Вы находитесь на странице: 1из 13

M. S.

Bruzzi
P. E. McHugh
Micromechanics Research Unit,
Department of Mechanical Engineering,
National University of Ireland, Galway, Ireland
Micromechanical Modeling of the
Static Loading of an Al 359-SiC
MMC
The objective of this work was to use micromechanical nite element models to simulate
the static mechanical behavior of a metal matrix composite: a cast Al 359 alloy reinforced
with 20% SiC particles, at two different temperatures: room temperature and 150C. In
the simulations, periodic unit cell models incorporating the explicit representation of the
matrix, reinforcing particles and precipitated primary silicon crystals in both 2D and 3D
were used. Micromechanical models with both idealized and realistic reinforcing particle
geometries and distributions were generated. The realistic particle geometries and distri-
butions were inferred from experimental SEM micrographs. The pattern and intensity of
the plastic deformation within the matrix was studied and the macroscale behavior of the
composite was inferred from average stress and strain values. In order to include the
effects of residual stresses due to the processing of the material, a quenching simulation
was performed, prior to mechanical loading, and its effects on the macroscopic and
microscopic behavior of the MMC was assessed. The effects of introducing the damage
mechanisms of ductile void growth and brittle failure of the reinforcing particles was also
investigated. The results of the simulations were compared with experimental results for
the MMC in terms of macroscopic tensile stressstrain curves and conclusions were
drawn. DOI: 10.1115/1.1839217
Introduction
This work is concerned with the study of the static tensile load-
ing of a particulate reinforced Al 35920% SiC metal matrix
composite material MMC. The overall objective of this work is
to apply a micromechanical unit cell modelling approach to inves-
tigate microstructural deformation in the MMC and to predict its
overall tensile stressstrain behavior. It is also an objective to
assess the effects of damage caused by void growth and particle
fracture, on the uniaxial stressstrain behavior of the composite.
The material under investigation in this work is a cast Al 359
T6 alloy reinforced with 20% by volume of SiC particles. As a
result of the processing of this material 1,2, porosity existed
within the MMC with a determined volume fraction of 0.96% 1.
Two classes of pores were identied: large blow-holes, which
contained entrapped gas 0.5% and inderdendritic shrinkage due
to contraction during solidication 0.46%. The SiC particles
were observed in the metallographic analysis of Forn et al. 1,2
to be fragmented in nature with a mean area of 29.4 m
2
and a
mean aspect ratio of 1.5. A micrograph indicating the typical mi-
crostructure is shown in Fig. 1. Other micrographs showing the
fragmented nature of the particles are presented in 2. It was also
observed from the microstructural investigation of Forn et al. 1
that large primary silicon Si crystals with an area fraction of
6.67% were precipitated into the microstructure during process-
ing. The Si crystals were slightly smaller than the SiC particles,
with a mean area of 17 m
2
, and existed as primary crystals. This
can be observed in Fig. 1. It is also worth noting that the area of
the SiC particles had a standard deviation of 514 m
2
, indicating
a wide distribution of large and small SiC particles. Many of these
particles were fragmented or cracked. The reinforcing particles
were observed to exist in clusters mingled with shrinkage poros-
ity, and the Al 359 matrix existed as large unreinforced areas
between these particle clusters. The blow-holes tended to exist
within the large areas of an unreinforced matrix.
As a result of the particle clustering and porosity within the Al
359 MMC the material exhibited a brittle behavior under static
tensile loading conditions. There was also very little change in this
brittle behavior with an increase in temperature of up to 200C
i.e., there was only a reduction in the tensile strength of up to
15% in going from room temperature to 200C) 2. This ob-
served brittle tensile behavior, along with an analysis of the frac-
ture surfaces of specimens loaded to failure 2,3, indicated that
failure was initiated by brittle facture within the particle clusters.
Particle cracking and possibly some interfacial failure were con-
cluded to be the dominant damage mechanisms in the Al 359
MMC.
With regards to the residual stress state within the MMC prior
to loading, it was found by means of neutron diffraction that the
macrostresses, within the specimens analyzed, were negligible
Fiori et al. 4 and Carrado et al. 5. The internal residual
stresses within the microstructure were microstresses alone. These
stresses were directly caused by the mismatch of thermal expan-
sion coefcients between the aluminum matrix and the reinforcing
SiC particles and also the precipitated primary Si during pro-
cessing. The particles Si and SiC were found to be in a net state
of compression with an average hydrostatic stress of 400 MPa,
and the matrix was found to exist in a net state of tension with an
average hydrostatic stress of 100 MPa. It is worth noting, how-
ever, that the experimental error in these measurements is esti-
mated to be of the order of tens of MPa.
In order to conrm that the stresses were mainly internal mi-
crostresses, the Al 359 MMC specimens were loaded to a strain to
0.2%, then unloaded, and the internal stresses were then measured
in 4. The internal microstresses were found to be lower in both
the matrix and the reinforcing particles. The particles had a re-
duced stress of 150 MPa and the matrix was found to have a
tensile stress of 20 MPa. This can be explained by the existence of
local plasticity induced within the matrix, around the reinforcing
particles, and hence the relaxation of the compressive stress
within the particles and the tensile stresses within the matrix.
Manuscript received April 14, 2004; revision received June 29, 2004. Review
conducted by: A. Pelegri.
106 Vol. 127, JANUARY 2005 Copyright 2005 by ASME Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
In this paper, a brief overview of the modeling of MMCs is rst
given, followed by a description of the periodic unit cell modeling
approach. The modeling of the Al 35920% SiC MMC is then
outlined. Models were developed based on both idealized and
realistic microstructural geometries, mostly in 2D but also in 3D,
and were used to attempt to predict tensile mechanical behavior, at
both room temperature and 150C. This higher temperature is
reexive of the operating conditions the material might experience
in an automotive application, for example. Following this, the
models were used to assess the effects of the introduction of re-
sidual stresses from processing, particle fracture and matrix fail-
ure on the predicted performance. In the nal section, the results
are discussed and conclusions made. It is important to emphasize
that the modeling was focused on predicting the constitutive be-
havior of the material prior to failure, i.e., the stiffening or soft-
ening effects of the various microstructural features present. The
models were not used to attempt to predict the macroscopic brittle
fracture point of the material.
Periodic Unit Cell Models
MMCs have been the subject of intense analytical and experi-
mental investigation over the last 15 to 20 years, to characterize
and understand their mechanical properties and performance, and
to assess the importance of variables such as reinforcement vol-
ume fraction and morphology. This is true for MMCs with various
types of reinforcement structures including particulate, short bre
and continuous bre reinforcements. As regards modelling of me-
chanical performance, both analytical and computational methods
have been used, including computational micromechanics, which
is based on explicitly representing the material microstructure in
computational models. Due to the fact that the complete analysis
of an MMC microstrucutre, in which the matrix and reinforcement
are represented separately, with realistic reinforcement geometries
and distribution, is still beyond reasonable computational capabili-
ties, many micromechanical descriptions of composites are based
on periodic unit cell approaches. Excellent reviews of investiga-
tions into these aspects of MMCs are to be found in Suresh et al.
6, Bohm 7,8 and Llorca 9.
In the periodic unit cell approach, the microstructural geometry
is divided into periodically repeating unit cells, to which the in-
vestigation may be limited without greatly reducing its generality.
For discontinuously reinforced MMCs a large body of work exists
in the literature showing how the periodic unit cell methodology
has been very successfully used to study the room temperature,
monotonic tensile loading behavior of the materials, based on
models with idealized and regular microstructural geometries, e.g.
McHugh et al. 10, Christman et al. 11, Bao et al. 12 and
Weissenbek et al. 13. Models of this type have also been used by
Llorca et al. 14, Bohm 15 and Bruzzi et al. 16, to study cy-
clic loading, and by Li and Ramesh 17 to study compressive
loading.
Models with realistic geometry, based on micrographs of the
material microstructure, have been developed and have proved
useful for quantifying local phenomena in MMCs e.g., Fis-
chmeister and Karlsson 18, Brockenbrough and Suresh 19,
Bruzzi et al. 16, Li et al. 20. An alternative approach to
achieving realistic microstructural geometries is to use the
Voronoi cell method, as implemented by Ghosh et al. 21 and
Ghosh and Moorthy 22.
For discontinuously reinforced materials, 3D unit cell models
are currently seeing increased utilization due to continuing im-
provements in computer performance. Both idealized and realistic
microstructural geometries have been used; for the latter the em-
phasis has been placed on including as many particles/short bers
within the unit cell as possible and achieving realistic reinforce-
ment distribution patterns e.g., Gusev 23, Watt et al. 24,
Bohm et al. 25, Bohm and Han 26, Hine et al. 27.
Material damage and microfracture has also been explicitly rep-
resented in periodic unit cell models. This by denition means that
the damage is itself periodic throughout the material. Some of the
most recent work has focused on 3D analyses using regular rein-
forcement geometries and realistic reinforcement distribution pat-
terns where reinforcement fracture e.g., Gall et al. 28, Eck-
schlager et al. 29, Seguardo et al. 30 and matrix void growth
and coalescence has been modeled e.g., Llorca and Segurado
31.
The majority of the MMC micromechanical modeling work
presented in the literature has been focused on representing mate-
rials with relatively well-dened two-phase microstructures. In
contrast to this, the focus in this work is on a cast MMC that has
a relatively complex multi-phase microstructure and signicant
porosity. It is of interest to examine how the periodic unit cell
modeling approach performs in this, somewhat different, context.
The emphasis of the work is not an exhaustive geometrical param-
eter study in 2D or 3D, as many previous studies have been, but
the development of periodic unit cell models incorporating many
realistic features of the Al 359 MMC, that attempt to predict their
macroscopic and microscopic behavior.
The Patran code was used for mesh generation and ABAQUS/
Standard was used as the nite element code for the simulations.
Tensile loading was simulated at both room temperature and at
150C. In 2D the simulations were made progressively more re-
alistic, starting with a 2D generalized plane strain model with
idealized microstructural geometries, followed by 2D models with
realistic geometries based on the actual SEM micrographs of
the material. A simple 3D model was also generated. Residual
stresses, due to the thermal expansion mismatch of the constitu-
ents during processing, were included into the models by simulat-
ing a quenching process. The microstresses within the microstruc-
ture were then compared to the values obtained by the neutron
diffraction experiments of 4,5.
Primary crystals of Si were included into both the idealistic and
realistic models for the Al 35920% SiC, due to the fact that the
Si was present in signicant amounts. The inuence of void
growth and particle cracking was also investigated. It was not the
purpose of this investigation, however, to model the details of the
fracture process of the MMC, but only to investigate the effects of
the inclusion of these types of damage on the macroscale tensile
stressstrain behavior of the composite.
Fig. 1 Micrograph of clustered SiC and primary Si crystals
within the microstructure of the Al 35920% SiC. Used with
the permission of A. Forn, UPC, Spain
Journal of Engineering Materials and Technology JANUARY 2005, Vol. 127 107
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
Model Description
The Al 359 MMC, under investigation in this work, was mod-
eled in both 2D and 3D. In both cases each MMC was initially
idealized as being composed of a periodic array of unit cells
square in 2D and cubic in 3D, each cell containing elastic par-
ticles and an elastic-plastic matrix. The interface between the ma-
trix and the reinforcing particles was assumed to be perfectly
bonded. Symmetry boundary conditions were assumed implying
that during tensile deformation the 2D unit cell was constrained to
remain rectangular and the 3D unit cell constrained to remain
orthorhombic. Consistent with these boundary conditions, the unit
cells themselves had internal symmetries.
In 2D the cell was composed of four square quadrants, each
quadrant being identical to the reection of its neighbors about
horizontal and vertical lines X-X and Y-Y, through the center of
the unit cell, as shown in Fig. 2a. Similarly in 3D, the cell was
composed of eight cubes, each being identical to the reection of
its neighbors in horizontal and vertical planes through the center
of the cube. Given these assumed symmetries, it was sufcient to
apply the nite element method to one quadrant of the unit cell in
the 2D case and one octant of the unit cell in the 3D case. The
boundary conditions imposed for the 2D case, shown in Fig. 2b,
were as follows: the bottom and left edge of the unit cell quadrant
were constrained to have zero displacements in the x
2
and x
1
direction, respectively, the top edge had a uniformly prescribed
displacement, U
2
and the right edge had a uniform solution de-
pendent displacement, U
1
.
Similarly in 3D, the left hand face, the bottom face and the back
face of the unit cell octant were constrained to have zero displace-
ments in the x
1
, x
2
and x
3
directions, respectively, where x
3
is the
out of plane axis. A prescribed uniform displacement, U
2
, was
applied to the top surface, and the right face and the front face
had uniform solution dependent displacements, U
1
and U
3
,
respectively.
Tensile loading in the vertical direction was achieved by apply-
ing uniform vertical displacements, U
2
, on the top edge of the 2D
quadrant and the top surface of the 3D octant, respectively. The
applied cell boundary displacements determined the average ten-
sile engineering strain,
22
ave
, for the material. The boundary reac-
tion forces were averaged over the top boundary to calculate the
average tensile engineering stress,
22
ave
, for the material.
In the 2D models, generalized plane strain deformation, which
allows uniform strain in the out of plane direction, i.e.,
33
A
where A is uniform across the x
1
x
2
plane, was assumed. In the
2D simulations, 4-noded, with full integration, or 8-noded, with
reduced integration, quadrilateral generalized plane strain ele-
ments were used to model the matrix. 3-noded or 6-noded trian-
gular generalized plane strain elements were used to mesh the
particles. In the 3D simulations, 8-noded linear or 20-noded qua-
dratic hexahedron elements were used to model both the matrix
and the particles.
Initially, three models were generated for the cast Al 359
20% SiC MMC; two 2D models and one in 3D, as shown in
Figs. 3a3c. This was done to assess the performance of mod-
els with different particle geometries and distributions and also to
provide a basis for assessing the effects of the residual stresses
and damage mechanisms to be discussed in the following sections.
Each model explicitly represented the 20% by volume of reinforc-
ing SiC particles, 6.67% by volume of the precipitated primary Si
crystals, and pores with a volume fraction equal to that of the
overall porosity experimentally determined as 0.96% in the Al
359 MMC.
The 2D-Ideal Geometry model, a quadrant of which is shown in
Fig. 3a, represented a simplistic 2D model where the SiC and Si
particles had idealized geometries. 940, 8-noded quadrilateral
generalized plane strain elements with reduced integration were
used to mesh this model. Figure 3b shows a quadrant of the
more complex 2D-Real Geometry model where realistic particle
geometries were represented, and the particles existed within
bands or clusters. This particular model was generated using the
particle geometry, orientation and distribution data of Forn 2.
The 2340, 8-noded quadrilateral generalized plane strain elements
using reduced integration were used to mesh this model. The two
different types of porosity found experimentally were included in
the models in a straightforward manner, i.e. the large blow-hole
was included within the large central matrix region, and shrinkage
porosity was included by including pores between the particles,
within the particle clusters. For simplicity, in each of the 2D mod-
els, the pores were assumed to be hexagonal in shape.
In 3D, one model, an octant of which is shown in Fig. 3c, was
generated with idealized particle geometries. In this model, one
particle represented the 20% SiC reinforcement and one particle
represented the precipitated primary Si crystals. A pore of volume
fraction 0.96% was also included. This model was meshed using
1260, 8-noded hexahedron brick elements.
Fig. 2 a 2D unit cell, containing a square reinforcing particle, with vertical and horizontal
lines of symmetry; b one quadrant of the 2D unit cell: sufcient for modeling
108 Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
Within the models, the SiC and Si particles were assumed to be
elastic, and the Al 359 alloy was assumed to be elastic-plastic. The
elastic constants for SiC and Si were obtained from literature,
Suresh et al. 6. The elastic constants and the plastic strain
hardening properties of the matrix were determined from the cal-
culated
true

true
curve obtained from the tensile test data of
Forn et al. 2 for the unreinforced Al 359 T6 alloy. This was done
for both room temperature and 150C.
A displacement U
2
, corresponding to the experimental fracture
strain of
22
ave
0.3%, was applied to the top boundary of the unit
cell quadrant, in order to apply a tensile load. This strain level was
imposed for both room temperature and 150C, since the fracture
strain was found to be the same in both cases.
Results
Contour plots of equivalent plastic strain for the quadrants of
the 2D-Ideal Geometry model and the 2D-Real Geometry model
at room temperature are shown in Figs. 4a and 4b, respec-
tively. It is clear that these contour plots show very little plastic
deformation within the matrix, due to the very low applied strain.
The plastic deformation that does occur is localized in bands ori-
ented at approximately 45 deg to the loading direction, as would
be expected. The strain localizations occur within the particle
clusters in the 2D-Real Geometry model, and also at the corners
of the hexagonally shaped pores, for both models. Similar results
were found when modeling the MMC at 150C.
The macroscale stressstrain (
22
ave

22
ave
) curves for the mod-
els are compared to the experimental stressstrain curve for the Al
359 MMC, at both room temperature and 150C, in Figs. 5a and
5b, respectively. It is interesting to note the very accurate mate-
rial stiffness prediction of the 2D models, especially at 150C,
and the relative over-prediction of the material stiffness by the 3D
model, which may be due partly to an increased level of geometri-
cal constraint on plasticity in 3D.
Fig. 3 a Unit cell quadrant with idealized particle geometries and particle distribution for the Al 359 MMC 2D-Ideal Geometry
model. Light gray particleSiC; dark gray particleSi, hexagonal pore included. b Unit cell quadrant with realistic particle
geometries and particle distribution for the Al 359 MMC 2D-Real Geometry model. Light gray particlesSiC, dark gray particles
Si, hexagonal pores included. c Unit cell octant with realistic particle geometries for the Al 359 MMC 3D-Real Geometry model
Journal of Engineering Materials and Technology JANUARY 2005, Vol. 127 109
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
As regards temperature dependence, it can be seen from the
experimental stressstrain curves in Fig. 5 that the stiffening ef-
fect of the reinforcement i.e., the fractional increase in stress
levels for the MMC relative to the unreinforced matrix is smaller
at the lower temperature in comparison to the higher temperature.
Part of the reason for this could be the more rapid softening of the
matrix due to both elastic softening and an earlier onset of plas-
ticity in comparison to the reinforcement as the temperature in-
creases. It is clear that this is only part of the reason since the
models are capable of accounting for these component stiffness
changes and Fig. 5 shows that the predicted macroscopic tempera-
ture dependence is not a strong as that observed experimentally.
It can be seen from Fig. 5 that there was very little difference
between the predictions of the 2D-Ideal Geometry model and the
2D-Real Geometry model, both at room temperature and at
150C. As a result, no signicant effect of reinforcement particle
clustering on the overall static mechanical behavior of the MMC
could be seen. This questions the benets of generating complex
realistic 2D models where computational costs can be quite high.
Fig. 4 Contour plot of equivalent plastic strain for the a 2D-Ideal Geometry
and b 2D-Real Geometry models for the Al 359 MMC at room temperature with

22
ave
0.3%
110 Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
Inclusion of Residual Stresses
As mentioned in the Introduction, there existed internal residual
microstresses within the Al 359 MMC, generated as a result of the
mismatch of coefcients of thermal expansion of the Al matrix
and the SiC and Si particles during its thermal processing. These
residual microstresses were included in the models by simulating
a quenching process from 530C casting temperature to room
temperature, before mechanical loading was applied. In each of
the models, thermal expansion coefcients of 22.510
6
C
1
for the Al 359 matrix, and 4.710
6
C
1
were used for the SiC
and Si particles. Contour plots of hydrostatic pressure, (
kk
,
the negative of hydrostatic stress and equivalent plastic strain, for
the 2D-Real Geometry model, when quenched to room tempera-
ture, are shown in Figs. 6a and 6b, respectively. It is clear that
after quenching the matrix exists in a net state of tension, and the
SiC and Si particles exist in a state of compression, as would be
expected. In addition, the plastic strain is localized in the matrix
within particle clusters. The peak level of effective plastic strain
of 6.9110
2
is much higher than those produced when the
model was strained to the fracture strain of 0.3%, without the
inclusion of residual stresses, as shown in Fig. 4b.
In order to compare the results obtained by this quenching
simulation with the experimental values obtained by neutron dif-
fraction as described above, the hydrostatic stress,
kk
, was av-
eraged for the Al 359 matrix and for all the SiC and Si particles
for the 2D-Real Geometry model. The comparison is shown for
the room temperature results in Table 1. It is encouraging to see
that the predicted stresses are in the right order of magnitude
although the levels are lower than those measured experimentally.
It should be noted however that the experimental error in the
neutron diffraction experiments was estimated to be in the tens of
MPa.
Also shown in Table 1 is the comparison with the experimental
values of the residual microstresses after a pre-load of 0.2% strain
was applied and removed. As expected, there was a net reduction
in the residual stresses within the material after application of the
pre-load. The tensile stress in the Al matrix decreased with the
increased plastic strain that occurred during loading. Also the high
compressive residual stresses in the reinforcing particles de-
creased with the decreased constraint of the surrounding matrix.
As shown in Table 1, the model predicted the correct trend in
stress reduction but underpredicted the magnitude of the reduc-
tion, probably due to the lack of representation of microscale
damage mechanisms, that would be active given that the applied
pre-strain of 0.2% is so close to the fracture strain of 0.3%, and
that would result in microscale stress relaxation.
Following the quenching simulation, as described above, a uni-
form strain,
22
ave
0.3%, was applied to each model. Contour
plots of equivalent plastic strain for the 2D-Ideal Geometry and
the 2D-Real Geometry models are shown in Figs. 7a and 7b,
respectively. It can be seen that the equivalent plastic strain levels
are signicantly higher than those shown in Fig. 4 due to the
residual stresses from the thermal processing. In Fig. 7b it is
clear that there is signicant plastic strain localization within the
particle clusters. While no damage mechanisms were included in
these simulations it is reasonable to assume that, as indicated by
the plastic strain intensity, fracture would take place within these
particle clusters.
Upon examination of the macroscale stressstrain curves pre-
dicted by the models for the Al 359 MMC in Figs. 8a and 8b,
it is seen that the inclusion of residual stresses due to the thermal
processing has the effect of reducing the predicted stiffnesses. A
more accurate prediction of the MMC behavior is obtained at
room temperature, whereas the model predictions under-estimates
the MMC tensile behavior at 150C. It is also seen that the inclu-
sion of residual stresses into the 3D model causes a greater reduc-
tion in the stiffness of the model predictions than in 2D. This is
true both at room temperature and at 150C, and is attributed to
the increased effective plastic strain of the matrix surrounding the
particle in all three dimensions, as opposed to only two dimen-
sions for the 2D generalized plane strain models.
Effect of Matrix Void Growth
The Gurson void growth model Gurson 32,33, HKS 34,35
was implemented into the models in order to investigate the effect
of introducing matrix damage by void growth. This is a continuum
damage model for ductile metals. In this case, it takes the form of
an extension of J
2
ow theory, where the relationship between
stress and strain at a material point is modied to represent, in a
continuum sense, the effects of the growth of microvoids in the
region of the material point. The stressstrain relationship in the
form of a plastic potential was rst developed analytically by
Gurson 32,33. As input, in addition to the normal elastic-plastic
material properties, the model requires an initial local microvoid
volume fraction. The initial void volume fraction introduced into
the model was the experimentally determined overall porosity of
0.96%. As before, a quenching simulation was rst performed in
order to introduce internal residual stresses due to the mismatch of
thermal expansion coefcients between the Al 359 matrix and SiC
and Si particles, as discussed in the last section. The models were
then loaded to a strain of 0.3%, as before.
Contour plots of the hydrostatic pressure,
kk
, and the void
volume fraction are shown in Figs. 9a and 9b, respectively, for
Fig. 5 Macroscale engineering stressstrain curves for the Al
359 MMC models: a room temperature and b 150C
Journal of Engineering Materials and Technology JANUARY 2005, Vol. 127 111
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
the 2D-Real Geometry model. The void growth mainly occurs in
regions of high hydrostatic stress within the matrix and also
around the pores within the model. These regions of high hydro-
static matrix stress exist predominantly within the particle clus-
ters, due to the plastic constraint caused by the particles. It is
interesting to note, by comparing Fig. 9 and Fig. 7b, that both
Fig. 6 Contour plots, immediately after the quenching simulation, for the 2D-Real
Geometry model: a hydrostatic pressure and b equivalent plastic strain
Table 1 A comparison of average internal hydrostatic stresses
at room temperature for the 2D-Real Geometry model
Al 35920% SiC
2D-Real Geometry
Model MPa
Experimental
MPa
Al 359 80 100
(SiCSi) 257 400
Al 3590.2% Strain 71 20
(SiCSi)0.2% Strain 232 150
112 Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
the areas of signicant matrix void growth and localized plastic
strain do not necessarily coincide, even though both are predomi-
nantly located within the particle clusters.
The effect of introducing damage, by means of void growth, is
a reduction in the macroscale stress level predictions of the mod-
els, as is shown in Fig. 10 for room temperature conditions. This
reduction increases with strain due to progressive matrix void
growth, and results in an under-prediction of the material stiffness
relative to the experimental data. It is interesting to note how the
three models coincide so closely in their predictions. If one com-
pares Fig. 10 with Fig. 5a and Fig. 8a, it can be seen that the
3D model predictions depend strongly on the presence of the re-
sidual stresses and weakly on the matrix void growth, whereas the
predictions of both 2D models depend almost equally on the re-
sidual stresses and matrix void growth.
No explicit material separation criteria were introduced into the
Fig. 7 Contour plots of equivalent plastic strain for the a 2D-Ideal Geometry and b
2D-Real Geometry Al 359 MMC models with
22
ave
0.3%
Journal of Engineering Materials and Technology JANUARY 2005, Vol. 127 113
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
simulations in addition to the basic Gurson model available in
ABAQUS/Standard, but elements with high void volume fractions
implicitly fail within the Gurson theory. This highlights a weak-
ness of the periodic cell modeling approach in that when one
element fails within a unit cell quadrant, it implicitly means that
there is failure in each cell within the material. This is statistically
unrealistic for a real material and contributes to the under-
prediction of the macroscopic stress levels that are observed.
However, including the void growth damage mechanism is still
useful in terms of investigating in what locations and under what
conditions damage by void growth would occur, and what effect
this would have on the macroscale tensile behavior of the Al 359
MMC. On this basis, it is clear that the models do predict micros-
cale failure at overall strains on the order of the measured experi-
mental fracture strain.
Effect of Particle Fracture
Particle fracture was the most dominant failure mechanism ob-
served experimentally for this material, as described in 3,36.
Also, many of the particles were cracked due to the processing of
the MMC, before loading. In order to investigate the effects of
fracturing particles on the predicted tensile behavior of the Al 359
MMC, aws were introduced into some of the larger SiC particles
in the 2D-Real Geometry model. The nodes ahead of these aws
within the reinforcing SiC particles were allowed to separate
along a predened path. These nodes only separate once a speci-
ed critical value of the maximum principle stress is reached at a
specied distance ahead of the aw tip. No data was available on
the SiC material that would provide a denite physical basis for
the selection of the initial aw size, critical stress and critical
distance ahead of the aw tip. What was done was to arbitrarily
choose reasonable values for these parameters so that particle
fracture would occur before the maximum applied strain of 0.3%
was reached. An initial aw size of 10% of the particle size and a
critical stress of 260 MPa at a critical distance of 1% of the mean
particle length 7 m were used. This choice allowed the inves-
tigation of the effects of particle fracture on the macroscale tensile
stressstrain behavior of the models.
The following point is worth emphasizing. One could argue that
careful microscale examination of the particles should have been
performed to give an indication on typical aw sizes, with a view
to improving the physical basis of the model. This, however,
would have been futile since all three model parameters aw
size, critical stress and critical distance would need to be explic-
itly determined from examination/testing of the material to de-
nitely specify the separation criterion. For example, a physically
determined aw size and critical stress could have been deter-
mined if it were possible to do so but the macroscopic load at
which the separation would occur particle would fracture would
still be arbitrary based on an arbitrary choice of the critical
distance.
A quench was simulated in order to introduce internal residual
stresses due to the processing of the MMC, as before. Also a strain
of 0.3% was applied to the unit cell quadrant, as described for the
models without the cracked particles. Contour plots of the effec-
tive plastic strain for two of the SiC fracturing particles in the
2D-Real Geometry model are shown in Figs. 11a11c. The
initial aw, which was assumed to be perfectly sharp, existed at
the surface of each particle before loading was applied. The initial
crack size was arbitrarily chosen to be 10% of the particle size as
measured along the predened crack path. This predened crack
path is shown as a line through the two large particles in Fig.
11a; the initial aw was on the right hand side of the left particle
and the left hand side of the right particle. During loading, the
aw causes a stress concentration within the particle and so the
critical stress is eventually reached, causing the crack to begin to
grow. Once started, the crack propagates rapidly, resulting in a
complete fracture of the particle due to the increasing stress inten-
sity associated with the lengthening crack. Figure 11b shows the
crack during this short propagation stage, and Fig. 11c shows the
complete fracture of the reinforcing particle at the maximum ap-
plied strain of 0.3%. It is seen by comparing Fig. 11c to the
contour plot of Fig. 7b that there is a large increase in the
equivalent plastic strain. The peak values of equivalent plastic
strain increase from 7.7310
2
to 3.3210
1
with the introduc-
tion of the awed and subsequently fractured particles.
In order to compare this fracturing process with the case where
the particles were initially fully cracked, the cracks were intro-
duced into the 2D-Real Geometry model in the same locations as
the pre-dened fracture paths, described above. These cracked
particles effectively meant that the particle behaved as two sepa-
rate entities with no matrix in between. A contour plot of the
equivalent plastic strain for a quadrant of the 2D-Real Geometry
model with cracked particles is shown in Fig. 12 at the maximum
applied strain of 0.3%. The peak value of equivalent plastic strain
of 3.3210
1
was the same as that shown in Fig. 11c, as ex-
pected, since all the relevant particles have fully cracked in both
the models at that strain.
The effects of fracturing particles and cracked particles on the
macroscale stressstrain behavior of the 2D-Real Geometry
Fig. 8 Macroscale engineering stressstrain curves for Al 359
MMC models with the inclusion of residual stresses at a room
temperature and b 150C
114 Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
model are shown in Fig. 13. The main observation is that there is
only a slight reduction in the stress level predictions of the model,
and the predictions are still very accurate. Indeed, the particle
fracture introduces a softening into the stressstrain response at a
higher strain that better represents the shape of the experimental
curve. The fracturing particles model initially produces the same
stressstrain behavior as the original 2D-Real Geometry model.
Once the particles crack, a process that happens quickly, the
stressstrain response corresponds to that of the model with ini-
tially cracked particles.
Discussion and Conclusions
This work has focused on a number of issues related to the
modeling of the mechanical behavior of the Al 35920% SiC
Fig. 9 Contour plots of a hydrostatic pressure and b Gurson void volume fraction
for the Al 359 MMC 2D-Real Geometry model at
22
ave
0.3%
Journal of Engineering Materials and Technology JANUARY 2005, Vol. 127 115
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
MMC, using computational micromechanics. The results of this
work show that quite a good prediction of the actual tensile me-
chanical behavior of the material up to the macroscopic failure
point can be achieved using periodic unit cell models and isotro-
pic matrix plasticity representations. This is true at both room
temperature and 150C, based on knowledge of the experimen-
tally measured tensile stressstrain behavior of the unreinforced
matrix. The accuracy of the 2D model predictions, bearing in
mind the obvious geometrical simplications of the 2D assump-
tion, is particularly noteworthy.
However, having said that, it is clear that the models do not
predict the temperature dependence of the stiffening effect of the
reinforcement as strongly as seen experimentally; the stiffening
effect is actually relatively high at the higher temperature. This
deciency could be due to the implicit assumption in the modeling
that the microstructure of the unreinforced alloy and the matrix in
the composite possess the same mechanical properties. This as-
sumption may break down after a prolonged period at 150C, as
accelerated precipitation may occur into the matrix Forn et al.
2.
Residual stresses due to thermal processing are seen to have a
signicant effect on model predictions, especially in 3D, and re-
sult in a better stressstrain curve prediction at room temperature.
The predicted stress magnitudes themselves compare favourably
with those measured using neutron diffraction.
In terms of the macroscale stiffening effect of the reinforce-
ment, the ideal and real geometry models predict practically iden-
tical behavior, suggesting that for predicting the stiffness of this
material and its stressstrain response up to the failure point, little
is to be gained by using complex unit cell geometry incorporating
particle clustering, etc. A very simple unit cell structure, such as
that in the 2D-Ideal Geometry model, with a small number of
particles, performs well.
It is interesting to see that even when microscale damage by
matrix void growth is introduced into the models the macroscale
effect is the same for both ideal and real geometry models. One
might expect the macroscale effects of microscale damage like
this to be very dependent on unit cell conguration. However, this
does not seem to be the case here. For example, even though there
is a 30% difference in peak plastic strain in the microstructure
between the two 2D models, as shown in Fig. 7, this does not
translate into a notable macroscopic effect, either with or without
matrix void growth.
Focusing on the 2D-Real Geometry model, the inclusion of
particle fracture produces accurate results and results in less ma-
terial softening in comparison to matrix void growth. This may be
explained by the failure of the interconnecting matrix that sup-
ports the loads for the matrix void growth case. Also, with the
void growth damage, a much greater volume of matrix fails in
comparison to the amount of reinforcing particles fracturing, at
least within the context of the model presented here.
An examination of the results leads to the conclusion that for
the prediction of the stiffness and tensile stressstrain behavior,
up to the failure point, of this material with its multiple phases
and porosity, the periodic unit cell modeling method performs
very well. The results also suggest that to achieve accuracy with
these models it is important to include thermal residual stresses,
and also a representation of particle fracture, which is known to
occur in this material. In addition, for an analysis at elevated
Fig. 10 Effect of void growth damage on the macroscale engi-
neering stressstrain curve for the Al 359 MMC 2D-Real Geom-
etry model at room temperature
Fig. 11 Contour plots of the effective plastic strain for two of
the SiC fracturing particles in the 2D-Real Geometry model, at
strain levels of a 0.1%, b 0.2% and c 0.3%
116 Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
temperatures, its attention would need to be given to the selection
of the matrix properties since the unreinforced alloy properties
might not be completely appropriate.
The representation of matrix void growth results in a response
that is much too soft, which highlights a weakness of the period-
icity assumption: when microdamage is distributed throughout the
volume of one cell, the material has effectively failed throughout
its entirety. This is unphysical when it comes to the representation
of a typical fracture process for such a material, which is charac-
terized by microdamage primarily concentrated in the neighbor-
hood of one, or at most a small number of, main cracks.
As has been stated, the models were not used to attempt to
predict the macroscopic failure point of this brittle material. How-
ever it is the authors contention that the modeling framework
presented here could be used to do so, through the representation
of particle fracture and subsequent matrix fracture. To do so
would more than likely require greater unit cell complexity in
terms of more particles and realistic particle shapes and distribu-
tion patterns. If the unit cell is large with many particles, damage
may be conned to sub-regions within the cell leaving relatively
large volumes of the material intact, and as a result not producing
high levels of articial softening. Llorca 9 estimates that in ex-
cess of 50 particles would be required to accurately predict mate-
rial behavior up to complete failure.
Acknowledgment
The authors thanks all partners of the EU BRITE/EURAM
project MISPOM Aerospatiale Matra, Defense Evaluation and
Research Agency, Universitat Politenica de Catalunya, Universita`
di Ancona, Erich Schmid Institut, Teksid, National University of
Ireland, Galway for their collaboration. They are indebted to the
European Community for nancial support, under Contract No.
BRPR-CT97-0396.
References
1 Forn, A., Martin, E., Baile, M. T., and Bastidas, J. M., 2002, Corrosion
Behavior of AA2124 and AA359 Reinforced with SiC Particles, 15th Inter-
national Corrosion Congress, Granada, Spain.
2 Forn, A., Martin, E., and Nogue, R., 2000, Characterization and Mechanical
Properties of 35920% SiC
p
, 8th National Con. Heat Treat. Surf., Barce-
lona, Spain.
3 Tabernig, B., and Pippan, R., 1998, Brite-Euram MISPOM Project Working
Paper UPC/WP/2/3 dated 31 March 1998.
4 Fiori, F., Girardin, E., Giuliani, A., Lorentzen, T., Pyzalla, A., Rustichelli, F.,
and Stanic, V., 2000, Neutron Diffraction Measurements For the Determina-
tion of Residual Stresses in MMC Tensile and Fatigue Specimens, Physica B,
276278, pp. 923924.
5 Carrado, A., Fiori, F., Girardin, E., Pirling, T., Powell, P., and Rustichelli, F.,
2001, Neutron Diffraction Measurements of Residual Stresses in Metal Ma-
trix Composite Samples, Radiat. Phys. Chem., 61, pp. 575577.
6 Suresh, S., Mortensen, A., and Needleman, A., eds., 1993, Fundamentals of
Metal Matrix Composites, Butterworth-Heinemann, Stoneham, MA.
7 Bohm, H. J., 1998, A Short Introduction to Basic Aspects of Continuum
Mechanics, CDL-FMD Report 3, Institute of Lightweight Structures and
Aerospace Engineering, Vienna University of Technology, Austria.
8 Bohm, H. J., 2003, Introduction to Continuum Mechanics, Course Notes
For CISM Course on Mechanics of Microstructured Materials, CISM, Udine,
Italy.
9 Llorca, J., 2003, Deformation and Damage in Particle Reinforced Compos-
Fig. 12 Contour plot of equivalent plastic strain for a quadrant of the 2D-Real Geom-
etry model with pre-cracked particles
Fig. 13 Effects of particle fracture on the macroscale engi-
neering stressstrain behavior of the 2D-Real Geometry model
at room temperature
Journal of Engineering Materials and Technology JANUARY 2005, Vol. 127 117
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms
ites: Experiments and Models, Course Notes For CISM Course on Mechanics
of Microstructured Materials, CISM, Udine, Italy.
10 McHugh, P. E., Asaro, R. J., and Shih, C. F., 1993, Computational Modeling
of Metal Matrix Composite MaterialsI. Isothermal Deformation Patterns in
Ideal Microstructures, Acta Metall. Mater., 41, pp. 14611476.
11 Christman, T., Needleman, A., and Suresh, S., 1989, Acta Metall. Mater., 37,
pp. 30293050.
12 Bao, G., McMeeking, R. M., and Hutchinson, J. W., 1991, Acta Metall. Mater.,
39, pp. 18711882.
13 Weissenbek, E., Bohm, H. J., and Rammerstorfer, F. G., 1994, Microme-
chanical Investigations of Arrangement Effects in Particle Reinforced Metal
Matrix Composites, Comput. Mater. Sci., 3, pp. 263278.
14 Llorca, J., Suresh, S., and Needleman, A., 1992, An Experimental and Nu-
merical Study of Cyclic Deformation in Metal Matrix Composites, Metall.
Trans. A, 23A, pp. 919934.
15 Bohm, H. J., 1991, Computer Based Micromechanical Investigations of the
Thermomechanical Behavior of Metal Matrix Composites, Fortschr.-Ber.
VDI Reihe 18 Nr. 101. Dusseldorf: VDI-Verlag.
16 Bruzzi, M. S., McHugh, P. E., ORourke, F., and Linder, T., 2001, Microme-
chanical Modeling of the Static and Cyclic Loading of an Al 2141SiC
MMC, Int. J. Plast., 17, pp. 565599.
17 Li, Y., and Ramesh, K. T., 1998, Inuence of Particle Volume Fraction,
Shape, and Aspect Ratio on the Behavior of Particle-Reinforced Metal Matrix
Composites at High Rates of Strain, Acta Mater., 46, pp. 56335646.
18 Fischmeister, H., and Karlsson, B., 1997, Plastizitatseigenschaften Grob-
zweiphasiger Werkstoffe, Z. Metallkd., 68, pp. 311327.
19 Brockenbrough, J. R., and Suresh, S., 1990, Plastic Deformation of Continu-
ous Fibre-Reinforced Metal-Matrix Composites: Effects of Fibre Shape and
Distribution, Acta Metall. Mater., 24, pp. 325330.
20 Li, M., Ghosh, S., Richmond, O., Weiland, H., and Rouns, T. N., 1999, Three
Dimensional Characterization and Modeling of Particle Reinforced Metal Ma-
trix Composites, Part I: Quantative Description of Microstructural Morphol-
ogy, Mater. Sci. Eng., A, 265, pp. 153173.
21 Ghosh, S., Lee, K. H., and Raghavan, P., 1996, Two Scale Analysis of Het-
erogeneous Elastic-Plastic Materials With Asymptotic Homogenization and
Voronoi Cell Finite Element Model, Comput. Methods Appl. Mech. Eng.,
132, pp. 63116.
22 Ghosh, S., and Moorthy, S., 1998, Particle Fracture Simulation in Non-
Uniform Microstructures of Metal-Matrix Composites, Acta Mater., 46, pp.
965982.
23 Gusev, A. A., 1997, Representative Volume Element Size for Elastic Com-
posites: A Numerical Study, J. Mech. Phys. Solids, 45, pp. 14491459.
24 Watt, D. F., Xu, X. Q., and Lloyd, D. J., 1996, Effects of Particle Morphology
and Spacing on the Strain Fields in a Plastically Deforming Matrix, Acta
Metall. Mater., 44, pp. 789799.
25 Bohm, H. J., Eckschlager, A., and Han, W., 2002, Multi-Inclusion Unit Cell
Models for Metal Matrix Composites With Randomly Oriented Discontinuous
Reinforcements, Comput. Mater. Sci., 25, pp. 4253.
26 Bohm, H. J., and Han, W., 2001, Comparisons Between Three-Dimensional
and Two-Dimensional Multi-Particle Unit Cell Models For Particle Reinforced
Metal Matrix Composites, Modell. Simul. Mater. Sci. Eng., 25, pp. 4765.
27 Hine, P. J., Lusti, H. R., and Gusev, A. A., 2002, The Numerical Simulation
of the Elastic and Thermoelastic Properties of Short Fibre Composites, Com-
pos. Sci. Technol., 62, pp. 14451453.
28 Gall, K., Horstemeyer, M., McDowell, D. L., and Fan, J., 2000, Finite Ele-
ment of the Stress Distributions Near Damaged Si Particle Clusters in Cast
AlSi Alloys, Mech. Mater., 32, pp. 277301.
29 Eckschlager, A., Han, W., and Bohm, H. J., 2002, A Unit Cell Model For
Brittle Fracture of Particles Embedded in a Ductile Matrix, Comput. Mater.
Sci., 25, pp. 8591.
30 Segurado, J., Gonzales, C., and Llorca, J., 2003, A Numerical Investigation
of the Effect of Particle Clustering on the Mechanical Properties of Compos-
ites, Acta Mater., 51, pp. 23552369.
31 Llorca, J., and Segurado, J., 2004, Three-Dimensional Multiparticle Cell
Simulations of Deformation and Damage in Sphere-Reinforced Composites,
Mater. Sci. Eng., A, 365, pp. 267274.
32 Gurson, A. L., 1975, Plastic Flow and Fracture Behavior of Ductile Materials
Incorporating Void Nucleation, Growth and Interaction, Ph.D. Thesis, Divi-
sion of Engineering, Brown University, Providence, RI.
33 Gurson, A. L., 1977, Continuum Theory of Ductile Rupture by Void Nucle-
ation and Growth: Part IYield Criteria and Flow Rules For Porous Ductile
Materials, J. Eng. Mater. Technol., 99, pp. 215.
34 HKS, 1997, ABAQUS Theory Manual Version 5.7, Hibbit, Karlsson and So-
rensen, Inc., Pawtucket, RI.
35 HKS, 2003, ABAQUS Users Manuals Version 6.4, Hibbit, Karlsson and So-
rensen, Inc., Pawtucket, RI.
36 Tabernig, B., Pippan, R., Bruzzi, M., McHugh, P., Foulquier, J., Rapoport, A.,
Fedrigoni, G. P., Powell, P. M., Cook, R., Forn, A., Martin, E., Fiori, F., and
Rustichelli, F., 2000, Modeling of In-Service Fatigue Performance of Particle
Reinforced Aluminum Alloys, Fatigue 2000, EMAS, M. R. Bache, P. A.
Blackmore, J. Draper, J. H. Edwards, P. Roberts, J. R. Yates, eds., pp. 93100.
118 Vol. 127, JANUARY 2005 Transactions of the ASME
Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 08/16/2013 Terms of Use: http://asme.org/terms

Вам также может понравиться