You are on page 1of 10

Tensile creep characteristics of Sn3.5Ag0.

5Cu (SAC355) solder


reinforced with nano-metric ZnO particles
A. Fawzy
a
, S.A. Fayek
b
, M. Sobhy
a,n
, E. Nassr
a
, M.M. Mousa
a
, G. Saad
a
a
Physics Department, Faculty of Education, Ain Shams University, Cairo, Egypt
b
Physics Department, National Centre for Radiation Research and Techno, Nasr City, Cairo, Egypt
a r t i c l e i n f o
Article history:
Received 21 January 2014
Received in revised form
17 February 2014
Accepted 18 February 2014
Available online 24 February 2014
Keywords:
Nano-metric
Composite solder
Structural renement
Creep resistance
a b s t r a c t
Recently, nano-composite solders have been developed in the electronic packaging materials industry to
improve the mechanical response of solder joints to be used in service under different conditions. In this
study mechanical mixing has been used to disperse nano-metric ZnO particles in Sn3.5Ag0.5Cu
(SAC355) solder at 420 1C for 2 h. In comparison with SAC355 solder, addition of nano-metric ZnO
particles effectively suppressed the formation and restricted the volume fraction of the Ag
3
Sn and Cu
6
Sn
5
intermetallic compound particles, lowering grain sizes and controlled the growth of -Sn grains in the
matrix. An improvement in tensile creep resistance of the reinforced SAC355 composite is noticed. This
improvement seems to be due to its effect in structural renement and makes the composite solder to
display a large creep life time. The addition of nano-metric ZnO particles keeps the melting temperature
nearly at the SAC355 level, indicating that the composite solder is t for the existing soldering process.
& 2014 Elsevier B.V. All rights reserved.
1. Introduction
Currently, concern over the toxicity of lead in eutectic SnPb
solders has prompted the development of new tin rich lead free
solder alloys for electronic packaging [14]. This has prompted
a search for lead-free solders and more attention in the research
activities in this eld. To date, several types of binary and
ternary Sn-based lead-free solders such as SnAg, SnCu, SnAu,
SnAgCu and SnZn have been developed and applied in the
electronic packaging industry [46]. Among the various types of
lead-free solders SnAgCu solder has been proposed as one of the
most promising substitutes for conventional SnPb solder [7,8].
However, many problems with the ternary SnAgCu lead-free
solders still exist such as an excessive reaction between SnAgCu
solder and substrates, the formation of large brittle intermetallic
compounds (IMCs) and the short creep-rupture life time in service.
In addition, the formation of large primary Sn dendrite reduces the
resistance to thermalmechanical fatigue [9]. In order to further
enhance the properties of SnAgCu solder alloys, alloying elements
such as rare earth, Bi, Sb, In, Co, Ga, Ni, Ge and nano-metric
particles were selected by many of researchers as alloys addition
into these alloys [1015].
Moreover, with the advancement of micro-/nano-systems
technology through the past years, microelectronic components
have evolved to become smaller, lighter and more functional.
Therefore, conventional solder technology can no longer guarantee
the solder joint reliability of electronic components due to the
higher diffusivity and the softening nature of the solder. To solve
such problems, a potentially viable and economically affordable
innovative approach to improve the mechanical properties of a
conventional solder is adding micro/nano-metric, non-reacting,
non-coarsening oxide particles to a solder matrix so as to form a
composite solder [16,17].
In the literature, composite solders have improved the reliability
of solder joints because the reinforcement particles suppress grain
boundary sliding, grain growth, retardation of the IMC formation and
redistribution of stress uniformly [18]. In general, the reliability of the
solder joints is mainly dependent on an interfacial IMC layer, the
difference in coefcients of thermal expansion, the yield strength,
elastic modulus, shear strength, fatigue and creep behavior. Under
usual operating conditions, the temperature of solder joints in
microelectronic packages ranges between 25 1C and 100 1C, which
corresponds to a high homologous temperature (0.60.8 T
m
). In this
temperature range, creep is the most important deformation
mechanism, which must be analyzed in order to understand the
reliability characteristics of solder joints [19]. However, knowledge
about the creep behavior of solders is very important for a successful
electronic product development. Tai et al. [20,21] reported adding Ag
nano-particles into SnCu solder affected creep life time signicantly.
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/msea
Materials Science & Engineering A
http://dx.doi.org/10.1016/j.msea.2014.02.061
0921-5093 & 2014 Elsevier B.V. All rights reserved.
n
Corresponding author. Tel.: 20 1200097710; fax: 20 22581243.
E-mail address: miladsobhym@yahoo.com (M. Sobhy).
Materials Science & Engineering A 603 (2014) 110
Their results showed that at different temperatures and stresses, Ag
nano-particles reinforced SnCu composite solder joints have higher
creep rupture life and creep activation energy than Sn0.7Cu solder
joints. Shi et al. [22] produced a series of composite solders by
blending Sn37Pb and Sn0.7Cu solder micro-sized powders with
different volume percentages of nano-sized Cu, Ag, Al
2
O
3
and TiO
2
reinforcement particles. According to their results, the creep resis-
tance of each composite solder was improved. Moreover, Tsao and
Chang [23] developed a series of Sn3.5Ag0.25Cu composite solders
reinforced with different weight percentages (0, 0.25, 0.5 and 1 wt%)
of TiO
2
nanoparticles. The Sn3.5Ag0.25Cu composite solders con-
taining 1 wt%TiO
2
nanoparticles indicated signicant improvements
in the mechanical strength.
In view of the above, the main characteristics of a Pb-free
solder is obtaining an acceptable level of mechanical and electrical
performance, reasonable melting temperature, be corrosion resis-
tant, harmless to health and the environment, and cost effective.
Literature survey revealed that no studies have been reported so
far on tensile creep behavior of the lead-free SnAgCu (SAC)
solder containing nano-metric ZnO particles. So, the objectives of
this work are to investigate the effect of 0.5 wt% of nano-sized ZnO
particles addition on the melting behavior, microstructure, ther-
mal and tensile creep properties of Sn3.5 wt% Ag0.5 wt% Cu
(SAC355) solder at different test conditions (testing temperature
and applied stresses).
2. Experimental procedures
As developed in our previous work [15], two lead-free solders
have the composition Sn3.5 wt% Ag0.5 wt% Cu (SAC355 plain
solder) and Sn3.5 wt% Ag0.5 wt% Cu0.5 wt% ZnO nanoparticles
(SAC355 composite solder) are designed to understand the effect
of nano-metric ZnO particles on the thermal and microstructure of
the plain Sn3.5Ag0.5Cu solder. These samples were prepared by
the same method used in the previous work [15]. Rod-like samples
were obtained with a diameter of approximately12 mm. These
rod-like samples were mechanically machined (cold drowning)
into wires of 0.8 mm in diameter and working length of 50 mm for
tensile creep tests and sheets of 0.4 mm thickness for structure
investigations. The small sized solder tensile test specimens used
in this work were more advantageous than the test standards
since the tensile 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 [24,25]. The microstructure
was carried out by using scanning electron microscopy (SEM)
JSM-5410, Japan. A solution of 80% glycerin, 10% nitric and 10%
acetic acid was prepared and used to etch the surface of 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 K

radiation and individual phases and


their crystal structures were identied by matching the character-
istic XRD peaks recorded from the bulk specimens against JCPDS
data. Differential scanning calorimetry (DSC) (shimadzu DSC-50)
was carried out to understand the melting process of the two,
plain and composite, solder alloys. Heating the specimens in DSC
was carried out at 10 K/min with heating rate.
Prior to the tensile creep testing, all specimens were heat-
treated at a temperature of 393 K for 2.5 h and then slowly cooled
to room temperature to stabilize the microstructure and to remove
the residual defects produced during the drawing process. In order
to understand the effect of adding nano-metric ZnO particles on
the tensile creep characteristics of SAC355 composite, tensile creep
tests were carried out at different testing temperatures ranging
from 323 to 398 K under the effect of different applied stresses
ranging from 7.5 to 13 MPa using a computerized tensile creep
testing machine described elsewhere [15]. Each datum represents
an average of three measurements. The environment chamber
temperature could be monitored using a thermocouple contacting
with the test sample.
3. Results and discussion
3.1. Melting characteristics
Fig. 1a shows the DSC curve of the SAC355 plain solder while
Fig. 1b shows the DSC curve of the SAC355 composite solder
specimen doped with 0.5 wt% nano-sized ZnO particles. For each
solder, only one peak is observed and the melting temperature of
the SAC355 composite solder was found to be 494.18 K while that
of the composite was found to be 495.26 K with a shift of about
1.1 K higher than that of the SAC355 plain solder. This observation
was found similar to those obtained in other studies on SAC
composite solders [17,23,26,27]. The slight increase in melting
point of the SAC355 composite solder can be attributed to the
effect of the nano-metric ZnO particles on the rate of solidication.
Such particles may serve as retardation sites for the solidication
process of the IMCs. The reinforcing nano-sized ZnO particles may
also change the surface instability and physical properties of the
grain boundary interfacial characteristics. This result was in good
agreement with those reported for Sn3.5 wt% Ag0.25 wt% Cu
unreinforced with nano-sized TiO
2
[26] and SAC355 with nano-
sized Al
2
O
3
particles [23]. X-ray diffraction (XRD), analysis is
performed to emphasize phase composition of the nano-metric
ZnO particles as shown in Fig. 2a. Fig. 2b represents a TEM image
for the nano-metric ZnO particles used in this study. It showed an
average size of nominally polyhedrons nano-sized ZnO particles
with 66 nm diameter.
3.2. Microstructure analysis
X-ray diffraction investigation of SAC355 plain and SAC355
composite solder as illustrated by the diffraction patterns shown
in Fig. 3a,b showed three types of phases: -Sn, Ag
3
Sn and Cu
6
Sn
5
phases. The diffraction patterns of both solders are found to have
nearly the same features. Besides, SEM images for the microstruc-
ture of SAC355 solder and SAC355 composite solder are shown in
Figs. 4 and 5. Fig. 4b shows a signicant decrease in the grain sizes
-60
-40
-20
0
20
160 180 200 220 240 260 280 300
-60
-40
-20
0
20
SAC
222.26C (495.26 K)
221.18C (494.18 K)
H
e
a
t

F
l
o
w

(
m
W
/
m
g
)
Temperature (C)
SAC+ZnO
Fig. 1. DSC curves of (a) SAC355 and (b) SAC composite solder alloys.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 2
of the -Sn matrix of the SAC355 composite solder compared with
those in the SAC plain solder alloy shown in Fig. 4(a and b). The
SAC355 composite solder has an average -Sn grain diameter of
33 m which is 23% smaller than the grain size of SAC355 plain
solder (43 m). The SEM observations (Fig. 5a and b) indicated that
the solidication process exhibited dendritic dark regions and
interdendritic bright regions. The SEM micrographs shown in
Fig. 5 present the conguration of the Ag
3
Sn and Cu
6
Sn
5
IMC
phases dispersed within the -Sn rich matrix consisting of needle-
like ne particles besides irregular polygon shapes, respectively.
EDS analysis conrmed that the dark dendritic arm is the -Sn
phase while the bright interdendritic regions are found to contain
Cu, Sn and Ag. Since the solders used were SAC355, the eutectic
mixture contains the needle-like ne Ag
3
Sn particles dispersed
within the Sn-rich matrix besides irregular polygons of Cu
6
Sn
5
IMCs. The presence of these IMCs was also conrmed by the XRD
patterns shown in Fig. 3. Addition of 0.5 wt% nano-metric ZnO
affects the microstructure of SAC355 plain solder. It was found to
control the melt crystallization of SAC355 plain solder. Fig. 5b
reveals the microstructure of the SAC355 composite solder.
It presents the same features in the microstructure of the SAC355
plain solder. It is clear from this gure that -Sn interdendritic
arms (eutectic mixture) as well as the IMC particles in the
composite solder are reduced in size, i.e. seem to be small and
ne compared with those in SAC355 plain solder alloy. According
to the EDS analysis, the eutectic areas were found to contain Zn, O,
Cu, Sn and Ag. Thus, it can be concluded that the network eutectic
areas are Cu
6
Sn
5
and Ag
3
Sn besides the ZnO particles. In order to
evaluate the evidence for the uniform distribution of ZnO nano-
particles inside the solder matrix, an over-etched SAC355 compo-
site solder was examined by SEM and EDS analysis, as shown in
Fig. 5c. It is clear that the ZnO nano-particles are dispersed
uniformly on the surface of the solder matrix as a ne dot-
shaped submicron precipitate. Besides, EDS analysis for the over-
etched specimen exhibited the same features shown in Fig. 4b. The
retardation effect of the ZnO nano-sized particles is similar to that
reported in other studies [2630]. This means that addition of
0.5 wt% nano-metric ZnO to SAC355 suppresses and controls the
-Sn grains, Ag
3
Sn needle-like and the Cu
6
Sn
5
polygon particles
yielding a uniform dispersion of these IMCs within the Sn-rich
mixture producing a ne network like microstructure with the
-Sn (Fig. 5b), which subsequently may affect its physical and
mechanical properties. They are important to understand the
tensile creep properties of SAC355 plain and composite solders,
i.e., the change in creep resistance usually attributed only to the
change of the microstructure.
3.3. Tensile creep response
It is well known that the microstructural characteristics of an
alloy determine its mechanical response. To evaluate the tensile
creep response of the SAC355 plain and SAC composite solder,
tensile creep tests were carried out at the temperatures 323, 348,
373 and 398 K under the effect of different applied stresses
ranging from 7.2 to 13 MPa. The presence of nano-metric ZnO
particles, testing temperature and the applied stress were found to
strongly affect the tensile creep behavior: Fig. 6a shows typical
representative creep curves of the SAC355 plain and SAC355
composite solder samples tested at 398 K under the effect of
11.5 MPa. Since the stress and temperature are constants, the
variations in the creep rate suggest a basic change in the internal
stress of the solder sample during time for both solders. As
indicated in Fig. 6a, it could be noticed that addition of 0.5 wt%
nano-metric ZnO particles resulted in increasing creep resistance
and life time of the composite solder depending on the applied
stress and testing temperature. Another set of representative creep
curves of both samples are stretched with 13 MPa at the tempera-
tures 323, 348, 373 and 398 K, showing that the creep curves are
strongly affected by the testing temperature as shown in Fig. 6b.
Creep behavior of both solders was found also to be strongly
affected by the applied stress as has been observed in Fig. 6c.










(
2
0
0
)










(
1
1
0
)










(
2
0
1
)










(
2
0
2
)










(
1
0
4
)

I
n
t
e
n
s
i
t
y

(
a
r
b
i
t
r
a
r
y

u
n
i
t
s
)
2 Theta










(
1
0
0
)











(
1
0
2
)

(
1
0
1
)










(
0
0
2
)











(
1
0
3
)

nano-sized
ZnO
Fig. 2. (a) XRD proles and (b) bright eld TEM of ZnO nano-particles.
2000
1000
0
3000
2000
1000
0
3000
30 40 50 60 70 80
30 40 50 60 70 80
2 Theta
I
n
t
e
n
s
i
t
y

(
a
r
b
.
u
n
i
t
s
)
Cu
6
Sn
5
Fig. 3. XRD proles of (a) SAC355 solder and (b) SAC355 composite solder alloys.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 3
This gure shows the representative creep curves for SAC355 and
SAC355 composite samples tested at 373 K under the effect of 7.2,
8.6, 10.1, 11.5 and 13.0 MPa. These isothermal creep curves of both
solders showed a monotonic shift towards higher strains and
lower fracture times on increasing the deformation temperature
and/or applied stress. Moreover, the level of creep strain for
SAC355 composite is generally lower than that of SAC355 plain
solder under the same testing conditions (Fig. 6a).
From this comparison, it is apparent that the SAC355 composite is
more creep resistant than that of SAC355 plain solder. Likewise, the
beginning of tertiary creep in SAC355 composite solder appears to
start at longer times than that of SAC355 plain solder which means
that the creep life time of SAC355 composite is longer than that of
SAC355 plain solder. Thus the composite solder has benet to enhance
its creep resistance in good agreement with the prediction of the
classical theory of dispersion strengthening created by the addition of
ZnO nanoparticles. This result is credible when the increased creep
resistance of SAC355 composite solder is mostly attributed to the
decrease of its grain sizes (Fig. 4) and renement of the IMC particles
Ag
3
Sn and Cu
6
Sn
5
(Fig. 5). Similar trends are also found in SAC solders
in other studies [2630]. They suggested that adding nano-metric
particles increased the strength of Sn-rich composite solders and
attributed this behavior to the dispersion strengthening of the ne
microstructure.
In the early stage of the time dependent deformation behavior,
transient creep strain,
tr
was found to obey a relation of the form
[31]

tr
t
n
1
where t is the transient creep time, and , n are constants
depending on the test conditions. From the relation between ln
tr
and ln t given in Fig. 7, as representative examples for both solders
at 373 K under different stresses ranged from 7.2 to 13 MPa, the
parameter n is obtained as the slope of the straight lines and is
deduced from the intercept with the ln
tr
axis. The value of n is
found independent of s but it varies with the testing temperature
as can be seen from Fig. 8. Fig. 8 shows a linear increase of average
value of n with increasing testing temperature for both solders.
At all testing temperatures n values for the composite solder were
found higher than those of the plain solder The increase of n with
increasing testing temperature can be interpreted in terms of the
fact that the transient creep time exponent n determines the
dependence of the density of mobile dislocations, , on the mean
internal stress, s
i
, and the relaxed shear modulus G according to
Fig. 4. SEM micrographs showing the grains of (a) SAC355 solder and (b) SAC355 composite solder alloys.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 4
the following relation [31,32]:
s
i
=nGb
2
2
where b is Burger's vector of the dislocations involved. Assuming that
the internal stress s
i
does not change at the same test conditions,
increasing the deformation temperature T leads to decreasing the
dislocation density due to their annihilation [31,32]. Consequently
the transient creep parameter n will increase with increasing testing
temperature as has been observed in Fig. 8. In addition, increasing
the testing temperature T resulted in: (i) increasing thermal agitation
Fig. 5. SEM micrographs showing the IMCs and the corresponding EDS curves in (a) SAC355 solder and (b) SAC355 composite solder alloys, and (c) over-etched SAC355
composite solder alloys.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 5
for the moving dislocations leading to faster annihilation and/or (ii)
decreasing the pinning effect of the pinning centers in each solder.
Hence the increase of n with T can be accounted for. Higher values of
n for the SAC355 composite solder compared with those in the
sAC355 plain solder can be rendered to the decrease of the grain
sizes in the composite solder compared with those in the SAC355
plain solder pointing to the integrated role of both grain size (Fig. 4)
and testing temperature in decreasing the dislocation density [33].
In addition the ner microstructure of the IMCs (Ag
3
Sn and Cu
6
Sn
5
)
is expected to be more effective in the SAC355 composite solder
more than that in the SAC355 plain solder; consequently their
pinning action will be higher than those in plain solder.
The minimum (steady state) creep rate
st
is one of the most
important creep parameters for fundamental and engineering
studies. Its stress dependence is often described by the power-
law equation: [33]
Cs
m
3
where C is a constant and m is the stress sensitivity parameter, and
both were found to depend on the experimental test conditions [33].
0 1000 2000 3000 4000 5000 6000
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
= 11.5 MPa
S
A
C
+
Z
n
O
S
A
C
T
testing
= 398 K
S
t
r
a
i
n
Time (Sec)
0 1000 2000 3000 4000 5000 6000
0.00
0.05
0.10
0.15
0.20
(MPa) :
1 : 7.2
2 : 8.6
3 : 10.1
4 : 11.5
5 : 13
S
t
r
a
i
n
Time (Sec)
SAC
T
testing
= 373 K
4
2
3
5
1
0 1000 2000 3000 4000 5000 6000
0.00
0.05
0.10
0.15
0.20
1 : 7.2
2 : 8.6
3 : 10.1
4 : 11.5
5 : 13
S
t
r
a
i
n
Time (Sec)
SAC+ZnO T
testing
= 373 K
4
2
3
5
1
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
S
t
r
a
i
n
Time (Sec)
SAC = 13 MPa
T
teasting
(K):
1 : 323
2 : 348
3 : 373
4: 398
4
2)
3
1
0 1000 2000 3000 4000 5000 6000
0 1000 2000 3000 4000 5000 6000
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
T
teasting
(K):
1 : 323
2 : 348
3 : 373
4: 398
4
= 13 MPa
S
t
r
a
i
n
Time (Sec)
SAC+ZnO
2
3
1
(MPa) :
Fig. 6. Representative straintime curves showing the effect of: (a) nano-metric ZnO particles addition to the SAC355 plain solder, (b) stresses at a constant testing
temperature, as indicated, and (c) testing temperature at constant applied stress, as indicated, for both SAC355 and SAC355 composite solders.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 6
The steady state creep rate
st
of the tested samples is calculated
from the slopes of the linear parts of the creep curves obtained at the
different test conditions. It increased on increasing both the testing
temperature and applied stress for both solders as observed from
Fig. 9. It is seen that under the same test conditions SAC355
composite yields lower creep rates compared with those of the
SAC355 plain solder. This can be rendered to change in the micro-
structure of the SAC355 plain solder into a ner microstructure in the
SAC355 composite solder.
3.3.1. Stress sensitivity parameter (m)
The relation between ln
st
and ln s was plotted and straight
lines for both solders were obtained (Fig. 10). Slopes of these
straight lines yielded values of the stress sensitivity parameter m
at different deformation temperatures. The temperature depen-
dence of m values given in Fig. 11 showed surprising increase with
increasing testing temperature. Moreover, values of m were found
higher in the SAC355 composite than in the SAC355 plain solder.
This means that the composite solder has longer creep life time
while retaining higher creep resistance, indicating that the com-
posite solder is suitable for solder fabrication.
The increased minimum creep rate
st
by increasing both
testing temperature and/or the applied stress (Figs. 9 and 10)
may be rendered to the formation of new dislocation sources in
the early stage of creep and the role of the applied stress which
increases the driving force for the mobile dislocations and
increases their density, which in turn contribute to the creep rate
(Figs. 9 and 10). This can be observed from increasing of the stress
exponent m with increasing temperature (Fig. 11).
It is generally known that the value of m in creep deformation,
upon the application of stress at different testing temperatures,
the obtained steady-state creep rate
st
seems to depend mainly on
the structure of the tested sample and the testing temperature. On
increasing the applied stress, the interaction between dislocations
and the other phases may lead to climb, glide and/or viscous
motion of the mobile dislocations on their slip planes. The stress
sensitivity parameter m helps in judging the controlling mechan-
ism. In the present study, m was increased with increasing the
testing temperature, from 2 to 4.4 in the composite and from 1.5 to
4 in the plain solder (Fig. 11), indicating that the rate controlling
mechanism may be glide and/or climb of dislocations along the
grain boundaries [31,33,34]. From Fig. 11, the obtained higher
values of m for the composite compared with those in the plain
solder is rendered to the renement of the microstructure and the
presence of the nano-metric ZnO particles in the -Sn matrix.
The value of the activation volume, V ln
st
/s, as derived
from the slopes of the straight lines of Fig. 12a relating ln
st
and
the applied stress s was found to increase with increasing
deformation temperature, as shown in Fig. 12b. The increase in
the activation volume by increasing temperature, Fig. 12b, can be
explained on the basis of the dislocation density because the
activation volume is directly proportional to the average spacing of
dislocations [33,35]. As thermal energy is provided, the dislocation
density decreases and the average distance between dislocations
will increase. Fig. 6a shows that the strain in plain solder samples
is higher than that for the composite solder sample under the
same test conditions. This indicates that plain solder is expected to
be softer than the composite one. However, this behavior is not
due to the increase in the average distance between dislocations
but due to the low dislocation mobility in composite solder as a
3.5 4.0 4.5 5.0 5.5 6.0 6.5
-8.5
-8.0
-7.5
-7.0
-6.5
-6.0
-5.5
-5.0
-4.5
-4.0
-3.5
-3.0
-2.5
T
testing
= 373 K T
testing
= 373 K
(Mp):
13
11.5
10.1
8.6
7.2
SAC
L
n

t
r
Ln t
3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5
-8.5
-8.0
-7.5
-7.0
-6.5
-6.0
-5.5
-5.0
-4.5
-4.0
-3.5
-3.0
-2.5
(Mp):
13
11.5
10.1
8.6
7.2
SAC+ZnO
L
n

t
r
Ln t
Fig. 7. Relation between the ln
tr
and ln t for both SAC355 and SAC355 composite solder alloys at different applied stresses.
320 340 360 380 400
0.5
0.6
0.7
0.8
0.9
n

a
v
e
r
a
g
e
T (K)
SAC
SAC-0.5 ZnO
Fig. 8. Temperature dependence of the creep parameter n under different stresses.
320 330 340 350 360 370 380 390 400
0.00000
0.00001
0.00002
0.00003
0.00004
0.00005
0.00006
0.00007
(MPa) :
1 : 7.2
2 : 8.6
3 : 10.1
4 : 11.5
5 : 13

.
s
t
T (K)
Solid lines :SAC
Dash lines :SAC+ZnO
(4)
(2)
(3)
(5)
(1)
Fig. 9. Temperature dependence of steady state creep rate
st
for both SAC355
(solid lines) and SAC355 composite solder alloys (dashed lines).
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 7
result of the high pinning action of its ne IMCs (Ag
3
Sn and
Cu
6
Sn
5
) besides the existence of the nano-metric Zno particles.
This is shown in Fig. 12b, which shows that the level of activation
volume values is slightly higher for the composite solder, even
with a higher pinning action of the ne microstructure.
As creep is a continuous thermally activated process, a para-
meter was reported to correlate transient creep with the steady
state creep rate
st
through the relation [31,33]:

o

st

4
where
o
is a constant, and the exponent ( ln / ln
st
) relates
both the transient and the steady creep stages. is the slope of the
straight lines relating ln and ln
st
given in Fig. 13 which shows
that is a stress independent ratio for both solders. For plain
SAC355 solder, assumed the value 0.8, and for SAC355 composite
solder it was 0.85. This behavior satises the concepts that creep
in both solders is a continuous process and the value of shows to
what extent the transient stage characteristics extend to the
steady state creep. The density of dislocations at the end of the
so-called sub-steady state creep in the middle of the primary creep
state [33,36] increases within grains and near grain boundaries.
These dislocations get into uniform distribution where the steady
state creep stage starts.
3.4. Role of nano-metric ZnO particles in renement mechanism
It is generally known that improvement of solder reliability in
electronic components is related to its microstructure stability. So,
the effect of microstructure, which may vary during the life of the
component, must be predictable. In short, the renement mechan-
ism of grains and suppression of the growth of IMCs in lead-free
solders during solidication and aging can be achieved by additions
of nano-metric, non-reacting, non-coarsening oxide particles [27,37].
In order to precisely evaluate the inuence of nano-metric ZnO
particles on the microstructure and tensile creep properties of the
solidied SAC355 solder we will refer to the SEM micrographs
shown in Figs. 4 and 5. It has been found that addition of a small
percentage of nano-metric ZnO resulted in renement of the
microstructure of the near eutectic SAC355 composite solder.
It appeared as dot-shaped rened Ag
3
Sn and polygons of Cu
6
Sn
5
precipitates around the eutectic as has been detected by EDS
investigations. This renement may be attributed to the adsorp-
tion of the nano-metric ZnO particles on these solidied grain
surfaces through the matrix during the solidication process.
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 2.7
-14
-13
-12
-11
-10
-9
T
testing
(K):
323
348
373
398
L
n

.
s
t
Ln
SAC
-14
-13
-12
-11
-10
-9
L
n

.
s
t
Ln
T
testing
(K):
323
348
373
398
SAC+ZnO
Fig. 10. Relation between the ln
st
and ln s at different testing temperatures for both SAC355 and SAC355 composite solder alloys.
320 330 340 350 360 370 380 390 400
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
5.5
m
T (K)
SAC
SAC+ZnO
Fig. 11. The dependence of the stress sensitivity parameter, m, on the testing
temperature, T, for both SAC355 and SAC355 composite solder alloys.
-14
-13
-12
-11
-10
-9
(MPa)
T
testing
(K):
323
348
373
398
L
n

s
t
SAC
6 8 10 12 14
320 330 340 350 360 370 380 390 400
0.15
0.20
0.25
0.30
0.35
0.40
0.45
S
A
C
+
Z
n
O
V

x
1
0
6
T (K)
S
A
C
Fig. 12. (a) Relation between the ln
st
and s at different testing temperature, T, for
SAC355 solder alloy. (b) The dependence of the activation volume, V, on the testing
temperature for both SAC355 and SAC355 composite solder alloys.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 8
The existence of such adsorbed nano-metric particles decreases
the surface energy of the IMC and subsequently decreases the
growth velocity of these IMC particles [37,38]. Moreover, Liu et al.
[39] have found that the adsorption of nano-metric Ag
3
Sn (IMC)
particles occurs during the solidication of SAC. This adsorption
decreased the surface energy of the other IMC (Cu
6
Sn
5
) and
retarded the growth of the whole IMC padticles.
For SAC355 composite solder, sizes of the IMC particles Ag
3
Sn and
Cu
6
Sn
5
ranged from 1 to 2 m and from 2 to 5 m, respectively,
as evaluated from electron micrographs. These mean values are
much larger than those of the nano-metric ZnO particles
(0.066 m). Experimental results here indicated that the existence
of such nano-metric particles renes the IMCs. The proposed
mechanism for the effect of nano-metric ZnO particles on renement
of the Ag
3
Sn IMC can be summarized as follows: during the
solidication process, nano-metric ZnO particles, which are mechani-
cally mixed in the molten SAC355 composite solder, cling to the
micrometric-metric (larger-sized) Ag3Sn particles just as spheres
cling to a plane. This can be simplied by treating the Ag
3
Sn crystal
surface as a plane and the nano-metric ZnO particles as spheres.
Accordingly adsorption of such nano-metric surface active material
can decrease the surface energy of the Ag
3
Sn crystal [26,39].
According to adsorption mechanism, Tsao et al. [26,27] reported that
the largest number of nano-Ag
3
Sn particles form in the eutectic area
and at the same time -Sn grains also become ner. Adsorption of
these nano-metric particles play an important role during the
solidication process of solder alloys and will greatly affect the
microstructure. First, when Ag
3
Sn grains had a relatively large surface
tension and grew rapidly during solidication, their surfaces had a
maximum adsorption of the nano-metric particles, and thus the
surface energy of Ag
3
Sn grains decreased. Therefore, the growth rate
of Ag
3
Sn grains was suppressed, resulting in a rened Ag
3
Sn phase.
Second, this interpretation explains that the surface energy of the
Cu
6
Sn
5
grains decreased and suppressed the growth of the whole of
the scallop-type Cu
6
Sn
5
particles. This entire interface would capture
both the Ag
3
Sn submicro-particles and the nano-metric particles
preferentially.
From this standpoint, the obtained microstructure of the
SAC355 composite solder can reect itself on the tensile creep
properties and improve the tensile creep resistance of the compo-
site solder. This is because of the: (i) pinning action of these nano-
metric ZnO particles which subsequently retards sliding of the
grain boundaries as can be expected from the results obtained in
Figs. 4 and 14 and (ii) the dispersion strengthening mechanism of
the matrix by the nely dispersed IMCs (Ag
3
Sn and Cu
6
Sn
5
) and
the nano-metric ZnO particles. These results were conrmed in a
previous study of the authors [15] as well as in other studies
[23,28,40].
Until now, it has been quite difcult to calculate the surface
energy in order to minimize the free energy of the Ag
3
Sn particles
entirely in SACx% nano-metric oxide composite solder. This is
consistent with the experimental results that the sizes of both
Ag
3
Sn and Cu
6
Sn
5
IMC particles decreased in the Sn3.5AgXCu
composite solder [27]. The morphology of Cu
6
Sn
5
grains affects the
adsorbed nano-Ag
3
Sn particles. Especially, this is the case in the
scallop-type Cu
6
Sn
5
grains that are formed by the ripening
reaction, which are likely to be captured by the large amount
of nano-Ag
3
Sn particles. These nano-metric particles decrease the
surface energy and hinder the growth of the Cu
6
Sn
5
IMC particles.
It is also expected that renement of -Sn grains can take place
during the solidication process. This is because from the results
obtained in Figs. 4, 5 and 14, it is interesting that a large number of
dot-shaped rened Ag
3
Sn and polygons of Cu
6
Sn
5
precipitates are
observed around the eutectic areas and seem to be captured by
these grain boundaries. This will decrease -Sn grain surface
energy and retard the growth of the whole grain areas of the -
Sn matrix. These ne dispersed IMC particles may pin grain
boundaries and help in controlling grain sizes and contribute very
signicantly to the material strength. In addition, grain size
strengthening seems to be a small component of the material
strength. It seems high with the intermetallic compounds in the
matrix strengthening. In other words, grain boundaries greatly
affect the creep resistance of materials. During creep test, grain
boundaries may cause strengthening or weakening depending on
the deformation temperature T and the applied stress s [41].
From the above analysis, it is not surprising that addition of
nano-metric ZnO particles to the SAC355 plain solder makes
composite solder joints more efcient in reducing the growth of
the overall IMC particles at the temperature range used in
this study.
3.4.1. Activation energy
In general, the mechanical response and performance of
materials change with increasing temperature. Some properties
and performance, such as strength and creep resistance decrease
with increasing temperature. We use an Arrhenius type equation
of the form [31]

st
Const: expE
st
=kT 6
where k and T are Boltzmann's constant and absolute temperature,
respectively. The relation between ln
st
and 1000/T as illustrated in
Fig. 15 gave straight lines for the different applied stresses. It shows
that the calculated energy needed for the operating mechanism in
the steady-state creep is 0.3 and 0.35 eV for the SAC355 plain and
-13.5 -13.0 -12.5 -12.0 -11.5 -11.0
-8.0
-7.5
-7.0
-6.5
-6.0
-5.5
-5.0
= 0.8
Ln
st
S
A
C
+
Z
n
O
L
n

K
S
A
C
= 0.85
Fig. 13. Relation between ln K and the ln
st
for both SAC355 and SAC355 composite
solder alloys.
Fig. 14. Dot-shaped rened Ag
3
Sn and Cu
6
Sn
5
precipitates around the eutectic
areas (magnied part from Fig. 5b).
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 9
composite solders, respectively. These values suggest that the creep
process in this stage is controlled by grain boundary sliding or
migration mechanisms, which agree with previous nding [33]. It
is clear that the activation energies for composite are higher than
that of the plain solder. This might be rendered to the rening effect
of nano-metric ZnO particles which is associated by suppressing the
formation of the IMCs, which forms with a large structure in the
SAC355 plain solder.
4. Conclusions
In this work, the inuence of nano-metric ZnO particles
addition on the microstructure, thermal behavior and correspond-
ing tensile creep properties of SAC355 plain solder was investi-
gated. The major ndings are:
(1) The nano-metric ZnO particles have been found to control the
microstructure of SAC355 pain solder.
(2) The tensile creep tests revealed an improvement in the creep
resistance of the SAC355 composite solder.
(3) The creep parameters n and showed marked increase with
increasing temperatures.
(4) The Orowan strengthening mechanism is the dominating
mechanism due to the presence of ZnO nanoparticles in the
solder matrix.
(5) Addition of nano-metric ZnO particles to the SAC355 plain
solder makes composite solder joints more efcient in redu-
cing the growth of the overall IMC particles.
References
[1] A.K. Gain, Y.C. Chan, W.K.C. Yung, Microelectron. Reliab. 51 (2011) 975.
[2] A.A. ElDaly, A.E. Hammad, A. Fawzy, D.A. Nasrallh, Mater. Des. 43 (2013) 40.
[3] A.A. El-Daly, A. Fawzy, S.F Mansour, M.J. Younis, Mater. Sci. Eng. A 578 (2013) 62.
[4] Jin Yu, D.K. Joo, S.W. Shin, Acta Mater. 50 (2002) 4315.
[5] A.A. El-Daly, A.E. Hammad, Mater. Des. 40 (2012) 292.
[6] A.A. El-Daly, A.E. Hammad, Mater. Sci. Eng. A 527 (2010) 5212.
[7] F.X. Che, W.H. Zhu, Edith S.W. Poh, X.W. Zhang, X.R. Zhang, J. Alloys Compd.
507 (2010) 215.
[8] S. Wiese, K.J. Wolter, Microelectron. Reliab. 47 (2007) 223.
[9] A.R. Geranmayeh, R. Mahmudi, J. Mater. Sci. 40 (2005) 3361.
[10] R. Mahmudi, S. Mahin-Shirazi, Mater. Des. 32 (2011) 5027.
[11] A.A. El-Daly, Y. Swilem, A.E. Hammad, J. Mater. Sci. Technol. 24 (2008) 921.
[12] A.A. El-Daly, Y. Swilem, A.E. Hammad, J. Alloys Compd. 471 (2009) 98.
[13] A.A. El-Daly, A.M. El-Taher, Mater. Des. 51 (2013) 789.
[14] L. Gao, S. Xue, L. Zhang, Z. Sheng, F. Ji, W. Dai, S. Yu, G. Zeng, Microelectron.
Eng. 87 (2010) 2025.
[15] A. Fawzy, S.A. Fayek, M. Sobhy, E. Nassr, M.M. Mousa, G. Saad, J. Mater. Sci.
Mater. Electron. 24 (2013) 3210.
[16] Z.D. Xia, Z.G. Chen, Y.W. Shi, N. Mu, N. Sun, J. Electron. Mater. 31 (2002) 564.
[17] S.Y. Chang, C.C. Jain, T.H. Chuang, L.P. Feng, L.C. Tsao, Mater. Des. 32 (2011) 4720.
[18] Q. Zeng, J. Guo, X. Gu, X. Zhao, X. Liu, J. Mater. Sci. Technol. 26 (2010) 156.
[19] K.J. Puttlitz, K.A. Stalter, Handbook of Lead-Free Solder Technology for
Microelectronic Assemblies, Marcel Decker Inc., New York, 2004.
[20] F. Tai, Fu Guo, Zhi-dong Xia, Yong-ping Lei, Yao-wu Shi, Int. J. Miner. Metall.
Mater. 16 (2009) 677.
[21] F Tai, F Guo, ZD Xia, YP Lei, YF Yan, JP Liu, et al., J. Electron. Mater. 34 (2005) 1357.
[22] Y.W. Shi, J.P. Liu, Y.F. Yan, Z.D. Xia, Y.P. Lei, F. Guo, et al., J. Electron. Mater.
37 (2008) 507.
[23] L.C. Tsao, S.Y. Chang, C.I. Lee, W.H. Sun, C.H. Huang, Mater. Des. 31 (2010) 4831.
[24] A.A. El-Daly, A.E. Hammad, A. Fawzy, D.A. Nasrallh, Mater. Des. 43 (2013) 40.
[25] S. Wiese, M. Roellig, M. Mueller, K.J. Wolter, Microelectron. Reliab. 48 (2008) 843.
[26] L.C. Tsao, S.Y. Chang, Mater. Des. 31 (2010) 990.
[27] Asit Kumar Gain, Y.C. Chan, Winco K.C. Yung, Microelectron. Reliab. 51 (2011) 975.
[28] Y. Shi, J. Liu, Z. Xia, Y. Lei, F. Guo, X. Li, J. Mater. Sci. Mater. El 19 (2008) 349.
[29] P. Liu, P. Yao, J. Liu, J. Electron. Mater. 37 (2008) 874.
[30] A.K. Gain, Y.C. Chan, W.K.C. Yung, Microelectron. Reliab. 51 (2011) 975.
[31] G. Saad, A. Fawzy, E. Shawky, J. Alloys Compd. 479 (2009) 844.
[32] G.S. Al-Ganainy, A. Fawzy, F. Abd El-Salam, Physica B 344 (2004) 443.
[33] F. Abd El-Salam, A.M. Abd El-Khalek, R.H. Nada, A. Fawzy, Mate. Character.
59 (2008) 9.
[34] C.H. Reader, D. Mitlin, R.W. Messler, J. Mater. Sci. 33 (1998) 4503.
[35] Bo Wang Liu, C. Tang, W. Qiu, F. Tien, Mater. Sci. Forum 175 (1995) 443.
[36] M. Suery, B.B. Baudelet, Philos. Mag. 41 (1980) 41.
[37] L.C. Tsao, J. Alloys Compd. 509 (2011) 2326.
[38] D.Q. Yu, L. Wang, C.M.L. Wu, C.M.T. Law, J. Alloys Compd. 389 (2005) 153.
[39] T.H. Chuang, L.C. Tsao, Chien-Han Chung, S.Y. Chang, Mater. Des. 39 (2012) 475.
[40] M.J. Esfandyarpour, R. Mahmudi, Mat. Sci. Eng. A 530 (2011) 402.
[41] A. Fawzy, J. Alloys Compd. 486 (2009) 768.
2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2
-14.0
-13.5
-13.0
-12.5
-12.0
-11.5
-11.0
-10.5
-10.0
-9.5
-9.0
Q= 0.3 eV
(Mp):
13
11.5
10.1
8.6
7.2
L
n

.
s
t
1000/T
SAC
2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2
-14.0
-13.5
-13.0
-12.5
-12.0
-11.5
-11.0
-10.5
-10.0
-9.5
-9.0
(Mp):
13
11.5
10.1
8.6
7.2
L
n

.
s
t
1000/T
Q= 0.35 eV SAC+ZnO
Fig. 15. Relation between the ln
st
and 1000/T for: (a) SAC355 and (b) SAC355
composite solder alloys.
A. Fawzy et al. / Materials Science & Engineering A 603 (2014) 110 10

You might also like