Sie sind auf Seite 1von 19

J Mater Sci: Mater Electron (2013) 24:172190

DOI 10.1007/s10854-012-0720-y

REVIEW

Reliability behavior of lead-free solder joints in electronic


components
Liang Zhang Ji-guang Han Cheng-wen He
Yong-huan Guo

Received: 9 December 2011 / Accepted: 5 April 2012 / Published online: 19 April 2012
Springer Science+Business Media, LLC 2012

Abstract With more consumer products moving towards


environmentally friendly packaging, making solder Pb-free
has become an urgent task for electronics assemblies.
Solder joints are responsible for both electrical and
mechanical connections. Solder joint does not have adequate ductility to ensure the repeated relative displacements due to the mismatch between expansion coefficients
of the chip carrier and the circuit board. Materials behavior
of solder joints involves a creepfatigue interaction, making it a poor material for mechanical connections. The
reliability of solder joints of electronics components has
been found playing a more important role in service for
microelectronics components and micro-electro-mechanical systems. So many researchers in the world investigated
reliability of solder joints based on finite element simulation and experiments about the electronics devices, such as
CR, QFP, QFN, PLCC, BGA, CSP, FCBGA and CCGA,
which were reviewed systematically and extensively.
Synchronously the investigation on reliability of solder
joints was improved further with the high-speed development of lead-free electronic packaging, especially the
constitutive equations and the fatigue life prediction
equations. In this paper, the application and research status
of constitutive equations and fatigue life prediction equations were reviewed, which provide theoretic guide for the
reliability of lead-free solder joints.
L. Zhang (&)  J. Han  C. He  Y. Guo
School of Mechanical and Electrical Engineering,
Jiangsu Normal University, Xuzhou 221116, China
e-mail: zhangliang@nuaa.edu.cn
L. Zhang
Provincial Key Laboratory of Advanced Welding,
Jiangsu University of Science and Technology,
Zhenjiang 212003, China

123

1 Introduction
Recent years, investigations on lead-free soldering have
been conducted worldwide according to the implementation of the directive on Waste Electrical and Electronic
Equipment (WEEE) and the directive and the Restriction of
the Use of Hazardous Substances in Electrical and Electronic Equipment (RoHS Directive) [13]. Therefore, many
Sn-rich alloy systems have been developed as alternative
candidate lead-free solders, which have attracted much
researches and trial applications in recent years [4]. Among
the various lead-free solder alloy choices as replacements
for conventional SnPb solders, SnAgCu (SAC) alloys are
currently the most popular because of their relatively good
soldering performance, excellent creep resistance, and
thermal fatigue reliability, as well as their compatibility
with current electronic components, meanwhile, SnAg,
SnCu, SnZn et al. are all proposed to replace the SnPb
alloys [5, 6]. However, understanding reliability of leadfree solder joints is still in its infancy compared to traditional SnPb solders, thus, further studies are needed.
Semiconductor packages are exposed to various environments, such as temperature, humidity, dust, shock and
vibration, the main cause for failure of electronic package
is temperature fatigue (55 %) and vibration fatigue (20 %),
in addition, humidity (19 %) and dust (6 %) also contribute
to the failure of electronic devices [7]. Since solder joints
provide the mechanical and electrical interconnect between
the package and the board in the electronic devices, they
are susceptible to failures during thermal cycling [8].
Temperature fluctuations caused by either power consumption or environmental changes, along with the
resulting thermal expansion mismatch between the various
package materials result in deformation stress in packages/
assemblies especially in solder interconnects [9], and

J Mater Sci: Mater Electron (2013) 24:172190

accumulation of stress and strain may induce fatigue cracks


in solder joints and eventually cause the failure of the
whole devices [10, 11]. Consequently, the reliability of
lead-free solder joints is of the main critical issue when
selecting a package for a particular application in electronic industry.
This paper intends to summarize the reliability behavior
of lead-free solder joints in different electronic components, review the stressstrain response of lead-free solder
joints under thermal cycles and analyze the constitutive
models and fatigue life equations of the lead-free solder
joints and SnPb used in electronic packages.

2 Different electronic components


2.1 Resistor packages
In order to understand and appreciate how mean values of
the creep strain and creep energy densities were determined
consider Fig. 1 [12] which depicts the highly stressed areas
of the solder joint, the so called standoff, which is located
directly beneath the CR. These areas are prone to damage
and cracking, which will finally result in component failure. For this reason the irreversibly accumulated strains
were obtained for exactly these areas. For ceramic resistor
devices, Xue [13, 14] found that the shear strengths of
SnAgCu solder joints is greater than that of conventional
eutectic SnPb solder joints. Hegde [15] deals with creep
damage of SnAgCu lead-free solder joints for power
cycling using finite element analysis with the consideration
of experimentally observed non-uniform temperature distributions in the 1,206 surface mount chip resistor. In
addition, a comparison is made for inelastic strain accumulation and fatigue life for creep damage study for spatially uniform and non-uniform temperature power cycling.
Moreover, an accelerated thermal cycling test to assess the
reliability of SnAgCu lead-free solders at chip resistor has
been conducted by Han [16]. In Qis research [17], test
results indicate that the life of lead-free solder at chip

Fig. 1 FEM and corresponding microsection of a destroyed solder


joint [12]

173

resistor (6,432 mm type) would be worse than that of SnPb


solder in the accelerated thermal cycling condition
(125 C/-50 C). Three conditions with two ramp rates
(14 and 95 C/min) and two temperature ranges
(DT = 0100 and -40125 C) were applied to resistor
2,512 test vehicles assembled with SnPb and Pb-free solders. The test results showed that the higher ramp rate
reduced the testing time while retaining the same failure
modes, and that the damage per cycle increased with the
temperature difference. For the resistors, the Pb-free solder
joints lasted longer than the SnPb joints at the smaller DT,
but were inferior at the larger DT.
2.2 QFP/QFN/PLCC/SOJ devices
Quad flat package (QFP) is one of the most common
package types in surface mount technology (SMT) which is
used in many large scale integration (LSI) and very large
scale integration (VLSI) [18]. Three dimensional (3D)
nonlinear finite element simulation methods were utilized
to analyze the stress-strain response in Sn37Pb and
Sn3.5Ag solder joints in QFP100 devices [19]. The simulated results indicate creep distribution of solder joints is
not uniform, the heel and toe of solder joints, the area
between solder joints and leads are the creep concentrated
sites. The creep strain value of Sn3.5Ag solder joints is
lower than that of Sn37Pb solder joints. And the reliability
of SnAgCu solder joints is higher than that of SnPb solder
joints [20]. A quarter of the whole QFP256 was selected for
finite element modeling using hexahedron elements in the
FE analysis due to the symmetry in geometry by Zhang and
Xue [21] as shown in Fig. 2. The simulated results indicate
that SnAgCuCe solders show a superior anti-creep property
over SnAgCu alloys. Using the symmetric geometry,
quarter models with sub-model were used for QFP176
assemblies as shown in Fig. 3 [22]. It is found that the
fatigue life of PQFP components is about 3 times as fatigue
life of PBGA component. The optimum simulation for QFP
devices is also researched [23]. The results indicate that

Fig. 2 Finite element model for QFP256 [21]

123

174

Fig. 3 Quarter FE global and submodel for QFP176 [22]

when the lead widths increase, while the reliability of the


solder joints reduces. When the lead widths are the same,
the maximum equivalent stress of the solder joints does not
decrease completely with the increasing of lead pitches, a
minimum value of the maximum equivalent stress value
exists in all the curves. Under this condition the maximum
equivalent stress of the solder joints is relatively the least,
the reliability of solder joints is high and assembly is
excellent. The simulating results indicate the best parameter: the lead width is 0.2 mm and lead pitch is 0.3 mm,
which are benefited for the reliability of QFP devices now.
In addition, Zhang [24] found that the maximum stress of
QFP lead-free solder joint concentrates on the sharp corner
of interior part of the joint with FR-4 substrate, and the
maximum stress focuses on the sharp corner of exterior
part of the joint with LTCC substrate and the largest stress
concentrated area is at the sharp corner between the solder
joint and lead with PTFE substrate.
Quad flat non-lead (QFN) packages are getting popular
as a low-cost solution for applications with low pin-count
requirements. QFN is also known as micro-lead frame
(MLF), micro-lead package (MLP), quad outline non-lead
(QON), or small outline non-lead (SON) package [25]. It is
a type of chip scale package and can be assembled and
molded in matrix form for saving costs. For quad flat nonlead (QFN) packages, board-level solder joint reliability
during thermal cycling test is a critical issue. A parametric
Fig. 4 Correlation of failure
analysis and SED distribution of
failed peripheral solder joint
[26]

123

J Mater Sci: Mater Electron (2013) 24:172190

3D FEA sliced model is established for QFN on board with


considerations of detailed pad design, realistic shape of
solder joint and solder fillet, and non-linear material
properties [26]. It has the capability to predict the fatigue
life of solder joint during thermal cycling test within
34 % error. The fatigue model applied is based on a
modified Darveauxs approach with non-linear viscoplastic
analysis of solder joints. A solder joint damage model is
used to establish a connection between the strain energy
density (SED) per cycle obtained from the FEA model and
the actual characteristic life during thermal cycling test.
For the test vehicles studied, the maximum SED is
observed mostly at the top corner of peripheral solder joint
as shown in Fig. 4. The modeling predicted fatigue life is
first correlated to thermal cycling test results using modified correlation constants, curve-fitted from in-house QFN
thermal cycling test data. Subsequently, design analysis is
performed to study the effects of 17 key package dimensions, material properties, and thermal cycling test condition. Generally, smaller package size, smaller die size,
bigger pad size, thinner PCB, higher mold compound CTE,
higher solder standoff, and extra soldering at the center pad
help to enhance the fatigue life. For QFN package, it seems
that the lead-free SnAgCu gives a much higher solder
reliability than its SnPb alternative [27].
The appearance of PLCC (Plastic Leaded Chip Carrier)
is similar to that of QFP devices. Zhang [28] conducted a
study on SnPb and lead-free solder joints reliability of
PLCC devices by finite element simulation, the results
indicated that the largest stress was at the area between the
solder joints and the leads, and analysis results showed that
the von Mises stress at the location slightly increased with
the increasing of lead counts. A comprehensive experimental and numerical study of solder joints for plastic
leaded chip carrier (PLCC) 84-pin, 1.27 mm pitch was
carried by Chen [29]. It was found that the Pb-rich phases
segregated gradually and formed a continuous layer adjacent to the intermetallic compound (IMC) layer with
increasing the number of thermal cycles, resulting in cracks
near the solder/lead interface. The response of stress and

J Mater Sci: Mater Electron (2013) 24:172190

strain was studied using nonlinear finite element method


(FEM), and results agreed well with the experimental data.
Furthermore, finite element method was used to simulate
the solder joint of J-lead of the SOJ (Small Outline J-lead
Package) by Wu [30, 31]. Results indicate that the distortion is obvious, the ceramic carrier with up-warp trend, the
J-lead with forth extension trend, but the PCB with little
distortion because of the rigidity displacement restrict. The
strain of the J-lead solder joint soldered heel is bigger than
the soldered toe and the center place has the least strain.
The stress of the J-lead soldered heel is bigger than the
other place of the solder joint and the stress distributes a
wide area, where the weakest place of the whole solder
joint, so it is easy to lose efficacy. The stress of the J-lead
soldered toe is less than the soldered heel, and the stress
distributing area is small. On the contrary, the J-lead solder
joint center place distributes a large area and the stress
value is one order less than the stress value of soldered heel
and soldered toe, so it is the safety place of the solder joint.
The calculated values are in agreement with the experimental results.
2.3 BGA/FCBGA/CSP devices
The ball grid array (BGA) is the prevalent interconnect
structure between electronic packages and printed circuit
boards (PCBs) [32]. BGAs provide structural and thermal

175

connection as well, acting as electrical-circuit components.


An energy partitioning damage evolution (EPDE) model is
proposed to describe the viscoplastic damage evolution.
Application of the EPDE model is demonstrated for
SnAgCu solder joints during thermal cycling of a ball grid
array (BGA) electronic assembly. The FEA analysis predicts the critical solder joint (joint that experiences maximum damage and first failure) to be at the corner of the
package, which agrees with experimental results. It also
predicts the initiation site to be at the component side,
which also was confirmed by experiment as seen in Fig. 5
[33]. In order to make all ball grid array (BGA) packages
on the test board subject to the uniform stress and strain
level during drop impact, a test board in round shape is
designed to conduct drop tests. In simulation, FEM is used
to study transient dynamic responses [34]. The peeling
stress which is considered as the dominant factor affecting
the solder joint reliability is used to identify location of the
failed solder joints. Simulation results show very good
correlation with experiment measurement in terms of
acceleration response and strain histories in actual drop
test. Crack occurred at intermetallic composite (IMC)
interface on the package side with some brittle features.
The position of maximum peeling stress in finite element
analysis (FEA) coincides with the crack position in the
cross section of a failed package, which validated the FEA.
Board-level reliability of low-temperature co-fired ceramic

Fig. 5 Crack observed in the


solder, close to pad. a Contour
plot of damage in local model of
critical solder joint: front view
(b), isometric view (c) [31]

123

176

(LTCC) modules with thermo-mechanically enhanced ballgrid-array (BGA) solder joint structure mounted on a
printed wiring board (PWB) was experimentally investigated by thermal cycling tests in the 0100 and
-40125 C temperature ranges [35]. After the thermal
cycling tests were completed, the test modules were
investigated by SEM. The investigation verified that the
primary failure mechanism was solder fatigue cracking in
both (all) test sets. It can also be noticed that the Sn90Pb
sphere joints are deformed whereas the PCSB joints have
retained their initial shape. The degree of deformation of
Sn90Pb solder joints correlated mainly with the magnitude
of the exposed temperature extremes and the joint distance
from the module neutral point. In the milder test condition,
secondary cracks appeared within the PCSB solder joints,
as shown in Fig. 6. Qu [36] found that reveal that the
failures mainly occur in the solderPCB interface in lower
load level based on experiment and simulation, the other
way round, in a higher load level, the cracks are more
possibly formed in solderpackage interface; comparing to
dropping in horizontal direction with package faces down,
the solder joints are much harder to fail when dropping in
vertical direction. An optimal height and smaller pitch of
solder joints lead to lowest SED and best reliability in the
drop test.
The need of ultra-dense and high I/O for integrated circuit
designs has driven the development of flip-chip technologies
in the contemporary electronic packaging industry. For flipchip ball grid array (FCBGA) package assemblies, the solder
interconnections serve as electrical and thermal paths as well
as structural supports [37]. Anand model was used to
establish the constitutive equation of Sn3.0Ag0.5Cu solder
by Zhang and Xue [38], the stress distribution of solder joints
was analyzed with underfill and without it. It is indicated that

Fig. 6 Fatigue failures in a test module with PCSBs after thermal


cycling at 0100 C [35]

123

J Mater Sci: Mater Electron (2013) 24:172190

the stress concentrates on the top surface of outermost solder


joint whether the underfill is here or not. Due to the using of
underfill, the stress decreases and distributes evenly on the
top surface of solder joint. The Modified Coffin-Mason
equation by Engelmaier was utilized to predict the fatigue
life of solder joint, it is found that the fatigue life of solder
joint with underfill is longer than that without it. The effects
of underfill properties parameters were investigated, it is
indicated that the CTE of underfill influences strongly the
fatigue life of solder joints, but the influence is little for
Youngs Modulus, which will provide a theory guide for
practical applications. Warpage issues are of concern for flip
chip Ball Grid Array (FCBGA) package reliability on thin
substrates [39]. Despite advantages in electrical performance
of thin substrates, FCBGA warpage hampers reliability of
the solder balls making up the ball grid array. FEM calculations of thermal deformation and equivalent plastic strain
of solder balls considered to determine overall reliability of
the solder balls. Solder ball plastic strain decreased with
substrate core thickness. Substrate surface warpage fluctuated significantly as core thickness decreased due to reduced
stiffness. In addition, it is possible that solder balls and surface mounted components can be affected by surface fluctuations. Although thinner substrates have advantages in
terms of electrical performance, high warpage is a critical
root cause of package failure. In addition, since the potential
occurrence of brittle solder joint on FCBGA substrates was a
big concern, there was a strong need to develop a suitable
metrology to detect brittle solder joints. Figure 7 [40] shows
a schematic of different failure modes for the peel test. There
are four different failure modes that can be observed during
the peel test-pad cratering in the substrate, brittle fracture at
the NiP layer in the solder joint at the substrate side, ductile
fracture in the solder ball, and pad cratering in the test coupon
(or mother-board).
The effect of PWB protective coatings and pad structure
(via-in-pad and no via-in-pad) on the reliability of chip
scale package (CSP) interconnections will be investigated
under the drop test conditions. The failure analysis revealed

Fig. 7 A schematic of different failure modes for the peel test [40]

J Mater Sci: Mater Electron (2013) 24:172190

four different modes [41]: (1) fracture in the component


side the CuNiSn layer, (2) fracture in the Cu Sn layer on
the PWB side (only in the Cu OSP assemblies), (3) fracture
below the Cu Ni Sn layer on the PWB side (only in the Ni/
P/Au assemblies), and (4) PWB failure in the RCC buildup
underneath the soldering pads. Modes (2) and (4) did not
cause the components to fail electrically. Thermomechanical stress and strain analysis was conducted by Gonzalez
[42], the trends obtained from modeling results have a good
agreement with the experimental data found in literature,
results also indicate that lead-free alloys can be used as
alternative solder to improve the resistance to fatigue when
compared with standards lead containing solder. Moreover,
effects of high-temperature storage (HTS) aging on the
fracture performance and joint performance of low-Ag
solders/UBM (Ni(P)-Au) are examined for wafer-level
chip-scale packages (WLCSP) [43]. For the Sn1Ag0.5Cu
(SAC 105), the formation and evolution of IMC is different
from SAC 101 through aging. Figure 8a, b shows the
evolution of IMC in SAC105/UBM before aging and after
200 C aging 280 h, respectively. Results shows that HTS
aging results in the change of IMC from one layer (Fig. 8a)
(Cu,Ni)3Sn4 IMC to two different layers IMC (Fig. 8b). In
Fig. 8b, lift part is (Cu,Ni)6Sn5 near bulk solder and the
right part is (Cu,Ni)3Sn4 near UBM. A boundary appears
clearly between the two phases. New phase (Cu,Ni)6Sn5
was formed during HTS between the (Cu,Ni)3Sn4 layer and
solder bulk. An empirical model for estimating the reliability of CSP solder joints is derived by correlating the
simulated strains to thermal cycling results for 20 different
sample configurations [44]. This empirical model translates
the inelastic strains calculated by nonlinear three dimensional (3D) finite element simulations into reliability estimation (N50 % or N100 ppm). By comparing with the results
of reliability tests, it can be concluded that this model is
accurate and consistent for analyzing the effect of solder
joint geometry. These parameters can have a tolerance
(=accuracy) of their nominal value, and it is shown that
these small tolerances can have a significant influence on
the solder joint reliability. To evaluate the various solder
balls in CSP package application, lead-free SnAg-X
(X = In, Cu, Bi) and Sn9Zn1Bi5In solder balls were
characterized by melting behavior, phases, interfacial
reaction, and solder joint reliability based on Yoons works
[45]. All testes lead-free solders showed better board level
solder joint reliability than Sn36Pb2Ag. Sn3.5Ag0.7Cu and
Sn9Zn1Bi5In solders showed 35 %, 100 % superior solder
joint reliability than Sn36Pb2Ag solder ball, respectively.
A three dimensional (3D) nonlinear finite element model of
an overmolded chip scale package (CSP) on flex-tape
carrier has been developed by Mertol [46]. Predicted
results indicate that for an optimum design, that is low
stress in the package and low package warpage, the

177

Fig. 8 The evolution of IMC in SAC105/UBM [43]

package should have smaller die with thicker overmold.


Meanwhile, Pang [47] investigated the effects of employing different two-dimensional (2D) and three-dimensional
(3D) finite element analysis (FEA) modes for analyzing the
solder joint reliability performance of a flip chip on board
assembly. The thermal fatigue life of Sn3.5Ag lead-free
solder joints of a chip scale package subjected to a thermal
cycling loaded is predicted by Xiao [48]. And Wang [49]
investigated the stress-strain response of solder joints in
CSP device, the result show that temperature distribution
and equivalent stress of solder joints with two different
models are basically same. The difference between the two
numbers of failure cycles is not more than 5 %.
2.4 CCGA devices
In many applications such as computers and telecommunications, the IC chip sizes are very large, the on-chip
frequency and power dissipation are very high, and the

123

178

number of chip I/Os is very large. The CCGA (ceramic


column grid array) package developed by IBM is one of the
best candidates for housing these kinds of chips [50, 51].
For the CCGA assembly with SnAgCu and SnPb solder
fillets based on finite element simulation, the thermalfatigue life of the solder columns with lead-free solder
fillets should be shorter than that with the SnPb solder
fillets, since the maximum creep strain energy per cycle in
the solder column with the lead-free solder paste is larger
than that with the SnPb solder paste. ANSYS FEA software
was employed to establish three-dimensional strip model of
CCGA624 device and optimized simulation of soldered
column dimension was also studied [52]. Results indicated
that the most dangerous soldered column located in the
furthest distance from the device center, the strain concentration occur in and around the interface between the
eutectic solder and the Sn3.5Ag solder column where will
be the foremost location for appearance of crack. Optimization results of soldered column dimensions show that the
value of strain increase with the increase of column pitch;
with increase of soldered column height the strain curves
present parabolic-shaped and can acquire the minimum
equivalent strain in the height of 2.07 mm; the value of
strain increase with the increase of soldered column
diameter and the curve shows a clear monotone trend. In
practical applications, we can choose proper soldered column dimension based on the principle of smaller strain.
The dominant deformation mode is bending of the column
due to thermal expansion mismatch between the module
and the printed circuit board (PCB) [53]. The results are
compared with those of tinlead CCGA tested under similar conditions. Unlike tinlead columns, where the failure
occurs at the column near the top of the solder fillet and
through the thickness of the column, in the CuCGA, the
failure is found to occur first in the solder fillet at the
solder/copper column interface and the crack propagates
along the periphery of the copper column. The accumulated plastic deformation per cycle is larger in tinlead
columns compared to the copper columns. Moreover,
thermal cycling behavior of ceramic packages (CCGA560)
commonly used for the aerospace applications was studied
by Ghaffarian [54]. Based on numerical FEM models
(Fig. 9) and experiments, the high cycle solder joint fatigue
life for the 90Pb10Sn solder in a CCGA electronic component was determined by Perkins [55]. The location of
solder joint failures was identified through a dye-and-pry
analysis and correctly predicted with the numerical FEM.
The solder joints close to the clamped boundaries experience the greatest stress amplitude and fail first. The
mechanism of failure was observed to be a microcrack
initiation at phase boundaries followed by microcrack
coalescence into macrocracks. The fatigue life of a CCGA
can be greatly enhanced by not using the solder joints

123

J Mater Sci: Mater Electron (2013) 24:172190

Fig. 9 The finite element model [55]

closest to constrained boundary conditions for electrical


purposes. Ceramic column grid array (CCGA) packages
have become increasing popular as an alternative to ceramic ball grid array (CBGA) packages for applications
requiring more than 625 interconnections [56]. As package
body size increases beyond 32.5 mm, the column structure
offers an essential improvement to thermal fatigue life of
the package solder joint for many application. Microelectronic packages are subjected to a combination of thermal
cycling, power cycling, and dynamic mechanical loading
such as shock and vibration besides humidity, pressure, and
other conditions [57, 58]. Therefore, effects of such multiple loading conditions on CCGA solder joints reliability
need to be assessed by applying the loads concurrently and/
or sequentially.

3 Constitutive models
3.1 Anand model
Anand and Brown et al. [5961] firstly proposed a simple
set of constitutive equations for materials with large, isotropic elastic-viscoplastic deformations but small elastic
deformations. There are two basic features in this Anand
model [62, 63]. First, this model needs no explicit yield
condition and no loading/unloading criterion. The plastic
strain is assumed to take place at all nonzero stress values,
although at low stresses the rate of plastic flow may be
immeasurably small. Second, this model employs a single
scalar as an internal variable to represent the isotropic
resistance to plastic flow offered by the internal state of the
material. For applying the Anand to describe the inelastic
deformation of solders, the physical mechanisms such as
temperature sensitivity, strain hardening, strain rate sensitivity, strain rate, and temperature history effects, etc.
should be taken into account.

J Mater Sci: Mater Electron (2013) 24:172190

179

The averaged isotropic resistance representing the isotropic strengthening mechanisms such as dislocation density, solid solution strengthening, subgrain, and grain size
effects is characterized by the internal variable s, which
depends on the temperature and rate-dependent strain history of the materials [64]. The flow function of the Anand
model is in hyperbolic sine shape:

 1=m
 
nr
Q
e_ P A sinh
exp
1
S
kT

From the above Anand constitutive model, there are


nine material parameters, the nine material constants A, Q,
_
n, M, s , n, ho, a and So can be determined directly from the
experimental data. The material parameters of Anand
constitutive relation for different lead-free solders and
conventional SnPb alloys are determined based on fitting
procedure by many researchers are listed in Table 1.
3.2 Steady-state creep model

The evolution equation for the internal variable s is


assumed to be of the form as


a B
s_ h0 jBj
2
e_ P
j Bj
s
B1 
3
s

 n
_p
Q
_ e
exp
s s
4
kT
A

Characteristics of time-dependent deformation and failure


for the electronics components are a major concern for
stability and safety. Creep under constant stress and
relaxation under constant strain are two typical timedependent phenomena of materials [80]. Because of the
low melting points of both lead-free and SnPb solders, even
room-temperature corresponding to a high homologous
temperature of the more than 0.5Tm at which, creep is the
most important deformation mechanism. Figure 10 [81]
shows a typical creep curve for lead-free solders, which
generally consists of three stages after the initial instantaneous strain when a constant load is applied. At the instant
the force is applied, the solder will experience a strain,
which is part elastic and part plastic. The elastic part can be
predicted given the Youngs Modulus of the solder as
discussed earlier. The plastic part is the instantaneous

where So, initial value of deformation resistance; Q, activation energy; R, universal gas constant; A, pre-exponential
factor; n, multiplier of stress; m, strain rate sensitivity of
_
stress; ho, hardening/softening constant; s , coefficient for
deformation resistance saturation value; n, strain rate sensitively of saturation (deformation resistance) value; a,
strain rate sensitively of hardening or softening.
Table 1 Material parameters of Anand constitutive relation
Solder alloys

A/s-1

Q/R/J mol-1

Sn3.8Ag0.7Cu [65]
Sn3.8Ag0.7Cu0.03Ce [65]

24300
21200

8710
8026

n
5.8
5

s /MPa

0.183
0.130

65.3
57.6

0.019
0.0175

52.4

Sn3.5Ag [66]

177016

85459

0.207

Sn3.8Ag0.7Cu [67]

15.773

9883

1.0673

0.3686

Sn1.0Ag0.5Cu [67]

3.773

8076

0.9951

0.4454

Sn3.0Ag0.5Cu [68]

717.260

6067

0.130

3.1505
3.5833
29.0

0.0177
0.0352
0.0120
0.0436

ho/MPa
3541.2
4352.6
27782

So/MPa

1.9
2.3

39.5
28.5

1.6

0.0673T ? 28.6

1076.9

1.6832

3.2992

4507.5

2.1669

2.3479

2.22

2.45

14560

Sn3.5Ag [68]

344.716

6538

0.143

26.0

0.0447

23241

1.46

0.65

Sn0.7Cu [68]

764.166

5837

0.123

26.4

0.0430

11656

2.33

4.43

Sn37Pb [69]

6220

6525

3.33

0.27

36.86

0.022

60599

1.7811

3.1522

Sn3.8Ag0.7Cu [70]

107.65

7619

59.36

4.03

86.28

0.0046

9002

1.30

22.64

Sn3.5Ag [71]

52690

8765

0.182

76.944

0.018

7205.18

1.2321

30.36

Sn37Pb [72]

26

5797

10

0.25

83.12

0.043

Sn36Pb2Ag [73]

23000000

11262

11

0.303

80.79

0.0212

Sn3.5Ag [73]

22300

8900

0.182

73.81

0.018

Sn4.0Ag0.5Cu [74]

325

10561

10

0.32

42.1

0.02

Sn3.8Ag0.7Cu [75]
Sn40Pb [76]

2.23
14900000

8900
10830

6
11

0.182
0.303

73.81
80.415

0.018
0.0231

92148
4121.31
3321.15
800000
3321.15
2640.75

37.9
42.32

1.82

39.09

2.57

20

1.82
1.34

39.09
56.33

Sn97.5Pb [77]

3.25E12

15583

0.143

72.73

0.00437

3.73

15.09

Sn3.8Ag0.7Cu [78]

862

15516

25.57

0.55

18.7

0.034

195109

2.02

28.81

Sn8.5Sb [78]

2.23E8

13816.8

13.85

0.417

76.0

0.01

291215

2.34

48.15

Sn92.5Pb2.5Ag [79]

105200

11010

0.241

41.63

0.002

1.3

33.07

1787.02

1.24
1.38

1432

123

180

J Mater Sci: Mater Electron (2013) 24:172190

Garofalo-Arrhenius model:


Q
n
e_ cr C sinhar exp 
RT
Dorn model:


Q
e_ cr Arn exp 
RT

Fig. 10 Diagram showing the response of solder during a creep test


[81]

plasticity, Darveaux et al. [82] and Wiese and Rzepka [83]


offer laws to predict this, although the distinction between
this instantaneous plasticity and the following primary
creep region is not well defined. As the test continues the
strain increases, first rapidly (primary region) and gradually
slowing to a steady strain rate (secondary region). This
additional strain is due to creep. After the secondary region
come the tertiary region during which the strain rate
increases until rupture. This occurs due to both necking
(shrinking of the cross section resulting in increased stress)
and damage (cracking) occurring in the solder. However,
the secondary, or steady-state creep is the dominant
deformation experienced by solder alloys [84]. In this
stage, the strain rate is retarded by strain hardening, which
decreases the deformation speed, while the associated
recovery and recrystallization (softening) tend to accelerate
the creep rate. The steady-state creep rate can be quantitatively estimated, and a series of constitutive models have
been proposed. The following two models are the most
widely accepted for the characterization of solder alloys by
considering the diffusion controlled creep deformation
mechanism (Fig. 11).

(a)

(b)

Fig. 11 Dislocation motion in the two stress regimes [108]. a LSR,


b HSR

123

where e_ cr is the creep rate, r is the applied stress, n is the


stress exponent, C and A are constants relevant to the
microstructure, Q is the creep-active energy, R is the gas
constant, T is absolute temperature. Q and n are two
important mechanical and thermodynamic parameters for
describing the creep deformation of solders. Equations (5)
and (6) can be further simplified as Eqs. (7) and (8).


C4
_ecr C1 sinhC2 rC3 exp 
7
T


C3
8
e_ cr C1 rC2 exp 
T
Based on above steady-state constitutive model, these
material parameters can be determined directly from the
experimental data and fitting. The material parameters of
the constitutive relation for different lead-free solders and
conventional SnPb alloys are determined based on fitting
procedure by many researchers are listed in Tables 2 and 3.
Creep deformation mechanisms are identified by the values
of the stress exponent n and the activation energy Q [85].
Dislocation creep occurs as a result of glide and climb of
edge dislocations.
3.3 Double power law creep model
Recent years, many efforts were made by researchers to
explore the constitutive response of the lead-free solders
and SnPb alloys during the creep. Among the proposed
constitutive equations, there are two widely accepted
models [102]. One is the hyperbolic sine law which was
suggested by Schubert et al. [103] to represent the high
stress region as a power law break down region. The other
is the double power law which was formulated by Wise
et al. [104] to describe the steady state creep of the solder
by clime controlled behavior at low stress and the combined glide/climb behavior at high stress (Eq. 9). For the
eutectic SnPb was found [105] that the creep behavior at
low and medium stresses can be divided into three distinct
regions: region I (low strain rates) = NabarroHeering
creep; region II (intermediate strain rates) = superplastic
behavior; region III (high strain rates) = climb controlled
dislocation creep. Although experimental results for region
I are contradictory, there is agreement about the creep
behavior in regions II and III, which is usually expressed

J Mater Sci: Mater Electron (2013) 24:172190

181

Table 2 Creep parameters for different solders (Eq. 7)


C1(1/s)

Solder alloys

C2(1/MPa)

C3

C4(K)

Sn1.3Ag0.2Cu0.05Ni [86]

8E ? 5

0.1

9020

Sn1.3Ag0.5Cu0.05Ni [86]

6E ? 4

0.11

8419

Sn2.7Ag0.4Cu0.05Ni [86]

2E ? 7

0.143

12026

Sn80Au [87]

4.62E ? 15

2E-5

2.07

12267

Sn3.5Ag0.7Cu [88]

4.61E ? 6

0.037

6.17

8400

Sn3.8Ag0.7Cu [17]

926(508-T)/T

(37.78 9 106-74414T)-1

3.3

6360

Sn3.5Ag [89]

98437

0.103

6.65

9561

Sn5Sb [89]

42409

0.04

4.5

6542

Sn9Zn [89]

81524

0.06

4.57

7841

Sn3.0Ag0.5Cu [90]

6.385E-6

0.08638

5.84

Sn3.8Ag0.7Cu [91]
Sn3.8Ag0.7Cu [92]

3.2E ? 4
441000

0.037
5E-9

5.1
4.2

159.97
6524
5412

Sn3.9Ag0.6Cu [93]

500000

0.01

5802

Sn3.5Ag [94]

18(553-T)/T

145.036 9 [1/(6386-11.55T)]

5.5

5802

Sn3.9Ag0.6Cu [95]

441000

0.005

4.2

5412

Sn36Pb2Ag [95]

462(508-T)/T

145.036 9 [1/(5478-10.79T)]

3.3

6360

100In [94]

40647(593-T)/T

145.036 9 [1/(274-0.47T)]

8356

Sn3.8Ag0.7Cu [96]

325000

0.05217

5.3

5800

SnAgCu0.03Ce [96]

284000

0.02432

6.1

6400

Sn3.5Ag [97]

900000

0.0653

5.5

8690

Sn37Pb [98]

10

0.2

5400

Table 3 Creep parameters for different solders (Eq. 8)


Solder alloys

C1(1/s)

C2

C3(K)
12267

Sn80Au [87]

8.37E ? 5

2.07

Sn3.8Ag0.7Cu [95]

2.6E-5

3.69

4330.046

Sn3.8Ag0.7Cu [96]

1.5E-9

8.2

8538.78

Sn3.8Ag0.7Cu0.03Ce [96]

6.2E-11

8.0

9621.17

Sn3.5Ag [98]

5E-6

11

1179

Sn4.0Ag0.5Cu [98]

2E-21

18

9994

Sn3.5Ag [99]

7.087E-8

5.5

Sn3.5Ag [100]

6.62E-3

12

13048.7

Sn9Zn [100]

9.27

8.1

12014.4

Sn3.5Ag2Bi [101]

0.679

3.304

4630.186

7357.907

by double power law equation [106]. Song et al. [107]


investigated the creep behavior of three tin-rich solders that
have become candidates for use in lead-free solder joints:
Sn3.5Ag, Sn3Ag0.5Cu, Sn0.7Cu. Their steady-state creep
rates are separated into two regimes with different stress
exponents. The low-stress exponents range from *3 to 6,
while the high-stress exponents are anomalously high
(712). Strikingly, the high-stress exponent has strong
temperature dependence near room temperature, increasing
significantly as the temperature drops from 95 to 60 C.
The anomalous creep behavior of the solders appears to be
due to the dominant tin constituent. Research on creep in

bulk samples of pure tin suggests that the anomalous


temperature dependence of the stress exponent may show a
change in the dominant mechanism of creep.

 n1

 n2
Q1
r
Q2
r
e_ cr A1 exp
A2 exp
9
rn
rn
RT
RT
where e_ cr is the steady state creep strain rate, A a constant,
Q the activation energy, R the Boltzmanns constant, T the
absolute temperature, r the applied stress, rn the deformation resistance relative parameter and n is the stress
component The subscript 1 denotes the parameters for
creep behavior at low stresses while the subscript 2 denotes
the parameters for creep behavior at high stresses.
The constitutive behavior for creep performance of
Sn3.5Ag0.7Cu lead-free solder joints was investigated by
Han [108]. It was observed that the stress exponent (n) can
be well defined into two stress regimes: low stress and high
stress. A new, improved constitutive model, which considered back stress, was proposed to describe the creep
behavior of SnAgCu solder joints. In this model, the back
stress, which is a function of the applied shear stress in the
low stress regime (LSR) and a function of the particle size,
volume fraction and coarsening of IMC particles in the
high stress regime (HSR), was introduced to construct the
relationship between the creep strain rate and the shear
stress. The creep mechanism in these two stress regimes
was studied in detail In the LSR [109], dislocations passed

123

182

through the matrix by climbing over the intermetallic


particles, while in the HSR [110], the dislocations were
glide-controlled. The double power law creep model of
Sn3.8Ag0.7Cu and Sn3.8Ag0.7Cu0.1RE solders by Chen
[111]. Results indicated that at low stress, the true creepactivation energy of SnAgCu and SnAgCuRE solder joints
is close to the lattice self-diffusion activation energy, so the
steady-state creep rates of these two solder joints are both
dominated by the rate of lattice self-diffusion. While at
high stress, the true creep-activation energy of SnAgCu and
SnAgCuRE solder joints is close to the dislocation-pipe
diffusion activation energy, so the steady-state creep rates
are dominated by the rate of dislocation-pipe diffusion. At
low stress, the best-fit stress exponent n of SnAgCu and
SnAgCuRE solder joints are 6.9 and 8.2, respectively, and
the true creep-activation energy of them both is close to
that of lattice self-diffusion. At high stress, it equals 11.6
and 14.6 for SnAgCu and SnAgCuRE solder joints,
respectively, and the true creep-activation energy for both
is close to that of the dislocation-pipe diffusion. Reynolds
[112] studied the creep behavior of In-Ag eutectic solder
joints. The high-stress mechanism is a bulk mechanism
with a thermal dependence dominated by the thermal
dependence of creep in the In-rich matrix. The low-stress
mechanism is a grain boundary mechanism.

4 Fatigue life equations


4.1 Coffin-Manson equation
For strain-based analysis, Coffin-Manson fatigue model is
widely used for low cycle fatigue analysis. More than
35 years have elapsed since Coffin and Manson [113, 114]
independently proposed the famous empirical law
Nf  Deb
P , which relates the number Nf of cycles to fracture in fatigue to the amplitude of the applied cyclic plastic
deformation Dep . The exponent b has been found remarkably universal and very close to 2 in single-phased metallic
materials, whatever their atomic and/or polycrystalline
structure. This law holds when Dep is not too small defining
the low-cycle-fatigue regime (LCF) [115]. Under LCF
conditions with or without hold time, the relation between
the number of cycles to failure (Nf ) and the plastic strain
range (Dep ) is known to have the following relationship:
 b
Nf C Dep
10
where Nf is the number of cycles to failure, C and b are the
materials constants, Dep is the plastic strain range.
In the LCF tests, as the loading hold time is increased,
the fatigue life is observed to be decreased at a fixed test
temperature and a given plastic strain range [116]. Solder

123

J Mater Sci: Mater Electron (2013) 24:172190

joints fatigue life subject to thermal cycling loading is


often characterized by a series of isothermal low cycle
fatigue (LCF) tests at different test temperatures and frequencies [117]. Although much information is available for
the low cycle fatigue behavior of Sn37Pb solder, less data
can be found for lead-free alloys. Chen [118] fitted CoffinManson equation of Sn4.0Ag0.5Cu lead-free solder joints
by using results of solder joints reliability test and finite
element analysis. Moreover, Sun [119] has determined the
Coffin-Manson (CM) equation constants for fatigue life
estimation of Sn8Zn3Bi solder joints, since Sn8Zn3Bi
solder has a melting temperature of around 199 C which is
close to that of the conventional SnPb solder which has
previously been used in the electronics assembly industry.
The CM equation for Sn8Zn3Bi solder joints was fitted to
the lifetimes measured and the shear strains simulated. The
constants were determined to be 0.0294 for C, the proportional constant, and for the fatigue exponent, b, -2.833.
Low cycle fatigue studies on Sn0.7Cu lead-free solders was
evaluated over a range of test temperatures (298, 348 and
398 K) and frequencies (10-3 to 1 Hz), tested at four
values of total strain range (2, 2.5, 5 and 7.5 %) by Pang
[120]. It is found that the fatigue exponent and ductility
coefficient in the Coffin-Manson model can be determined
at three temperatures of 298, 348, 398 K and at the three
frequencies of 10-3, 10-2, and 1 Hz. In addition, the low
cycle fatigue behavior of the eutectic solder was found by
Shi [121] to be strongly dependent on test temperature and
frequency. If the CoffinManson model is used to describe
such fatigue behavior, the fatigue exponent and ductility
coefficient in the model are found to be a function of
temperature and frequency rather than numerical constants.
The plastic flow law was employed to explain the temperature and frequency dependence. The frequency-modified CoffinManson model was tried and found to be able
to eliminate the frequency dependence of the numerical
constants but not the temperature dependence. To have a
full description of the temperature- and frequency-dependent fatigue behavior, a set of empirical formulae was
derived based on the frequency-modified CoffinManson
model.
The low cycle fatigue lives under constant-amplitude
uniaxial loading can be corrected by the Basquin-CoffinManson equation as follows [122]:
b

c
De r0f 

2Nf e0f 2Nf


2
E

11

In this expression, r0f and e0f are constants known as the


fatigue strength coefficient and the fatigue ductility
coefficient, respectively, and b and c are known as the
fatigue strength exponent and the fatigue ductility
exponent, respectively. 2Nf is a reversal and is defined to

J Mater Sci: Mater Electron (2013) 24:172190

183

Fig. 12 Total strain versus life equation [126]

allow for analysis of complex loading spectra. This fatigue


model includes fatigue damage contributions from plastic
deformation (CoffinManson LCF curve) and elastic
deformation (Basquin HCF curve) [123]. As seen in
Eq. (11), there are four model constants altogether that
need to be estimated. The plastic (LCF) part of the
durability model has been investigated in previous studies
for SAC305 [124] and for Sn37Pb [125], based on 50 %
load-drop under quasistatic mechanical cycling on a shear
specimen. This fatigue model is an improvement over the
Coffin-Manson equation in that it also accounts for the
elastic contribution to fatigue failure. As can be seen in
Fig. 12 [126], the low cycle region to the left of Nf is
governed by the plastic-strain amplitude (Coffin-Manson
equation), and the high cycle region to the right of Nf is
governed by the elastic-strain amplitude (Basquins
equation).

where e_ f is fatigue ductility coefficient, Nf is the mean


cycles to failure. c is the fatigue ductility exponent,
c 0:442  6  104 TS 1:74  102 ln1 f , TS is
the mean cyclic solder joint temperature, C, f is the cyclic
frequency, 1  f  1; 000 cycles/day. e0f is the fatigue ductility coefficient, 2e0f 0:65, f is the cyclic frequency.
For eutectic SnPb solder the two fatigue parameters in
Eq. (11) can be determined from the experiments by
making assumptions about the correlations between the
fatigue ductility exponent and both the solder joints temperatures and the appropriate behavior of the exponent
within the ranges, a linear temperature correlation and a
logarithmic frequency correlation appear to describe this
behavior best [130]. Shin [131] predicted the thermal
fatigue life of a lBGA (Micro Ball Grid Array) solder
joints based on Engelmaier fatigue life equation. Thermal
fatigue life of four different solder alloys shows an
increasing order: as Sn3.5Ag3In0.5Bi [ Sn3.5Ag0.7Cu [
Sn3.5Ag [ Sn37Pb at the low temperature range, and
Sn3.5Ag [ Sn3.5Ag0.7Cu [ Sn37Pb [ Sn3.5Ag3In0.5Bi
at the high temperature range. The fatigue life of PBGA
was predicted using Engelmaier model by Tong [132]. The
simulation result of the model shows that the position of
the critical soldered position of a plastic ball grid array
component is right below the edge of its die, but not the
outboard solders. This result is helpful to improve the
thermal fatigue reliability of plastic ball grid array
components. And Ma [133] found that compared with
Engelmaier solder joints prediction fatigue life model,
Coffin-Manson model is highly evaluated than that of
practical experimental fatigue life in flip chip.
4.3 Darveauxs energy method

4.2 Engelmaier equation


Werner Engelmaier developed his semi-empirical solder
joint fatigue model during the early 1980s [127]. The
model is based on generic fatigue models for metals. By
testing a considerable number of test devices, Engelmaier
could turn the generic, somewhat abstract parameters into
easily measurable ones. The formulation used is based on
the shear strain assumption resulting in easily measurable
parameters, such as solder joint height and distance to
neutral point. The Engelmaier fatigue model is shown in
Eq. (12) [128, 129]. The total number of cycles to failure is
related to the total shear strain Dc, the fatigue ductility
coefficient e_ f , and the variable, c, which is a function of
frequency and temperature.
1 Dc
Nf
2 2e0f

!1=c
12

There have been several models describing the relationship


between the physical constants and the real lifetime of the
solder joints under different thermal or mechanical stress
conditions, based on stress, strain or energy Among them,
the Darveaux model [36] is the most popular one because
of its good compatibility with ANSYS software. In Darveaus energy method, to predict characteristic fatigue life
of each solder joints, accumulated plastic work per cycle is
averaged across the elements along the solder joint interface where the crack propagates. The equation for
stabilized change in average plastic work is given in
Eq. (13).
P
DW  V
P
DWave
13
V
Calculation of thermal cycles to crack initiation and crack
propagation rate per thermal cycle are given by Eqs. (14)
and (15)

123

184

NO K1 DWave K2
da
K3 DWave K4
dN
a
Nf No
da=dN

J Mater Sci: Mater Electron (2013) 24:172190

Nf C0 eacc

15

Nf W 0 wacc

16

where K1 , K2 , K3 , K4 are the experimental constants related to


failure mechanism; No is the number of cycles when the
failure appears; Nf is the number of cycles when the failure
occurs; a is the critical length for fracture; da=dN is the speed
of crack expansion; DWave is the average strain energy density.
Characteristic solder joint fatigue life can then be calculated by summing the cycles to crack initiation with the
number of cycles taken by crack to propagate across the
entire solder joint diameter shown in Eq. (17)
a
a NO
17
da=dN
To evaluate characteristic life, DWave, i.e., element
volumetric average of the stabilized change in plastic work
within the controlled soldered element thickness was used.
But, maximum accumulated plastic work per cycle was
used in place of DWave. This approach was used to achieve
more realistic crack initiation across the solder
interconnects since averaging of plastic work has a
tendency to overpredict the life of solder interconnects
[134]. A 3D finite element model has been developed by
Chen [135] based on Darveauxs energy method to study
the reliability of BGA solder joints, it is found that solder
joints fatigue life is dependent on the its location, namely
the outmost solder joints fatigue life is shortest.
4.4 Syed model
Many thermo-mechanical fatigue life prediction models
have been developed, recently, for SnPb and lead-free
solder interconnects either based on strain range, accumulated creep strain or based on the accumulated strain
energy density during a temperature cycle. Among those
life models, the models, proposed by Syed [136, 137] based
on accumulated creep strain and accumulated creep strain
density, however, are worth to be explained. According to
Syed, creep is considered as the primary damage mechanism for SnAgCu solder during thermal cycling and is
solely used to simulate the material behavior. The life
prediction basis, therefore, should to be creep deformation
also. The cyclic loading is considered as a special case of
creep due to sequence of loading which are repeated in a
cyclic manner to derive the life prediction model parameters. The life prediction model parameters were determined
by curve fitting creep strain and accumulated energy density per cycle calculated from simulations. The life prediction models [102] are given as

123

1

14

18

1

19

where Nf is the number of repetitions or cycles to failure,

eacc the accumulated creep strain per cycle, C0 1 ef the


inverse of creep ductility and ef is the constant which gives
the creep ductility or the strain at the onset of failure, wacc
the accumulated creep strain energy density per cycle and
W 0 is the creep energy density for failure.
In addition, Wise [83, 104, 138, 139] described the
steady state creep behavior of lead-free solder both by
dislocation climb controlled process at low stress and the
combined dislocation glide/climb process at high stress.
Therefore, in order to predict the fatigue life of lead-free
solder joints, it is necessary to take the two factors into
consideration. For these reasons, Syed A [140] gave the
fatigue life prediction based on accumulated creep strain
and creep strain energy density as follows:
1
Nf C1 eIacc C2 eP
acc

1
Nf WI wIacc WP wP
acc

20
21

where Nf is the number of repetitions or cycles to failure,


C1 and C2 are materials creep constants, wIacc and wP
acc are
I
P
the creep strain energy density per cycle, eacc and eacc are
the accumulated creep strain parameter per cycle under the
conditions of low stress and high stress, respectively.
4.5 Others
The fatigue life of solder joints can be described with a
two-parameters Weibull equation [141].
F x 1  ex=h

22

where F x is the probability of failure after x cycles, b is


the shape parameter, h is the mean life (63 % failure).
The critical plane approach proposed by Brown and Miller
[142] is given in the generalized functional form of shear
strain and normal strain on the maximum shear strain plane.
b

c
s0f 
Dcmax
ken
2Nf o c0f 2Nf o
2
G

23

where Dcmax is the maximum shear strain range, en is the


normal strain range acting on the maximum shear strain
plane, and k is a material constant that can be determined
by merging axial and torsional fatigue data together.
Also it is found that fatigue life has a power relationship
with ratcheting strain rate. For this reason, one proposes the
range of maximum shear strain range (Dcmax ) and axial
ratcheting strain rate (_er ) as damage parameters for the new
model. The new parameter is given by the following
equation [143].

J Mater Sci: Mater Electron (2013) 24:172190

b

c
s0f 
Dcmax
o
S_ec
2Nf o c0f 2Nf o

r
2
G

185

24

where Dcmax is the maximum shear strain range, e_ r is saturation of ratcheting strain rate, co is the shear fatigue
ductility exponent, and S is a material constant that can be
determined by uniaxial ratchetingfatigue tests.
A fatigue life prediction methodology was developed by
Pierce [144], based on stress-strain, creep, and isothermal
fatigue data. The proposed fatigue life prediction methodology builds on current practices in fatigue prediction for
solder alloys, particularly the concepts of unpartitioned
energy methods in finite element analysis (FEA) and continuum damage mechanics.
pl

da

dNi

ZWf 
W1pl

i


NCS;f
X  da
da
pl

T;
a

dW

T;
a

DWjpl
pl
dW pl
dW
i
j
jN
CS;1

25
where da=dNi is the crack growth rate per cycle for the ith
cycle. Increments of plastic work are integrated from the
initial viscoplastic work of the ith cycle, W1pl , to the final
viscoplastic work of the same cycle, Wfpl . On the right hand
side is the numerical implementation of the crack growth
rate per cycle. The summation for this numerical estimate
is done from the first converged substep of the ith cycle,
i
i
NCS;1
, to the final converged substep of the ith cycle, NCS;f
.
Nozaki [145] studies the creepfatigue life evaluation of
Sn3.5Ag solder under push-pull loading using fast-fast,
fast-slow, slow-fast, slow-slow, and strain-hold strain
waves. The grain boundary sliding model (GBSM)
assumed that the grain boundary sliding rate is proportional
to the creepfatigue damage rate, so cycles to failure in any
strain wave (Nf ) is related to cycles to failure in respective
strain waves by
1
1
1
1
1

26
Npc Ncp Ncc Nth Nf

where 1 Npc , 1 Ncp , 1=Ncc and 1=Nth are the creepfatigue


damage rates in pc, cp, cc, and th tests, respectively; pc, the
unsymmetrical triangular waves with fast tension-going/
slow compression-going strain rates; cp, slow tensiongoing/fast compression-going strain rates; cc, wave
includes creep deformation in both tension and compression going stages; th, trapezoidal strain wave with tension
hold-times.
An innovative methodology to predicting solder joint
fatigue life that combines the strength of mechanics-based
finite-element formation, sets of experimental data, and
design of simulations has been developed by Perkins
[146].

Universal CBGA predictive equation for thermal


cycling:
N ATC Weibull
uWeibull  AF
uAF  N50
uN 50



 
1=3
ln1  F%  0:01 1=c
135 1:9 f ATC


ln 0:5
DT ATC
2

1
0
1414

B
 @e

1
1
T ATC 383
peak

C
A  12476  0:241A  719:9B

1728C  1381D  12456E 78:7BC



72:4BD 220CD
kuN ATC
27
A, substrate size (mm), 2532.5 mm; B, CTE mismatch,
7.411.2 ppm/C; C, substrate thickness, 0.82.9 mm; D,
PWB thickness, 1.570.8 mm; E, pitch, 1.001.27 mm;
s
2
u
 u 2 u
Weibull 2
AF
N 50
ATC
uN ATC N
28

Weibull
AF
N 50
DT, temperature range (K); f, frequency (cph); Tpeak ,
peak temperature (K); F%, failure percentage, 0 \ F%
\ 100 %; c, shape parameter, 411 for CBGA; uWeibull ,
estimated uncertainty of the Weibull factor; uAF , estimated
uncertainty of the AF factor; uN 50 , estimated uncertainty
of N50 regression equation.
A new and simple empirical equation for predicting the
thermal-fatigue life of solder bumped flip chip on low-cost
PCB has been presented by Lau [147].
da
C1 10C2 C3 C1 C4 C4 =C1 C3 DW C1 1=C1 C3
dN

29

where a is the crack length, N is the number-of-cycles, C1 ,


C2 , C3 , C4 are constants, and da=dN is the fatigue crack
growth rate of the corner solder joints. DW is the average
strain energy range.

5 Conclusion
Miniaturization and high levels circuit integration are
driving forces for surface mount technology (SMT) devices
in finer pitches, higher speed, and greater packaging density. However, the reliability of solder joints has been
became a serious concern because solder joints in electronics components provide not only the mechanical
attachments but also electrical connections were investigated systematically. In addition, the transition for leadfree solder and lead-free process will bring many new
reliability problems, especially for the constitutive relation
models and fatigue life prediction equation. Upon
increasing to the concentration further, until now there are

123

186

few reports about the strain to failure of new lead-free


solder joints, such as SnAgCu or SnZn lead-free solder
joints with the addition of alloying elements. The mechanisms of failure for the observed behavior are not clear and
have not been fully discussed. In addition, in service, solder
joints experience high homologous temperatures that promote time-dependent creep and microstructural change,
and an important area of investigation is the formulation of
constitutive equations that describe the resultant deformation in terms of external variables: applied stress, temperature, and time. These relationships are essential for
assessing lead-free alloys performance under likely inservice conditions, and they provide crucial input to finite
element models that examine the structural integrity of
lead-free solder joints. Moreover, this paper presents a
critical review of fatigue life prediction methodology of
solder joints. Approaches of fatigue life prediction are
developed based on experimental stress/strain/energy density and so on. A number of fatigue life prediction methods
have been confirmed during the past few years. However,
there are few literatures reported the fatigue life model for
lead-free solder joints. Therefore, the reliability of leadfree solder joints need to be investigated systematically,
especially the constitutive relation models and fatigue life
prediction equations.
Acknowledgments The present work was carried out with the supported by the Jiangsu University of Science and Technology: Provincial
Key Lab of Advanced Welding Technology Foundation (JSAWS-1103) and the Xuzhou Normal University Foundation (11XLR16).

References
1. J.O. Kim, J.P. Jung, J.H. Lee, J. Suh, H.S. Kang, Effects of laser
parameters on the characteristics of a Sn-3.5wt.%Ag solder
joint. Met. Mater. Int. 15(1), 119123 (2009)
2. M.N. Wang, J.Q. Wang, H. Feng, W. Ke, Effects of microstructure and temperature on corrosion behavior of Sn-3.0Ag0.5Cu lead-free solder. J. Mater. Sci.: Mater. Electron. 23(1),
148155 (2012)
3. C.D. Zou, Y.L. Gao, B. Yang, Q.J. Zhai, Nanoparticles of
Sn3.0Ag0.5Cu alloy synthesized at room temperature with large
melting temperature depression. J. Mater. Sci.: Mater. Electron.
23(1), 27 (2012)
4. W.H. Zhong, Y.C. Chan, M.O. Alam, B.Y. Wu, J.F. Guan,
Effect of multiple reflow processes on the reliability of ball grid
array (BGA) solder joints. J. Alloy. Compd. 414(12), 123130
(2006)
5. C. Yu, Y. Yang, P.L. Li, J.M. Chen, H. Lu, Suppression of
Cu3Sn and Kirkendall voids at Cu/Sn-3.5Ag solder joints adding
a small amount of Ge. J. Mater. Sci.: Mater. Electron. 23(1),
5660 (2012)
6. L. Zhang, S.B. Xue, L.L. Gao, G. Zeng, Z. Sheng, Y. Chen, S.L.
Yu, Effects of rare earths on properties on properties and
microstructures of lead-free solder alloys. J. Mater. Sci.: Mater.
Electron. 20(8), 685694 (2009)
7. D.H. Kim, Reliability Study of SnPb and SnAg Solder Joints in
PBGA Packages. The University of Texas at Austin (2007)

123

J Mater Sci: Mater Electron (2013) 24:172190


8. A. Schubert, R. Dudek, H. Walter, E. Jung, A. Gollhardt, B.
Michel, H. Reichl, Reliability assessment of flip-chip assemblies
with lead-free solder joints, in Proceedings of 52nd Electronic
Components and Technology Conference, pp 12461255 (2002)
9. A. Shawkret, Study on the high temperature reliability of leadfree surface mount solder joints (Shanghai Institute of Metallurgy, Chinese Academy of Sciences, 2001)
10. S.B. Xue, Y.F. Hu, S.L. Yu, Reliability of CBGA solder joint
under thermal cycling. Trans. China Weld. Inst. 26(10), 8183
(2005)
11. C. Kanchanomai, Y. Miyashita, Y. Mutoh, Low cycle fatigue
behavior and mechanisms of a eutectic Sn-Pb solder 63Sn/37Pb.
Int. J. Fatigue 24(6), 671683 (2002)
12. T. Hannach, H. Worrack, W.H. Muller, T. Hauck, Creep in
microelectronic solder joints: finite element simulations versus
semi-analytical methods. Arch. Appl. Mech. 79(67), 605617
(2009)
13. S.B. Xue, H. Wang, Y.F. Hu, Effects of thermal cycling on shear
strength of solder joints of laminar ceramic resistor. Trans.
Nonferrous Soc. China 15(S3), 305310 (2005)
14. S.B. Xue, Z.J. Han, H. Wang, J.X. Wang, Fracture mechanism
of lead-free solder joints of rectangular chip component. Trans.
China Weld. Inst. 27(8), 2326 (2006)
15. P. Hegde, D.C. Whalley, V.V. Silberschmidt, Creep damage
study at powercycling of lead-free surface mount device.
Comput. Mater. Sci. 45(3), 638645 (2009)
16. C. Han, B. Song, Development of life prediction model for leadfree solder at chip resistor, in 8th Electronics Packaging Technology Conference, pp. 781786 (2006)
17. Y. Qi, R. Lam, H.R. Ghorbani, P. Snugovsky, J.K. Spelt,
Temperature profile effects in accelerated thermal cycling of
SnPb and Pb-free solder joints. Microelectron. Reliab. 46(24),
574588 (2006)
18. Z.J. Han, S.B. Xue, J.X. Wang, X. Zhang, L. Zhang, S.L. Yu, H.
Wang, Mechanical properties of QFP micro-joints soldered with
lead-free solders using diode laser soldering technology. Trans.
Nonferrous Metals Soc. China 18(4), 814818 (2008)
19. L. Zhang, S.B. Xue, Z.J. Han, J.X. Wang, L.L. Gao, Z. Sheng,
Mechanical properties of fine pitch devices solder joints based
on creep model. Chin. J. Mech. Eng. 21(6), 8285 (2008)
20. L. Zhang, S.B. Xue, G. Zeng, Z.J. Han, S.L. Yu, Study on
mechanical properties and numerical simulation of fine pitch
devices solder joints. Trans. China Weld. Inst. 29(10), 8992
(2008)
21. L. Zhang, S.B. Xue, L.L. Gao, Z. Sheng, G. Zeng, Y. Chen, S.L.
Yu, Properties of SnAgCu/SnAgCuCe solder joints for electronic packaging. J. Mater. Sci.: Mater. Electron. 21(6), 635642
(2010)
22. F.X. Chen, J.H.L. Pang, B.S. Xiong, L.H. Xu, T.H. Low, Lead
free solder joint reliability characterization for PBGA, PQFP
and TSSOP assembilities, in Electronic Components and
Technology Conference (2005)
23. Y.X. Wu, S.B. Xue, Z.J. Han, J.X. Wang, Effects of lead widths
and pitches on reliability of quad flat package (QFP) solder
joints. Chin. J. Mech. Eng. 20(4), 4043 (2007)
24. L. Zhang, S.B. Xue, F.Y. Lu, Z.J. Han, Numerical simulation on
solder joints of QFP device with different substrate materials
and thicknesses. Trans. China Weld. Inst. 29(1), 3539 (2008)
25. T.Y. Tee, Z.W. Zhong, Integrated vapor pressure, hygroswelling, and thermo-mechanical stress modeling of QFN package
during reflow with interfacial fracture mechanics analysis. Microelectron. Reliab. 44(1), 105114 (2004)
26. T.Y. Tee, H.S. Ng, D. Yap, Z.W. Zhong, Comprehensive boardlevel solder joint reliability modeling and testing of QFN and
powerQFN packages. Microelectron. Reliab. 43(8), 13291338
(2003)

J Mater Sci: Mater Electron (2013) 24:172190


27. B. Vandevelde, M. Gonzalez, P. Limaye, P. Ratchev, E. Beyne,
Thermal cycling reliability of SnAgCu and SnPb solder joints: a
comparison for several IC-packages. Microelectron. Reliab.
47(23), 259265 (2007)
28. L. Zhang, S.B. Xue, F.Y. Lu, Z.J. Han, J.X. Wang, Numerical
simulation of solder joints and reliability analysis of PLCC
components with J-shape leads. China Weld. 17(2), 3741
(2008)
29. H.T. Chen, C.Q. Wang, M.Y. Li, Y.H. Tian, Experimental and
finite element method studies of J-lead solder joint reliability.
J. Mater. Sci. Technol. 21(3), 419422 (2005)
30. Y.X. Wu, S.B. Xue, Y.F. Hu, Finite element analysis on reliability of solder joint of J-lead. Trans. China Weld. Inst. 26(12),
8588 (2005)
31. Y.X. Wu, Research on the reliability of solder joints and optimum simulation for QFP gull wing lead sizes (Nanjing University of Aeronautics and Astronautics, Nanjing, 2007)
32. W. Dauksher, J. Lau, A finite-element-based solder-joint fatigue-life prediction methodology for Sn-Ag-Cu ball-grid-array
packages. IEEE Trans. Device Mater. Reliab. 9(2), 231235
(2009)
33. L.J. Ladani, A. Dasgupta, A meso-scale evolution model for
cyclic fatigue of viscoplastic materials. Int. J. Fatigue 31(4),
703711 (2009)
34. F. Liu, G. Meng, M. Zhao, J.F. Zhao, Experimental and
numerical analysis of BGA lead-free solder joint reliability
under board-level drop impact. Microelectron. Reliab. 49(1),
7985 (2009)
35. T. Kangasvieri, O. Nousiainen, J. Putaala, R. Rautioaho, J.
Vahakangas, Reliability and RF performance of BGA solder
joints with plastic-core solder balls in LTCC/PWB assemblies.
Microelectron. Reliab. 46(8), 13351347 (2006)
36. X. Qu, Z.Y. Chen, B. Qi, T. Lee, J.J. Wang, Board level drop
test and simulation of leaded and lead-free BGA-PCB assembly.
Microelectron. Reliab. 47(12), 21972204 (2007)
37. T.H. Wang, Y.S. Lai, J.D. Wu, Effect of underfill thermomechanical properties on thermal cycling fatigue reliability of flipchip ball grid array. J. Electron. Packag. 126(4), 560564 (2004)
38. L. Zhang, S.B. Xue, Z.J. Han, F.Y. Lu, S.L. Yu, Z.M. Lai,
Fatigue life prediction of SnAgCu solder joints of FCBGA
device. Trans. China Weld. Inst. 29(7), 8588 (2008)
39. S. Cho, J. Choi, H. Kim, Study on the behavior characteristics of
solder balls for FCBGA package. Met. Mater. Int. 15(2),
299305 (2009)
40. J.L. Wang, H.K. Lim, H.S. Lew, W.T. Saw, C.H. Tan, A testing
method for assessing solder joint reliability of FCBGA packages. Microelectron. Reliab. 44(5), 833840 (2004)
41. T.T. Mattila, P. Marjamaki, J.K. Kivilahti, Reliability of CSP
interconnections under mechanical shock loading conditions.
IEEE Trans. Compon. Packag. Technol. 29(4), 787795 (2006)
42. M. Gonzalez, B. Vandevelde, E. Beyne, Thermo-mechanical
analysis of a chip scale package (CSP) using lead free and lead
containing solder materials, in European Microelectronics and
Packaging Symposium, pp. 247252 (2004)
43. P. Sun, P. Hochstenbach, W.D.V. Driel, G.Q. Zhang, Fracture
morphology and mechanism of IMC in low-Ag SAC solder/
UBM (Ni(P)-Au) for WLCSP. Microelectron. Reliab. 48(89),
11671170 (2008)
44. B. Vandevelde, E. Beyne, K. Zhang, J. Caers, D. Vandepitte, M.
Baelmans, Parameterized modeling of thermomechanical reliability for CSP assemblies. J. Electron. Packag. 125(4), 498505
(2003)
45. S.W. Yoon, C.J. Park, S.H. Hong, J.T. Moon, I.S. Park, H.S.
Chun, Interfacial reaction and solder joint reliability of Pb-free
solders in lead fram chip scale package (LF-CSP). J. Electron.
Mater. 29(10), 12331240 (2000)

187
46. A. Mertol, Application of the Taguchi method to chip scale
package (CSP) design. IEEE Trans. Adv. Packag. 23(2),
266276 (2000)
47. J.H.L. Pang, D.Y.R. Chong, Flip chip on board solder joint
reliability analysis using 2-D and 3-D FEA models. IEEE Trans.
Adv. Packag. 24(4), 499506 (2001)
48. X. Han, H. Ding, X.J. Sheng, B. Zhang, Thermal fatigue life
time prediction of Sn-3.5Ag lead-free solder joint for chip scale
package. Chin. J. Semiconduct. 27(9), 16951700 (2006)
49. Q. Wang, H. Liang, Y.J. Xu, Y. Liu, Solder joint fatigue life
predictions for CSP package under power cycling. J. Zhejiang
Univ. Technol. 34(2), 157161 (2006)
50. J. Lau, W. Dauksher, Reliability of an 1657CCGA (ceramic
column grid array) package with 95.5Sn3.9Ag0.6Cu lead-free
solder paste on PCBS (printed circuit boards). J. Electron.
Packag. 127(2), 96105 (2005)
51. B.Z. Hong, S.K. Ray, Ceramic column grid array technology
with coated solder columns, in Electronic Components and
Technology Conference (2000)
52. L.L. Gao, S.B. Xue, L. Zhang, Z. Sheng, Finite element analysis
on influencing factors of soldered column reliability in a CCGA
device. Trans. China Weld. Inst. 29(7), 9396 (2008)
53. S.B. Park, R. Joshi, Comparison of thermo-mechanical behavior
of lead-free copper and tin-lead column grid array packages.
Microelectron. Reliab. 48(5), 763772 (2008)
54. R. Ghaffarian, Thermal cycle reliability and failure mechanisms
of CCGA and PBGA assemblies with and without corner staking. IEEE Trans. Compon. Packag. Technol. 31(2), 285296
(2008)
55. A. Perkins, S.K. Sitaraman, Analysis and prediction of vibration-induced solder joint failure for a ceramic column grid array
package. J. Electron. Packag. 130(1), 111 (2008)
56. E.M. Ingalls, M. Cole, J. Jozwiak, C. Milkovich, J. Stack, in
Improvement in reliability with CCGA column density increase
to 1 mm pitch. Electronic Components and Technology Conference (1998)
57. A. Perkins, S.K. Sitaraman, in A study into the sequencing of
thermal cycling and vibration tests. Electronic Components and
Technology Conference, pp. 584592 (2008)
58. R.T. Winslow, Challenges in modification of electronic components. IEEE Trans. Compon. Packag. Technol. 30(2),
361363 (2007)
59. R. Wu, F.P. Mccluskey, in Constitutive relations of indium
solder joint in cold temperature electronic packaging based on
Anand model. 11th Intersociety Conference on Thermal and
Thermomechanical Phenomena in Electronic Systems, Florida,
USA, May, pp. 683686 (2008)
60. L. Anand, Constitutive equations for hot-working of metals. Int.
J. Plast 1(3), 213231 (1985)
61. S.B. Brown, K.H. Kim, L. Anand, An internal variable constitutive model for hot working of metals. Int. J. Plast 5(3), 95130
(1989)
62. G.Z. Wang, Z.N. Chen, Viscoplastic Anand constitutive relations of tin-lead solder alloy. Chin. J. Appl. Mech. 17(3),
133139 (2000)
63. Y.X. Wu, S.B. Xue, L. Zhang, X. Huang, Optimum simulation
and prediction on thermal fatigue life of solder joints of QFP
devices. Trans. China Weld. Inst. 27(8), 99102 (2006)
64. L. Zhang, S.B. Xue, S.L. Yu, Z.J. Han, L.L. Gao, F.Y. Lu, Z.
Sheng, Application of FEM analysis in reliability of microsolder joints. Electr. Weld. Mach. 38(9), 1321,72 (2008)
65. L. Zhang, S.B. Xue, L.L. Gao, Y. Chen, S.L. Yu, Z. Sheng,
Determination of Anand parameters for SnAgCuCe solders.
Modell. Simul. Mater. Sci. Eng. 17(7), 075014 (2009)
66. X. Chen, G. Chen, M. Sakane, Prediction of stress-strain relationship with an improved Anand constitutive model for lead-

123

188

67.

68.

69.

70.

71.

72.

73.

74.

75.

76.

77.

78.

79.

80.

81.

82.

83.

84.

J Mater Sci: Mater Electron (2013) 24:172190


free solder Sn-3.5Ag. IEEE Trans. Compon. Packag. Technol.
28(1), 111116 (2005)
D. Bhate, D. Chan, G. Subbarayan, C. Chiu, V. Gupta, D.R.
Edwards, Constitutive behavior of Sn3.8Ag0.7Cu and
Sn1.0Ag0.5Cu alloys at creep and low strain rate regimes. IEEE
Trans. Compon. Packag. Technol. 31(3), 622633 (2008)
N. Bai, X. Chen, H. Gao, Simulation of uniaxial tensile properties for lead-free solders with modified Anand model. Mater.
Des. 30(1), 122128 (2009)
L. Zhang, X. Chen, H. Nose, M. Sakane, Stress-strain behaviors
of 63Sn37Pb solder simulated by Anand model. J. Mech.
Strength 26(4), 447450 (2004)
B. Rodgers, B. Flood, F. Waldron, in Experimental determination and finite element model validation of the Anand viscoplasticity model. Proceedings of the 6th International
Conference on Thermal, Mechanical and Multi-Physics Simulation and Experiments in Micro-electronics and Micro-systems,
490496 (2005)
P. Zhou, B.T. Hu, J.M. Zhou, Y. Yang, Parameter fitting of
constitutive model and FEM analysis of solder joint thermal
cycle reliability for lead-free solder Sn-3.5Ag. J. Central South
Univ. Technol. 16(3), 339343 (2009)
A. Yeo, C. Lee, J.H.L. Pang, Flip chip solder joint reliability
analysis using viscoplastic and elastic-plastic-creep constitutive
models. IEEE Trans. Compon. Packag. Technol. 29(2), 355363
(2006)
J.G. Bai, J.N. Calata, G.Q. Lu, Discussion on the reliability
issues of solder-bump and direct-solder bonder power device
packages having double-sided cooling capability. J. Electron.
Packag. 128(3), 208214 (2006)
Q. Wang, L.H. Liang, X.F. Chen, X.H. Weng, in Experimental
determination and modification of Anand model constants for
Pb-free materials 95.5Sn4.0Ag0.5Cu. International Conference
on Thermal, Mechanical and Multi-Physics Simulation Experiments in Microelectronics and Micro-System, 19 (2007)
W. Wang, Z.G. Wang, A.P. Xian, J.K. Shang, Microstructure
and fracture of Pb-free solder interconnects in CBGA packages
under thermal cycling. Acta Metall. Sin. 42(6), 647652 (2006)
L. Zhang, S.B. Xue, F.Y. Lu, Z.J. Han, Finite element analysis
on solder joint reliability of QFP devices with different lead
materials. Trans. China Weld. Inst. 28(6), 6568 (2007)
G.Z. Wang, Z.N. Cheng, K. Becker, J. Wilde, Applying Anand
model to represent the viscoplastic deformation behavior of
solder alloys. J. Electron. Packag. 123(3), 247253 (2001)
X.F. Chen, L.H. Liang, Y. Liu, Q. Wang, Mechanical properties
and parameter determination of Anand viscoplastic constitutive
model for Pb-free alloys. Chin. J. Appl. Mech. 26(2), 248252
(2009)
Z.N. Cheng, G.Z. Wang, L. Chen, J. Wilde, K. Becker, Viscoplastic Anand model for solder alloys and its application. Solder.
Surface Mount Technol. 12(2), 3136 (2000)
J.H. Choi, A.H.M.F. Anwar, Y. Ichikawa, Observation of timedependent local deformation of crystalline rocks using a confocal laser scanning microscope. Int. J. Rock Mech. Mining Sci.
45(3), 431441 (2008)
S. Ridout, C. Bailey, Review of methods to predict solder joint
reliability under thermo-mechanical cycling. Fatigue Fract. Eng.
Mater. Struct. 30(5), 400412 (2006)
R. Darveax, K. Banerji, A. Mawer, G. Doddy, Reliability of
plastic ball grid array assembly in ball grid array technology,
New York (1995)
S. Wiese, S. Rzepka, Time-independent elastic-plastic behaviour
of solder materials. Microelectron. Reliab. 44(12), 18931900
(2004)
H.T. Ma, Constitutive models of creep for lead-free solders.
J. Mater. Sci. 44(14), 38413851 (2009)

123

85. R.J. McCabe, M.E. Fine, Creep of tin, Sb-solution-strengthened


tin, and SbSn-precipitate-strengthened tin. Metallurg. Mater.
Trans. A 33(5), 15311539 (2002)
86. S. Wiese, M. Roellig, M. Mueller, S. Bennemann, M. Petzold,
K.J. Wolter, in The size effect on the creep properties of
SnAgCu solder alloy. Electronic Components and Technology
Conference (2007)
87. G.S. Zhang, H.Y. Jing, L.Y. Xu, Y.D. Han, Creep behavior of
eutectic 80Au/20Sn solder alloy. J. Alloy. Compd. 476(12),
138141 (2009)
88. J.W. Kim, J.K. Jang, S.O. Ha, S.S. Ha, D.G. Kim, S.B. Jung,
Effect of high-speed loading conditions on the fracture mode of
the BGA solder joint. Microelectron. Reliab. 48(1112),
18821889 (2008)
89. B.Z. Hong, in Thermal fatigue analysis of a CBGA package with
lead-free solder fillets. InterSociety Conference on Thermal
Phenomena, pp. 205211 (1998)
90. J.W. Kim, S.B. Jung, Design of solder joint structure for flip
chip package with an optimized shear test method. J. Electron.
Mater. 36(6), 690696 (2007)
91. F.X. Chen, J.H.L. Pang, in Thermal fatigue reliability analysis
for PBGA with Sn-3.8Ag-0.7Cu solder joints. Electronics
Packaging Technology Conference, pp. 787792 (2004)
92. Y. Qi, H.R. Ghorbani, J.K. Spelt, Thermal fatigue of SnPb and
SAC resistor joints: analysis of stress-strain as a function of
cycle parameters. IEEE Trans. Adv. Packag. 29(4), 690700
(2006)
93. J. Lau, W. Dauksher, in Effects of ramp-time on the thermalfatigue life of SnAgCu lead-free solder joint. Electronic Components and Technology Conference, pp. 12921298 (2005)
94. J.H. Lau, S.H. Pan, C. Chang, Creep analysis of wafer level chip
scale package (WLCSP) with 96.5Sn-3.5Ag and 100In lead-free
solder joints and microvia build-up printed circuit board.
J. Electron. Packag. 124(2), 6976 (2002)
95. W.R. Jong, H.C. Tsai, H.T. Chang, S.H. Peng, The effects of
temperature cyclic loading on lead-free solder joints of wafer
level chip scale package by Taguchi method. J. Electron.
Packag. 130(1), 110 (2008)
96. L. Zhang, S.B. Xue, L.L. Gao, G. Zeng, Y. Chen, S.L. Yu, Z.
Sheng, Creep behavior of SnAgCu solders with rare earth Ce
doping. Trans. Nonferrous Metals Soc. China 19(6), 753778
(2009)
97. S. Ridout, M. Dusek, C. Bailey, C. Hunt, Assessing the performance of crack detection tests for solder joint. Microelectron.
Reliab. 46(12), 21222130 (2006)
98. S. Wiese, E. Meusel, Characterization of lead-free solders in flip
chip joints. J. Electron. Packag. 125(4), 531538 (2003)
99. D. Pan, I. Dutta, A mechanics-induced complication of
impression creep and its solution: application to Sn-3.5Ag solder. Mater. Sci. Eng., A 379(12), 154163 (2004)
100. H. Mavoori, J. Chin, S. Vaynman, B. Moran, L. Keer, M. Fine,
Creep, stress relaxation, and plastic deformation in Sn-Ag and SnZn eutectic solders. J. Electron. Mater. 26(7), 783790 (1997)
101. M. Zeng, Z.Z. Chen, B.L. Shen, D.F. Xu, Indentation creep
behavior of Sn-3.5Ag-2Bi lead-free solder. Chin. J. Nonferrous
Metals 18(4), 620625 (2008)
102. X.Y. Li, Z.S. Wang, Thermo-fatigue life prediction methodologies for SnAgCu solder joints in flip chip assemblies. J. Mech.
Strength 28(6), 893898 (2006)
103. A. Schubert, R. Dudek, E. Auerswald, G.B. Michel, H. Reichl,
in Fatigue life models for SnAgCu and SnPb solder joints
evaluated by experiments and simulation. Proceedings of 53rd
Electronic Components and Technology Conference, New
Orleans, USA, May, pp. 603610 (2003)
104. S. Wiese, E. Meusel, K. Wolter, in Microstructural dependence
of constitutive properties of eutectic SnAg and SnAgCu solders.

J Mater Sci: Mater Electron (2013) 24:172190

105.
106.

107.

108.

109.

110.

111.

112.

113.

114.
115.

116.

117.
118.

119.

120.
121.

122.

123.

124.

125.

Proceeding of 53rd Electronic Components and Technology


Conference (2003)
D.H. Avery, W.A. Backofen, Structural basis for superplasticity.
Trans. Am. Soc. Metals 58(4), 551562 (1965)
S. Wiese, F. Feustel, E. Meusel, Characterisation of constitutive
behavior of SnAg, SnAgCu and SnPb solder in flip chip joints.
Sensors Actuator A Phys. 99(12), 188193 (2002)
H.G. Song, J.W. Morris, F. Hua, The creep properties of leadfree solder joints. J. Minerals Metals Mater. Soc. 54(6), 3032
(2002)
Y.D. Han, H.Y. Jing, S.M.L. Nai, C.M. Tan, J. Wei, L.Y. Xu,
S.R. Zhang, A modified constitutive model for creep of Sn3.5Ag-0.7Cu solder joints. J. Phys. D Appl. Phys. (2009). doi:
10.1088/0022-3727/42/12/125411
J. Rosler, Back-stress calculation for dislocation climb past noninteracting particles. Mater. Sci. Eng., A 339(12), 334339
(2003)
M. Kerr, N. Chawla, Creep deformation behavior of Sn-3.5Ag
solder/Cu couple at samll length scales. Acta Mater. 52(15),
45274535 (2004)
Z.G. Chen, Y.W. Shi, Z.D. Xia, Constitutive relations on creep
for SnAgCuRE lead-free solder joints. J. Electron. Mater. 33(9),
964971 (2004)
H.L. Reynolds, S.H. Kang, J.W. Morris, The creep behavior of
In-Ag eutectic solder joints. J. Electron. Mater. 28(1), 6975
(1999)
S.S. Manson, M.H. Hirshberg, in Fatigue, an interdisciplinary
approach. Proceedings of 10th Sagamore Army Materials
Research Conference, pp. 133178 (1964)
L.F. Coffin, A note on low cycle fatigue laws. J. Mater. 6(2),
388402 (1971)
D. Sornette, T. Magnin, Y. Brechet, The physical origin of the
Coffin-Manson law in low-cycle fatigue. Europhys. Lett. 20(5),
433438 (1992)
S.W. Nam, Y.C. Yoon, B.G. Choi, J.M. Lee, J.W. Hong, The
normalized Coffin-Manson plot in terms of a new damage
function based on grain boundary cavitation under creep-fatigue
condition. Metallurg. Mater. Trans. A 27(5), 12731281 (1996)
J.H.L. Pang, B.S. Xiong, T.H. Low, Low Cycle fatigue models
for lead-free solders. Thin Solid Films 462463, 408412 (2004)
S. Chen, P. Sun, X.C. Wei, Z.N. Cheng, F. Liu, Coffin-Manson
equation of Sn-4.0Ag-0.5Cu solder joint. Solder. Surface Mount
Technol. 21(2), 4854 (2009)
P. Sun, C. Andersson, X.C. Wei, Z.N. Cheng, D.K. Shanguan, J.
Liu, Coffin-Manson constant determination for a Sn-8Zn-3Bi
lead-free solder joint. Solder. Surface Mount Technol. 18(2),
411 (2006)
J.H.L. Pang, Low cycle fatigue study of lead free 99.3Sn-0.7Cu
solder alloy. Int. J. Fatigue 26(8), 865872 (2004)
X.Q. Shi, H.L.J. Pang, W. Zhou, Z.P. Wang, Low cycle fatigue
analysis of temperature and frequency effects in eutectic solder
alloy. Int. J. Fatigue 22(3), 217228 (2000)
X. Chen, J. Song, K.S. Kin, Low cycle fatigue life prediction of
63Sn-37Pb solder under proportional and non-proportional
loading. Int. J. Fatigue 28(7), 757766 (2006)
Y. Zhou, M.A. Bassyiouni, A. Dasgupta, Vibration durability
assessment of Sn3.0Ag0.5Cu and Sn37Pb solders under harmonic excitation. J. Electron. Packag. 131(1), 011016 (2009)
G. Cuddalorepatta, A. Dasgupta, in Cyclic mechanical durability
of Sn3.0Ag0.5Cu Pb-free solder alloy. Proceeding of ASME
International Mechanical Engineering Congress and RD&D
Expo, pp. 511 (2005)
P. Haswell, A. Dasgupta, in Durability properties characterization of Sn36Pb2Ag solder alloy. Proceedings of the ASME
International Mechanical Engineering Congress and Exposition,
pp. 181187 (2000)

189
126. W.W. Lee, L.T. Nguyen, G.S. Selvaduray, Solder joint fatigue
models: review and applicability to chip scale packages. Microelectron. Reliab. 40(2), 231244 (2000)
127. O. Salmela, K. Andersson, A. Perttula, J. Sarkka, M. Tammenmaa, Modified Engelmaiers model taking account of different stress levels. Microelectron. Reliab. 48(5), 773780 (2008)
128. L. Zhang, S.B. Xue, F.Y. Lu, Z.J. Han, S.L. Yu, Z.M. Lai,
Fatigue life prediction for fine pitch device solder joints based
on creep model. J. Mech. Eng. 45(9), 279284 (2009)
129. N. Paydar, Y. Tong, H.U. Akay, A finite element study of factors affecting fatigue life of solder joints. J. Electron. Packag.
116(4), 265273 (1994)
130. W. Engelmaier, Fatigue life of leadless chip carrier solder joints
during power cycling. IEEE Trans. Components Hybrids
Manufact. Technol. 6(3), 232237 (1983)
131. Y.E. Shin, J.H. Lee, Y.W. Koh, C.W. Lee, A study on lBGA
solder joints reliability using lead-free solder materials. KSME
Int. J. 16(7), 919926 (2002)
132. C. Tong, S.K. Zeng, Y.X. Chen, Finite element analysis simulations of life prediction for PBGA solder joints under thermal
cycling. Trans. China Weld. Inst. 28(10), 8992 (2007)
133. X.S. Ma, J.J. Chen, Affects of low thermal expansion coefficient
underfill on reliability of flip chip solder joint. Mach. Des. Res.
21(3), 6870 (2005)
134. S.B. Park, I.Z. Ahmed, Shorter field life in power cycling for
organic packages. J. Electron. Packag. 129(1), 2834 (2007)
135. S. Chen, T. Lee, J. Lee, N.F. Feng, in Solder joint thermal
fatigue analysis of 48-FCBGA. 7th International Conference on
Electronics Packaging Technology (2006)
136. A. Syed, in Updated life prediction models for solder joints with
removal of modeling assumptions and effect of constitutive
equations. 7th International Conference on Thermal, Mechanical
and Multiphysics Simulation and Experiments in Micro-Electronics and Micro-Systems (2006)
137. X.Y. Li, Z.S. Wang, Thermo-fatigue life evaluation of SnAgCu
solder joints in flip chip assemblies. J. Mater. Process. Technol.
183(1), 612 (2007)
138. S. Wiese, A. Schubert, H. Walter, R. Dudek, E. Feustel Meusel, B.
Michel, in Constitutive behaviour of lead-free solders vs. leadcontaining solders -experiments on bulk specimens and flip-chip
joints. Electronic Components and Technology Conference (2001)
139. S. Wiese, K.J. Wolter, Creep of thermally aged SnAgCu-solder
joints. Microelectron. Reliab. 47(23), 223232 (2007)
140. A. Syed, in Accumulated creep strain and energy density based
thermal fatigue life prediction models for SnAgCu solder joints.
54th Electronic Components and Technology Conference, Las
Vegas, NV, pp. 737746 (2004)
141. K. Andersson, O. Salmela, A. Perttula, J. Sarkka, M. Tammenmaa, in Measurement of acceleration factor for lead-free
solder (SnAg3.8Cu0.7) in thermal cycling test of BGA components and calibration of lead-free solder joint model for life
prediction by finite element analyses. 6th International Conference on Thermal, Mechanical and Multiphysics Simulation and
Experiments in Micro-Electronics and Micro-Systems (2005)
142. M.W. Brown, K.J. Miller, A theory for fatigue failure under
multi-axial stress-strain conditions. Proc. Inst. Mech. Eng.
187(65), 745755 (1973)
143. H. Gao, X. Chen, Effect of axial ratcheting deformation on
torsional low cycle fatigue life of lead-free solder Sn-3.5Ag. Int.
J. Fatigue 31(2), 276283 (2009)
144. D.M. Pierce, S.D. Sheppard, P.T. Vianco, A general methodology to predict fatigue life in lead-free solder alloy interconnects. J. Electron. Packag. 131(1), 011008 (2009)
145. M. Nozaki, M. Sakane, Y. Tsukada, H. Nishimura, Creep-fatigue life evolution for Sn-3.5Ag solder. J. Eng. Mater. Technol.
128(2), 142150 (2006)

123

190
146. A. Perkins, S.K. Sitaraman, Universal fatigue life prediction
equation for ceramic ball grid array (CBGA) packages. Microelectron. Reliab. 47(12), 22602274 (2007)

123

J Mater Sci: Mater Electron (2013) 24:172190


147. J.H. Lau, S.H. Pan, C. Chang, A new thermal-fatigue life prediction model for wafer level chip scale (WLCSP) solder joints.
J. Electron. Packag. 124(3), 212220 (2002)

Das könnte Ihnen auch gefallen