You are on page 1of 6

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/231754516

Modeling of Solder Fatigue Failure due to


Ductile Damage
ARTICLE NOVEMBER 2010
DOI: 10.1017/S1727719100004627

CITATIONS

DOWNLOADS

VIEWS

106

3 AUTHORS, INCLUDING:
Yu-Lin Shen
University of New Mexico
127 PUBLICATIONS 1,897 CITATIONS
SEE PROFILE

Available from: Yu-Lin Shen


Retrieved on: 15 September 2015

Technical Note

MODELING OF SOLDER FATIGUE FAILURE DUE


TO DUCTILE DAMAGE
K. Aluru *
Department of Mechanical Engineering
University of New Mexico
Albuquerque, NM 87131, U.S.A.
F.-L. Wen **
Department of Mechanical and Computer-Aided Engineering
St. Johns University
Taipei, Taiwan 25135, R.O.C.
Y.-L. Shen ***
Department of Mechanical Engineering
University of New Mexico
Albuquerque, NM 87131, U.S.A.
ABSTRACT
A numerical study is undertaken to simulate failure of solder joint caused by cyclic shear deformation.
A progressive ductile damage model is incorporated into the rate-dependent elastic-viscoplastic finite
element analysis, resulting in the capability of simulating damage evolution and eventual failure through
crack formation. It is demonstrated that quantitative information of fatigue life, as well as the temporal
and spatial evolution of fatigue cracks, can be explicitly obtained.
Keywords : Solder joint, Fatigue, Ductile damage, Plastic deformation, Finite element analysis.

1.

INTRODUCTION

Fatigue failure of solder joints in microelectronic


packages has long been a reliability concern [1-3].
Aside from the traditional slow-rate thermo-mechanical
fatigue [4,5], solder joints may also experience severe
cyclic loading, such as vibrational fatigue, in many
military and civilian applications. Damage can be
generated by the excessive plastic deformation and/or
creep inside the solder alloy. For analysis and design
purposes, it is of interest to develop numerical simulation capabilities which can quantitatively predict failure
in solder.
Numerical prediction of solder fatigue typically involves the calculation of stress or plastic strain ranges,
which can then be substituted to the empirical Basquine
[6] or Coffin-Manson [7,8] types of fatigue criteria to
obtain the number of cycles to failure [2,9-14]. This
approach, while indirect, is largely necessitated by the
impracticality of simulating direct material failure in a
realistic setting. However, since numerical methodologies exist for modeling damage and failure in ductile
*

Graduate Student

**

Professor

***

metals, the present study thus aims at incorporating


failure into the finite element analysis and simulating
fatigue performance of a model joint structure. We
attempt to test the feasibility of applying this direct approach to address the solder fatigue problem.
The model utilizes a simple lap-shear test scheme.
A damage model is built into the analysis, which, in
conjunction with the element removal process, is capable of simulating direct failure of the solder material.
Attention is devoted only to the evolution of ductile
damage, without the possible interference of interfacial
or intermetallic fracture during deformation of the solder joint.

2.

NUMERICAL APPROACH

Figure 1 shows the computational model. The solder material is bonded to two copper substrates. The
width (w) and thickness (h) of the solder joint are taken
to be 1mm and 0.5mm, respectively.
The substrate

Professor, corresponding author

Journal of Mechanics, Vol. 26, No. 4, December 2010

N23

d p
e f ( p ) R
,
dt

Copper
H
w

solder

y
h

vx

x
W
Copper
Fig. 1

Schematic of the solder joint model along with


the boundary conditions (plot not to scale).
The applied velocity vx gives rise to the nominal shear deformation of the solder. The dimensions used in the finite element analysis
are: w 1mm, h 0.5mm, H 2.5mm and W
0.5mm

dimensions H and W are 2.5mm and 0.5mm, respectively. Note that W represents the part of substrate
outside the solder width. At each interface between
the solder and substrate, there is a 5m-thick intermetallic layer included in the model (not specifically shown
in Fig. 1). Deformation of solder is induced by prescribing a constant boundary velocity, vx, at the far right
edge of the lower substrate. This gives rise to a nominal shear deformation. For reversed loading back to
the zero nominal shear strain, the boundary velocity is
then vx. During deformation the x-direction movement of the far left edge of the upper Cu is forbidden,
but movement in the y-direction is allowed. The top
boundary of the upper Cu is fixed in the y-direction but
their x-movement is allowed. The bottom boundary of
the lower Cu is constrained to remain horizontal but its
nodal displacements in x and y are otherwise free. The
calculations are based on the plane strain condition,
which effectively simulates the nominal simple shearing
mode of the solder [15-17].
In the model the copper substrate is taken to be isotropic linear elastic, with Youngs modulus 114GPa,
Poissons ratio 0.31 and density 8930kg/m3. The
Cu6Sn5 intermetallic layer is also assumed to be isotropic linear elastic, with Youngs modulus 85.5GPa,
Poissons ratio 0.28 and density 8280kg/m3. The solder material, taken to be the Sn-1.0Ag-0.1Cu alloy [18],
is treated as an isotropic elastic-viscoplastic solid, with
Youngs modulus 47GPa, Poissons ratio 0.36 and density 5760kg/m3. Its yielding and strain hardening response follows the experimental stressstrain curves for
different strain rates [18]. At the strain rate of 0.005s1
or below, the initial yield strength is 20MPa; the flow
strength increases to a peak value of 36MPa at the plastic strain of 0.15, beyond which a perfectly plastic behavior is assumed. This slow-rate form is considered
as the static response. The rate-dependent plastic
flow strength follows

N24

(1)

where e is the von Mises effective stress, f (a function


of equivalent plastic strain p ) is the static plastic
stressstrain response, and R (a function of plastic strain
rate d p / dt ) defines the ratio of flow strength at
higher strain rates to the static flow strength where R
equals unity. Compared to rate-independent plasticity,
this formulation utilizes the scaling parameter R to
quantify the strain rate hardening effect. The R values of the solder alloy are 1.0, 1.9, 2.4, 2.8, 3.1, 3.4 and
3.5 at the plastic strain rates of, respectively, 0.005, 0.5,
6, 50, 100, 200 and 300 s1.
A progressive ductile damage model is utilized to
simulate failure of the solder alloy. Figure 2 shows a
schematic of the stressstrain curve which includes the
damage response (solid curve) [19]. The damage
process is quantified by a scalar damage parameter D,
with
(1 D) ,

(2)

where is the current flow stress and is the flow


stress in the absence of damage. In addition to leading
to softening of the plastic stress, damage is also manifested by the degradation of elastic modulus as shown
by the dashed unloading/reloading line in Fig. 2. The
equivalent plastic strains at the onset of damage (D 0)
and failure (D 1) are 0pl and fpl , respectively. A
material element loses its capability to carry stress when
its D attains unity, at which point the element will be
removed from the mesh so a void thus develops.
Cracking is then a consequence of linking multiple adjacent voids in the model.
In general 0pl can be made a function of the stress
triaxiality, hyd / e, where hyd = 1/3(xx + yy + zz) is
the hydrostatic stress and e is the von Mises effective
stress. In the present study 0pl is assumed to be independent of stress triaxiality due to the lack of experimental data that may be used for defining the functional
form. Upon damage initiation, strain softening and
thus strain localization set in, which displays a strong
mesh dependency. To alleviate the problem, a characteristic length L is used in the model, with
u p L p

(3)

where u p represents a plastic displacement quantity,


and L is defined as the square root of the integration
point area in each finite element. The softening phenomenon is now expressed as a stressdisplacement
relationship [20]. Prior to the initiation of damage,
u p 0 ; after damage initiation Eq. (3) starts to take

Journal of Mechanics, Vol. 26, No. 4, December 2010

D0

(undamaged response)

enced by solder joints. The nominal shear strain range


used in the simulation was arbitrarily chosen to be between 0 and 0.01, and the simulations were conducted
until the entire joint was severed.

0
3.

D1

(1D)E

Fig. 2

RESULTS AND DISCUSSION

pl
0

pl
f

Schematic showing the ductile damage response, in terms of the uniaxial stressstrain
curve

effect. Failure (removal of the element) occurs when


u p reaches the specified failure value, u fp . The
evolution of the damage parameter is taken to follow a
linear form,
D

up
.
u fp

(4)

The damage response is thus completely specified by


the two parameters 0pl and u fp . They are chosen to
be 0.18 and 3m, respectively [21]. It is noted that the
chosen value of u fp corresponds to a fpl value of
approximately 0.5. These parameters were largely
based on measured tensile stressstrain curves of bulk
pure Sn or Sn-rich solder alloy. It is understood that
bulk materials and actual solder joints have significantly
different physical sizes and microstructure and thus different constitutive responses. In addition, formation
and coalescence of microvoids tend to occur more readily under tensile loading. The present set of damage
parameters should be viewed only as a lower-limit approximation. Nevertheless, for the present work emphasizing qualitative comparison, the choice of damage
parameters has a benefit of reducing the number of fatigue cycles, thus computing time, needed to cause failure.
The finite element program ABAQUS [19] was employed for the modeling. The solder/intermetallic
structure was discretized into 5000 four-noded linear
elements, and each Cu substrate was discretized into
2400 elements. The mesh convergence was checked
by another set of preliminary calculations using twice
the number of elements in the model. In this study the
nominal shear strain rate imposed on the solder joint is
defined to be vx / h (see Fig. 1). We consider three
different nominal shear strain rates: 1s1, 10s1 and
100s1. It is noted that these strain rates are pertinent
to the range typical of vibrational fatigue, rather than
the much slower thermo-mechanical fatigue, experi-

Journal of Mechanics, Vol. 26, No. 4, December 2010

We first present a representative result of fatigue


cracking configuration. Figures 3(a) and 3(b) show the
contour plots of equivalent plastic strain along with the
cracks, after 11.5 and 16.5 cycles, respectively, of deformation under the 100s1 shear strain rate. For clarity only a small portion of each copper substrate is included in the figures. It can be seen that extensive
plastic deformation has occurred. The initiation of
cracks appears in the corner region of the solder joint.
Crack propagation follows the path of greatest equivalent plastic strain, which in turn is evolving during the
deformation and damage processes.
Figure 3(b) corresponds to the final failure state where
a major crack traversing the entire span of the solder joint
has just formed. Here two major cracks can be seen:
The upper one has traversed the entire joint and the lower
one also has propagated over a significant distance.
Therefore, although only one crack is responsible for the
final failure, the damage process itself is not associated
with only one dominant crack. We attribute this to the
steadily progressive accumulation of plastic strain during
the cyclic deformation, which leads to significant strain
rate hardening so the two main cracks had to take turns
to continue with the propagation process. It is worth
pointing out that, for the most part, the cracking path is
near and parallel to, but not at, the interface between the
solder alloy and the intermetallic layer. This is in consistence with experimental observations of ductile failure
in fatigued solder joints, as discussed in ref. [16].
In the present work two other applied shear strain
rates, namely 1 and 10s1, were also simulated. Similar cracking morphologies were obtained. Figure 4
shows the number of cycles required for the initiation of
fatigue cracks and final failure, as a function of the applied shear strain rate. It can be seen that crack
propagation generally involves more cycles than those
needed to initiate the first crack. Crack initiation is not
a strong function of applied strain rate. However, in
the case of 100s1 strain rate, the crack spends a much
shorter time (number of cycles) in traversing the entire
solder joint than the cases of slower rates. This observation suggests the detrimental effect on the life of
solder joint caused by high frequency cyclic loading.
The numerical results presented above have demonstrated the feasibility of incorporating damage models
for direct simulation of fatigue failure in solder joints.
It should be noted that the material damage parameters
chosen for the current work were based on an idealized
condition, that of a simple tensile stress-strain response
without consideration of the effect of stress triaxiality

N25

etc. Therefore the numerically predicted fatigue life


should only be considered as a lower-limit estimate.
For future investigations, it is suggested that the damage
parameters of solder alloys, needed as input for the
simulation, be experimentally characterized in detail.
The brittle fracture features, such as decohesion along
the interface between the solder and intermetallic layer,
can also be incorporated into the modeling framework.
A more comprehensive modeling capacity can thus be
facilitated for the quantitative prediction of solder failure.
(a)

4.

CONCLUSIONS

Numerical finite element modeling of solder joint fatigue was carried out. The analysis incorporated a
ductile damage model capable of simulating material
failure through the element removal process. It was
illustrated that the technique can be employed to directly model fatigue failure of solder joint induced by
cyclic shear. Fatigue cracks were seen to propagate
primarily near the interface but inside the solder. The
number of cycles to failure can be influenced by the
applied strain rate.

ACKNOWLEDGMENTS
(b)

Fig. 3

Contour plots showing the equivalent plastic


strain and crack profile for the case of 100s1
strain rate at (a) 11.5 cycles and (b) 16.5 cycles

This work was supported in part by the National


Science Council of Taiwan, under contract # NSC 982221-E-129-005 and contract # NSC 99-2811-E-129001.

REFERENCES
50
final fracture

Number of cycles

40
30
20

crack initiation

10
0
1

10

100

Strain rate (1/s)


Fig. 4

N26

Numbers of cycles for the initiation of fatigue


cracks and final failure, as a function of applied shear strain rates obtained from the modeling

1. Plumbridge, W. J., Solders in Electronics, Journal of


Materials Science, 31, pp. 25012514 (1996).
2. Lee, W. W., Nguyen, L. T. and Selvaduray, G. S.,
Solder Joint Fatigue Models: Review and Applicability
to Chip Scale Packages, Microelectronics Reliability,
40, pp. 231244 (2000).
3. Atluri, V. P., Mahajan, R. V., Patel, P. R., Mallik, D.,
Tang, J., Wakharkar, V. S., Chrysler, G. M., Chiu, C.-P.,
Choksi, G. N. and Viswanath, R. S., Critical Aspects
of High-Performance Microprocessor Packaging, MRS
Bulletin, 28, pp. 2134 (2003).
4. Lee, C. F., Lee, Z. H. and Qu, S. H., The Endochronic
Viscoplasticity for Sn/3.9Ag/0.6Cu Solder under Low
Strain Rate Fatigue Loading Coupled with Thermal Cycling, Journal of Mechanics, 25, pp. 261270 (2009).
5 Lee, C. F. and Lee, Z. H., Predicting Fatigue Initiation
Life of Sn/3.8Ag/0.7Cu Solder Using Endochronic Cyclic Damage-Coupled Viscoplastic Theory, Journal of
Mechanics, 24, pp. 369377 (2008).
6. Basquin, O. H., The Exponential Law of Endurance
Tests, Proceedings of the American Society for Testing
and Materials, 10, pp. 625630 (1910).

Journal of Mechanics, Vol. 26, No. 4, December 2010

7. Coffin, Jr., L. F., A Study of the Effects of Cyclic


Thermal Stresses on a Ductile Metal, Transactions of
ASME, 76, pp. 931950 (1954).
8. Manson, S. S., Behavior of Materials under Conditions
of Thermal Stress, Proceedings of the Heat Transfer
Symposium, University of Michigan Engineering Research Institute, pp. 975 (1953).
9. Hong, B. Z., Finite Element Modeling of Thermal Fatigue and Damage of Solder Joints in a Ceramic Ball
Grid Array Package, Journal of Electronic Materials,
26, pp. 814820 (1997).
10. Li, Y., Mahajan, R. L. and Subbarayan, G., The Effect
of Stencil Printing Optimization on Reliability of
CBGA and PBGA Solder Joints, Journal of Electronic
Packaging, 120, pp. 5460 (1998).
11. Gu, Y. and Nakamura, T., Interfacial Delamination and
Fatigue Life Estimation of 3D Solder Bumps in
Flip-Chip Packages, Microelectronics Reliability, 44,
pp. 471483 (2004).
12. Kim, Y. B., Noguchi, H. and Amagai, M., Vibration
Fatigue Reliability of BGA-IC Package with Pb-free
Solder and Pb-Sn Solder, Microelectronics Reliability,
46, pp. 459466 (2006).
13. Ghorbani, H. R. and Spelt, J. K., An Analytical ElastoCreep Model of Solder Joints in Leadless Chip Resistors: Part 2Applications in Fatigue Reliability Predictions for SnPb and Lead-free Solders, IEEE Transactions on Advanced Packaging, 30, pp. 695704 (2007).
14. Lee, C.-C., Lee, C.-C., Ku, H.-T., Chang, S.-M. and
Chiang, K.-N., Solder Joints Layout Design and Reliability Enhancements of Wafer Level Packaging Using
Response Surface Methodology, Microelectronics Reliability, 47, pp. 196204 (2007).

Journal of Mechanics, Vol. 26, No. 4, December 2010

15. Shen, Y.-L., Chawla, N., Ege, E. S. and Deng, X., Deformation Analysis of Lap-Shear Testing of Solder
Joints, Acta Materialia, 53, pp. 26332642 (2005).
16. Moy, W. H. and Shen, Y.-L., On the Failure Path in
Shear-Tested Solder Joints, Microelectronics Reliability, 47, pp. 13001305 (2007).
17. Shen, Y.-L., Externally Constrained Plastic Flow in
Miniaturized Metallic Structures: A Continuum-Based
Approach to Thin Films, Lines, and Joints, Progress in
Materials Science, 53, pp. 838891 (2008).
18. Wong, E.-H., Selvanayagam, C. S., Seah, S. K. W., van
Driel, W. D., Caers, J. F. J. M., Zhao, X. J., Owens, N.,
Tan, L. C., Frear, D. R., Leoni, M., Lai, Y.-S. and Yeh,
C.-L., StressStrain Characteristics of Tin-Based Solder Alloys for Drop-Impact Modeling, Journal of
Electronic Materials, 37, pp. 829836 (2008).
19. Abaqus 6.8, Users Manual, Dassault Systmes Simulia
Corp., Providence, RI.
20. Hillerborg, A., Modeer, M. and Petersson, P. E.,
Analysis of Crack Formation and Crack Growth in
Concrete by Means of Fracture Mechanics and Finite
Elements, Cement and Concrete Research, 6, pp.
773781 (1976).
21. Shen, Y.-L. and Aluru, K, Numerical Study of Ductile
Failure Morphology in Solder Joints under Fast Loading
Conditions, Microelectronics Reliability, in press.

(Manuscript received October 25, 2010,


accepted for publication November 12, 2010.)

N27

You might also like