You are on page 1of 13

Science in China Series G: Physics, Mechanics & Astronomy

www.scichina.com
2008 SCIENCE IN CHINA PRESS phys.scichina.com
www.springerlink.com
Springer-Verlag

Calculation of transient dynamic stress


intensity factors at bimaterial interface
cracks using a SBFEM-based frequency-
domain approach
Z. J. YANG1,2 & A. J. DEEKS2
1
College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310027, China;
2
School of Civil and Resource Engineering, University of Western Australia, 35 Stirling Highway, Crawley,
WA 6009, Australia

A frequency-domain approach based on the semi-analytical scaled boundary finite


element method (SBFEM) was developed to calculate dynamic stress intensity
factors (DSIFs) at bimaterial interface cracks subjected to transient loading. Be-
cause the stress solutions of the SBFEM in the frequency domain are analytical in
the radial direction, and the complex stress singularity at the bimaterial interface
crack tip is explicitly represented in the stress solutions, the mixed-mode DSIFs
were calculated directly by definition. The complex frequency-response functions
of DSIFs were then used by the fast Fourier transform (FFT) and the inverse FFT to
calculate time histories of DSIFs. A benchmark example was modelled. Good re-
sults were obtained by modelling the example with a small number of degrees of
freedom due to the semi-analytical nature of the SBFEM.

scaled boundary finite element method, bimaterial interface crack, frequency domain, dynamic stress
intensity factors, fast Fourier transform

1 Introduction
Bimaterial interfaces exist in many engineering structures such as composites, sandwich plates,
welded joints, and structures strengthened by adhesively-bonded advanced composite materials.
Because the interfaces are generally weaker than the constituent materials, they are susceptible to
fracture loadings that may lead to cracks on the interfaces. Once cracks on the interfaces develop,
critically evaluating the fracture resistance of cracked interfaces is vital to the structural integrity,
reliability and safety.

Received August 30, 2006; accepted May 25, 2007


doi: 10.1007/s11433-008-0057-y

Corresponding author (email: zjyang@liv.ac.uk)


Supported by the Scientific Research Foundation for the Returned Overseas Chinese Scholars, Ministry of Education of China (Grant No.
J20050924) and the Australian Research Council Discovery Project (Grant No. DP0452681)

Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
For bimaterial structures with interface cracks subjected to dynamic loadings, dynamic stress
intensity factors (DSIFs) are the most important parameters in the linear elastic fracture mechanics
for predicting when and in which direction a new crack initiates or an existing crack propagates.
Both analytical methods and numerical methods, such as the finite difference method (FDM), the
finite element method (FEM) and the boundary element method (BEM), have been widely applied
to evaluating the DSIFs. In terms of how the time histories of DSIFs are computed, these methods
can be classified into time-domain methods based on explicit or implicit time integration schemes,
and integral transform methods including the Laplace transform and the Fourier transform (i.e., the
frequency-domain method based on frequency analysis and the discrete or fast Fourier transform
(FFT)). For example, Chen and Wu[1] used the FEM with the Newmark integration scheme to
model bimaterial interface facture. Special crack-tip meshes composed of singular elements em-
bedding proper singularities were used to model the crack. Geubelle[2] developed a numerical
scheme based on a spectral representation of boundary integral relations for modelling homoge-
neous and interface dynamic fracture problems. Marur and Tippur[3] used the FEM and the dis-
placement extrapolation technique to calculate DSIFs of bimaterial and graded interface cracks
under impact loading with an explicit time-integration scheme. Dineva et al.[4] used the BEM to
calculate DSIFs of bimaterial interface cracks under time-harmonic loadings. Lira-Vergara and
Rubio-Gonzalez[5] used the Laplace and Fourier transform to solve the equation of motion and the
numerical Laplace transform inversion to obtain the time histories of DSIFs of cracks at the in-
terface of two orthotropic half-space. Modelling of crack propagation on bimaterial interfaces was
also carried out, e.g., by Xu and Needleman[6] and Nishioka et al.[7] using the time-domain FEM,
Dornowski and Perzyna[8] using the time-domain FDM, and Lei et al.[9] using a time-domain BEM.
In general, analytical solutions can be obtained only for simple problems and the FDM is suitable
only for problems with regular geometries. The FEM has much wider applicability but needs fine
crack-tip meshes to accurately represent the stress singularities[1,2,6,7]. On the contrary to the
square-root singularity at homogeneous cracks, the stress singularity of bimaterial interface cracks
is a complex value (dependent upon material properties), which needs complicated shape functions
to represent. For example, Chen and Wu[1] used eighteen functions in the singular elements for
displacement interpolation. The BEM discretizes domain boundaries only and can thus represent
crack propagation in a simpler way. However, the need of fundamental solutions limits its appli-
cability considerably. In addition, most of the existing numerical studies used the displacement
extrapolation technique to calculate DIFs but uncertainty exists in selecting optimal locations on
the crack surfaces for best accuracy[4]. Therefore, improvements are still needed to model the in-
terface fracture problems more effectively and efficiently.
The scaled boundary finite-element method (SBFEM), developed recently by Wolf and
Song[10,11], is a semi-analytical method combining the advantages of the FEM and the BEM, and
also possessing its own attractions, including: (i) it reduces the modelled spatial dimensions by one
and discretizes the boundaries only as the BEM, but does not need fundamental solutions; (ii) no
discretization of crack surfaces, boundaries and bimaterial interfaces that are connected to scaling
centres is needed, potentially reducing the computational cost significantly; (iii) the displacement
and stress fields are analytical in the radial direction. The analytical stress field explicitly repre-
sents stress singularities at crack tips, which allows accurate SIFs to be computed directly from the
definition. This has been well demonstrated by recent studies calculating static SIFs for isotropic
materials[12] and anisotropic materials[13]. The above features of the SBFEM have also been ex-

520 Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
ploited in developing a very simple remeshing procedure for automatic modelling of mixed-mode
crack propagation problems[14]. The only application of the SBFEM to the dynamic fracture
problems appears to be the work of Song in 2004[15] using a standard time-integration scheme. The
time-domain method requires a static stiffness matrix and a mass matrix. The stiffness matrix can
be calculated readily after an eigenproblem is solved. The mass matrix derived by Song[15] corre-
sponds to the low-frequency expansion of the dynamic stiffness matrix and thus can only accu-
rately represent the inertial effects at low frequencies. This requires that the size of the subdomains
or super-elements be small enough to account for the highest frequency component of interest. This
may lead to considerable computational cost in solving a significant number of eigenproblems in
each time step.
A frequency-domain approach has certain advantages over the time-domain approach, such as
no need for a mass matrix, so that coarser meshes may be used, and once a complex fre-
quency-response function is obtained, it can be used in combination with FFT and inverse FFT
(IFFT) to calculate transient responses for various forms of dynamic loads. However, in the fre-
quency domain, the SBFEM has a nonhomogeneous second-order differential governing equation
system which is difficult to solve analytically. One procedure currently available in the frequency
domain was developed by Song and Wolf[16]. This procedure first transforms the second order
differential equations to first order equations with the number of degrees of freedom (DOFs)
doubling. This is followed by a complicated treatment involving a sequence of matrix function
operations, leading to an analytical solution consisting of infinite series, the number of terms in
which must be selected by the analyst. The resultant analytical formulae are difficult to interpret
physically. As an alternative, the authors recently developed a new analytical series solution in the
frequency domain using the Frobenius procedure[17]. Compared with Song and Wolfs procedure[16],
our solution procedure is very easy to follow. Like the static solution, the frequency-domain solu-
tion has clear physical meaning, and calculating the stress field is trivial. This makes calculating
DSIFs from the definition as simple as from the static stress field. In addition, the dynamic stiffness
matrix is explicitly formed and the number of series terms in the solution is determined by the
desired accuracy.
This study aims at developing a new frequency-domain approach using the SBFEM to model
transient dynamic bimaterial interface fracture problems. The newly-developed Frobenius solu-
tion[16] in the frequency domain is employed to calculate complex frequency-DSIFs functions,
which are subsequently used with FFT and IFFT to calculate time histories of DSIFs. The fol-
lowing sections present the fundamentals of the SBFEM, the approach to extract DSIFs from the
solution and the procedure to calculate time histories of DSIFs, followed by a detailed modelling of
a benchmark bimaterial interface fracture problem and discussion of results.

2 The scaled boundary finite element method


The essential features of the SBFEM are illustrated by Figure 1(a) and 1(b). Figure 1(a) shows a
domain of an arbitrary shape, modelled by three subdomains. Any scheme of subdivision, with
various numbers, shapes and sizes of subdomains, can be used, as long as a scaling centre for each
subdomain can be found from which the subdomain boundary is fully visible. Figure 1(b) shows
the details of Subdomain 1. The subdomain is represented by scaling a defining curve S relative to
a scaling centre. The defining curve is usually taken to be the domain boundary, or part of the
boundary. A normalised radial coordinate is defined, varying from zero at the scaling centre and

Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531 521
unit value on S. A circumferential coordinate s is defined around the defining curve S. A curve
similar to S defined by = 0.5 is shown in Figure 1(b). The coordinates and s form the local
coordinate system, which is used in all subdomains. The governing equilibrium equations of the
SBFEM for elastostatics and elastodynamics can then be derived using a Galerkin approach within
the context of virtual work principle[16,18]. This will not be repeated here.

Figure 1 The concept of the scaled boundary finite-element method. (a) Subdomaining of a domain; (b) subdomain 1.

In a frequency-domain analysis, the first step is to compute the complex frequency-response


functions. The governing equilibrium equations of the SBFEM under a harmonic excitation with
constant frequency f (Hz), e.g. p = p0ei2f t are[16]
p0 = E 0 u( ), + E 1T u( ) =1 , (1)

E 0 2 u( ), + ( E 0 + E 1T E 1 ) u( ), E 2 u( ) + (2f ) 2 2 M 0 u( ) = 0, (2)
where p0 are the magnitudes of the equivalent nodal loads. E 0, E 1, E 2 and M 0 are coefficient ma-
trices that are dependent on the geometry of the subdomain boundaries and the material properties.
u( ) represent the magnitudes of the nodal displacements and are analytical functions of the radial
coordinate . Eq. (1) and eq. (2) first apply to subdomains. The contributions of subdomains to the
coefficient matrices and equivalent nodal loads are then assembled in the same way as the con-
tributions of finite elements in FEM. The assembled equation system for the whole domain has the
same form as eq. (1) and eq. (2).

3 The frobenius solution procedure to eq. (2)


Eq. (2) is a second-order nonhomogeneous Euler-Cauchy differential equation system with respect
to the radial coordinate . The non-homogeneity complicates its solution considerably. An
easy-to-follow analytical solution to eq. (2) was recently developed by the authors[17] using the
Frobenius procedure. Interested readers are referred to ref. [16] for detailed derivation. The solu-
tion with (k + 1) series is

522 Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
n n n k +1 n
u = ci ( i ) i + ( ci ( 2 gi ) + " ( ci ( k +1 gi ) = ci ( ( ),
1 2 k +1 j
k +1 i ) i ) i ) j
i (3)
i =1 i =1 i =1 j =1 i =1

where
k +1 k +1
gi = Gi i ; (4a)
k +1 k +1 k 2
Gi = Bi Bi " Bi ; (4b)
2 1
k +1
Bi = (2f ) 2 ( k +1i ) 2 E 0 + ( k +1i )( E1T E1 ) E M 0 ; (4c)
k +1 k +1
di = Bi ( k d i ); (4d)
1
d i = ci i ; (4e)
k +1
i = k i + 2; (4f )

j i for j = 1,
i = j (4g)
gi for j = 2, 3, " , k + 1,
where the subscript i ranges from 1 to n for all the variables and n is the degrees of freedom (DOFs)
of the problem. Eq. (3) shows that the solution to elastodynamics consists of (k+1) summations of
series. The first summation is the solution to the corresponding homogeneous equations of eq. (2),
i.e., the governing equations for elastostatics. The added k summations account for dynamic effects.
The more series terms are added, the more accurately the solution represents the dynamic effects.
1
i = i and i are the positive eigenvalues and corresponding eigenvectors of the standard linear
eigenproblem[18] formed from the elastostatic governing equations. ci are constants determined by
boundary conditions.
Eqs. (4a)(4f) describe an iterative process in which the number of added summation terms k in
eq. (3) is determined by a convergence criterion[16]:
max (|R|)< ; (5a)
n
R = (2f )2 M 0 (
k +2
i ) k +1
( d i ), (5b)
i =1
where R can be regarded as the residual vector and is the convergence tolerance.
Setting = 1 in eq. (3) leads to
k +1
u =1 = ub = 1c or c = 11ub , (6)
where ub is the nodal displacements on the boundary, c = {c1 c2 cn}T and the matrix
k +1 k +1 k +1
1 = j 1 j 2 " j n . (7)
j =1 j =1 j =1
Substituting eq. (3) into eq. (1) leads to
p0 = E 0 2 c + E1T1c, (8)
where the matrix
k +1 k +1 k +1
2 = j 1 j 1 j 2 j 2 " j n j
n . (9)
j =1 j =1 j =1
Substituting eq. (6) into eq. (8) gives

Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531 523
p0 = ( E 0 211 + E 1T )ub = K d ub . (10)
Therefore, the dynamic stiffness matrix of the domain with respect to DOFs on the boundary is
K d = E 0 211 + E 1T . (11)
The nodal displacement vector ub can be calculated by eq. (10) by applying boundary conditions
on ub and loading conditions on p0. The integration constants c are then obtained using eq. (6).
Assuming the (k+1)th solution meets the criterion eq. 5a, the displacement field is recovered as
k +1 n
u( , s ) = N ( s) ci ( i ) j i ,
j
(12)
j =1 i =1
where N(s) is the matrix of shape functions at the circumferential direction, which are constructed
as in FEM, typically using polynomial functions.
The stress field is then obtained as
k +1 n k +1 n
( , s ) = DB1 ( s ) ci ( ji ) ( ( ) + DB (s) c ( ) ,
j
i 1) j 2 ( j i 1) j
i i i (13)
j =1 i =1 j =1 i =1
where D is the elasticity matrix and B1(s) and B2(s) are relevant matrices[11].
Eq. (13) can be rewritten as
k +1 n
( , s ) = ci ( ( (s) ),
j
i 1) j
i (14)
j =1 i =1

where each term in eq. (14) can be interpreted as a stress mode where ji(s) depends on the
circumferential coordinate s
j x ( s )i

i ( s ) = j y ( s )i = D ji B1 ( s ) + B 2 ( s ) j i . ( ) (15)
j
xy ( s )i
There are totally n(k+1) stress modes in the stress solutions eq. (14).

4 Calculation of dynamic stress intensity factors


Figure 2 shows a bimaterial domain with an interface crack modeled by the SBFEM. The scaling
center is always placed at the crack tip. Various definitions of DSIFs have been proposed based on
the asymptotic expansions of the stress fields around the bimaterial crack tip[7,19,20]. This study uses
the following definition
i 0.5 + i
( K I + iK II ) r ( K I + iK II ) r
( y + i xy ) =0 = = , (16)
2r l 2l l
where r and are the polar coordinates with the origin at the crack tip as shown in Figure 2, and l is
a characteristic length which is usually taken to be the entire crack length l = 2a, where a is the half
crack length. The oscillation index is defined as
1 1 / G1 + 1/ G2
= ln , (17)
2 2 / G2 + 1/ G1
where

524 Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
Figure 2 A bimaterial domain with an interface crack modelled by the SBFEM.

3 4 j , for plane strain,



j = 3 j (18)
1 + , for plane strain,
j

where j and G j ( j = 1, 2) are Poissons ratios and shear moduli of the two materials, respectively.
The singular stresses in eq. (16) can be rewritten in the following matrix form:
r r
cos sin
y =0 1 l l KI
= (19)
xy =0 2r r r K II
sin cos
l l
or
r r
cos l sin l
KI y =0
= 2r . (20)
K II r r xy =0
sin cos
l l
In the SBFEM, the singular stresses along the radial line at = 0 are determined from eq. (14).
As the crack-tip stress field in a homogeneous material[13 15], only two stress modes in eq. (14)

contribute to the stress singularity. The eigenvalues of these two stress modes are a pair of com-
plex conjugates whose real part is 0.5 and imaginary part is . All the other stress modes in eq. (14)
have eigenvalues whose real parts are greater than 1.0 and thus are all zero for 0 or r0. De-
noting these two singular stress modes as mode I and mode II (with eigenvalues I = 0.5+ i and II
= 0.5 i ), the singular stresses along the radial line at = 0 (from eq. (14)) are

Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531 525
y =0 1 x ( s = s A )
= ci i , (21)
xy i = I,II
=0 xy ( s = s A ) i
where s = sA corresponds to = 0 (Figure 2).
Note the relationship between r and is ( = 0 at the crack tip and 1 at the boundary, see Figure
2)
r = L( ), (22)
where L( ) is the distance between the crack tip and the intersection point of the polar line r and the
domain boundary.
Substituting eq. (21) and eq. (22) into eq. (20) leads to
L L
cos 0 sin 0
KI l l 1 x ( s = s A )
= 2 L0 ci i , (23)
K II L0 L0 i = I,II xy ( s = s A ) i
sin cos l
l
where L0 = L( = 0) is the distance between the crack tip and the intersection point of the crack
direction and the boundary. Eq. (23) is valid for any (0< 1) which means the stress solutions
at any point on the radial line at = 0 can be used to calculate the DSIFs. For simplicity, the stress
modes of the intersection point A with = 1 on the crack domain boundary as shown in Figure 2 are
used, leading to the following DSIFs of bimaterial interface cracks:
L L
cos 0 sin 0
KI l l x ( s = s A )
= 2L0
ci .
L0 i = I,II xy ( s = s A ) i
(24)
K II L0
sin cos
l l
It should be noted that the above definition is based on the local coordinate system where the
cracking direction is assumed to coincide with the global x axis. For a domain with a crack surface
in an arbitrary direction, as usually occurs in crack propagation modeling[14], the stress terms x
and xy in eq. (24) should first be transformed by the standard procedure to normal (mode I) stress
n and shear (mode II) stress s on the cracking surface plane at point A so that eq. (24) becomes
L L
cos 0 sin 0
KI l l n ( s = s A )
= 2L0
ci .
L0 i = I,II s ( s = s A ) i
(25)
K II L0
sin cos
l l
It should also be noted that point A does not need to be an existing node because the stresses at
any point of the domain can be calculated by eq. (14).
A total dynamic stress intensity factor K0 can be defined as
2 2

K0 = K I2 + K II2 = 2L0 [ ci n ( s = s A )i ] + [ ci s ( s = s A )i ] . (26)
i = I,II i = I,II

526 Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
5 Calculation of time histories of DSIFs
The time histories of DSIFs are computed by the following procedure. The complex fre-
quency-response functions (CFRFs) for a wide range of frequencies are first computed using the
Frobenius solution procedure described in sec. 3. The dynamic stress intensity factors KI and KII are
then extracted directly from the stress responses, as presented in the previous section. This is fol-
lowed by a FFT of the transient load and an inverse FFT of the CFRFs to calculate the time histo-
ries of DSIFs. The readers are referred to [21] for details of the FFT. The functions fft() and ifft() in
MATLAB[22] are used to conduct FFT and IFFT, respectively.
Damping plays a vital role in calculating accurate time responses using frequency-domain
methods. Too low damping may lead to CFRFs with pitches representing resonance, which in turn
leads to spurious oscillations in the time histories, whereas too high damping may underestimate
CFRFs, which in turn results in underestimation of time histories. The material damping effect is
taken into account by modifying the elastic moduli to incorporate an internal damping coefficient ,
forming the complex Youngs modulus Ec = E(1+2 i) and the complex shear modulus Gc = G(1 +
2 i), where E and G are the elastic Youngs modulus and the elastic shear modulus, respectively.
The damping coefficients ranging from 0 to 0.05 are modelled.
A convergence tolerance of = 1103 in eq. (5a) is used for all the analyses. Quantitative
comparisons of the proposed method with the FEM, the BEM and the time-domain SBFEM[15] are
not conducted with respect to the computational costs, as the results would be influenced greatly by
the algorithms used in many different steps in the respective method.

6 Numerical example, results and discussion


The example solved here is a rectangular bimaterial plate with a central crack subjected to uniform
tractions on its upper and lower surfaces. The geometry, boundary and loading conditions are
shown in Figure 3(a). The crack length is 2a = 4.8 mm. Both materials are isotropic. The upper part
of the plate is quartz (Material 1) and the lower part is copper (Material 2). The elastic material
properties are: shear modulus G1 = 47.9 GPa, Poissons ratio 1 = 0.058 and density 1 = 2650
kg/m3; G2 = 48.1 GPa, 2 = 0.2976 and 2 = 8960 kg/m3. The oscillation index is 0.04696 ac-
cording to eq. (17) and eq. (18). This problem under a time-harmonic load with amplitude P0 was
modeled by Dineva et al.[4] using the BEM. The time histories of DSIFs of this problem under a
ramped form of time-dependent uniform traction p(t) with a maximum value P0 as shown in Figure
3(a) was modelled by Song[15] using the time-domain SBFEM. All SIFs reported below are nor-
malised by a factor P0(a)1/2 where a is the half crack length.
A plane strain condition is assumed. Only half of the plate is modelled due to symmetry with
two-noded linear line elements. Three meshes as shown in Figure 3(b), 3(c) and 3(d) with 33 nodes,
57 nodes and 89 nodes, respectively, are modelled to investigate mesh convergence. The domain is
divided into three subdomains. The subdomain with the crack has its scaling centre at the crack tip.
The scaling centres of the other two subdomains are placed in the middle of right edges which are
not discretized. It may also be noted that the crack faces and the material interface are not discre-
tized. The stresses at point A in Figure 3(a) are used to extract DSIFs according to eq. (25) and eq.
(26).
The CFRFs of KI, KII, K0 and the crack mouth opening displacements (CMOD) were calculated
using the frequency-domain SBFEM for f = 0 Hz to 300000 Hz with an interval f = 1000 Hz.

Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531 527
Figure 3 A rectangular bimaterial plate with a central crack. (a) Dimensions and loading conditions; (b) 33 nodes; (c) 57 nodes;
(d) 89 nodes.

Figure 4 shows the modelled normalised dimensionless frequency a0 (=2fa/cs1)-normalised K0


curves using the coarsest mesh with 33 nodes for the damping coefficients = 0 where cs1 =
(G1/ 1)1/2 is the shear wave velocity. Excellent agreement can be observed between the present
results and those from [15] which reported results up to a0 = 0.6. It should be noted that the present
method directly uses the dynamic stiffness matrix eq. (11) which is implicitly analytical to the
frequency, and the method in [15] uses mass matrix and stiffness matrix which correspond to a
low-frequency expansion of the dynamic stiffness matrix, although both methods use the SBFEM.
This may have led to discrepancies for higher frequencies. The CFRFs of CMOD normalized by

Figure 4 Dimensionless frequency a0 = 2fa/cs1-magnitude of normalized K0 curves.

528 Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
G1/(aP0) from these two methods are shown in Figure 5. Again excellent agreement can be seen. It
may be noted that the present results were obtained from the coarsest mesh (Figure 3(b)) with 3
subdomains and only 33 nodes whereas the results in [15] were calculated from 11 subdomains and
more than 100 nodes.

Figure 5 Dimensionless frequency a0 = 2fa/cs1-magnitude of normalized CMOD curves.

Figure 6 and Figure 7 show the CFRFs of the real part and the imaginary part of KI respectively
from different meshes using the present method for = 0.01. The dashed lines represent the results
from the coarsest mesh with 33 nodes. It can be seen that the two finer meshes led to virtually
identical results and those from the coarsest mesh are very close to the former. This indicates that
the mesh in Figure 3(b) with only 33 nodes is sufficiently fine to calculate accurate results. Other
values of were also modelled and same conclusions can be drawn. The CFRFs as in Figures 6 and
7 were used in FFT and IFFT to calculate time histories of DSIFs.
Figure 8 and Figure 9 present the modelled time histories of normalised CMOD and of nor-
malised DSIFs KI and KII respectively, under the ramped transient loading (Figure 3(a)) using =
0.01 and the coarsest mesh, compared with those from the time-domain SBFEM[15]. It can be seen
that the overall agreement is very good.

Figure 6 Dimensionless frequency a0 = 2fa/cs1-real part of KI curves using = 0.01.

Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531 529
Figure 7 Dimensionless frequency a0 = 2fa/cs1-imaginary part of KI curves using = 0.01.

Figure 8 Time histories of normalized CMOD using = 0.01.

Figure 9 Time histories of normalized DSIFs KI and KII using = 0.01.

530 Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531
7 Conclusions
This paper has developed a SBFEM-based frequency-domain method for calculating DSIFs at
bimaterial interface cracks subjected to transient loading. The recently developed Frobenius solu-
tion procedure in the frequency domain for solving the governing equations of SBFEM is used to
calculate the DSIFs KI and KII directly from the stress solutions, leading to complex fre-
quency-DSIFs response functions, which are subsequently used with the FFT and the IFFT to
calculate time histories of the DSIFs. Numerical results of a benchmark fracture problem demon-
strate that this method is capable of computing time histories of DSIFs at bimaterial interface
cracks accurately and effectively with a small number of degrees of freedom.

1 Chen W H, Wu C W. On elastodynamic fracture mechanics analysis of bimaterial structures using finite element method.
Engin Fract Mech, 1981, 15(1-2): 155168[DOI]
2 Geubelle P H. A numerical method for elastic and viscoelastic dynamic fracture problems in homogeneous and bimaterial
systems. Computat Mech, 1997, 20: 2025[DOI]
3 Marur P R, Tippur H V. Dynamic response of bimaterial and graded interface cracks under impact loading. Int J Fract, 2000,
103: 95109[DOI]
4 Dineva P, Gross D, Rangelov T. Dynamic behavior of a bimaterial interface-cracked plate. Engin Fract Mech, 2002, 69(11):
11931218[DOI]
5 Lira-Vergara E, Rubio-Gonzalez C. Dynamic stress intensity factor of interfacial finite cracks in orthotropic materials. Int J
Fract, 2005, 135: 285309[DOI]
6 Xu X P, Needleman A. Numerical simulations of dynamic crack growth along an interface. Int J Fract, 1996, 74: 289324 [DOI]
7 Nishioka T, Hu Q H, Fujimoto T. Component separation method of the dynamic J integral for evaluating mixed-mode stress
intensity factors in dynamic interfacial fracture mechanics problems. JSME Int J Ser A-Solid Mech Mater Engin, 2002, 45(3):
395406
8 Dornowski W, Perzyna P. Numerical analysis of macrocrack propagation along a bimaterial interface under dynamic loading
processes. Int J Solids Struct, 2002, 39: 49494977[DOI]
9 Lei J, Wang Y S, Gross D. Time-domain BEM analysis of a rapidly growing crack in a bimaterial. Int J Fract, 2004, 126(2):
103121[DOI]
10 Wolf J P, Song C M. Finite-element Modelling of Unbounded Media. Chichester: John Wiley and Sons, 1996
11 Wolf J P. The Scaled Boundary Finite Element Method. Chichester: John Wiley and Sons, 2003
12 Deeks A J, Chidgzey S R. Determination of coefficients of crack tip asymptotic fields using the scaled boundary finite
element method. Engin Fract Mech, 2005, 72(13): 20192036[DOI]
13 Song C M, Wolf J P. Semi-analytical representation of stress singularities as occurring in cracks in anisotropic
multi-materials with the scaled boundary finite-element method. Comput & Struct, 2002, 80(2): 183197[DOI]
14 Yang Z J. Fully automatic modelling of mixed-mode crack propagation using scaled boundary finite element method. Engin
Fract Mech, 2006, 73(12): 17111731[DOI]
15 Song C M. A super-element for crack analysis in the time domain. Int J Num Methods Engin, 2004, 61(8): 13321357 [DOI]
16 Song C M, Wolf J P. The scaled boundary finite-element method: Analytical solution in frequency domain. Comput Methods
Appl Mech Engin, 1998, 164(1-2): 249264[DOI]
17 Yang Z J, Deeks A J, Hao H. A Frobenius solution to the scaled boundary finite element equations in frequency domain. Int J
Num Methods Engin, 2007, 70(12): 13871408[DOI]
18 Deeks A J, Wolf J P. A virtual work derivation of the scaled boundary finite-element method for elastostatics. Comput Mech,
2002, 28(6): 489504[DOI]
19 Yang W, Suo Z, Shih C F. Mechanics of dynamic debonding. Proc Roy Soc London Ser A-Math Phys Engin Sci, 1991,
433(1889): 679697
20 Sun C T, Jih C J. On strain energy release rates for interfacial cracks in bimaterial media. Engin Fract Mech, 1987, 28(1):
1320[DOI]
21 Cooley J W, Tukey J W. An algorithm for the machine calculation of complex Fourier series. Math Comput, 1965, 19: 297
301[DOI]
22 MATLAB V7.1. The MathWorks, Inc, 2005

Z. J. YANG et al. Sci China Ser G-Phys Mech Astron | May 2008 | vol. 51 | no. 5 | 519-531 531

You might also like