You are on page 1of 20

Available online at www.sciencedirect.

com
ScienceDirect

Comput. Methods Appl. Mech. Engrg. 347 (2019) 1085–1104


www.elsevier.com/locate/cma

A two-set order parameters phase-field modeling of crack


deflection/penetration in a heterogeneous microstructure
Hu Chena,b , Chi Zhanga ,∗, Qianli Lub,c , Hao Chena , Zhigang Yanga , Youhai Wend ,
Shenyang Hue , Lei Chenb ,∗
a Key Laboratory of Advanced Materials of Ministry of Education, School of Materials Science and Engineering, Tsinghua
University, Beijing 100084, China
b Department of Mechanical Engineering, Mississippi State University, MS 39762, USA
c State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum University, Chengdu 610500, China
d U.S. DOE National Energy Technology Laboratory, Albany, OR 97321, USA
e Pacific Northwest National Laboratory, Richland, WA 99352, USA

Received 26 June 2018; received in revised form 31 October 2018; accepted 8 January 2019
Available online 21 January 2019

Highlights
• Competition between crack penetration and deflection at an interface is studied.
• Interface and bulk fracture toughness are incorporated through a unified form.
• Effect of grain boundary energy and interface fracture toughness on crack deflection.
• Crack morphology and fracture toughness in polycrystals are simulated.

Abstract
Modeling of crack deflection/penetration or manifested as intergranular/transgranular fracture in polycrystals has long been
a challenge in both fracture mechanics and materials science. A phase-field model formulated with two-set order parameters
describing the crack field and the microstructure field respectively, is herein established to investigate the competition between crack
penetration and deflection at an interface. The advantages of the proposed model are three-fold: (1) the fracture toughness of grain
boundaries and/or interfaces, besides single crystal properties (e.g., Young’s modulus, Poisson ratio, and fracture toughness), can
be automatically incorporated, (2) cohesive interfaces are treated in one phase-field framework in comparison with the integrated
phase-field and cohesive zone model in which cohesive elements have to be manually placed at the interface, and (3) two-set order
parameters enable to describe the concurrent crack propagation and microstructure evolution. The proposed model is validated by
comparing the simulated critical deflection angles with the results from theoretical analysis of the linear elastic fracture mechanics
and the experiments. The model is then applied to a heterogeneous Silicon Nitride polycrystal to determine the effect of interface
fracture toughness on the intergranular/transgranular fracture transition and the effective toughness of the materials.
⃝c 2019 Elsevier B.V. All rights reserved.

Keywords: Phase-field modeling; Crack deflection; Interface fracture toughness; Heterogeneous microstructure

∗ Corresponding authors.
E-mail addresses: chizhang@tsinghua.edu.cn (C. Zhang), chen@me.msstate.edu (L. Chen).

https://doi.org/10.1016/j.cma.2019.01.014
0045-7825/⃝ c 2019 Elsevier B.V. All rights reserved.
1086 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

1. Introduction
Modeling of intergranular and transgranular fracture in heterogeneous microstructures has gained extensive
attention from both fracture mechanics and materials science communities, because the crack morphology in a brittle
heterogeneous material is the key element for determining the effective toughness [1,2]. The nature of the transition
from the intergranular to transgranular fracture can be simplified to a classical fracture mechanics problem, a crack
impinging on an interface with an inclined angle.
In the linear elastic fracture mechanics (LEFM), He and Hutchinson [3] had demonstrated theoretically that the
competition between crack deflection and penetration depends on (1) the ratio of the interface fracture toughness to
the bulk fracture toughness, and (2) the angle of the crack at which is impinging on the interface. This problem had
been further re-examined from crack topologies [4], failure mechanism [5], stress singularity [6,7], dynamic crack
[8], etc. Martı́nez et al. [4] presented the correct solutions for the doubly deflected case, which indicated that the ratio
of energy release rate for the interface to that of the bulk in the doubly deflected case is lower than the corresponding
value for the singly deflected case. As the model proposed by He and Hutchinson sometimes failed to predict the
interfacial deflection or interfacial penetration, Leguillon et al. [5] extend the crack deflection model by taking into
the interface debonding ahead of the crack tip to enable crack deflection more favorable than He and Hutchinson’s
criterion. To tackle the singularity of the stress field at the interface, asymptotic methodologies were extensively
utilized [6,7]. Xu et al. [8] explored the equivalent dynamic problem by examining the deflection/penetration behavior
of dynamic cracks propagating at various speeds towards inclined weak interfaces. Although some experimental
studies confirmed the model proposed by He and Hutchinson [3], more complex crack propagation behavior appeared
in experimental investigations was not expected to be reproduced by the above analytical model. For example,
simultaneous occurrence of crack penetration and deflection was observed in the bi-material system [9], which is
out of the capability of the above model. The phenomenon that the crack path, manifested as the percentage of
intergranular and transgranular fracture, of polycrystal ceramic was found to be determined by the shape of the grain,
grain size and chemical concentration of the interface [10–12] also cannot be captured by the above model.
Some numerical methods, such as extended-finite element method (XFEM) [13–18], and cohesive zone model [2,
19,20], have thus been established to tackle the complex crack propagation problems that cannot be solved by the
analytical model. In the XFEM, ad hoc parameters are needed for the crack initiation and propagation, as well as
the crack growth direction [19]. Moreover, the XFEM presents significant implementational difficulties, such as
remeshing, state variables transfer [21] and integration [22]. A widely used alternative to simulate the transition
from the intergranular to transgranular fracture is the cohesive zone model, which has achieved notable implantation
in ceramic materials, such as Silicon Nitride (Si3 N4 ) [2,19]. Dynamic crack propagation in the Si3 N4 microstructure
containing elongated grains was studied through the insertion of cohesive elements [2,19], which can give accurate
prediction for the percentage of transgranular fracture and intergranular fracture at different loading conditions. The
cohesive zone model’s extension to unknown crack paths entails the insertion of cohesive elements between each pair
of continuum elements in the finite element mesh, thus having a strong mesh dependency and an over-estimation of
the crack area [22]. Another concern is that adaptive meshing has to be performed in the cohesive zone model to
alleviate the cracks’ tendency to propagate along the existing element boundaries [19].
Phase-field model has been widely used in predicting the crack propagation behavior recently [1,23–29], in which
order parameters are employed to define the damaged and undamaged materials. Compared to the above sharp
interface models [2,13–20], the phase-field model has a decisive advantage as the explicit interface tracking can
be avoided [1]. The existing researches about the phase-field cracking mainly focused on material anisotropy [30,31],
plasticity [22,32–34] and proposing new energy formulations [35–37]. Karma et al. [30] firstly proposed a phase-field
model to predict the crack path in anisotropic brittle materials and found that the path is uniquely determined by
the directional anisotropy of fracture energy, independent of details of the failure process. Li et al. [31] presented
a phase-field model for strongly anisotropic fracture with smooth local maximum entropy approximants in a direct
Galerkin method, contrast to the previous framework only allowing weak anisotropy. Another research focus is to
integrate different plastic constitutive model, such as the elasto-plasticity [22] and viscoplasticty [34], into the fracture
phase-field model. Shanthraj et al. [32] further proposed a finite-strain anisotropic phase field method by considering
both anisotropic fracture energy and plastic constitutive models to simulate the cleavage fracture with a high order
crystal symmetry. In addition, much effort has been put into developing new formulations and algorithms to reduce
the computational cost [35], to tackle nonlinear behavior [36] and to improve the convergence rate of the numerical
solution [37]. Only few works studied the crack deflection/penetration by the phase-field model. Paggi et al. [38] had
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1087

successfully integrated the phase-field model and the cohesive zone model [19], in which the phase-field model was
used to simulate the crack propagation in the bulk, and the cohesive zone model for the crack deflection at interfaces.
In addition, the approach developed in [38] had been further extended to ceramic laminates [39], and anisotropic
polycrystals [40]. However, the results of such an integrated model are sensitive to the interactions between the two
fracture length scales of the phase-field model and the cohesive zone model [38]. In addition, the integrated method
inevitably inherits the drawbacks of the cohesive zone model. Schneider et al. [1] and Oshima et al. [24] recently
proposed to combine a phase-field crack model with an established multiphase-field model. The combined model was
then utilized to investigate the effect of grain boundary energy on the crack deflection/penetration behavior and the
crack path in multiphase systems. However, the simulation results could not be compared with the LEFM predictions
and the experimental results directly, as the interface fracture toughness, not the grain boundary energy, is the
parameter commonly used in the traditional crack deflection/penetration simulations and experiments [8,11,12,41,42].
In the present work, a two-set order parameters phase-field model is herein developed to study the crack
deflection/penetration behavior in heterogeneous microstructures. One set of order parameters describes the crack field
and while the other for the microstructure field. The main difference between the present model and Paggi’s [38] is
the description of the interface fracture toughness. In [38], the interface properties are illustrated by the cohesive zone
model, in which the traction–separation profile is the crucial parameter. In the present model, however, the interface
fracture toughness is a function of the adjacent grains’ fracture toughness and the interface energy, which indicates that
the main difference between the interface and the bulk materials is the atomic arrangement and the resulted interface
energy. In contrast to Schneider’s model [1], the interface fracture toughness is theoretically correlated with the grain
boundary energy and/or interface energy through an interface weakening parameter in a thermodynamically consistent
manner in the present model. The interface fracture toughness and the bulk fracture toughness are automatically
incorporated through a unified form, facilitating the comparison with the theoretical LEFM predictions and the
experimental measurements. The phase-field model is then applied to a heterogeneous Si3 N4 polycrystal to explore the
effect of interface fracture toughness on the intergranular/transgranular fracture transition and the effective toughness
of the system. The established model allows modeling the complex crack patterns in the polycrystalline structure
involving grain boundaries and/or interfaces and can be easily applied to investigate the interaction between the crack
propagation and the microstructure evolution.

2. Analytical model of crack deflection/penetration


Crack propagation behaviors at interfaces have been intensively studied [3,4,8,38,43–47]. Based on the LEFM,
He and Hutchinson [3] proposed a general model for predicting the propagation direction of a crack approaching an
interface between elastic materials. They argued that cracks deflect at the interface if the following critical condition
is satisfied (otherwise, cracks penetrate into the bulk) [3],
G i (θ) Gi
b
> bc (1)
G Gc
where G ic and G bc are the fracture toughness of the interface and the bulk materials, respectively. G i (θ ) and G b are the
energy release rates of crack deflecting along the interface and penetrating into the bulk, respectively.
In the case of a crack penetrating into the bulk, the energy release rate can be calculated with the mode I stress
intensity factor K I as,
1 − ν2 2
Gb = KI (2)
E
where E is the Young’s modulus and ν is the Poisson ratio of the material.
For the case of a crack deflecting along the interface at an angle of θ , the energy release rate can be given by,
1 − ν2 [ i 2
G i (θ ) = K I (θ ) + K Ii I (θ )2
]
(3)
E
where K Ii (θ) and K Ii I (θ ) are the mode I and mode II stress intensity factors for a mixed-mode crack after the deflection,
respectively. They can be calculated by [2],
3 θ 1 3θ
K Ii (θ) = K I ( cos + cos ) (4)
4 2 4 2
1088 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

Fig. 1. The critical defection angle as a function of interface/bulk fracture toughness ratio according to linear elastic fracture mechanics [3].

1 θ 1 3θ
K Ii I (θ) = K I ( sin + sin ) (5)
4 2 4 2
With Eqs. (2)–(5), the ratio of G i (θ ) to G b can be expressed by,
[( ) ]
G i (θ) θ 3θ 2 θ 3θ 2
) (
1
= 3 cos + cos + sin + sin (6)
Gb 16 2 2 2 2
From Eqs. (1) and (6), the criterion of crack deflection/penetration [38] depends on the deflection angle and the
fracture toughness ratio, but not the stress intensity factor. The critical angle and fracture toughness ratio at which the
crack deflection transforms to the crack penetration are plotted by the solid line in Fig. 1. The solid line separates the
map into the deflection and penetration regions. For example, when the interface/bulk fracture toughness ratio is 0.5,
the critical deflection angle is 65◦ , i.e., the crack will be deflected by the interface with an impinging angle smaller
than 65◦ .

3. Phase-field model
3.1. Regularized framework for crack propagation in homogeneous media

We start from a brief introduction of the regularized formulation for phase-field modeling of crack propagation in
homogeneous media. The foundation of the phase-field model for brittle fracture is the variational principle based on
energy minimization as proposed by Francfort and Marigo∏ [48]. For an arbitrary brittle body Ω with a boundary ∂Ω
as shown in Fig. 2(a) [38], the total energy functional (u, Γ ) can be given by [23,38,49],
∏ ∏ ∏ ∫ ∫
(u, Γ ) = (u, Γ ) + (Γ ) = ψ(u)dV + G c dΓ (7)
Ω Γ Ω Γ

where ψ(u) is the elastic energy density and G c is∏the fracture toughness. The term Ω (u, Γ ) describes the elastic

strain energy stored in the damaged body, while Γ (Γ ) is the fracture resistance term associated with the energy
required to create new crack surfaces according to the Griffith criterion [50].
The crack initiation, propagation and branch are then determined by a minimization principle for the total energy
functional. However, the numerical treatment is rather complicated as a minimization has to be performed on a set
of possibly discontinuous functions [51]. To overcome the difficulties, the total energy functional is approximated
through smearing the discrete crack, as shown in Fig. 2(b), based on a regularized framework [38],
∏ ∫ ∫
(u, φc ) = h(φc ) f el (u)dV + G c d(φc , ∇φc )dV (8)
Ω Ω
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1089

Fig. 2. Schematic illustration of an arbitrary body with a crack and an interface/grain boundary [38]: (a) discrete crack surface and discrete interface
embedded into the solid; (b) smeared crack and diffusive interface; (c) distribution of two-set order parameters in the polycrystal domain with five
different grain orientations.

where φc is a phase-field order parameter describing a continuous transition between the intact solid state (φc = 0)
and the fully broken state (φc = 1). h(φc ) = (1 − φc )2 is a degradation function modeling the stiff loss between the
intact and the broken material [49], and the strain energy density is given by f el (u) = 12 εe (u) : Cεe (u) and depends on
the symmetric local elastic strain εe (u), where C is the elastic stiffness tensor of the solid phase.
In the regularized formulation, the crack surface density function, which is a convex function composed of a
quadratic term of φc and another quadratic term involving its gradient, is also introduced [23,38],
1 2
d(φc , ∇φc ) = φ + εc |∇φc |2 (9)
4εc c
The above energy functional is thermodynamic consistence and the integral of d(φc , ∇φc ) over the fractured
domain converges to the surface measure of the crack set as the crack length scale εc → 0 [23]. Fig. 3 shows the
equilibrium profiles of crack at different values of εc in a one-dimensional domain. Large values of εc can smoothen the
crack field, whereas the limit εc → 0 yields a discontinuous crack surface [49]. The convex quadratic potential cannot
simulate the irreversibility of fracture and additional constraints for the fracture field need to be formulated [25]. Hence
the homogeneous Dirichlet type conditions are imposed on the fracture field in the present simulation as follows [25],
φc (x, t) = 1 for t > tx∗ if φc (x, tx∗ ) = 1 (10)

3.2. Two-set order parameters phase-field model for crack propagation in heterogeneous media

The crack propagation model formulated in the homogeneous media is then extended to the heterogeneous media.
Here the regularized crack propagation model is combined with a polycrystal phase-field model to build a two-set
order parameters model by introducing additional microstructure order parameters φs , where s = 1 . . . N and N is the
number of grain orientations. In a polycrystal with five grain orientations as illustrated in Fig. 2(c), the microstructure
1090 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

Fig. 3. The equilibrium profiles of crack at different values of εc in a one-dimensional domain with a length of 2L.

order parameters φs are one inside the grains and gradually decrease to zero across the grain boundaries. The chemical
free energy density f chem (φc , φs ) related to microstructure then has to be added and the total energy functional
becomes,
∏ ∫
(u, φc , φs ) = h(φc ) f el (u, φs ) + G c (φs )d(φc , ∇φc ) + f chem (φc , φs )dV (11)

3.2.1. Chemical free energy


Based on phase-field models of grain growth [52–55], the chemical free energy density of a polycrystalline
microstructure can be expressed as,
N
∑ κs
f chem (φc , φs ) = f 0 (φc , φs ) + (∇φs )2 (12)
s=1
2
where κs is a positive gradient coefficient associated with the interfacial inhomogeneity, and f 0 (φc , φs ) is the local
free energy densities, which is defined as,
N
( N )2
α∑ 2 β ∑ 2
f 0 (φc , φs ) = − φ + φ
2 s=1 s 4 s=1 s
) (∑N ∑ N
) N
(13)
β
( ∑
+ γ− φs φg + γ φc
2 2 2
φs
2
2 s=1 g>s s=1

where α, β and γ are constants; for α = β > 0 and γ > β/2, the local free energy density has 2N degenerate
minima located at (φ1 , φ2 , . . . , φ N ) = (±1, 0, . . . , 0), (0, ±1, . . . , 0), . . . , (0, 0, . . . , ±1). The minima associated
with φs = −1 are eliminated by setting the value of each order parameter equal to its absolute value during the initial
stages of the simulation [54]. The additional coupling term of the crack field and microstructure field is added as
shown by the last term of Eq. (13), which will contribute to the potential at the interaction regions of grain boundary
and crack surface, so that the conversion in these regions becomes energetically unfavorable [1,56]. The form of
the coupling term is adapted from the one proposed by Chen et al. [54] in describing the interaction between the
recrystallization and deformed phases. At present, only above coupling form is utilized. In the following work, the
effect of coupling forms will be studied by trying different coupling forms, such as the one proposed by Schneider
et al. [1].

Remark. In the present phase-field model, the convex quadratic potential for crack is coupled with the double-well
potential for microstructure. The reason why we use the convex quadratic potential, φc2 , instead of the double well
potential, φc2 (1 − φc )2 , for crack is that with a double-well potential, crack field φc will converge to the mean-curvature
motion of its level lines [25,57]. It means if there exists curvature in the level line of crack field φc , the crack field will
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1091

continue to evolve. Hence the crack field φc may continue to propagate away from the crack tip due itself curvature,
even the external load is removed. In a curve crack, the φc may evolve in the lateral direction to form a “thick”
domain. Although the evolution rate of this effect is slow, the accuracy of the approximation of the crack surface
will be unacceptable in the case of the infinitely slow quasi-static crack propagation [57]. To avoid the problems, the
convex quadratic potential is adopted. Because we couple two different energy potentials to formulate the crack and
the microstructure fields that are featured with different physical behaviors, the setup of two-set order parameters is
utilized, and the proposed model is termed as “two-set order parameters phase-field model”.

3.2.2. Elastic energy


In the two-set order parameters model, the elastic energy density also depends on the local microstructure and can
be described by,
1
f el (u, φs ) = Ci jkl (φs )εiej (u)εkl
e
(u) (14)
2
where Ci jkl (φs ) denotes the microstructure-dependent elastic stiffness tensor.
Since the grains are rotated with respect to a fixed coordinate system, the elastic stiffness tensor for each grain is
obtained by transforming the tensor with respect to the fixed coordinate system. If C pqr t represents the stiffness tensor
for a single grain in its crystal coordinate frame, the microstructure-dependent elastic stiffness tensor for the entire
polycrystal in terms of the microstructure fields is given by [54],

Ci jkl (φs ) = p(φs )aisp a sjq akr
s s
alt C pqr t (15)
s

where p(φs ) = φs3 (6φs2


− 15φs + 10) is a interpolation function, and a is the transformation matrix representing the
rotation of the crystal coordinate system defined on a given grain ‘s’ with respect to the fixed coordinate system, which
is expressed in terms of the symmetric Euler angles {Ψ , Θ, φ} (Kocks convention) (in three dimensions) [54]:
− sin φsinΨ − cos φcosΨ cosΘ sin φcosΨ − cos φsinΨ cosΘ cos φsinΘ
⎛ ⎞

a = ⎝ cos φsinΨ − sin φcosΨ cosΘ − cos φcosΨ − sin φsinΨ cosΘ sin φsinΘ ⎠ (16)
cosΨ sinΘ sinΨ sinΘ cosΘ
where 0 ≤ Ψ ≤ 2π, 0 ≤ Θ ≤ π, 0 ≤ φ ≤ 2π.

3.2.3. Interface fracture toughness and grain boundary energy


To simulate the crack deflection, the fracture resistance term in Paggi’s work [38] had been split into the
corresponding counterparts associated with the dissipated energy of cracks in the bulk Γb and of cracks along the
existing interface Γi as follows,
∏ ∏ ∏ ∫ ∫
b
(Γ ) = + = G c dΓ + G ic dΓ (17)
Γ Γb Γi Γb Γi

where G bc is bulk fracture toughness and G ic is the interface fracture toughness ruled by a linear cohesive zone
model [38].
In the present two-set order parameters phase-field model (Fig. 2(c)), the interface behavior can be incorporated
into the bulk in a unified form through the microstructure order parameters φs . Hence the fracture resistance term can
be rewritten as,
∏ ∏ ∏ ∫
(Γ ) = + = G c (φs )dΓ (18)
Γ Γb Γi Γb +Γi

where the fracture toughness G c (φs ) depends on the local microstructures,


N

G c (φs ) = k(φs ) p(φs )G sc (19)
s=1

in which p(φs ) is the interpolation function same as that used in the elastic energy density and G sc is the bulk fracture
toughness of grain s. An interface weakening coefficient, k(φs ), is originally proposed in the present model to integrate
1092 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

the interface fracture toughness and the bulk fracture toughness through a unified form. The main difference between
the intrinsic interface, formed during phase transformation, and the bulk is the atomic arrangement, so a unified form
of the fracture toughness of interface and bulk is assumed in the present study. Due to the different or irregular atomic
arrangement of the interface, the energy of the interface may be higher than that of grain interior with periodic atomic
arrangements. Physically speaking, the release of extra interface energy will promote the break of chemical bonds and
the formation of fresh surface during fracture, leading to the weakening of the interface.
In the elastic-brittle Griffith model, all plastic flow processes are neglected. The critical driving force for crack
propagation in the homogeneous material is G c = 2E sur f , where E sur f is the intrinsic surface energy [58]. In the
heterogeneous material, where grain boundary and grain boundary energy exist, the crack propagation criteria may be
changed due to the interaction between the crack surface and the grain boundary. Considering an ideal condition that
the grain boundary energy E gb is totally released to form the new crack surfaces, the critical driving force for crack
propagation becomes G c = 2E sur f − E gb . Therefore, the lower bound of k(φs ) can be given by,
∑N
p(φs )G sc − E gb
klower (φs ) = s=1∑N (20)
s
s=1 p(φs )G c
Inside the solid phase region, where grain boundary energy is zero, k(φs ) = 1, which means no weakening effect.
At the interface region, klower (φs ) ≤ k(φs ) < 1. The smaller the interface weakening coefficient, the weaker the
interface is. Also, it is clear that the interface weakening coefficient and the interface fracture toughness both will
decrease as the increase of the grain boundary energy.
In addition, the grain boundary energy can be related to the gradient coefficient κs through the energy barrier height
w and interface thickness 2λ by [59,60],

κs · w
E gb = √ (21)
3 2
√ κs

2λ = a 2 (22)
w
where a is a constant which depends on the definition of the interface. The interface thickness is defined as a measure
of the changes of the microstructure field φs from 0.1 to 0.9, which can be considered as the length-scale of the
microstructure.

3.3. Evolution equations

With the total energy functional discussed above, kinetics of the crack propagation and microstructure evolution
can be obtained by solving the Allen–Cahn [61] equations,
G c (φs )
⎡ ⎤
−2ε G
c c s (φ )∆φ c + φ c
∂φc δ 2εc
∏ ⎢ ⎥
= −Mc = −Mc ⎢ ∂h(φ ) (23)
⎢ N

∂t δφc c
∑ ⎥
⎣+ f el (u, φs ) + 2γ φc φs
2⎦
∂φc s=1
⎡ ⎤
∂G c (φs ) 2 φ 2
c
∂φs δ
∏ (εc |∇φc | + )
= −Ms ⎢ ∂φs
⎢ ⎥
= −Ms
⎢ 4εc ⎥
(24)
∂t δφs

⎣ ∂ f el (u, φs ) ∂ f 0 (φc , φs ) ⎦
+h(φc ) + − κs ∆φs
∂φs ∂φs
where Mc and Ms are the mobility of crack surface and grain boundaries, respectively. To avoid crack healing, only
physically relevant changes, ∂φ
∂t
c
≥ 0, are accepted [1]. The partial differential for the local free energy density can be
written by,
N
∂ f 0 (φc , φs ) ∑
= −αφs + βφs3 + 2γ φs φg2 + 2γ φc2 φs (25)
∂φs g̸=s
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1093

Fig. 4. Sketch of the simulation domain and the loading conditions for crack deflection and penetration.

The elastic energy density used in the Allen–Cahn equations are obtained from the displacement fields,
∂ 2u − → du
ρ = ∇ · σ + η∇ 2 (26)
∂t 2 dt
where ρ is the density, η is the viscosity, and σ i j = h(φc )Ci jkl (φs )εkl
e
(u) are the components of Cauchy stress tensor.
The crack propagation and microstructure evolution are then simulated by solving the coupled equations for the
two-set order parameters and the displacement fields.

3.4. Numerical implementation

In the simulations, a rectangular domain of dimension Lx × L y , where Lx = 300 µm and L y = 200 µm, under plane
strain condition is considered as illustrated in Fig. 4. The heterogeneous microstructures including the matrix and a
layer of second phase, are firstly generated [60]. For studying the competition between deflection and penetration of a
crack impinging on an inclined interface, a pre-crack is initialized, terminating at an interface at the angle θ from the
horizontal axis.
The system is subjected to a uniform displacement ∆ on the lower and upper boundaries. A traction-free boundary
condition (σx x = σx y = 0) is applied to the left and right sides [62]. The Young’s modulus of the matrix is
E m = 210 GPa and the Poisson ratio is ν = 0.3. The modulus of the second phase layer is either set equal to
that of the matrix for purely studying the effect of interface fracture toughness on crack deflection/penetration or set
as a small or large value for studying the effect of modulus mismatch. In the current work, the fracture toughness of
the matrix and the second phase layer are set to be the same while the interface fracture toughness is controlled by the
interface weakening coefficient k(φs ). As the main purpose of the present study is to demonstrate the capability of the
proposed phase-field model in predicting the transition of crack deflection/penetration, the evolution of microstructure
is temporarily ignored. The dynamic fracture is analyzed by integrating Eqs. (23) and (26) through a staggered solution
scheme by using a finite difference scheme according the Murali’s work [62].

4. Results and discussions


4.1. Competition between crack penetration and deflection

The transition of crack types is shown in Fig. 5 as the decrease of the interface weakening coefficient with an
inclined interface of 30◦ . When the interface fracture toughness is equal to the bulk fracture toughness (k(φs ) = 1.0),
the crack penetrates straightly into the inclined layer as no interface exists. For a smaller interface fracture toughness
(k(φs ) = 0.75), small deviation in crack path is observed at the interface between the matrix and the second phase
layer. When k(φs ) = 0.5, an obvious crack deflection occurs. With the further decrease of the interface fracture
toughness, Fig. 5(e) and (f), the crack propagates along the interface for a relative large distance before it turns into
1094 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

Fig. 5. (a–f) Crack penetration to defection as the decrease of interface weakening coefficient with an interface inclination angle of 30◦ , the first
deflection indicated by the blue arrows and second deflection indicated by the white arrows; (g) and (h) crack deflection at the large elongated
grains in Si3 N4 ceramics observed by Sun et al. [11] . (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)

the second phase layer again as indicated by the blue arrows. When the crack meets the second interface pointed
by the white arrows, it is found to be deflected again as shown in Fig. 5(d), (e) and (f). The predicted double-
deflection behavior is similar to the experimental results shown in Fig. 5(g) and (h), which is observed in Si3 N4
with large elongated second phase grains by in situ scanning electron microscopy (SEM) observation [11]. However,
a quantitative comparison becomes somehow impossible, as the absolute value of the interface fracture toughness
of Si3 N4 in above experiments is unknown. The force–displacement curve of the simulation condition of Fig. 5(e),
i.e. with the interface weakening coefficient k = 0.375 and the interface inclination angle of 30◦ , is shown in Fig. 6.
As the propagation of crack, the force starts to decrease and the crack deflected by the interface will not change
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1095

Fig. 6. Force–displacement curve of the simulation of the interface weakening coefficient k = 0.375 and an interface inclination angle of 30◦ .

Fig. 7. Deflection vs. penetration at different interface/bulk fracture toughness ratios according to the present phase field model, linear elastic
fracture mechanics and experiments [41].

the descending trend. The force shows two plateaus when the crack type changes from deflection to penetration as
indicated by the arrows in Fig. 6, with the corresponding crack state identified by the cycles in Fig. 5(e).
The deflection/penetration transitions are then simulated for five different interface inclination angles, 15◦ , 30◦ ,
45 , 60◦ and 75◦ , under three different applied strains. The predicted critical deflection/penetration angles are plotted

in Fig. 7. For demonstrating the capability of present model in predicting crack deflection, the bulk fracture toughness
of the matrix and the second phase layer are set to be the same in the simulation. The interface fracture toughness
corresponding to different interface weakening coefficients is firstly calculated and the fracture toughness ratio is
further obtained by interface fracture toughness/bulk fracture toughness. Then the deflection/penetration behavior
of the present model can be compared with the prediction by LEFM and the results of experiments for the same
fracture toughness ratio. It can be seen that the critical deflection/penetration angles almost linearly increase as the
interface/bulk fracture toughness ratio decreases for all applied strains. For the comparison, the deflection/penetration
transition results predicted by the LEFM and measured from experiments are also plotted in Fig. 7. It is found that
(1) under the applied strain of 0.015 and 0.016, the critical deflection angle is about 25◦ smaller than that predicted
by the LEFM. And (2) as the applied strain decreases the critical deflection angle approaches to LEFM criterion. The
1096 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

threshold value for the start of crack propagation is defined as the critical strain which is 0.014 at the present setup
and is determined by slowly increasing the applied strain.
Mirkhalaf’s experiments about the bio-materials [41] show the evidence that, for G ic /G bc = 0.25, the measured
deflection to penetration transition occurs at the angle between 60◦ to 70◦ . As depicted in Fig. 7, however, the predicted
critical deflection angle from the LEFM criterion is 90◦ , which is considerably larger than the experimental result.
A similar trend was observed in the Si3 N4 ceramics by Kadin et al. [10], in which G ic /G bc = 0.37 and the critical
deflection angle was supposed to be about 77◦ based on the LEFM criterion; however, the crack was not deflected
at the angle of 40◦ according to the SEM observations. It can possibly be attributed to that the applied strain in
experiments could not be the critical strain exactly, so the LEFM criterion might overestimate the tendency of the
crack deflection when compared with the actual experimental conditions. On the contrary, the experimental result is
in a good agreement with the simulation results with a relative large strain.
The above relation between the present phase-field model and the LEFM criterion can be explained by the following
virtual experiments, in which the interface fracture toughness, G ic , and the bulk fracture toughness, G bc , are set as 0.5
and 1.0, respectively.
(1) At a small strain, the energy release rate for the crack deflection along the interface, G i , is set as 0.6 and the
energy release rate for crack penetration into the bulk, G b , is 1.0.
(2) With the increase of the applied strain, G i and G b will increase simultaneously to keep the ratio of 0.6 according
to Eq. (6), and the absolute values of G i and G b are assumed to be 0.9 and 1.5, respectively.
From the LEFM criterion, the crack will be deflected at the interface. Since the present phase-field model is based
on the Griffith theory [50], the crack will propagate along the direction at which the energy release rate is fastest,
which means if (G i − G ic ) > (G b − G bc ), the crack will propagate along the interface and if (G i − G ic ) < (G b − G bc ),
the crack will penetrate into the bulk. For the small strain condition (G i − G ic = 0.6 − 0.5 = 0.1) > (G b − G bc =
1.0 − 1.0 = 0), crack will be deflected to propagate along the interface; however, for the above large strain condition
(G i − G ic = 0.9 − 0.5 = 0.4) < (G b − G bc = 1.5 − 1.0 = 0.5), the crack will penetrate into the bulk. In short, the
effect of the applied strain is not considered in the LEFM model; however, the deflection/ penetration transition curve
can be shifted by the applied strain in the present phase-field model based on the Griffith theory, which means that the
LEFM criterion is a static and critical case of the present phase-field model.

4.2. Effect of modulus mismatch

In the simulations above, the Young’s modulus of the second phase layer is equal to that of the matrix. The effect of
Young’s modulus of the second phase layer on the crack propagation path will be studied in this section, and the results
are presented in Fig. 8. In the first case, interface weakening is not considered. If the Young’s modulus of the second
phase layer, E h , is larger than that of the matrix, E m , i.e. the modulus mismatch ratio, E h /E m , is larger than 1.0, the
crack goes downward at the first interface while goes upward at the second interface, as shown in Fig. 8(a) and (b). On
the contrary, if the Young’s modulus of the second phase layer is smaller than that of the matrix, the crack will first be
deflected upward then downward, as shown in Fig. 8(c) and (d). It is concluded that when the crack propagates from
the stiff material to the soft material, it will be attracted to the interface; however, when the crack propagates from the
soft material to the stiff material, it will be repulsed from the interface. The extent of the deflection is sensitive to the
value of the modulus mismatch ratio.
The mechanism of crack deflection related to the modulus mismatch can be explained from the change of the stress
distribution at the crack tip. For the condition at which the modulus of the second phase layer is larger than that of
the matrix, when the crack approaches the second phase layer, the stress concentration at the crack tips and the stress
concentration in the second phase layer due to the large modulus will interact with each other. The stress distribution
at the crack tip will be changed, hence, the stress strength factors. To minimize the total energy of the system, the
crack propagation path will incline to the stress distortion direction as indicated by the black arrow in Fig. 9(a). When
the crack tip finally meets the interface, the crack will penetrate into the second phase layer as observed in Fig. 9(b).
Fig. 9(c) shows that the stress distortion will be reversed when the crack propagates from the second phase layer to
the matrix. When the second phase layer is softer than the matrix, a different stress distortion and thus an opposite
deflection trend will be witnessed as shown in Fig. 9(d–f). It is noteworthy that the modulus mismatch will also
promote the crack deflection, analogous to the interface weakening.
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1097

Fig. 8. Effect of modulus mismatch on the crack propagation path, deflection direction indicated by the arrows.

Fig. 9. Evolution of the distribution of Syy at different modulus mismatch ratios as the crack approaching and penetrating the interface, stress
distortion direction indicated by the arrows.

One can infer that if the interface weakening is combined with the modulus mismatch, the crack deflection will be
intensified. The crack paths of k(φs ) = 0.5 with different modulus mismatch ratios are shown in Fig. 10. When no
modulus mismatch exists, the crack deflects slightly at the interfaces and the extent of the defections is same for the
1098 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

Fig. 10. Combined effect of modulus mismatch and interface weakening on the crack deflection.

first and the second interfaces, as marked by circles in Fig. 10(b). When the modulus mismatch ratio is 0.5 as shown
in Fig. 10(a), the crack goes upward at the first interface and then propagates downward in the second phase layer
due to the stress distortion effect resulted from the modulus mismatch. As the crack meets the second interface, it
will propagate along the interface for a long distance under the combined effect of modulus mismatch and interface
weakening. For the modulus mismatch of 2.0, in Fig. 10(c), the stress distortion effect is also observed; however, the
main deflection position is the first interface instead of the second one. Both the decrease and increase of the modulus
of the second phase layer can enhance the crack deflection, and the deflection position depends on relative value of
the moduli.

4.3. Crack morphology and fracture toughness of Si3 N4 polycrystal

The crack morphology, which is mainly controlled by the interface/bulk fracture toughness ratio and the grain
structure in the heterogeneous microstructure, is the key element for determining the effective fracture toughness [2].
Based on the above deflection/penetration simulations, the influence of interface/bulk fracture toughness ratio and
grain structure on the crack morphology can be qualitatively illustrated as shown in Fig. 11. When the interface/bulk
fracture toughness ratio is relatively high, the crack propagates in a mixing mode, i.e., intergranular and transgranular
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1099

Fig. 11. Crack morphology governing effects: (a) comparison between large and small interface/bulk fracture toughness ratios, grain square
describes the crack penetrating at high angle boundary; (b) and (c) grain size on crack deflection, green squares identify the maximum deflection
length at this condition.

cracking. The crack penetrates into the grains when the angle between the grain boundary and the crack is larger
than the critical deflection angle, as indicated by the green square in Fig. 11(a). With the decrease of interface
fracture toughness and thus the increase of critical deflection angle, a full intergranular fracture may be observed.
The crack propagation may also depend on grain morphology including grain shape and size. The effect of grain
size on crack propagation is shown in Fig. 11(b) and (c). The results show that the crack propagation mode changes
from a fully intergranular fracture to a mixing mode as the grain size increases. This phenomenon can be explained
by considering the fact that the crack deflection changes the stress strength factor and the driving force for the
crack propagation. As described in Fig. 5, the extent of deflection depends on the values of the interface fracture
toughness and inclination angles. Consequently, the crack may stop after it propagates a certain distance and change
the propagation direction. For small grains, the deflected crack can easily reach next grain boundary and continue to
propagate along grain boundaries. However, for the material with large grains, the deflected crack may stop at the
middle of the grain boundaries and penetrates into the bulk again, which will decrease the fraction of intergranular
fracture.
To demonstrate the effect of interface fracture toughness and grain sizes on crack prorogation behavior, the
developed phase-field model is applied to study the crack propagation in a heterogeneous polycrystal. As the main
focus of this part is to show the capability of the proposed model in predicting the transition of transgranular fracture to
intergranular fracture in polycrystals, the effect of grain orientations is temporarily ignored. The initial grain structure
and loading conditions are shown in Fig. 12. The crack propagation path in the polycrystal for different interface
fracture toughness are presented in Fig. 13(a)–(f). When the interface weakening coefficient is equal to 0.75, almost
1100 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

Fig. 12. Sketch of the simulation domain and the loading conditions for polycrystal.

Fig. 13. Crack morphology in polycrystal with different interface weakening coefficients, the major intergranular fracture identified by the circles
and the crack bulges indicated by the arrows.

no deflection is observed as shown in Fig. 13(a). With the interface weakening coefficient further decreasing to 0.625,
two obvious intergranular fractures at inclined angles about 25◦ and 15◦ occur as shown in Fig. 13(b). Also, the crack
has a tendency of branch as indicated by the small bulges on the crack in Fig. 13(e). For the interface weakening
coefficient of 0.125 in Fig. 13(f), large angle intergranular fracture, about 80◦ , can be activated. With the decrease of
the interface weakening coefficient, the crack shows a continuous transition of transgranular fracture to intergranular
fracture.
Kadin et al. [10] measured the R-curves of Si3 N4 ceramics with different grain sizes. They found that with a
20% increase in the representative length of the grains, the effective fracture toughness would increase about 19%.
To compare with these experimental results, the energy dissipated during the crack propagation is calculated on two
systems with different grain sizes, in which the equivalent grain size of the large one is 1.2 times of that of the small
one. As shown in Fig. 14, the relative dissipated energy shows a linear decrease with the decrease of the interface
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1101

Fig. 14. Relative dissipated energy in polycrystals of different grain sizes with the decrease of interface weakening coefficients.

Fig. 15. Intergranular (region 1, 3, 5, 7) and transgranular (region 2, 4, 6) fracture transition in polycrystals.

weakening coefficient. As expected, the dissipated energy of the large grain is higher than that of the small grain
for all interface weakening coefficients, which means that the fracture toughness of the large grain material is higher
than that of the small grain one. Considering that the interface weakening coefficient of Si3 N4 is about 0.30 – 0.40
in Kadin’s experiments [10,42], the calculated effective fracture toughness ratio of different grain size systems is
between 1.12 to 1.16, similar to that of Kadin’s experimental measurement [10].
For understanding the toughening mechanism of the grain size, a detailed observation of the crack path is
conducted. From Fig. 15, the crack path can be divided into the intergranular fracture mode, regions 1, 3, 5, 7, and
the transgranular fracture mode, regions 2, 4, 6. The transgranular fractures are always related to the large grains,
marked as grain A, grain B and grain C. When the crack approaches these grains, the crack is first deflected and
propagates along the grain boundaries. As described above, the driving force may be not large enough for a large
distance deflection, so transition from the intergranular to transgranular fracture occurs at the middle of these large
grain boundaries. If the grain size is relatively small, the crack may propagate straight along an inclined boundary
before it propagates along the boundaries paralleled to the x axis, as shown in region 3 of Fig. 15. As a consequence,
higher proportion of transgranular fracture and lower proportion of low angle intergranular fracture can be expected
with the increase of grain size.
1102 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

If the crack propagates along the grain boundaries and/or the interfaces, then the crack enters a mixed mode
condition, which will lead to an increase in apparent fracture toughness G ac [41],
G ac 1 G ic
= (27)
cos 4 θ2 G bc
( )
G bc
When θ is close to 0◦ , i.e. the weak interface is almost paralleled to x axis, the apparent fracture toughness is equal
to the interface fracture toughness. With the increase of the inclined angle θ , the apparent fracture toughness shows
a considerable increase with respect to the interface fracture toughness. At the critical deflection angle, the apparent
fracture toughness can even reach the bulk fracture toughness. The larger the deflection angle is, the smaller the
fracture toughness loss is. Hence the intergranular crack can be divided into two classifications, deflection at angles
closed to the critical deflection angle and deflection at angles closed to the x axis, based on the negative effect on
the fracture toughness loss. Hence crack propagating along the weak boundaries paralleled to the x axis is the most
detrimental fracture mode for the material toughness; however, the fracture toughness loss is relative small for large
angle intergranular crack. In summary, the decreasing proportion of low angle intergranular fracture and increasing
proportion of transgranular fracture in large grain microstructure accounts for its relative good fracture resistance.

5. Conclusions
In the present study, a two-set order parameters phase-field model is proposed to study the crack deflec-
tion/penetration behavior by integrating the interface fracture toughness and the bulk fracture toughness through a
unified form. The model is validated by comparing with the LEFM predictions and the existed experimental results.
The following conclusions can be drawn:
• The proposed phase-field model can simulate the crack penetration and deflection at the interface as well as
the transition from transgranular fracture to intergranular fracture in the polycrystal. And the present model can
capture the prediction of LEFM criterion naturally.
• The modulus mismatch combined with the interface weakening can facilitate the crack deflection at the interface;
however, the deflection position, at first or second interface, depends on the relative value of the moduli.
• The crack morphology of intergranular/transgranular fracture and thus the effective fracture toughness of
polycrystal brittle material can be predicted by the present model. The toughening mechanism of large grain
size is attributed to the decrease of low angle intergranular fracture and the increase of transgranular fracture.
The proposed model is a very promising simulation tool that allows modeling complex crack patterns in systems
involving grain boundaries and/or interfaces. Within the two-set order parameters setup, the present model can
incorporate the real material microstructure automatically and offers the possibility to investigate the crack
propagation and microstructure evolution process simultaneously, which will be present in our forthcoming paper.

Acknowledgments
Chi Zhang acknowledges the financial support from the National Natural Science Foundation of China (Grant No.
51771097), the National Program on Key Basic Research Project, China (973 Program, Grant No. 2015CB654802),
the National Magnetic Confinement Fusion Energy Research Project of China (Grant No. 2015GB118001) and the
Science Challenge Project, China (Grant No. TZ2018004). The authors also gratefully acknowledge the financial
support provided by Tsinghua University Scholarship Council for Graduate Overseas Studies, China. Lei Chen is
grateful for the financial support by National Science Foundation, United States under CBET-1604104.

References
[1] D. Schneider, E. Schoof, Y. Huang, M. Selzer, B. Nestler, Phase-field modeling of crack propagation in multiphase systems, Comput. Methods
Appl. Mech. Engrg. 312 (2016) 186–195.
[2] S.T. Mousavi, N. Richart, C. Wolff, J.-F. Molinari, Dynamic crack propagation in a heterogeneous ceramic microstructure, insights from a
cohesive model, Acta Mater. 88 (2015) 136–146.
[3] H. Ming-Yuan, J.W. Hutchinson, Crack deflection at an interface between dissimilar elastic materials, Int. J. Solids Struct. 25 (9) (1989)
1053–1067.
[4] D. Martínez, V. Gupta, Energy criterion for crack deflection at an interface between two orthotropic media, J. Mech. Phys. Solids 42 (8)
(1994) 1247–1271.
H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104 1103

[5] D. Leguillon, C. Lacroix, E. Martin, Interface debonding ahead of a primary crack, J. Mech. Phys. Solids 48 (10) (2000) 2137–2161.
[6] W. Yong-li, Crack tip stress singularities in bimaterial with an inclined interface, Int. J. Fract. 54 (4) (1992) R65–R72.
[7] M. Paggi, A. Carpinteri, On the stress singularities at multimaterial interfaces and related analogies with fluid dynamics and diffusion, Appl.
Mech. Rev. 61 (2) (2008) 020801.
[8] L.R. Xu, Y.Y. Huang, A.J. Rosakis, Dynamic crack deflection and penetration at interfaces in homogeneous materials: experimental studies
and model predictions, J. Mech. Phys. Solids 51 (3) (2003) 461–486.
[9] W. Lee, Y.-H. Yoo, H. Shin, Reconsideration of crack deflection at planar interfaces in layered systems, Compos. Sci. Technol. 64 (15) (2004)
2415–2423.
[10] Y. Kadin, S. Strobl, C. Vieillard, P. Wijnbergen, V. Ocelik, In-situ observation of crack propagation in silicon nitride ceramics, Procedia
Struct. Integr. 7 (2017) 307–314.
[11] E.Y. Sun, P.F. Becher, K.P. Plucknett, C.H. Hsueh, K.B. Alexander, S.B. Waters, K. Hirao, M.E. Brito, Microstructural design of silicon
nitride with improved fracture toughness: II, effects of yttria and alumina additives, J. Amer. Ceram. Soc. 81 (11) (1998) 2831–2840.
[12] H.J. Kleebe, G. Pezzotti, G. Ziegler, Microstructure and fracture toughness of Si3N4 ceramics: combined roles of grain morphology and
secondary phase chemistry, J. Amer. Ceram. Soc. 82 (7) (1999) 1857–1867.
[13] J. Dolbow, T. Belytschko, A finite element method for crack growth without remeshing, Internat. J. Numer. Methods Engrg. 46 (1) (1999)
131–150.
[14] J. Mediavilla, R. Peerlings, M. Geers, Discrete crack modelling of ductile fracture driven by non-local softening plasticity, Internat. J. Numer.
Methods Engrg. 66 (4) (2006) 661–688.
[15] Y. Jiang, T. Tay, L. Chen, Y. Zhang, Extended finite element method coupled with face-based strain smoothing technique for three-
dimensional fracture problems, Internat. J. Numer. Methods Engrg. 102 (13) (2015) 1894–1916.
[16] L. Chen, T. Rabczuk, S.P.A. Bordas, G. Liu, K. Zeng, P. Kerfriden, Extended finite element method with edge-based strain smoothing
(ESm-XFEM) for linear elastic crack growth, Comput. Methods Appl. Mech. Engrg. 209 (2012) 250–265.
[17] Y. Jiang, T. Tay, L. Chen, X. Sun, An edge-based smoothed xfem for fracture in composite materials, Int. J. Fract. 179 (1–2) (2013) 179–199.
[18] L. Chen, G. Liu, K. Zeng, A combined extended and edge-based smoothed finite element method (ES-XFEM) for fracture analysis of 2D
elasticity, 8 (04) (2011) 773–786.
[19] S.M.T. Mousavi, B. Hosseinkhani, C. Vieillard, M. Chambart, P.J.J. Kok, J.-F. Molinari, On the influence of transgranular and intergranular
failure mechanisms during dynamic loading of silicon nitride, Acta Mater. 67 (2014) 239–251.
[20] L. Chen, K. Yeap, K. Zeng, G. Liu, Finite element simulation and experimental determination of interfacial adhesion properties by wedge
indentation, Phil. Mag. 89 (17) (2009) 1395–1413.
[21] J. Mediavilla, R.H.J. Peerlings, M.G.D. Geers, Discrete crack modelling of ductile fracture driven by non-local softening plasticity, Internat.
J. Numer. Methods Engrg. 66 (4) (2006) 661–688.
[22] M. Ambati, T. Gerasimov, L. De Lorenzis, Phase-field modeling of ductile fracture, Comput. Mech. 55 (5) (2015) 1017–1040.
[23] C. Miehe, F. Welschinger, M. Hofacker, Thermodynamically consistent phase-field models of fracture: Variational principles and multi-field
FE implementations, Internat. J. Numer. Methods Engrg. 83 (10) (2010) 1273–1311.
[24] K. Oshima, T. Takaki, M. Muramatsu, Development of multi-phase-field crack model for crack propagation in polycrystal, Int. J. Comput.
Mater. Sci. Eng. 03 (02) (2014) 1450009.
[25] C. Kuhn, A. Schlüter, R. Müller, On degradation functions in phase field fracture models, Comput. Mater. Sci. 108 (2015) 374–384.
[26] S. Biner, S.Y. Hu, Simulation of damage evolution in composites: a phase-field model, Acta Mater. 57 (7) (2009) 2088–2097.
[27] P. Chakraborty, Y. Zhang, M.R. Tonks, Multi-scale modeling of microstructure dependent intergranular brittle fracture using a quantitative
phase-field based method, Comput. Mater. Sci. 113 (2016) 38–52.
[28] S. Zhou, X. Zhuang, H. Zhu, T. Rabczuk, Phase field modelling of crack propagation, branching and coalescence in rocks, Theor. Appl.
Fract. Mech. 96 (2018) 174–192.
[29] X.-H. Guo, S.-Q. Shi, L.-J. Qiao, Simulation of hydrogen diffusion and initiation of hydrogen-induced cracking in PZT ferroelectric ceramics
using a phase field model, J. Amer. Ceram. Soc. 90 (9) (2007) 2868–2872.
[30] V. Hakim, A. Karma, Crack path prediction in anisotropic brittle materials, Phys. Rev. Lett. 95 (23) (2005) 235501.
[31] B. Li, C. Peco, D. Millán, I. Arias, M. Arroyo, Phase-field modeling and simulation of fracture in brittle materials with strongly anisotropic
surface energy, Internat. J. Numer. Methods Engrg. 102 (3–4) (2015) 711–727.
[32] P. Shanthraj, B. Svendsen, L. Sharma, F. Roters, D. Raabe, Elasto-viscoplastic phase field modelling of anisotropic cleavage fracture, J.
Mech. Phys. Solids 99 (2017) 19–34.
[33] D. Schneider, F. Schwab, E. Schoof, A. Reiter, C. Herrmann, M. Selzer, T. Böhlke, B. Nestler, On the stress calculation within phase-field
approaches: a model for finite deformations, Comput. Mech. 60 (2) (2017) 203–217.
[34] N. Mozaffari, G.Z. Voyiadjis, Coupled gradient damage – viscoplasticty model for ductile materials: phase field approach, Int. J. Plast. 83
(2016) 55–73.
[35] M. Ambati, T. Gerasimov, L. De Lorenzis, A review on phase-field models of brittle fracture and a new fast hybrid formulation, Comput.
Mech. 55 (2) (2014) 383–405.
[36] C. Hesch, K. Weinberg, Thermodynamically consistent algorithms for a finite-deformation phase-field approach to fracture, Internat. J.
Numer. Methods Engrg. 99 (12) (2014) 906–924.
[37] M.J. Borden, T.J.R. Hughes, C.M. Landis, C.V. Verhoosel, A higher-order phase-field model for brittle fracture: formulation and analysis
within the isogeometric analysis framework, Comput. Methods Appl. Mech. Engrg. 273 (2014) 100–118.
[38] M. Paggi, J. Reinoso, Revisiting the problem of a crack impinging on an interface:a modeling framework for the interaction between the
phase field approach for brittle fracture and the interface cohesive zone model, Comput. Methods Appl. Mech. Engrg. 321 (2017) 145–172.
[39] V. Carollo, J. Reinoso, M. Paggi, Modeling complex crack paths in ceramic laminates: a novel variational framework combining the phase
field method of fracture and the cohesive zone model, J. Eur. Ceram. Soc. 38 (8) (2018) 2994–3003.
1104 H. Chen, C. Zhang, Q. Lu et al. / Computer Methods in Applied Mechanics and Engineering 347 (2019) 1085–1104

[40] M. Paggi, M. Corrado, J. Reinoso, Fracture of solar-grade anisotropic polycrystalline silicon: a combined phase field–cohesive zone model
approach, Comput. Methods Appl. Mech. Engrg. 330 (2018) 123–148.
[41] M. Mirkhalaf, A.K. Dastjerdi, F. Barthelat, Overcoming the brittleness of glass through bio-inspiration and micro-architecture, Nature
Commun. 5 (2014) 3166.
[42] L. Porz, S. Wei, J. Zhao, E.A. Patterson, B. Liu, Characterizing brittle fracture by modeling crack deflection angles from the microstructure,
J. Amer. Ceram. Soc. 98 (12) (2015) 3690–3698.
[43] T. Cook, F. Erdogan, Stresses in bonded materials with a crack perpendicular to the interface, Internat. J. Engrg. Sci. 10 (8) (1972) 677–697.
[44] F. Erdogan, V. Biricikoglu, Two bonded half planes with a crack going through the interface, Internat. J. Engrg. Sci. 11 (7) (1973) 745–766.
[45] J.W. Hutchinson, Z. Suo, Mixed mode cracking in layered materials, Adv. Appl. Mech. 29 (1991) 63–191.
[46] W. Lee, S. Howard, W. Clegg, Growth of interface defects and its effect on crack deflection and toughening criteria, Acta Mater. 44 (10)
(1996) 3905–3922.
[47] X. Zeng, Y. Wei, Crack deflection in brittle media with heterogeneous interfaces and its application in shale fracking, J. Mech. Phys. Solids
101 (2017) 235–249.
[48] G.A. Francfort, J.-J. Marigo, Revisiting brittle fracture as an energy minimization problem, J. Mech. Phys. Solids 46 (8) (1998) 1319–1342.
[49] C. Kuhn, R. Müller, A continuum phase field model for fracture, Eng. Fract. Mech. 77 (18) (2010) 3625–3634.
[50] A.A. Griffith, M. Eng, VI, The phenomena of rupture and flow in solids, Philos. Trans. R. Soc. Lond. A Math. Phys. Eng. Sci. 221 (582–593)
(1921) 163–198.
[51] C. Kuhn, R. Müller, A phase field model for fracture, Pamm 8 (1) (2008) 10223–10224.
[52] L.-Q. Chen, Phase-Field models for microstructure evolution, Ann. Rev. Mater. Res. 32 (1) (2002) 113–140.
[53] S. Hu, L. Chen, A phase-field model for evolving microstructures with strong elastic inhomogeneity, Acta Mater. 49 (11) (2001) 1879–1890.
[54] L. Chen, J. Chen, R.A. Lebensohn, Y.Z. Ji, T.W. Heo, S. Bhattacharyya, K. Chang, S. Mathaudhu, Z.K. Liu, L.Q. Chen, An integrated fast
Fourier transform-based phase-field and crystal plasticity approach to model recrystallization of three dimensional polycrystals, Comput.
Methods Appl. Mech. Engrg. 285 (2015) 829–848.
[55] C. Krill Iii, L.-Q. Chen, Computer simulation of 3-D grain growth using a phase-field model, Acta Mater. 50 (12) (2002) 3059–3075.
[56] B. Nestler, H. Garcke, B. Stinner, Multicomponent alloy solidification: phase-field modeling and simulations, Phys. Rev. E 71 (4) (2005)
041609.
[57] B. Bourdin, C.J. Larsen, C.L. Richardson, A time-discrete model for dynamic fracture based on crack regularization, Int. J. Fract. 168 (2)
(2010) 133–143.
[58] D. Maugis, Subcritical crack growth, surface energy, fracture toughness, stick–slip and embrittlement,, J. Mater. Sci. 20 (9) (1985)
3041–3073.
[59] S.Y. Hu, J. Murray, H. Weiland, Z.K. Liu, L.Q. Chen, Thermodynamic description and growth kinetics of stoichiometric precipitates in the
phase-field approach, CALPHAD 31 (2) (2007) 303–312.
[60] H. Chen, Y. Ji, C. Zhang, W. Liu, H. Chen, Z. Yang, L.-Q. Chen, L. Chen, Understanding cementite dissolution in pearlitic steels subjected
to rolling-sliding contact loading: a combined experimental and theoretical study, Acta Mater. 141 (2017) 193–205.
[61] S.M. Allen, J.W. Cahn, A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening, Acta Metall.
27 (6) (1979) 1085–1095.
[62] P. Murali, T.K. Bhandakkar, W.L. Cheah, M.H. Jhon, H. Gao, R. Ahluwalia, Role of modulus mismatch on crack propagation and toughness
enhancement in bioinspired composites, Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 84 (1 Pt 2) (2011) 015102.

You might also like