Sie sind auf Seite 1von 10

Downloaded from https://www.cambridge.org/core.

Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article
DOI: 10.1557/jmr.2019.60

3D multiscale modeling of fracture in metal matrix


composites
Yan Li1,a) , Leon Phung1, Cyril Williams2
1
Department of Mechanical and Aerospace Engineering, California State University, Long Beach, California 90840, USA
2
US Army Research Laboratory, Aberdeen Proving Ground, Adelphi, Maryland 21005, USA
a)
Address all correspondence to this author. e-mail: yan.li@csulb.edu
Received: 5 December 2018; accepted: 31 January 2019

Metal matrix composites (MMCs) have great potential to replace monolithic metals in many engineering
applications due to their enhanced properties, such as higher strength and stiffness, higher operating
temperature, and better wear resistance. Despite their attractive mechanical properties, the application of
MMCs has been limited primarily due to their high cost and relative low fracture toughness and reliability.
Microstructure determines material fracture toughness through activation of different failure mechanisms. In
this paper, a 3D multiscale modeling technique is introduced to resolve different failure mechanisms in MMCs.
This approach includes 3D microstructure generation, meshing, and cohesive finite element method based
failure analysis. Calculations carried out here concern Al/SiC MMCs and focus on primary fracture mechanisms
which are correlated with microstructure characteristics, constituent properties, and deformation behaviors.
Simulation results indicate that interface debonding not only creates tortuous crack paths via crack deflection
and coalescence of microcracks but also leads to more pronounced plastic deformation, which largely
contributes to the toughening of composite materials. Promotion of interface debonding through
microstructure design can effectively improve the fracture toughness of MMCs.

Introduction negatively influence the fracture toughness of MMCs. Alaneme


Metal matrix composites (MMCs) exhibit excellent material and Aluko [7] reported that interface debonding is a beneficial
properties such as higher specific strength, operating temper- failure mechanism which can lead to improved fracture
toughness of MMCs. This is because the propagation of
ature, and wear resistance than monolithic metals. These
interface microcracks to the ductile matrix causes more pro-
properties make MMCs attractive for many structural applica-
nounced plastic deformation which largely contributes to the
tions. To facilitate successful application of MMCs, an in-depth
toughening of composite materials. In addition to the exper-
understanding of the microstructure–property relationship is
imental studies, a number of numerical investigations have
required. It has been reported that microstructure and constit-
been performed, largely with the intent of predicting the
uent properties combine to determine the overall fracture
fracture behavior of MMCs given the mechanical properties
toughness of MMCs through the activation of different fracture
of the matrix and reinforcement phases. Ayyar and Chawla [8]
mechanisms [1, 2, 3]. Specifically, interface debonding and
j Journal of Materials Research j www.mrs.org/jmr

conducted a series of 2D microscale simulations of MMCs by


particle cracking are two competing fracture mechanisms considering the effect of particle orientation, shape, and size on
during the crack–reinforcement interactions [4]. Qian et al. crack growth. They found that a homogenous particle distri-
[5] experimentally evaluated the fracture toughness of 6061 Al bution with minimum particle clustering would be preferential
MMCs and concluded that MMCs with large reinforcement for enhancing the crack growth resistance. Sozhamannan et al.
particles exhibit lower initiation fracture toughness, crack [9] studied the fracture behavior of MMCs based on a 2D
propagation energy, and total absorbed energy. Jarzabek et al. elastic–plastic finite element model. Their simulation results
[6] further observed that large particles correspond to higher confirm that large particles are more prone to fracture due to
interface strength which discourages interface debonding and the large stress concentration and high strain hardening rate

ª Materials Research Society 2019 cambridge.org/JMR 1


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

around the particles and increased area of contact. It is worth


noting that fracture is inherently a 3D problem. 2D models,
which assume plane strain conditions, cannot capture the 3D
morphology of microstructural details or realistically track the
crack–material interactions due to nonplanar crack extension.
Romanova et al. [1] studied the influence of reinforcing particle
shape and interface strength on the fracture behavior of
a MMC through 3D calculations. This work only focused on
idealized microstructure configurations and fracture patterns.
Li et al. [10] developed a 3D model which allows representation
and characterization of realistic 3D microstructures of particle
reinforced MMCs. However, this model does not explicitly
account for microstructural level crack propagation.
In this paper, a 3D multiscale modeling technique is
introduced to resolve different failure mechanisms in MMCs.
Figure 1: Scheme of specimen configuration in the analysis.
This approach includes 3D microstructure generation, meshing,
and cohesive finite element method based micromechanical
simulations. The 3D model developed here allows representation Constitutive modeling of bulk matrix
of both realistic and idealized microstructure configurations. It Johnson–Cook model is employed to describe the constitutive
also allows explicit tracking of crack propagation in the micro- behavior of bulk aluminum matrix. This model considers the
structure. For good delineation of microstructural features and separated effects of strain hardening and thermal softening as
realistic prediction of crack trajectories, unstructured tetrahedral h  n i m
meshes are generated using the open source code iso2mesh [11].  ¼ A þ B epl
r 1  T^ ; ð1Þ
Cohesive elements with bilinear traction-separation laws are
embedded in the microstructure region so as to resolve different where r  and epl are the static yield stress and equivalent plastic
failure mechanisms and microcrack patterns. This 3D multiscale strain, respectively. A, B, m, and n are the material parameters,
framework allows quantification of energy dissipations through which are determined based on the flow stress data obtained
explicit simulation of fracture process in the microstructure from mechanical tests. T^ is the nondimensional temperature
region. The modeling approach developed here provides new defined as
insights into the physical aspects of competition between 8
<0 for T , Ttran
different failure mechanisms and their correlations with plastic T^ ¼ ðT  Ttran Þ=ðTmelt  Ttran Þ for Ttran # T # Tmelt ;
:
deformation and microstructural properties. 1 for T > Tmelt
ð2Þ

Cohesive finite element based 3D multiscale Here, T is the current temperature, Tmelt is the melting
modeling temperature, and Ttran is the transition temperature defined
An edge-cracked specimen under Mode I static tensile loading is as the one at or below which there is no temperature de-
illustrated in Fig. 1. This model includes both the microstructure pendence of the yield stress [12]. It should be noted that A, B,
region and the homogenized region. The microstructure region is m, and n are measured at or below the transition temperature
a cube with dimension of 100 lm. It is embedded in the Ttran. The material parameters are summarized in Table I [13].
homogenized region, which is also a cube with a dimension of The Young’s modulus Em and Poisson’s ratio mm of Al matrix
300 lm. The precrack plane has a length of 125 lm and a width are chosen as 73 GPa and 0.33, respectively.
j Journal of Materials Research j www.mrs.org/jmr

of 300 lm. A boundary velocity of v 5 5 mm/s is imposed at the


top and bottom surfaces of the specimen to effect symmetric
Constitutive modeling of bulk reinforcements
Mode I tensile loading. The remaining surfaces of the specimen Bulk SiC reinforcements follow isotropic linear elastic consti-
are traction-free. Conditions of plane strain are assumed to tutive relations as
prevail. The configuration illustrated in Fig. 1 allows explicit Ep mp Ep
representation of microstructures and account for microstruc- r¼ eþ   trðeÞI ; ð3Þ
1 þ mp 1 þ mp 1  2mp
tural level of deformation, damage, and failure mechanisms, while
allowing macroscopic conditions such as controlled loading and where r and e are the stress and strain tensors, respectively. tr
structural response to be considered at the same time. (e) is the trace of e and I is the second order identity tensor.

ª Materials Research Society 2019 cambridge.org/JMR 2


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

The Young’s modulus Ep and Poisson’s ratio mp of SiC p2 þ ½ðD  ðD  nÞnÞ  q2 ¼ jD  ðD  nÞnj, respectively, de-
reinforcement are chosen as 410 GPa and 0.14, respectively. note the normal and tangential components of D, with n being
the unit normal, and p and q being the two unit tangential
Constitutive modeling of cohesive elements vectors. Note that n, p, and q are mutually orthogonal to each
other and form a right-handed triad. Dnc is the critical normal
Cohesive finite element method has been extensively used to separation at which the cohesive strength of an interface
model crack propagation in different types of materials [14, 15, vanishes under conditions of pure normal deformation (Dt 5
16, 17]. Generally speaking, there are two approaches exist for 0). Similarly, Dtc is the critical tangential separation at which
embedding cohesive elements in the bulk elements when the the cohesive strength of an interface vanishes under conditions
crack paths are unknown a priori. One method is to insert of pure shear deformation (Dn 5 0). Tmax represents the
cohesive elements in front of the crack tip as the crack develops maximum traction that the cohesive element can sustain at the
[18, 19]. This method can avoid cohesive surface-induced onset of irreversible separation.
stiffness reduction of the overall model. However, it is In order to account for the irreversibility of separations,
computationally expensive and requires specific fracture initi- a parameter g 5 max{g0, kul} is defined. As illustrated in
ation criteria that are extrinsic to the overall finite element Fig. 2, g0 is the initial value of g which defines the stiffness of
model. Another method is to embed cohesive elements in any the original undamaged cohesive surface, while kul is the
surface which is shared by two bulk elements [20, 21]. Cohesive hitherto maximum value of k at which an unloading process
elements become part of the physical model as fracture emerges was initiated. It should be noted that kul is associated with the
as a natural outcome of the deformation process without the onset of an unloading event and is not necessarily the hitherto
use of any failure criterion. In this paper, we employ the second maximum value of k. kul represents the current (reduced)
method by permeating cohesive elements in the microstructure
stiffness of the cohesive surfaces after damage and unloading
region. The cohesive constitutive law takes the form of a bi- have occurred. Furthermore, g0 represents the characteristic
linear relation between traction and separation, as illustrated in value of effective separation k at which the effective traction r
Fig. 2. This is an idealization that can be modified to for a cohesive surface pair reaches the strength reaches the
accommodate more detailed and quantitative information strength Tmax of the undamaged surface. kul stands for the
regarding, for example, interfacial property by adjusting the critical level of k at which r reaches the reduced strength Tmax
cohesive strength or separation energy. The cohesive law is (1  g)/(1  g0) of the hitherto damaged cohesive surface
derived from a potential, U, which is a function of the pair. The specific expression for potential U is of the form
separation vector D through
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi a state variable defined as
k ¼ ðDn =Dnc Þ2 þ ðDt =Dtc Þ2 . This variable describes the ef- 8  1g  2 
>
> U k
if 0#k#g
< 0 1g0  g 
fective instantaneous state qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
of mixed-mode separations. Here,
 2  2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi U ¼ Uðk; gÞ ¼ 1g ð1kÞ2 :
Dn 5 nD and Dt ¼ Dp þ Dq ¼ ½ðD  ðD  nÞnÞ > U0 1 if g#k#1
>
: 1g0 1g
0 if k>1
ð4Þ
TABLE I: Johnson–Cook material parameters for Al matrix.

A (MPa) B (MPa) n M Tmelt (°C) Ttran (°C) As indicated in Fig. 2, separation occurs elastically and the
cohesive energy stored (work done in causing separation) is
324 114 0.42 1.34 925 293.2
fully recoverable between A and B (0 # k # g0), and damage
j Journal of Materials Research j www.mrs.org/jmr

Figure 2: Bilinear traction-separation law for cohesive elements.

ª Materials Research Society 2019 cambridge.org/JMR 3


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

in the form of microcracks and other small-scale defects does cohesive elements are inserted and sorted out all the shared
not occur. Between B and C (g0 # k # 1), material faces and related element indices; (iv) duplicate the nodes in
degradation causes progressive reduction in the strength of the shared faces and redistribute the node label to the elements
the cohesive surfaces. This represents a phenomenological that share the face. Update the ABAQUS inp file by adding the
account of the effects of microcracks and other defects not new node information to the *Node section and updating the
explicitly modeled in the cohesive finite element model. node connectivity in the *Element section. As illustrated in
Unloading from any point P follows path PA and subsequent Fig. 3, element 1 and element 2 share the same face with node
reloading follows AP and then PC. Part of the work expended 2, 3, and 4. Assume N is the current total node number. Node
on causing the separation in this regime is irreversible, as N 1 1, N 1 2, and N 1 3 are cloned from node 2, 3, and 4,
indicated by the hysteresis loop ABP which implies dissipation respectively. When distributing the duplicated nodes to ele-
during the softening process. Correspondingly, there is a de- ments (element 1 and element 2 in this case), the node labels in
crease in the maximum tensile strength of the cohesive surface. each element are only allowed to update once. For example,
This is reflected in the elastic reloading of the interface along node 2 can be the member of another shared face in addition to
AP and further softening along path PC. face (2 3 4). Once node 2 in element 1 is replaced with node N
Introducing cohesive surfaces to the complex 3D micro- 1 1, it is no longer allowed to be replaced by another
structure meshes requires extensive node relabeling and duplicated node afterward. This duplicated node can only be
changes in elemental connectivity. An algorithm is developed redistributed to the other element which shared the face. The
to automatically insert 3D cohesive element in the microstruc- cohesive element is generated by connecting the six nodes in
ture region and assign corresponding cohesive law based on the a counterclockwise direction. As illustrated in Fig. 3, each
material phase property. In a two-phase MMC microstructure, cohesive element follows the numbering pattern as [2 3 4 N 1
there are three cohesive laws exist: cohesive laws in the matrix, 1 N 1 2 N 1 3]. It is assumed that all the 6-node cohesive
in the reinforcement, and along the interface. This process elements (COH3D6) have zero thickness.
includes the following steps: (i) extract the existing nodal
coordinates and element arrangement of the microstructure;
(ii) separate all nodes and elements in three sets: matrix, Constitutive modeling in the homogenized region
reinforcement, and interface. Further separate the reinforce- The homogenized region as shown in Fig. 1 does not include
ment set for each particle; (iii) define the location where any cohesive element. The effective properties in this region are

j Journal of Materials Research j www.mrs.org/jmr

Figure 3: Algorithm for insertion of cohesive elements in the microstructure.

ª Materials Research Society 2019 cambridge.org/JMR 4


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

estimated by using the Mori–Tanaka method [22]. Specifically,


dc ¼ del þ dpl ; ð9Þ
the effective bulk and shear moduli are
8 Here, dc is the critical CTOD. del and dpl represent the elastic
<K  ¼ K0 þ f ðK1 K0 Þð3K0 þl0 Þ
3K0 þ4l0 þ3ð1f ÞðK1 K0 Þ
; ð5Þ and plastic parts of the CTOD, respectively.
:l 5f l0 ðl1 l0 Þð3K0 þ4l0 Þ
 ¼ l0 þ 5l ð3K0 þ4l
0 Þþ6ð1f Þðl l ÞðK0 þ2l Þ
0 1 0 0 The elastic part of CTOD del can be computed by the
following:
Here, f is the volume fraction of reinforcement particles. Kr and  
lr represent the bulk and shear moduli for matrix (r 5 0) and del ¼ KI2 1    y ;
m2 =2Er ð10Þ
reinforcements (r 5 1), respectively.
where the effective Young’s modulus E  and effective Poisson’s
The homogenized Young’s modulus E  and Poisson’s ratio 
m
ratio 
m of MMCs are calculated according to Eq. (6). The initial
are calculated according to
8 yield stress of MMCs ry is estimated based on Eq. (8). KI is the
l
<E  ¼ 9K stress intensity factor and can be numerically calculated
3Kþ l
: ð6Þ
: m ¼ 3K2
 l according to
 l
6Kþ2
h i
KI ¼ rij ð2pr Þ1=2 =gIij ðhÞ ; ð11Þ
It is assumed that the yield criteria in the homogenized
region follow J2 plasticity without considering strain rate
where gIij ðhÞ is the function of h and rij is the stress value of the
dependency or the influence of temperature [23]. The yield
node having maximum displacement at radius r and angle h.
function f is formulated as
The plastic portion of CTOD dpl is determined by
rffiffiffi
2 assuming that the uncracked ligament works like a plastic
f ðsÞ ¼ ksk  ry ; ð7Þ
3 hinge with its center at a distance rplb from the crack tip.
Specifically, dpl is calculated as
Here, s is the deviatoric stress tensor and ksk denotes the norm
 
of the stress deviator. ry is the initial yield stress of MMCs. dpl ¼ rpl b= rpl b þ a Vpl ; ð12Þ
According to Nardone and Prewo [24], ry can be estimated by
considering the microstructure effect as Here, rpl, b, and a are defined as the plastic rotational factor,
ligament length, and precrack length, respectively. In this
ry ¼ rm
y ½f ðS þ 2Þ=2 þ ð1  f Þ ; ð8Þ paper, rpl 5 0.5, b 5 200 lm, and a 5 125 lm are considered.
The crack mouth opening displacement Vpl can be extracted
where rm y is the yield stress of matrix. In the following from the load–displacement curve as discussed in detail in
y ¼ 177:5 MPa is employed. S is defined as the
calculations, rm ASTM E1290-08e1 [25].
aspect ratio. For spherical SiC particles, S 5 1 is considered. f
represents the volume fraction of SiC particles. The homogenized
region and microstructure region are connected using mesh tie Results and discussions
constraint which requires no conformity of node connections. In the following analysis, the fracture toughness KIC of Al
Although a simplified constitutive model is employed, this matrix and SiC particles are selected as 25 MPa m1/2 and 4.6
framework allows the microstructure effect on the overall MPa m1/2, respectively [27, 28]. The fracture energy rate U of
response of MMCs to be quantified. A more advanced each phase is estimated according to
homogenization method, though more realistic and considers   2
U ¼ J ¼ 1  m2 KIC =E : ð13Þ
more factors, would also complicate an initial study aimed at
developing a 3D approach for relating microstructure to
Based on the material properties provided in sections
macroscopic fracture toughness because of its large number
“Constitutive modeling of bulk matrix” and “Constitutive
of parameters. But the capability of the current framework can
j Journal of Materials Research j www.mrs.org/jmr

modeling of bulk reinforcements”, the fracture energy rates


be extended to address hardening and material rate depen-
in Al matrix and SiC particles are calculated as 7.63  103 N/m
dency during the deformation and failure process.
and 50.6 N/m, respectively. The maximum traction Tmax is
obtained from Tmax ¼ D2Unc ¼ aD
2U
tc
, where a is defined as Dnc/Dtc
Evaluation of fracture toughness from crack tip and is assumed to be 1. It should be noted that both the normal
opening displacement and shear separations should not exceed the element size in the
Crack tip opening displacement (CTOD) has been widely used microstructure region so that cohesive surfaces can achieve
to measure material fracture toughness. According to ASTM complete debonding during the fracture process. In our model,
E1290-08e1 [25] and Panontin et al. [26], Dnc 5 Dtc 2 (1, 5) lm is a reasonable separation range.

ª Materials Research Society 2019 cambridge.org/JMR 5


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

In the following calculations, interfacial fracture energy rate precrack plane as the particle volume fraction increases. The
U 5 10.12 N/m is considered. 3D microstructures with crack trajectory is relatively localized when f 5 10%. As
spherical particles are systematically generated by choosing the particle volume fraction decreases, the crack trajectory
a particle radius of 8 lm and a volume fraction f of 1%, 5%, and becomes more tortuous as microcracks start to initiate at places
10%, respectively. In order to account for stochastic variation of that are further away from the precrack plane. It is also noted
fracture behavior at the microstructural level, 20 random from Fig. 5 that more intensive interface debonding is observed
microstructure instantiations (samples) under each volume when the particle volume fraction is low. Here, SDEG
fraction are generated and implemented in the multiscale (a computational variable in the ABAQUS software) is a pa-
model. Figure 4 shows a few microstructure instantiations rameter, which quantifies the degree of damages in cohesive
which are selected from each volume fraction group. As elements. Specifically, SDEG 5 1 represents fully debonded
indicated in Fig. 5, the crack tends to propagate along the cohesive surfaces when the traction stress drops to 0 (point C

j Journal of Materials Research j www.mrs.org/jmr

Figure 4: Illustration of selective 3D microstructures with different particle volume fractions.

ª Materials Research Society 2019 cambridge.org/JMR 6


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

Figure 5: Crack propagation in microstructures and damages in particles and along the interface.

takes place. The simulation results indicate that particles with


complete interface debonding can effectively survive fracture.
This trend is supported by a few experimental results of MMCs
[29, 30, 31, 32]. As shown in Fig. 7, the surfaces near the
precrack tip of two fractured compact tension (CT) MMC
specimens are polished so that both the microstructure features
and the crack trajectory are well observed from the microscope.
It is noticed that the specimen with lower particle volume
fraction exhibit more tortuous crack path and a higher per-
centage of interface debonding. In contrast, the increase of
particle volume fraction leads to straighter crack trajectory and
more intensified particle cracking. From the energy point of
view, a crack can grow only when the energy available at the
Figure 6: Effect of particle volume fraction on the percentage of fully
debonded interface. crack tip is sufficient to balance out the energy required for
crack propagation. For brittle materials such as ceramic
composites, the most effective way to improve fracture tough-
in Fig. 2). It signifies complete failure of the cohesive element ness is to create tortuous crack paths since, in the absence of
and formation of a crack. Regions with SDEG , 1 do not plasticity, the total energy released is transformed into surface
represent actual crack formation but can be considered as energy alone. For MMCs, the fracture resistance depends on
j Journal of Materials Research j www.mrs.org/jmr

potential sites for microcracks. According to Fig. 6, the average the sum of energy spent on both surface generation and plastic
percentage of fully debonded interface is 95.3% when the deformation in the ductile matrix. As indicated in Fig. 8, higher
corresponding particle volume fraction f 5 1%. This percent- volume fraction of the brittle SiC particles leads to lower plastic
age decreases to 68.7% when f increases to 10%. It is also worth energy dissipation and therefore, lower fracture resistance of
noting that interface debonding and particle cracking are two the entire materials. This result agrees well with a few exper-
competing failure mechanisms. As shown in Fig. 5, when f 5 imental findings that particle cracking suppresses plastic de-
1%, no particle cracking is observed when nearly complete formation in the ductile matrix and negatively influences the
interface debonding is activated. As f increases, particle material fracture toughness [6, 9, 33, 34]. According to section
cracking starts to emerge and gradually interface debonding “Evaluation of fracture toughness from CTOD”, the CTOD

ª Materials Research Society 2019 cambridge.org/JMR 7


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

Figure 7: Experimentally observed crack trajectory of Al/SiC MMCs under different SiC volume fractions.

Figure 8: Plastic energy dissipation at various particle volume fractions. Figure 9: Effect of particle volume fraction on CTOD.

particles lose the capability to carry load. Interface debonding, on


values measured from each particle volume fraction group are the other hand, alleviates stress concentration in the particle
summarized in Fig. 9. It is noted that CTOD decreases with phase and promotes microcrack initiation and propagation to
particle volume fraction f. This indicates that the material the matrix phase, leading to more intensified plastic energy
response is more brittle with increasing f. The higher volume dissipation. In addition, the coalescence of microcracks can lead
fraction of brittle phase leads to lower level plastic deformation to higher surface energy dissipation as a result of tortuous crack
j Journal of Materials Research j www.mrs.org/jmr

in the matrix; therefore, it creates favorable condition for paths. The above conclusions can be supported by our simula-
particle cracking which negatively influences the material tion results and other experimental observations [5, 6, 7].
fracture toughness. In addition, the CTOD values tend to be
more scattered when f decreases. This is expected as the crack
trajectory becomes more localized when particle cracking wins Summary
the competition during the crack–particle interactions. A cohesive finite element based multiscale framework is
It can be concluded that interface debonding is a beneficial developed to simulate crack propagation in Al/SiC MMCs.
mechanism for both strengthening and toughening of MMCs. This framework allows explicit representation of microstruc-
Particle cracking adversely affects material strength since cracked ture and resolution of different failure mechanisms. Simulation

ª Materials Research Society 2019 cambridge.org/JMR 8


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

results indicate that interface debonding can positively con- 10. M. Li, S. Ghosh, O. Richmond, H. Weiland, and T.N. Rouns:
tribute to the toughening of MMCs. This failure mechanism Three dimensional characterization and modeling of particle
not only creates tortuous crack paths via crack deflection and reinforced metal matrix composites part II: Damage
coalescence of microcracks with the precrack but also leads to characterization. Mater. Sci. Eng., A 266, 221–240 (1999).
more pronounced plastic deformation, which largely contrib- 11. Q. Fang and D.A. Boas: Tetrahedral mesh generation from
utes to the toughening of composite materials. The methodol- volumetric binary and gray-scale images. In Proceedings of the
ogy is potentially useful for development of high toughness Sixth IEEE international conference on Symposium on Biomedical
MMCs through selection of materials and tailoring of Imaging: From Nano to Macro, Boston, Massachusetts, 2009.
microstructures. 12. ABAQUS 6-14 manual (Simulia, Providence, Rhode Island, 2016).
13. A. Ghandehariun: Mechanics of machining metal matrix
Acknowledgment composites: Analytical modeling and finite element simulation.
Doctoral Dissertation, University of Ontario Institute of
This research is primarily supported by the Army Research
Technology, 2015. Retrieved from http://hdl.handle.net/10155/514.
Office through contract No. W911NF-16-1-0541. The financial
14. Y. Li and M. Zhou: Effect of competing mechanisms on fracture
and technical support from the Impact Physics Branch of the
toughness of metals with ductile grain structures. Eng. Fract. Mech.
Army Research Lab at Aberdeen Proving Ground is highly
appreciated. The funding support from HSI STEM Program 205, 14–27 (2019).

and ORSP Summer Research Program at California State 15. X. Guo, K. Chang, L.Q. Chen, and M. Zhou: Determination of

University, Long Beach is acknowledged as well. fracture toughness of AZ31 Mg alloy using the cohesive finite
element method. Eng. Fract. Mech. 96, 401–415 (2012).
16. A. Needleman, S.R. Nutt, S. Suresh, and V. Tvergaard: Matrix,
References
reinforcement, and interfacial failure. In Fundamentals of Metal-
1. V.A. Romanova, R.R. Balokhonov, and S. Schmauder: The Matrix Composites, S. Suresh, A. Mortensen, and A. Needleman,
influence of the reinforcing particle shape and interface strength on
eds. (Butterworth-Heinemann, Boston, 1993); ch. 13, pp. 233–250.
the fracture behavior of a metal matrix composite. Acta Mater. 57,
17. Y. Li and M. Zhou: Prediction of fracture toughness of ceramic
97–107 (2009).
composites as function of microstructure: I. Numerical
2. K.K. Chawla: Metal matrix composites. In Composite Materials
simulations. J. Mech. Phys. Solids 61, 472–488 (2013).
(Springer, New York, 1998).
18. O. Shor and R. Vaziri: Adaptive insertion of cohesive elements for
3. J. Llorca, A. Needleman, and S. Suresh: An analysis of the effects
simulation of delamination in laminated composite materials. Eng.
of matrix void growth on deformation and ductility in metal-
Fract. Mech. 146, 121–138 (2015).
ceramic composites. Acta Metall. Mater. 39, 2317–2335 (1991).
19. A.O. Pandolfi: An efficient adaptive procedure for three-
4. L. Babout, Y. Brechet, E. Maire, and R. Fougères: On the
dimensional fragmentation simulations. Eng. Comput. 18, 148
competition between particle fracture and particle decohesion in
(2002).
metal matrix composites. Acta Mater. 52, 4517–4525 (2004).
20. J. Zhai, V. Tomar, and M. Zhou: Micromechanical simulation of
5. L. Qian, T. Kobayashi, H. Toda, T. Goda, and Z.G. Wang:
dynamic fracture using the cohesive finite element method. J. Eng.
Fracture toughness of a 6061Al matrix composite reinforced with
Mater. Technol. 126, 179–191 (2004).
fine SiC particles. Mater. Trans. 43, 2838–2842 (2002).
21. Y. Li, D. McDowell, and M. Zhou: A multiscale framework for
6. D.M. Jarzabek, M. Chmielewski, J. Dulnik, and A. Strojny-
Nedza: The influence of the particle size on the adhesion between predicting fracture toughness of polycrystalline metals. Mater.

ceramic particles and metal matrix in MMC composites. J. Mater. Perform. Charact. 3, 1–16 (2014).

Eng. Perform. 25, 3139–3145 (2016). 22. J. Qu and M. Cherkaoui: Fundamentals of Micromechanics of

7. K.K. Alaneme and A.O. Aluko: Fracture toughness (K1C) and Solids (John Wiley, Hoboken, NJ, 2007), pp. 120–153.
23. R.I. Borja: Plasticity Modeling and Computation (Springer-Verlag,
j Journal of Materials Research j www.mrs.org/jmr

tensile properties of as-cast and age-hardened aluminium (6063)–


silicon carbide particulate composites. Sci. Iran. 19, 992–996 Berlin, Heidelberg, 2013).
(2012). 24. J.L. York Duran, C. Kuhn, and R. Müller: Modeling of the
8. A. Ayyar and N. Chawla: Microstructure-based modeling of crack effective properties of metal matrix composites using
growth in particle reinforced composites. Compos. Sci. Technol. 66, computational homogenization. Appl. Mech. Mater. 869, 94–111
1980–1994 (2006). (2017).
9. G.G. Sozhamannan, S.B. Prabu, and R. Paskaramoorthy: 25. ASTM E1290-08e1: Standard test method for crack-tip opening
Failures analysis of particle reinforced metal matrix composites by displacement (CTOD) fracture toughness measurement. Am. Soc.
microstructure based models. Mater. Des. 31, 3785–3790 (2010). Test. Mater. (2011).

ª Materials Research Society 2019 cambridge.org/JMR 9


Downloaded from https://www.cambridge.org/core. Eugene McDermott Library, University of Texas at Dallas, on 11 Mar 2019 at 05:45:34, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2019.60

Article

26. T.L. Panontin, A. Makino, and J.F. Williams: Crack tip opening 31. J.N. Hall, J.W. Jones, and A.K. Sachdev: Particle size, volume
displacement estimation formulae for C(T) specimens. Eng. Fract. fraction and matrix strength effects on fatigue behavior and
Mech. 67, 293–301 (2000). particle fracture in 2124 aluminum-SiCp composites. Mater. Sci.
27. R.J. Bucci: Selecting aluminum alloys to resist failure by fracture Eng., A 183, 69–80 (1994).
mechanisms. Eng. Fract. Mech. 12, 407–441 (1979). 32. Y. Li, J. Cao, and C. Williams: Competing failure mechanisms in
28. https://accuratus.com.
metal matrix composites and their effects on fracture toughness.
29. C. Wu, K. Ma, J. Wu, P. Fang, G. Luo, F. Chen, Q. Shen,
Materialia 5, 100238 (2019).
L. Zhang, J.M. Schoenung, and E.J. Lavernia: Influence of particle
33. M. Song and B. Huang: Effects of particle size on the fracture
size and spatial distribution of B4C reinforcement on the
toughness of SiCp/Al alloy metal matrix composites. Mater. Sci.
microstructure and mechanical behavior of precipitation strengthened
Al alloy matrix composites. Mater. Sci. Eng., A 675, 421–430 (2016). Eng., A 488, 601–607 (2008).

30. M.T. Milan and P. Bowen: Tensile and fracture toughness 34. L. Babout, W. Ludwig, E. Maire, and J.Y. Buffière: Damage

properties of SiCp reinforced Al alloys: Effects of particle size, assessment in metallic structural materials using high resolution
particle volume fraction, and matrix strength. J. Mater. Eng. synchrotron X-ray tomography. Nucl. Instrum. Methods Phys. Res.,
Perform. 13, 775–783 (2004). Sect. B 200, 303–307 (2003).

j Journal of Materials Research j www.mrs.org/jmr

ª Materials Research Society 2019 cambridge.org/JMR 10

Das könnte Ihnen auch gefallen