Вы находитесь на странице: 1из 10

Materials and Design 43 (2013) 4049

Contents lists available at SciVerse ScienceDirect

Materials and Design


journal homepage: www.elsevier.com/locate/matdes

Microstructure, mechanical properties, and deformation behavior


of Sn1.0Ag0.5Cu solder after Ni and Sb additions
A.A. El-Daly a,, A.E. Hammad a, A. Fawzy b, D. A. Nasrallh a
a
b

Physics Department, Faculty of Science, Zagazig Univ., Zagazig, Egypt


Physics Department, Faculty of Education, Ain Shams Univ., Cairo, Egypt

a r t i c l e

i n f o

Article history:
Received 23 May 2012
Accepted 27 June 2012
Available online 5 July 2012
Keywords:
Lead-free solder
SnAgCu alloys
Non-ferrous alloys
Microstructure
Mechanical properties

a b s t r a c t
Minor alloying addition to solders has been an important strategy to improve the integrity and reliability
of Pb-free solders joint. In this study, the effects of 0.06Ni and 0.5Sb additives on the microstructure and
solidication behavior as well as the creep properties of Sn1.0Ag0.5Cu (SAC105) alloys were investigated. Results show that alloying of Ni and Sb resulted in considerably reduced undercooling, increased
eutectic area and extended volume fraction of proeutectic Sn of which the dendritic size was rened.
Moreover, with the addition of Ni and Sb into SAC105, signicant improvement in creep resistance of
(210%) and (350%) is realized when compared with the SAC105 solder alloy. Likewise, the creep life time
of SAC105 alloys was remarkably enhanced (23 times) with the minor alloying additions. An analysis of
the creep behavior at elevated temperatures suggested that the presence of hard Ni3Sn4 IMC particles and
the solid solution hardening effects which appeared, respectively, in the Ni-doped and Sb-doped alloys
could increase the resistance to dislocation movement, which improves the creep properties.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
The development of lead-free solders has become impartment
for material researchers due to health and environmental concerns
regarding the high toxicity of lead. The eutectic SnAgCu (SAC)
solders are considered one of the most popular solders for surface
mount technology (SMT) or ball grid array (BGA) applications. Because of their high moduli, SAC solders are not very satisfactory for
certain applications requiring high impact asset, such as in mobile
electronics [1]. In addition, their high joint strength often yields a
serious chip joining problem such as chip-to-package interaction
(CPI) or white bump defects in a back-end-of-line (BEOL) structure.
Consequently, the performance of SAC solder joints under high
strain rate and large temperature ranges typical of drop impact situations is a major concern. To solve these problems, efforts have
been made to develop solders with a low melting point, higher
strength, better microstructure properties, and high creep resistance. Recently, SAC solders with a low Ag or Cu content have been
identied as promising candidates to replace the traditional SnPb
solder for ipchip interconnects [2]. The Sn1.0Ag0.5Cu
(SAC105) solder is a low-Ag lead-free solder alloy in the SAC family. It can offer a better resistance to drop failure in the interconnections. Besides, SAC105 solder possesses a major advantage in
that it provides a thinner brittle Ag3Sn IMC layer thickness during
Corresponding author.
E-mail addresses: dreldaly11@yahoo.com, dreldaly@zu.edu.eg (A.A. El-Daly).
0261-3069/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.matdes.2012.06.058

soldering process owing to their low Ag content [3]. In addition to a


reduced Ag content, micro-alloying is an effective method to modify the properties of lead-free solder alloys, but adding of these elements also could affect their microstructure and other properties
[4,5]. For this reason, some metal additives, such as Bi, Sb, In, Co,
Ga, Ni, Ge and nanoparticles have been introduced into SAC solders
to rene the microstructures and reduce the IMC growth [69].
One of the more noteworthy alloying elements is Ni. It was
reported that Ni addition to Sn-3.5Ag in amounts as small
as 6 0.1 wt.% could substantially hinder the Cu3Sn growth during
soldering as well as during the following solid-state aging.
Recently, some works also revealed that a small amount of Ni addition to the SnCu or SnAgCu solders could improve the solder
properties, such as the mechanical strength and wettability [10].
Although nickel has a strong inuence on solderability and
mechanical properties of solders, the detailed mechanism is not
fully understood. Interestingly, it was also reported that addition
of elements such as Sb could improve the mechanical properties,
mainly due to the solid solution hardening effects of Sb, formation
of the SnSb particles, and the presence of Sb could suppress the
coarsening the of b-Sn and rene the Ag3Sn precipitate, and thus,
improve the mechanical and thermal properties of the Sn-based
solders [1113].
For applications requiring dimensionally stable solders, a very
low creep rate at high homologous temperatures is required.
Understanding of creep properties of Sn-based alloys, which are
widely used in electronic products, is important for evaluating

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

41

Table 1
Chemical composition of the solders studied (wt.%).
Alloy

Cu

Ag

Ni

Sb

Sn

SAC(105)
SAC(105)0.06Ni
SAC(105)0.5Sb

0.5
0.5
0.5

1.0
1.0
1.0

0.06

0.5

Bal.
Bal.
Bal.

Fig. 1. XRD pattern for (a) SAC (105), (b) SAC-0.06Ni and (c) SAC-0.5Sb solder alloys.

the thermo-mechanical reliability of solder joints [14]. Creep


deformation of SAC-based solders is controlled by a combined
function of creep in the b-Sn dendrite matrix and the brittle IMC
particles. Since many electronic components undergo wide ranges
of temperature with different stresses, it is important to evaluate
the creep behavior of the used solders at various applied stresses
over different temperature ranges. Thus, the aim of this study is
to investigate the inuence of 0.06Ni and 0.5Sb additions on the
microstructure, solidication behavior and creep properties of
Sn1.0Ag0.5Cu solder alloy. The test conditions cover the temperatures and applied stresses, which are important for the assessment of solder joint reliability. The key factors that affect the
creep behavior of the solder alloy are discussed.
2. Experimental details
The nominal compositions of the three alloys are given in Table 1.
The three lead-free solder alloys Sn1.0Ag0.5Cu, Sn1.0Ag0.5Cu
0.5Sb and Sn1.0Ag0.5Cu0.06Ni are designed to understand the
effect of Ni and Sb additions on the microstructure, solidication
behavior and creep properties of Sn1.0Ag0.5Cu alloy. Hereafter,
these alloys are assigned as; SAC(105), SAC(105)0.06Ni and
SAC(105)0.5Sb, respectively. The lead-free solders were prepared
from Sn, Cu, Ag, Ni and Sb (purity 99.99%) as raw materials. The process of melting was carried out in a vacuum arc furnace under high
purity argon atmosphere to produce rod-like specimen with a diameter of approximately10 mm.The melt was held at 500 C for 1 h to
complete the dissolution of Sn, Cu, Ag, Ni and Sb and then poured in
a steel mold to prepare the chill cast ingot. A cooling rate of 68 C/s

Fig. 2. OM microstructures for (a) SAC (105), (b) SAC-0.06Ni and (c) SAC-0.5Sb
solder alloys.

was achieved, so as to create the ne microstructure typically found


in small solder joints in microelectronic packages. The microstructure was examined by optical microscopy (OM) and scanning electron microscopy (SEM) JSM-5410, Japan. A solution of 3%HCl,
2%HNO3 and 95% (vol.%) Ethyl alcohol was prepared and used to
etch the samples. Phase identication was based on two complimentary techniques, namely X-ray diffraction (XRD) and Energy
Dispersive X-ray Spectrometry (EDS). For XRD, we used a Philips diffractometer with monochromatic Cu Ka radiation and individual
phases and their crystal structures were identied by matching
the characteristic XRD peaks against JCPDS data. The X-ray diffractograms were recorded from bulk specimens and thus texture effects were likely. Differential scanning calorimetry (DSC)
(shimadzu DSC-50) was carried out to understand the melting and
freezing processes of the three solder alloys. Heating and cooling

42

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

(a)

(d)

(b)

(e)

(c)

(f)

Fig. 3. SEM microstructures of (a) SAC (105), (b) SAC-0.06Ni and (c) SAC-0.5Sb solder alloys and the corresponding EDS analysis of some locations.

the specimens in DSC were carried out at 5 C/min of heating rate in


Ar ow. Because of the different approaches in specimen design, test
setups and experimental methodology are necessary to investigate
solder creep behavior at different volumes. The present work represents one of these footsteps. The homogenized cast ingots were then
mechanically machined into a wire samples with a gauge length
marked 4  102 m for each samples and 2.5  103 m in diameter,
as developed in our previous work [15]. The small sized solder specimens used in this work had more advantageous than the test standards since the creep behavior follows a simple power law. Besides,
the metallization material changes not only the absolute creep
strength, but also the stress sensitivity (stress exponent) of the
creep behavior. Such effects could not be found in bulk specimens.
Details are described in [15,16]. To obtain samples containing the
fully precipitated phases and free from any plastic strain accumulation during machining, the samples were annealed at 130 C for
30 min, then left to cool slowly to room temperature. In order to
understand the effect of Ni and Sb additions on the mechanical
properties of SAC (105), tensile creep tests were carried out with a
tensile testing machine (Instron 3360 Universal Testing Machine).
Creep testing was performed at 25, 70, and 110 C after waiting time

of 5 min for the test temperatures to be reached. Each datum represents an average of three measurements. The environment chamber
temperature could be monitored by using a thermocouple contacting with specimen.
3. Results and discussion
3.1. X-ray diffraction analysis
XRD was conducted to identify the phase structures of
SAC(105), SAC(105)0.06Ni and SAC(105)0.5Sb solders and the corresponding patterns are presented in Fig. 1. In Fig. 1a, the Ag3Sn,
g-Cu6Sn5 IMC particles and the b-Sn phase were detected in
SAC(105) solder. From XRD analysis, while all Cu6Sn5 found in this
study was g-Cu6Sn5 having hexagonal lattice structure, the Cu3Sn
was not found in SAC(105) alloy samples. It is well known that
Cu3Sn growth is usually linked to the formation of Kirkendall voids,
which in turn increases the potential of interfacial brittle fracture,
whereas Cu6Sn5 does not induce the Kirkendall voids formation
[17]. For that reason, Cu6Sn5 is preferred to Cu3Sn based on the reliability of the solder. According to the phase diagram between Cu

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

43

Fig. 4. EDS analyses of (a) b-Sn, (b) Ni3Sn4 and (c) Ag3Sn particles in the respective alloys.

and Sn [18], g-Cu6Sn5 is the IMC found at a higher temperature,


while the monoclinic g0 -Cu6Sn5 IMC phase is the low-temperature
Cu6Sn5 phase. The excessive g-Cu6Sn5 growth during soldering and
the absence of g0 -Cu6Sn5 IMC phase in this study suggested that gCu6Sn5 does not transform to g0 -Cu6Sn5 during cooling period of
the solder alloys. Fig. 1b shows that Ni addition has a positive effect on the growth of Ni3Sn4 IMC. Ni is known to be one dominant
and widely used doping material in SAC solders due to its excellent
performance in improving solder microstructure and increasing
the drop lifetime of electronic assembly [19]. However, Fig. 1c
shows that no new peaks are detected with the addition of Sb to
the SAC(105) solder alloy. This implies that the Sb content of the
alloy predominantly dissolves in the b-Sn matrix to form a solid
solution.
3.2. Effect of Ni and Sb additions on the solidication microstructures
OM and SEM images are employed to analyze the microstructure of as-cast SAC(105), SAC(105)0.06Ni and SAC(105)0.5Sb alloys. Fig. 2a shows the typical OM microstructure of the
SAC(105) solder with the bright and dark phases being the primary
b-Sn phases and eutectic SAC phases, respectively. According to the
XRD analysis, the eutectic areas were found to contain Ag3Sn and
Cu6Sn5 IMCs. Addition of a small percentage of Ni and Sb elements
to the SAC(105) solder was observed to alter the as-solidied condition OM microstructure, as shown in Fig. 2b and c. The volume
fraction of the primary b-Sn phase of the SAC(105) alloys decreased
while the eutectic area increased. In Fig. 2b, the eutectic phase contained b-Sn, Cu6Sn5 ,Ni3Sn4 and the needle-like Ag3Sn phase as
conrmed by XRD analysis. With the addition of Sb, the grain size

of dendritic b-Sn phase and size of IMCs could be rened to get a


more stable microstructure and better mechanical properties.
The low and high magnication SEM microstructures of (a)
SAC(105), (b) SAC(105)0.06Ni and (c) SAC(105)0.5Sb solder alloys
are shown in Fig. 3. In SAC(105) specimen (Fig. 3a and d), the dark
primary b-Sn cells or grains with the average diameter of 60
90 lm are surrounded by light-gray eutectic regions. The microstructure of IMCs in the eutectic regions has ner dot-shaped
precipitates morphology and more needle-like morphology. The
dense IMCs precipitates in between b-Sn were identied as Cu6Sn5
and Ag3Sn eutectic according to EDS analysis shown in Fig. 4.
The addition of Ni rened the b-Sn cells to some extents. The
morphology of eutectic IMCs was also inuenced by Ni alloying.
According to microstructure, it can be noted that the coarse ber-like IMCs and ner dot-shaped precipitates rather than needle-like morphology can occur at the surface of b-Sn matrix
easily. As a result, it is expected that Ni3Sn4 IMC reinforced
SAC(105) solder alloy has experienced less serious surface defects,
thus providing the justication for its better performance as compared to the SAC(105) solder, and could restrict the dislocation motion (Fig. 3b and e.)
Conversely, alloying of Sb triggers the morphology change of
b-Sn grains into light-gray b-Sn primary dendrites surrounded by
dark eutectic regions of rened Cu6Sn5 and Ag3Sn IMC phases
(Fig. 3c and f). Due to the relatively low level of Sb, no SnSb IMC
phases were detected. Instead, Sb dissolves in the Sn matrix and
strengthens the alloy by solid solution hardening. However, it
can also be seen from Fig. 3c that b-Sn dendrites are associated
with the secondary dendrite arm spacing of 4 0.5 lm. Therefore,
the b-Sn dendrites are the predicted primary phase in SAC

44

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

Fig. 5. DSC results of SAC (105), SAC (105)0.06Ni and SAC (105)0.5Sb solder alloys
(a) during heating (endothermal) and (b) cooling (exothermal).

Table 2
Comparison of solidus temperature (Tonset) and liquidus temperature (Tend) for
SAC(105), SAC(105)0.06Ni and SAC(105)0.5Sb solder alloys from heating curve.
Alloy

(Tonset)(C)

Tend(C)

Pasty range
(Tend-Tonset)(C)

SAC(105)
SAC(105)0.06Ni
SAC(105)0.5Sb

216.4
217.0
219.5

234.5
233.6
237.6

18.1
16.6
18.1

Melting
temperature(C)
P1

P2

226.0
220.8
227.4

226.6

Table 3
Comparison of solidus temperature (Tonset) during heating, liquidus temperature
(Tonset) during cooling and undercooling range for SAC(105), SAC(105)0.06Ni and
SAC(105)0.5Sb solder alloys.
Alloy

(Tonset)
heating(C)

Tonset
cooling(C)

Undercooling (Th
Tc)(oC)

SAC(105)
SAC(105)0.06Ni
SAC(105)0.5Sb

216.4
217.0
219.5

180.8
190.6
192.7

35.6
26.4
26.8

Fig. 6. (a) Comparison of creep curves, (b) creep ratetime curves and (c) creep
ratestrain curves at T = 25 C and r = 24.2 MPa of SAC(105), SAC(105)0.06Ni,
SAC(105)0.5Sb solder alloys.

(105)0.5Sb and the ne dendrite arm spacing implies that they


grew at a higher velocity. Chen and Li [20] reported that the addition of Sb into SnAgCu solder can effectively improve the properties of solders. Their results show that the Sn3.5Ag0.7Cu with
about 1.0 wt.% Sb solder could exhibit the smallest growth rate

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

and gives the most prominent effect in retarding IMC growth and
rening IMC particle size. Hence, a heterogeneous nucleation effect
for retarding the IMC growth due to Sb addition is proposed.
3.3. Melting, solidication and undercooling of SAC(105)-based solders
To identify the fundamental thermal reactions of the solder
alloys during heating and cooling, the as-cast SAC(105), SAC
(105)0.06Ni, and SAC(105)0.5Sb alloys were analyzed by DSC, as
shown in Fig. 5. The DSC results are summarized in Tables 2 and
3. During heating process, one may notice that there was an elevated melting temperature less than 1.5 C for the SAC (105) solders
doped with Ni and Sb. The melting point (226.0 C) of SAC (105)
solder is slightly increased to 226.6 and 227.4 C, respectively,
when 0.06 Ni wt.% and 0.5Sb wt.% were added to the solder. From
these DSC proles, it was not required to make any changes to the
existing solder process parameters such as the reow temperature
when applying these SAC (105) solders for SMT or BGA applications. The pasty range, which is the difference between the solidus
(TS) and liquidus (TL) temperatures, for SAC (105), SAC (105)0.06Ni
and SAC (105)0.5Sb, respectively, was 18.1, 16.6 and 18.1 C. That
means the pasty range also was slightly decreased or quite similar
to the lead-free SAC (105) solder.
In the onset-to onset method, undercooling is calculated as the
onset temperature during heating minus the onset temperature
associated with cooling. Due to undercooling, the exothermic
peaks upon cooling for all samples appeared at a lower temperature compared to their endothermal peak, but the reductions in

45

temperature were not identical as seen in Fig. 5. Remarkably, the


onset temperatures associated with cooling were 180.8, 190.6
and 192.7 C for SAC (105), SAC(105)0.06Ni and SAC(105)0.5Sb,
respectively, which leads to undercooling of 35.6, 26.4 and
26.8 C for the three solders. It is worthy of notice that the degree
of undercooling for proeutectic Sn was strongly affected by the
composition, in the decreasing order was SAC(105), SAC(105)0.5Sb
and SAC(105)0.06Ni. The high value of undercooling of SAC(105)
could result in ne structures and thus, the same alloy may behave
differently with addition of small amount of Ni and Sb compared to
SAC(105) bulk specimen. The primary phases of SAC(105) may not
have provided suitable sites for heteronucleation of Sn, and thus
increasing the value of undercooling. The addition of 0.06Ni and
0.5Sb could reduce the undercooling much enough, so that each
peak was dened clearly during the freezing sequence. This reduction can be attributed to the effect of alloying elements on the rate
of solidication. Such elements may serve as extra heterogeneous
nucleation sites and thus promote solidication process. These results are quite consistent with the explanation proposed by Chen
et al. [2] that alloying elements with SnAg or SnCu solders tend
to be associated with such small undercooling. This phenomenon
also appeared in SAC(305) and SAC(105) solders as reported by
Lin et al.[21].
3.4. Creep performance
To evaluate the creep behavior of the three solder alloys, the
creep property testing was tested at 25 C with an applied stress

Fig. 7. Relationship between r and ln (e_ ) at T = 25, T = 70 and T = 110 C for (a) SAC(105), (b) SAC(105)0.06Ni and (c) SAC(105)0.5Sb solder alloys.

46

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

of 24.2 MPa. Fig. 6a shows the typical creep curves of SAC(105),


SAC(105)0.06Ni, and SAC(105)0.5Sb solder materials. Since the
stress and temperature are constants, the variations in creep rate
e_ suggest a basic change in the internal stress of the alloy sample
during time. This implies that the hardening of the matrix was
recovered immediately and balanced at an extended deformation
rate [14]. However, additions of 0.06Ni or 0.5Sb resulted in
decreasing the creep rates and thus improve creep resistance.
To compare the creep resistance of the entire alloy samples,
their computed creep rates, e_ , were plotted under the same conditions as illustrated in Fig. 6b. It is seen that the average strain ratetime history usually exhibited classical primary, secondary and
tertiary strain behavior. Remarkably, the steady-state creep strain
rate of SAC(105), SAC(105)0.06Ni and SAC(105)0.5Sb alloys,
respectively is 5.9  105 s1, 3.5  105 s1and 1.7  105 s1.
Hence, the steady state creep strain rate was lower in the
SAC(105)0.5Sb solder alloy than in the other two solder alloys.
From this comparison, it is apparent that the SAC(105) 0.5Sb solder
is much more creep resistant (350%) than that of SAC(105)solder,
but for SAC(105)0.06 Ni solder joint the creep resistance is 210%
larger than that of SAC(105) solder. Likewise, the beginning of tertiary creep in SAC(105) Sb solder appears to start at about 3 times
longer than that of SAC(105) solder, whereas the creep life of
SAC(105) Ni is about 2 times longer than that of SAC(105) solder.
It indicated that the Sb and Ni are benecial to improve the creep
strength of the SAC(105) solder and could enhance its creep resistance due to solid solution hardening of Sb and particle hardening

of Ni3Sn4 IMC reinforced SAC(105) solder alloy. Assuming that the


size of precipitates affects the creep resistance, this result is credible when the results from Fig. 6ac are taken into account. The increased creep resistance of SAC(105)0.06Ni or SAC(105)0.5Sb
alloys is mostly attributed to the change in the size of IMCs and
morphology of SAC(105) from needle-like to ber-like IMCs or ne
eutectic regions surrounding the b-Sn primary dendrites in the
SAC(105)0.06Ni and SAC(105)0.5Sb alloys, respectively.
Similar trends are also found for the SAC solders by Li et al. [22],
who suggested that adding Sb could increase ultimate tensile
strength (UTS) than Sb-free solder alloys due to solid solution
hardening and particle hardening. They also found that the tensile
strength of the Sn3.5Ag0.7Cu solder joint reaches the highest value when 1.0 wt.% Sb is added. Besides, the present results are consistent with the testing results given by Che et al. [19] which
focused on two Ni doped solders of SAC105Ni0.02 and SAC105Ni0.05, and implying that adding Ni particles in the SAC solder
could improve the solder joint drop reliability, since Ni additive
could suppress the Cu3Sn IMC growth and improves the solder
microstructures and tensile properties.
The creep strain vs. time data was converted into a strain rate
vs. strain plot, as this provides the best representation to compare
the deformation kinetics of the three solders with each other, as
exemplied in Fig. 6c. Unlike the SAC(105) specimen which
showed smallest strain, the largest strains were observed for the
SAC(105)0.5Sb solder followed by SAC(105)0.06Ni. Furthermore,
identication of the beginning of tertiary creep is found to be near

Fig. 8. Relationship between ln (r) and ln (e_ ) at T = 25, T = 70 and T = 110 C for (a) SAC(105), (b) SAC(105)0.06Ni and (c) SAC(105)0.5Sb solder alloys.

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

47

Fig. 9. Relationship between ln [sinh (a r)] and ln (e_ ) for determination stress exponent (n) values at T = 25, T = 70 and T = 110 C for (a) SAC(105), (b) SAC(105)0.06Ni and (c)
SAC(105)0.5Sb solder alloys.

an average creep strain of 0.4 for SAC(105)0.5Sb but for


SAC(105)0.06Ni and SAC(105) solders, the tertiary creep appears
to start at a slightly smaller strain of 0.3 and 0.2, respectively.
3.5. Constitutive creep equation and parameters
Due to the high homologous temperature (T/Tm > 0.5) of
SAC(105) based-solders, creep is considered as the dominant
deformation-controlling mechanism. Considering that e_ takes up
the most of the creep time of nal failure, the steady state creep region is chosen to represent the whole process in most of the constitutive models. It is generally accepted that the following sine
hyperbolic equation, which combines the power law and the exponential law in low and high stress ranges, respectively, is suitable
for constitutive analysis over wide ranges of deformation temperatures and strain rates [23]:

e_ Asinharn expQ =RT

where A (s1), a (MPa1), are material constants independent of


temperature, a is stress level parameter, R is the universal gas constant, n is the stress exponent constant related to strain rate, T (K) is
thermodynamic deformation temperature; r (MPa) is the steady
state ow stress and Q (kJ/mol) is the activation energy of deformation. The logarithm was taken and Eq. (1) was rearranged:

ln e_ ln A n lnsinhar  Q =RT

The value of a represents the stress reciprocal at which the material


deformation changes from power to exponential stress dependence.
An approximate value of a was calculated by a = b/n1, where n1 and b
are the average slopes of ln e_  ln(r) and ln e_  r lines at constant
T, respectively. The plots of ln e_  ln(r) and ln e_  r lines at temperature of 25, 70 and 110 C for the three alloys are given in Figs. 7
and 8. The slope of the ln e_ against ln [sinh(ar)] plot then gave the
value of n (Fig. 9). The value of ln(A) could then be obtained from
the intercept of the ln e_ against ln [sinh(ar)] plot and the activation
energy Q can be expressed as the slope of ln e_ against 1/T (Fig. 10).
The results, given in Fig. 10 and Table 4, show that the stress exponents decrease, respectively, from 7.0 to 5.3, 8.4 to 6.4 and 9.3 to
7.2 for SAC(105), SAC(105)0.06Ni, and SAC(105)0.5Sb alloys, respectively, as the temperature increases from 25 to 110 C. A signicant
decrease in the stress exponent with increasing temperature has
been related to the instability of the microstructure, which occurs
during high temperature deformation [24]. The difference in stress
exponent values suggests that creep deformation in the solder
alloy is inuenced by the solid solution hardening effects of Sb and
formation of Ni3Sn4 IMC that could restrict the dislocation climbing
during creep. Dislocation climbing is controlled by the dislocation
pipe diffusion and the lattice self-diffusion of the corresponding
solder matrix. In addition, the activation energy of SAC (105) for

48

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049

Eqs. (3)(5) give a good t for the temperature normalized creep


strain rate versus the stress data. Hence, the steady-state creep
strain rate and some temperatures can be accurately predicted for
the three solders by Eqs. (3)(5).
However for SAC(105) based alloys, unlike the large solid solubility of Sb in b-Sn, the extremely small solid solubility of Ni in bSn limits its diffusion. Hence, a mass of defects (e.g. dislocations
and vacancies) at the interface boundary can enhance the diffusion
of Ni atoms along the interface boundary. As a result, it can be inferred that the applied external mechanical stress on the solder alloy and a large number of excess vacancies were produced at the
phase interface under local static stress, which enhanced the formation of ber-like particles [26]. Nevertheless, the precipitation-strengthening effect of the second phase is greater at the
lower temperature range (2570 C) for SAC(105) based alloys,
suggesting that the dispersion precipitation strengthening of IMCs
is more dominant in this temperature range.
Fig. 10. Temperature dependence of steady state creep rate of the three solder
alloys.

Table 4
Activation energy (Q) and stress exponent (n) values for SAC(105), SAC(105)0.06Ni
and SAC(105)0.5Sb solder alloys.
Alloy

Q (kJ/mol)

Temperature (C)

a (MPa1)

SAC(105)

40.7

25
70
110

0.03739
0.05375
0.07231

7.0
6.0
5.3

SAC(105)0.06Ni

46.0

25
70
110

0.04135
0.05543
0.07000

8.4
7.1
6.4

SAC(105)0.5Sb

54.4

25
70
110

0.43243
0.05500
0.06667

9.3
8.0
7.2

corresponding strain rate, temperature and strain level was found to


be 40.7 kJ/mol, which is slightly increased to 46.0 kJ mol1 with Ni
addition, but clearly increased to 54.4 kJ mol1with the addition of
Sb. Accordingly, the addition of Sb apparently increased the barrier
to IMCs formation and may have enhanced the solubility of Cu and
Ag in SAC(105)0.5Sb solder. Hence, it can suppress the growth of
Ag3Sn and Cu6Sn5 IMCs, yielding ne IMCs particles. From Table 4,
our results suggest that the activation energy of creep for SAC(105)
based solders is close to the activation energy reported for the grain
boundary diffusion of b-Sn (approximately 34.665 kJ/mol) and is almost half the value of that for lattice diffusion (approximately
100 kJ/mol) [24,25]. Therefore, the creep deformation mechanism
is a power law creep by pipe diffusion, which is short-circuiting diffusion. The variations in Q and n values with the other reported values can be explained by differences in testing methods,
microstructure, specimen preparation, measuring errors, and data
processing method.
By determining A, Q, n and a values for SAC(105), SAC(105)0.06Ni and SAC(105)0.5Sb alloys, the creep strain rate, respectively, can
be estimated from the constitutive equations represented by:

e_ 4:16  103 sinh0:05448r6:1 exp


e_ 1:12  103 sinh0:05559r7:3 exp
e_ 4:24  104 sinh0:18470r8:2 exp






40:7
RT


46:0
RT

54:4
RT


5

4. Conclusions
The effects of 0.06Ni and 0.5Sb additives on the microstructure
and solidication behavior of Sn1.0Ag0.5Cu alloys (SAC105) as
well as the creep properties have been investigated. Signicant
conclusions obtained in this research are as follows:
(1) The addition of Sb to SAC(105) alloy suppresses the formation of large b-Sn grains and large needle-like precipitates
and instead favors the formation of small b-Sn dendrites surrounded by ne eutectic regions of b-Sn, Ag3Sn and Cu6Sn5
IMC phases.
(2) The addition of Ni favors the formation of ber-like IMCs and
ner dot-shaped precipitates on the surface of b-Sn matrix
rather than needle-like morphology.
(3) Alloying of Ni and Sb resulted in considerably reduced the
undercooling and slightly decreased the pasty range, while
elevated the melting temperature less than 1.5 oC.
(4) With the addition of Ni and Sb into SAC105, signicant
improvement in creep resistance of (210%) and (350%) is
realized when compared with the SAC105 solder alloy The
SAC(105) 0.5Sb solder alloy had better creep property.
(5) The creep life time of SAC105 alloys was remarkably
enhanced 3 times with the 0.5Sb alloying additions.
(6) According to the obtained stress exponents and activation
energies, it is proposed that the dominant deformation
mechanism in SAC(105) solders is dislocation climb over
the whole temperature range investigated.
References
[1] Luo ZB, Zhao J, Gao YJ, Wang L. Revisiting mechanisms to inhibit Ag3Sn plates
in SnAgCu solders with 1 wt.% Zn addition. J Alloys Compd 2010;500:3945.
[2] Chen WM, Kang SK, Kao CR. Effects of Ti addition to SnAg and SnCu solders. J
Alloys Compd 2012;520:2449.
[3] Kanlayasiri K, Ariga T. Inuence of thermal aging on microhardness and
microstructure of Sn0.3Ag0.7CuxIn lead-free solders. J Alloys Compd
2010;504:159.
[4] Kanlayasiri K, Mongkolwongrojn M, Ariga T. Inuence of indium addition on
characteristics of Sn0.3Ag0.7Cu solder alloy. J Alloys Compd
2009;485:22530.
[5] Xie HX, Chawla N, Shen YL. Mechanisms of deformation in high-ductility Cecontaining SnAgCu solder alloys. Microelectron Reliab 2011;51(6):11427.
[6] Gain AK, Chan YC, Yung WKC. Microstructure, thermal analysis and hardness
of a SnAgCu1 wt% nano-TiO2 composite solder on exible ball grid array
substrates. Microelectron Reliab 2011;51:97584.
[7] Ventura T, Terzi S, Rappaz M, Dahle AK. Effects of solidication kinetics on
microstructure formation in binary SnCu solder alloys. Acta Mater
2011;59(4):16518.
[8] Gao L, Xue S, Zhang L, Sheng Z, Ji F, Dai W, et al. Effect of alloying elements on
properties and microstructures of SnAgCu solders. Microelectron Eng
2010;87:202534.

A.A. El-Daly et al. / Materials and Design 43 (2013) 4049


[9] Zhu F, Zhang H, Guan R, Liu S. Effects of temperature and strain rate on
mechanical property of Sn96.5Ag3Cu0.5. J Alloys Compd 2007;438:1005.
[10] Wang CH, Shen HT. Effects of Ni addition on the interfacial reactions between
SnCu solders and Ni substrate. Intermetallics 2010;18:61622.
[11] Mahmudi R, Mahin-Shirazi S. Effect of Sb addition on the tensile deformation
behavior of lead-free Sn3.5Ag solder alloy. Mater Des 2011;32:502732.
[12] El-Daly AA, Swilem Y, Hammad AE. Inuence of Ag and Au Additions on
Structure and Tensile of Sn-5Sb Lead free Solder Alloy. J Mater Sci Technol
2008;24(6):9215.
[13] El-Daly AA, Swilem Y, Hammad AE. Creep properties of SnSb-based lead-free
solder alloys. J Alloys Compd 2009;471:98.
[14] El-Daly AA, Mohamad AZ, Fawzy A, El-Taher AM. Creep behavior of nearperitetic Sn-5Sb solders containing small amount of Ag and Cu. Mater Sci Eng
A 2011;528:105562.
[15] El-Daly AA, Hammad AE. Enhancement of creep resistance and thermal
behavior of eutectic SnCu lead-free solder alloy by Ag and In-additions. Mater
Des 2012;40:2928.
[16] Wiese S, Roellig M, Mueller M, Wolter KJ. The effect of downscaling the
dimensions of solder interconnects on their creep properties. Microelectron
Reliab 2008;48:84350.
[17] Mookam N, Kanlayasiri K. Effect of soldering condition on formation of
intermetallic phases developed between Sn0.3Ag0.7Cu low-silver lead-free
solder and Cu substrate. J Alloys Compd 2011;509:62769.
[18] Laurila T, Vuorinen V, Paulasto-Krckel M. Impurity and alloying effects on
interfacial reaction layers in Pb-free soldering. Mater Sci Eng R 2010;68:138.

49

[19] Che FX, Zhu WH. Poh Edith SW, Zhang XW, Zhang XR. The study of mechanical
properties of SnAgCu lead-free solders with different Ag contents and Ni
doping under different strain rates and temperatures. J Alloys Compd
2010;507:21524.
[20] Chen BL, Li GY. Inuence of Sb on IMC growth in SnAgCuSb Pb-free solder
joints in reow process. Thin Solid Films 2004;462:395.
[21] Lin LW, Song JM, Lai YS, Chiu YT, Lee NC, Uan JY. Alloying modication of Sn
AgCu solders by manganese and titanium. Microelectron Reliab
2009;49:23541.
[22] Li GY, Chen BL, Shi XQ, Wong SCK, Wang Z. Effects of Sb addition on tensile
strength of Sn15Ag0.7Cu solder alloy and joint. Thin Solid Films
2006;504:4215.
[23] Zhang J, Chen B, Zhang B. Effect of initial microstructure on the hot
compression deformation behavior of a 2219 aluminum alloy. Mater Des
2012;34:1521.
[24] Geranmayeh AR, Nayyeri G, Mahmudi R. Microstructure and impression creep
behavior of lead-free Sn5Sb solder alloy containing Bi and Ag. Mater Sci Eng A
2012;547:1109.
[25] Esfandyarpour MJ, Mahmudi R. Microstructure and tensile behavior of Sn5Sb
lead-free solder alloy containing Bi and Cu. Mater Sci Eng A 2011;530:40210.
[26] Lin F, Bi W, Ju G, Wang W, Wei X. Evolution of Ag3 Sn at Sn3.0Ag0.3Cu
005Cr/Cu joint interfaces during thermal aging. J Alloy Compd
2011;509:666672.

Вам также может понравиться