Sie sind auf Seite 1von 9

Journal of Alloys and Compounds 457 (2008) 97105

Review

Tensile properties and microstructural characterization of Sn0.7Cu0.4Co


bulk solder alloy for electronics applications
Cristina Andersson a, , Peng Sun a,b , Johan Liu a,b
a

Department of Microtechnology and Nanoscience, Bionano Systems Laboratory, Chalmers University of Technology, and Sino Swedish Microsystem
Integration Technology (SMIT) Center, Kemivagen 9, Se-412 96, Gothenburg, Sweden
b Key State Lab for New Displays and System Applications and SMIT Center, Shanghai University, Box 282,
Yanchang Road 149, Shanghai 200072, PR China
Received 22 November 2006; received in revised form 1 March 2007; accepted 2 March 2007
Available online 12 March 2007

Abstract
The ternary SnCuCo system eutectic composition was obtained by means of CALPHAD (CALculation of PHAse Diagram) methodology and
it was found to be 0.4% Co and 0.7% Cu (wt%) with a melting point of 224 C. The tensile behavior of this alloy was investigated at different strain
rates (105 , 104 and 103 s1 ) and compared to both Sn37Pb and Sn4.0Ag0.5Cu. The Sn4.0Ag0.5Cu alloy depicts the highest ultimate tensile
strength (UTS) followed by the Sn37P and finally the Sn0.7Cu0.4Co system. The elastic modulus was also higher for the Sn4.0Ag0.5Cu
followed by the Sn0.7Cu0.4Co and last the Sn37Pb. The microstructure of the Sn0.7Cu0.4Co alloy is composed of two types of intermetallic
phases, (Cu,Co)6 Sn5 and (Co,Cu)Sn2 dispersed in a Sn-rich matrix. The microstructure of this alloy proved to be very stable, after aging at 150 C for
24 h. The eutectic Sn0.7Cu0.4Co solder alloy can therefore be a very good alternative for the SAC alloys for surface mount technology applications.
2007 Elsevier B.V. All rights reserved.
Keywords: Tensile properties; Lead-free; Solder alloy; SnCuCo; Strain rate

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Calculation of eutectic point of the SnCuCo system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2. Solder specimen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Test procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1. Elongation to failure (EF) tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2. Strain rate change (SRC) tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1. Microstructural analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Stability of the microstructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3. Elongation to failure (TF) tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1. Strain hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4. Strain rate change (SRC) tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5. Fracture surface analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author at: Room A518, Bionano Systems Laboratory, MC2, Chalmers University of Technology, Kemivagen 9, 412 96 Goteborg, Sweden.
Tel.: +46 31 772 3074; fax: +46 31 772 3622.
E-mail address: cristina.andersson@chalmers.se (C. Andersson).
0925-8388/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2007.03.028

98
98
98
98
99
99
99
99
99
100
100
102
103
104
104
105
105

98

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

1. Introduction
Many lead-free solder alloys have already been proposed as
alternatives for the conventional eutectic Sn37Pb, and many
of them have already been greatly emphasized [13]. Lead-free
solders span from binary to ternary systems, and are generally
based on tin (Sn) with small additions of alloying elements such
as silver (Ag), copper (Cu), zinc (Zn), indium (In), etc. Some
popular binary systems already investigated include SnAg,
SnCu, SnZn, SnBi, and SnIn, while the ternary systems
include SnAgCu, SnAgBi, SnZnBi, among others. As
a replacement for the Sn37Pb solder alloy for surface mount
technology (SMT) applications, the International Electronics
Manufacturing Initiative (iNEMI) recommended in the year of
2000 the ternary Sn3.9Ag0.6Cu (SAC) alloy, based on availability, patent considerations and compatibility issues [4]. The
new lead-free alloys present, however, some disadvantages
compared to the Sn37Pb alloy. Apart from issues related to
physical and mechanical properties, the material cost is also a
disadvantage. The majority of all lead-free solder alloys are more
expensive than the lead containing ones, mainly because many of
these new alloys contain valuable and expensive metals, such as
silver.
The main purpose of this work was to investigate the
tensile properties of the newly developed Sn0.7Cu0.4Co leadfree solder alloy at different strain rates and to compare the
results with both Sn4.0Ag0.5Cu and eutectic Sn37Pb. The
microstructure of this alloy was also thoroughly investigated in
the context of this work.
By using the proposed alloy composition, and by substituting
the very expensive Ag for the much cheaper Co, will result in
cost savings for the whole electronics manufacturing industry.
As shown in Fig. 1, the price of 1 kg of Sn4.0Ag0.5Cu metal
alloy is 2.49 times that of Sn0.7Cu0.4Co, and 3.7 times that
of Sn37Pb. One kg of Sn0.7Cu0.4Co, on the other hand, is
merely 1.5 times that of Sn37Pb. The different element prices,
in USD/kg, used to calculate the alloy prices were the following: 9.23 for Sn, 7.81 for Cu, 359.25 for Ag and 33.63 for
Co [5].

Fig. 1. Price in USD per kg for Sn0.7Cu0.4Co, Sn4.0Ag0.5Cu and


Sn37Pb alloys.

Fig. 2. Binary phase diagram at the Sn-rich corner for CoCu interaction.

2. Experimental
2.1. Calculation of eutectic point of the SnCuCo system
In order to find the eutectic composition of the SnCuCo solder system, the
SnCo binary system was thoroughly assessed using CALPHAD (CALculation
of PHAse Diagram) methodology. The composition of the Sn-rich eutectic point
of the SnCuCo system was found to be 0.4 wt% Co and 0.7 wt% Cu with a
melting temperature of 224 C (see Fig. 2). All the detailed calculations have
been previously published in [6].

2.2. Solder specimen


The solder materials tested in the present work were the Sn0.7Cu0.4Co,
the Sn4.0Ag0.5Cu and the eutectic Sn37Pb as reference (all the alloy compositions in the context of this work are given in wt%). The dog-bone specimens
were prepared by machining ingot bars into standard testing samples [79]. The
ingot bars were manufactured by melting the bulk materials and pouring them
into moulds of plaster. The cooling was performed in air and at room temperature. A schematic of the dog-bone bulk specimen used is shown in Fig. 3, where
a is the thickness of the reduced section, b is the width of the reduced section,
c is the overall length, d is the length of the grip section, e is the width of the
grip section and L0 is the gage length. The dimensions of the specimen are
shown in Table 1. The specimens have a rectangular cross-sectional area and
a radius of curvature of 5 mm to prevent any stress concentration due to sharp
corners.

Fig. 3. Schematic and geometry of bulk sample.

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105


Table 1
Dimensions of bulk specimen (mm)
a
b
c
d
e
L0

99

3. Results
6
8
60
10
12
30

2.3. Test procedure


Two types of tensile tests were performed in order to measure the strain rate
sensitivity of the alloys, namely, elongation to failure (EF) and strain rate change
(SRC) tests. Both types of tests were carried out with an Instron 4505 testing
machine, at room temperature and at three different strain rates (103 , 104 and
105 s1 ). The strain rate was controlled by a crosshead displacement rate. The
SRC tests were performed in order to better analyze the influence of the strain
rate on the stress response without the effect of sample-to-sample geometry
variation, which causes scatter in the tensile tests results. Therefore, the strain
rate sensitivity analysis is performed using the same sample.
2.3.1. Elongation to failure (EF) tests
For the EF tests, the strain rate was maintained until specimen failure
occurred. Four to five specimens were tested under each test condition (45
replications).
2.3.2. Strain rate change (SRC) tests
For the SRC tests the strain rate was kept at the lowest value up to a total plastic strain of approximately 5% and then increased in two steps to the subsequent
strain rate test levels. Each subsequent strain rate level was also maintained for
approximately 5% strain deformation. The SRC started with the lowest strain
rate value of 105 s1 and then changed to 104 and 103 s1 , respectively.
Four to five specimens were tested (45 replications).

3.1. Microstructural analysis


The microstructure of the studied alloys was analyzed both
by means of optical microscopy (OM) and scanning electron
microscopy (SEM).
Fig. 4 shows the EDX elemental mapping analysis of the
IMCs, including Sn, Ag and Cu found in the Sn4.0Ag0.5Cu
microstructure. For this alloy, two different groups of intermetallics existed in the -Sn-rich solder matrix. By means
of EDX, the larger particles were identified to be a SnAg
compound. The average compositional value of this SnAg
phase was Sn:Ag = 29.3:70.7 and therefore, these IMCs could
be denoted as Ag3 Sn. The morphology of the Ag3 Sn intermetallics depends on the cooling rate; slow cooling rate results
in the formation of larger needle-like Ag3 Sn particles dispersed
through the Sn-rich matrix, while fast cooling rate results in finer
Ag3 Sn particles. Small Ag3 Sn particles have been identified to
reinforce the solder matrix [10] and to improve the mechanical properties of the solder alloy, however, reducing the solder
joints ductility, inducing brittle fracture, influencing crack initiation and accelerating the fatigue crack growth kinetics due
to de-cohesion of large Ag3 Sn particles especially when having
a branch-like morphology that deteriorates the homogeneity of
the mechanical properties [11,12]. The finer particles dispersed
in the SAC alloy were identified as being a CuSn phase composed of 53.6 at% Cu and 46.4 at% Sn, indicating the Cu6 Sn5
phase.

Fig. 4. SEM/EDX elemental mapping of solder matrix in Sn4.0Ag0.5Cu bulk solder alloy (as-solidified): (a) SEM picture of the area analyzed, (b) elemental
distribution of Sn, (c) elemental distribution of Ag, and (d) elemental distribution of Cu.

100

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

Fig. 5. SEM/EDX elemental mapping in Sn0.7Cu0.4Co bulk solder alloy: (a) SEM picture of the area analyzed, (b) elemental distribution of Sn, (c) elemental
distribution of Co, and (d) elemental distribution of Cu.

Fig. 5 shows the EDX-elemental mapping analysis performed on the microstructure of the Sn0.7Cu0.4Co alloy,
showing the IMCs and including Sn, Cu and Co elements.
According to the EDX analysis, the atomic composition of the
Co-rich particles was Sn:Co:Cu, 70:26:4 at% corresponding to
the CoSn2 , with the substitution of Cu (solid solution) into the
CoSn2 phase, resulting in (Co,Cu)Sn2 . Different Co-rich particles were analyzed and all showed some Cu substitute, in
the range of 2.655.86 at%, inside the CoSn2 phase. The Curich intermetallics showed an atomic ratio between Sn:Cu:Co of
59.5:34.5:6, and can therefore be denoted as the (Cu,Co)6 Sn5
IMC phase with the substitution of Co into the Cu6 Sn5 phase.
The fact that Co does not form any intermetallics with either
Ag or Cu, and that the only intermetallics found between the
Co and Sn were rod-like CoSn2 has also been acknowledged
by other researchers [13,14]. The reason for this behavior was
the fact that Co has little solubility in the -Sn matrix, Ag and
Cu. This behavior was also confirmed by using CALFAD, which
showed that the only two types of stable IMCs that can exist in the
eutectic SnCuCo alloy are the phases Cu6 Sn5 and CoSn2 [5].
According to Fig. 5, it is also difficult to distinguish a shape
difference between the Cu6 Sn5 and the CoSn2 IMCs found in
the Sn0.7Cu0.4Co alloy. In contrast to the SAC alloys, where
there is a visible and evident difference in shape between the
Ag3 Sn and the Cu6 Sn5 IMCs, a chemical analysis is necessary for the SnCuCo alloy in order to identify the different
IMCs.

3.2. Stability of the microstructure


The three alloys were aged at 150 C for 24 h to study
their microstructural stability. The Sn37Pb alloy shows a
visible coarsening, depicting larger Pb-rich domains. The
Sn4.0Ag0.5Cu alloy also presents some kind of intermetallic
coarsening and dispersing. As shown in Fig. 6 (ad), the elongated Ag3 Sn and the smaller rounder Cu6 Sn5 particles appear
enlarged/coarsened and at the same time one can see larger
Sn-rich regions.
Regarding the Sn0.7Cu0.4Co alloy, no apparent changes
were observed between the as-solidified and the aged material state. The Cu6 Sn5 and the CoSn2 intermetallics seem to
be of the same size and shape, and have the same distribution
in the Sn-rich matrix. This alloy presents therefore a very stable
microstructure (for the aging conditions used).
3.3. Elongation to failure (TF) tests
All the three alloys investigated exhibited indistinct yield,
however, distinct work hardening and significant strain to failure
(ductility). They also showed strong visco-plastic behavior and
relatively strong strain rate sensitivity. Fig. 7 shows three stress
strain curves for the Sn0.7Cu0.4Co alloy for the three different
strain rates. As shown, the mechanical properties of this alloy are
strongly dependent on the strain rate applied, and increasing the
strain rate from 105 to 103 s1 results in approximately a two-

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

101

Fig. 6. Optical micrographs at 100 magnification of the microstructure of: as-solidified (a) SAC, (b) Sn0.7Cu0.4Co and (c) Sn37Pb, and after aging at 150 C
for 24 h (d) SAC, (e) Sn0.7Cu0.4Co and (f) Sn37Pb.

Fig. 7. Three engineering stressstrain curves for three Sn0.7Cu0.4Co samples tested at 105 , 104 and 103 s1 , respectively.

fold increase in UTS [15,16]. Furthermore, the total elongations


increased in average with a decrease in strain rate. The other two
alloys followed the same trend as the Sn0.7Cu0.4Co.
The E-modulus, defined as the slope of the linear part in the
stressstrain relationship, was also analyzed. The elastic modulus is a material property and has therefore to have a constant
value. When measuring the E-modulus it is important to be aware
that solders have both elastic and plastic response at the instant of
load application. When tensile tests are performed at low strain
rates, the apparent linearity of the stress versus strain curve is not

accurate, because both plastic and elastic strains are present. To


be sure that the E-modulus is correctly measured, the test should
be performed at a strain rate value for which creep deformation
can be neglected (low stresses so that minimal yielding occurs
and a rate high enough to eliminate any visco-plastic effects). To
be sure that high enough strain rate is used the stressstrain slope
must be obtained for increasing loading rates. When there is no
change in the measured slope with further increases in rate, then
it is safe to assume that the solder is undergoing purely elastic
deformations [17]. For the Sn40Pb, the strain rate at which no
creep takes place has been identified as 2 102 s1 [18]. This
means that strain rates larger than 2 102 s1 are necessary in
order to obtain time independent tensile properties for that alloy.
The effect of strain rate is probably one of the reasons why there
is such large scattering in tensile properties data of solder alloys.
Therefore, and for the strain rates analyzed in this work, the values for the E-modulus are taken from the tests performed with
a strain rate value of 103 s1 .
Table 2 shows the tensile test results for the three solder
alloys tested. All the measured properties, except for E-modulus
and uniform strain, increase with faster strain rate (for the
strain rate values investigated). Similar trends have also been
reported by other researchers [15,19]. The E-modulus is in average 20 GPa for the Sn37Pb, 40 GPa for the SAC and 37.1 GPa
for the Sn0.7Cu0.4Co. The SAC depicts therefore the highest
value followed by the Sn0.7Cu0.4Co and then the eutectic

Table 2
Maximum load, UTS, yield stress, E-modulus and uniform strain for the three alloys tested
Sn37Pb

Sn4.0Ag0.5Cu

Sn0.7Cu0.4Co

103 s1

104 s1

105 s1

103 s1

104 s1

105 s1

103 s1

104 s1

105 s1

Maximum load (N)


UTS (MPa)
Yield stress (MPa)
E-Modulus (GPa)

2052
42.7
31.3
20

1697
35.4
26.6
(28)

1225
25.5
22.5
(30)

2271
47.3
33.7
40

1892
39.4
28.8
(43.8)

1585
33
23
(50)

1460
30.4
19.1
37.1

1116
23.4
18.1
(40.7)

861
17.3
10.9
(42.6)

Uniform strain (%)

9.8

9.5

8.6

7.1

7.5

102

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

Table 3
Average UTS, yield stress and E-modulus, and ratios between the three alloys
tested, for the strain rate of 103 s1
Alloy

UTS

Ratio

Yield stress

Ratio

E-Modulus

Ratio

Sn37Pb
Sn4.0Ag0.5Cu
Sn0.7Cu0.4Co

42.7
47.3
30.4

1
1.11
0.71

31.3
33.7
19.1

1
1.08
0.61

20
40
37.1

1
2
1.86

Sn37Pb. The E-modulus of the SAC alloy is in average 2 times


that of Sn37Pb and 1.08 times that of Sn0.7Cu0.4Co. The
E-modulus of the Sn0.7Cu0.4Co is in average 1.85 that of
Sn37Pb (see Table 3).
Therefore, taking the strain rate value of 103 s1 as reference, in terms of tensile strength, the Sn0.7Cu0.4Co alloy has
an average UTS value of 30.4 MPa, which is 0.71 times that
of Sn37Pb and 0.64 that of SAC. For the same strain rate, the
yield stress of the Sn0.7Cu0.4Co alloy is in average 19.1 MPa
which is 0.61 times that of Sn37Pb and 0.57 times that of SAC.
The SAC alloy shows the highest tensile strength of the three
solders investigated. It is 1.11 times that of Sn37Pb and 1.56
times that of Sn0.7Cu0.4Co. The yield stress of this alloy is
also the highest of the three, being 1.08 times that of Sn37Pb
and 1.76 times that of Sn0.7Cu0.4Co (see Table 3).
Fig. 8 shows that there is a linear relationship between the
UTS and the strain rate (loglog plot), and this relationship can
be expressed by Eq. (1):
= Cm ,

(1)

where is the tensile strength (UTS), the strain rate, m the strain
rate sensitivity (given by the slope of each one of the lines) and
C is a constant. The m values calculated for each alloy were the
following: 0.078 for SAC, 0.10 for Sn0.7Cu0.4Co and 0.112
for Sn37Pb. These values are in good agreement with those
presented by [7], where m values of 0.09 and 0.08 were calculated for the eutectic Sn37Pb and for the Sn3.5Ag0.75Cu,
respectively. The m value describes the capacity to resist necking
[20]. From these values we could then expect that the lead-free
solders would present a lower resistance to necking compared
to the Sn37Pb, however, there are other factors that also affect

this behavior, namely, strain hardening, which therefore, is also


analyzed in the context of this work.
According to [20], super-plasticity is only observed when the
m value is greater than 0.3. Since the m values obtained in this
work are lower than 0.3, super-plasticity was not observed for
the alloys tested under the present conditions. Super-plasticity
has, however, been observed for Sn37Pb tested at 150 C [20].
Super-plasticity takes place normally when the deformation temperature is higher than 0.5Tm .
3.3.1. Strain hardening
The stressstrain curves, in the region of uniform plastic
deformation, do not increase proportionally with strain. The
material is then said to strain harden. Strain hardening opposes
the plastic instability of the material and is the phenomenon
whereby a ductile material becomes harder and stronger as it is
plastically deformed. For super-plastic materials, for example,
the deformation hardening is very small. When one loads a
sample beyond the yield stress, such that plastic deformation
occurs, the material will strain harden, and this will occur until
the material reaches the ultimate tensile strength. During strain
hardening the dislocation density increases with deformation
resulting in further hindering of the dislocation movement. As
the dislocation density increases, and the resistance to their
motion increases, the imposed stress that is necessary to deform
the material increases with increased strain hardening [21]. It is
known that most materials strain harden at room temperature,
and solders, being ductile materials, also follow this rule. There
is a price for strain hardening, that is, ductility. The higher the
strain hardening (measured in percent) the lower the ductility
(% elongation).
A popular empirical mathematical description of strain hardening has been developed by Holloman in 1945 which describes
the shape of the engineering stressstrain curve [22]. The relationship between true stress and true strain is given by Eq. (2):
= kn ,

(2)

where and are the true stress and the true strain, k the strength
coefficient and n is the strain-hardening exponent which can be
obtained from a loglog plot of true stress versus true strain. The
strain hardening exponent is equal to zero for perfectly plastic
solids, equal to 1 for perfectly elastic solids, and for the majority
of all metals this value is between 0.1 and 0.5.
Table 4 lists the n and k values for the three solder alloys
tested. As shown in Table 4, the highest strain hardening exponent is depicted by the SAC followed by the Sn0.7Cu0.4Co
and last the Sn37Pb. This behavior is in good agreement with
the values of the shear modulus measured, which was higher for
the SAC, followed by the Sn0.7Cu0.4Co and last the Sn37Pb
alloy.
Table 4
Strain hardening exponent n and strength coefficient k for the three alloys tested

Fig. 8. Engineering tensile stress vs. strain rate for the Sn37Pb, SAC and
Sn0.7Cu0.4Co alloys.

103 s1

Sn0.7Cu0.4Co

Sn4.0Ag0.5Cu

Sn37Pb

0.13

26.9

0.26

36.5

0.11

49.5

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

103

Table 5 lists the stress exponent w and the strain rate sensitivity m for each solder alloy tested, plus the m values obtained
from the EF tests. As shown in Table 5, the m values calculated from the EF tests are slightly lower compared to the ones
obtained from the SRC tests. The stress exponents w for the
three solder alloys are quite high, with values between 7.94 for
the Sn0.7Cu0.4Co and 12.48 for the SAC. The w value for
the Sn0.7Cu0.4Co is very close to that of pure tin, 7.6 [23],
which is a result of the high tin content of the Sn0.7Cu0.4Co
alloy, namely 98.9 wt%.

Fig. 9. True stress vs. true strain data according to the SRC method for the
Sn0.7Cu0.4Co alloy (initial strain rate of 105 s1 ).

Fig. 10. Strain rate vs. true stress obtained from the SRC tests, for the three
alloys tested.

3.4. Strain rate change (SRC) tests


For the SRC tests the strain rate was maintained at the lowest value of 105 s1 for approximately 5% strain and then it
was increased to the subsequent strain rates of 104 and 103 ,
respectively (see Fig. 9). Fig. 10 shows the relationship between
the strain rate and true stress obtained from the SRC test curves.
The slope of each line (fitted to the experimental data) gives the
so called stress exponent, w, and the strain rate sensitivity, m, is
given by 1/w [7]. The m value can also be calculated from the
curves represented in Fig. 9, by using Eq. (3):
m=

log(2 /1 )
,
log(2 /1 )

(3)

where m is the strain rate sensitivity, 1 , 1 and 2 , 2 are the


stress and strain rate values for the SRC curves at strain rate
level 1 and strain rate level 2, respectively.
Table 5
Stress exponent w and strain rate sensitivity m from SRC and EF tests
Alloy

Sn37Pb
Sn4.0Ag0.5Cu
Sn0.7Cu0.4Co

SRC

EF

12.48
7.94
8.73

0.080
0.126
0.115

0.078
0.102
0.112

Fig. 11. SEM micrographs of the fracture surface of three Sn0.7Cu0.4Co


specimens tested at strain rates of (a) 103 s1 , (b) 104 s1 , and (c) 105 s1 .

104

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

Fig. 12. Percentage in area reduction at fracture surface site as a function of


strain rate for all three alloys tested.

3.5. Fracture surface analysis


The majority of all specimens tested under the present conditions fractured during test. The fracture surfaces of the fractured
specimens were analyzed by means of SEM to determine the
causes of failure.
Fig. 11 shows the SEM pictures of the fracture surfaces of
three Sn0.7Cu0.4Co specimens tested at different strain rates.
The analysis was performed in order to investigate how the
fracture surface changes as a function of strain rate. At 500
magnification, it is visible that the fracture surface is dependent of strain rate. With slower strain rate, the fracture surface
shows a more typical ductile fracture surface with dimples. The
dimples are a result of micro-void coalescence. For the typical
cup and cone fracture mode that normally happens in ductile
materials, dimples are normally equiaxed in shape in the sample centre and elongated, pointing toward the origin of failure,
on the shear edges. For the 103 s1 strain rate, the fracture
surface does not show the typical surface with small dimples
but, it shows signals of faster material rupture. No SEM pictures are shown for the other two materials analyzed since there
are already extensive amount of published data on those alloys.
The authors can, however, state that the fracture surface of the
SAC alloy also presented a dimple rupture mode, with a mix-

Fig. 13. OM image of surface near fracture site of Sn0.7Cu0.4Cos specimen


tested at a strain rate of 104 s1 (the arrow represents the load direction).

ture of both large and small dimples, which is typical of ductile


fracture.
The percentage area reduction was also analyzed, both as a
function of solder material and strain rate. As shown in Fig. 12,
there is an increase in area reduction (at the necked region) with
decreased strain rate. The cross-sectional area of the specimens
tested at 105 s1 is smaller compared to the cross-sectional
area of the specimens tested at 103 s1 , proving that a larger
elongation took place for the slowest strain rate (corroborating
the preliminary results from the stressstrain curves). The alloy
showing the largest area reduction is the Sn37Pb followed by
the SAC and last the Sn0.7Cu0.4Co. This is a proof that the
eutectic Sn37Pb is more ductile compared to the other two
materials.
Fig. 13 shows an optical micrograph of the surface near
the fracture site of a Sn0.7Cu0.4Co specimen, depicting a
striation-like pattern, which is an indication of grain boundary
sliding taking place during the tensile test. This is also an indication of intergranular fracture, meaning that cracks grew and
propagated along the grain boundaries of the solder material
[24].

4. Conclusions
For the tensile strain rates investigated, and the test parameters used, the following conclusions can be drawn: the strain rate
has a substantial effect on the measured tensile strength of the
solder material; an increase in strain rate results in an increase
in strength.
For all the strain rates analyzed the strongest alloy is
the Sn4.0Ag0.5Cu, followed by the Sn37Pb and the
Sn0.7Cu0.4Co. The highest E-modulus is also shown by
the SAC followed by the Sn0.7Cu0.4Co alloy and last the
Sn37Pb.
The microstructure of the Sn0.7Cu0.4Co alloy is composed of two different intermetallic phases, the (Co,Cu)Sn2
phase and (Cu,Co)6 Sn5 dispersed in a Sn-rich matrix. The
Sn0.7Cu0.4Co alloy presents a very stable microstructure (for
the aging conditions used).
From the strain rate sensitivity values calculated, it can be
concluded that both lead-free solders present a lower resistance
to necking compared to the Sn37Pb, however, the highest strain
hardening exponent n, which decreases with increased strain
rate, is depicted by the SAC followed by the Sn0.7Cu0.4Cu
and finally the Sn37Pb. These n values also prove that the most
ductile alloy is the Sn37Pb, followed by the SAC and finally
the Sn0.7Cu0.4Co.
The fracture surface of all the alloys analyzed in this work is
dependent of strain rate. With slower strain rate (105 s1 ), the
fracture surface shows a more typical ductile fracture surface
with dimples. There is also an increase in area reduction (at the
necked region) with decreased strain rate. The alloy showing
the largest area reduction is the Sn37Pb, followed by the SAC
and last the Sn0.7Cu0.4Co. This corroborates once again that
the eutectic Sn37Pb is more ductile compared to the other two
alloys.

C. Andersson et al. / Journal of Alloys and Compounds 457 (2008) 97105

Acknowledgments
The authors are grateful for the financial support provided
by the EU project, Flex-Eman under the contract no: COOPCT-2003-507983 and the Swedish Foundation for International
Cooperation in Research and Higher Education (STINT), and
last but not least all the SMIT-centre member companies.
References
[1] M. Abtew, G. Selvaduray, Mater. Sci. Eng. 27 (2000) 95141.
[2] J.H.L. Pang, B.S. Xiong, T.H. Low, Int. J. Fatigue 26 (2004) 865872.
[3] C.M.L. Wu, D.Q. Yu, C.M.T. Law, L. Wang, Mater. Sci. Eng. R44 (2004)
144.
[4] http://www.nemi.org/projects/ese/if assembly.html.
[5] http://www.metalprices.com/FreeSite/metals/ag/ag.asp.
[6] L. Liu, C. Andersson, J. Liu, J. Electron. Mater. 33 (9) (2004) 935939.
[7] Zugproben, DIN 50 125, ISO 6892-1984.
[8] Standard methods of tension testing of metallic foil, E345-93, Annual Book
of ASTM standards, vol. 03.01, 1993.
[9] Standard methods of tension testing of metallic materials, E8-93, Annual
Book of ASTM standards, vol. 03.01, 1993.
[10] A. Sharif, Y.C. Chan, M.N. Islam, M.J. Rizvi, J. Alloys Compd. 388 (1)
(2005) 162167.
[11] P.L. Liu, J.K. Shang, J. Electron. Mater. 29 (5) (2000) 622627.

105

[12] K.S. Kim, S.H. Huh, K. Suganuma, Mater. Sci. Eng. A333 (1/2) (2003)
223236.
[13] K.S. Kim, S.H. Hug, K. Suganuma, Microelectron. Reliab. 43 (2003)
259267.
[14] P. Sun, C. Andersson, X. Wei, Z. Cheng, D. Shangguan, J. Liu, Proceedings
of the Conference on High Density Microsystem Design and Packaging and
Component Failure Analysis (HDP06), Shangai, China, 2006.
[15] I. Shohji, T. Yoshida, T. Takahiko, S. Hioki, Mater. Sci. Eng. A366 (2004)
5055.
[16] I. Shohji, T. Yoshida, T. Takahashi, S. Hioki, J. Mater. Sci.: Mater. Electron.
15 (2004) 219223.
[17] T.J. Kilinski, J.R. Lesniak, B.I. Sandor, in: J.H. Lau (Ed.), Solder Joint
ReliabilityTheory and Applications, Van Nostrand Reinhold, NY, 1991.
[18] H. Nose, M. Sakane, Y. Tsukada, H. Nishimura, J. Electron. Pack. 125
(2003) 5966.
[19] W.J. Plumbridge, C.R. Gagg, J. Mater. Sci.: Mater. Electron. 10 (1999)
461468.
[20] F. Lang, H. Tanaka, O. Munegata, T. Taguchi, T. Narita, Mater. Charact.
54 (2005) 223229.
[21] W.D. Callister Jr. (Ed.), Materials Science and EngineeringAn Introduction, third ed., John Willey & Sons, Inc., NY, 1994 (Chapter: Dislocations
and Strengthening Mechanisms, pp. 149180).
[22] J.H. Holloman, Trans. Am. Inst. Min. Metall. Petrol. Eng. 162 (1945)
268272.
[23] R.J. McCabe, M.E. Fine, Metall. Mater. Trans. 33A (2002) 15311539.
[24] F. Zhu, H. Zhang, R. Guan, S. Liu, Proceedings of the IEEE 6th International
Conference on Electronic Packaging Technology, 2005, pp. 466470.

Das könnte Ihnen auch gefallen