.40
ξcl = 0.30
100 ξcl = 0.20
ξcl = ξ = 0.15

0
0 0.01 0.02 0.03 0.04 0.05
Strain

Fig. 3. Evolution of the average Von Mises equivalent stress in


the composite, r eq , (i.e. the applied stress) and of the average
Fig. 4. Contour plot of the maximum tensile stress in the
eq
equivalent stress in the particles, r p , as a function of the applied spherical reinforcements of the composite with ncl = 0.40. The
strain in composites with homogeneous (ncl = 0.15) and inhomo-
applied strain in the arrow direction was 5%. Notice the stress
geneous particle distributions (ncl = 0.20, 0.30 and 0.40).
concentrations in the spheres close to each other and oriented
along the loading axis.
Twelve stress–strain curves in tension (four RVEs
deformed along the three perpendicular cube axes)
were computed for each microstructure; they where ri and Vi stand, respectively, for the stress
showed very limited scatter when plastic deforma- tensor and the volume associated with Gauss point
tion became dominant. The maximum difference in i in the sphere. The average values of r eq
p for the
stresses between two RVEs of the same microstruc- spheres in each microstructure are also plotted in
ture was small (around 4% at a tensile strain of 0.05) Fig. 3 as a function of the applied strain. They show
and the standard deviation of the stress was always greater differences between the homogeneous and
below 1.3%. The average stress–strain curves of the the inhomogeneous materials, an indication that
four microstructures are plotted in Fig. 3 (true stres- the average stresses in the particle do increase with
ses and logarithmic strains are used throughout the clustering although the effective properties are not
paper1) and they show that, as in the elastic regime, significantly modified by the particle spatial distri-
the composite with ncl = 0.30 presented the stiffest bution (as in the elastic regime). Moreover, the con-
effective behavior, although the overall behavior of tour plot of the maximum principal stresses in the
composites with homogeneous and heterogeneous spheres (Fig. 4) show that the stresses within the
reinforcement distributions was very similar, the spheres are highly inhomogeneous. Maximum ten-
maximum difference in stresses being 2%. In addi- sile stresses (in excess of 1 GPa) are found in the
tion, the equivalent stress in each sphere, r eq spheres when they are very close to each other and
p was
computed from the volume-averaged stress tensor oriented along the loading axis, while isolated
in each sphere rp , which is given by spheres bear much lower loads and the maximum
P tensile stresses do not exceed 0.6 GPa. From the
ri V i practical viewpoint, these results point out that—
p ¼ Pi
r ð2Þ
iV i in the absence of damage—particle clustering
does not determine the composite behavior. On
1
the other hand, the local values of stress and strains
The true stress along the loading axis was computed as the are higher in the inhomogeneous microstructures
total force divided by the actual cross-section. The corresponding
logarithmic strain was computed from the displacement along the
and this is likely to influence the onset and propaga-
same axis between paired nodes in opposite surfaces, see tion of damage. This is assessed in the following
Segurado et al. (2003) for details. sections.
878 J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883

3.3. Damage by particle cracking 0.12

Particle fracture is an important damage mecha-


nism in particle-reinforced composites (LLorca and

Sphere fracture probability


0.09 m=5
González, 1998; González and LLorca, 2000), and
the influence of the particle spatial distribution on σ0 = 1.5 GPa
the onset of damage can be studied from the magni- σ0 = 2.0 GPa
tude of the stress fields at the local level provided by 0.06
the finite element simulations of the RVE. It is well
established that the strength of brittle particles fol-
lows the Weibull statistics (LLorca et al., 1993;
Lewis and Withers, 1995; LLorca, 1995) and thus 0.03
the fracture probability, Fi, of a region of volume
Vi subjected to a constant stress field given by the
tensor ri is expressed by
  m  0
V i rIi 0.1 0.2 0.3 0.4
F i ¼ 1  exp  ð3Þ (a) ξcl
V 0 r0
where rIi is the maximum principal stress in the re- 0.12
gion, m and r0 stand, respectively, for the Weibull σ0 = 2.0 GPa
modulus and the characteristic strength of the m=3
ceramic material, and V0 is an arbitrary reference m=5
Sphere fracture probability

volume. In our models, each sphere can be consid- 0.09 m=7


ered an aggregate of regions of volume Vi, which
correspond to the volume associated to the Gauss
points in the finite element discretization. Each re- 0.06
gion is subjected to a stress field ri, and V0 is the
sphere volume. The brittle sphere will break when-
ever any of the small regions fails, so the sphere
fracture probability is given by 0.03

Y
N   m 
V i rIi
F ¼1 exp  ð4Þ
i¼1
V 0 r0
0
0.1 0.2 0.3 0.4
rIi
where is the maximum principal stress in the vol- ξcl
(b)
ume i. Neglecting the stress redistribution associated
with the fracture of spheres, F provides an estimate Fig. 5. Estimations of the fraction of broken spheres in
for the fraction of broken spheres as a function of composites with different particle spatial distributions as a
function of the degree of clustering represented by the local
the stresses in the spheres, which can be obtained
particle volume fraction in the clusters, ncl. (a) m = 5 (b)
from the finite element simulations of the RVE. r0 = 2.0 GPa. The applied strain was 5%.
They are plotted in Fig. 5 for the composites with
homogeneous and inhomogeneous particle distribu-
tions when the applied tensile strain was 5%. The ture greatly reduces the flow stress of the composite,
results in Fig. 5(a) show the influence of the Weibull eliminating the slight strength increment provided
modulus m and those in Fig. 5(b) illustrate the effect by the inhomogeneous reinforcement distribution.
of the characteristic strength r0. Both results point In addition, this leads to a marked drop in the strain
out that even small amounts of clustering (ncl = hardening capacity, and the onset of plastic instabil-
0.20) led to a significant increase in the sphere frac- ity (which controls the tensile ductility of the com-
ture probability and thus demonstrated the negative posites, see LLorca and González, 1998) occurs at
effects of inhomogeneous particle distributions on very low strains. These numerical predictions have
the composite behavior. The rapid fracture of the been confirmed recently by experimental results on
spheres in the materials with a clustered microstruc- Al-based composites with controlled particle spatial
J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883 879

distribution (Prasad et al., 2002; Spowart et al.,


2003): composites with homogeneous particle distri- 1
butions presented much larger ductility and lower
damage rates by particle fracture. The influence of
particle strength (r0) on damage, shown in 0.8
Fig. 5(b), follows the expected trends with the
sphere fracture probability increasing rapidly as r0

σ (λ) / t c
0.6
decreases. The effect of the Weibull modulus de-
picted in Fig. 5(a) is more complex, as it depends Ki
on the presence of stress concentrations in the 0.4
spheres (Fig. 4): no definitive conclusions can be
drawn. However, damage seems to be maximum
in (or around) ncl  0.3–0.4, and this follows the 0.2
trends observed in the elastic and plastic regime, 1
which also showed a maximum influence of the
inhomogeneous particle distribution on the effective 0
0 0.2 0.4 0.6 0.8 1
properties in composites with this degree of
λ
clustering.
Fig. 6. Function r(k). Notice the degradation of the element
stiffness once the maximum interfacial strength has been reached.
3.4. Damage by interface decohesion

Interface decohesion is another well-known dam- 1


age mechanism in particle-reinforced composites Cin ¼ tc Duc . ð7Þ
2
(Whitehouse and Clyne, 1993; Babout et al., 2004).
It was incorporated in the numerical simulations of The cohesive model includes the progressive deg-
the cubic RVE by including a three-dimensional radation of the interface stiffness and strength once
interface element at the sphere surfaces. The element the maximum interfacial strength has been reached,
formulation is detailed elsewhere (Segurado and as indicated in the unloading and reloading paths of
LLorca, 2004) and need not be repeated here. Inter- Fig. 6. The initial stiffness Ki is a numerical param-
face decohesion was simulated through a cohesive eter that is adjusted to ensure that the presence of
crack model, where the normal and tangential stress the interface elements does not perturb the stress
transferred by the interface were derived from a fields around the spherical particles in the absence
potential U given by of damage.
Z k The influence of the particle spatial distribution
UðDun ; Dut1 ; Dut2 Þ ¼ Duc rðk0 Þ dk0 ð5Þ on the mechanical behavior of composites including
0 interface decohesion was studied in the two limiting
where Dun, Dut1 and Dut2 stand for the normal and cases of homogeneous and highly clustered micro-
tangential relative displacements between the crack structures. The cubic RVEs of the homogeneous
faces and Duc is the critical normal (or tangential) composite contained 35 spherical particles while
displacement between the crack faces at which all the inhomogeneous ones were formed by seven
interaction vanishes. k is the generalized crack open- spherical clusters with five spheres each. The RVE
ing displacement expressed as size was a compromise between the minimum one
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi necessary to obtain accurate results during elasto-
2  2  2ffi plastic deformation of sphere-reinforced composites
Dun Dut1 Dut2
k¼ þ þ . ð6Þ (Segurado et al., 2003) and the maximum imposed
Duc Duc Duc
by the computing power available. The reduced
The function r(k) (which stands for the normal scatter among the simulations of different cubic cells
stress transferred through the crack in the absence with the same microstructure (either homogeneous
of tangential displacements) is plotted in Fig. 6. or clustered) showed a posteriori that the size
The interface properties are given by the interfacial selected was adequate. The average sphere volume
strength, tc, and the fracture energy, Cin, which is fraction was n = 0.15, while ncl = 0.37. The interface
the area enclosed under the r(k) function fracture energy Cin was 400 J/m2, and the interface
880 J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883

strength tc was assumed to vary from particle to materials, respectively. The differences between the
particle according to curves of the same microstructure and failure mech-
anism (either none or interface decohesion) showed
tc ¼ tc0 ð lnð1  N ÞÞ0:2 ð8Þ
very little scatter, so the size of the RVEs was large
where 0 6 N 6 1 is a randomly generated number enough to reproduce accurately the effective com-
and tc0 = 0.5 GPa. These values are sensible for Al posite properties. The figures show the part played
alloys reinforced with brittle ceramic particles by the particle spatial distribution in the devel-
(Evans et al., 1999; Babout et al., 2004). opment of damage by interface decohesion. Homo-
Four tensile stress–strain curves were computed geneous microstructures (Fig. 7(a)) showed a
for each material (2 RVEs deformed along two minimum reduction in the strain hardening capacity
perpendicular cube axes) and they are plotted in as a result of the progressive interface fracture but
Fig. 7(a) and (b) for the homogeneous and clustered the effects of damage were evident in the very early
stages of the deformation in the clustered micro-
350 structure, Fig. 7(b), and led to a significant reduc-
tion of the flow stress (11% at a far-field strain
300 of 8%, while the decrease was only of 4% in the
homogeneous composite). In fact, the slight harden-
250 ing effect of clustering in the absence of damage
(Fig. 3) was reversed by the stress relaxation trig-
gered by damage, and the flow stress of the inhomo-
Stress (MPa)

200
geneous microstructure was lower than that of the
150
homogeneous one. It should be noted that the
reduction in composite strain hardening, although
not dramatic, precipitates the onset of plastic insta-
100 Homogeneous, ξ = 0.15
bility, which controls the strain-to-failure of these
No damage ductile-matrix composites (LLorca and González,
50
Interface decohesion 1998; Segurado and LLorca, 2005b). Thus, the drop
in ductility due to damage by interface decohesion
0 was very large: from  14% in the undamaged com-
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
(a) Strain
posites to 7.5 ± 0.2% in the homogeneous and
6.6 ± 0.5% in the clustered materials.
350 The numerical simulations of the RVE also pro-
vided information about the critical locations at
300

250
Stress (MPa)

200

150
Clustered
100 ξ = 0.15, ξcl = 0.37

No damage
50
Interface decohesion

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
(b) Strain Fig. 8. Deformed shape of the RVE face of a clustered material
(ncl = 0.37). The applied strain was 7.8% along the horizontal
Fig. 7. Effect of interface decohesion on the tensile stress–strain direction. The arrows indicate the locations of interface decohe-
curves. (a) Homogeneous particle distribution and (b) inhomo- sion. The detail of decohesion in one particle is shown in the right
geneous particle distribution. inset.
J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883 881

which damage developed earlier during deforma- particle spatial distribution, and they are relieved
tion. This is shown in Fig. 8, where the deformed by interfacial fracture. These patterns follow exper-
shape of a cube face of an RVE corresponding to imental observations in the literature (Whitehouse
the clustered microstructure is shown for an axial and Clyne, 1993; Babout et al., 2004).
strain of 7.8%. Decohesion began at the poles of The main consequences of interface fracture were
the spherical particles and grew along the interface, a reduction in the average stress carried by the
although no complete decohesions were observed. spherical reinforcements and an increment in the
Decohesion was localized between particles very volumetric strain as a result of the void formation.
close to each other and oriented along the loading These two variables can be easily computed from
axis within a conic angle of ±45°. Large stress con- the numerical simulations and their average values
centrations are generated at the interface by this are plotted in Fig. 9(a) and (b). The evolution of
eq
r p with the applied strain in Fig. 9(a) shows very
clearly the contrasting effect of particle clustering
600
and damage on the particle stresses. Clustering
without damage leads to an increase in the stress
500 carried by the particles and improves the composite
flow stress. Interface decohesion reduced the aver-
400
age stress in the particles, and eventually the load
borne by the reinforcements reached a maximum
and began to decrease. The reduction in r eq
p was
σpeq (MPa)

300 more pronounced in the clustered material and, as


a result, the average stress in the spherical reinforce-
200 ments was higher in the homogeneous material than
in the inhomogeneous if damage occurs. The onset
Homogeneous, no damage and evolution of damage is well described by the
100 Homogeneous, interface decohesion
Clustered, no damage
volumetric strain in Fig. 9(b), which shows that
Clustered, interface decohesion interface decohesion not only began earlier in the
0 clustered material but grew more rapidly, leading
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
(a) Strain

0.005
90

Homogeneous
0.004 Clustered
85
Elastic modulus (GPa)
Volumetric strain

0.003
80

0.002

75

0.001

70 Homogeneous, ξ = 0.15
0 Clustered, ξ = 0.15, ξcl = 0.37
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
(b) Strain 65
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Fig. 9. (a) Evolution of the volume-averaged equivalent stress in Strain
the spherical reinforcements, r eq
p , as a function of the applied
tensile strain. (b) Evolution of the volumetric strain as a function Fig. 10. Evolution of the composite elastic modulus as a function
of the applied tensile strain. of the applied tensile strain.
882 J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883

to the reduction in the effective properties depicted and 40%) was 3–6 times higher than in the homoge-
in Fig. 7. neous one. Damage by particle/matrix decohesion
Finally, the reduction in the interface stiffness was simulated by introducing interface elements at
due to damage led to a progressive decline in the the sphere surfaces. Damage by decohesion led to
composite elastic modulus, which was measured an important degradation of the composite strain
by carrying out periodic unloads during the simula- hardening capacity and stiffness, which was more
tion of the tensile deformation. The variation of the pronounced in the inhomogeneous microstructures,
average elastic modulus was computed from the where damage was nucleated earlier and progressed
four simulations carried out in each microstructure at higher rates, and induced a significant reduction
and the results are plotted in Fig. 10 for the dam- of the strain at the onset of plastic instability, which
aged composites with homogeneous and clustered controls the tensile ductility. These simulations
microstructures. The inhomogeneous composite showed that damage (either by particle or interface
was slightly stiffer at the beginning of the deforma- decohesion) was preferentially nucleated between
tion but the higher damage rates reversed the situa- particles very close to each other and oriented along
tion very rapidly, showing again the higher damage the loading axis within a conic angle of ±45°, in
rates in the clustered microstructure. agreement with the experimental patterns of inter-
face decohesion.
4. Conclusions
Acknowledgments
The effect of the reinforcement spatial distribu-
tion on the mechanical behavior was studied in a The financial support from the Spanish Ministry
model composite material made up of an elasto-plas- of Education and Science through grants MAT
tic matrix reinforced with 15 vol.% of elastic, stiff 2003-5202-C02 and MAT 10589E is gratefully
spherical particles. The analyses were carried out acknowledged.
within the framework of computational microme-
chanics, and the composite behavior was obtained References
from the finite element simulations of cubic RVEs
containing several dozens of particles with periodic Abaqus, 2004. UsersÕ Manual. Hibbitt, Karlsson, and Sorensen,
boundary conditions. Composites with four different Inc.
Babout, L., Maire, E., Fougeres, R., 2004. Damage initiation in
random and isotropic particle spatial distributions model metallic materials: X-ray tomography and modelling.
were considered: a homogeneous one and three het- Acta Materialia 52, 2055–2063.
erogeneous, where the particles were clustered in Benveniste, Y., 1987. A new approach to the application of
spherical regions with increasing local reinforcement Mori–TanakaÕs theory in composite materials. Mechanics of
volume fractions ncl = 20%, 30% and 40%. It was Materials 6, 147–157.
Böhm, H.J., 2004. Continuum models for the thermomechanical
found that the particle spatial distribution had a very behavior of discontinuously reinforced materials. Advanced
weak effect on the effective properties in the elastic Engineering Materials 6, 626–633.
and plastic regimes. Maximum values of the stiffness Böhm, H., Han, W., 2001. Comparisons between three-dimen-
and strain hardening rates were found in the com- sional and two-dimensional statistics-based unit cell models
posites with an intermediate degree of clustering for particle-reinforced MMCs. Modeling and Simulation in
Materials Science and Engineering 9, 47–65.
(ncl = 30%) but the differences with the homoge- Böhm, H.J., Han, W., Eckschlager, A., 2004. Multi-inclusion unit
neous one were almost negligible. However, the local cell studies of reinforcement stresses and particle failure in
stresses in the spherical reinforcements were notably discontinuously reinforced ductile matrix composites. Com-
higher in the clustered microstructures, and this con- puter Modeling in Engineering & Sciences 5, 5–20.
Cantwell, W.J., Roulin-Moloney, A.C., 1989. Fractography and
siderably affected damage.
failure mechanisms of unfilled and particulate filled epoxy
A damage indicator of particle fracture was cre- resins. In: Fractography and Failure Mechanisms of Unfilled
ated using the local values of the maximum princi- and Particulate Filled Epoxy Resins. pp. 234–289.
pal stress in the spherical reinforcements and the Christensen, R.M., Lo, K.H., 1979. Solutions for effective shear
Weibull model for fracture of brittle materials. It properties of three phase sphere and cylinder models. Journal
was found the even small amounts of inhomogene- of the Mechanics and Physics of Solids 27, 315–330.
Drugan, W.J., Willis, J.R., 1996. A micromechanics-based
ity in the particle spatial distribution increased nonlocal constitutive equation and estimates of the represen-
significantly the particle fracture probability, which tative volume element size for elastic composites. Journal of
in very inhomogeneous microstructures (ncl = 30% the Mechanics and Physics of Solids 44, 497–524.
J. Segurado, J. LLorca / Mechanics of Materials 38 (2006) 873–883 883

Evans, A.G., Hutchinson, J., Wei, Y., 1999. Interface adhesion: Nemat-Nasser, S., Hori, M., 1999. Micromechanics: Overall
effects of plasticity and segregation. Acta Materialia 47, 4093– Properties of Heterogeneous Materials. North-Holland.
4113. Poza, P., LLorca, J., 1996. Fracture toughness and fracture
González, C., LLorca, J., 2000. A self-consistent approach to the mechanisms of Al/Al2O3 composites at cryogenic and ele-
elasto-plastic behavior of two-phase materials including vated temperature. Materials Science and Engineering A 206,
damage. Journal of the Mechanics and Physics of Solids 48, 183–193.
675–692. Prasad, V.V.B., Bhat, B.V.R., Mahajan, Y.R., Ramakrishnan,
González, C., Segurado, J., LLorca, J., 2004. Numerical simu- P., 2002. Structure–property correlation in discontinuously
lation of elasto-plastic deformation of composites: evolution reinforced aluminium matrix composites as a function of the
of stress microfields and implications for homogenization relative particle size ratio. Materials Science and Engineering
models. Journal of the Mechanics and Physics of Solids 52, A 337, 179–186.
1573–1593. Rintoul, M.D., Torquato, S., 1997. Reconstruction of the
Gusev, A.A., 1997. Representative volume element size for elastic structure of dispersions. Journal of Colloid and Interface
composites: a numerical study. Journal of the Mechanics and Science 186, 467–476.
Physics of Solids 45, 1449–1459. Segurado, J., LLorca, J., 2002. A numerical approximation to the
Lewandowski, J.J., Liu, C., Hunt, W.H., 1989. Effects of matrix elastic properties of sphere-reinforced composites. Journal of
microstructure and particle distribution on fracture of an the Mechanics and Physics of Solids 50, 2107–2121.
aluminum metal matrix composite. Materials Science and Segurado, J., LLorca, J., 2004. A new three-dimensional interface
Engineering A 107, 241–255. finite element to simulate fracture in composites. Interna-
Lewis, C.A., Withers, P.J., 1995. Weibull modelling of particle tional Journal of Solids and Structures 41, 2977–2993.
cracking in metal matrix composites. Acta Metallurgica et Segurado, J., LLorca, J., 2005a. Unpublished results.
Materialia 43, 3685–3699. Segurado, J., LLorca, J., 2005b. A computational micromechan-
LLorca, J., 1995. An analysis of the influence of reinforcement ics study of the effect of interface decohesion on the
fracture on the strength of discontinuously-reinforced metal– mechanical behavior of composites. Acta Materialia 53,
matrix composite. Acta Metallurgica et Materialia 43, 181–192. 4931–4942.
LLorca, J., González, C., 1998. Microstructural factors control- Segurado, J., LLorca, J., González, C., 2002. On the accuracy of
ling the strength and ductility of particle-reinforced metal– mean-field approaches to simulate the plastic deformation of
matrix composites. Journal of the Mechanics and Physics of composites. Scripta Materialia 46, 525–529.
Solids 46, 1–28. Segurado, J., González, C., LLorca, J., 2003. A numerical
LLorca, J., Segurado, J., 2004. Three-dimensional multiparticle investigation of the effect of particle clustering on the
cell simulations of deformation and damage in sphere- mechanical properties of composites. Acta Materialia 51,
reinforced composites. Materials Science and Engineering A 2355–2369.
365, 267–274. Spowart, J.E., Ma, Z.-Y., Mishra, R.S., 2003. The effect of
LLorca, J., Martı́n, A., Ruiz, J., Elices, M., 1993. Particulate friction stir processing on the spatial heterogeneity of
fracture during deformation in a spray-formed metal–matrix discontinuously-reinforced aluminum (DRA) microstruc-
composite. Metallurgical Transactions A 24, 1575–1588. tures. In: Friction Stir Welding and Processing.
Michel, J.C., Moulinec, H., Suquet, P., 1999. Effective properties Torquato, S., 1998. Effective stiffness tensor of composite media:
of composite materials with periodic microstructure: a com- II. Applications to isotropic dispersions. Journal of the
putational approach. Computational Methods in Applied Mechanics and Physics of Solids 46, 1411–1440.
Mechanical Engineering 172, 109–143. Torquato, S., 2001. Random Heterogeneous Materials. Springer.
Mori, T., Tanaka, K., 1973. Average stress in the matrix and Whitehouse, A.F., Clyne, T.W., 1993. Cavity formation during
average elastic energy of materials with misfitting inclusions. tensile straining of particulate and short fiber metal matrix
Acta Metallurgica et Materialia 21, 571–574. composites. Acta Metallurgica et Materialia 41, 1701–1711.

You might also like