Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Engineering Fracture Mechanics 264 (2022) 108305

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Thermoelastic fracture analysis of functionally graded materials


using the scaled boundary finite element method
M.D. Iqbal a , C. Birk a ,∗, E.T. Ooi b , A.L.N. Pramod c , S. Natarajan d , H. Gravenkamp e ,
C. Song f
a
Institute of Structural Analysis of Plates and Shells, University of Duisburg–Essen, 45141 Essen, Germany
b
School of Engineering, Information Technology and Physical Sciences, Federation University Australia, Ballarat VIC 3350, Australia
c
Department of Mechanical Engineering, Indian Institute of Technology Bombay, Mumbai 400076, India
d Department of Mechanical Engineering, Indian Institute of Technology Madras, Chennai 600036, India
e International Centre for Numerical Methods in Engineering, 08034 Barcelona, Spain
f School of Civil and Environmental Engineering, University of New South Wales, Sydney NSW 2052, Australia

ARTICLE INFO ABSTRACT

Keywords: The scaled boundary finite element method is extended to model fracture in functionally graded
Thermoelasticity materials (FGM) under coupled thermo-mechanical loads. The governing equations of coupled
Stress intensity factors thermo-mechanical equilibrium are discretized using scaled boundary shape functions enriched
T -stress
with the thermal load terms. The material gradient is modeled as a series of power functions,
Functionally graded materials (FGM)
and the stiffness matrix is calculated semi-analytically. Stress intensity factors and 𝑇 −stress
SBFEM
are directly calculated from their definition without any need for additional post-processing
techniques. Arbitrary-sided polygon elements are employed for flexible mesh generation. Several
numerical examples for isotropic and orthotropic FGMs are presented to validate the proposed
technique.

1. Introduction

Functionally graded materials (FGMs) are a class of engineered materials that are custom designed by gradually varying the
volume fractions of its constituents (usually a ceramic-rich phase in combination with a metallic-rich phase) within the solid.
The resulting composite exhibits both continuously and smoothly varying material parameters that can be tailored to suit specific
applications, e.g. as thermal barrier coatings, wear-resistant coatings and bio-medical materials [1,2]. This distinguishing feature
of FGMs usually results in superior thermal and wear-resistant qualities when proper design conditions are met, compared with
the more common laminated composite materials. For the latter, abrupt changes in the material properties are usually observed at
the interface between the different material phases, which often lead to high stress concentrations that compromise the structural
integrity of the structure. Consequently, FGMs are usually utilized under conditions of extreme thermo-mechanical loading. The
investigation of fracture in FGMs subjected to thermo-mechanical loads is thus crucial for the identification of reliability and
durability of such materials.
Early studies on the thermo-mechanical fracture of FGMs were performed analytically, often involving simple geometries,
material profiles and boundary conditions, e.g. embedded or surface cracks in unconstrained elastic layers [3], circumferential cracks
in a thin-walled cylinder under quasi-static thermal load [4], edge crack in a FGM strip with varying thermal properties subjected

∗ Correspondence to: Institute of Structural Analysis of Plates and Shells, University of Duisburg–Essen, Germany.
E-mail addresses: muhammad.iqbal@uni-due.de (M.D. Iqbal), carolin.birk@uni-due.de (C. Birk), e.ooi@federation.edu.au (E.T. Ooi), snatarajan@iitm.ac.in
(S. Natarajan), hgravenkamp@cimne.upc.edu (H. Gravenkamp), c.song@unsw.edu.au (C. Song).

https://doi.org/10.1016/j.engfracmech.2022.108305
Received 19 September 2021; Received in revised form 9 January 2022; Accepted 2 February 2022
Available online 12 February 2022
0013-7944/© 2022 Elsevier Ltd. All rights reserved.
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

to transient thermal loading [5] and embedded crack in FGM plates and cylindrical shells [6]. While the closed-form solutions of
the asymptotic stress distribution and stress intensity factors obtained from analytical techniques are valuable, it is a challenging
task to extend such techniques to arbitrary geometries, material profiles and loading conditions. A more pragmatic approach lies in
the use of numerical techniques.
Numerical simulations that involve deformable bodies are usually performed using the finite element method (FEM). For
applications to FGMs, graded finite elements [7] which necessitate a slight modification in the calculation of the finite element
stiffness matrix were proposed. Successful applications of graded finite elements in thermal stress analyses have been reported
e.g. in [8–10]. When applied to modeling of fracture with or without consideration of thermal loads, the graded finite elements,
like the conventional finite elements, suffer from their inability to accurately model the asymptotic stress field in the vicinity of
stress singularities as indicated by the size of the meshes used in the simulations in [10]. This is because, like the conventional
FEM, the graded FEM employs polynomial functions to represent the unknown fields. This modeling aspect can be improved by
employing special-purpose elements such as quarter-point elements [11], specially designed finite elements incorporating analytical
base functions [12] or enriching the elements in the vicinity of the crack tip within the context of the extended finite element method
(XFEM) [13–16]. When employing the XFEM, the choice of enrichment functions need to be derived from the near tip temperature
distribution or an equivalent asymptotic analysis, particularly when considering orthotropic or anisotropic FGMs.
Other alternative techniques that have been employed to model fracture in FGMs under the combined thermo-mechanical loading
include the boundary element method (BEM) [17], meshless methods [18,19] and numerical manifold method (NMM) [20]. The
BEM offers some appealing advantages over the FEM in terms of mesh generation and smaller problem size by reducing the spatial
dimension of the problem. However, its extension to FGMs is still very restricted due to the fact that the required fundamental
solutions are either unavailable or involve very complex mathematical analysis. Meshless methods, as the name suggests, do not
require a formal definition of a mesh, however, special requirements need to be adhered to in the representation of cracks, integration
of bilinear/linear form and imposition of boundary conditions. Further, to capture the singularity, the meshfree approximation
is augmented with enrichment functions similar to the XFEM and a path-independent integral is adopted to compute the stress
intensity factors. The numerical manifold method (NMM) [20] adopts a dual cover, viz., a mathematical and a physical cover that
facilitates the modeling of fracture-related problems. An accurate representation of the singularity of a stress field at the crack
tip is realized through the introduction of enrichment terms into the approximation similar to the XFEM. Several investigations
of thermoelastic fracture in FGMs using the NMM have been reported in literature including the calculation of stress intensity
factors [21,22], transient analysis [23] and calculation of 𝑇 −stress [24]. Similar to other approaches, the calculation of the stress
intensity factors using NMM is usually carried out with the aid of path-independent integrals.
The scaled boundary finite element method (SBFEM) [25] is a semi-analytical technique that is based entirely on finite elements
but with discretization restricted only to boundaries of the (sub)domains. In two dimensions, for example, only the boundaries of
(sub)domains are discretized using one-dimensional line elements. An analytical solution is then sought within the domain and
expressed in terms of power functions of the radial coordinate 𝜉 with its origin at a point (within the domain) known as the scaling
center. Of particular interest is the case when the (sub)domain boundary does not form a closed loop. This results in a singularity
at the scaling center that is automatically captured by the SBFEM, enabling it to model such class of problems efficiently [26].
This advantage further results in the development of efficient techniques to model crack initiation [27] and crack propagation,
e.g. [28–31].
The extension of the SBFEM to model fracture in FGMs was presented by Chiong et al. [32]. In order to capture the material
gradient across the domain, the SBFEM was expressed first in the form of scaled boundary shape functions. The material gradient
is then expressed in the form of a power function in terms of the radial coordinate 𝜉, so that the integration can be carried out
semi-analytically. Unlike homogeneous materials where only few large (sub)domains are required in a fracture simulation, FGMs
necessitate the use of a sufficiently fine mesh to accurately represent the material inhomogeneity. Both polygon meshes [33] and
quadtree meshes [34] can be adopted for this purpose. The SBFEM admits both polygon and quadtree discretizations without
requiring any change in the construction of its shape functions. Note that through-thickness power-law heterogeneities can also
be included in the semi-analytical treatment of functionally graded plates by the SBFEM [35].
In this study, the SBFEM will be further extended to model thermoelastic fracture in FGMs. The spatial discretization is based
on the scaled boundary shape functions [36], and the material gradient is represented as a power function in the radial coordinate,
such that the resulting stiffness matrix can be evaluated semi-analytically. The proposed methodology retains all the salient features
of the SBFEM including accurate modeling of asymptotic stress fields in the vicinity of singularities and the ability to be formulated
over arbitrary polytopes. Further, it does not require a priori knowledge of enrichment functions as in the Partition of Unity Methods
and special post-processing techniques to extract the stress intensity factors.
The paper is organized as follows: Section 2 contains a brief summary of the considered problem and the proposed solution
strategy. Section 3 describes the discretization of elasticity and heat conduction problems by SBFEM. Furthermore, it covers the
construction of SBFEM polygon shape functions for thermoelasticity. In Section 4, the thermoelastic fracture modeling of FGMs
based on SBFEM is discussed in detail. Section 5 presents the semi-analytical calculations of stress intensity factors and 𝑇 −stress
using SBFEM. Constitutive relations for coupled thermoelasticity are explained in Section 6. Finally, Section 7 illustrates the proposed
procedure through several numerical examples in the context of thermoelastic fracture in FGMs.

2
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

2. Problem statement and solution strategy

In the following, we consider uni-directionally coupled, steady-state thermoelasticity in a two-dimensional domain 𝛺. The
corresponding governing equations are expressed as

∇T𝐿 (𝜿∇𝐿 𝛩) + 𝑝 = 0, (1a)


∇T𝑢 𝝈 + 𝐛 = 𝟎, (1b)
( )
𝝈 = 𝐃 ∇𝑢 𝐮 − 𝜷𝛩 . (1c)

In Eq. (1), the symbols 𝜿, 𝛩, 𝑝, 𝝈 and 𝐛 denote the thermal conductivity matrix, temperature change, heat source, components of the
stress tensor in Voigt notation and body loads, respectively. 𝐃 and 𝐮 are the elasticity matrix and displacement vector, respectively
and 𝜷 is computed using elastic parameters and components of the symmetric second-order tensor of thermal expansion coefficients.
For isotropic material behavior, it is defined as
{
[𝛼 𝛼 0]T for plane stress,
𝜷= (2)
(1 + 𝜈)[𝛼 𝛼 0]T for plane strain,

with coefficient of thermal expansion 𝛼 and Poisson’s ratio 𝜈. For orthotropic material behavior, 𝜷 is expressed as
{
[𝛼 𝛼2 0]T for plane stress,
𝜷= [ 1 ]T (3)
𝜈31 𝛼3 + 𝛼1 𝜈32 𝛼3 + 𝛼2 0 for plane strain,

with Poisson’s ratios 𝜈𝑖𝑗 and coefficients of thermal expansion 𝛼𝑖 with respect to principal axes 𝑖, 𝑗, see also Section 6. Considering
any prescribed heat flux 𝑞 and tractions 𝐩𝑡 on boundaries 𝛤 of 𝛺 and using e.g. the method of weighted residuals, the corresponding
weak form of Eq. (1) is

∇𝐿 𝛿𝛩𝜿∇𝐿 𝛩 d𝛺 = 𝛿𝛩𝑝 d𝛺 + 𝛿𝛩𝑞 d𝛤 , (4a)


∫𝛺 ∫𝛺 ∫𝛤

(∇𝑢 𝛿𝐮)T 𝐃∇𝑢 𝐮 d𝛺 − (∇𝑢 𝛿𝐮)T 𝐃𝜷𝛩 d𝛺 = 𝛿𝐮T 𝐛 d𝛺 + 𝛿𝐮T 𝐩𝑡 d𝛤 . (4b)


∫𝛺 ∫𝛺 ∫𝛺 ∫𝛤
In Eq. (4), the 𝛿𝛩 and 𝛿𝐮 are the temperature and displacement weight functions, respectively. In a classical finite element
framework, both the unknown temperature change and displacement field and the corresponding weight functions are approximated
by interpolating nodal values using two-dimensional shape functions. These are typically based on polynomial functions and defined
on standard element shapes such as triangles or quads. In the following, we propose to use shape functions that have been constructed
using the SBFEM — so called scaled boundary shape functions. Such shape functions can be constructed on star-convex polygons
with an arbitrary number of edges, including open polygons to represent stress singularities accurately. The scaled boundary shape
functions for the temperature change are obtained by solving Laplace equation, i.e. Eq. (1a) with 𝑝 = 0, on a homogeneous domain
using SBFEM. The scaled boundary shape functions for the displacement field are obtained from the semi-analytical solution of
Eqs. (1b) and (1c) for constant material parameters. The construction of these shape functions is outlined in Section 3. They will
then be used as two-dimensional shape functions when discretizing Eq. (4), similar to standard FEM. However, the integration and
thus also the evaluation of the functional grading will be performed semi-analytically. This is explained in detail in Section 4.

3. Summary of the scaled boundary finite element method

This section briefly outlines the scaled boundary finite element formulation focusing on the derivation of the shape functions to
be applied in a thermoelastic analysis. The fundamental equations, i.e. the scaled boundary transformation and the scaled boundary
finite element equations in temperature and displacement, respectively, are summarized in Section 3.1. For detailed derivations, the
reader is referred to [37–39]. The scaled boundary shape functions for thermoelasticity are recapped in Section 3.2.

3.1. Scaled boundary transformation and resulting ordinary differential equations for linear elasticity and heat conduction

In 2D, the scaled boundary finite element method can be formulated on star-convex polygons with an arbitrary number of sides.
The scaled boundary coordinate system (𝜉, 𝜂) is shown in Fig. 1 for a 6-sided polygon. The scaled boundary transformation that
relates the coordinates (𝜉, 𝜂) to the Cartesian coordinates (𝑥, 𝑦) is expressed as

𝑥(𝜉, 𝜂) = 𝑥0 + 𝜉𝐍(𝜂)𝐱𝑏 = 𝑥0 + 𝜉𝑥𝜂 (𝜂), (5a)


𝑦(𝜉, 𝜂) = 𝑦0 + 𝜉𝐍(𝜂)𝐲𝑏 = 𝑦0 + 𝜉𝑦𝜂 (𝜂). (5b)

Here, the scaled boundary coordinates 𝜉 and 𝜂 are defined pointing outwards from the scaling center 𝑂 and in the circumferential
direction, respectively. The symbols 𝐱𝑏 , 𝐲𝑏 denote the Cartesian coordinates of nodes located on the line element. The Cartesian
coordinates on the boundary are denoted as 𝑥𝜂 , 𝑦𝜂 and obtained by interpolating the nodal coordinates 𝐱𝑏 , 𝐲𝑏 using one-dimensional
shape functions 𝐍(𝜂). The scaling center 𝑂 is located at (𝑥0 , 𝑦0 ).

3
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 1. Polygonal domain modeled by SBFEM.

Using the scaled boundary finite element transformation (5), the differential operators ∇𝐿 and ∇𝑢 in Eq. (1) can be expressed in
terms of the scaled boundary coordinates as
𝜕 𝜕
∇𝐿 = 𝐛1 (𝜂) + 𝜉 −1 𝐛2 (𝜂) , (6a)
𝜕𝜉 𝜕𝜂
𝜕 𝜕
∇𝑢 = 𝐛1 (𝜂) + 𝜉 −1 𝐛2 (𝜂) . (6b)
𝜕𝜉 𝜕𝜂

Detailed expressions of 𝐛𝑖 and 𝐛𝑖 (for 𝑖 = 1, 2) are given in Appendix A. The temperature change 𝛩 and displacement field 𝐮 are
expressed in a semi-discrete form as

𝛩(𝜉, 𝜂) = 𝐍(𝜂)𝜣 ℎ (𝜉), (7a)


𝐮(𝜉, 𝜂) = 𝐍(𝜂)𝐮ℎ (𝜉), (7b)

with 𝐍(𝜂) = [ 𝑁̄ 1 ⋅ 𝐈 𝑁̄ 2 ⋅ 𝐈 ⋯ 𝑁̄ 𝑙 ⋅ 𝐈 ] where 𝐈 is a 2 × 2 identity matrix and 𝑙 denotes the number of nodes on a line
element. Note that the same shape functions 𝑁̄ 𝑖 are used in the circumferential interpolation of the geometry, the temperature and
displacement field. Here and in the following, the overbar indicates association with the temperature field. Substituting Eqs. (6a),
(7a) and (6b), (7b) into Eqs. (1a) and (1b)–(1c), respectively, and applying the method of weighted residuals with respect to 𝜂 yields
the scaled boundary finite element equations in temperature change (8a) and displacements (8b), respectively [40].
( T)
𝐄0 𝜉 2 𝜣 ℎ (𝜉),𝜉𝜉 + 𝐄0 − 𝐄1 + 𝐄1 𝜉𝜣 ℎ (𝜉),𝜉 −𝐄2 𝜣 ℎ (𝜉) + 𝐟(𝜉) = 𝟎. (8a)
( T
)
𝐄0 𝜉 2 𝐮ℎ (𝜉),𝜉𝜉 + 𝐄0 − 𝐄1 + 𝐄1 𝜉𝐮ℎ (𝜉),𝜉 − 𝐄2 𝐮ℎ (𝜉) + 𝐟(𝜉) = 𝟎. (8b)

The coefficient matrices 𝐄𝑖 , 𝐄𝑖 (𝑖 = 0, 1, 2) depend on the geometry, material, and the discretization of the boundary only and
can be computed using standard finite element technology. Detailed definitions are given in Appendix B for completeness. The
inhomogeneous terms 𝐟(𝜉) and 𝐟(𝜉) are defined as

𝐟(𝜉) = 𝜉𝐅𝑞 + 𝜉 2 𝐅𝑝 , (9a)


𝑡 2 𝑏
𝐟(𝜉) = 𝜉𝐩0 + 𝜉𝐅 + 𝜉 𝐅 . (9b)

In Eq. (9a), the terms 𝐅𝑞 and 𝐅𝑝 correspond to prescribed normal flux amplitudes on the boundary segments passing through the
scaling center and to internal heat source amplitudes, respectively. Since only the homogeneous part of Eq. (8a) is used to derive
the polygon shape functions for the temperature field, these terms will not be considered further in the construction of the scaled
boundary shape functions for the temperature change. Similarly, the terms 𝐅𝑡 and 𝐅𝑏 in Eq. (9b) corresponding to prescribed tractions
on boundary segments passing through the scaling center and internal body forces, respectively, will be neglected in the following.
The term 𝐩0 corresponds to the contribution of thermal stress 𝝈 0 = −𝐃𝜷𝛩 (c.f. Eq. (1c)). It is taken into account in the derivation
of displacement shape functions for thermal stress analysis and expressed as
( T )
𝐩0 (𝜉) = 𝐁1 (𝜉𝝈 0 ),𝜉 −𝐁T2 𝝈 0 𝐽 d𝜂, (10)
∫𝜂

where 𝐽 is the Jacobian determinant. Detailed definitions of the Jacobian matrix 𝐉 and of the matrices 𝐁𝑖 are given in Appendices A
and B, respectively, for completeness.

4
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

3.2. Scaled boundary shape functions for thermoelasticity

As explained in Ref. [40], the polygon shape functions for the temperature field are constructed using the homogeneous solution
of Eq. (8a). The latter requires the evaluation of the following eigenvalue decomposition:
[ ] [ ][ ]
𝐔(𝑢)
𝑛 𝐔(𝑢)
𝑝 𝐔(𝑢)
𝑛 𝐔(𝑢)
𝑝 𝐒𝑛
𝐙 = , (11)
𝐔(𝑞)
𝑛 𝐔(𝑞)
𝑝 𝐔(𝑞)
𝑛 𝐔(𝑞)
𝑝 𝐒𝑝

with
[ −1 T −1 ]
𝐄0 𝐄1 −𝐄0
𝐙= −1 T −1 . (12)
−𝐄2 + 𝐄1 𝐄0 𝐄1 −𝐄1 𝐄0

Here, the eigenvalues of the Hamiltonian matrix 𝐙 are sorted in ascending order, where the diagonal blocks 𝐒𝑛 and 𝐒𝑝 correspond
to the eigenvalues with negative and positive real parts, respectively. Using the eigenvalues 𝐒𝑛 and the eigenvectors 𝐔(𝑢)
𝑛 , the
temperature change is expressed as

𝛩(𝜉, 𝜂) = 𝜱(𝜉, 𝜂)𝜣 0 . (13)

In Eq. (13), 𝜣 0 and 𝜱(𝜉, 𝜂) denote the nodal values of temperature change and the corresponding polygon shape functions,
respectively, with
( )(𝑒) −𝐒 ( (𝑢) )−1
𝜱(𝜂, 𝜉) = 𝐍(𝜂) 𝐔(𝑢)
𝑛 𝜉 𝑛 𝐔𝑛 . (14)

Note that Eq. (14) is obtained by assembling the contributions from the line elements that form the polygon. That is, only the rows
in 𝐔(𝑢)
𝑛 corresponding to a line element (𝑒) are required.
The displacement field is expressed by interpolating the parameters 𝐮0 as

𝐮(𝜉, 𝜂) = 𝜱(𝜉, 𝜂)𝐮0 , (15)

with
[ ] [ ]T
𝜱(𝜉, 𝜂) = 𝜱(ℎ) (𝜉, 𝜂) 𝜱(𝑏) (𝜉, 𝜂) , 𝐮0 = 𝐮ℎ (𝜉 = 1) 𝐜̄ . (16)

The parameters 𝐮0 contain both the desired nodal displacements 𝐮ℎ (𝜉 = 1) and additional variables 𝐜̄ corresponding to the thermal
load. The size of 𝐮0 is 3𝑚, where 𝑚 is the number of nodes in a polygon element. Contrary to the shape functions for the temperature,
the polygon shape functions for displacements contain two components. The block 𝜱(ℎ) (𝜉, 𝜂) corresponds to the homogeneous
solution of Eq. (8b). It is obtained by constructing a block-diagonal Schur decomposition of 𝐙 [41],
̂
𝐙𝐕 = 𝐕𝐒, (17)

with
( [ ] )
𝐒̂ 𝑁 𝐈
𝐒̂ = diag 𝐒̂ 1 , … , 𝐒̂ 𝑁−1 , , ̂ 𝑁+2 , … , 𝐒̂ 2𝑁 ,
𝐒 (18)
0 𝐒̂ 𝑁+1
[ (𝑢) ] [ (𝑢) (𝑢)
]
𝐕 𝐕𝑛 𝐕𝑝
𝐕= = , (19)
𝐕(𝑞) 𝐕(𝑞)
𝑛 𝐕(𝑞)
𝑝
[ ]
𝐄−10
𝐄T1 −𝐄−1
0
𝐙= −1 T T −1 , (20)
−𝐄2 + 𝐄1 𝐄0 𝐄1 𝐄1 − 𝐄1 𝐄0
as
( )(𝑒) −𝐒̂ ( (𝑢) )−1
𝜱(ℎ) (𝜉, 𝜂) = 𝐍(𝜂) 𝐕(𝑢)
𝑛 𝜉 𝑛 𝐕𝑛 . (21)

In Eq. (18), 𝑁 is the number of diagonal blocks in 𝐒. ̂ In the event where a complete block diagonalization is performed, 𝑁 = 2𝑚 is
the number of degrees of freedom in a polygon with number of nodes 𝑚.
On the other hand, the term 𝜱(𝑏) (𝜉, 𝜂) is a supplementary shape function, which corresponds to the particular solution of Eq. (8b)
due to the thermal stress. It is expressed as [40]
(( )(𝑒) 𝐒 ( )(𝑒) −𝐒̂ ( (𝑢) )−1 ( (𝑢) )(𝑒) )
𝜱(𝑏) (𝜉, 𝜂) = 𝐍(𝜂) 𝐕(𝑢) 𝐀𝜉 𝑏 − 𝐕(𝑢)
𝑛 𝜉 𝑛 𝐕𝑛 𝐕 𝐀 , (22)

with
( (1) (2) (𝑚)
)
𝐒𝑏 = diag −𝑆 𝑛 + 1, −𝑆 𝑛 + 1, … , −𝑆 𝑛 + 1 , (23)
[ ]
𝐀 = 𝐀(1) 𝐀(2) ⋯ 𝐀(𝑚) , (24)
(𝑗)
𝐀(𝑗) = −(𝐒̂ 𝑖 + (−𝑆 𝑛 + 1)𝐈)−1 𝐆(𝑗) . (25)

5
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

The matrix 𝐆 is computed for each diagonal block 𝑖 in 𝐒̂ prior to assembling the contributions of each block into a polygon as
(𝑢) T
𝐆𝑖 = −𝐇−T
𝑖 (𝐕𝚤 ) 𝐐0 (26)
where the index 𝚤 indicates the block conjugate,

𝚤 = 2𝑁 + 1 − 𝑖. (27)
The matrix 𝐇𝑖 is defined as
{
𝐇𝑖 when 𝑗 = 𝚤
𝐕T𝑖 𝐉2𝑛 𝐕𝑗 = −𝐕T𝑗 𝐉2𝑛 𝐕T𝑖 = , (28)
𝟎 when 𝑗 ≠ 𝚤
where the matrix 𝐉2𝑛 is
[ ]
𝟎 𝐈
𝐉2𝑛 = , (29)
−𝐈 𝟎
and 𝐈 is an identity matrix. The matrix 𝐐0 is computed for each line element on the polygon boundary as
1[ ]
𝐐0 = −𝐁T1 (𝜂)𝐃𝜷𝐍(𝜂)(𝐔(𝑢)
𝑛 )(𝑒) (−𝐒𝑛 + 𝐈)𝐁T2 (𝜂)𝐃𝜷𝐍(𝜂)(𝐔(𝑢)
𝑛 )(𝑒) 𝐽 (𝜂) d𝜂. (30)
∫−1
For detailed derivations, we refer to Ref. [40].

4. Thermoelastic fracture modeling in functionally graded materials

In the following, the scaled boundary shape functions (14) and (16) will be used to interpolate the nodal temperatures and
displacements in a thermo-mechanical fracture analysis of functionally graded materials (c.f. Eqs. (13) and (15)). The weak form
of the governing equations of uni-directionally coupled steady-state thermoelasticity is described in Eq. (4). The temperature and
displacement weight functions 𝛿𝛩 and 𝛿𝐮, respectively, are interpolated using the same shape functions as for the temperature
change and displacement field, i.e. (14) and (16), respectively. The gradients of the temperature and displacement fields (along
with their respective weight functions) are obtained using Eqs. (6a) and (6b)

∇𝐿 𝛩 = 𝐁(𝜉, 𝜂)𝜣 0 , (31a)


∇𝐿 𝛿𝛩 = 𝐁(𝜉, 𝜂)𝛿𝜣 0 , (31b)
∇𝑢 𝐮 = 𝐁(𝜉, 𝜂)𝐮0 , (31c)
∇𝑢 𝛿𝐮 = 𝐁(𝜉, 𝜂)𝛿𝐮0 , (31d)

with
𝐁(𝜉, 𝜂) = 𝜳 𝑞 (𝜂)𝜉 −𝐒𝑛 −𝐈 (𝐔(𝑢) −1
𝑛 ) , (32a)
𝐒−𝐈
𝐁(𝜉, 𝜂) = 𝜳 𝜖 (𝜂)𝜉 𝐓. (32b)

The terms 𝜳 𝑞 (𝜂) and 𝜳 𝜖 (𝜂) are referred to as the heat flux modes and strain modes [40] and expressed as

𝜳 𝑞 (𝜂) = −𝐁1 (𝜂)(𝐔(𝑢) (𝑒) (𝑢) (𝑒)


𝑛 ) 𝐒𝑛 + 𝐁2 (𝜂)(𝐔𝑛 ) , (33a)
𝜳 𝜖 (𝜂) = 𝐁1 (𝜂)𝜳 (𝑒) 𝐒 + 𝐁2 (𝜂)𝜳 (𝑒) , (33b)
where the matrices 𝐒, 𝐓, and 𝜳 are defined as
[ ]
−𝐒̂ 𝑛 𝟎
𝐒= , (33c)
𝟎 𝐒𝑏
[ ( )−1 ( )−1 ]
𝐕(𝑢)
𝑛 − 𝐕(𝑢)
𝑛 𝐕(𝑢) 𝐀
𝐓= , (33d)
𝟎 𝐈
[ ]
𝜳 = 𝐕(𝑢)
𝑛 𝐕(𝑢) 𝐀 . (33e)
Substituting the interpolations of the temperature change Eq. (13) and displacements Eq. (15) with their respective weight functions
as well as Eqs. (31) into Eqs. (4a) and (4b) yields
T 𝜿 T 𝑝 T 𝑞
𝐁 𝐁 d𝛺 𝜣 0 = 𝜱 d𝛺 + 𝜱 d𝛤 , (34)
∫𝛺 𝑇0 ∫𝛺 𝑇0 ∫𝛤 𝑇0

𝐁T 𝐃𝐁 d𝛺 𝐮0 − 𝐁T 𝐃𝜷𝜱 d𝛺 𝜣 0 = 𝜱T 𝐛 d𝛺 + 𝜱T 𝐩𝑡 d𝛤 . (35)
∫𝛺 ∫𝛺 ∫𝛺 ∫𝛤

Eqs. (34) and (35) can be cast in a matrix form,

𝐊𝛩 𝜣 0 = 𝐅𝛩 , (36a)

6
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

𝐊𝑢 𝐮0 = 𝐊𝑐 𝜣 0 + 𝐅𝑢 , (36b)

where

𝐊𝑢 = 𝐁T 𝐃𝐁 d𝛺, (37a)
∫𝛺
T
𝐊𝛩 = 𝐁 𝜿𝐁 d𝛺, (37b)
∫𝛺

𝐊𝑐 = 𝐁T 𝐃𝜷𝜱 d𝛺, (37c)


∫𝛺

𝐅𝑢 = 𝜱T 𝐛 d𝛺 + 𝜱T 𝐩𝑡 d𝛤 , (37d)
∫𝛺 ∫𝛤
T T
𝐅𝛩 = 𝜱 𝑝 d𝛺 + 𝜱 𝑞 d𝛤 . (37e)
∫𝛺 ∫𝛤
In a functionally graded medium, the material properties vary as functions of the spatial coordinates (𝑥, 𝑦). In order to model this
material variation, it is assumed that the constitutive matrices can be represented locally in each polygon by a polynomial surface
in Cartesian coordinates (𝑥, 𝑦),

𝐃(𝑥, 𝑦) = 𝐃0 + 𝐃1 𝑥 + 𝐃2 𝑦 + 𝐃3 𝑥2 + 𝐃4 𝑥𝑦 + 𝐃5 𝑦2 + ⋯ , (38a)
𝜿(𝑥, 𝑦) = 𝜿 0 + 𝜿 1 𝑥 + 𝜿 2 𝑦 + 𝜿 3 𝑥2 + 𝜿 4 𝑥𝑦 + 𝜿 5 𝑦2 + ⋯ , (38b)
𝐃𝑐 (𝑥, 𝑦) = 𝐃𝑐0 + 𝐃𝑐1 𝑥 + 𝐃𝑐2 𝑦 + 𝐃𝑐3 𝑥2 + 𝐃𝑐4 𝑥𝑦 + 𝐃𝑐5 𝑦2 + ⋯ , (38c)

where the material properties 𝐃, 𝜿 and 𝐃𝑐 are interpolated values of elasticity matrix, conductivity and 𝐃𝜷. Eq. (38) can be
transformed to scaled boundary coordinates (𝜉, 𝜂),

𝑛
𝐃(𝜉, 𝜂) = 𝐃(0) (𝜂)𝜉 0 + 𝐃(1) (𝜂)𝜉 1 + 𝐃(2) (𝜂)𝜉 2 + ⋯ = 𝐃(𝑟) (𝜂)𝜉 𝑟 , (39a)
𝑟=0

𝑛
𝜿(𝜉, 𝜂) = 𝜿 (0) (𝜂)𝜉 0 + 𝜿 (1) (𝜂)𝜉 1 + 𝜿 (2) (𝜂)𝜉 2 + ⋯ = 𝜿 (𝑟) (𝜂)𝜉 𝑟 , (39b)
𝑟=0
∑𝑛
𝐃𝑐 (𝜉, 𝜂) = 𝐃(0) 0 (1) 1 (2) 2
𝑐 (𝜂)𝜉 + 𝐃𝑐 (𝜂)𝜉 + 𝐃𝑐 (𝜂)𝜉 + ⋯ = 𝐃(𝑟) 𝑟
𝑐 (𝜂)𝜉 . (39c)
𝑟=0

The coefficient matrices 𝐊𝑢 , 𝐊𝛩 and 𝐊𝑐 in Eq. (36) are functions of the scaled boundary coordinates 𝜉 and 𝜂. Expressing the
differentials d𝛺 and d𝛤 in scaled boundary coordinates and substituting the scaled boundary shape functions and their derivatives
in Eq. (37)(a)–(c) results in
( T −𝐈
( (∑
𝑛
) ) )
𝐊𝑢 = 𝐓T 𝜉𝐒 𝜳 T𝜖 𝐃(𝑟) (𝜂) 𝜉 𝑟 𝜳 𝜖 𝐽 d𝜂 𝜉 𝐒 d𝜉 𝐓, (40a)
∫𝜉 ∫𝜂
𝑟=0
( T ( (∑
𝑛
) ) )
𝐊𝛩 = (𝐔(𝑢)
𝑛 )
−T
𝜉 −𝐒𝑛 −𝐈 𝜳 T𝑞 𝜿 (𝑟) (𝜂) 𝜉 𝑟 𝜳 𝑞 𝐽 d𝜂 𝜉 −𝐒𝑛 d𝜉 (𝐔(𝑢) −1
𝑛 ) , (40b)
∫𝜉 ∫𝜂
𝑟=0
( T
( (∑
𝑛
) ) )
𝐊𝑐 = 𝐓T 𝜉𝐒 𝜳 T𝜖 𝐃(𝑟) 𝑟 (𝑢)
𝑐 (𝜂) 𝜉 𝐍𝐽 d𝜂𝐔𝑛 𝜉 −𝐒𝑛 d𝜉 (𝐔(𝑢) −1
𝑛 ) . (40c)
∫𝜉 ∫𝜂
𝑟=0

For further use, the matrices 𝐘(𝑟)


𝑘𝑢
, 𝐘(𝑟)
𝑘𝛩
and 𝐘(𝑟)
𝑘𝑐
are introduced for each term of the material interpolation 𝑟,

(𝑟)
𝐘𝑘𝑢 = 𝜳 T𝜖 𝐃(𝑟) (𝜂)𝜳 𝜖 𝐽 d𝜂, (41a)
∫𝜂
(𝑟)
𝐘𝑘𝛩 = 𝜳 T𝑞 𝜿 (𝑟) 𝜳 𝑞 𝐽 d𝜂, (41b)
∫𝜂
(𝑟)
𝐘𝑘𝑐 = 𝜳 T𝜖 𝐃(𝑟) (𝑢)
𝑐 𝐍𝐽 d𝜂𝐔𝑛 . (41c)
∫𝜂

Note that the matrices defined in Eq. (41) depend on the cirumferential coordinate 𝜂 only. Moreover, we define the matrices 𝐗(𝑟)
𝑘𝑢
,
𝐗(𝑟)
𝑘𝛩
(𝑟)
and 𝐗𝑘𝑐 for each term 𝑟,

(𝑟) T +(𝑟−1)𝐈
𝐗𝑘𝑢 = 𝜉𝐒 𝐘(𝑟) 𝜉 𝐒 d𝜉, (42a)
∫𝜉 𝑘𝑢
T
(𝑟)
𝐗𝑘𝛩 = 𝜉 −𝐒𝑛 +(𝑟−1)𝐈 𝐘(𝑟) 𝜉 −𝐒𝑛 d𝜉, (42b)
∫𝜉 𝑘𝛩

(𝑟) T +𝑟𝐈 (𝑟) −𝐒𝑛


𝐗𝑘𝑐 = 𝜉𝐒 𝐘𝑘𝑐 𝜉 d𝜉. (42c)
∫𝜉

7
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 2. Scaled boundary polygon element containing a singularity.

Using the properties of matrix power functions and integrating by parts, it can be verified that the integral equations (42) result in
the Sylvester equations:

𝐗(𝑟)
𝑘𝑢
𝐒 + (𝐒T + 𝑟𝐈)𝐗(𝑟)
𝑘𝑢
= 𝐘(𝑟)
𝑘𝑢
, (43a)
T
𝐗(𝑟)
𝑘𝛩
(−𝐒𝑛 ) + (−𝐒𝑛 + 𝑟𝐈)𝐗(𝑟)
𝑘𝛩
= 𝐘(𝑟)
𝑘𝛩
, (43b)
𝐗(𝑟)
𝑘𝑐
(−𝐒𝑛 + 𝐈) + (𝐒T + 𝑟𝐈)𝐗(𝑟)
𝑘𝑐
= (𝑟)
𝐘𝑘𝑐 . (43c)

Substituting Eq. (42) into Eq. (40) and using (41), the coefficient matrices 𝐊𝑢 , 𝐊𝛩 , 𝐊𝑐 can be calculated by summing up the
contributions of all coefficient matrices 𝐗(𝑟)
𝑖𝑗 ,

(∑
𝑛 )
𝐊𝑢 = 𝐓T 𝐗(𝑟)
𝑘𝑢
𝐓, (44a)
𝑟=0
(∑
𝑛 )
𝐊𝛩 = (𝐔(𝑢)
𝑛 )
−T
𝐗(𝑟)
𝑘𝛩
(𝐔(𝑢) −1
𝑛 ) , (44b)
𝑟=0
(∑
𝑛 )
𝐊𝑐 = 𝐓T 𝐗(𝑟)
𝑘𝑐
(𝐔(𝑢) −1
𝑛 ) . (44c)
𝑟=0

Thus, the nodal temperature change 𝜣 0 and displacement 𝐮0 are calculated by successively solving the algebraic Eqs. (36).

5. Calculation of stress intensity factors and 𝑻 −stress

This section presents the calculation of the stress intensity factors and 𝑇 −stress in the context of the SBFEM. Fig. 2 shows a scaled
boundary polygon element with a re-entrant corner which does not form a closed loop of line elements. Here, the scaling center is
placed at the point of singularity to semi-analytically capture the singular solution characteristics in the radial direction. The local
coordinate system 𝑥̃ − 𝑦̃ is introduced, where 𝑥̃ is parallel to a boundary intersecting the scaling center/point of singularity. For a
crack-like singularity, this corresponds to a crack face. Note that the SBFEM is equally applicable to various types of singularities
corresponding, e.g. to crack-like or wedge-like discontinuities as well as to material interfaces. Following the procedure outlined
in [40], the stress field in the scaled boundary polygon element 𝝈(𝜉, 𝜂) is a power function of the coordinate 𝜉 and is defined as

𝝈(𝜉, 𝜂) = 𝐃𝜳 (𝑒)
𝜖 (𝜂)𝜉
𝐒−𝐈
𝐓𝐮0 − 𝐃𝜷𝐍(𝜂)(𝐔(𝑢) (𝑒) −𝐒𝑛
𝑛 ) 𝜉 (𝐔(𝑢) −1
𝑛 ) 𝜣0. (45)

The SIFs and 𝑇 −stress are commonly used parameters in the characterization of fracture phenomena. In the special case of a crack
in an isotropic homogeneous medium, the asymptotic stress field near the crack tip can be expressed as a series of power functions
with real exponents [42]. The first term of the series leads to the stress singularity at the crack tip, commonly known as square root
singularity. The second term corresponds to the constant stress parallel to the crack surface known as 𝑇 −stress. One of the main
features of SBFEM is that the SIFs and 𝑇 −stress can be calculated directly from their definitions without requiring the knowledge
of higher order terms in the series [43].
In the more general case of an open polygon with scaling center at a point of singularity, one diagonal block of size 2 × 2 in
𝐒 corresponds to eigenvalues with −1 < Re(𝐒) ̂ < 0, and leads to a stress singularity when 𝜉 approaches the point of singularity,

8
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

i.e. 𝜉 ≈ 0. All the eigenvalues −𝑆 𝑛 ≥ 0, and do not contribute to singularity [41]. The diagonal block leading to singularity is termed
as 𝐒(𝑠) , and the singular stress corresponding to this block is expressed as
(𝑠) −𝐈
𝝈 (𝑠) (𝜉, 𝜂) = 𝐃𝜳 (𝑠)
𝜖 (𝜂)𝜉
𝐒
𝐜(𝑠) (46)

where 𝐜(𝑠)
is obtained from the terms in the product 𝐓𝐮0 corresponding to the singular block 𝐒(𝑠) and 𝜳 (𝑠)
𝜖 (𝜂) are the singular strain
modes defined as

𝜳 𝜖(𝑠) (𝜂) = 𝐁1 (𝜂)(𝜳 (𝑠) )(𝑒) 𝐒(𝑠) + 𝐁2 (𝜂)(𝜳 (𝑠) )(𝑒) . (47)

To compute the generalized stress intensity factors, the singular stresses in Eq. (46) are first transformed to polar coordinates (𝑟, 𝜑)
( )𝐒(𝑠) −𝐈
𝑟
𝝈 (𝑠) (𝑟, 𝜑) = 𝐃𝜳 (𝑠)
𝜖𝐿
(𝜑) 𝐜(𝑠) , (48)
𝐿
where a characteristic length 𝐿 is introduced to make the SIFs independent of the system of units. The term 𝜳 (𝑠)
𝜖𝐿
(𝜑) indicates the
singular strain modes at characteristic length 𝐿 and is expressed as
( )𝐒(𝑠) −𝐈
(𝑠) 𝐿
𝜳 𝜖𝐿 (𝜑) = 𝜳 (𝑠)
𝜖 (𝜂(𝜑)) 𝑟(𝜑) . (49)

The generalized stress intensity factors at angle 𝜑 are defined in [44] as


( )𝐒(𝑠) −𝐈 ( )−1
1 𝑟
𝝈 (𝑠) (𝑟, 𝜑) = √ 𝐃𝜳 (𝑠)
𝜖𝐿
(𝜑) 𝐃𝜳 (𝑠)
𝜖𝐿
(𝜑) 𝑲(𝜑). (50)
2𝜋𝐿 𝐿
After selecting the relevant components of singular stress in Eq. (48) as
[ ]T
(𝑠) (𝑠)
𝝈 (𝑠) (𝑟, 𝜑) = 𝜎𝜑𝜑 (𝑟, 𝜑) 𝜎𝑟𝜑 (𝑟, 𝜑) , (51)

the corresponding singular strain modes 𝜳 (𝑠)


𝜖𝐿
(𝜑) become a 2 × 2 matrix, which contains only the rows corresponding to 𝜎𝜑𝜑 (𝑠)
(𝑟, 𝜑)
(𝑠)
and 𝜎𝑟𝜑 (𝑟, 𝜑). Comparing Eqs. (50) and (48), the stress intensity factors are calculated directly from their definition as

𝑲(𝜑) = 2𝜋𝐿𝐃𝜳 (𝑠)𝜖𝐿
(𝜑)𝐜(𝑠) . (52)

Similar to stress intensity factors, 𝑇 −stress is also calculated directly from its definition. In the following, we consider the first
and second term of Eq. (45) separately (multiplied by 𝐮0 and 𝜣 0 , respectively). The first term yields constant stress when a diagonal
block of 𝐒̂ 𝑛 is associated with an eigenvalue with Re(𝐒̂ 𝑛 ) = −1 or when an eigenvalue 𝑆̄𝑛(𝑖) = 0, such that the corresponding term
in 𝐒𝑏 equals one. These respective blocks in 𝐒 are referred to as 𝐒(𝑡) . In addition, the second term of Eq. (45) is constant when a
diagonal block of 𝐒̄ 𝑛 is associated with an eigenvalue 𝑆̄𝑛(𝑖) = 0. The corresponding blocks in 𝐒̄ 𝑛 are referred to as 𝐒̄ (𝑡)
𝑛 . The stresses
related to the blocks 𝐒(𝑡) and 𝐒̄ (𝑡)
𝑛 are expressed in scaled boundary coordinates as
( (𝑢),(𝑡) )(𝑒) (𝑡)
𝝈 (𝑡) (𝜉, 𝜂(𝜑)) = 𝐃𝜳 (𝑡) (𝑡)
𝜖 (𝜂(𝜑))𝐜 − 𝐃𝜷𝐍(𝜂) 𝐔𝑛 𝐝 , (53)

where 𝐜(𝑡) contains the terms in the product 𝐓𝐮0 corresponding to 𝐒(𝑡) and 𝐝(𝑡) contains the terms in the product ((𝐔(𝑢) −1
𝑛 ) 𝜣0)
corresponding to 𝐒̄ 𝑛(𝑡) . The term 𝜳 (𝑡)
𝜖 (𝜂(𝜑)) contains the strain modes corresponding to the block 𝐒(𝑡) and is expressed as

𝜳 𝜖(𝑡) (𝜂(𝜑)) = 𝐁1 (𝜂(𝜑))(𝜳 (𝑡) )(𝑒) 𝐒(𝑡) + 𝐁2 (𝜂(𝜑))(𝜳 (𝑡) )(𝑒) . (54)

The term 𝐔𝑛(𝑢),(𝑡)


denotes the eigenvectors of the thermal problem that are associated with zero eigenvalues, see Eq. (11). In [43],
𝑇 −stress is evaluated at the boundary, at (𝑟 = 𝐿, 𝜑 = 0). Evaluating Eq. (53) at the same point, the component 𝜎𝑥(𝑡)
̃ 𝑥̃ results in 𝑇 −stress.

6. Constitutive relations

This section presents the constitutive relations for FGMs under thermal loads. For isotropic FGMs, standard definitions of both
plane stress and plane strain states are used. In both states under isotropic conditions, only one value of Young’s modulus, Poisson’s
ratio, and coefficient of thermal expansion is necessary. However, for orthotropic FGMs the required number of thermoelastic
parameters is not the same in plane stress and plane strain cases [45]. In plane stress state, six constants are needed while in
plane strain state 10 constants are required to model the stress–strain relation.
Considering the principal axes 1 and 2 of orthotropy to be aligned with 𝑥 and 𝑦, respectively; the constitutive relation for the
plane stress state is expressed as
⎡𝜖11 ⎤ ⎡ 1∕𝐸1 −𝜈12 ∕𝐸1 0 ⎤ ⎡𝜎11 ⎤ ⎡𝛼1 𝛩⎤
⎢𝜖 ⎥ = ⎢−𝜈 ∕𝐸 1∕𝐸2 0 ⎥ ⎢𝜎22 ⎥ + ⎢𝛼2 𝛩⎥ . (55)
⎢ 22 ⎥ ⎢ 12 1 ⎥⎢ ⎥ ⎢ ⎥
⎣𝛾12 ⎦ ⎣ 0 0 1∕𝐺12 ⎦ ⎣𝜎12 ⎦ ⎣ 0 ⎦
For the plane strain state, information on the third principal direction is also required, and the constitutive relation is described as
⎡𝜖11 ⎤ ⎡ (1 − 𝜈31 𝜈13 )∕𝐸1 −(𝜈12 + 𝜈13 𝜈32 )∕𝐸1 0 ⎤ ⎡𝜎11 ⎤ ⎡(𝜈31 𝛼3 + 𝛼1 )𝛩⎤
⎢𝜖 ⎥ = ⎢−(𝜈 + 𝜈 𝜈 )∕𝐸 (1 − 𝜈23 𝜈32 )∕𝐸2 0 ⎥ ⎢𝜎22 ⎥ + ⎢(𝜈32 𝛼3 + 𝛼2 )𝛩⎥ . (56)
⎢ 22 ⎥ ⎢ 12 13 32 1
⎥⎢ ⎥ ⎢ ⎥
⎣𝛾12 ⎦ ⎣ 0 0 1∕𝐺12 ⎦ ⎣𝜎12 ⎦ ⎣ 0 ⎦

9
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 3. (a) Geometry and boundary conditions and (b) mesh at 𝑎∕𝑊 = 0.5 of a plate with an edge crack under thermal loads.

Table 1
Material parameters with linear gradation of a plate with an edge crack.
Young’s modulus Poisson’s ratio Coefficient of thermal
𝑥 [MPa] expansion [1∕K]
0 𝐸1 = 1 × 105 𝜈1 = 0.3 𝛼1 = 1.67 × 10−5
𝑊 𝐸2 = 0.5 × 105 𝜈2 = 0.35 𝛼2 = 1 × 10−5

The above constants are related as

𝜈21 ∕𝐸2 = 𝜈12 ∕𝐸1 , 𝜈31 ∕𝐸3 = 𝜈13 ∕𝐸1 , 𝜈32 ∕𝐸3 = 𝜈23 ∕𝐸2 . (57)

Here, 𝐸𝑖 , 𝜈𝑖𝑗 , 𝐺𝑖𝑗 , 𝛼𝑖 , and 𝛩 are Young’s moduli, Poisson’s ratios, shear moduli, coefficients of thermal expansion and temperature
change, respectively. The indices 𝑖𝑗 describe the principal direction. Note that the material properties used in the calculation of SIFs
and 𝑇 −stress are evaluated at the scaling center.

7. Numerical examples

In this section, the performance of the developed method in evaluating the stress intensity factors and 𝑇 −stress for FGM plates
under coupled thermo-mechanical loads is examined. For this purpose, three isotropic FGM and one orthotropic FGM benchmark
examples are considered. The findings are compared with available numerical results. The three examples addressing isotropic FGMs
consider different types of grading. A plate with an edge crack and linearly graded material is modeled in Section 7.1. In Section 7.2 a
plate with an inclined center crack and exponential grading is considered. The third example (c.f. Section 7.3) illustrates the effect
of an edge crack in a thermal barrier coating (TBC). An orthotropic FGM with gradation perpendicular to the crack surface is
addressed in Section 7.4. The crack is assumed to be insulated in all examples, and the polygon edges are discretized with one cubic
line element in general. Each edge on the crack polygon boundary is further divided into four line elements to capture the angular
variation of stress intensity factors accurately.

7.1. Plate with an edge crack: linear material gradient

In the first example, the validity of the proposed method in calculating the stress intensity factors and the 𝑇 −stress is assessed.
For this purpose, a plate with an edge crack illustrated in Fig. 3(a) is considered. Linear functions of 𝑥 describe the variation of
Young’s modulus 𝐸, Poisson’s ratio 𝜈 and coefficient of thermal expansion 𝛼. A plane strain state is considered.
The geometry of the plate is defined as 𝐿 = 𝑊 = 1 m with varying crack length 𝑎. The right side of the plate is constrained in the
𝑥−direction. The midpoint of the right side is also constrained in the 𝑦−direction. The steady-state temperature change is achieved
by maintaining the temperature of the left side at reference 𝛩0 = 𝛩1 = 𝛩(𝑥 = 0) = 0 ◦ C and increasing the temperature of the right

10
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Table 2
Number of polygons and nodes at varying crack length 𝑎∕𝑊 .
𝑎∕𝑊 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Polygons 2254 2320 2355 2417 2438 2421 2410 2407
Nodes 18691 19246 19535 20067 20238 20120 20050 20044

Fig. 4. (a) Mode-I stress intensity factor and (b) 𝑇 −stress in linear FGM under thermal loads.

side to 𝛩2 = 𝛩(𝑥 = 𝑊 ) = 1 ◦ C. The crack, top, and bottom edges are assumed to be insulated. The material properties1 of the plate
are summarized in Table 1.
Table 2 presents the number of polygons as well as nodes employed at each 𝑎∕𝑊 . Fig. 3(b) shows the mesh at 𝑎∕𝑊 = 0.5.
The applied boundary conditions result in mode-I fracture. Mode-I stress intensity factors 𝐾𝐼 and 𝑇 −stress for varying 𝑎∕𝑊 are
compared with the reference solution [8] and illustrated in Fig. 4.
It is observed that the values of 𝐾𝐼 increase until 𝑎∕𝑊 = 0.3 for homogeneous material and until 𝑎∕𝑊 = 0.2 for FGM. After
that, 𝐾𝐼 starts to decrease with increasing 𝑎∕𝑊 . The decrement of 𝐾𝐼 is observed due to the constrained boundary conditions on
the right-side. The 𝑇 −stress continues to increase with increasing 𝑎∕𝑊 for both homogeneous material and FGM. Furthermore, for
all the 𝑎∕𝑊 , the absolute values of 𝐾𝐼 and 𝑇 −stress obtained for FGM are lower than those obtained for a homogeneous material.
The results are in good agreement with the reference solution [8].
For 𝐾𝐼 , in the FGM case, the average difference of 0.1% is noted compared to the reference. These results are obtained for
all values of 𝑎∕𝑊 except 𝑎∕𝑊 = 0.8, which corresponds to a very small value of 𝐾𝐼 . For the homogeneous case, all 𝑎∕𝑊 are
considered, and the average difference of 0.087% is observed. It is also realized that the 𝑇 −stress strongly depends on material
gradation. Average differences of 2.26% and 8.3% are observed for homogeneous and FGM cases, respectively. Small values of
𝑇 −stress are excluded in the calculations.

7.2. Plate with an inclined center crack: exponential material gradient

In this example, mixed-mode fracture of a plate with an inclined center crack as shown in Fig. 5(a) is studied. The geometry of
the plate is defined as 𝑊 = 𝐿 = 20 cm and 𝑎 = 1 cm. Different crack orientations of 𝜑 = 0◦ to 90◦ measured counter clock-wise from
the 𝑥-axis are considered. The plate is assumed to be in a plane stress state and subjected to a steady-state thermal load. Fig. 5(b)
illustrates the mesh of the plate at 𝜑 = 45◦ with 1506 polygons. Young’s modulus follows an exponential function, while Poisson’s
ratio is kept constant. The material properties and the temperature boundary conditions for the isotropic case are defined as

𝐸(𝑥) = 𝐸 0 e(𝛽𝑥) , 𝜈(𝑥) = 𝜈, 𝜖 = −𝛼(𝑥)𝛩(𝑥) = 1,

with 𝐸 0 = 1 MPa and 𝜈 = 0.3. The non-homogeneity parameter 𝑎𝛽 = 0.5 is considered.


The top and bottom sides of the plate are constrained in the 𝑦 direction. Also, the midpoints of both sides are constrained in the
𝑥 direction. The left, right, and bottom sides and the crack edge are assumed to be insulated. The applied boundary conditions of
𝑚𝑒𝑐ℎ (𝑥, 𝑦) = 𝜖 in a setting corresponding to an uncracked structure. For
steady-state thermal load result in uniform mechanical strain 𝜖𝑦𝑦

1 For homogeneous material 𝐸 = 𝐸1 , 𝜈 = 𝜈1 , and 𝛼 = 𝛼1 are considered.

11
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 5. (a) Geometry and boundary conditions and (b) mesh at 𝜑 = 45◦ of a plate with an inclined center crack.

Fig. 6. Normalized SIFs 𝐾 = 𝐾∕𝐾0 for non-homogeneity parameter 𝑎𝛽 = 0.5 in isotropic FGM for an inclined center crack. (a) Right crack tip and (b) left crack
tip.

Table 3
Average percent differences in normalized SIFs compared to the reference
solutions.
𝐾+ 𝐾−

References 𝐾𝐼+ +
𝐾𝐼𝐼 𝐾𝐼− −
𝐾𝐼𝐼
Konda et al. [46] 1.37 2.94 1.84 2.93
Amit et al. [8] 1.37 2.89 1.65 2.04
Hosseini et al. [13] 1.37 4.01 1.01 3.20

simplicity, the coefficient of thermal expansion and temperature change are chosen as 𝛼(𝑥) = 𝛼 = 1 K1 and 𝛩(𝑥) = −1 ◦ C, respectively.

The stress intensity factors are normalized by a factor 𝐾0 = 𝜖𝐸 0 𝜋𝑎, and the 𝑇 −stress is normalized by 𝜎0 = 𝜖𝐸 0 .
Fig. 6 illustrates the behavior of the normalized stress intensity factors 𝐾 = 𝐾∕𝐾0 compared with the solutions in Refs. [8,13,46].
It is noted that the mode-I stress intensity factors decrease as the crack inclines towards the vertical axis. In general, higher values
of stress intensity factors are observed for the right crack tip because the stiffness increases along the 𝑥−direction.
Table 3 shows the average differences of normalized SIFs for crack inclination angles 𝜑 in comparison to the reference
solutions [8,13,46]. For both cracks, 𝐾 is close to zero when 𝜑 = 90◦ , and 𝐾𝐼𝐼 is also close to zero when 𝜑 = 0◦ . These are
excluded from the calculations of average difference. The maximum average difference of 4.01% is noted between SBFEM and
+
reference solution of 𝐾𝐼𝐼 in [13]. The minimum average difference of 1.01% is observed for 𝐾𝐼− between SBFEM and reference
solution in [13].
Fig. 7 presents 𝑇 −stress for both crack tips compared to the reference solutions [8,47]. 𝑇 −stress is observed to grow with
increasing crack orientation angle. For the right crack tip, average differences of 8.9% and 8.0% are calculated compared to the

12
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 7. Normalized 𝑇 −stress 𝑇 = 𝑇 ∕𝑇0 for non-homogeneity parameter 𝑎𝛽 = 0.5 in isotropic FGM for an inclined center crack. (a) Right crack tip and (b) left
crack tip.

Fig. 8. Thermal barrier coating with periodic cracks.

Table 4
Number of polygons and nodes at varying crack length 𝑎∕𝑊1 .
𝑎 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Polygons 1448 1510 1532 1584 1808 1745 1721 1663
Nodes 12084 12586 12780 13208 15009 14526 14361 13891

Refs. [8,47], respectively. For the left crack tip these values are 4.36% and 3.0%, respectively. At 𝜑0 = 45◦ , the values of 𝑇 −stress for
both crack tips are close to zero, hence excluded from the calculations. A higher value of percent difference is observed for the right
crack tip. In general, the results are in good agreement with the reference solutions [8,47]. Note that the results in Refs. [46,47]
are obtained by applying an equivalent mechanical loading.

7.3. FGM plate with thermal barrier coating (TBC)

In this example, a plate coated with a functionally graded thermal barrier, as illustrated in Fig. 8 is considered [48]. The plate
comprises of an isotropic FGM coating made of 100% Zirconia–Yttria at 𝑥 = 0 and 100% nickel–chromium–aluminum–zirconium
(NiCrAlY) isotropic bond coat at 𝑥 = 𝑊1 . The metallic substrate is made of nickel-based super alloy. Hyperbolic tangent functions
describe the steep transition of material properties between the bond coat and metal substrate. The plate also contains periodic
cracks at the edge. Due to the cracks’ symmetry, only one of them is modeled. Fig. 9(a) shows the boundary conditions of the plate
after exploiting symmetry.
The geometry of the plate is defined as 𝑏 = 2 m, 𝑊1 = 1 m, 𝑊2 = 0.5 m, and 𝑊3 = 5 m. An edge crack with varying length 𝑎
is assumed. The top and bottom sides of the plate are constrained in the 𝑦 direction. The corner points of the right side are also
constrained in the 𝑥 direction. The plate is considered to be at a reference temperature 𝛩0 = 1000 ◦ C. The temperature change of
𝛩1 = 𝛩(𝑥 = 0) = 0.2 𝛩0 is applied at the left side, while the temperature change of 𝛩2 = 𝛩(𝑥 = 𝑊1 + 𝑊2 + 𝑊3 ) = 0.5 𝛩0 is applied at
the right side. Plane strain state is considered. Fig. 9(b) illustrates the mesh of the plate at 𝑎∕𝑊1 = 0.5. Table 4 presents the number
of polygons and nodes used to discretize the plate at varying crack length 𝑎∕𝑊1 .

13
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Table 5
Material parameters of isotropic FGM TBC.
Material FGM Bond coat Metal substrate
parameters (Zirconia–Yttria) (NiCrAlY) (Ni)
Young’s modulus [GPa] 𝐸 𝑐 = 27.6 𝐸 𝑏𝑐 = 137.9 𝐸 𝑠 = 175.8
Poisson’s ratio 𝜈 𝑐 = 0.25 𝜈 𝑏𝑐 = 0.27 𝜈 𝑠 = 0.25
Coeff. of thermal expansion [1∕◦ C] 𝛼 𝑐 = 10.01 × 10−6 𝛼 𝑏𝑐 = 15.16 × 10−6 𝛼 𝑠 = 13.91 × 10−6
Thermal conductivity [W∕(m K)] 𝜅𝑐 = 1 𝜅 𝑏𝑐 = 25 𝜅𝑠 = 7

Fig. 9. (a) Boundary conditions after applying symmetry and (b) mesh at 𝑎∕𝑊1 = 0.5 of a plate with thermal barrier coating.

Fig. 10. Mode-I stress intensity factor 𝐾𝐼 and 𝑇 −stress of isotropic TBC under thermal loads.

Table 5 presents the material properties of isotropic FGM coat, bond coat, and metal substrate. For isotropic FGM coating region,
the grading functions are defined as

𝐸(𝑥) = 𝐸 𝑐 + (𝐸 𝑏𝑐 − 𝐸 𝑐 )𝑥2 , 𝜈(𝑥) = 𝜈 𝑐 + (𝜈 𝑏𝑐 − 𝜈 𝑐 )𝑥,


𝛼(𝑥) = 𝛼 + (𝛼 𝑐 𝑏𝑐 𝑐
− 𝛼 )𝑥, 𝜅(𝑥) = 𝜅 𝑐 + (𝜅 𝑏𝑐 − 𝜅 𝑐 )𝑥2 .

The functions describing the steep gradation of material properties between the bond coat and metal substrate are assumed to be
hyperbolic tangent functions defined as follow

𝑃 𝑠 + 𝑃 𝑏𝑐 𝑃 𝑠 − 𝑃 𝑏𝑐
𝑃 (𝑥) = + tanh(𝛽𝑥),
2 2
1
where 𝛽 = 100 m
and 𝑃 = 𝐸, 𝜈, 𝛼, and 𝜅. The values of 𝑃 𝑠 and 𝑃 𝑏𝑐 are taken from Table 5.
The applied thermal loads result in mode-I fracture. The results of 𝐾𝐼 and 𝑇 −stress are compared with reference solutions [8,49]
and presented in Fig. 10.
It is noted that the 𝐾𝐼 increases, while 𝑇 −stress decreases with increasing crack length 𝑎. The average differences of 0.26% and
0.58% are noted for 𝐾𝐼 compared to [8,49], respectively. For 𝑇 −stress, these values are 13.37% and 5.05%. It is again observed
that the 𝑇 −stress are highly influenced by the material gradation.

14
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 11. Orthotropic FGM plate with center crack.

Fig. 12. (a) Boundary conditions after exploiting symmetry and (b) mesh with 𝑏∕𝑊 = 0.4, 𝑏1 ∕𝑊 = 0.3 and 𝑎∕𝑊 = 0.1.

Table 6
Material parameters of orthotropic ceramic at (𝑦 = 𝑏).
Alumina (Al2 O3 )
Modulus [GPa] 𝐸1𝑐 = 90.43 𝐸2𝑐 = 116.36 𝐺12𝑐
= 38.21
Coeff. of thermal expansion [1∕◦ C] 𝛼1𝑐 = 8 × 10−6 𝛼2𝑐 = 7.5 × 10−6 𝛼3𝑐 = 9 × 10−6
Thermal conductivity [W∕(m K)] 𝜅1𝑐 = 21.25 𝜅2𝑐 = 29.82
𝑐 𝑐
Poisson’s ratio 𝜈12 = 0.22 𝜈13 = 0.14
𝑐 𝑐
𝜈31 = 0.14 𝜈32 = 0.21

7.4. Plate with orthotropic material gradient perpendicular to crack surface

In the last example, an orthotropic FGM plate with a center crack reported in [50] is considered, as illustrated in Fig. 11. In this
example, material gradation and temperature distribution are applied perpendicular to the crack surface resulting in mixed-mode
fracture.
The bottom side (𝑦 = 0) is assumed to be 100% metal (nickel Ni), and the top side (𝑦 = 𝑏) is assumed to be 100% orthotropic
ceramic (alumina Al2 O3 ). The plate is initially at a reference temperature of 𝛩0 = 1100 K. Steady-state temperature distribution is
applied by maintaining the temperature of the bottom edge at reference 𝛩(𝑦 = 0) = 𝛩0 , and increasing the temperature of the top
edge to 𝛩(𝑦 = 𝑏) = 2𝛩0 . The right and left sides are assumed to be insulated. A plane strain state is considered.
The temperature boundary conditions and the geometry of the plate are symmetrical about the vertical axis. By exploiting the
symmetry only half of the plate is discretized. Fig. 12(a) shows the geometry of the plate after employing symmetry. Fig. 12(b)
presents the mesh of the plate at 𝑎∕𝑊 = 0.1, 𝑏1 ∕𝑊 = 0.3, and 𝑏∕𝑊 = 0.4. The transition in material properties from metal to
ceramic is represented by a power-law as follows

𝑃 (𝑦) = 𝑃 𝑚 + (𝑃 𝑐 − 𝑃 𝑚 )(𝑦∕𝑏)𝑛 . (58)

where 𝑃 (𝑦) is the variation of the material properties 𝐸1 , 𝐸2 , 𝐺12 , 𝜈12 , 𝜈13 , 𝜈31 , 𝜈32 , 𝛼1 , 𝛼2 , 𝛼3 , 𝜅1 , and 𝜅2 . The superscript 𝑛 defines
the corresponding exponent of these properties as 𝛾1 , 𝛾2 , 𝛾12 , 𝛽12 , 𝛽13 , 𝛽31 , 𝛽32 , 𝛿1 , 𝛿2 , 𝛿3 , 𝜔1 , and 𝜔2 , respectively. 𝑃 𝑐 and 𝑃 𝑚 indicate
the properties of ceramic and metal, respectively. The above power-law functions Eq. (58) are widely used to model the behavior of
FGMs. The exponents are positive constants, 𝑛 < 1 indicates the metal-rich profile while 𝑛 > 1 indicates the ceramic-rich profile [50].
The isotropic material properties 𝑃 𝑚 of metal (Ni) at (𝑦 = 0) are taken as
1 W
𝐸 𝑚 = 204 GPa, 𝜈 𝑚 = 0.31, 𝛼 𝑚 = 13.3 × 10−6 ◦C , 𝜅 𝑚 = 70 mK
. (59)

The properties 𝑃𝑐 of 100% orthotropic ceramic (alumina Al2 O3 ) at (𝑦 = 𝑏) are listed in Table 6.

15
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

Fig. 13. (a) Normalized mode-I SIF 𝐾𝐼 and, (b) normalized mode-II SIF 𝐾𝐼𝐼 for 𝑎∕𝑊 = 0.1 and 𝑏∕𝑊 = 0.4 for varying 𝜔2 versus vertical location of the crack
𝑏1 ∕𝑊 .

Fig. 14. Normalized 𝑇 for 𝑎∕𝑊 = 0.1 and 𝑏∕𝑊 = 0.4 for varying 𝜔2 versus vertical location of the crack 𝑏1 ∕𝑊 .

The remaining properties 𝐸3𝑐 , 𝜈21


𝑐 , and 𝜈 𝑐 can be calculated using orthotropic material identities Eq. (57).
23
The effects of all the material properties on SIFs and 𝑇 −stress can be considered independently. For simplicity and comparison
purpose, the influence of varying thermal conductivity’s exponent in the 𝑦−direction (𝜔2 ) versus the vertical location of the crack
𝑏1 on thermal fracture parameters are studied. All the other exponents 𝑛 are constant. Their values are 𝛾1 = 𝛾2 = 𝛾12 = 2,

𝛽12 = 𝛽13 = 𝛽31 = 𝛽32 = 1.5, 𝛿1 = 𝛿2 = 𝛿3 = 3, and 𝜔1 = 4. The normalized SIFs are defined as 𝐾 = 𝐾∕𝐾0 , with 𝐾0 = 𝜎0 𝜋𝑎. The
𝑐 𝑐
normalized 𝑇 −stress is defined as 𝑇 = 𝑇 ∕𝜎0 , where 𝜎0 = 𝛼1 𝐸1 𝛩0 . Figs. 13(a), 13(b), and 14 present the comparison of normalized
𝐾𝐼 , 𝐾𝐼𝐼 , and 𝑇 −stress with the reference solutions reported in [50].
It is noted that 𝐾𝐼 increases with increasing 𝑏1 ∕𝑊 , when 𝜔2 is kept constant. On the other hand, 𝐾𝐼𝐼 decreases with increasing
𝑏1 ∕𝑊 when 𝜔2 is constant. In general, higher values of 𝐾𝐼𝐼 are observed compared to 𝐾𝐼 for all 𝑏1 ∕𝑊 . The 𝑇 −stress is noted
to be compressive for 𝑏1 ∕𝑊 = 0.2, 0.25, 0.3 and tensile for 𝑏1 ∕𝑊 = 0.35, respectively. The magnitude of 𝑇 −stress decreases with
increasing 𝜔2 . The results of presented fracture parameters are in good agreement with the solutions reported in Ref. [50].

16
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

8. Conclusions

The scaled boundary finite element method is developed to model fracture in FGMs under thermo-mechanical loads. SBFEM
shape functions for thermoelastic equilibrium are constructed and partitioned into homogeneous and non-homogeneous parts.
The homogeneous part maps the general mechanical behavior, while the non-homogeneous part represents the thermal stress
contribution. These shape functions are expressed in the form of power functions.
The stiffness matrix containing the material gradient is constructed by exploiting the semi-analytical nature of SBFEM shape
functions. Owing to the semi-analytical solution of stresses, the thermal fracture parameters, i.e. stress intensity factors and 𝑇 −stress,
are extracted from their definitions directly. The semi-analytical nature of stresses obviates the need for additional post-processing
techniques completely. Moreover, polygon elements with an arbitrary number of edges are employed for robust mesh generation.
Validation of the proposed method is demonstrated using several numerical examples. The results of fracture parameters have
shown good agreement. In summary, the scaled boundary finite element method is an attractive alternative to other numerical
methods for computing thermally-induced fracture parameters in functionally graded materials.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared
to influence the work reported in this paper.

Acknowledgments

The research reported herein was partially performed within the scope of the Australia–Germany Joint Research Cooperation
and DAAD-PPP (Australia) Schemes. The financial support of Universities Australia and the German Federal Ministry of Education
and Research represented by the German Academic Exchange Service are gratefully acknowledged.

Appendix A. Scaled boundary transformation

1 [ ]T
𝐛1 (𝜂) = 𝑦𝜂,𝜂 (𝜂) −𝑥𝜂,𝜂 (𝜂) , (A.1a)
𝐽 (𝜂)
1 [ ]T
𝐛2 (𝜂) = −𝑦𝜂 (𝜂) 𝑥𝜂 (𝜂) . (A.1b)
𝐽 (𝜂)

⎡ 𝑦 (𝜂) 0 ⎤
1 ⎢ 𝜂,𝜂 ⎥,
𝐛1 (𝜂) = 0 −𝑥𝜂,𝜂 (𝜂) (A.2a)
𝐽 (𝜂) ⎢ ⎥
⎣ −𝑥𝜂,𝜂 (𝜂) 𝑦𝜂,𝜂 (𝜂) ⎦
⎡ −𝑦𝜂 (𝜂) 0 ⎤
1 ⎢ ⎥,
𝐛2 (𝜂) = 0 𝑥𝜂 (𝜂) (A.2b)
𝐽 (𝜂) ⎢ ⎥
⎣ 𝑥𝜂 (𝜂) −𝑦𝜂 (𝜂) ⎦
where 𝐽 (𝜂) is the determinant of the Jacobian matrix 𝐉(𝜂) with
[ ]
𝑥𝜂 (𝜂) 𝑦𝜂 (𝜂)
𝐉(𝜂) = . (A.3)
𝑥𝜂,𝜂 (𝜂) 𝑦𝜂,𝜂 (𝜂)

Appendix B. Coefficient matrices of scaled boundary finite element equations

The coefficient matrices for steady-state heat conduction are expressed as


1 T
𝐄0 = 𝐁1 (𝜂)𝜿𝐁1 (𝜂)𝐽 (𝜂)𝑑𝜂, (B.1a)
∫−1
1 T
𝐄1 = 𝐁2 (𝜂)𝜿𝐁1 (𝜂)𝐽 (𝜂)𝑑𝜂, (B.1b)
∫−1
1 T
𝐄2 = 𝐁2 (𝜂)𝜿𝐁2 (𝜂)𝐽 (𝜂)𝑑𝜂, (B.1c)
∫−1

where 𝜿 is the thermal conductivity matrix, and the 𝐁𝑖 matrices are defined as

𝐁1 =𝐛1 (𝜂)𝐍(𝜂), (B.2a)


𝐁2 =𝐛2 (𝜂)𝐍(𝜂),𝜂 . (B.2b)

17
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

The coefficient matrices for elasticity are expressed as


1
𝐄0 = 𝐁T1 (𝜂)𝐃𝐁1 (𝜂)𝐽 (𝜂)𝑑𝜂, (B.3a)
∫−1
1
𝐄1 = 𝐁T2 (𝜂)𝐃𝐁1 (𝜂)𝐽 (𝜂)𝑑𝜂, (B.3b)
∫−1
1
𝐄2 = 𝐁T2 (𝜂)𝐃𝐁2 (𝜂)𝐽 (𝜂)𝑑𝜂, (B.3c)
∫−1

where 𝐃 is the elasticity matrix, and the 𝐁𝑖 matrices are defined as

𝐁1 =𝐛1 (𝜂)𝐍(𝜂), (B.4a)


𝐁2 =𝐛2 (𝜂)𝐍(𝜂),𝜂 . (B.4b)

References

[1] Salimi Esmaeil. Functionally graded calcium phosphate bioceramics: An overview of preparation and properties. Ceram Int 2020;46:19664–8.
[2] Saleh Bassiouny, Jiang Jinghua, Fathi Reham, Al-hababi Tareq, Xu Qiong, Wang Lisha, Song Dan, Ma Aibin. 30 years of functionally graded materials:
An overview of manufacturing methods, applications and future challenges. Composites B 2020;201:108376.
[3] Erdogan F, Wu BH. Crack problems in FGM layers under thermal stress. J Therm Stresses 1996;19(3):237–65.
[4] Dag S, Kadioglu S, Yahsi OS. Circumferential crack problem for an FGM cylinder under thermal stresses. J Therm Stresses 1999;22(7):659–87.
[5] Jin ZH, Paulino GH. Transient thermal stress analysis of an edge crack in a functionally graded material. Int J Fract 2001;107(1):73–98.
[6] Guo LC, Noda N, Ishihara M. Thermal stress intensity factors for a normal surface crack in a functionally graded coating structure. J Therm Stresses
2008;31(2):149–64.
[7] Kim JH, Paulino GH. Isoparametric graded finite elements for nonhomogeneous isotropic and orthotropic materials. J Appl Mech 2002;69(4):502–14.
[8] Amit KC, Kim JH. Interaction integral for thermal fracture of functionally graded materials. Eng Fract Mech 2008;75:2542–65.
[9] Burlayenko VN, Altenbach H, Sadowski T, Dimitrova SD, Bhaskar A. Modelling functionally graded materials in heat transfer and thermal stress analysis
by means of graded finite elements. Appl Math Model 2017;45:422–38.
[10] Nojumi MM, Wang X. Analysis of crack problems in functionally graded materials under thermomechanical loading using graded finite elements. Mech
Res Commun 2020;106:103534.
[11] Kim JH, Paulino GH. Mixed-mode fracture of orthotropic functionally graded materials using finite elements and the modified crack closure method. Eng
Fract Mech 2002;69(14–16):1557–86.
[12] Dag S, Yildirim B, Sarikaya D. Mixed-mode fracture analysis of orthotropic functionally graded materials under mechanical and thermal loads. Int J Solids
Struct 2007;44:7816–40.
[13] Hosseini SS, Bayesteh H, Mohammadi S. Thermo-mechanical XFEM crack propagation analysis of functionally graded materials. Mater Sci Eng A
2013;561:285–302.
[14] Goli E, Bayesthe H, Mohammadi S. Mixed-mode fracture analysis of adiabatic cracks in homogeneous materials and nonhomogeneous materials in the
framework of partition of unity and the path independent interaction integral. Eng Fract Mech 2014;131:100–27.
[15] Bouhala L, Makradi A, Belouettar S. Thermo-anisotropic crack propagation by XFEM. Int J Mech Sci 2015;103:235–46.
[16] Bayat SH, Nazari MB. Thermal fracture analysis in orthotropic materials by XFEM. Theor Appl Fract Mech 2021;112:102843.
[17] Ekhlakov AV, Khay OM, Zhang C, Sladek J, Sladek V. A BDEM for transient thermoelastic crack problems in functionally graded materials under thermal
shock. Comput Mater Sci 2012;57:30–7.
[18] Belytschko T, Lu YY, Gu L. Element-free Galerkin methods. Internat J Numer Methods Engrg 1994;37:229–56.
[19] Atluri SN, Zhu TL. The meshless local Petrov-Galerkin (MLPG) approach for solving problems in elasto-statics. Comput Mech 2000;25:169–79.
[20] Shi GH. Manifold method of material analysis. In: Transactions of the 9th army conference on applied mathematics and computing. Minnesota, Minneapolis:
University of Minnesota; 1991, p. 57–76.
[21] Zhang HH, Ma GW, Ren F. Implementation of the numerical manifold method for thermo-mechanical fracture of planar solids. Eng Anal Bound Elem
2014;44:45–54.
[22] Zhang HH, Liu SM, Han SY, Fan SL. Modeling of 2D crack FGMs under thermo-mechanical loadings with the numerical manifold method. Int J Mech Sci
2018;148:103–17.
[23] Zhang HH, Liu SM, Han SY, Fan LF. The numerical manifold method for crack modeling of two-dimensional functionally graded materials under thermal
shock. Eng Fract Mech 2019;1:90–106.
[24] Zhang HH, Ji XL, Han SY, Fan LF. Determination of 𝑇 −stress for thermal cracks in homogeneous and functionally graded materials with numerical manifold
method. Theor Appl Fract Mech 2021;113:102940.
[25] Song C, Wolf JP. The scaled boundary finite-element method–alias consistent infinitesimal finite-element cell method–for elastodynamics. Comput Methods
Appl Mech Engrg 1997;147(4–7):329–55.
[26] Song C, Ooi ET, Natarajan S. A review of the scaled boundary finite element method for two-dimensional linear elastic fracture mechanics. Eng Fract
Mech 2018;187:45–73.
[27] Dölling S, Bremm S, Kohlstetter A, Felger J, Becker W. Predicting thermally induced edge-crack initiation using finite fracture mechanics. Eng Fract Mech
2021;252:107808.
[28] Yang ZJ. Fully automatic modelling of mixed-mode crack propagation using scaled boundary finite element method. Eng Fract Mech 2006;73(12):1711–31.
[29] Ooi ET, Song C, Tin-Loi F, Yang ZJ. Polygon scaled boundary finite elements for crack propagation modelling. Internat J Numer Methods Engrg
2012;268:905–37.
[30] Dai Shangqiu, Augarde Charles, Du Chengbin, Chen Denghong. A fully automatic polygon scaled boundary finite element method for modelling crack
propagation. Eng Fract Mech 2015;133:163–78.
[31] Zhang Peng, Du Chengbin, Zhao Wenhu, Sun Liguo. Dynamic crack face contact and propagation simulation based on the scaled boundary finite element
method. Comput Methods Appl Mech Engrg 2021;385:114044.
[32] Chiong I, Ooi ET, Song C, Tin-Loi F. Scaled boundary polygons with application to fracture analysis of functionally graded materials. Internat J Numer
Methods Engrg 2014;98(8):562–89.
[33] Ooi ET, Natarajan S, Song C. Crack propagation modeling in functionally graded materials using scaled boundary polygons. Int J Fract 2015;192:87–105.
[34] Chen X, Luo T, Ooi ET, Ooi EH, Song C. A quadtree-polygon-based scaled boundary finite element method for crack propagation modelling in functionally
graded materials. Theor Appl Fract Mech 2018;94:120–33.

18
M.D. Iqbal et al. Engineering Fracture Mechanics 264 (2022) 108305

[35] Liu Jun, Hao Congkuan, Ye Wenbin, Yang Fan, Lin Gao. Free vibration and transient dynamic response of functionally graded sandwich plates with
power-law nonhomogeneity by the scaled boundary finite element method. Comput Methods Appl Mech Engrg 2021;376:113665.
[36] Ooi ET, Iqbal MD, Birk C, Natarajan S, Ooi EH, Song C. A polygon scaled boundary finite element formulation for transient coupled thermoelastic fracture
problems. Eng Fract Mech 2020;240:107300.
[37] Wolf John P. The scaled boundary finite element method. Chichester: John Wiley & Sons Ltd; 2003.
[38] Song Chongmin, Wolf John P. The scaled boundary finite element method–alias consistent finite element cell method–for diffusion. Internat J Numer
Methods Engrg 1999;45:1403–31.
[39] Song Chongmin. The scaled boundary finite element method: Introduction to theory and implementation. John Wiley & Sons Ltd; 2018.
[40] Ooi ET, Iqbal MD, Birk C, Natarajan S, Ooi EH, Song C. A polygon scaled boundary finite element formulation for transient coupled thermoelastic fracture
problems. Eng Fract Mech 2020;240:107300.
[41] Song C. Analysis of singular stress fields at multi-material corners under thermal loading. Internat J Numer Methods Engrg 2006;65:620–52.
[42] Williams Max L. On the stress distribution at the base of a stationary crack. J Appl Mech 1957;24:109–14.
[43] Song C. Evaluation of power-logarithmic singularities, 𝑇 −stresses and higher order terms of in-plane singular stress fields at cracks and multi-material
corners. Eng Fract Mech 2005;72(10):1498–530.
[44] Song C, Tin-Loi F, Gao W. A definition and evaluation procedure of generalized stress intensity factors at cracks and multi-material wedges. Eng Fract
Mech 2010;77(12):2316–36.
[45] Dag S. Thermal fracture analysis of orthotropic functionally graded materials using an equivalent domain integral approach. Eng Fract Mech
2006;73(18):2802–28.
[46] Konda N, Erdogan F. The mixed mode crack problem in a nonhomogeneous elastic medium. Eng Fract Mech 1994;47(4):533–45.
[47] Paulino GH, Kim J-H. A new approach to compute 𝑇 −stress in functionally graded materials by means of the interaction integral method. Eng Fract Mech
2004;71(13–14):1907–50.
[48] Yildirim B. An equivalent domain integral method for fracture analysis of functionally graded materials under thermal stresses. J Therm Stresses
2006;29(4):371–97.
[49] Guo F, Huang K, Guo L, Bai X, Zhong S, Yu H. An interaction energy integral method for 𝑇 −stress evaluation in nonhomogeneous materials under thermal
loading. Mech Mater 2015;83:30–9.
[50] Dag S, Arman Erhan E, Yildirim B. Computation of thermal fracture parameters for orthotropic functionally graded materials using 𝐽𝑘 -integral. Int J Solids
Struct 2010;47(25–26):3480–8.

19

You might also like