You are on page 1of 16

Mechanics of Materials 188 (2024) 104867

Contents lists available at ScienceDirect

Mechanics of Materials
journal homepage: www.elsevier.com/locate/mecmat

Research paper

A residual stress-dependent mixed-mode phase-field model: Application to


assessing the role of tailored residual stresses on the mechanical performance
of ceramic laminates
Akash Kumar Behera a , Mohammad Masiur Rahaman a ,∗, Debasish Roy b
a
School of Infrastructure, Indian Institute of Technology Bhubaneswar, Odisha 752050, India
b Centre of Excellence in Advanced Mechanics of Materials; Indian Institute of Science, Bangalore 560012, India

ARTICLE INFO ABSTRACT

Keywords: Ceramics offer several attractive properties of industrial relevance, e.g. high strength and hardness, low
Ceramic laminates thermal conductivity, and chemical inertness in critical environments. There is thus interest among researchers
Tailored residual stress to improve the fracture toughness of ceramics, which is generally low and considered a major drawback.
Fracture toughness
Numerous experimental studies in the literature report on the enhancement of mechanical performance of
Mixed-mode
ceramic laminates, especially fracture toughness, by tailoring the residual stress, resulting in crack deflection
Phase field
Thermodynamic consistency
or bifurcation. However, there is a dearth of computational models that can reliably predict the crack path and
accurately quantify the improved fracture toughness. In this article, we propose a residual stress-dependent
mixed-mode phase-field model within a small deformation set-up. The proposed model is efficient in including
any energy dissipation effect in a consistent manner. The model can be exploited as a tool to study the effect
of tailored residual stresses on the fracture toughness of ceramic laminates. We have validated our proposal
by reproducing the results for a few problems that require the incorporation of residual stresses within the
formulation. By conducting a set of four-point bending tests on notched composite beams made of alternating
layers of Al2 O3 with 5% tetragonal ZrO2 (ATZ) and Al2 O3 with 30% monoclinic ZrO2 (AMZ), we demonstrate
how tailored residual stresses could indeed influence the mechanical performance of ceramic laminates.

1. Introduction chemical reactions, etc. leads to the appearance of residual stresses


in ceramic laminates (Sánchez-Herencia et al., 1999; Bermejo et al.,
Ceramics remain a focal point in contemporary research on com- 2006a). It has been realized that one may tailor the residual stresses
putational and experimental mechanics of materials, mainly owing to by controlling different design variables such as material properties,
such desirable properties as high strength and hardness, temperature thicknesses of alternating layers etc., in such a way that the surface
stability, chemical inertness, and high resistance to wear. Even with compressive stresses significantly increase fracture toughness, stress
a range of superior mechanical properties, the inherent brittleness of corrosion cracking, wear and impact damage resistance compared to
ceramics has limited their applications mainly to cases that predom- monolithic materials of the same composition.
inantly involve compressive loading. Targeting a wider application of In this context, we present a brief overview of the relevant literature
ceramic materials, several studies have been carried out to enhance the
on the mechanical behavior and design principles of ceramic laminates
fracture toughness through phase transformations, addition of fibers,
with residual stresses. Orlovskaya et al. (2001) conducted an inves-
control or design of micro-structures, and introduction of secondary
tigation on a silicon nitride-based ceramic laminate with controlled
phases (Yang et al., 2012; Bucevac et al., 2010; Hua et al., 2010; Becher
compressive and tensile stresses in separate layers. They observed crack
et al., 1998). However, these methods often entail incompatibly higher
bifurcation resulting from the existing residual stresses leading to an
costs of manufacturing vis-á-vis the benefits of increased fracture tough-
ness (Liu et al., 2016). A promising yet inexpensive way to enhance the increased toughness of the laminate. Further, Blattner et al. (2001)
mechanical properties including fracture toughness is the introduction examined the impact of the thickness ratio of compressive and tensile
of surface compressive stresses during the production of ceramic lami- layers in alumina–zirconia composites through three-point bending
nates. The mismatch in thermal expansion coefficients, Young’s moduli, tests on single-edge notched beam specimens. Qin et al. (2003) used the

∗ Corresponding author.
E-mail address: masiurr@iitbbs.ac.in (M.M. Rahaman).

https://doi.org/10.1016/j.mechmat.2023.104867
Received 15 September 2023; Received in revised form 10 November 2023; Accepted 12 November 2023
Available online 13 November 2023
0167-6636/© 2023 Elsevier Ltd. All rights reserved.
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

finite element method (FEM) for evaluating the residual stress within 2. A phase-field model with residual stresses
the layered ceramic laminate. They found a good agreement between
the residual stress obtained by using FEM and that by using X-ray stress In this section, we describe our residual stress-dependent mixed-
analysis. Lugovy et al. (2004) employed a compliance method to assess mode phase-field model in a small deformation setup. There are of
apparent fracture toughness and found it to align well with the exper- course several phase-field models that have been independently de-
imental findings. Gogotsi et al. (2004) utilized the R-curve method to veloped by researchers in the field of continuum mechanics (Francfort
investigate the relationship between the crack length parameter and and Marigo, 1998; Bourdin et al., 2000, 2008; Kuhn and Müller, 2008;
apparent fracture toughness for ceramic laminates. Bermejo and Danzer Amor et al., 2009; Kuhn and Müller, 2010; Miehe et al., 2010a; Borden
(2010) suggested that for increased toughness of layered ceramics with et al., 2014; Ambati et al., 2015; Dhas et al., 2018; Zhang et al., 2017;
a strong interface, a combination of crack bifurcation and delamination Spetz et al., 2021) and physics (Aranson et al., 2000; Karma et al., 2001;
Hakim and Karma, 2009; Spatschek et al., 2011; Eastgate et al., 2002;
should be pursued. Subsequently, several researchers (Orlovskaya et al.,
Henry and Levine, 2004). Adopting the idea of a mixed-mode phase-
2005; Sglavo et al., 2005; Sglavo and Bertoldi, 2006; de Portu and
field model in Spetz et al. (2021), we extend the work of Dhas et al.
Micele, 2007) proposed robust procedures for designing and manufac-
(2018) to integrate the effect of residual stresses within the model. It
turing ceramic laminates with higher compressive stress and reduced
also provides a facility to incorporate strictly dissipative effects in a
tensile stress within the layers to achieve increased toughness.
thermodynamically consistent manner whenever needed.
Available experimental evidence shows that the crack propagation
behavior in ceramic laminates, when strongly bonded interfaces are 2.1. Kinematics
introduced, is very different from that in a monolithic specimen of
the same composition and geometric dimensions (Liu et al., 2017; Consider a deformable body, the reference configuration of which is
Grigoriev et al., 2021; Wei et al., 2021). While it is recognized that an open set 𝛺 with a Lipschitz smooth boundary 𝜕𝛺. The infinitesimal
the presence of internal residual stresses in ceramic laminates leads to strain tensor (𝝐) can be expressed in terms of the gradient of the
drastic changes in crack propagation patterns, a model to reliably pre- displacement vector as
dict crack paths and accurately quantify the enhancement in fracture 1
toughness visá-vis monolithic samples seems to be missing. Typically, 𝝐= (∇𝒖 + ∇𝒖𝑇 ), (1)
2
numerical tools employed for crack path prediction fall into two cate- where 𝒖 denotes the displacement vector and ∇ denotes the gradient
gories: discrete and continuum approaches. Discrete approaches, such operator. To account for any pre-existing residual strain in the material
as linear elastic fracture mechanics (LEFM) (Pak, 1992; Suo et al., body, we assume an additive decomposition of strain as
1992; Gao et al., 1997; McMeeking, 2001; Zhang et al., 2002) and
cohesive zone modeling (Arias et al., 2006; Verhoosel et al., 2011; 𝝐 = 𝝐 elas + 𝝐 res , (2)
Linder et al., 2011) require explicit tracking of displacement discon- where 𝝐 elas is the elastic part of strain and 𝝐 res the residual strain
tinuities, making it challenging to capture complex crack paths. In Zobeiry and Poursartip (2015). Though a generalization to the finite-
contrast, continuum approaches adopt a diffuse crack representation deformation case, involving a multiplicative decomposition of the de-
in which the displacement field remains continuous but with reduced formation gradient, should be straightforward, it is not attempted here.
stiffness in the cracked regions. Consequently, continuum approaches As seen in Fig. 1, we introduce a length scale parameter 𝑙𝑠 , and
can predict crack paths by solving a set of governing partial differential phase field variable 𝑠 to represent crack as a continuous field within
equations without requiring an explicit tracking of cracks. Among the the domain. The phase-field 𝑠 ∈ [0, 1] may be loosely interpreted
various numerical methods in this category, a widely used technique as quantifying damaged, undamaged, or partially damaged states of
is that of phase field. This technique employs a phase field variable matter as follows: 𝑠 = 1 denotes the undamaged state, 𝑠 = 0 the fully
and a length scale parameter to provide a diffuse representation of damaged state and 𝑠 ∈ (0, 1) partially damaged state. The crack set
cracked surfaces; see Ambati et al. (2015) for a detailed review on may thus be defined as 𝛺𝑠 = {𝒙 ∈ 𝛺|𝑠(𝒙) = 0}. Reversal of 𝑠 from
phase field modeling of fracture. Since it is tricky to give a geometric the damaged state to partially damaged or undamaged states is not
or physically meaningful interpretation of the phase field variable in permissible, that is if 𝒙 ∈ 𝛺𝑠 at time 𝑡0 , then 𝒙 ∈ 𝛺𝑠 for all 𝑡 ≥ 𝑡0 .
the context of fracture in solids, we presently treat it just as an internal The deformed (and possibly damaged) state of the material body may
variable. The objective of this article is to develop a thermodynamically then be described by treating 𝑠 as an additional descriptor alongside
the displacement vector 𝒖.
consistent phase-field model that can serve as a computational tool
for predicting crack paths and possibly improved fracture toughness in
ceramic laminates with tailored residual stresses. 2.2. Force balances

To develop a model for residually stressed ceramic laminates, we


To describe the deformed state of the material body under quasi-
consider an additive decomposition of strain into elastic and residual
static loading, we consider the twin balances of the so-called macro-
strain parts. We adopt a mixed-mode phase field model, suitably modi-
and micro-forces derivable from a virtual power principle (Gurtin,
fied to include the effect of residual stresses. In the current formulation,
1996). While the stress tensor 𝝈, traction vector 𝒕(𝒏) and body force 𝒃
an additional driving term related to the residual stresses appears in the together define the macro-force system, the micro-force system consists
governing partial differential equations, thereby altering the evolution of scalar micro-traction 𝜒(𝒏), vector micro-stress 𝝃 and scalar micro-
of displacement as well as phase field. Section 2 describes the proposed stress 𝜋. The term ‘micro-force’ could potentially be misleading, as it
model. Section 3 presents a finite element discretization for numerically exists within the same continuum as the macro-system. As noted, one
implementing the model. In Section 4, several numerical examples are may derive the force balance equations via a virtual power principle.
provided to validate and assess the model. Furthermore, a set of four- Identifying the velocity vector 𝒖̇ (the time derivative of displacement
point bending tests on a ceramic laminate of alternating layers of ATZ 𝒖) as power conjugate to 𝒕(𝒏) and 𝒃, and 𝑠̇ as power conjugate to 𝜒(𝒏),
and AMZ is considered to study the effect of residual stress, number of one expresses the external power 𝑃 ext expended over any arbitrary part
layers, and thickness ratio between ATZ to AMZ on the performance of of the body  ⊂ 𝛺 as
ceramic laminates. Finally, the outcomes of this work are summarized [ ]
and conclusions are drawn in Section 5. 𝑃 ext = 𝒕 (𝒏) ⋅ 𝒖̇ + 𝜒(𝒏) 𝑠̇ 𝑑𝐴 + 𝒃 ⋅ 𝒖̇ 𝑑𝑉 , (3)
∫𝜕 ∫

2
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 1. A composite defined by an open set 𝛺 = 𝛺𝐴 ∪ 𝛺𝐵 ∪ 𝛺𝐶 , where the open sets 𝛺𝐴 , 𝛺𝐵 and 𝛺𝐶 define three constituents 𝐴, 𝐵 and 𝐶, respectively. The composite body is
subjected to Dirichlet and traction boundary conditions on 𝜕𝛺𝑢 and 𝜕𝛺𝑡 , respectively. Sub-figure (a) illustrates a sharp crack depiction within the composite, while sub-figure (b)
showcases a smoothed representation of the crack achieved through the utilization of parameters 𝑙𝑠 and 𝑠. The variable 𝑠 can vary between 0 and 1; 𝑠 = 0 indicates a damaged
state, 𝑠 = 1 indicates an undamaged state, and any intermediate value of 𝑠 indicates a partially damaged state.

with 𝒏 denoting the unit normal vector to the boundary 𝜕. In Eq. (3), Since Eq. (11) holds for all admissible 𝑠̃ and over any , Eq. (11) may
𝑑𝐴 and 𝑑𝑉 are the area and volume measures on 𝜕 and , respec- be localized as:
tively. The internal power 𝑃 int can be expressed as the summation of
𝜒(𝒏) = 𝝃 ⋅ 𝒏 on 𝜕, (12)
power expenditures of 𝝈 over ∇𝒖, ̇ 𝝃 over ∇𝑠̇ and 𝜋 over 𝑠,
̇ i.e.
and
𝑃 int = (𝝈 ∶ ∇𝒖̇ + 𝝃 ⋅ ∇𝑠̇ + 𝜋 𝑠)
̇ 𝑑𝑉 . (4)
∫ ∇ ⋅ 𝝃 − 𝜋 = 0 on . (13)
Introducing 𝒖̃ and 𝑠̃ as the virtual counterparts of 𝒖̇ and 𝑠, ̇ respectively, One may consider Eq. (12) as micro-traction and Eq. (13) as the
we define a set  = (𝒖, ̃ which may be referred to as the generalized
̃ 𝑠), micro-force balance, respectively.
virtual velocity vector. Then the external virtual power 𝑃̃ ext (; ) and
the internal virtual power 𝑃̃ int (; ) are given by 2.2.3. Thermodynamics and constitutive modeling
[ ] Constitutive relations for the macro- and micro-stresses must be
𝑃̃ ext (; ) = 𝒕 (𝒏) ⋅ 𝒖̃ + 𝜒(𝒏) 𝑠̃ 𝑑𝐴 + 𝒃 ⋅ 𝒖̃ 𝑑𝑉 , (5) arrived at in line with the first and second laws of thermodynamics.
∫𝜕 ∫
Assuming isothermal conditions, the second law of thermodynamics
and (free energy inequality) for  ⊂ 𝛺 may be stated as:

𝑃̃ int (; ) = (𝝈 ∶ ∇𝒖̃ + 𝝃 ⋅ ∇𝑠̃ + 𝜋 𝑠)


̃ 𝑑𝑉 . (6) 𝑑
∫ 𝜓𝑑𝑉 ≤ 𝑃 ext , (14)
𝑑𝑡 ∫
By utilizing Eqs. (5) and (6) and applying the virtual power principle, with 𝜓 denoting the Helmholtz free-energy density. Using Eq. (3),
which states 𝑃̃ ext (; ) = 𝑃̃ int (; ), we obtain Eq. (4), Eq. (14) and the power balance, i.e. 𝑃 ext = 𝑃 int in the last
[ ] inequality, one may arrive at
𝒕 (𝒏) ⋅ 𝒖̃ + 𝜒(𝒏) 𝑠̃ 𝑑𝐴 + 𝒃 ⋅ 𝒖̃ 𝑑𝑉 = ̃ 𝑑𝑉 . (7)
(𝝈 ∶ ∇𝒖̃ + 𝝃 ⋅ ∇𝑠̃ + 𝜋 𝑠)
∫𝜕 ∫ ∫
𝜓𝑑𝑉
̇ ≤ (𝝈 ∶ ∇𝒖̇ + 𝝃 ⋅ ∇𝑠̇ + 𝜋 𝑠)
̇ 𝑑𝑉 . (15)
∫ ∫
The macro- and micro-force balance equations can be derived by
appropriate substitution of  = (𝒖, ̃ in Eq. (7).
̃ 𝑠) Since 𝝈 is symmetric, we immediately have 𝝈 ∶ ∇𝒖̇ = 𝝈 ∶ 𝝐̇ = 𝝈 ∶ 𝝐̇ elas
(using Eq. (2) and considering 𝝐 res as time invariant). Substituting
𝝈 ∶ ∇𝒖̇ = 𝝈 ∶ 𝝐̇ elas in Eq. (15) and invoking the localization theorem,
2.2.1. Macro-force balance
the following inequality results.
Since 𝒖̃ and 𝑠̃ are virtual power conjugates to macro- and micro-
( )
forces, the macro-force balance equation may be obtained by letting 𝜓̇ − 𝝈 ∶ 𝝐̇ elas + 𝝃 ⋅ ∇𝑠̇ + 𝜋 𝑠̇ ≤ 0. (16)
 = (𝒖,
̃ 0) i.e., by substituting 𝑠̃ = 0 in Eq. (7), leading to
One must satisfy the inequality given in Eq. (16) whilst formulating the
constitutive relations of the fluxes 𝝈, 𝝃 and 𝜋.
(𝒕 (𝒏) − 𝝈𝒏) ⋅ 𝒖̃ 𝑑𝐴 = − (∇ ⋅ 𝝈 + 𝒃) ⋅ 𝒖̃ 𝑑𝑉 . (8)
∫𝜕 ∫
Since Eq. (8) holds for all admissible 𝒖,
̃ Eq. (8) may be localized as: 2.3. Constitutive response functions

𝒕 (𝒏) = 𝝈𝒏 on 𝜕, (9) Let the free energy density 𝜓 depend on 𝝐 elas , ∇𝑠, 𝑠 so that it is
expressed as
conventionally is recognized as the macro-traction condition and
̂ elas , ∇𝑠, 𝑠).
𝜓 = 𝜓(𝝐 (17)
∇ ⋅ 𝝈 + 𝒃 = 0 on , (10)
Applying the chain rule in Eq. (17), we get the rate of free energy as
recognized as the macro-force balance.
𝜓̇ = 𝜕𝝐elas 𝜓 ∶ 𝝐̇ elas + 𝜕∇𝑠 𝜓 ⋅ ∇𝑠̇ + 𝜕𝑠 𝜓 𝑠,̇ (18)

2.2.2. Micro-force balance with 𝜕(.) denoting the partial derivative of a function with respect to
For deriving the micro-force balance we consider  = (𝟎, 𝑠).
̃ Hence, the argument in the suffix whilst keeping others fixed. Assuming 𝜋 =
by substituting 𝒖̃ = 𝟎 in Eq. (7), we get 𝜋 en + 𝜋 dis , where (.)en and (.)dis are used for energetic and a dissipative
[ ] part of its argument, and substituting Eq. (18) in Eq. (16), we get
𝜒(𝒏) − 𝜉 ⋅ 𝒏 𝑠̃ 𝑑𝐴 = − (∇ ⋅ 𝝃 − 𝜋) 𝑠̃ 𝑑𝑉 . (11) (𝝈 − 𝜕𝝐elas 𝜓) ∶ 𝝐̇ elas + (𝝃 − 𝜕∇𝑠 𝜓) ⋅ ∇𝑠̇ + (𝜋 en − 𝜕𝑠 𝜓)𝑠̇ + 𝜋 dis 𝑠̇ ≥ 0. (19)
∫𝜕 ∫

3
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Utilizing the Coleman–Noll procedure (Coleman and Noll, 1974), on In Eq. (28), the term 𝑠2 is the degradation function that considers
Eq. (19), we obtain the decrease in material stiffness due to damage. In Eq. (29), and
Eq. (30), 𝑘𝑛 = 𝜆 + 2𝜇 𝑛
, where 𝜆, and 𝜇 are Lamè parameters and 𝑛 is
𝝈 = 𝜕𝝐 elas 𝜓, (20)
the dimensional order of the problem. The Lamé parameters can be
evaluated as 𝜆 = 𝐸𝜈∕ ((1 + 𝜈)(1 − 2𝜈)) and 𝜇 = 𝐸∕ (2(1 + 𝜈)), where
𝝃 = 𝜕∇𝑠 𝜓, (21) 𝐸 represents Young’s modulus and 𝜈 represents Poisson’s ratio. We
postulate the fracture energy density 𝜓 frac as
𝜋 en = 𝜕𝑠 𝜓, (22) ( )
(1 − 𝑠)2 𝑙𝑠
𝜓 frac (∇𝑠, 𝑠) = 𝐺𝑐 + ∇𝑠 ⋅ ∇𝑠 , (31)
The inequality in Eq. (19) then reduces to 2𝑙𝑠 2

𝜋 dis 𝑠̇ ≥ 0. (23) with 𝐺𝑐 denoting the critical energy release rate. Using Eqs. (20), (21),
(22), (29), (30) and (31), we get 𝝈, 𝝃 and 𝜋 en as
The inequality constraint given by Eq. (23) can be satisfied by choosing ( )
𝜋 dis = ̂𝑠,
̇ where ̂ is a constitutive function which should satisfy ̂ ≥ 0 𝝈 = 𝜕𝝐elas 𝜓 = 𝑠2 𝑘𝑛 ⟨tr(𝝐 elas )⟩+ + 2𝜇𝝐 elas
dev
+ 𝑘𝑛 ⟨tr(𝝐 elas )⟩− + 2𝜇𝝐 elas
dev
(32)
+ −
as we must comply with 𝑠̇ ≤ 0.
𝝃 = 𝜕∇𝑠 𝜓 = 𝐺𝑐 𝑙𝑠 ∇𝑠, (33)
2.4. Specialized constitutive relations
and
To evaluate the thermodynamic forces in terms of the kinematic 𝐺𝑐
forces, we specialize the constitutive relations by explicitly postulating 𝜋 en = 𝜕𝑠 𝜓 = 2𝑠𝜓+elas (𝝐 elas ) − (1 − 𝑠) . (34)
𝑙𝑠
the free energy density 𝜓 in terms of 𝝐 elas , ∇𝑠 and 𝑠. We assume that
Here, we use a hybrid approach suggested by Ambati et al. (2015),
𝜓(𝝐 elas , ∇𝑠, 𝑠) can be written as a sum of the elastic energy density
following which we express 𝝈 as
𝜓 elas (𝝐 elas , 𝑠) and the fracture energy density 𝜓 frac (∇𝑠, 𝑠),
( )
𝜓(𝝐 elas , ∇𝑠, 𝑠) = 𝜓 elas (𝝐 elas , 𝑠) + 𝜓 frac (∇𝑠, 𝑠). (24) 𝝈 = 𝑠2 𝜆tr(𝝐 elas ) + 2𝜇𝝐 elas , (35)

To ensure that cracks do not propagate under pure compression, re- and the tension–compression asymmetry is accounted only in the eval-
searchers generally adopt a volumetric–deviatoric decomposition (Amor uation of 𝜓 elas , appearing in the expression of 𝜋 en . Substituting the
et al., 2009), or a spectral decomposition (Miehe et al., 2010a) of strain above expressions in the macro- and micro-force balance equations, the
tensor. However, both of these models come with their own set of lim- governing partial differential equations can be derived in terms of the
itations. For instance, the volumetric–deviatoric decomposition-based state variables 𝒖 and 𝑠.
approach may yield inaccurate results when subjected to transverse
compression, and the spectral decomposition-based model may produce 3. Boundary value problem and finite element discretization
incorrect outcomes under pure shear, as reported in the literature
(Prakash et al., 2023; Liu et al., 2021; Agrawal, 2016). In this work, Having presented the strong form of the boundary value prob-
we employ a recent decomposition proposed by Spetz et al. (2021), lem described through partial differential equations with appropriate
which combines aspects of both volumetric–deviatoric and spectral boundary conditions, we provide the finite element formulation for the
decompositions to provide a more accurate approximation of crack model using the weak form of the governing equations. Substituting the
propagation in mixed-mode conditions. We decompose 𝝐 into its vol- expression of 𝝈 in Eq. (35) in Eq. (10), we get
umetric component (𝝐 elas ) and deviatoric component (𝝐 elas ) and then ( )
vol dev ( )
decompose the deviatoric component into its corresponding positive ∇ ⋅ 𝑠2 𝜆tr(𝝐 elas ) + 2𝜇𝝐 elas + 𝒃 = 0. (36)
(𝝐 elas
dev
) and negative (𝝐 elas
dev
) parts.
+ −
In the absence of body forces, Eq. (36) may be modified using Eq. (2)
elas
𝝐 = 𝝐 elas
vol
+ 𝝐 elas
dev+
+ 𝝐 elas
dev−
, (25) as
( ) ( )
where, ( ) ( )
∇ ⋅ 𝑠2 𝜆tr(𝝐) + 2𝜇𝝐 = ∇ ⋅ 𝑠2 𝜆tr(𝝐 res ) + 2𝜇𝝐 res . (37)
𝝐 elas
vol
= Pvol 𝝐 elas , and 𝝐 elas
dev
= Pdev 𝝐 elas (26)
Similarly, substituting the value of 𝝃 and 𝜋 en in Eq. (13), we have
and ( )
𝐺
∇ ⋅ 𝐺𝑐 𝑙𝑠 ∇𝑠 − 2𝑠 𝜓+elas (𝝐 elas ) + 𝑐 (1 − 𝑠) − ̂𝑠̇ = 0. (38)
𝝐 elas
dev
elas
= ⟨𝜖dev ⟩+ 𝒏𝑖 ⊗ 𝒏𝑖 , and 𝝐 elas
dev
elas
= ⟨𝜖dev ⟩− 𝒏𝑖 ⊗ 𝒏𝑖 . (27) 𝑙𝑠
+ 𝑖 − 𝑖

In (26), Pvol and Pdev are fourth-order tensors utilized for the pro- Eq. (38) may further be modified as
jection of second-order tensors onto their volumetric and deviatoric ( )
𝜓+elas (𝝐 elas ) 1
elas are the principal strains of
constituents, respectively. In Eq. (27) 𝜖dev ∇ ⋅ 𝑙𝑠 ∇𝑠 − 2𝑠 + (1 − 𝑠) − 𝑠̇ = 0, (39)
𝑖 𝐺𝑐 𝑙𝑠
𝝐 elas
dev
, 𝒏𝑖 are the corresponding principal strain directions, ⟨𝜖dev
elas ⟩
+ = 𝑖
elas + |𝜖 elas |), and ⟨𝜖 elas ⟩ = (1∕2)(𝜖 elas − |𝜖 elas |). We postulate with  = ∕𝐺 ̂ 𝑐 . In order to impose the irreversibility constraint as
(1∕2)(𝜖dev dev𝑖 dev𝑖 − dev𝑖 dev𝑖
𝑖
the elastic part of the free energy density as noted in Section 2.3, we use a method suggested by Miehe et al.
( ) ( ) ( ) (2010a). Specifically, we introduce (), a history function of  =
𝜓 elas 𝝐 elas = 𝑠2 𝜓+elas 𝝐 elas + 𝜓−elas 𝝐 elas , (28) 𝜓+elas (𝝐 elas )∕𝐺𝑐 . Mathematically the history function can be expressed
elas
( elas ) as
where 𝜓+ 𝝐 contributes towards crack propagation and is given
as (𝑓 (𝑡)) = max 𝑓 (𝜏), (40)
( ) 1 𝜏∈[0,𝑡]
𝜓+elas 𝝐 elas = 𝑘𝑛 ⟨tr(𝝐 elas )⟩2+ + 𝜇(𝝐 elas
dev+
∶ 𝝐 elas
dev+
), (29)
2 for any time-dependent function 𝑓 . Adopting the idea in Spetz et al.
( ) (2021), we further modify  to accommodate mixed-mode effects
whereas 𝜓−elas 𝝐 elas does not contribute towards crack development
and is given as involving mode-I and mode-II. We express  as
( ) 1 1 𝜇
𝜓−elas 𝝐 elas = 𝑘𝑛 ⟨tr(𝝐 elas )⟩2− + 𝜇(𝝐 elas ∶ 𝝐 elas ). (30) = 𝑘 ⟨tr(𝝐 elas )⟩2+ + (𝝐 elas ∶ 𝝐 elas ), (41)
2 dev− dev− 2𝐺cI 𝑛 𝐺cII dev+ dev+

4
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

where the critical energy release rates 𝐺cI and 𝐺cII are for mode-I Table 1
Material properties of ATZ and AMZ.
and mode-II, respectively. Using the history function, Eq. (39) can be
expressed as Material Property ATZ AMZ
( ) 𝐸 (GPa) 390 280
1 𝜈 0.22 0.22
∇ ⋅ 𝑙𝑠 ∇𝑠 − 2𝑠() + (1 − 𝑠) − 𝑠̇ = 0. (42)
𝑙𝑠 𝛼 × 10−6 (K−1 ) 9.8 8.0
𝜅𝐼𝐶 (MPa m−1∕2 ) 3.2 2.6
Note that the history-dependent term renders Eq. (42) to be different
𝐺𝑐 (J∕m2 ) 25 23
from Eq. (39) which was derived based on thermodynamic consistency.
In this article, we have opted for the history function-based approach
for the numerical implementation of the model as it effectively han-
dles the irreversibility of crack propagation (Miehe et al., 2010a,b). Table 1. Our aim is to demonstrate how tailoring residual stresses
However, for the imposition of the irreversibility of crack propagation, could improve mechanical performance. We use thermal residual stress,
there are other methods (Gerasimov and De Lorenzis, 2019) which which is induced due to a mismatch in the thermal expansion coeffi-
we plan to explore in a separate study. Eqs. (37) and (42) along with cients (𝛼) of ATZ and AMZ during the manufacturing process, as an
their boundary conditions can be considered the strong form of the input parameter to the model. Thus we study how thermal residual
governing equations for 𝒖 and 𝑠. The equations given by Eqs. (37) and stresses affect the fracture response of ceramic laminates. The calcu-
(42) are coupled and subject to boundary conditions, viz. prescribed lation of residual stress for any change in temperature (𝛥𝑇 ) is briefly
displacement and traction on 𝜕𝛺𝑢 and 𝜕𝛺𝑡 , respectively. described in Appendix A. The validation of our model is undertaken by
Using the strong form as above, we may obtain the corresponding reproducing the results in Salvati (2021) for tension test on a notched
weak form of the governing equations, thus setting the stage for a plate with a circular inclusion exhibiting known residual strain. Sub-
finite element discretization. We regard 𝒖̃ and 𝑠̃ as the test functions sequently, we attempt to capture the experimental fracture pattern of
associated with the 𝒖 and 𝑠, respectively. The trial space, considering ATZ–AMZ ceramic laminates with residual stress as reported in the
the prescribed displacement boundary condition can be given as literature, followed by a parametric study to examine how factors such
as residual stress magnitude, number of layers, and thickness ratio of
𝐻𝑢 = {𝒖 ∈ 𝐻 1 (𝛺); 𝒖 = 𝒖̄ on 𝜕𝛺𝑢 }, (43)
ATZ and AMZ layers influence the mechanical performance.

𝐻𝑠 = {𝑠 ∈ 𝐻 1 (𝛺)}, (44) 4.1. Numerical validation of the proposed model


with 𝒖̄ being the prescribed displacement on 𝜕𝛺𝑢 . The test space is
defined as First, we reproduce the findings of Salvati (2021). Thus, we con-
sider a notched plate exhibiting residual strain on a circular inclusion,
𝑉𝑢 = {𝒖̃ ∈ 𝐻 1 (𝛺); 𝒖̃ = 0 on 𝜕𝛺𝑢 }, (45) supported at the bottom and loaded at the top edge. Fig. 2 describes
the test setup and a representative finite element mesh considered for
𝑉𝑠 = {𝑠̃ ∈ 𝐻 1 (𝛺)}. (46) numerical simulation. The circular section in the test setup is subjected
to a residual strain (𝝐 res ); three different cases of hydrostatic residual
The finite element sub-spaces of test and trial functions may be intro- strain with magnitudes (𝜖 ∗ ) of −0.0015, 0.0, and 0.0015 are considered.
duced as 𝐻𝑢ℎ ⊂ 𝐻𝑢 , 𝐻𝑠ℎ ⊂ 𝐻𝑠 , 𝑉𝑢ℎ ⊂ 𝑉𝑢 , and 𝑉𝑠ℎ ⊂ 𝑉𝑠 . Now the weak For the numerical work, we use a plane stress approximation and
forms of Eqs. (37) and (42) may be written as
choose the modulus of elasticity (𝐸) as 70 GPa, and Poisson’s ratio
( )
( ) (𝜈) as 0.3. For this problem, we consider the same critical energy
(𝑠ℎ )2 𝜆tr(𝝐 ℎ ) + 2𝜇𝝐 ℎ ∶ 𝝐̃ ℎ 𝑑𝛺 = 𝝈 res ∶ 𝝐̃ ℎ 𝑑𝛺, (47)
∫𝛺 ∫𝛺 release rate for both fracture modes i.e, 𝐺cI = 𝐺cII = 4.096 N/mm; the
and length scale parameter (𝑙s ) is 1 mm. We do not include any dissipative
( ) effect in this example ( = 0 Sec./mm). Crack profiles for different
1 1 ℎ
𝑙 ∇𝑠ℎ ⋅ ∇𝑠̃ℎ + 2𝑠ℎ 𝑠̃ℎ ( ℎ ) + 𝑠ℎ 𝑠̃ℎ + 𝑠̇ℎ 𝑠̃ℎ 𝑑𝛺 = 𝑠̃ 𝑑𝛺, (48) cases are shown in Fig. 3. In the presence of hydrostatic compressive
∫𝛺 𝑠 𝑙𝑠 ∫𝛺 𝑙𝑠
residual strain (𝜖 ∗ = −0.0015), the crack profile deviates from the
for 𝒖ℎ ⊂ 𝐻𝑢ℎ , 𝒖̃ ℎ ⊂ 𝑉𝑢ℎ , 𝑠ℎ ∈ 𝐻𝑠ℎ and 𝑠̃ℎ ∈ 𝑉𝑠ℎ . In Eqs. (47) and (48), circular section, effectively avoiding it. Conversely, with hydrostatic
𝝐 ℎ = 21 (∇𝒖ℎ + (∇𝒖ℎ )𝑇 ), and tensile residual strain (𝜖 ∗ = 0.0015), the crack profile exhibits a slight
( ) bending to accommodate passage through the circular inclusion. Fig. 4
𝝈 res = (𝑠ℎ )2 𝜆tr(𝝐 res ) + 2𝜇𝝐 res , (49) shows the normalized load–displacement curves for all cases. In these
In this article, we have used an open-source finite element package, curves, the maximum load attained by the specimen with zero residual
Gridap (Badia and Verdugo, 2020; Verdugo and Badia, 2022) in Ju- strain is taken as the reference load (𝑃𝑐 ). Correspondingly, the applied
lia for the numerical implementation of the proposed model using a displacement (𝑢̄ 2 ) at which the zero-residual-strain specimen achieves
staggered scheme (see Julia codes for a phase-field fracture modeling 𝑃𝑐 is taken as the reference applied displacement (𝑢̄ 2𝑐 ). As seen from
of ceramic laminates under residual stress). One may consider the the figure, the specimen with negative residual strain attains a higher
implementation of the proposed model as a non-trivial extension of load, presumably due to the need for an additional force to counter
the existing implementations (Rahaman, 2022; Pillai et al., 2023) to the compressive residual stain. In contrast, the specimen with tensile
accommodate tailored residual stress. We refer to recent articles on residual strain reaches a lower maximum load whilst experiencing two
the use of Gridap in modeling brittle fracture under different types of peaks, distinguishing it from the other two scenarios. The normalized
loading (Rahaman, 2022; Behera et al., 2023b,a; Das et al., 2023). In load–displacement curves via our model and the ones reported by
any case, for the sake of completeness, we provide a brief account of Salvati (2021) are in good agreement.
the finite element implementation of the present model in Appendix B.
4.2. Comparison with an experimental crack pattern
4. Numerical simulations
In this section, we consider a four-point bending test proposed
We provide numerical simulations for four-point bending tests on by Bermejo et al. (2006b) on a nine-layered ATZ–AMZ ceramic lam-
notched ceramic composite beams made of thin layers of AMZ (Al2 O3 inate for which a schematic loading arrangement is shown in Fig. 5.
with 30% monoclinic ZrO2 ) and thick layers of ATZ (Al2 O3 with 5% We assume that the ATZ and AMZ layers are perfectly bonded with dis-
tetragonal ZrO2 ). The material properties of ATZ and AMZ are in placement continuity at the interfaces. The thickness ratio is 𝑡ATZ ∕𝑡AMZ =

5
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 2. Sub-figure (a) shows the test-set up for the tension test on a notched plate with the circular inclusion under residual strain (𝝐 res ); all dimensions in millimeters. Sub-figure
(b) demonstrates a representative finite element mesh.

Fig. 3. Damage profile for the tension test on a notched plate with the circular inclusion under hydrostatic residual strains of magnitude (a) −0.0015, (b) 0.0, and (c) 0.0015.

AMZ layers, 𝛥𝑇 is taken as -1250 ◦ C. Residual stresses are compressive


in AMZ layers and tensile in ATZ layers. We assume plane strain
conditions pertaining to a two-dimensional approximation. Of course,
during the manufacturing process, residual stresses would develop in
both X- and Z-directions. However, we hypothesize that the crack
may propagate across the entire width (in the Z-direction) and that
the residual stress in the Z-direction would not affect the fracturing
process. Nevertheless, incorporating a three-dimensional analysis might
enhance the accuracy of the results and allow for the examination of
problems with complex boundary conditions. However, we keep a more
detailed study out of the present scope.
For numerical simulation, we use a non-uniform finite element mesh
consisting of constant strain triangular (CST) elements. The phase field
length scale parameter (𝑙s ) is taken as 0.02 mm. Note that the element
size in the region where crack propagation might occur should not
be greater than half of the designated 𝑙s value. Material properties
for the ATZ and AMZ layers are in Table 1. The simulation has been
Fig. 4. Normalized load–displacement plots for the tension test with a circular inclusion
under hydrostatic residual strain of different magnitudes. Load–displacement plots from conducted as per the staggered approach given in Appendix B. We start
the proposed model are shown against results reported by Salvati (2021). with the assumption that the critical energy release rate is the same
for mode-I and mode-II fracture, i.e, 𝐺cI = 𝐺cII = 𝐺c and there is no
dissipative effect, i.e.  = 0 Sec./mm. We explore two scenarios: one
5.7, where 𝑡ATZ denotes the thickness of ATZ layers and 𝑡AMZ that of with no residual stress and another with residual stress induced due to a
AMZ layers. For the introduction of residual stresses within the ATZ and temperature change of 𝛥𝑇 = −1250 ◦ C. Notably, the crack pattern in the

6
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 5. Loading arrangement for four-point bending test on a nine-layered ATZ–AMZ ceramic laminate. All dimensions are in millimeters.

Fig. 6. Crack profiles of a nine-layered ATZ–AMZ ceramic laminate. Sub-figures (a) and (b) illustrate the damage profile without residual stress considering 𝐺cII = 𝐺cI and
𝐺cII = 10𝐺cI respectively. Sub-figures (c) and (d) illustrate the damage profile with residual stress induced for 𝛥𝑇 = −1250 ◦ C with 𝐺cII = 𝐺cI and 𝐺cII = 10𝐺cI respectively. Sub-figure
(e) displays the experimental crack profile reported by Bermejo et al. (2006b).

absence of residual stress appears straight (Fig. 6(a)). However, upon


introducing residual stress, a series of vertical cracks emerge in the AMZ
layer (Fig. 6(c)). As the applied load increases, the crack originates at
the notch tip, and as it reaches the AMZ layer, it follows the path of
one of the vertical cracks before continuing straight within the next
ATZ layer. These vertical cracks in the AMZ layer may be due to the
shear stress generated by tensile stresses in ATZ layers positioned above
and below AMZ layers exhibiting compressive residual stress. However,
such vertical cracks are absent in the experimentally reported damage
profile (Fig. 6(e)). As a result, we substantially increase the strain
energy release rate for mode-II fracture: 𝐺cII = 10𝐺cI . This adjustment
aims at preventing the formation of vertical cracks within the AMZ
layer. The simulated damage profiles for 𝐺cII = 10𝐺cI without and with
residual stress (induced for 𝛥𝑇 = −1250 ◦ C) are shown in Figs. 6 (b) and
(d) respectively. With a higher 𝐺cII , the damage profile for no residual
stress remains similar to that for the case of 𝐺cII = 𝐺cI . However,
when residual stress is present, vertical cracks in the AMZ layer are
significantly reduced with higher 𝐺cII . The obtained damage profile is
Fig. 7. Load–deflection curves for a nine-layered notched composite beam (four AMZ
layers sandwiched between five ATZ layers with the thickness ratio 𝑡ATZ ∕𝑡AMZ = 5.7) for similar to the experimentally reported crack pattern (Fig. 6(e)), where
four different cases. The first and second cases are without residual stress considering the crack initiates from the notch tip and, upon reaching the AMZ
𝐺cII = 𝐺cI , and 𝐺cII = 10𝐺cI respectively. The second and the third cases involve layer, bifurcates into two branches. Load–displacement plots for the
consideration of residual stress resulting from a temperature change of 𝛥𝑇 = −1250 ◦ C four considered cases are demonstrated in Fig. 7. Observing the graphs,
with the third case being 𝐺cII = 𝐺cI , and the fourth case being 𝐺cII = 10𝐺cI .
it is apparent that, in the absence of residual stress, the responses for

7
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 8. Crack profiles of a nine-layered ATZ–AMZ ceramic laminate. Sub-figures (a), (b), and (c) demonstrate the crack profile for the laminate with  = 0.25, 0.50, and 0.75
Sec./mm respectively.

the laminates. We take notched composite beams of constant height


𝐻 and varying thickness ratios 𝑡ATZ ∕𝑡AMZ as shown in Fig. 10. For
numerical simulations on the composite beam, we take the length 𝐿
as 10 mm, height 𝐻 as 2 mm, distance between two loading points 𝐿2
as 4.5 mm, distance between two supports 𝐿1 as 9 mm, notch height
ℎ𝑛 as 0.3 mm and notch width 𝑤𝑛 as 0.075 mm. Assuming that plane
strain conditions hold, we carry out finite element simulations for the
two-dimensional model using CST elements and a staggered scheme
described in Appendix B. We consider the regularization parameter 𝑙𝑠
(= 0.015 mm) the same for both the ATZ and AMZ layers. A non-
uniform finite element mesh is used with a finer mesh in regions where
cracks may possibly propagate (Fig. 11). We consider mainly three
different cases for tailoring the residual stresses. First, we vary the
residual stress values by changing 𝛥𝑇 for a fixed geometric configu-
ration of the notched composite beam. Due to a mismatch in thermal
Fig. 9. Load–deflection curves for a nine-layered notched composite beam (four AMZ
expansion coefficients of ATZ and AMZ ( Table 1), tensile residual
layers sandwiched between five ATZ layers with the thickness ratio 𝑡ATZ ∕𝑡AMZ = 5.7)
with different energy dissipation effect. stress is induced in the ATZ layers whereas compressive residual stress
is induced in the AMZ layers for a negative 𝛥𝑇 (see Fig. 10 for a
schematic representation). As the second case, we consider different
number of layers, viz. 3, 5 and 7 in the notched composite beam for a
fixed ratio 𝑡ATZ ∕𝑡AMZ of 12. As the third case, we vary the layer thickness
the cases with 𝐺cII = 𝐺cI and with 𝐺cII = 10𝐺cI yield similar responses. by considering different thickness ratios 𝑡ATZ ∕𝑡AMZ , viz. 4, 8, and 12.
However, when including residual stress within the laminate, opting Unless explicitly stated, dissipative effects are not taken into account
for 𝐺cII = 10𝐺cI leads to considerably higher ultimate load and tough- (i.e.  = 0.0 Sec./mm), and identical critical energy release rates for
ness vis-á-vis the choice of 𝐺cII = 𝐺cI . Hence, for such problems, a mode I and mode II (𝐺cII = 𝐺cI = 𝐺c ) are employed.
mixed-mode formulation is necessary. We allow displacement-controlled loading and plot the damage pro-
Given the possibility of additional energy dissipation mechanisms file of a three-layered composite beam made of two ATZ layers and one
apart from crack formation, we incorporate non-zero dissipative effects AMZ layer with the thickness ratio (𝑡ATZ ∕𝑡AMZ ) being 12. The damage
into the model to observe the resulting behavior. We consider a nine- profiles are generated at different stages of applied displacement for
layered laminate (Fig. 5) with residual stress induced for 𝛥𝑇 = −1250 ◦ C two different scenarios: one with no residual stress (𝛥𝑇 = 0 ◦ C) as
with 𝐺cII = 10𝐺cI and  = 0.25, 0.5, and 0.75 Sec./mm. The damage shown in Fig. 12, and another with residual stress resulting from 𝛥𝑇 =
profile and load–displacement responses for all these cases are reported −1000 ◦ C, illustrated in Fig. 13. We observe that the damage profile in
in Fig. 8 and Fig. 9 respectively. One notices that damage profiles for all the residually stressed composite beam (Fig. 13) is very different from
the considered  cases are more or less similar. In each case, the crack that in a beam under no residual stress (Fig. 12). The composite beam
originates at the notch tip and then divides into two branches upon without residual stress fails rather suddenly at an applied displacement
entering the AMZ layer, where compressive residual stress is preva- of 0.003 mm, and the crack propagation is along a straight path. In
lent. Throughout the initial six layers, the manner in which the crack contrast, for the composite beam with residual stress, the propagating
propagates remains nearly uniform across all these scenarios. However, crack bifurcates as it reaches the AMZ layer due to the compressive
within the final three layers, the trajectory of crack propagation differs residual stress, so the final failure of the specimen gets delayed.
depending on the  values. Load–displacement responses for different We investigate a seven-layered composite beam (Fig. 10(c)) for
 values also follow a similar pattern. However, with increased , the a thickness ratio of 𝑡ATZ ∕𝑡AMZ = 12 with various 𝛥𝑇 , e.g. 𝛥𝑇 =
specimen attains higher ultimate load and toughness. 0 ◦ C, −500 ◦ C, −800 ◦ C, and − 1000 ◦ C, so that changes in the response
of the specimen due to changes in the magnitude of the residual stress
4.3. Parametric study can be analyzed. Load–displacement responses for different 𝛥 𝑇 are
shown in Fig. 14. The corresponding damage profiles are illustrated
We now consider variations in the magnitude of residual stress, in Fig. 15. From the load–displacement plots, it is seen that, for
number of layers in the laminate, and thickness ratio of ATZ and AMZ higher 𝛥𝑇 , i.e., with higher residual stress, the specimen becomes
layers, to understand their effect on the mechanical performance of tougher as the area under the curve increases, and the failure of the

8
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 10. Three-point bending test set-up for notched composite beams made of alternating layers of ATZ and AMZ shown in sub-figures for (a) three, (b) five and (c) seven layered
beams. On the right within a small segment of the composite beam, outward arrows in ATZ layers indicate that ATZ layers will be under tensile residual stress and inward arrows
in AMZ layers indicate that AMZ layers will be under compressive residual stress due to the manufacturing process.

Fig. 11. Representative finite element mesh for three-layered notched composite beam (one AMZ layer is sandwiched between two ATZ layers). The mesh is finer near material
interfaces and very fine around the middle region where crack may propagate.

9
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 12. Damage profiles for a notched composite beam (one AMZ layer and two ATZ layers) under no residual stress at different deformation stages shown in sub-figures at an
applied displacement of (a) 0.003 mm, (b) 0.0033 mm, (c) 0.0035 mm and (d) 0.006 mm.

Fig. 13. Damage profile for a three-layered notched composite beam (one AMZ layer and two ATZ layers) with residual stress at different deformation stages shown in sub-figures
at an applied displacement of (a) 0.003 mm, (b) 0.0033 mm, (c) 0.0035 mm and (d) 0.006 mm. Residual stress applied to the composite beam corresponds to a temperature difference
𝛥𝑇 = −1000 ◦ C.

Fig. 14. Load–deflection curves for a seven-layered notched composite beam (three AMZ layers sandwiched between four ATZ layers with the thickness ratio 𝑡ATZ ∕𝑡AMZ = 12) with
induced residual stresses pertaining to 𝛥𝑇 values of 0 ◦ C, −500 ◦ C, −800 ◦ C and −1000 ◦ C under four-point loading arrangement.

10
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 15. Comparison of crack patterns in a seven-layered notched composite beam (three AMZ layers sandwiched between four ATZ layers with the thickness ratio 𝑡ATZ ∕𝑡AMZ = 12)
for different residual stresses induced with 𝛥T values of (a) 0 ◦ C (no residual stress) (b) −500 ◦ C (c) −800 ◦ C and (d) −1000 ◦ C.

Fig. 16. Load–deflection curve for a notched composite beam consisting of three, five, and seven layers (alternating layers of ATZ and AMZ with the thickness ratio (𝑡ATZ ∕𝑡AMZ = 12))
when the beam is under no residual stress, i.e. 𝛥𝑇 = 0 ◦ C and with thermal residual stress induced for 𝛥𝑇 = −1000 ◦ C.

specimen is delayed. Such responses are consistent with the damage load–displacement curves (Fig. 16) that, for all the cases i.e., for the
profile, where the bifurcation of the crack in the AMZ layer, exhibiting 3-, 5- and 7-layered beams, there is an increment in the ultimate load
compressive residual stress, is more pronounced with higher residual as well as toughness due to the specimen being residually stressed. The
stress magnitude (higher 𝛥𝑇 ) against a straight crack propagation for corresponding damage profiles are shown in Fig. 17. Without residual
the case of no residual stress. However, for very high 𝛥𝑇 , the residual stress, the crack propagation is straight; however, damage profiles in
stress may be so high that the specimen could fail under the action of the composite beams with residual stress tend to bifurcate as they
residual stress alone. Hence, for designing such a composite section, an reach the AMZ layer exhibiting compressive residual stress. Composite
optimum 𝛥𝑇 must be arrived at. beams with residual stress have a longer overall crack length, resulting
In the second case, we consider notched beams with three, five, in a higher ultimate load and tougher load–displacement response.
and seven layers (see Fig. 10 for a schematic representation) under no In this particular case, the three-layered beam outperforms five- or
residual stress (𝛥𝑇 = 0 ◦ C) and under residual stress resulting from 𝛥𝑇 seven-layered beams, regardless of whether residual stress is present or
of −1000 ◦ C. By this, we wish to explore if we can change the nature of not. However, it should not be concluded that three-layered composite
the failure of ceramic composites from brittle to relatively ductile by beams are always superior to those with more layers. The optimal
introducing thermally induced residual stresses. We observe from the number of layers in a composite may depend on various factors such as

11
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 17. Crack patterns in a notched composite beam consisting of three, five, and seven layers when the beam is under no residual stress i.e. 𝛥𝑇 = 0 ◦ C (sub-figures (a) 3 layers,
(c) 5 layers and (e) 7 layers) and under thermal residual stress induced for 𝛥𝑇 = −1000 ◦ C (sub-figures (b) 3 layers, (d) 5 layers and (f) 7 layers) at an applied displacement of
0.006 mm. For all the cases, 𝑡ATZ ∕𝑡AMZ = 12.

Fig. 18. Load–deflection curve for a seven-layered notched composite beam (three AMZ layers sandwiched between four ATZ layers) with varying layer thickness ratios 𝑡ATZ ∕𝑡AMZ
of 4, 8 and 12 under no thermal residual stress i.e. 𝛥𝑇 = 0 ◦ C, and under thermal residual stress induced with 𝛥𝑇 = −1000 ◦ C.

specimen dimensions, constituent materials of the composite, loading The specimen with thickness ratio 𝑡ATZ ∕𝑡AMZ = 4 attains maximum ulti-
rate, magnitude of the residual stress, etc. Hence, the number of layers mate load and toughness for the considered loading set-up, and as the
in a composite must be optimized for specific scenarios to achieve best thickness ratio increases, both the ultimate load and toughness reduce.
performance. From the damage profiles, one observes that with increasing 𝑡ATZ ∕𝑡AMZ
Further, we explore the possibility of an optimal thickness ratio ratio, the bifurcation of cracks in the AMZ layer is less prominent.
between the ATZ and the AMZ layers by analyzing a notched composite This might be the reason that 𝑡ATZ ∕𝑡AMZ = 4 attains maximum load
beam under two conditions: one with no residual stress (i.e., 𝛥𝑇 = compared to other cases. These findings strongly suggest the existence
0 ◦ C), and the other with thermal residual stress corresponding to of an optimal thickness ratio that may enable superior mechanical
𝛥𝑇 = −1000 ◦ C. The thickness ratios 𝑡ATZ ∕𝑡AMZ investigated are 4, 8, and performance. Note that the optimal thickness ratio for a seven-layered
12 under same loading rate. The simulated load–displacement responses composite should not be assumed as 4 based solely on this example. The
and damage profiles are shown in Figs. 18 and 19 respectively. The ideal thickness ratio may vary depending on other factors, geometric
load–displacement responses reveal that all the thickness ratios produce or loading-related. Therefore, when designing composite beams for a
similar responses when there is no residual stress. However, when given scenario, one must optimize the thickness ratio in conjunction
residual stresses are present, responses exhibit noteworthy differences. with other parameters for that specific case.

12
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. 19. Comparison of crack patterns in a seven-layered composite beam for different thickness ratios when the composite beam is under no residual stress, i.e. 𝛥𝑇 = 0 ◦ C
(sub-figures (a) 𝑡ATZ ∕𝑡AMZ = 4, (c) 𝑡ATZ ∕𝑡AMZ = 8, and (e) 𝑡ATZ ∕𝑡AMZ = 12) and under thermal residual stress induced with 𝛥𝑇 = −1000 ◦ C (sub-figures (b) 𝑡ATZ ∕𝑡AMZ = 4, (d)
𝑡ATZ ∕𝑡AMZ = 8, and (f) 𝑡ATZ ∕𝑡AMZ = 12).

5. Concluding remarks downloaded as jupyter notebook files from Julia codes for a phase-field
fracture modeling of ceramic laminates under residual stress.
We have developed a residual stress-dependent mixed-mode phase-
field model that can also account for strictly dissipative effects. The Funding information
effectiveness of the model is verified by comparing the simulations
with existing findings in the literature. We have demonstrated the The corresponding author, Mohammad Masiur Rahaman acknowl-
working of the proposed approach in reproducing experimentally ob- edges the support from DST-SERB under the grant SRG/2021/000254.
served crack patterns in ceramic laminates under residual stress fields.
We have also shown how the response of these laminates alters with
CRediT authorship contribution statement
additional dissipative effects. Utilizing the proposed model, we have
explicated on possible enhancements in mechanical performance of
ceramic laminates through tailored residual stresses. Numerical sim- Akash Kumar Behera: Conceptualization, Methodology, Formal
ulations of four-point bending tests on a notched ceramic composite analysis, Investigation, Writing – original draft, Writing – review & edit-
beam with a fixed number of layers provide insightful information ing, Visualization. Mohammad Masiur Rahaman: Conceptualization,
on how crack paths change as the thermally induced residual stress Methodology, Resources, Writing – original draft, Writing – review &
increases, leading to superior mechanical performance with higher editing, Supervision, Funding acquisition. Debasish Roy: Conceptual-
fracture toughness and load-carrying capacity. While optimizing the ization, Methodology, Writing – review & editing, Supervision.
number of layers in the laminate might yield elevated mechanical
performance, tailoring the residual stress (by changing the thickness Declaration of competing interest
ratio of layers of two different materials) For a fixed number of layers,
might also achieve the same goal. This suggests a possible optimization The authors declare that they have no known competing finan-
of the thickness ratio to maximize mechanical performance. Thus the cial interests or personal relationships that could have appeared to
methodology, in principle, might be used as a numerical tool to design influence the work reported in this paper.
high-performance ceramic laminates.
While the phase-field scheme that forms the backbone of this work Data availability
is popular for simulating fracture, it is not without its share of inade-
quacies and even inconsistencies. Leaving aside the issue of a physically No data was used for the research described in the article.
meaningful or geometric interpretation of the phase-field variable, the
use of a history function for the microforce balance or the energy split
Appendix A. Computation of thermally induced residual stress
to accommodate tension–compression asymmetry is also questionable.
Indeed, the very use of a diffused form of the crack surface area in the
phase field is susceptible to criticism as Griffith’s law may not hold in In order to study the effect of residual stresses on the mechanical
many quasi-brittle regimes. A kinematically transparent, more gener- performance of ceramic laminates, we have assumed that the residual
ally applicable, and thermodynamically informed phase field approach stress can be computed a-priori based on a somewhat simplistic analy-
is certainly needed so that simulations can be better trusted in the sis (Kotoul et al., 2012; Ševeček et al., 2013) and is given as known
absence of elaborate experimental evidence. input to the model. In the present study, we have considered only
thermal residual stresses induced due to a mismatch in the thermal
Data accessibility expansion coefficients of the layer materials during the manufacturing
process. The thermally induced residual stress is computed based on
Source codes for the implementation of the proposed phase-field the compatibility of strain between the layers. When the material body
fracture model for ceramic laminates under residual stress can be is under only thermal loading and all the layers in the laminate are

13
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

perfectly bonded together, one may require that the total thermal strain Similarly, the test functions and their derivatives can be defined as
in all the layers be the same, so that
(𝒖̃ ℎ )𝑒 = 𝖭𝑒𝑢 𝗎̃ 𝑒 , (𝑠̃ℎ )𝑒 = 𝖭𝑒 𝗌̃𝑒 , (B.4)
res res
𝛼ATZ 𝛥𝑇 + 𝜖ATZ = 𝛼AMZ 𝛥𝑇 + 𝜖AMZ , (A.1)
where, 𝜖AMZres res represent the residual strains, and 𝛼
and 𝜖ATZ AMZ and (̃𝝐 ℎVoigt )𝑒 = 𝖡𝑒𝑢 𝗎̃ 𝑒 , ∇(𝑠̃ℎ )𝑒 = 𝖡𝑒𝑠 𝗌̃𝑒 , (B.5)
𝛼ATZ are the thermal coefficients of AMZ and ATZ layers, respectively.
Assuming equal bi-axial stresses along 𝑥 and 𝑧 directions (𝜎𝑥𝑥res = 𝜎 res = where, 𝗎̃ 𝑒 and 𝗌̃𝑒 are the test functions at the 𝑒th element. We can
𝑧𝑧
𝜎 res ) and zero stress along 𝑦 direction (𝜎𝑦𝑦res = 0), the stress strain define the domain for each element by an open set 𝛺𝑒 ⊂ 𝛺 such that,
𝑁𝑒
𝐸 ∪𝑒=1 𝛺𝑒 = 𝛺, where 𝑁𝑒 is the total number of elements. For the domain
relation is given as 𝜎 res = 𝐸 ′ 𝜖 res where 𝐸 ′ = 1−𝜈 with 𝐸 denoting
the modulus of elasticity and 𝜈 the Poisson ratio of the material. Note 𝛺𝑒 , we can write the discretized form of Eq. (47) as
that the relation is valid for both ATZ and AMZ layers with respective
(𝐵𝑢𝑒 )𝑇 𝖣 𝖡𝑒𝑢 𝗎𝑒 𝑑𝛺𝑒 = (𝖡𝑒𝑢 )𝑇 𝝈 res
Voigt
𝑑𝛺𝑒 , (B.6)
material properties. Considering 𝑛 ATZ layers having individual layer ∫𝛺𝑒 ∫𝛺𝑒
thickness of 𝑡ATZ and (𝑛 − 1) AMZ layers having individual layer thick-
where 𝖣 is the reduced matrix representation of fourth order tensors
ness of 𝑡AMZ , a constant width along the 𝑧 direction and 𝜎 res = 𝐸 ′ 𝜖 res ,
Cmod . We express Cmod as
the force equilibrium equation for the laminate under thermal stress
can be expressed as (A.2), Cmod = 𝑠2 (𝜆𝑰 ⊗ 𝑰 + 𝝁I), (B.7)
res ′ res ′
𝑛𝑡ATZ 𝜖ATZ 𝐸ATZ + (𝑛 − 1)𝑡AMZ 𝜖AMZ 𝐸AMZ = 0. (A.2) where, 𝑰 is second-order identity tensor and I is fourth-order identity
Employing Eqs. (A.1) and (A.2), thermal residual strain in ATZ and tensor. Now Eq. (B.6) can be further reduced to the form
AMZ layers can be evaluated as
𝖪𝑒𝑢 𝗎𝑒 = 𝖿𝑢𝑒 , (B.8)

𝐸AMZ (𝑛 − 1)𝑡AMZ
res
𝜖ATZ = (𝛼AMZ − 𝛼ATZ )𝛥𝑇 ′ ′
, (A.3) where
𝐸ATZ 𝑛𝑡ATZ + 𝐸AMZ (𝑛 − 1)𝑡AMZ

res
𝐸ATZ 𝑛𝑡ATZ 𝖪𝑒𝑢 = (𝖡𝑒𝑢 )𝑇 𝖣 𝖡𝑒𝑢 𝑑𝛺𝑒 , (B.9)
𝜖AMZ = −(𝛼AMZ − 𝛼ATZ )𝛥𝑇 ′ . (A.4) ∫𝛺𝑒

𝐸ATZ 𝑛𝑡ATZ + 𝐸AMZ (𝑛 − 1)𝑡AMZ
res = 𝜖 res = 𝜖 res and for every
and
In every ATZ layer we may consider 𝜖𝑥𝑥 𝑧𝑧 ATZ
res = 𝜖 res = 𝜖 res . One can find 𝜖 res by satisfying the
AMZ layer 𝜖𝑥𝑥 𝑧𝑧 AMZ 𝑦𝑦 𝖿𝑢𝑒 = (𝖡𝑒𝑢 )𝑇 𝝈 res
Voigt
𝑑𝛺𝑒 . (B.10)
res = 0. Hence, following Hooke’s law, and setting 𝜎 res = 0, ∫𝛺𝑒
constraint 𝜎𝑦𝑦 𝑦𝑦
res for both ATZ and AMZ layers can be derived as
the expression of 𝜖𝑦𝑦 In Eqs. (B.9) and (B.10), 𝖪𝑒𝑢 and 𝖿𝑢𝑒 represent the element stiffness matrix
−2𝜈𝜖 res and the element load vector, respectively for the equation Eq. (B.6).
res
𝜖𝑦𝑦 = . (A.5) Appropriately mapping the local element degrees of freedom to the
1−𝜈
With known residual thermal strain components, the residual stress corresponding global degrees of freedom (see Reddy, 2010), one can
tensor 𝝈 res can be computed using the fourth order elasticity tensor as assemble the element matrices and determine the global ones. For
C𝝐 res . instance, introducing the global stiffness matrix 𝖪𝑢 and the global force
vector 𝖿𝑢 , we can write
Appendix B. Finite element implementation of the proposed 𝑁𝑒 𝑁𝑒
∑ ∑
model 𝖪𝑢 (𝐼, 𝐽 ) = 𝖪𝑒 (𝑖, 𝑗), and 𝖿𝑢 (𝐼) = 𝖿𝑢𝑒 (𝑖), (B.11)
𝑒=1 𝑒=1
In this section, we provide a brief description of the implementation
where, 𝐼 and 𝐽 are the global degrees of freedom corresponding to their
of the finite element formulation discussed in Section 3. We have im-
local counterparts 𝑖 and 𝑗, respectively. After assembling and imposing
plemented the finite element formulation using two-dimensional plane
the boundary conditions, we can get the global form 𝖪𝑢 𝗎 = 𝖿𝑢 , the
strain approximation. In the planar two-dimensional case, we express
solution of which gives the global displacement vector 𝗎. Similarly,
the stress tensor 𝝈 in terms of a vector by using Voigt notation 𝝈 Voigt
[ ]𝑇 Eq. (48) for the 𝑒th element can be expressed as
where 𝝈 Voigt = 𝜎𝑥𝑥 𝜎𝑦𝑦 𝜎𝑥𝑦 with (⋅)𝑇 denoting the transpose of the
argument. Similarly, the strain tensor 𝝐 is expressed by using its cor- ( 𝑒𝑇 𝑒 1  )
[ ]𝑇 (𝖡𝑠 ) 𝑙𝑠 𝖡𝑠 + (𝖭𝑒 )𝑇 2(( ℎ )𝑒 )𝖭𝑒 + (𝖭𝑒 )𝑇 𝖭𝑒 + (𝖭𝑒 )𝑇 𝖭𝑒 𝗌𝑒 𝑑𝛺𝑒 =
responding vector 𝝐 Voigt = 𝜖𝑥𝑥 𝜖𝑦𝑦 𝜖𝑥𝑦 . For the discretization, we ∫𝛺𝑒 𝑙𝑠 𝛥𝑡
have used three-noded[ triangular element, for which
] the interpolation (1 𝑒 𝑇  )
function 𝖭𝑒 (𝑥, 𝑦) = 𝑁1𝑒 (𝑥, 𝑦) 𝑁2𝑒 (𝑥, 𝑦) 𝑁3𝑒 (𝑥, 𝑦) , where 𝑁1𝑒 , 𝑁2𝑒 , and (𝖭 ) + (𝖭𝑒 )𝑇 𝖭𝑒 𝗌𝑒𝑛−1 𝑑𝛺𝑒 ,
∫𝛺𝑒 𝑙𝑠 𝛥𝑡
𝑁3𝑒 are the shape functions associated with the three nodes of the 𝑒th
element. For the 𝑒th element, the primary variables 𝒖ℎ (𝑥, 𝑦) and 𝑠ℎ (𝑥, 𝑦) (B.12)
can be expressed as
where ( ℎ )𝑒
is a function of 𝝐 elas
which can be expressed as 𝝐 elas =
ℎ 𝑒 res res
(𝒖 ) (𝑥, 𝑦) = 𝖭𝑒𝑢 (𝑥, 𝑦) 𝗎𝑒 , ℎ 𝑒 𝑒 𝑒
and (𝑠 ) (𝑥, 𝑦) = 𝖭 (𝑥, 𝑦) 𝗌 , (B.1) 𝑒 𝑒
𝖡𝑢 𝗎 − 𝝐 Voigt with 𝝐 Voigt denoting the vector corresponding to the
residual strain tensor 𝝐 res in Voigt notation, 𝛥𝑡 is the pseudo time
where
[ ] increment, and 𝗌𝑒𝑛−1 is the nodal phase field vector for the previous
𝖭𝑒 (𝑥, 𝑦) 𝟢 loading step. Now, the reduced from of Eq. (B.12) can be expressed
𝖭𝑒𝑢 (𝑥, 𝑦) = , (B.2)
𝟢 𝖭𝑒 (𝑥, 𝑦) as
[ ]𝑇
𝗎𝑒 = 𝗎𝑒𝑥 𝗎𝑒𝑦 is the nodal displacement vector and 𝗌𝑒 the nodal phase 𝖪𝑒𝑠 𝗌𝑒 = 𝖿𝑠𝑒 , (B.13)
field vector at the 𝑒th element. Using Eq. (B.1), we write the finite
element approximation for the element strain vector (𝝐 ℎVoigt )𝑒 (𝑥, 𝑦) and where
∇𝑠𝑒 (𝑥, 𝑦) as 𝖡𝑒𝒖 (𝑥, 𝑦) 𝗎𝑒 and 𝖡𝑒𝒔 (𝑥, 𝑦) 𝗌𝑒 , respectively, where 𝖡𝑒𝒖 (𝑥, 𝑦) and ( 𝑒𝑇 1  )
𝖡𝑒𝒔 (𝑥, 𝑦) are defined by 𝖪𝑒𝑠 = (𝖡𝑠 ) 𝑙𝑠 𝖡𝑒𝑠 + (𝖭𝑒 )𝑇 ( ℎ )𝖭𝑒 + (𝖭𝑒 )𝑇 𝖭𝑒 + (𝖭𝑒 )𝑇 𝖭𝑒 𝑑𝛺𝑒
∫𝛺𝑒 𝑙𝑠 𝛥𝑡
⎡ 𝖭𝑒 [ ] (1 𝑒 𝑇 )
𝟢 ⎤ 𝖿𝑠𝑒 =

(𝖭 ) + (𝖭𝑒 )𝑇 𝖭𝑒 𝗌𝑒𝑛−1 𝑑𝛺𝑒
⎢ ,𝑥 ⎥ 𝖭𝑒,𝑥 ∫𝛺𝑒 𝑙𝑠 𝛥𝑡
𝖡𝑒𝑢 (𝑥, 𝑦) =⎢ 𝟢 𝑒 𝑒
𝖭,𝑦 ⎥ , 𝖡𝑠 (𝑥, 𝑦) = . (B.3)
⎢ 𝖭𝑒 𝖭𝑒,𝑦
⎣ ,𝑦 𝖭𝑒,𝑥 ⎥⎦ (B.14)

14
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Fig. B.20. Implementation of the phase-field model with residual stress using a staggered scheme.

In Eq. (B.14), 𝖪𝑒𝑠 and 𝖿𝑠𝑒 represent element stiffness matrix and element Aranson, I., Kalatsky, V., Vinokur, V., 2000. Continuum field description of crack
load vector, respectively, for Eq. (B.12). Upon assembly, we get propagation. Phys. Rev. Lett. 85 (1), 118.
Arias, I., Serebrinsky, S., Ortiz, M., 2006. A phenomenological cohesive model of
𝑁𝑒 𝑁𝑒
∑ ∑ ferroelectric fatigue. Acta Mater. 54 (4), 975–984.
𝖪𝑠 (𝐼, 𝐽 ) = 𝖪𝑒𝑠 (𝑖, 𝑗), and 𝖿𝑠 (𝐼) = 𝖿𝑠𝑒 (𝑖), (B.15) Badia, S., Verdugo, F., 2020. Gridap: An extensible finite element toolbox in julia. J.
𝑒=1 𝑒=1 Open Source Softw. 5 (52), 2520.
where, 𝖪𝑠 and 𝖥𝑠 represent global stiffness matrix and global load Becher, P.F., Sun, E.Y., Plucknett, K.P., Alexander, K.B., Hsueh, C.-H., Lin, H.-T., Wa-
vector respectively. Using 𝖪𝑠 and 𝖥𝑠 , global phase field vector 𝗌 can ters, S.B., Westmoreland, C.G., Kang, E.-S., Hirao, K., et al., 1998. Microstructural
design of silicon nitride with improved fracture toughness: I, effects of grain shape
be obtained.
and size. J. Am. Ceram. Soc. 81 (11), 2821–2830.
A robust, staggered scheme is adopted for the numerical solution. Behera, A.K., Sudeep, K.H., Rahaman, M.M., 2023a. Thermodynamically consistent
We have the following steps to update the solution from pseudo-time volumetric–deviatoric decomposition-based phase-field model for thermo-electro-
𝑡𝑛 to 𝑡𝑛+1 (Miehe et al., 2010a). mechanical fracture. Eng. Fract. Mech. 290, 109468.
Behera, A.K., Unnikrishna Pillai, A., Rahaman, M.M., 2023b. A phase-field model for
1. Let the global nodal displacement vector, nodal phase-field vec- electro-mechanical fracture with an open-source implementation of it using Gridap
tor and the history function be known at 𝑡𝑛 and given by 𝗎𝑛 , 𝗌𝑛 in Julia. Math. Mech. Solids 28 (8), 1877–1908.
and 𝑛 , respectively. Bermejo, R., Danzer, R., 2010. High failure resistance layered ceramics using crack
bifurcation and interface delamination as reinforcement mechanisms. Eng. Fract.
2. Using 𝗎𝑛 , 𝗌𝑛 and 𝑛 in Eq. (B.14), solve Eq. (B.13) and update Mech. 77 (11), 2126–2135.
the nodal phase-field vector at 𝑡𝑛+1 as 𝗌𝑛+1 . Bermejo, R., Torres, Y., Sanchez-Herencia, A., Baudín, C., Anglada, M., Llanes, L.,
3. Use 𝗎𝑛 , 𝑛 and 𝗌𝑛+1 for the evaluation of 𝖣 in Eq. (B.9) and solve 2006a. Fracture behaviour of an Al2O3–ZrO2 multi-layered ceramic with residual
Eq. (B.8) to update the nodal displacement vector at 𝑡𝑛+1 as 𝗎𝑛+1 . stresses due to phase transformations. Fatigue Fract. Eng. Mater. Struct. 29 (1),
71–78.
4. Once 𝗎𝑛+1 and 𝗌𝑛+1 are determined, update the energy history
Bermejo, R., Torres, Y., Sanchez-Herencia, A., Baudín, C., Anglada, M., Llanes, L.,
function at 𝑡𝑛+1 by defining it as 2006b. Residual stresses, strength and toughness of laminates with different layer
{ thickness ratios. Acta Mater. 54 (18), 4745–4757.
 ℎ for  ℎ > 𝑛
𝑛+1 = (B.16) Blattner, A.J., Lakshminarayanan, R., Shetty, D.K., 2001. Toughening of layered ceramic
𝑛 otherwise. composites with residual surface compression: effects of layer thickness. Eng. Fract.
Mech. 68 (1), 1–7.
We provide a flow chart, shown in Fig. B.20, which depicts the steps Borden, M.J., Hughes, T.J., Landis, C.M., Verhoosel, C.V., 2014. A higher-order phase-
followed in implementing the model using a staggered scheme. field model for brittle fracture: Formulation and analysis within the isogeometric
analysis framework. Comput. Methods Appl. Mech. Engrg. 273, 100–118.
References Bourdin, B., Francfort, G.A., Marigo, J.-J., 2000. Numerical experiments in revisited
brittle fracture. J. Mech. Phys. Solids 48 (4), 797–826.
Agrawal, V., 2016. Multiscale phase-field model for phase transformation and fracture Bourdin, B., Francfort, G.A., Marigo, J.-J., 2008. The variational approach to fracture.
(Ph.D. thesis). Carnegie Mellon University. J. Elasticity 91 (1–3), 5–148.
Ambati, M., Gerasimov, T., De Lorenzis, L., 2015. A review on phase-field models of Bucevac, D., Boskovic, S., Matovic, B., Krstic, V., 2010. Toughening of SiC matrix with
brittle fracture and a new fast hybrid formulation. Comput. Mech. 55 (2), 383–405. in-situ created TiB2 particles. Ceram. Int. 36 (7), 2181–2188.
Amor, H., Marigo, J.-J., Maurini, C., 2009. Regularized formulation of the variational Coleman, B.D., Noll, W., 1974. The thermodynamics of elastic materials with heat
brittle fracture with unilateral contact: Numerical experiments. J. Mech. Phys. conduction and viscosity. In: The Foundations of Mechanics and Thermodynamics.
Solids 57 (8), 1209–1229. Springer, pp. 145–156.

15
A.K. Behera et al. Mechanics of Materials 188 (2024) 104867

Das, S., Pillai, A.U., Chemenghat, A., Rahaman, M.M., 2023. Determining the role Miehe, C., Welschinger, F., Hofacker, M., 2010b. Thermodynamically consistent phase-
of microstructural topology on the mechanical performance of nacre-inspired field models of fracture: Variational principles and multi-field FE implementations.
composites using a phase-field model. Mater. Today Commun. 37, 107453. Internat. J. Numer. Methods Engrg. 83 (10), 1273–1311.
de Portu, G., Micele, L., 2007. Lamination process to obtain structure with tailored Orlovskaya, N.A., Kuebler, J., Subotin, V.I., Lugovy, M., 2001. High toughness ceramic
residual stress distribution. In: Key Engineering Materials, Vol. 333. Trans Tech laminates by design of residual stresses. MRS Online Proc. Lib. 702 (1), 1–6.
Publ, pp. 27–38. Orlovskaya, N., Lugovy, M., Subbotin, V., Radchenko, O., Adams, J., Chheda, M.,
Dhas, B., Rahaman, M., Akella, K., Roy, D., Reddy, J., et al., 2018. A phase-field damage Shih, J., Sankar, J., Yarmolenko, S., 2005. Robust design and manufacturing of
model for orthotropic materials and delamination in composites. J. Appl. Mech. 85 ceramic laminates with controlled thermal residual stresses for enhanced toughness.
(1). J. Mater. Sci. 40 (20), 5483–5490.
Eastgate, L., Sethna, J., Rauscher, M., Cretegny, T., Chen, C.-S., Myers, C., 2002. Pak, Y., 1992. Linear electro-elastic fracture mechanics of piezoelectric materials. Int.
Fracture in mode I using a conserved phase-field model. Phys. Rev. E 65 (3), J. Fract. 54 (1), 79–100.
036117. Pillai, A.U., Behera, A.K., Rahaman, M.M., 2023. Combined diffused material interface
Francfort, G.A., Marigo, J.-J., 1998. Revisiting brittle fracture as an energy minimization and hybrid phase-field model for brittle fracture in heterogeneous composites. Eng.
problem. J. Mech. Phys. Solids 46 (8), 1319–1342. Fract. Mech. 277, 108957.
Gao, H., Zhang, T.-Y., Tong, P., 1997. Local and global energy release rates for an Prakash, V., Behera, A.K., Rahaman, M.M., 2023. A phase-field model for
electrically yielded crack in a piezoelectric ceramic. J. Mech. Phys. Solids 45 (4), thermo-mechanical fracture. Math. Mech. Solids 28 (2), 533–561.
491–510. Qin, S., Jiang, D., Zhang, J., Qin, J., 2003. Design, fabrication and properties of layered
Gerasimov, T., De Lorenzis, L., 2019. On penalization in variational phase-field models SiC/TiC ceramic with graded thermal residual stress. J. Eur. Ceram. Soc. 23 (9),
of brittle fracture. Comput. Methods Appl. Mech. Engrg. 354, 990–1026. 1491–1497.
Gogotsi, G., Lugovoi, N., Slyunyaev, V., 2004. Fracture resistance of residually-stressed Rahaman, M.M., 2022. An open-source implementation of a phase-field model for brittle
ceramic laminated structures. Strength Mater. 36 (3), 291–303. fracture using Gridap in Julia. Math. Mech. Solids 27 (11), 2404–2427.
Grigoriev, O., Stepanenko, A., Vinokurov, V., Neshpor, I., Mosina, T., Silvestroni, L., Reddy, J., 2010. An Introduction to the Finite Element Method, Vol. 1221. McGraw-Hill
2021. ZrB2–SiC ceramics: Residual stresses and mechanical properties. J. Eur. New York.
Ceram. Soc. 41 (9), 4720–4727. Salvati, E., 2021. Residual stress as a fracture toughening mechanism: A Phase-Field
Gurtin, M.E., 1996. Generalized Ginzburg-Landau and Cahn-Hilliard equations based study on a brittle material. Theor. Appl. Fract. Mech. 114, 103021.
on a microforce balance. Physica D 92 (3–4), 178–192. Sánchez-Herencia, A.J., Pascual, C., He, J., Lange, F.F., 1999. ZrO2/ZrO2 layered
Hakim, V., Karma, A., 2009. Laws of crack motion and phase-field models of fracture. composites for crack bifurcation. J. Am. Ceram. Soc. 82 (6), 1512–1518.
J. Mech. Phys. Solids 57 (2), 342–368. Ševeček, O., Bermejo, R., Kotoul, M., 2013. Prediction of the crack bifurcation in
Henry, H., Levine, H., 2004. Dynamic instabilities of fracture under biaxial strain using layered ceramics with high residual stresses. Eng. Fract. Mech. 108, 120–138.
a phase field model. Phys. Rev. Lett. 93 (10), 105504. Sglavo, V.M., Bertoldi, M., 2006. Design and production of ceramic laminates with high
Hua, Y., Zhang, L., Cheng, L., Li, Z., Du, J., 2010. Microstructure and high temperature mechanical reliability. Composites B 37 (6), 481–489.
strength of SiCW/SiC composites by chemical vapor infiltration. Mater. Sci. Eng. A Sglavo, V.M., Paternoster, M., Bertoldi, M., 2005. Tailored residual stresses in
527 (21–22), 5592–5595. high reliability alumina-mullite ceramic laminates. J. Am. Ceram. Soc. 88 (10),
Karma, A., Kessler, D.A., Levine, H., 2001. Phase-field model of mode III dynamic 2826–2832.
fracture. Phys. Rev. Lett. 87 (4), 045501. Spatschek, R., Brener, E., Karma, A., 2011. Phase field modeling of crack propagation.
Kotoul, M., Sevecek, O., Vyslouzil, T., 2012. Crack growth in ceramic laminates with Phil. Mag. 91 (1), 75–95.
strong interfaces and large compressive residual stresses. Theor. Appl. Fract. Mech. Spetz, A., Denzer, R., Tudisco, E., Dahlblom, O., 2021. A modified phase-field fracture
61, 40–50. model for simulation of mixed mode brittle fractures and compressive cracks in
Kuhn, C., Müller, R., 2008. A phase field model for fracture. In: PAMM: Proceedings porous rock. Rock Mech. Rock Eng. 54, 5375–5388.
in Applied Mathematics and Mechanics, Vol. 8, No. 1. Wiley Online Library, pp. Suo, Z., Kuo, C.-M., Barnett, D., Willis, J., 1992. Fracture mechanics for piezoelectric
10223–10224. ceramics. J. Mech. Phys. Solids 40 (4), 739–765.
Kuhn, C., Müller, R., 2010. A continuum phase field model for fracture. Eng. Fract. Verdugo, F., Badia, S., 2022. The software design of Gridap: a finite element package
Mech. 77 (18), 3625–3634. based on the Julia JIT compiler. Comput. Phys. Comm. 276, 108341.
Linder, C., Rosato, D., Miehe, C., 2011. New finite elements with embedded strong Verhoosel, C.V., Scott, M.A., De Borst, R., Hughes, T.J., 2011. An isogeometric approach
discontinuities for the modeling of failure in electromechanical coupled solids. to cohesive zone modeling. Internat. J. Numer. Methods Engrg. 87 (1–5), 336–360.
Comput. Methods Appl. Mech. Engrg. 200 (1–4), 141–161. Wei, C., Liu, Z., Wu, Y., Liu, Y., Zhang, H., Wang, P., Sun, Q., Zhou, L., 2021. Toughness
Liu, Y., Cheng, C., Ziaei-Rad, V., Shen, Y., 2021. A micromechanics-informed phase field and R-curve behaviour of laminated Si3N4/SiCw ceramics. Ceram. Int. 47 (13),
model for brittle fracture accounting for unilateral constraint. Eng. Fract. Mech. 18693–18698.
241, 107358. Yang, W., Fuso, L., Biamino, S., Vasquez, D., Bolivar, C.V., Fino, P., Badini, C., 2012.
Liu, S., Li, Y., Chen, P., Li, W., Gao, S., Zhang, B., Ye, F., 2016. Residual stresses and Fabrication of short carbon fibre reinforced SiC multilayer composites by tape
mechanical properties of. casting. Ceram. Int. 38 (2), 1011–1018.
Liu, S., Li, Y., Chen, P., Li, W., Gao, S., Zhang, B., Ye, F., 2017. Residual stresses Zhang, X., Sloan, S.W., Vignes, C., Sheng, D., 2017. A modification of the phase-field
and mechanical properties of Si3N4/SiC multilayered composites with different SiC model for mixed mode crack propagation in rock-like materials. Comput. Methods
layers. Boletin de la Sociedad Espanola de Ceramica Y Vidrio 56 (4), 147–154. Appl. Mech. Engrg. 322, 123–136.
Lugovy, M., Slyunyayev, V., Subbotin, V., Orlovskaya, N., Gogotsi, G., 2004. Crack Zhang, T.-Y., Zhao, M., Tong, P., 2002. Fracture of piezoelectric ceramics. Adv. Appl.
arrest in Si3N4-based layered composites with residual stress. Compos. Sci. Technol. Math. 38, 147–289.
64 (13–14), 1947–1957. Zobeiry, N., Poursartip, A., 2015. The origins of residual stress and its evalua-
McMeeking, R.M., 2001. Towards a fracture mechanics for brittle piezoelectric and tion in composite materials. In: Structural Integrity and Durability of Advanced
dielectric materials. Int. J. Fract. 108 (1), 25–41. Composites. Elsevier, pp. 43–72.
Miehe, C., Hofacker, M., Welschinger, F., 2010a. A phase field model for rate-
independent crack propagation: Robust algorithmic implementation based on
operator splits. Comput. Methods Appl. Mech. Engrg. 199 (45–48), 2765–2778.

16

You might also like