You are on page 1of 49

Journal Pre-proof

Accurate 3D thermal stress analysis of thermal barrier coatings

Qiuhua Li , Pengfei Hou , Shouming Shang

PII: S0020-7403(21)00722-0
DOI: https://doi.org/10.1016/j.ijmecsci.2021.107024
Reference: MS 107024

To appear in: International Journal of Mechanical Sciences

Received date: 17 July 2021


Revised date: 17 December 2021
Accepted date: 19 December 2021

Please cite this article as: Qiuhua Li , Pengfei Hou , Shouming Shang , Accurate 3D thermal stress
analysis of thermal barrier coatings, International Journal of Mechanical Sciences (2021), doi:
https://doi.org/10.1016/j.ijmecsci.2021.107024

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier Ltd.


Accurate 3D thermal stress analysis of thermal barrier coatings
Qiuhua Lia,b, Pengfei Hou*a,b , Shouming Shanga,b
a
HNU College of Mechanical and Vehicle Engineering, Hunan University, Changsha 410082, PR China
b
Department of Engineering Mechanics, Hunan University, Changsha 410082, PR China

Graphical Abstract

1
Highlights:

(1) An accurate and efficient calculation method is proposed for 3D full-field thermal stress of
coating structure.
(2)
(3) The 3D full field analytical solution is given in the form of concise elementary functions.
(4)
(5) The effects of material parameters and coating thickness on the stress field are investigated
(6)
(7) The interface failure problem and material mismatch problem are analyzed

(8) This work would provide a guideline for the design of coating structure.

Authorship contribution statement


Qiuhua Li: Methodology, Software, Investigation, Formal Analysis, Writing-Original
Draft; Pengfei Hou: Conceptualization, Supervision, Writing-Review & Editing;
ShouMing Shang: Visualization, Software, Investigation;

Abstract
An accurate and efficient calculation method for the thermoelastic field of thermal barrier
coating structures has a great contribution to their accurate design and quality promotion. However,
because of the thin coating thickness and large stress gradient in the coatings, it is too hard to
guarantee the calculation accuracy especially on the interface by using the traditional commercial
software. With this motivation, an accurate method for 3D thermoelastic field of thermal barrier
coating structure under arbitrary heat loads is proposed in this paper. The numerical results reveal
that the method has good convergence, high accuracy, and high stability. Based on this method, the
effects of material parameters and coating thickness on the stress field are investigated to provide a

2
refined guideline for the selection of coating materials. Furthermore, the optimal thickness of coating
can be designed accurately by considering the interfacial failure criterion. Numerical results also
reveal that the appropriate coating materials and coating thickness can largely reduce coating and
interface stress. These studies indicate that the highly precise calculation method is useful for
extremely fine structure design, parameter analysis and failure evaluation of the thermal barrier
coating structures and enhancing their quality.

Keywords: thermal barrier coatings; thermal stress; heat source; transversely isotropic; Green’s
function

1. Introduction

The improvement of the service life of various structural elements is one of the most important
tasks to be addressed by the modern industry. As an alternative, various types of protective coatings
formed with complex functionally-graded or layered structures are used for these purposes [1]–[3]
For example, hundreds of coatings are used to protect a variety of engineering structural materials
from corrosion, wear, and erosion and to provide lubrication or thermal insulation[4], [5]. Typically,
thermal barrier coating (TBC) materials have important applications in the protection of the
engineering components operating in the high-temperature environments [6]–[9], e.g., the ceramic
liners of high thrust–weight ratio engines[10]. How to accurately calculate the stress field and choose
the appropriate coating material determine the service performance of the coating structure.
Materials with different thermal and elastic parameters may be selected for coatings that protect
structures (termed substrates) avoiding thermal load. The thermal stress caused by thermal load
during the coating preparation process cannot be ignored[11], [12]. The thermal stresses are directly
related with the structural strength of the coating and substrate[13], [14]. Due to the differences of
thermal and elastic parameters, thermal loads may cause disturbance in the heat flux. The thermal
expansion mismatch between the substrate and coating structures may cause thermal stresses[15].
The properties of coatings are also closely related to the elastic modulus mismatch between coating
and substrate [16]. Thus, the precise selection of the coating thickness and coating material
parameters to simultaneously ensure adequate coating strength and production economy is important
during the design stage of the coating/substrate structure[17]. Apart from the traditional structural
strength problem, interfacial failure is a critical drawback for these layered structures[10], [18].
Owing to the complex thermal stress state between the coating and substrate, the interfacial coupling
field of a TBC’s structure would result in the germination of interfacial cracks. The expansion of
interfacial cracks would result in the delamination failure of the thermal barrier coating structure[14],

3
[19]–[21]. In addition, because the types of heat sources in actual engineering problems vary, the
relationship between interfacial stress and the different heat source profiles in coated structures needs
to be determined[22]. Thus, a precise analysis of the thermal stress characteristics of the coating and
substrate is important for structural design and performance evaluation. Thermal stresses
(particularly for interface thermal stress) play an important role in the design and service lifetime of
coating–substrate systems. The method for calculating the thermal coupling field under a thermal
load is the necessary foundation for an accurate analysis and effective design of various TBCs or
other coating structures.
Over the past several decades, many studies have been performed to investigate the mechanical
behavior of coating–substrate systems. Owing to the complexity of the coating structures, studies on
the mechanical characteristics using the experimental approach were conducted. These studies
yielded a variety of useful mechanical parameters[23]–[25]. However, the effects of the parameters
(which include coating thickness, coating thermal conductivity parameters, and mechanical
parameters) on the thermoelastic response of coating structures are diverse. It is difficult to quickly
determine the influence of certain key parameters on the thermal mechanical behaviors using
experimental research. This is because of the practical constraints on the controllability of certain
experimental parameters. In addition, due to the thin coating, especially the coating structure under
thermal environment, it is very difficult to achieve in-situ measurement. In-situ stress measurement
of coating structure in thermal–mechanical coupled high temperature environment needs to be
studied. Apart from experimental methods, certain completely numerical methods have been
introduced. As a classical approach, the finite element method (FEM) is widely used to analyze
coating structures[26]–[30]. Substantial differences in thermo-mechanical material properties
between the coating and substrate materials are necessary to provide effective thermal protection.
These differences may induce abrupt variations in the thermal and mechanical properties of coating
structures across the interfaces[31]. The gradient of the coupling stress field is exceptionally high
[32], [33]. Similarly, it is difficult to prevent the interfacial stress concentration and high gradient
stress because of the material mismatch and the thin coating thickness[34], [35]. To obtain accurate
solutions for the stress field near the heat load or interface using the FEM, a high-density grid is
required near the interface to capture the high gradient stress fields[36], [37]. On the one hand, the
use of a high-density grid may result in the consumption of a substantial amount of time. On the
other hand, it generally yields ill-conditioned equations. Thus, it is challenging to obtain accurate
solutions for coating–substrate structures which are important to evaluate coating crack and other
problems by traditional FEM. In addition, problems like those discussed above may be encountered
while determining the exact solution for the coupling field, because most coatings are very thin. The
4
boundary element method (BEM) is another method for analyzing coating–substrate
systems[38]–[41]. Unlike the FEM, the BEM requires meshes only on the boundaries of the
calculation region. The field results can be obtained by integrating along the boundary. However, the
calculation around certain singular points by using BEM may encounter certain difficulties owing to
singularity integration[42]. In addition, the corresponding Green’s functions must be obtained before
using the BEM[43]. Certain scholars studied thermoelastic problems using analytical solutions,
which differ from the completely numerical method mentioned above. This approach is convenient
for determining the mechanism underlying thermoelastic responses. The most popular methods are
the transfer matrix method and integral transform method, which are effective for solving for layered
structures. Fourier transforms are used to obtain solutions for the temperature increment and stress
fields of transversely isotropic substrate–coating systems[44]. Liu.et studied the 3D contact problem
of coating structures by Hank’s integral transformation method. Lin used the transfer matrix method
to analyze layered transversely isotropic media[45]. It is challenging to obtain accurate solutions for
coating structures using this method. The small layer thickness also adversely affects accuracy. Apart
from the analytical method for coated structures discussed above, the general solution method is a
potential approach to studying coating–substrate systems, which are ordinarily expressed in the form
of general functions[46]. Green’s function of exponentially graded transversely isotropic
substrate–coating system is presented[47]. With regard to infinite layer systems, Hou et al. have
conducted certain work on Green’s function of thermal elastic structures based on the general
solutions[48], [49]. Jiang[50] and Tong[51] obtained the 3D and 2D coating isotropic thermoelastic
analytical forms, respectively, under a point heat source on a free surface. However, thermal elastic
analysis of a transversely isotropic substrate–coating structure has not been performed. Once this
problem is solved, it will be helpful to analyze the thermodynamic behavior of coated structures such
as TBCs.
Thermal barrier coatings have a ceramic-metal configuration and may not be isotropic.
Plasma-sprayed coatings have exhibited transversely isotropic symmetry[52], [53]. In this study, a
highly efficient and accurate 3D calculation method for transversely isotropic coating structures
especially for thermal barrier coating structure is proposed. In addition, the thermoelastic analysis of
coating structures is conducted to provide certain design guidelines for coating materials. First,
compact potential functions in a harmonic form with unknown constants are constructed based on the
theory of differential operators. Finally, the unknown constants are obtained by considering the
equilibrium, boundary conditions, and compatibility conditions at the interface using Matlab
software platform. The analytical fundamental solutions expressed as harmonic functions under point
heat source are determined based on the general solution of 3D transversely thermoelastic structures,
5
which is named Green’s function. Furthermore, the Green’s solutions for an arbitrary heat source can
also be obtained based on the superposition principle and using the numerical integration method. In
the description of the numerical simulations, the convergence and computational speed are discussed
first. The effects of the thermal and mechanical parameters on thermoelasticity are analyzed. Then,
the stress fields of coating and substrate structures under distributed heat source are analyzed. Finally,
the interfacial failure problem is discussed based on the above method, and the optimal thickness of
the coating material is evaluated based on the interfacial failure criterion. The method proposed in
this study presents an accurate and efficient tool for designing and optimizing coating–substrate
systems, which have a substantial engineering application value.
The remainder of this paper is structured as follows. For completeness, the 3D governing
equations and the general solution for a 3D transversely thermoelastic material are presented in
Section 2. In Section 3, the potential functions with undetermined constants are constructed, and
Green’s function representing the elastic field and temperature increment field of a coating–substrate
structure (thermal barrier coating structure) under a point heat source is derived. The solutions for
arbitrary heat sources are also determined in this section. In Section 4, the effects of the coating
thermal parameters, coating elastic modulus, and coating thickness on the interface stress field with a
point heat source are presented. In Section 5, the thickness of coating material is optimized based on
the interfacial failure criterion. Finally, certain concluding comments are presented in Section 6.
Some notations in this paper are shown in Table 1.
Table 1 The definitions of the notations

Notations Definition
r, , z Cylindrical coordinates  r,, z 
 j , j  j  1, 2,3 Potential functions satisfying the Laplace equations
cij , ij  i, j  1, 2,3, 4,5,6  Elastic moduli and thermal moduli of the material
Temperature increment for the coating  0  z  h  and the substrate  z  h  ,
 , 
respectively
Displacement components of the coating  0  z  h  and the substrate
um , um  m  r,  , z 
 z  h  in cylindrical coordinates  r,, z  , respectively
Stress components of the coating  0  z  h  in cylindrical
 m , mn  m, n  r, , z 
coordinates  r,, z 
  m, n  r, , z 
 m , mn Stress components of the substrate  z  h  in cylindrical coordinates  r,, z 
h Thickness of the coating
H Intensity of the heat source
Elastic moduli of the coating  0  z  h  and the substrate  z  h  ,
cij , cij i, j  1, 2,3, 4,5,6
respectively
6
Thermal moduli of the coating  0  z  h  and the substrate  z  h  ,
ij , ij i, j  1, 2,3, 4,5,6
respectively
Thermal conduction coefficients the coating  0  z  h  in cylindrical
11 , 33
coordinates  r,, z 
Thermal conduction coefficients the substrate  z  h  in cylindrical
11 , 33
coordinates  r,, z 
Eigenvalues of the coating material and substrate material which are the
s j , sj  j  1, 2
roots whose real part are positive of Eq.(6), respectively
s3 Material characteristic constants s3  11 33 for the coating  0  z  h 
s3 Material characteristic constants s3  11 33 for the coating  z  h 
n Number of iterations: n  1, 2,3 

h0 Length characteristic scales


 , , Non-dimensional cylindrical coordinates:   r h0 ,    ,   z h0
 , , Non-dimensional Cartesian coordinates:   x h0 ,   y h0 ,   z h0
 k , kl (k , l   ,  ,  ) Non-dimensional stress components
 Non-dimensional temperature increment
 Non-dimensional thickness of the coating  = h h0
Dnjkl ,Dnjkl , Dnjkl , Dnjkl Undetermined constants
 Non-dimensional point heat source
E0 , Es Equivalent elastic modulus for substrate and coating, respectively
nc Required mirror order to achieve convergence
2. General solution for transversely isotopic thermoelastic material
When the coating materials are deposited on a substrate, they often form polycrystalline films
and coatings which are transversely isotropic, i.e. the thermal and mechanical properties are isotropic
in the coating plane but differ from those in the direction normal to the plane [43]. The general
solution for transversely isotopic thermoelastic material is introduced in the cylindrical
coordinates ( r ,  , z ) , where the r plane is parallel to the plane of isotropy. The constitutive
relationships of a transversely isotropic thermoelastic material are

7
ur   u 1 u
u z  1 ur u u 
 r  c11  c12  r 
  c13  11 ,  r  c66    ,
r   r r 
z  r  r r 
u  u 1 u  u z  u 1 u z 
   c12 r  c11  r    c13  11 ,   z  c44    , (1)
r  r r   z  z r  
 u u 1 u  u z  u u 
 z  c13  r  r    c33  33 ,  zr  c44  z  r  ,
 r r r   z  r z 

where um (m  r ,  , z ) are the components of the displacement,  m and  mn (m, n  r ,  , z ) are

the components of the normal and shear stresses, respectively,  is the temperature increment, cij

and ii (i, j  1, 2, , 6) are the elastic moduli and thermal moduli, respectively. The relationship

c66  (c11  c12 ) / 2 holds for materials with transverse isotropy.


In the absence of body forces, the mechanical equilibrium equation and heat equilibrium
equation are
 r 1  r  zr  r   
   0,
r r  z r
 r 1     z 2 r
    0,
r r  z r

 z r 1   z   z  z r
   0, (2a)
r r   z r

  2   2   2
11  2     33 2  0 , (2b)
 r rr r 2 2   z

where 11 and  33 are the coefficients of heat conduction.


The general solution of Eqs. (1) and (2) for a spatial axisymmetric problem can be expressed by
three potential functions as [53, 54]
3  j 3  j  2 3
ur   , u  0 , u z   s j k1 j ,   k23 , (3a)
j 1 r j 1  zj z32

1  j 3 2   j
3 2

 r  2c66    s j j ,
j 1 r r j 1 z 2j

1  j 3 2 j
3 2

   2c66    ( s j  j  2c66 ) 2 ,
j 1 r r j 1 z j
3  2 j 3  2 j
 z   j ,  r    z  0 ,  zr   s j j , (3b)
j 1 z 2j j 1 rz j

where potential functions  j ( j  1, 2,3) satisfy the following Laplace equations


8
 2  2 
 2   2  j  0 , ( j  1, 2,3) , (4)
 r r r z j 

and
z j  s j z , ( j  1, 2,3) , (5)

where s3  11 / 33 , s j ( j  1, 2) are two roots, whose real parts are larger than zero, of the

following fourth-degree polynomial equation:


as 4  bs 2  c  0 , (6)
where,

a  c33c44 , b  c11c33  c44   c13  c44  , c  c11c44 .


2 2
(7)

In addition,
 j  c44 (1  k1 j )  c13  c33s 2j k1 j  33k2 j  (c11  c13s 2j k1 j  11k2 j ) / s 2j ,

a2  b2 s 2j as34  bs32  c
k21  k22  0 , k1 j   , k23  ,
a1  b1s 2j a1  b1s32

a1  11c44 , b1  33 (c13  c44 )  11c33 ,

a2  33c11  11 (c13  c44 ) , b2  33c44 . (8)


It should be noted that the general solution given in Eq. (3) is valid for the most common case
when the three eigenvalues s j ( j  1, 2,3) are distinct. The thermoelastic field can be obtained based

on this general solution once the potential functions are determined. It will be found that the
construction and determination of the potential functions are the most important and difficult step in
search for the thermoelastic field.

3. The Green’s function of thermal barrier coating structure

In this section, the Green’s function for a point heat source on the surface of thermal barrier
coating structure is derived based on above general solution. As shown in Fig.1, the coating
0  z  h is perfectly bonded to the substrate z  0 . A point heat source H is applied at the point
(0,  , h) on the free and thermally insolated surface z  h of a thermal barrier coating structure.

9
Fig. 1 Thermal barrier coating structure geometrical model of single layered coating system without
imperfections which is applied by heat source H on the coating surface z  h . Cartesian coordinates
oxyz and cylindrical coordinates or z are introduced to describe this problem.

This is a spatial axisymmetric problem. It is addressed as follows: the primed quantities refer to
the variables in the substrate z  0 , and the un-primed quantities refer to those in the coating
0  z  h.
The boundary conditions for the free and thermally insolated surface z  h are in form of

 z (r , h)  0 ,  zr (r , h)  0 , ( r , h)  0 . (9)
z
Considering the mechanical and thermal equilibrium of the layer   z  h (0    h) (Fig.
1), the following equilibrium equations should be satisfied:


  (r,  )rdr  0 ,
0
z
(10a)



2 33  (r ,  )rdr  H . (10b)
0
z

Consider the case where the thermal barrier coating structure has no initial defects, i.e., the
coating 0  z  h and substrate z  0 are perfectly bonded. The compatibility conditions at the
interface z  0 should be satisfied as follows:
ur (r , 0)  ur (r , 0) , uz (r ,0)  uz (r ,0) ,

 z (r ,0)   z (r ,0) ,  zr (r , 0)   zr (r , 0) ,


  
 ( r,0)   ( r,0) , 33 (r , 0)  33 (r , 0) , (11)
z z

10
where  33 and  33 are the coefficients of heat conduction of the coating 0  z  h and substrate

z  0 , respectively.
After the boundary conditions, equilibrium equations, and compatibility conditions are
illustrated, the potential functions will be derived below. First, the following quantities are
introduced to simplify the notations:
z j  s j z , z j  sj z , h j  s j h ,

hnkl  (2n  k )h1  (k  l )h2  (l  1)h3 ,

znjkl  z j  hnkl , Rnjkl  r 2  znjkl


2 *
, Rnjkl  Rnjkl  znjkl ,

znjkl  z j  hnkl , Rnjkl  r 2  znjkl


2 *
, Rnjkl  Rnjkl  znjkl ,

  z j  hnkl , Rnjkl
znjkl   r 2  znjkl *  Rnjkl
2 , Rnjkl   znjkl
 ,

( n  1, 2, ,  ; j  1, 2,3 ; k  1, 2, , 2n ; l  1, 2, , k ). (12)

where s j and s j are the eigenvalues of the materials in the coating 0  z  h and substrate z  0 ,

respectively. These can be obtained by using Eq. (6).


The harmonic potential functions for the coating 0  z  h can be assumed using the mirror
image method as follows:

 j   ( nj  nj ) , ( j  1, 2,3) , (13)
n 1

where
2n k
 nj   D njkl ( znjkl ln Rnjkl
*
 Rnjkl ),
k 1 l 1

2n k
 nj   D njkl ( znjkl ln Rnjkl
*
 Rnjkl ),  n  1, 2, , , j  1, 2,3 , (14)
k 1 l 1

* *
where D njkl and D njkl are coefficients to be determined. Rnjkl , Rnjkl , Rnjkl , and Rnjkl are defined

in Eq. (12).
The corresponding thermoelastic field in coating can be determined by substituting Eqs. (13)
and (14) into the general solution (Eq. (3)) as follows:
   
ur   (urn  urn ), u z   (u zn  u zn ),    ( n   n ), u z   (u zn  u zn ),
n 1 n 1 n 1 n 1

   
 r   ( rn   rn ),     (  n    n ),  z   ( zn   zn ), zr   ( zrn   zrn ), (15)
n 1 n 1 n 1 n 1

where

11
3 2n k 3 2n k
r r
urn   D njkl *
, u rn    D njkl * ,
j 1 k 1 l 1 Rnjkl j 1 k 1 l 1 Rnjkl
3 2n k 3 2n k
uzn   s j k jDnjkl ln Rnjkl
*
, uzn   s j k jDnjkl ln Rnjkl
*
,
j 1 k 1 l 1 j 1 k 1 l 1

2n k 2n k
1 1
n  k23  Dn3kl , n  k23  Dn3kl ,
k 1 l 1 Rn3kl k 1 l 1 Rn3kl
3 2n k 3 2n k
1 1
 rn  2c66  Dnjkl *
  s 2j  j Dnjkl ,
j 1 k 1 l 1 Rnjkl j 1 k 1 l 1 Rnjkl
3 2n k 3 2n k
1 1
 rn  2c66  Dnjkl *
  s 2j  j Dnjkl ,
j 1 k 1 l 1 Rnjkl j 1 k 1 l 1 Rnjkl
3 2n k 3 2n k
1 1
  n  2c66  Dnjkl *
  ( s 2j  j  2c66 )Dnjkl ,
j 1 k 1 l 1 Rnjkl j 1 k 1 l 1 Rnjkl
3 2n k 3 2n k
1 1
  n  2c66  Dnjkl *
  ( s 2j  j  2c66 )Dnjkl ,
j 1 k 1 l 1 Rnjkl j 1 k 1 l 1 Rnjkl
3 2n k 3 2n k
1 1
 zn    jDnjkl ,  zn    j D njkl ,
j 1 k 1 l 1 Rnjkl j 1 k 1 l 1 Rnjkl
3 2n k 3 2n k
r r
 zrn   s j jDnjkl *
,  zrn   s j j Dnjkl *
,
j 1 k 1 l 1 Rnjkl Rnjkl j 1 k 1 l 1 Rnjkl Rnjkl

( n  1, 2, ,  ). (16)
The harmonic potential functions for the substrate can be assumed as follows:

 j   nj , ( j  1, 2,3), (17)
n 1

where
2n k
 nj   D njkl
 ( znjkl
 ln Rnjkl  ),  n  1, 2,
*  Rnjkl , , j  1, 2,3  , (18)
k 1 l 1

where D njkl * and Rnjkl


 are coefficients to be determined. Rnjkl  are defined in Eq. (12).

The corresponding thermoelastic field in substrate can be obtained by substituting Eqs. (17) and
(18) into the general solution Eq. (3) yields
  
ur   urn , u z   u zn ,      n ,
n 1 n 1 n 1

   
 r    rn ,      n ,  z    zn ,  zr    zrn
 , (19)
n 1 n 1 n 1 n 1

where
12
3 2n k 3 2n k
r
urn   D njkl
 , u    sj k jDnjkl
 ln Rnjkl
* ,
j 1 k 1 l 1 *
Rnjkl
zn
j 1 k 1 l 1

2n k
1
  Dn3kl
n  k23 ,
k 1 l 1 Rn3kl
3 2n k 3 2n k
1 1
  Dnjkl
 rn  2c66    sj 2 j Dnjkl
 ,
j 1 k 1 l 1 *
Rnjkl j 1 k 1 l 1 
Rnjkl
3 2n k 3 2n k
1 1
  D njkl
 n  2c66    ( sj 2 j  2c66
 )D njkl
 ,
j 1 k 1 l 1 *
Rnjkl j 1 k 1 l 1 
Rnjkl
3 2n k 3 2n k
1 r
 zn    jDnjk
    sj j Dnjkl
,  zrn  ,
j 1 k 1 l 1 
Rnjkl j 1 k 1 l 1  Rnjkl
Rnjkl *

( n  1, 2, ,  ). (20)
Considering the form of Eq. (15), the boundary conditions in Eq. (9) for the free surface z  h
can be separated as follows:
1
 z1 (r ,h ) ,0  zr1 (r , h)  0 , ( r , h)  0 , (21a)
z
 zn (r , h)   z ( n 1) (r , h)  0 ,

 zrn (r , h)   zr ( n 1) (r , h)  0 ,

 n 
(r , h)  n 1 (r , h)  0 , ( n  1, 2, ,  ). (21b)
z z
Considering the forms of Eqs. (15) and (19), the compatibility conditions in Eq. (11) for the
interface z  0 can be rewritten as
urn (r ,0)  urn (r ,0)  urn (r,0) , uzn (r ,0)  uzn (r ,0)  uzn (r ,0) ,

 zn (r,0)   zn (r,0)   zn (r,0) ,  zrn (r , 0)   zrn (r , 0)   zrn


 (r , 0) ,

     
 n (r , 0)   n (r , 0)   n (r , 0) , 33  n (r , 0)  n (r, 0)   33 n (r, 0) ,
 z z  z

( n  1, 2, ,  ). (22)
The substitution of Eq. (16) into Eq. (21a) yields
D1121  D1122  D1211 = D1222 = D1311 = D1321 = 0 , (23a)

1D1111  2D1221  3D1322  0 , (23b)

s11D1111  s22D1221  s33D1322  0 . (23c)

Considering z njkl and z njkl in Eq. (12), the following identities can be obtained:

13
z( n 1)1(2 n  2)  (h)  h1  (2n  2   )h2  (  1)h3 ,

z( n 1)2  (h)  (2n  2   )h1  h2  (  1)h3 ,

z( n 1)31 (h)  (2n  2   )h1  (  1)h2  h3 , ( n  1, 2, ,  ;   1, 2, , 2n  2 ), (24a)

zn111 (h)   z( n 1)111 (h)   z( n 1)221 (h)   z( n 1)322 (h)  2nh1 ,

zn 2(2 n )1 (h)   z( n 1)1(2 n 1)1 (h)   z( n 1)2(2 n  2)1 (h)   z( n 1)3(2 n 2)2 (h)  2nh2 ,

zn 3(2 n )(2 n ) (h)   z( n 1)1(2 n 1)(2 n 1) (h)   z( n 1)2(2 n  2)(2 n 1) (h)   z( n 1)3(2 n  2)(2 n  2) (h)  2nh3 ,

zn1( m1)1 (h)  zn 2 m1 (h)   z( n 1)1( m 1)1 (h)   z( n 1)2( m 2)1 (h)   z( n 1)3( m2)2 (h)  (2n  m)h1  mh2 ,

zn1( m1)( m1) (h)  zn3mm (h)   z( n 1)1( m 1)( m 1) (h)   z( n 1)2( m  2)( m 1) (h)   z( n 1)3( m 2)( m 2) (h)

 (2n  m)h1  mh3 ,

zn 2(2 n )( m1) (h)  zn 3(2 n ) m (h)   z( n 1)1(2 n 1)( m 1) (h)   z( n 1)2(2 n  2)( m1) (h)   z( n 1)3(2 n  2)( m 2) (h)
 (2n  m)h2  mh3 ,

zn1( 1)(  1) (h)  zn 2 (  1) (h)  zn3 (h)   z( n 1)1( 1)(  1) (h)   z( n 1)2( 2)(  1) (h)

  z( n 1)3(  2)(   2) (h)  (2n   )h1  (   )h2   h3 ,

( n  1, 2, ,  ; m  1, 2, , 2n  1 ;   2,3, , 2n  1 ;   1, 2, ,   1 ), (24b)
The substitution of Eq. (16) into Eq. (21b) using the above identities yields
D( n 1)1(2 n  2)   D( n 1)2  D( n 1)31  0 , ( n  1, 2, ,  ;   1, 2, , 2n  2 ), (25a)

1D( n 1)111  2D( n 1)221  3D( n 1)322  1D n111 ,

1D( n 1)1(2 n 1)1  2D( n 1)2(2 n 2)1  3D( n 1)3(2 n 2)2  2D n 2(2 n)1 ,

1D( n 1)1(2 n 1)(2 n 1)  2D( n 1)2(2 n  2)(2 n 1)  3D( n 1)3(2 n 2)(2 n 2)  3D n 3(2 n )(2 n ) ,

1D( n 1)1( m1)1  2D( n 1)2( m  2)1  3D( n 1)3( m 2)2  1D n1( m 1)1  2D n 2 m1

1D( n1)1( m1)( m1)  2D( n1)2( m 2)( m1)  3D( n1)3( m 2)( m2)  1D n1( m1)( m1)  3D n 3mm

1D( n1)1(2 n1)( m1)  2D( n1)2(2 n 2)( m1)  3D( n1)3(2 n2)( m2)  2D n 2(2 n )( m1)  3Dn 3(2 n ) m ,

1D( n 1)1( 1)(  1)  2D( n 1)2( 2)(  1)  3D( n1)3( 2)(  2)  1D n1( 1)(  1)  2D n 2 (  1)  3D n3 ,

( n  1, 2, ,  ; m  1, 2, , 2n  1 ;   2,3, , 2n  1 ;   1, 2, ,   1 ), (25b)

s11D( n 1)111  s22D( n 1)221  s33D( n 1)322   s11D n111 ,

s11D( n 1)1(2 n 1)1  s22D( n 1)2(2 n  2)1  s33D( n 1)3(2 n 2)2   s22D n 2(2 n)1 ,

s11D( n 1)1(2 n 1)(2 n 1)  s22D( n 1)2(2 n 2)(2 n 1)  s33D( n 1)3(2 n 2)(2 n  2)   s33D n 3(2 n )(2 n ) ,

14
s11D( n 1)1( m1)1  s22D( n 1)2( m 2)1  s33D( n 1)3( m2)2   s11D n1( m1)1  s22D n 2 m1 ,

s11D( n 1)1( m1)( m1)  s22D( n 1)2( m 2)( m1)  s33D( n 1)3( m 2)( m 2)   s11D n1( m 1)( m 1)  s33D n 3mm ,

s11D( n 1)1(2 n 1)( m1)  s22D( n 1)2(2 n 2)( m1)  s33D( n 1)3(2 n 2)( m2)   s22D n 2(2 n )( m1)  s33D n 3(2 n ) m ,

s11D( n1)1( 1)(  1)  s22D( n1)2(  2)(  1)  s33D( n1)3(  2)(   2)
,
 s11Dn1( 1)(  1)  s22Dn 2 (  1)  s33Dn3

( n  1, 2, ,  ; m  1, 2, , 2n  1 ;   2,3, , 2n  1 ;   1, 2, ,   1 ), (25c)

D ( n 1)3 2  0 , D ( n 1)3( k  2)( l  2)  D n 3kl ,

( n  1, 2, ,  ;   2, , 2n  2 ; k  1, 2, , 2n ; l  1, 2, , k ). (25d)
By substituting Eqs. (16) and (20) into Eq. (22) using the identities of
 (0)  hnkl ,
znjkl (0)   znjkl (0)  znjkl

 (r,0)  r 2  hnkl
Rnjkl (r,0)  Rnjkl (r,0)  Rnjkl 2
,

*
Rnjkl (r,0)  Rnjkl
*
* (r,0)  r 2  hnkl
(r,0)  Rnjkl 2
 hnkl ,

( n  1, 2, ,  ; j  1, 2,3 ; k  1, 2, , 2n ; l  1, 2, , k ), (26)
one can obtain

 D njkl  D njkl    D njkl


3 3
 ,
j 1 j 1

  s k     s j k1 jDnjkl ,
3 3

j 1j Dnjkl  sj k1jDnjkl


j 1 j 1

  D njkl   j Dnjkl     j Dnjkl


3 3

j
  ,
j 1 j 1

  s j jDnjkl  sjjDnjkl
    s j jDnjkl ,
3 3

j 1 j 1

 Dn3kl  k23Dn3kl ,
k23Dn3kl  k23

k23 s3 33Dn3kl  k23


 s3 33 Dn3kl  k23 s3 33Dn3kl ,

( n  1, 2, ,  ; k  1, 2, , 2n ; l  1, 2, , k ). (27)
The mechanical equilibrium equation Eq. (10a) is automatically satisfied when Eqs. (15) and
(16) are substituted into it. By substituting Eqs. (15) and (16) into the thermal equilibrium equation
Eq. (10b) using the following integral:
zn 3kl zn 3kl zn 3kl zn 3kl
R 3
n 3 kl
rdr  
Rn 3kl
, R 3
n 3 kl
rdr  
Rn 3kl
, (28)

15
One can obtain
r 
 2n k
 zn3kl ( ) z ( )  H
  Dn3kl
n 1 k 1 l 1  Rn3kl (r ,  )
 Dn3kl n3kl 
Rn3kl (r ,  )  r 0

2 k23 1133
. (29)

Using the following limits:


zn 3kl ( ) z ( ) z ( ) z ( )
lim  lim n 3kl  0 , lim n 3kl  1 , lim n 3kl 1,
n 3 kl ( r ,  ) n 3 kl ( r ,  ) n 3 kl ( r ,  ) n 3 kl ( r ,  )
r  R r  R r 0 R r 0 R

( n  1, 2, ,  ; k  1, 2, , 2n ; l  1, 2, , k ), (30)
Eq. (29) can be transformed into
 2n k
H
 D
n 1 k 1 l 1
n 3 kl  D n 3kl   
2 k23 1133
. (31)

Using Eqs. (23a), (25a) and (25d) and the following identity,
 2n k 2 k  2 n 2  2 n 2

 D
n 1 k 1 l 1
n 3kl  Dn3kl    D13kl    D( n1)31    D( n 1)3 2
k 1 l 1 n 1  1 n 1   2


   D ( n 1)3( k  2)(l  2)  D n 3kl  ,
2n k
(32)
n 1 k 1 l 1

Eq. (31) can be simplified as


H
D1322   . (33)
2 k23 1133

 for the different order n can be determined by


The coefficients D njkl , D njkl , and D njkl

following recursive procedure. The unknown coefficients D1 jkl ( j  1, 2,3; k  1, 2; l  1, k ) can be

determined by nine equations of Eqs. (23) and (33). The coefficients D1 jkl and D1 jkl

( j  1, 2,3; k  1, 2; l  1, k ) can be determined by Eq. (27). D 2 jkl ( j  1, 2,3; k  1, 2,3, 4; l  1, 2, k)

can be determined by Eq. (25). Then, for an arbitrary n , D njkl and

 ( j  1, 2,3; k  1, 2,3, 4; l  1, 2,
D njkl k ) can be determined by the recursive equations in Eqs. (25)

and (27). In addition, D njkl ( j  1, 2,3; k  1, 2, , 2n; l  1, 2, k ) can be determined by the recursive

equations in Eq. (25). The iterative calculations of undetermined coefficients are completed using
Matlab software platform. Once all the unknown coefficients are determined for an adequately large
n to satisfy convergence, the Green’s function solutions representing the thermoelastic field of Eqs.
(15) and (16) in the coating 0  z  h and thermoelastic field of Eqs. (19) and (20) in the substrate
z  0 can be obtained.
The Green’s function solutions derived above are also applicable to problems other than point

16
heat source problems. For an arbitrary heat sources in actual engineering, according to the
superposition, the corresponding thermoelastic field can be obtained by using the fundamental
solutions for a point heat source.

4. Study on thermal elastic field of coating structures

In this section, the foregoing analytical results will be presented numerically. The computational
convergence of the theoretical solutions and the material mismatch assessment is verified in Section
4.1. Based on the analytical solutions, the three dimensional contours of the stress fields for thermal
barrier coating structure under different loads including distributed heat sources are obtained in
Section 4.2. Specially, the influences of important coating physical quantities upon the coating
structure are discussed in Section 4.3 detailed. In chapter 4.3.1, the influences of the coating thermal
parameters on its stress distribution are analyzed, and then the guidelines for selecting and designing
coating property are presented. The influences of the elastic modulus on the stress distributions are
investigated in chapter 4.3.2. Finally, the influences of coating thickness on stresses are analyzed,
and the suitable thickness is recommended in chapter 4.3.3.
For the purpose, some coating structure materials are used in the present numerical calculations
whose properties are shown in Table 2. The coating thickness h0 , elastic modulus c33 , thermal

expansion coefficient  r , and heat conduction coefficient  33 are selected as the characteristic
parameters. The following calculations are normalized. And the non-dimensional components are
followings.
In the cylindrical coordinates  r, , z  ,
r z h
= ,    ,   ,    r ,   ,
h0 h0 h0
(34)
m  mn
k  ,  kl   m, n  r ,  , z ; k , l   ,  ,   .
c33 c33

In the Cartesian coordinates  x, y, z  ,


x y z  
= ,   ,   ,  k  m , kl  mn  m, n  x, y, z  . (35)
h0 h0 h0 c33 c33
In this case, Eq. (33) is in the following dimensionless form,

D1322   , (36)
2 s3 k23

where  is a non-dimensional point heat source as follows,

17
Hr
 . (37)
h0  33

Here let  =1 act at the point (0, 0,   1) , which is corresponding to the point heat source
h0  33
H .
r
Table 2 Material properties of the transversely isotropic thermoelastic material

Material property Hexagonal Zn ZnO


c11 162.8 209.7
c12 50.8 121.1
Elastic constants
(× 109 Nm-2)
c13 36.2 105.1
c33 62.7 210.9
c44 38.5 42.5
11 124 30
Heat conduction coefficients (WK-1m-1)
33 124 30
r 5.818 29.98
Thermal expansion coefficient (10-6K-1)
z 15.35 29.98
Thermal moduli 11  (c11  c12 ) r  c13 z 17.9839 130.6828
(105 NK-1m-2) 33  2c13 r  c33 z 13.8367 126.2458

4.1 Verification of the analytical solutions

Example 1: In order to validate the method and theory presented in this study, the coating
structure problem is firstly degenerated to a semi-infinite structure under a point heat load. In this
case, all parameters and conditions are taken according to the Ref.[54]. The coating and the substrate
are the same material which is Hexagonal Zn in Table 2. Comparisons of stress on section  =0 for
semi-infinite structure are depicted by two Green’s function solutions as shown in Fig. 2. One can
find that the results of the present solution (PS) agree well with those of the previous solutions [54].
It is obvious that the results of the two models are consistent which shows the correctness of the
harmonic function constructed in the present model. It is noteworthy that the coatings model can be
degraded to a semi-infinite body model.

18
Fig. 2 Comparison between the results of the present solutions (PS) and those for a semi-infinite
structure[54] within 0    3 and  =0 , (a) shear stress   102 and (b) axial stress   102 . The
materials of the structure are as follows: elastic constants 109 Nm 2  : c11  162.8; c12  50.8; c13  36.2;
c33  62.7; c44  38.5; the thermal parameters are as follows: 11 =124WK 1m 1 ;33 =124WK 1m 1 ;
 r  5.818 106 K 1 ;  z  15.35  106 K 1 ; 11 =17.9839 105 NK 1m 2 ; 33 =13.8367 105 NK 1m 2 .

Example 2: Based on the analytical solutions obtained, the comparative analysis with the
numerical results using finite element method (FEM) is obtained. Finite cylinder 0    50 and
50    1 are taken for axisymmetric FEM to calculate the thermoelastic field in the coating layer
0    1 and the semi-infinite substrate   0 . About eighty thousand quadrilateral isoparametric

elements are meshed. Fig. 3 shows the non-dimensional Green’s stress components   and   for

a ZnO/Zn coating–substrate structure on the interface   0 with two different calculation methods.

In the figure, n  k  k  1, 2,3,   represents the contribution value of the stress function taking

into account the mirror point of order k . It can be found that the differences between the stress
components of two adjacent n decrease dramatically with the increase of n . This implies that
these Green’s stress components converge rapidly. The shear stress   tends to be stable when

n  5 , whereas the axial stress   tends to be invariable when n  6 . This implies that the

convergence order n is different for the different stress components. When considering n=7 , all
stress components tend to be convergence and the results obtained by presented Green’s function
method and finite element method (FEM) are almost the same. When n=7 , the computer time using
the Green’s function method is 192.2s, while the computer time using FEM is almost 10 minutes for
two-dimensional model under the same computer configuration. This indicates that the analytical
method proposed in this study has high computational efficiency.

19
  102 and (b) axial stress   102 for
Fig. 3 The convergence of non-dimensional (a) shear stress
ZnO/Zn coatings on the interface   0 with different mirror order n in solution Eq. (15) under a
non-dimensional point heat source   1 . And the convergence results are in good agreement with the
results using FEM.

Example 3: In this example, it mainly explores the influence of different material combination
on convergence. Keep the parameters of material for substrate (Zn) constant whose equivalent elastic
modulus is denoted as E0 , where E0   c112 c33  2c11c132  c122 c33  2c12c132   c11c33  c132  . The coating’s

equivalent elastic modulus is denoted as Es , and other coating material’s parameters are kept

consistent with the substrate. The convergence of non-dimensional shear stress   102 on the

interface  =0 for different material combination is shown in Fig. 4. nc denotes the required

mirror order to achieve convergence. Fig. 4 shows that the convergence order nc is different for

different material combinations and variation trend of shear stress   on the interface   0 .

Generally speaking, the greater the differences of materials, the higher the required mirror order.
Meanwhile, it can be seen from the figure, the greater the difference of material parameters, the
greater the peak value of non-dimensional shear stress   102 on the interface  =0 . For thermal

barrier coatings, the thermal expansion coefficient of the coating is smaller than substrate to keep the
thermal protection function strong. Therefore, coating and substrate materials with similar elastic
molulus should be selected in the design of coating parameters, which is effective to reduce the
interface stress and improve the safety performance of the overall structure. A large number of
calculation results all show that the convergence rate is very fast when the difference of bonding
material parameters is within the engineering allowable range. In fact, when the material parameters
differ greatly, the analytical convergence is very slow and even divergence of numerical calculation
may occur in rare case.

20
Fig. 4 The convergence of non-dimensional shear stress   102 on the interface  =0 for different
material combination in solution Eq. (15) under a non-dimensional point heat source   1 .

4.2 Three dimensional stress fields analysis

Having derived the solutions, steady-state global thermal stress analyses are executed to study
the effect of various thermal sources on thermal barrier coating structure composing of ZnO and
Hexagonal Zn. The variations of temperature increment and thermal stress under two typical heat
source loads are discussed systematically. Thermal stress analyses are performed by using the
Green’s function solutions in Section. 3. The two heat source intensities of point heat source and
distributed heat source are guaranteed to be the same which are listed in Table 3. In order to illustrate
the thermoelastic fields of a coating–substrate structure, certain typical planes are selected to show
thermoelastic stress contours. Three coordinate planes passing the coordinates origin point
 0,0,0 are selected for analysis.
Table 3 Two types of dimensionless heat source loads
Case 1 Case 2
Load type Point heat source Asymmetric distributed heat source
H 0 ( , )  1   2
 0.2 2 
Q( , )  e

Distribution form

Loading area  =0,   0  2 + 2  1 1    0

21
4.2.1 Case 1: Point heat source

A non-dimensional point heat source H 0  ,    1 is loaded on the surface of the coating layer

whose non-dimensional coating thickness   1 . The corresponding contours of the temperature


increment  and stress components   ,   ,   ,   for a transversely isotropic thermoelastic

coating structure on the sections through the coordinates origin point  0,0,0 are plotted in Figs.

5–9.
The temperature increments contours are shown in Fig.4. As expected, the high temperature
increments are observed near the heat source, since it is subjected to the heat flux circumferentially.
As illustrated in Fig. 5, the temperature increment contours are continuous, but the gradients of these
contours are discontinuous on the interface z  0 . The temperature gradient discontinuity is mainly
caused by the difference of thermal conductivity between the coating and substrate. It is noteworthy
that most of the heat energy are bounded in the coating layer. This just makes more sense that ZnO is
a bad heat conductor which can provide effective thermal resistance.
Fig. 6 and Fig. 7 show that the shear stress   and axial stress   are continuous on the

interface, respectively. It is notable that the stress gradients are discontinuous on the interface, which
is caused by the material properties mismatch between the coating and substrate. As expected, the
high stress occurs at the interface near the heat source. These numerical results reveal that the
interface stress near the point heat source is in a tensile and shear stress state, which greatly increases
the risk of coating peeling and shear damage.
Fig. 8 and Fig. 9 show that the radial stress   and hoop stress   are discontinuous on the

interface. The magnitudes of the stresses decrease from the vicinity of the heat source to the
periphery, and the stress states transform from compressive stress to tensile stress. Also, on the
interface, the stresses are compressive at coating side and most area is in tensile stress state at
substrate side. The stress states of the hoop stress   and radial stress   are different on both

sides of the interface which may be a reason for the separation of coating from the substrate material
as a failure mechanism in thermal barrier coating system.
Based on above figures analysis, it can be revealed that all components of a coupled field have a
singular point at the loading position of the point heat source and tend to be zero at an infinite
distance. A comparison of all the stress components indicates that the stresses in the coating layer are
significantly higher than those in the substrate in Figs. 5-9. This reveals that the density of elastic
energy in the coating layer is also significantly higher than that in the substrate. It is obvious that a
high strength level of coating is required for a good coating structure.
22
Fig. 5 Temperature increment contours   10 on the section through the axis of symmetry under a
2

non-dimensional point heat source for Case 1, in which a non-dimensional thickness   1 is considered.

Fig. 6 Shear stress contours   102 on the section through the axis of symmetry under a
non-dimensional point heat source for Case 1, , in which a non-dimensional thickness   1 is considered.

23
Fig. 7 Axial stress contour   102 on the section through the axis of symmetry under a non-dimensional
point heat source for Case 1, in which a non-dimensional thickness   1 is considered.

Fig. 8 Radial stress contours   102 on the section through the axis of symmetry under a
non-dimensional point heat source for Case 1, in which a non-dimensional thickness   1 is considered.

24
Fig. 9 Hoop stress contour   102 on the section through the axis of symmetry under a non-dimensional
point heat source for Case 1, in which a non-dimensional thickness   1 is considered.

4.2.2 Case 2: Asymmetric distributed heat source


In this section, the solution for a coating material under an arbitrary distributed heat source is
obtained which is a major advantage of the analytical method proposed in this paper. For the
convenience of describing and addressing distributed heat sources, thermal stress analysis under
distributed heat load are discussed in a Cartesian coordinate system  x, y, z  . For arbitrary heat

sources in actual engineering problems, the stress fields can be obtained based on the superposition
principle by using the Green’s function solutions for a point heat source. In engineering, a tip heating
distribution form is generally modeled as an asymmetric heat source. For example, the front-end load
generated on aircraft by air friction can be simulated as an asymmetric load and rapid laser scanning
also produces asymmetric loads, etc. A dimensionless asymmetric distribution heat source described
in Table 3 is chosen on coating surface in the following form,
S ( , ) :  2   2  1. (38)
The intensity of the non-dimension asymmetric distributed heat source is
   0.2 2  1    0, 1    1 .
 
2

Q( , )  e (39)

Results obtained from the thermal stress analyses based on the Green’s function and
superposition principle, the temperature increment, shear stress and axial stress along the three axis

25
are discussed systematically. The overall contours of the temperature increment and stress
components are depicted in Figs. 10–14. As can be seen from the figures, the thermoelastic field in
coating system does not change much in comparison with those under a point heat source (in Case 1).
But the distribution trends of thermoelatic field are different near the heat source due to the different
distribution forms of thermal loads.
A contour-plot of temperature increment along the three coordinate planes is shown in Fig. 10.
As expected, the high temperature increment is at the center of the heat source. In comparison with
Fig. 10(a) and Fig. 10(b), it can be seen that the temperature has different distribution trends on the
  xz  and   yz  coordinate planes which is mainly affected by the distributed heat source

form. Fig. 10(c) displays the maximum of temperature increment on   xy  plane occurs below

the center of the heat source. These analyses show that the trend of temperature increment is closely
related to the form of heat source. The temperature increment is largest in the region near the heat
source, where the failure is easy to occur.
Fig. 11 depicts the distribution trend of shear stress   on three coordinate planes which

shows the shear stress value is large near the coating interface. From the figure, the shear stress
varies greatly along the thickness direction, especially in coating region. It can be inferred that the
interface is prone to shear delamination for thin coating. Fig. 12 shows the variation trend of axial
stress   in z direction. It can be obtained that the axial stress   is continuous at interface

z  0 , which is consistent with the continuity condition of the interface. The maximum axial stresses
occur near the interface which is tensile stress. Thus, this is the most probable area of tensile failure
which produces delamination cracks parallel to the interface or on the interface. Fig. 13 and Fig. 14
show the variation trend of axial stress in x and y directions respectively. It is obvious that both
stress are not continuous on the interface and they are different on both sides of the interface, which
could be a reason for the separation of coating from the based material as a failure mechanism in
thermal barrier coating system.
From Figs. 10-14 for thermoelastic fields, one can observe that the gradients of the all contour
line are not continuous on the interface. This is caused by the inconsistency in material properties
between the coating and matrix. It is apparent from the full-field stress contours that all the stresses
and temperature increment fields satisfy the free surface conditions and interface conditions, which
can effectively reflect the distribution law of each component in the coupling field. There is another
valuable phenomenon that the maximum shear stresses and normal stresses occur near the interface
which is the weak area in the coating structure. Thus it is very important to analyze the interface

26
stress accurately.

Fig. 10 Temperature increment   10 on the three different surfaces parallel to the coordinate planes
2

respectively through the position  0,0,0 under asymmetric heat load for Case 2: (a)   xy  plane; (b)
  yz  plane; (c)   xz  plane

Fig. 11 Shear stress   102 on the three different surfaces parallel to the coordinate planes respectively
through the position  0, 0, 0 under asymmetric heat load for Case 2: (a)   xy  plane; (b)   yz 
plane; (c)   xz  plane

27
Fig. 12 Axial stress   102 on the three different surfaces parallel to the coordinate planes respectively
through the position  0,0,0 under asymmetric heat load for Case 2: (a)   xy  plane; (b)   yz 
plane; (c)   xz  plane

Fig. 13 Radial stress   102 on the three different surfaces parallel to the coordinate planes respectively
through the position  0, 0, 0 under asymmetric heat load for Case 2: (a)   xy  plane; (b)   yz 
plane; (c)   xz  plane

28
Fig. 14 Radial stress   102 on the three different surfaces parallel to the coordinate planes respectively
through the position  0, 0, 0 under asymmetric heat load for Case 2: (a)   xy  plane; (b)   yz 
plane; (c)   xz  plane

4.3 Influences of coating parameters on thermal stress

For thermal barrier coatings, it is always easier for the first failure to occur at or near the
junction interface. Therefore, the strength of a bonding structure depends on the interface strength.
The key problem of coating structure is essentially the interface strength caused by bond interface.
Generally, the normal stress of vertical interface (called peel stress) and shear stress acting on
interface are usually used as the evaluation parameters of interface strength. Due to the mismatch of
thermal parameters and elastic moduli between the coating and the substrate, internal cracking or
interface cracking will occur when the coating is subjected to thermal or force load. Once the
interface cracking occurring, the coating will peel off and the substrate will be exposed to high
temperature environment. It is also found that the coating thickness has a great influence on the
mechanical properties of thermal barrier coating system, and seriously affects the damage evolution
law and failure mechanism of the coating system. In this chapter, numerical tests designed discuss
the influence of coating material parameters and coating thickness on the key stresses of the interface
under the action of thermal load systematically which can provide some guidance for the design of
coating system.

4.3.1 Influences of coating thermal parameters on thermal stress

The failure of thermal barrier coating systems is influenced by many factors including thermal
parameter mismatch [55]. The stresses caused by thermal parameter mismatch is the key factor that
causes spallation and delamination failure of entire coated structures[56], [57]. Meanwhile, thermal

29
expansion coefficient mismatch is the cause of high thermal stress, which plays an important role in
the failure of coatings. The thermal conductivity parameters mainly affect the distribution of
temperature increment gradient field, which could affect the stress field directly. A numerical
experiment is designed and implemented in this section to study the influence of thermal parameters
under point heat source on thermoelastic field.
In this set of analyses, various characteristics are varied, including thermal parameters as listed
in Table 4 and Table 5. Like any parametric analysis, while studying the influence of a given thermal
parameter, the other features are kept constant. Four types of coating material are employed and the
coating structure parameters of ZnO/Zn are used as a reference case. The variable heat conduction
coefficients and thermal expansion coefficients are half or two times of the reference coating
material. The designed material parameters are listed in Table 4 and Table 5.
Table 4 Designed heat conduction coefficients of coating layer
Material property Beta Beta1 Beta2 Beta3 Beta4

Heat conduction 11 30 3 15 60 124


coefficients (WK-1m-1) 33 30 3 15 60 124

Table 5 Designed thermal expansion coefficients of coating layer


Material property Alpha Alpha1 Alpha 2 Alpha 3 Alpha 4

Thermal expansion r 29.98 2.998 14.99 44.97 59.96


coefficients (10-6K-1) z 29.98 2.998 14.99 44.97 59.96

Results obtained from the thermoelastic field analyses, the temperature increment and key stress
components variations on the interface (  0) are discussed systematically. The variations of

temperature increment with  under different thermal parameters are shown in Fig. 15. It is
obvious that the thermal conduct coefficient is the primary factor causing the temperature increment
field in Fig. 15(a). The temperature increment on the interface decreases with the increase of the heat
conduction coefficient. This implies that the thermal barrier coating material with a low heat
conduction coefficient has superior heat resistance. This is consistent with the properties of thermal
resistance materials. Because this is a unidirectional coupling problem in this paper, so the
temperature increment is independent of the thermal expansion coefficient in Fig. 15(b).

30
Fig. 15 The temperature increment  on the interface (  0) along the  axis for different coating
thermal parameters; (a) the effect of heat conduction coefficients; (b) the effect of thermal expansion
coefficients.

Fig. 16 The shear stress   on the interface (  0) along the  axis for different coating thermal
parameters; (a) the effect of heat conduction coefficients; (b) the effect of thermal expansion coefficients.

Fig. 16 demonstrates the distributions of the non-dimensional shear stress   along the axial

direction at the interface  =0 for different thermal parameters. Fig. 16(a) indicates that the peak

value of   is positively correlated with the heat conduction coefficient, i.e. the shear stress would

be at a lower stress level when selecting a coating material with a smaller thermal conductivity. Fig.
16(b) illustrates the effect of thermal expansion coefficients on the shear stress at interface. From Fig.
16(b), the thermal expansion coefficients not only change the peak value, but also change the peak
position of the shear stress. It can be concluded that excessive heat conduction coefficient or thermal
expansion coefficient would make the interface shear stress at a higher stress level. For thermal
barrier coatings, the thermal conductivity determines the insulation effect and the thermal expansion
coefficient is the main cause of residual stress.

31
Fig. 17 The vertical stress   on the interface (  0) along the  axis for different coating thermal
parameters; (a) the effect of heat conduction coefficients; (b) the effect of thermal expansion coefficients.

Fig. 17 studies effect of thermal parameters of coating material on non-dimensional normal


stress   at the interface  =0 . Fig. 17(a) shows that the normal stress   is tension stress whose

peak values increase with the decrease of the coating thermal conductivity. This also indicates that
the coating failure is generally tensile failure when the coating thermal conductivity is much less
than substrate, which is consistent with the literature [58]. A comparison of Fig. 15(a) and Fig. 17(a)
also reveals that a high heat conduction coefficient causes a higher stress, whereas it could cause a
higher temperature increment. Fig. 17(b) displays the normal stress   is also tension stress whose

peak values increase with the increase of the coating thermal expansion coefficient. From the Fig.
17(b), the normal stress increases slowly when the thermal expansion coefficient of the coating
increases, when the thermal expansion coefficient decreases, the stress decreases rapidly.
These results help engineers gain a general understanding of the stress distribution on interfaces
before carrying out complex calculations or experimental studies. As illustrated in Figs. 14-16, it can
be concluded that the coating thermal properties exert significant influences on the thermal stress
field and temperature increment. Comparison of Fig. 16 and Fig. 17, one can observe the influence
trend of thermal conductivity on shear stress and normal stress is different. A lower thermal
conductivity is beneficial to reduce the risk of coating shear delamination, but it also increases the
risk of tensile delamination. The influence trend of thermal expansion coefficient on normal stress
and shear stress is basically the same. These results obviously imply that it is very important to select
suitable coating materials for different engineering conditions.

4.3.2 Influences of coating stiffness on thermal stress

In this section, the influences of coating stiffness on coating interface are discussed
systematically. For convenience, the elastic modulus of the substrate is expressed as being half and
1.5 times of ZnO, while the substrate is assumed to remain unchanged. A special solution has been
32
considered wherein the elastic moduli of the coating are identical to those of the substrate. The
material parameters are listed in Table 6. Since the elastic filed does not affect the temperature field,
the change of elastic constant does not cause the change of the temperature field. In this section,  

and   which are the main causes of interface failure are analyzed as typical cases.

Table 6 Designed elastic moduli of coating layer


Material property ZnO C11L C11S C12L C12S C13L
c11 209.7 104.85 341.55 209.7 209.7 209.7
c12 121.1 121.1 121.1 60.55 181.65 121.1
9 -2
Elastic constants (× 10 Nm ) c13 105.1 105.1 105.1 105.1 105.1 52.55
c33 210.9 210.9 210.9 210.9 210.9 210.9
c44 42.5 42.5 42.5 42.5 42.5 42.5

Material property C13S C33L C33S C44L C44S CZn

c11 209.7 209.7 209.7 209.7 209.7 162.8


c12 121.1 121.1 121.1 121.1 121.1 50.8
9 -2
Elastic constants (× 10 Nm ) c13 157.65 105.1 105.1 105.1 105.1 36.2
c33 210.9 105.45 316.35 210.9 210.9 62.7
c44 42.5 42.5 42.5 21.25 63.75 38.5

Fig. 18 shows the amplitudes of   on the interface (   0 ) with distance  for the elastic

moduli of different coating layers. It is noted that   is sensitive to the variation of elastic modulus,

particularly c33 . One can observe a lower c33 has a substantial influence on the peak value of the

shear stress and causes the shear stress to fluctuate severely. The elastic moduli c12 and c13 change
the peak values and cause the position of the peak values to shift forward or backward. The elastic
modulus c44 has negligible influence on the shear stress. When the coating and substrate are of an
identical material, the stress is higher than these. When the coating is made of a thermal resistance
coating material, the internal stresses reduce significantly. The influence of the stiffness on the stress
is more apparent, particularly near the coating. When the coating is consistent with the base material,
the stress would vary from a positive value to a negative value. Finally, one can obtain that the
coating layer with higher elastic moduli c11 , c12 and lower elastic moduli c13 , c33 can reduce the
shear stress and protect the substrate.

33
Fig. 18 The distributions of non-dimensional shear stress   102 on the interface   0 under
different elastic constants; the effect of the following elastic constants: (a) c11 , (b) c12 , (c) c13 , (d) c33 , (e)
c44 .

The non-dimensional vertical stresses   on the interface (   0 ) for the elastic moduli of

different coating layers are displayed in Fig. 18. It is observed that the stress is higher when the
coating and substrate are an identical material (Zn). When the coating is made of a thermal resistance
coating material (ZnO), the stress reduces significantly. The stress is highly sensitive to the variation
in the elastic modulus c33 . In particular, the peak value of the stress appears when c33 decreases

(C33L). The stress nearly remains unaltered when c33 increases. It is apparent that the variations in

the other elastic constants do not alter the trend of the stress field. The stress   increases with the
34
decrease in the elastic modulus c11 , and decreases with the increase in the elastic modulus c12 . The

elastic modulus c13 can alter the maximum stress value. The decrease of the elastic modulus c44
has a negligible impact on the stress.
Therefore, it is important to study the relationship between elastic modulus and interfacial stress
distribution. It provides guidance for the selection and design of coatings. The calculation results in
this section show that whereas anisotropic materials were considered to be isotropic materials in
previous research [59]. The above relationships between the elastic moduli and stresses can be used
to predict the shear stress   and vertical stress   on the interface. This enables engineers to

obtain a general estimate of the stress distributions before tedious calculations. It would be
convenient for engineers to select a suitable coating material.

35
Fig. 19 The distributions of non-dimensional aixal stress   102 on the interface   0 under different
elastic constants; the effect of the following elastic constants: (a) c11 , (b) c12 , (c) c13 , (d) c33 , (e) c44 .

4.3.3 Influences of coating thickness on thermal stress

In engineering, a lighter, inexpensive, and adequately strong structure can be obtained with a
suitable coating thickness. The mechanical properties and failure modes of entire systems depend on
coating thickness[60],[61]. In this section, the variation trends of interface temperature, interface
displacement and interface stress with different coating thickness are discussed for ZnO/Zn coated
structures. Fig. 20 and Fig. 21 show the variation trend of each component along and perpendicular
to the thickness of the coating.
Fig. 20(a) shows temperature increment increases with the decrease along coating thickness on
the interface. The peak value of the axial displacement increases with the decrease in thickness,
whereas the displacement variation at a distance from the loading point is converse to that in Fig.
20(b). Fig. 20(c-d) display that the stresses   and   have a similar variation tendency wherein

the slope of the curves increases as  decreases. The stress gradient is high at the interface when
the coating thickness is exceptionally thin. It is evident that the thickness of the coating alters only
the peak value of all the components, whereas their trend remains unaltered.
Fig. 21 depicts the variations in temperature increment, displacement and key stress components
with   z  for different coating thickness  at the position  0.5,   . It is apparent that each

component satisfies the interface and surface conditions in Eqs. (9) and (11). Fig. 21(a-b) show the
increases in the peak values of temperature increment and the displacement increase with the
increase of coating thickness and the increase rate is larger for a higher value of coating thickness  .
Fig. 21(c) shows the variation trend of the tangential stress   with thickness. The figure illustrates

that the tangential stress varies dramatically when the coating thickness is marginal. Fig. 21(d) shows
that the stress state of the axial stress   varies with the coating thickness.   is a compressive

stress for a thin coating, whereas the converse behavior is observed for a thick coating. With the
increase in coating thickness, the axial stress in the coating transforms from compressive stress to
tensile stress. The thinner the coating, the higher is the residual compressive stress, which is also
revealed by the experiment in the Ref. [10]. It is noteworthy that the maximum values of shear stress
and axial stress move to the coating with the increase in thickness.
It is concluded from the above observations that the thickness of the coating has significant
influences on the coating structure. The stress at the interface would vary rapidly when the coating
layer is too thin. This would result in a high likelihood of tension and shear stress, which, in turn,
36
would cause interface debonding and the failure of the entire structure. Thus, it is reasonable to
recommend that an appropriate coating thickness could be adopted [59]. Based on the accurate and
efficient method proposed in this paper, a reasonable thickness of coating layer for the specified
loading can be obtained for an effective trade-off among cost, weight, and strength of the entire
structure.

Fig. 20 Comparison between different coating thickness  with  on the interface  =0 : (a)
temperature increment  vary with  for different coating thickness; (b) axial displacement u vary
with  for different coating thickness; (c) shear stress   vary with  for different coating thickness; (d)
axial stress   vary with  for different coating thickness.

37
Fig. 21 Comparison between different coating thickness  with  at the position   0.5  0.5,  
along the thickness direction : (a) temperature increment  vary with  for different coating thickness;
(b) axial displacement u vary with  for different coating thickness; (c) shear stress   vary with  for
different coating thickness; (d) axial stress   vary with  for different coating thickness.

5. Optimal coating thickness and evaluation criteria for the interface failure

The main form of failure of a coated structure is its peeling. Therefore, an accurate evaluation of
the interface bonding strength between the coating and substrate material is the key to ensure the
safety and reliability of the structure. The interface failure problem has been considered for
improving the design requirements and reliability of the structure further. An exact solution of a
coated structure is helpful for determining the critical loads and predicting the size of the interface
failure area. Interface failure is a common phenomenon because of the stress concentration, low
bonding strength, and brittleness of the coating material. Herein, peeling of coating is the main form
of failure of thin-film coating materials. The evaluation of the interface bonding strength of the
coating and matrix structure is the key problem in the prediction of coating peeling failure and in the
optimization of the size of coating material. In this section, an effective failure evaluation criterion is
used to study interface delamination under a thermal load.
In this section, physical quantities in real scale are discussed to reveal certain coating failure
problems in real engineering. The problem is discussed in the Cartesian coordinate system  x, y, z  .

The interface failure criterion proposed by Xu et al. [62] is introduced to evaluate coating failure:
r c , (40)

where  c denotes the critical shear strength.  r is the equivalent stress and can be expressed as

 r    k z ,    zx2   zy2 , (41)

where  zx and  zy are the shear stress in the x- and y-directions at an arbitrary point ( x, y,0) on

the interface.  z is the normal stress at this point. k is a constant that represents the interface

38
strength characteristics, which is related to the material characteristics of the coating and substrate.
As discussed above, most materials that are used in coated structures in real engineering have
transversely isotropic properties. At present, however, the experimental test for transversely isotropic
material is very difficult to conduct. To our knowledge, no experimental test is available for
transversely isotropic materials. Thus, the material constant k for transversely isotropic materials is
unknown, and it is unfeasible to carry out the failure evaluation of transversely isotropic coated
structures. To illustrate the evaluation capability of the analytical method proposed in this study, the
degenerate solution for isotropic materials is determined as an example to reveal a certain
mechanism of the failure problem. The isotropic coating structure (TiN/high speed steel) is selected
for the illustration. Its material properties are listed in Table 7.
Table 7 Material properties of isotropic coated structure (TiN/high speed steel)
Terms Elastic modulus (GPa) Poisson ratio
Coating material: TiN 437 0.29
Substrate material: High speed steel 210 0.29

The critical shear strength and material characteristic constant for the coating structure
(TiN/high speed steel) are as follows [62]:
 c  0.56(GPa ), k  0.577. (47)
The failure problem is studied for different types of heat sources to investigate the influence of
the type of heat source on coating failure. The stress components on the interface corresponding to
different heat source types can be calculated using the fundamental solution in Section 3. In this
section, two types of heat source are considered. For an effective comparison, an identical total
strength value is applied for the two types. The total strength of the heat sources remains the same
 60W  and the two types of heat sources are listed in Table 8.
Table 8 Two types of heat source loads
Case 1 Case 2
Load type Point heat source Asymmetric distributed load
H01 ( x, y)  60  0,0   x2 0.2 y2 
W m2  H02 ( x, y) 9.64 1010 e

x2  y 2  h2  h  x  0
Distribution area
h=20 m Coating thickness

Fig. 22 displays the cloud map of the maximum equivalent stress  r for different coating
thicknesses and different intensities of heat source under point heat source. The black line labeled

39
with 0.56 denotes the boundary of the critical shear strength. Fig. 21 shows that the equivalent stress
in the lower region of the map is substantially high. This indicates that the strong heat source and
small coating thickness increases the equivalent stress, which is similar to the scenario in engineering.
The analysis is highly significant for the design of coated structures. Similarly, the elastic modulus
and thickness of coating vary depending on the applications and environment [44].
Fig. 23 shows the cloud maps of the maximum equivalent stress  r for a fixed coating
thickness under the two types of heat sources. The delamination area is clearly visible inside the
contour lines  c  0.56 , where  r   c . In Fig. 23 (a)-(b), the failure zone is a ring whose inner area

is at a safe state (  r   c ). A comparison of the two subfigures in Fig. 23 reveals that the stress at the
points near the source (i.e., around the central area) is significantly higher than that for the
distributed heat sources. Meanwhile, the failure areas are almost similar under different heat loading
conditions. The above results indicate that the distribution types of the heat source may affect only
the marginal local stress in the acting area. Owing to the differences in the loading area position for
the asymmetrically distributed load shown in Fig. 23(b), the failure area shifts to the left of the
origin.
It is noteworthy that the above analysis for certain simple structures and heat sources is an
example. It indicates that the method proposed in this paper can be used to conduct failure and
optimization analysis.

Fig. 22 Cloud map of equivalent stress  r on the interface: the relationship between equivalent stress and
coating thickness and heat strength in which the black line denotes that the equivalent stress is equal to
failure critical value  r =0.56 .

40
Fig. 23 Cloud map of equivalent stress  r on the interface  =0 for the coatings whose coating thickness
is h  20 m : (a) the point heat source is H01 ( x, y)  60 0,0 for Case 1; (b) the asymmetric distributed heat
  x 0.2 y 
2 2

source is H 02 ( x, y) 9.64 1010 e for Case 2; the total energy of the both heat source is the same (the value is
60W )

6. Concluding remarks

In this study, a highly efficient and accurate calculation method for 3D full-field thermal stress
analysis is proposed to provide a guideline for the design of the transversely isotropic thermoelastic
coated medium such as thermal barrier coating structure. All the expressions of the thermoelastic
fields in the full space are obtained in terms of elementary functions, which can facilitate their
further usage. The temperature and stress fields are calculated based on the exact solutions. In the
numerical analysis, the effects of the thermal parameters, elastic modulus, and coating thickness on
the interface stress field under heat source are analyzed. The characteristics of stress distribution and
the design guidance of coating structure are obtained. Some significant conclusions can be drawn as
follows:
(1) For thermal barrier coating structures, coating thermal conductivity determines the
thermal insulation performance of the coating and affects the stress values directly. Meanwhile,
coating thermal expansion coefficient is the key for interfacial stress which can change the form of
the interface stress. The greater difference between the thermal expansion coefficient of coating and
substrate, the range of stress change greater.
(2) The stiffness of the coating in different directions has different effects on the stress field
of the coatings. If the coating stiffness and the substrate stiffness are quite different, the stress on the
interface will be complex and it would make the interface easier to be destroyed.
(3) The coating thickness has a great influence on the maximum stress on the interface. The

41
maximum interfacial stress decreases with the increasing of coating thickness.
(4) Based on the interfacial failure criteria, the optimal thickness and the coating material
could be easy to design using the Green’s function solution obtained in this paper.
The analytical method presented in this paper can provide primary guidance for the design of
thermal barrier coatings and provide useful information for many engineering practices related to
interface analysis. The analytical method in this paper has been verified by finite element simulation,
and our team will combined with experimental verification to solve more complex coating structures
in our future work.

Conflict of interest
We declare that we do not have any commercial or associative interest that represents a conflict
of interest in connection with the work submitted.

Acknowledgments

The authors thankfully acknowledge the financial support from National Natural Science
Foundation of China (NO. 11572119), Natural Science Foundation of Hunan province
(NO.2019JJ50633).

42
References

[1] A. S. Vasiliev, S. S. Volkov, A. A. Belov, S. Y. Litvinchuk, and S. M. Aizikovich, “Indentation


of a hard transversely isotropic functionally graded coating by a conical indenter,” Int. J. Eng. Sci.,
vol. 112, pp. 63–75, Mar. 2017, doi: 10.1016/j.ijengsci.2016.12.002.
[2] J. Bang et al., “Advances in protective layer-coating on metal nanowires with enhanced stability
and their applications,” Appl. Mater. Today, vol. 22, p. 100909, Mar. 2021, doi:
10.1016/j.apmt.2020.100909.
[3] E. Medvedovski and T. Dudziak, “Protective coatings for high-temperature steam oxidation in
coal-fired power plants,” Surf. Coatings Technol., vol. 369, pp. 127–141, Jul. 2019, doi:
10.1016/j.surfcoat.2019.04.049.
[4] A. A. E.-H. Hamada, “Vibration and damping analysis of beams with composite coats,”
Compos. Struct., vol. 32, no. 1–4, pp. 33–38, Jan. 1995, doi: 10.1016/0263-8223(95)00054-2.
[5] M. Ashokkumar, D. Thirumalaikumarasamy, P. Thirumal, and R. Barathiraja, “Influences of
Mechanical, Corrosion, erosion and tribological performance of cold sprayed Coatings A review,”
Mater. Today Proc., Feb. 2021, doi: 10.1016/j.matpr.2021.01.664.
[6] N. P. Padture, “Thermal Barrier Coatings for Gas-Turbine Engine Applications,” Science (80-.
)., vol. 296, no. 5566, pp. 280–284, Apr. 2002, doi: 10.1126/science.1068609.
[7] D. N. Arnold, “An Interior Penalty Finite Element Method with Discontinuous Elements,”
SIAM J. Numer. Anal., vol. 19, no. 4, pp. 742–760, Aug. 1982, doi: 10.1137/0719052.
[8] P. Hansbo, C. Lovadina, I. Perugia, and G. Sangalli, “A Lagrange multiplier method for the
finite element solution of elliptic interface problems using non-matching meshes,” Numer. Math.,
vol. 100, no. 1, pp. 91–115, 2005, doi: 10.1007/s00211-005-0587-4.
[9] V. S. Godiganur, S. Nayaka, and G. N. Kumar, “Thermal barrier coating for diesel engine
application – A review,” Mater. Today Proc., Dec. 2020, doi: 10.1016/j.matpr.2020.10.112.
[10] S. Dhomne and A. M. Mahalle, “Thermal barrier coating materials for SI engine,” J. Mater.
Res. Technol., vol. 8, no. 1, pp. 1532–1537, Jan. 2019, doi: 10.1016/j.jmrt.2018.08.002.
[11] G. M. Smith, E. J. Gildersleeve, X.-T. Luo, V. Luzin, and S. Sampath, “On the surface and
system performance of thermally sprayed carbide coatings produced under controlled residual
stresses,” Surf. Coatings Technol., vol. 387, p. 125536, Apr. 2020, doi:
10.1016/j.surfcoat.2020.125536.
[12] N. Ahmed et al., “Residual Stress in Electrical Discharge Coatings,” Surf. Coatings Technol., p.
127156, Apr. 2021, doi: 10.1016/j.surfcoat.2021.127156.
[13] W. D. Nix, “Mechanical properties of thin films,” Metall. Trans. A, vol. 20, no. 11, pp.

43
2217–2245, Nov. 1989, doi: 10.1007/BF02666659.
[14] W. Zhu, L. Yang, J. W. Guo, Y. C. Zhou, and C. Lu, “Determination of interfacial adhesion
energies of thermal barrier coatings by compression test combined with a cohesive zone finite
element model,” Int. J. Plast., vol. 64, pp. 76–87, Jan. 2015, doi: 10.1016/j.ijplas.2014.08.003.
[15] K. Qi, Y. Yang, G. Hu, X. Lu, and J. Li, “Thermal expansion control of composite coatings on
42CrMo by laser cladding,” Surf. Coatings Technol., vol. 397, p. 125983, Sep. 2020, doi:
10.1016/j.surfcoat.2020.125983.
[16] X. Huang, I. Etsion, and T. Shao, “Effects of elastic modulus mismatch between coating and
substrate on the friction and wear properties of TiN and TiAlN coating systems,” Wear, vol.
338–339, pp. 54–61, Sep. 2015, doi: 10.1016/j.wear.2015.05.016.
[17] A. Ziaei-Asl and M. T. Ramezanlou, “Thermo-mechanical behavior of gas turbine blade
equipped with cooling ducts and protective coating with different thicknesses,” Int. J. Mech. Sci.,
vol. 150, pp. 656–664, Jan. 2019, doi: 10.1016/j.ijmecsci.2018.10.070.
[18] K. M. Kim, S. Shin, D. H. Lee, and H. H. Cho, “Influence of material properties on temperature
and thermal stress of thermal barrier coating near a normal cooling hole,” International Journal of
Heat and Mass Transfer, vol. 54, no. 25–26. pp. 5192–5199, 2011, doi:
10.1016/j.ijheatmasstransfer.2011.08.026.
[19] C.-W. Wu, G.-N. Chen, K. Zhang, G.-X. Luo, and N.-G. Liang, “The effect of periodic
segmentation cracks on the interfacial debonding: Study on interfacial stresses,” Surf. Coatings
Technol., vol. 201, no. 1–2, pp. 287–291, Sep. 2006, doi: 10.1016/j.surfcoat.2005.11.115.
[20] P. Jiang, X. Fan, Y. Sun, D. Li, B. Li, and T. Wang, “Competition mechanism of interfacial
cracks in thermal barrier coating system,” Mater. Des., vol. 132, pp. 559–566, Oct. 2017, doi:
10.1016/j.matdes.2017.07.018.
[21] Y. F. Chen and F. Erdogan, “The interface crack problem for a nonhomogeneous coating
bonded to a homogeneous substrate,” J. Mech. Phys. Solids, vol. 44, no. 5, pp. 771–787, May 1996,
doi: 10.1016/0022-5096(96)00002-6.
[22] J. L. Barber and L. G. Hector, “Thermoelastic Contact Problems for the Layer,” J. Appl. Mech.,
vol. 66, no. 3, p. 806, 1999, doi: 10.1115/1.2791759.
[23] C. Zhou, Q. Zhang, and Y. Li, “Thermal shock behavior of nanostructured and microstructured
thermal barrier coatings on a Fe-based alloy,” Surf. Coatings Technol., vol. 217, pp. 70–75, Feb.
2013, doi: 10.1016/j.surfcoat.2012.11.074.
[24] A. K. Krella, A. T. Sobczyk, A. Krupa, and A. Jaworek, “Thermal resistance of Al2O3 coating
produced by electrostatic spray deposition method,” Mech. Mater., vol. 98, pp. 120–133, Jul. 2016,
doi: 10.1016/j.mechmat.2016.05.002.
44
[25] H. Toyama, M. Niwa, J. Xu, and A. Yonezu, “Failure assessment of a hard brittle coating on a
ductile substrate subjected to cyclic contact loading,” Eng. Fail. Anal., vol. 57, pp. 118–128, Nov.
2015, doi: 10.1016/j.engfailanal.2015.07.039.
[26] W. Zhu, L. Yang, J. W. Guo, Y. C. Zhou, and C. Lu, “Determination of interfacial adhesion
energies of thermal barrier coatings by compression test combined with a cohesive zone finite
element model,” Int. J. Plast., vol. 64, pp. 76–87, 2015, doi: 10.1016/j.ijplas.2014.08.003.
[27] K. Sfar, J. Aktaa, and D. Munz, “Analysing the Failure Behaviour of Thermal Barrier Coatings
Using the Finite Element Method,” pp. 203–211.
[28] I. A. Roberts, C. J. Wang, R. Esterlein, M. Stanford, and D. J. Mynors, “A three-dimensional
finite element analysis of the temperature field during laser melting of metal powders in additive
layer manufacturing,” Int. J. Mach. Tools Manuf., vol. 49, no. 12–13, pp. 916–923, 2009, doi:
10.1016/j.ijmachtools.2009.07.004.
[29] C. Espinoza, E. Ramos-Moore, and D. Celentano, “Effect of Microstructure on Thermoelastic
Stresses in Al2O3/Ti(C,N,O)/Ti(C,N) Coating Systems Studied by Finite Element Method
Simulations,” Mater. Lett., p. 129777, Mar. 2021, doi: 10.1016/j.matlet.2021.129777.
[30] A. Abdelgawad and K. Al-Athel, “Effect of TGO thickness, Pores, and Creep on the Developed
Residual Stresses in Thermal Barrier Coatings under Cyclic Loading using SEM Image-Based Finite
Element Model,” Ceram. Int., Apr. 2021, doi: 10.1016/j.ceramint.2021.03.336.
[31] A. Masud and P. Chen, “Thermoelasticity at finite strains with weak and strong discontinuities,”
Comput. Methods Appl. Mech. Eng., vol. 347, pp. 1050–1084, Apr. 2019, doi:
10.1016/j.cma.2018.12.024.
[32] D. Y. Liu and W. Q. Chen, “Thermal stress analysis of a trilayer film/substrate system with
weak interfaces,” Compos. Part B Eng., vol. 43, no. 8, pp. 3445–3452, Dec. 2012, doi:
10.1016/j.compositesb.2012.01.041.
[33] Y. Huang, D. Ngo, and A. J. Rosakis, “Non-uniform, axisymmetric misfit strain: in thin films
bonded on plate substrates/substrate systems: the relation between non-uniform film stresses and
system curvatures,” Acta Mech. Sin., vol. 21, no. 4, pp. 362–370, Aug. 2005, doi:
10.1007/s10409-005-0051-9.
[34] W. Chen, E. Pan, H. Wang, and C. Zhang, “Theory of indentation on multiferroic composite
materials,” J. Mech. Phys. Solids, vol. 58, no. 10, pp. 1524–1551, 2010, doi:
10.1016/j.jmps.2010.07.012.
[35] L. L. HENCH and E. C. ETHRIDGE, “Biomaterials—The Interfacial Problem,” in Advances in
Biomedical Engineering, Elsevier, 1975, pp. 35–150.
[36] D. Gasiorek, P. Baranowski, J. Malachowski, L. Mazurkiewicz, and M. Wiercigroch,
45
“Modelling of guillotine cutting of multi-layered aluminum sheets,” J. Manuf. Process., vol. 34, pp.
374–388, Aug. 2018, doi: 10.1016/j.jmapro.2018.06.014.
[37] L. Wang, C. Deng, K. Ding, S. Guo, Z. Li, and X. Lin, “Model construction and effect of
thermally grown oxide dynamic growth on distribution of thermal barrier coatings,” Ceram. Int.,
Mar. 2021, doi: 10.1016/j.ceramint.2021.03.161.
[38] L.-L. Ke, J. Yang, S. Kitipornchai, and Y.-S. Wang, “Electro-mechanical frictionless contact
behavior of a functionally graded piezoelectric layered half-plane under a rigid punch,” Int. J. Solids
Struct., vol. 45, no. 11–12, pp. 3313–3333, Jun. 2008, doi: 10.1016/j.ijsolstr.2008.01.028.
[39] R. Kouitat Njiwa and J. von Stebut, “Boundary element numerical modelling as a surface
engineering tool: application to very thin coatings,” Surf. Coatings Technol., vol. 116–119, pp.
573–579, Sep. 1999, doi: 10.1016/S0257-8972(99)00230-3.
[40] Y. Zhang, Y. Gu, and J.-T. Chen, “Boundary element analysis of the thermal behaviour in
thin-coated cutting tools,” Eng. Anal. Bound. Elem., vol. 34, no. 9, pp. 775–784, Sep. 2010, doi:
10.1016/j.enganabound.2010.03.014.
[41] J. Vallepuga Espinosa and A. Foces Mediavilla, “Boundary element method applied to three
dimensional thermoelastic contact,” Eng. Anal. Bound. Elem., vol. 36, no. 6, pp. 928–933, Jun. 2012,
doi: 10.1016/j.enganabound.2011.12.010.
[42] X. Qin, J. Zhang, G. Xie, F. Zhou, and G. Li, “A general algorithm for the numerical evaluation
of nearly singular integrals on 3D boundary element,” J. Comput. Appl. Math., vol. 235, no. 14, pp.
4174–4186, May 2011, doi: 10.1016/j.cam.2011.03.012.
[43] F. Wu, X.-Y. Li, W.-Q. Chen, G.-Z. Kang, and R. Müller, “Indentation on a transversely
isotropic half-space of multiferroic composite medium with a circular contact region,” Int. J. Eng.
Sci., vol. 123, pp. 236–289, Feb. 2018, doi: 10.1016/j.ijengsci.2017.11.013.
[44] Z. Shi and S. Ramalingam, “Thermal and mechanical stresses in transversely isotropic
coatings,” Surf. Coatings Technol., vol. 138, no. 2–3, pp. 173–184, Apr. 2001, doi:
10.1016/S0257-8972(00)01167-1.
[45] L. Y. Bahar, “Transfer matrix approach to elastodynamics of layered media,” J. Acoust. Soc.
Am., vol. 57, no. 3, pp. 606–609, Mar. 1975, doi: 10.1121/1.380476.
[46] J.-Q. XU, Y. MUTOH, and L.-D. FU, “Theoretical Solution for Concentrated Force on the Free
Surface of a Coating Material. 1st Report. On the Two Dimensional Solutions.,” Trans. Japan Soc.
Mech. Eng. Ser. A, vol. 68, no. 672, pp. 1259–1265, 2002, doi: 10.1299/kikaia.68.1259.
[47] F. Akbari, A. Khojasteh, and M. Rahimian, “Asymmetric Green’s functions for exponentially
graded transversely isotropic substrate–coating system,” J. Cent. South Univ., vol. 25, no. 1, pp.
169–184, Jan. 2018, doi: 10.1007/s11771-018-3727-6.
46
[48] P. F. Hou and Y. Zhang, “An accurate and efficient method for the piezoelectric coated
functional devices based on the three-dimensional Green’s functions under a tangential point force,”
Int. J. Mech. Sci., vol. 133, no. May, pp. 387–415, 2017, doi: 10.1016/j.ijmecsci.2017.08.053.
[49] P. F. Hou, W. H. Zhang, and J.-Y. Chen, “Three-dimensional exact solutions of homogeneous
transversely isotropic coated structures under spherical contact,” Int. J. Solids Struct., vol. 161, pp.
136–173, Apr. 2019, doi: 10.1016/j.ijsolstr.2018.11.013.
[50] P. F. Hou, H. Y. Jiang, J. Tong, and S. M. Xiong, “Study on the coated isotropic thermoelastic
material based on the three-dimensional Green’s function for a point heat source,” Int. J. Mech. Sci.,
vol. 83, pp. 155–162, 2014, doi: 10.1016/j.ijmecsci.2014.03.019.
[51] J. Tong, Z.-L. Xu, J.-P. Li, Y. Zhang, and P.-F. Hou, “Green’s function for a line heat source
acting on the surface of a coated isotropic thermoelastic material,” J. Therm. Stress., vol. 42, no. 2,
pp. 279–293, Feb. 2019, doi: 10.1080/01495739.2018.1461582.
[52] M. Cerit and M. Coban, “Temperature and thermal stress analyses of a ceramic-coated
aluminum alloy piston used in a diesel engine,” Int. J. Therm. Sci., vol. 77, pp. 11–18, Mar. 2014,
doi: 10.1016/j.ijthermalsci.2013.10.009.
[53] Y. Tan, A. Shyam, W. B. Choi, E. Lara-Curzio, and S. Sampath, “Anisotropic elastic properties
of thermal spray coatings determined via resonant ultrasound spectroscopy,” Acta Mater., vol. 58,
no. 16, pp. 5305–5315, Sep. 2010, doi: 10.1016/j.actamat.2010.06.003.
[54] P.-F. Hou, A. Y. T. Leung, and C.-P. Chen, “Fundamental solution for transversely isotropic
thermoelastic materials,” Int. J. Solids Struct., vol. 45, no. 2, pp. 392–408, Jan. 2008, doi:
10.1016/j.ijsolstr.2007.08.024.
[55] A. Karlsson, “A fundamental model of cyclic instabilities in thermal barrier systems,” J. Mech.
Phys. Solids, vol. 50, no. 8, pp. 1565–1589, Aug. 2002, doi: 10.1016/S0022-5096(02)00003-0.
[56] Q. Chen, P. Hu, J. Pu, T. Zhang, and J. H. Wang, “Interfacial interaction and roughness
parameters effects on the residual stresses in DCL-TBC system with different thickness
distributions,” Ceram. Int., Sep. 2020, doi: 10.1016/j.ceramint.2020.09.132.
[57] J. AKTAA, K. SFAR, and D. MUNZ, “Assessment of TBC systems failure mechanisms using a
fracture mechanics approach,” Acta Mater., vol. 53, no. 16, pp. 4399–4413, Sep. 2005, doi:
10.1016/j.actamat.2005.06.003.
[58] L.Wang et al., “Finite element simulation of stress distribution and development in 8YSZ and
double-ceramic-layer La2Zr2O7/8YSZ thermal barrier coatings during thermal shock,” Appl. Surf.
Sci., vol. 258, no. 8, pp. 3540–3551, Feb. 2012, doi: 10.1016/j.apsusc.2011.11.109.
[59] J. H. Wang, “Substrate effects on piezoresponse force microscopy electromechanical responses
of piezoelectric thin films,” Int. J. Solids Struct., vol. 128, pp. 149–159, Dec. 2017, doi:
47
10.1016/j.ijsolstr.2017.08.024.
[60] C. He, Z. Xie, Z. Guo, and H. Yao, “Fracture-mode map of brittle coatings: Theoretical
development and experimental verification,” J. Mech. Phys. Solids, vol. 83, pp. 19–35, Oct. 2015,
doi: 10.1016/j.jmps.2015.06.005.
[61] Y. L. Bai, J. Bai, H. L. Li, F. J. Ke, and M. F. Xia, “Damage evolution, localization and failure
of solids subjected to impact loading,” Int. J. Impact Eng., vol. 24, no. 6–7, pp. 685–701, Jul. 2000,
doi: 10.1016/S0734-743X(99)00151-7.
[62] P.-F. Hou, W.-H. Zhang, and J.-Y. Chen, “Three-dimensional exact solutions of transversely
isotropic coated structures under tilted circular flat punch contact,” Int. J. Mech. Sci., vol. 151, pp.
471–497, Feb. 2019, doi: 10.1016/j.ijmecsci.2018.11.031.

48

You might also like