You are on page 1of 18

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/265969021

Review on the effect of alloying element and


nanoparticle additions on the properties of SnAg-Cu solder alloys
Article in Soldering and Surface Mount Technology May 2014
DOI: 10.1108/SSMT-02-2014-0001

CITATIONS

READS

50

2 authors, including:
Ervina Efzan
Multimedia University
36 PUBLICATIONS 39 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Ervina Efzan


Retrieved on: 31 July 2016

Soldering & Surface Mount Technology


Review on the effect of alloying element and nanoparticle additions on the properties of Sn-Ag-Cu solder
alloys
Ervina Efzan Mhd Noor Amares Singh

Article information:

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

To cite this document:


Ervina Efzan Mhd Noor Amares Singh , (2014),"Review on the effect of alloying element and nanoparticle additions on the
properties of Sn-Ag-Cu solder alloys", Soldering & Surface Mount Technology, Vol. 26 Iss 3 pp. 147 - 161
Permanent link to this document:
http://dx.doi.org/10.1108/SSMT-02-2014-0001
Downloaded on: 23 November 2015, At: 18:31 (PT)
References: this document contains references to 77 other documents.
To copy this document: permissions@emeraldinsight.com
The fulltext of this document has been downloaded 140 times since 2014*

Users who downloaded this article also downloaded:


Ervina Efzan Mhd Noor, Amares Singh, Yap Tze Chuan, (2013),"A review: influence of nano particles reinforced on solder
alloy", Soldering & Surface Mount Technology, Vol. 25 Iss 4 pp. 229-241 http://dx.doi.org/10.1108/SSMT-11-2012-0026
Barrie D. Dunn, Grazyna Mozdzen, (2014),"Tin oxide coverage on tin whisker surfaces, measurements and implications
for electronic circuits", Soldering & Surface Mount Technology, Vol. 26 Iss 3 pp. 139-146 http://dx.doi.org/10.1108/
SSMT-12-2013-0040
Yingxin Goh, A.S.M.A. Haseeb, Mohd Faizul Mohd Sabri, (2013),"Electrodeposition of lead-free solder alloys", Soldering &
Surface Mount Technology, Vol. 25 Iss 2 pp. 76-90 http://dx.doi.org/10.1108/09540911311309031

Access to this document was granted through an Emerald subscription provided by emerald-srm:530717 []

For Authors
If you would like to write for this, or any other Emerald publication, then please use our Emerald for Authors service
information about how to choose which publication to write for and submission guidelines are available for all. Please visit
www.emeraldinsight.com/authors for more information.

About Emerald www.emeraldinsight.com


Emerald is a global publisher linking research and practice to the benefit of society. The company manages a portfolio of
more than 290 journals and over 2,350 books and book series volumes, as well as providing an extensive range of online
products and additional customer resources and services.
Emerald is both COUNTER 4 and TRANSFER compliant. The organization is a partner of the Committee on Publication Ethics
(COPE) and also works with Portico and the LOCKSS initiative for digital archive preservation.
*Related content and download information correct at time of download.

Review on the effect of alloying element and


nanoparticle additions on the properties of
Sn-Ag-Cu solder alloys
Ervina Efzan Mhd Noor and Amares Singh

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Faculty of Engineering and Technology, Multimedia University, Malacca, Malaysia


Abstract
Purpose The aim of the present study was to gather and review all the important properties of the SnAgCu (SAC) solder alloy. The SAC solder
alloy has been proposed as the alternative solder to overcome the environmental concern of lead (Pb) solder. Many researchers have studied the
SAC solder alloy and found that the properties such as melting temperature, wettability, microstructure and interfacial, together with mechanical
properties, are better for the SAC solder than the tin lead (SnPb) solders. Meanwhile, addition of various elements and nanoparticles seems to
produce enhancement on the prior bulk solder alloy as well. These benefits suggest that the SAC solder alloy could be the next alternative solder
for the electronic packaging industry. Although many studies have been conducted for this particular solder alloy, a compilation of all these
properties regarding the SAC solder alloy is still not available for a review to say.
Design/methodology/approach Soldering is identified as the metallurgical joining method in electronic packaging industry which uses filler metal, or
well known as the solder, with a melting point 425C (Yoon et al., 2009; Ervina and Marini, 2012). The SAC solder has been developed by many methods
and even alloying it with some elements to enhance its properties (Law et al., 2006; Tsao et al., 2010; Wang et al., 2002; Gain et al., 2011). The development
toward miniaturization, meanwhile, requires much smaller solder joints and fine-pitch interconnections for microelectronic packaging in electronic devices
which demand better solder joint reliability of SAC solder Although many studies have been done based on the SAC solder, a review based on the important
characteristics and the fundamental factor involving the SAC solder is still not sufficient. Henceforth, this paper resolves in stating all its important properties
based on the SAC solder including its alloying of elements and nanoparticles addition for further understanding.
Findings Various Pb-free solders have been studied and investigated to overcome the health and environmental concern of the SnPb solder. In
terms of the melting temperature, the SAC solder seems to possess a high melting temperature of 227C than the Pb solder SnPb. Here, the melting
temperature of this solder falls within the range of the average reflow temperature in the electronic packaging industry and would not really affect
the process of connection. A good amendment here is, this melting temperature can actually be reduced by adding some element such as titanium
and zinc. The addition of these elements tends to decrease the melting temperature of the SAC solder alloy to about 3C. Adding nanoparticles,
meanwhile, tend to increase the melting temperature slightly; nonetheless, this increment was not seemed to damage other devices due to the very
slight increment and no drastic changes in the solidification temperature. Henceforth, this paper reviews all the properties of the Pb-free SAC solder
system by how it is developed from overcoming environmental problem to achieving and sustaining as the viable candidate in the electronic
packaging industry. The Pb-free SAC solder can be the alternative to all drawbacks that the traditional SnPb solder possesses and also an upcoming
new invention for the future needs. Although many studies have been done in this particular solder, not much information is gathered in a review
to give better understanding for SAC solder alloy. In that, this paper reviews and gathers the importance of this SAC solder in the electronic
packaging industry and provides information for better knowledge.
Originality/value This paper resolves in stating of all its important properties based on the SAC solder including its alloying of elements and
nanoparticles addition for further understanding.
Keywords Pb-free, Solder joints, Solder, Sn-Ag-Cu, Interconnections
Paper type Literature review

and the printed circuit board (PCB) (Lee, 2002; Gao et al., 2009).
The process used to connect the chips/components with the PCB is
by soldering, which was first implemented many years ago (Zhang
et al., 2010b). Soldering is identified as the metallurgical joining
method in the electronic packaging industry which uses a filler
metal, i.e. solder, with a melting point 425C (Yoon et al., 2009;
Ervina and Marini, 2012. Basically, this solder acts as the joining
material that connects between the device and the substrate to
ensure electrical, mechanical and thermal continuity of the
connection (Abtew and Selvaduray, 2000; Lee, 2002; Shen and
Chan, 2009; Sugunama and Dekker, 2004).
Because the application of the soldering process in this
industry, the material that has commonly been used istin
lead (SnPb) solder (Whalley, 2004; Gao et al., 2010). The
SnPb solder system has been shown to possess some beneficial

1. Introduction
Electronic packaging is a current manufacturing process used to
enable any electronic functional-based application (Abtew and
Selvaduray, 2000; Gao et al., 2009). The packaging system provides
a medium for electronic interconnections between the components

The current issue and full text archive of this journal is available at
www.emeraldinsight.com/0954-0911.htm

Soldering & Surface Mount Technology


26/3 (2014) 147161
Emerald Group Publishing Limited [ISSN 0954-0911]
[DOI 10.1108/SSMT-02-2014-0001]

147

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

properties. For instance, the SnPb solder has a very low


melting temperature of 183C (Abtew and Selvaduray, 2000;
Gao et al., 2010), low cost (Law et al., 2006) and possesses
good mechanical properties such as high shear strength
(Chang et al., 2006). In addition, the SnPb solder is known to
reduce the surface tension of pure tin (Sn) and forms a
well-defined intermetallic compound (IMC) for good joint
reliability (Abtew and Selvaduray, 2000; Zhang et al., 2012a).
These advantages were the main criteria for the application of
SnPb as the solder in the electronic packaging industry.
In recent years, environmental concerns regarding the use
of Pb-containing solder in electronic products were raised as a
vital issue (Lee, 2002). Parallel to that, SnPb solders were
banned due to their hazardous effects on both humans and the
environment (Abtew and Selvaduray, 2000; Ervina and
Marini, 2012). The European Commission introduced the
RoHS Directive that restricted the use of Pb in electronics
(Roubaud et al., 2001). Owing to this, alternative Pb-free
solders have been widely researched (Zhang et al., 2010b;
El-Daly and Hammad, 2009; Zhou et al., 2005).
To succeed as a good solder, several properties should be met,
as well as having the ability to withstand stress during service and
being reliable over long-term usage (Humpston and Jacobson,
2004; Lee, 2002). Properties such as a well-defined
microstructure (El-Daly and Hammad, 2010), appropriate
melting temperature (Mayappan and Ahmad, 2010; Yoon et al.,
2009), wettability (Yu et al., 2004; Noor et al., 2010) and
mechanical properties (Law et al., 2006; Anderson et al., 2001;
Shalaby, 2010) are the main concerns of a novel solder. The
mechanical properties of an Pb-free solder must be reliable
during service lifetime to ensure the reliability and integrity of the
function for the electrical component (Lee, 2002; Chang et al.,
2011). Hence, subsequently here, a solder should not satisfy only
environmental concerns (Law et al., 2006; Shen and Chan,
2009). Related to the microstructural evolution are the IMCs
which form due to the interaction between the solder alloy and
the substrate (Roubaud et al., 2001; Gong et al., 2009; Morando
et al., 2012). The IMC is important in providing better joint
reliability during long-term usage (Gong et al., 2008; El-Daly and
Hammad, 2010). All these factors are crucial in producing a
solder alloy that can be used for electronics packaging.
Therefore, to overcome the problems with the use of Pb-based
solders and also keeping the properties in a good manner, several
Sn-based Pb-free solder systems have been used as alternative
solders (Humpston and Jacobson, 2004; Jeon et al., 2008; Lin
and Chuang, 2008; El-Daly and Hammad, 2010). According to
Ervina and Singh (2012); Snzinc (Zn) solders have been found
to produce high strength with high shear strengths of 109 N/mm2
and tensile strengths of 30.2 N/mm2. In the same research,
adding bismuth (Bi) to SnZn solder, produced a low melting
point of 198C, which is near to that of SnPb (183C)
(Mayappan and Ahmad, 2010; Duan et al., 2004). Other
research also suggested that SnZn solder was a viable candidate
with the same explanation (Zhang et al., 2010b). Nonetheless,
this solder exhibits poor wetting properties and is prone to
oxidation and corrosion (Laurila et al., 2005; Haseeb et al., 2012;
Yoon et al., 2006; Mahmudi et al., 2008; Islam et al., 2005). All
the drawbacks of the Sn-based solder caused doubts about its
long-term use and the cost of repair and maintenance due to this
deterioration (Whalley, 2004; Mayappan and Ahmad, 2010).

Hence, several studies have recommend the SAC solder


system as a good candidate, and SAC alloys are now widely
used in the electronics industry due to their benefits compared
to other solders (Zhang et al., 2012a; Morando et al., 2012;
Sundelin et al., 2008; Chuang et al., 2012). The eutectic SAC
solder can be categorized based on three different
compositions proposed by three countries which have differed
its composition: Japan (Sn3.0 weight per cent Ag0.5 weight
per cent copper [Cu]), Europe (Sn3.8 weight per cent silver
[Ag]0.7 weight per cent Cu) and the USA (Sn3.9 weight per
cent Ag0.6 weight per cent Cu) (Zhang et al., 2012a).
SAC solders have been developed by many methods such as
alloying it with additional elements to enhance its properties
(Law et al., 2006; Tsao et al., 2010; Wang et al., 2002; Gain et al.,
2011). The development toward miniaturization, meanwhile,
requires much smaller solder joints and fine-pitch
interconnections for microelectronic packaging in electronic
devices and these demand better solder joint reliability from the
SAC solders (Lee, 2002; Shen and Chan, 2009; Gao et al.,
2010). Parallel to that, to provide this particular need, smaller
particles, such as nanoparticle reinforcements, have been added
to SAC solders to provide the mechanical integrity needed for a
Pb-free solder, as well as preserving the solder as environmental
friendly (Shen and Chan, 2009; Tsao et al., 2010). In addition,
nanoparticle reinforcements have been found to enhance the
properties of SAC solders (Shen and Chan, 2009; Tsao et al.,
2010; Gain et al., 2011 Haseeb and Leng, 2011; Chang et al.,
2011; Ervina et al., 2013). The difference here is that added
elements tend to undergo a diffusion process which produces an
alloy of the SAC systems or acts as a solid solution strengthener
(Wiese and Wolter, 2004; Gao et al., 2010; Kotadia et al., 2012),
while the addition of nanoparticles does not involve a diffusion
process that produces an alloy, instead the nanoparticles act as
discrete particles in the solder system (Shen and Chan, 2009;
Ervina et al., 2013). Although many studies have been performed
on SAC solders, a review of the important characteristics and the
fundamental factors involving the SAC solder is still not
sufficient. Hence, this paper resolves the issue by stating all the
important properties of SAC solders, including alloying with
additional elements and by nanoparticle additions.

2. Significant properties of SAC solder


In this part, a review of SAC solder based on its properties will be
presented. Combinations of some main properties will be explained
and portrayed based on the SAC solder and its alloying property.
2.1 Melting temperature properties of AnAgCu
Deceptively, the main concern when it comes to electronic
packaging is the melting temperature which comprises a liquidus
and solidus temperature (Morando et al., 2012; Ervina and
Marini, 2012). Melting temperature is important for producing
better joints at a lower temperature and also to avoid creep
(Manko, 2001). The melting temperature is basically determined
by the liquidus temperature, which occurs below 225C (Ervina
and Marini, 2012; Ma and Suhling, 2009). The liquidus
temperature is the state of temperature at which the solder alloy
is in the molten form and is important to avoid any thermal
damage to the device during the soldering process (Zhou et al.,
2005). The solidus temperature is the highest temperature at
which an alloy remains solid, and it is also crucial in providing
148

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

enhanced integrity between the solder and the substrate (Mei,


1996; Kanlayasari et al., 2009; Lee, 2002). Relating to this,
studies have also shown that rapid solidification enhances the
microstructure and provides a better bond between the solder
and the substrate compared to slow solidification (Shalaby,
2010; Li and Wang, 2006; Prabhu et al., 2012; Kanlayasari et al.,
2009). It is due to the development of fine-grained transition
structures which, in return, improve the creep resistance
(Shalaby, 2010; Prabhu et al., 2012). This information primarily
shows the importance of the melting temperature.
Typically, SnPb solder melts at about 183C (Jeon et al.,
2008; Duan et al., 2004). Therefore, this has been the
benchmark for new solders, or it is in the range of about 20C
(Whalley, 2004; Ma and Suhling, 2009; Gao et al., 2009).
The melting temperature of SAC solder ranges between 217
to 221C depending on the composition of Ag and Cu (Gao
et al., 2010; Chuang et al., 2012; Chang et al., 2011; Tsao
et al., 2010). Normally, eutectic SAC solders are developed by
adding 0.7 weight per cent of Cu into the eutectic Sn-3.5Ag
solders (Zhang et al., 2012a; Kotadia et al., 2012; Yang
et al., 2010). According to Zhang et al., 2012a and Chuang
et al., 2012; eutectic SAC solder showed a single
endothermic peak, and the melting range was short, within
3.90-4.66C, indicating the alloy existed as part liquid for a
short time during solidification. This criterion was

identified to form reliable joints during the reflow process


(Zhang et al., 2012a; Chuang et al., 2012; Kotadia et al.,
2012; Chang et al., 2011).
2.1.1 Melting temperature properties of SAC added with elements
SAC solders have already given a first impression of beneficial
influence compared to the SnPb solder in terms of melting
properties. Although, to further enhance the properties of the
SAC solder, several researchers have added elements such as
titanium (Ti) (Chuang et al., 2012), Bi (Zhang et al., 2012a), Zn
(Zhang et al., 2012a), magnesium (Mg) (Chuang et al., 2012),
cerium (Ce) (Gao et al., 2010; Zhang et al., 2012a), manganese
(Mn) (Zhang et al., 2012a), lanthanum (La) (Gao et al., 2010;
Law et al., 2006) and cobalt (Co) (Anderson et al., 2001).
Elements such as Ti, Bi, Zn and Mg affect the melting
temperature of SAC solder (Zhang et al., 2012a; Chuang et al.,
2012). By adding Ti to Sn-3.5Ag-0.5Cu solder, the melting
temperature tends to decrease by about 3C, as the weight
percentage of Ti increases from 0.25 to 1 per cent (Table I)
(Chuang et al., 2012). A short melting range of 2.88C was
also observed for the 1 weight per cent of Ti compared to the
4.66C melting range of bare Sn-3.5Ag-0.5Cu (SAC) solder.
Adding Zn also showed a similar reduction of about 2C
compared to the bulk Sn-3.8Ag-0.7Cu (SAC) solder (Kotadia
et al., 2012). Comparable reductions in the melting

Table I Melting temperature of SAC, SAC-X and SAC nanoparticles


Solder (weight per cent)

Melting
temperature, TM (C)

Solidus
temperature, TS (C)

Liquidus
temperature, TL(C)

60Sn-40Pb
Sn-3.5Ag (SA)

183
223

183
221

187
221

Sn-3.8Ag-0.7Cu (SAC)

217

217

217

Sn-3.9Ag-0.1Cu (SAC)
Sn-3.0Ag-0.5Cu
Sn.3.5Ag-0.5Cu (SAC)
Sn-3.8Ag-0.7Cu (SAC)
Sn-1.1Ag-0.45Cu-0.25Mn
Sn-1.07Ag-0.58Cu-0.037Ce
Sn-3.8Ag-0.7Cu-0.5Zn
Sn-3.8Ag-0.7Cu-1Zn
Sn-3.8Ag-0.7Cu-1.5Zn
Sn-3.0Ag-0.5Cu-0.25Al2O3
Sn-3.0Ag-0.5Cu-0.5Al2O3
Sn-3.0Ag-0.5Cu-1Al2O3
Sn-3.8Ag-0.7Cu-2Bi
Sn-3.8Ag-0.7Cu-4Bi
Sn-3.0Ag-0.5Cu 0.25Ti
Sn-3.0Ag-0.5Cu-0.5Ti
Sn-3.0Ag-0.5Cu-1Ti
Sn-3.5Ag-0.7Cu-0.5TiO2
Sn-3.0Ag-0.5Cu-0.5ZrO2
Sn-3.0Ag-0.5Cu-1ZrO2
Sn-3.0Ag-0.5Cu-3ZrO2
Sn-3.0Ag-0.5Cu-0.5SrTiO3

220.40
217.64
221.58
217.73

216.23
216.65
216.86
222.3
222.7
223.0
213.08
206.48
220.95
220.86
219.47
224.1

217.7

217.31
217.64

217.39
217.65
212.06
212.61
213.32

217.7
217.08
217.12
217.25

222.71
217.64
216.92

227.63
226.14

216.73
216.75
216.59

221.63
221.65
221.95

149

References
(Abtew and Selvaduray, 2000)
(Jeon et al., 2008; Yang et al.,
2010; Yang et al., 2013)
(Nai et al., 2006; Ervina and
Marini, 2012; Jeon et al.,
2008; Yang et al., 2010; Yang
et al., 2013)
(Ervina and Tan, 2013)
(Lee, 2002)
(Chuang et al., 2012)
(Kotadia et al., 2012)
(Law et al., 2006; Lin and
Chuang, 2010)
(Tsao et al., 2010)

(Chang et al., 2011)

(Law et al., 2006)


(Chuang et al., 2012)

(Wang et al., 2002)


(Gao et al., 2009)

(Fouzder et al., 2011)

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

temperature were found by adding Ce (Gao et al., 2010; Law


et al., 2006), La (Law et al., 2006), Bi (Zhang et al., 2012a)
and Mg (Chuang et al., 2012) compared with the bulk SAC
solder systems. The possible explanation of this reduction is
probably due to the dissolution or precipitation of the
elements as a lower melting point stage than the eutectic
matrix besides the reduction in the crystalline size (Zhang
et al., 2012a; Gain et al., 2011).

(Naisbitt, 2006; Noor et al., 2010; Nai et al., 2006). Wetting


is usually measured by using two methods, one by dipping and
another by the sessile drop method (Satyanarayan and
Prabhu, 2011). The dipping method involves measuring
contact angles when a sample of solid is brought into contact
with a test liquid (Shang et al., 2008; Xu et al., 2008).
Meanwhile, the sessile drop test is performed with a
goniometer, whereby a liquid sample is placed onto a surface
with a syringe, which can then be analyzed by using a
microscope (Shang et al., 2008).
In the wetting analysis, the principle used to evaluate the
wettability is by measuring the contact angle formed at the
joint of a solid and liquid (Naisbitt, 2006; Xu et al., 2008;
Noor et al., 2010). The surface tension equation (2) is usually
used to determine the contact angle (Satyanarayan and
Prabhu, 2011; Nai et al., 2006). According to Noor et al.,
2010; Youngs equation, below, is used to determine the
contact angle (Nai et al., 2006):

2.1.2 Melting temperature properties of SAC with added


nanoparticles
Another approach in the SAC solder development is reducing
its size from bulk to nanoparticles (Shen and Chan, 2009;
Gao et al., 2010). Reducing Sn-3.0Ag-0.5Cu (SAC) to a
nanoparticle size by using a consumable-direct current arc
technique was found to decrease the melting temperature to as
low as 179C, which is almost similar to SnPb solder (Gao
et al., 2010). Decreasing the size of the particles caused the
reduction of melting temperature due to an increase in the
surface instability (Tsao et al., 2012; Gao et al., 2010). Similar
reductions in melting temperatures ranging from 3.1 to 4.2C
were found in Sn-3.0Ag-0.5Cu (SAC) nanoparticle solders
prepared by a chemical reduction method (Gao et al., 2010).
Meanwhile, adding nanoparticles into SAC solder also
affects the melting temperature of the base solder (Chang
et al., 2011; Tsao et al., 2010). In their research, Tsao et al.
(2010) found that the addition of nanoparticles of aluminium
oxide (Al2O3) to the Sn-3.0Ag-0.5Cu (SAC) solder increased
the melting temperature slightly by about 0.5C; however, this
slight increase did not appear to deleteriously affect the
performance of the solder. Similar results were found by
adding zirconia (ZrO2) (Chang et al., 2011) and titanium
dioxide (TiO2)(Gain et al., 2011) in to Sn-3.0Ag-0.5Cu
(SAC) solder separately. Likewise, Fouzder et al. (2011)
found that the addition of rhodium or strontium titanate
(SrTiO3) nanoparticles increased the melting temperature of
the Sn-3.0Ag-0.5Cu (SAC) solder from 217 to 217.7C. In
contrast to the slight increase in melting temperature, the
addition of these nanoparticles did provide some
enhancement in the strength and microstructure.
Table I shows some of the melting temperatures of SAC with
different compositions of Ag and Cu, alloying of Sn-Ag-Cu-X
(SAC-X) with some elements such as Ce (Zhang et al., 2012a),
Bi (Zhang et al., 2012a; Abtew and Selvaduray, 2000), Mn
(Zhang et al., 2012a) and Ti (Chuang et al., 2012) and also for
nanoparticle additions into SAC nanoparticles by TiO2
(Chang et al., 2011; Tsao et al., 2012), Al2O3 (Tsao et al., 2010),
ZrO2 (Gain et al., 2011) and SrTiO3 (Fouzder et al., 2011).

Fw P

(1)

where, (FW) is the withdrawal force, () is the surface tension


and (P) is the specimen perimeter (Nai et al., 2006).
sv ls lvcos

(2)

where, sv surface tension of the solid, ls interfacial


energy, lv the surface tension of the liquid and contact
angle (Abtew and Selvaduray, 2000; Noor et al., 2010).
Generally, it has been concluded that the smaller the contact
angle and the higher the wetting force (1), the better the wettability
of a solder (Nai et al., 2006; Kanlayasari et al., 2009).
Relating to the wettability properties of the SAC solder
system, Sn-3.5Ag-0.7Cu (SAC) solder alloy has a higher
wetting force than the Sn-3.5Ag (SA) solder due to its lower
melting temperature than the latter (Gao et al., 2010). Based
on the research of Tsao et al. (2010), Sn-3.5Ag-0.5Cu (SAC)
solder produces a contact angle of about 33. In other
research, the wetting force was increased by the addition of Cu
in the SA solder (Zhang et al., 2012a). This correlated with the
formation of eutectic Sn-3.5Ag-0.5Cu (SAC) solder by the
addition of Cu in the SA solder (Zhang et al., 2012a; Kotadia
et al., 2012). Meanwhile, the overall contact angle for
Sn-3.1Ag-0.9Cu (SAC) was within 20 to 30, and the SAC
solder showed better wettability at a reflow temperature of
250C by producing a larger spreading area and lower contact
angle (Ervina and Tan, 2013). Although the wettability of the
SAC was 50, and this angle was sufficient to produce better
joints (Noor et al., 2010), little research has been done based
on the wettability of the SAC solder system.

2.2 Wettability properties of SAC


Wettability is the capability of a solder to spread on the
substrate to form a metallurgical bond (Satyanarayan and
Prabhu, 2011; Nai et al., 2006). To form an appropriate
metallurgical bond, the wetting process must occur (Shen and
Chan, 2009). The ability of molten solders to flow during
soldering is important in forming a proper intermetallic bond
(Noor et al., 2010). Thus, Nai et al., 2006 stated that this
proper intermetallic joint was crucial to attain a perfect joint
(Noor et al., 2010; Nai et al., 2006).
Wetting is a measure of the ability of a material, generally a
liquid, to spread over another material, usually a solid

2.2.1 Wettability properties of SAC with added elements


The SAC solder systems wettability properties are also
influenced by the addition of some other elements (Zhang
et al., 2012a; Wang et al., 2002). Elements such as Ce
(Zhang et al., 2010b; 2012a), indium (In) (Kanlayasari
et al., 2009), La (Zhang et al., 2010b), nickel (Ni) (Yoon
et al., 2009), diamond (Zhang et al., 2012a), erbium (Er)
(Gao et al., 2010), Mg (Lu et al., 2008) and Bi (Zhang et al.,
2012a) have been added and their influence on SAC solders
was studied in terms of wetting properties. The addition of
some elements tended to create better wettability properties
150

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Figure 2 Wetting force vs temperature of SAC and SAC-Ni

compared to SAC solders (Wang et al., 2002; Peng et al.,


2007; Zhang et al., 2012a).
Adding In into Sn-0.3Ag-0.7Cu (SAC) decreased the contact
angle together with giving a shorter wetting time (Figure 1)
(Kanlayasari et al., 2009). These results were mainly due to the
decrease of surface tension by the In addition, which suggests that
this enhanced the wettability of the Sn-0.3Ag-0.7Cu (SAC) (Nai
et al., 2006) and satisfied the ideal wettability criteria mentioned
earlier (Kanlayasari et al., 2009; Satyanarayan and Prabhu, 2011).
Another reason behind this is that adding rare elements, such as In,
increases the surface instability with the higher surface-free energy
rendered by the addition and this enables the solder to spread and
wet in a short time (Tsao et al., 2012).
In the study by Yoon et al., 2009; it was found that
Sn-3.0Ag-0.5Cu (SAC) solder provided better wettability than
Sn-3.0Ag-0.5Cu (SAC-Ni) solder, where the wetting force was
higher and the wetting angle was smaller, as shown in Figures 2
and 3 (Yoon et al., 2009). This result was supported by Zhang
et al. (2012a) who stated that a high wetting force or low contact
angle, as well as a fast wetting time, represented good wetting
(Nai et al., 2006; Noor et al., 2010; Yoon et al., 2009).
Law et al., 2006 examined the wettability based on the
addition of Ce and La to Sn-3.5Ag-0.7Cu (SAC) solder. The
appropriate additions of 0.1 weight per cent Ce and La
resulted in lower wetting times and increasing wetting force.
The reason is that these elements tend to decrease the surface
tension of the solder and this, consequently, reduces the
contact angle, enhancing the spreading area (Wang et al.,
2002; Zhang et al., 2012a). Nevertheless, excessive amounts
of these elements negatively impacted the wetting properties
by lowering the wetting force and increasing the solder
viscosity (Law et al., 2006; Tay et al., 2013).
Similarly, an over addition of elements will increase the
wetting angle, adversely impacting joint reliability (Yu et al.,
2004; Nai et al., 2006). The reason is that too large an addition
tends to increase the viscosity, making it harder for the liquid
solder to spread (Tay et al., 2013). Figures 4 and 5 show the
influence of elemental Ce and La on Sn-3.5Ag-0.7Cu (SAC)
solder in terms of the wetting angle and wetting force.
Further, adding diamond enhances the wetting properties of
Sn-3.0Ag-0.5Cu (SAC) solder (Zhang et al., 2012a). One of the
reasons behind this is that diamond has a density of 3.5 g/cm3
and is lower than the density of Sn at 7.3 g/cm3, which promotes
the diamond to move upward during wetting, discharge from the
solder and reduces the surface tension (Zhang et al., 2012a). Gao
et al., 2010 indicated that the addition of 0.25 weight per cent

Figure 3 Wetting time vs temperature of SAC and SAC-Ni

Er increased the spreading area for Sn-3.8Ag-0.7Cu (SAC)


solder. There are also other studies revealing that the addition of
Bi, gallium and germanium (Ge) to SAC provided better wetting
properties due to them behaving as surface-active elements (Gao
et al., 2010; Yu et al., 2004).
However, the over addition of elements ( 3 weight per
cent) has been found to degrade the wetting properties (Yu
et al., 2004; Tsao et al., 2012). Although most of these
elements gives beneficial results in terms of wetting
properties, Mg, meanwhile, degrades the wettability
properties due to its vulnerability to oxidation, which
therefore increases the surface tension and thus inhibits
solder spreading (Lu et al., 2008). Table II shows a
summary of the influence on wetting properties of SAC
solders by the addition of elements.

Figure 1 Wetting time of SAC with In added

2.2.2 wettability properties of SAC added with nanoparticles


Alternatively, the addition of nanoparticles to SAC has also
revealed some good wetting properties (Ervina et al., 2013). In
their research, Tsao et al., 2010 added nanoparticles of Al2O3
into Sn-3.5Ag-0.5Cu (SAC), and it enhanced the wetting
properties by reducing the contact angle to almost 29 with
0.5 weight per cent of Al2O3.
Figure 6 shows the contact angle of the Sn-3.5Ag-0.5Cu
(SAC) as the nanoparticles of Al2O3 are added. Comparable
to the addition of other elements, excessive addition of
151

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Figure 4 Wetting angle of SAC with different Ce and La additions

2.3 Microstructural properties of SAC


In the soldering processes, microstructural evaluation has been
found to be crucial for assessing strengthening effects in SAC solders
(El-Daly and Hammad, 2010; Garcia et al., 2010). The
microstructure serves as the initial property for an observation to
show how a structure provides a better joint (Garcia et al., 2010;
Duan et al., 2004). A fine and well-defined or distributed
morphology of the structure is important to provide better
strengthening of the solder joint (Roubaud et al., 2001; Laurila et al.,
2005). A uniform microstructure ensures a good and strong bond,
which also contributes to the solder joint reliability (Laurila et al.,
2005; Shen and Chan, 2009; Harrison, Vincent and Steen).
Therefore, the most important aspect that influences the strength is
the microstructure of the solder alloy (Gao et al., 2010; Gong et al.,
2009).
In the SAC solder system, the microstructure of the solder
mainly consists of primary Sn grains, eutectic structure particles
and intermetallic compound (IMC) particles (Gao et al., 2010;
Keller et al., 2011). A sufficient driving force will build up during
solidification, and then Sn nucleation will occur (Keller et al.,
2011; Shalaby, 2010). This is followed by growth of Sn dendrites
in the remaining molten solder (Gong et al., 2009). Meanwhile,
the dendrite structure formations are dependent on the content
of the Ag and Cu (Gong et al., 2009; Keller et al., 2011; Shalaby,
2010). The Sn dendrites will form earlier than the Sn matrix in
the eutectic phase, whereas in the Sn matrix, the dendrites grow
quickly, forming a large dendritic structure (Gong et al., 2009;
Wiese and Wolter, 2004).
The SAC eutectics is restrained by Sn atoms, and therefore
the growth of Sn crystals will not be significantly slowed down
by the depletion of the adjacent Sn atoms (Wiese and Wolter,
2004; Morando et al., 2012). Because Ag atoms have a higher
diffusion rate in the molten solder, they can diffuse out of the
way, enabling the Sn dendrites to grow (Wiese and Wolter,
2004). Also, Che et al. (2010) found that Sn dendrite
precipitation was finer with an increase in the Ag content
which directly increased the mechanical properties of the
solder. Hereafter, it is known that the Sn matrices in SAC
eutectics were formed due to the existing Sn dendrites (Wiese
and Wolter, 2004; Gong et al., 2009; Che et al., 2010).

Figure 5 Wetting force of SAC with different Ce and La additions

nanoparticles, for example the addition of 1 weight per cent


of Al2O3 appears to degrade the wetting properties of SAC
solder (Tay et al., 2013).
Particularly for the addition of nanoparticles to SAC, not
many studies have investigated the wettability properties
(Ervina et al., 2013). However, a preliminary knowledge
regarding the addition of Al2O3 nanoparticles is somewhat
important for further studies in terms of the wettability of
different nanoparticle additions to the SAC solder system.

2.3.1 Intermetallic and interfacial properties of properties of SAC


During soldering, an interfacial layer exists, and it is defined by
the IMC based on the reaction of the elements in the solder
(Anderson et al., 2001; Kikuchi et al., 2001). When a molten
solder alloy reacts with the substrate, a reaction occurs at the

Table II Influence on wetting properties of the addition of elements to SAC


Solder

Element added

Influence in wetting properties

Reference

Sn-0.3Ag-0.7Cu
Sn-3.5Ag-0.7Cu

Bi
Ce and La

(Gao et al., 2010)


(Law et al., 2006, Yu et al., 2004)

Sn-3.5Ag-0.7Cu

Mg

Sn-3.0Ag-0.5Cu

Ni

Sn-0.3Ag-0.7Cu
Sn-3.8Ag-0.7Cu

In
Er

Low Contact Angle


High Wetting Force
Low Contact Angle
Oxidize
Increase surface tension
Low Wetting Force
Longer Wetting Time
Low Contact Angle
Low Contact angle

152

(Gao et al., 2010)


(Yoon et al., 2009)
(Kanlayasari et al., 2009)
(Gao et al., 2010)

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Figure 6 Effect on contact angle of adding Al2O3 nanoparticles to


SAC

IMC. The composition of this area was analyzed by energy


dispersive X-ray to find the IMCs that had formed. The size
difference of the secondary domains (Figure 7) showed that
they were at different stages of growth and that their
distribution exhibited nucleation sites. The growth of the
interfacial site was divided into two processes as follows (Gong
et al., 2009; Gong et al., 2008):
1 they nucleate at different locations on the Cu surface first,
and then grow to form secondary domains; and
2 as soon as these domains are in contact, they combine to
form the main domain. The existence of small areas of
bare Cu in the main domain due to incomplete
coalescence proves this process.
The parts inside the domains are clearly composed of a large
number of coarse particles, which seem to have a certain crystal
structure (Gong et al., 2009). Continuing in the same study, it
was shown that the composition of the particles was close to
Cu6Sn5. The actual site was along the edge, or boundary, of the
domain, which was composed of a large number of fine particles
and named the transition zone in the same research.
When the reflow process continues, the entire Cu substrate is
covered with coarse particles (Gong et al., 2008; Roubaud et al.,
2001), indicating that the edge of the domain moves toward the
bare Cu if the liquid/solid reaction continues (Gong et al., 2009;
Yang et al., 2010). In this process, the current transition zone
consisting of fine particles will merge into a field and form coarse
particles that already exist inside the domain, and, meanwhile,
the new fine particles will be nucleated at the edge of a domain
(Gong et al., 2008; Gong et al., 2009). Therefore, the growth of
the domain is initiated from new and fine particles at the
transition zone (Gong et al., 2008). This actually simply
describes the process of the initial Cu6Sn5 IMC formation (Law
et al., 2006; Yang et al., 2010). The atomic ratios were measured,
and the planar intermetallic layer next to the solder was known to
possess compositions of Cu:Sn 6.1:5, which corresponds to
the Cu6Sn5 phase (Zhang et al., 2012a).
The IMC formation is vital for the solder to possess good joint
reliability, but excessive IMC formation is detrimental (Roubaud
et al., 2001). Relating to this, several studies have found that Cu3Sn
forms together with the Cu6Sn5 IMC, a later evolution of the
Cu6Sn5 IMC (Chang et al., 2006; Anderson et al., 2001; Liu et al.,
2009). This particular IMC is perceived to grow at the expense of
Cu6Sn5 and cause a thick intermetallic layer between the substrate
and the solder (Yoon et al., 2009). The growth of the Cu3Sn occurs
after aging and implies that the growth of Cu3Sn is much faster than
that of Cu6Sn5 and Ni Cu Sn IMCs for the Ni-plated solder
substrates (Rao et al., 2010). It has also been observed that the
growth of Cu6Sn5 intermetallic layers is retarded as time increases at
150C due to the formation Cu3Sn, i.e. at the expense of the
Cu6Sn5 layer (Gong et al., 2008). Once the Cu3Sn initiates at the
Cu/Cu6Sn5 interface, the growth rate of Cu6Sn5 will be controlled
by the limited amount of Cu in the solder (Gong et al., 2008; Peng
et al., 2007). Consequently, during isothermal aging, the Cu3Sn
layer appears to grow faster as it expands on both sides, resulting in
shifting of the Cu/Cu3Sn interface toward the Cu substrate
(Morando et al., 2012; Rao et al., 2010; Peng et al., 2007). From the
research of Yang et al. (2010), the chemical reaction (3) that follows
the formation of Cu6Sn5 is as follows:

interface, and this reactions product is the intermetallic property


(Laurila et al., 2005; Law et al., 2006). A thin IMC layer is
preferable, as it produces strong bonding at the interface (Laurila
et al., 2005; Tay et al., 2013). This IMC is crucial in enhancing
joint strength during service (Wiese and Wolter, 2004).
However, the downside is that an excessive IMC layer
thickness degrades the mechanical properties of the solder
joint and can promote a brittle failure mode due to crack
propagation (Gao et al., 2010; Law et al., 2006). Hence, it has
been found that the addition of an optimum composition of
some elements (Law et al., 2006; Anderson et al., 2001; Yu
et al., 2004) and nanoparticles (Chang et al., 2011; Tsao et al.,
2010; Gain et al., 2011; Fouzder et al., 2011; Haseeb et al.,
2012) can avoid this weakness. Further information regarding
this particular study will be discussed later in this paper.
Based on most studies, the IMC appeared to spread
between the solder and substrate in the SAC solder (Anderson
et al., 2001; Yoon et al., 2009). The intermetallics that coexist
with the SAC solder are Cu6Sn5, Cu3Sn and Ag3Sn (Zhang
et al., 2012a; Gong et al., 2008, Gong et al., 2009; Anderson
et al., 2001; Yoon et al., 2009). Most studies reveal that the
Cu6Sn5 exists in a scallop shape and between the solder and
the substrate (Gao et al., 2010; Zhang et al., 2012a; Yoon
et al., 2009). Gong et al., 2009 conducted an experiment,
revealing the process of Cu6Sn5 IMC formation. In the study,
the interfacial reactants were found for the first time at a
temperature around 232C (Wiese and Wolter, 2004). It was
observed that the interface comprised one main area domain
with several associated secondary domains, as shown in Figure
7. (Wiese and Wolter, 2004). This domain was indicated as
the area where the solder and substrate reacted to produce the
Figure 7 Interfacial formation in SAC solder with a Cu substrate

5Cu3Sn Cu6Sn5 9Cu


153

(3)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Likewise, this equation was supported by Peng et al. (2007).


In that study, the Cu3Sn was understood to be formed due to
solid-state reaction, where a new layer of stack was seen to be
formed between the solder and substrate which consisted of
solder/Cu6Sn5 and Cu3Sn/Cu. The Cu3Sn IMC grew ten
times faster than Cu6Sn5, and the increase of total IMC
thickness was mainly due to the growth of Cu3Sn rather than
Cu6Sn5 (Peng et al., 2007). The explanation is that after the
Cu atoms arrive at the interface of Cu3Sn/Cu6Sn5 by diffusion
through the grain boundaries of the Cu3Sn layer, the following
interfacial reaction (4) takes place:

works (Liu et al., 2009; Morando et al., 2012). In an


experiment conducted by Liu et al., 2009 using an organic
solderability preservative surface finish, the Ag3Sn tended to
adsorb at the Cu6Sn5 surface due to the larger size and fast
growth of this IMC, and, apparently, the Ag3Sn restricted
further growth of the Cu6Sn5 IMC to provide a better joint. In
the same perspective, SAC with a lower Ag content gave better
sparsely distributed Ag3Sn IMC particles (Che et al., 2010).
The Ag3Sn IMC appeared like a plate-type shape in the solder
matrix (Yoon et al., 2009; Keller et al., 2011). In another
experiment, the formation of Ag3Sn due to a slow cooling rate
was found at the interface of Cu6Sn5 by Kotadia et al. (2012).
This result was similar to the study of Che et al. (2010). The
Ag3Sn seemed to form as larger plates in the microstructure
which could degrade the strength (Kotadia et al., 2012).
Meanwhile, the Sn-rich intermetallic phases are known to
have nominal compositions of Ag3Sn, with the concentration
of Sn varying between 25.5 and 26 weight per cent, and the
solid solubility of Ag in Sn was 0.1 weight per cent (Abtew
and Selvaduray, 2000; Wiese and Wolter, 2004).

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Cu6Sn5 9Cu 5Cu3Sn

(4)

Here, Cu6Sn5 was converted to Cu3Sn at the interface (Zhang


et al., 2012a; Peng et al., 2007). Due to this reaction, the amount
of Cu atoms that can diffuse to the interface of Cu6Sn5/solder is
greatly reduced (Zhang et al., 2012a; Morando et al., 2012; Rao
et al., 2010; Peng et al., 2007). As a result, Cu3Sn grows rapidly
with increasing temperature and time by consuming Cu6Sn5 at
the interface of Cu3Sn/Cu6Sn5 (Rao et al., 2010; Peng et al.,
2007; Yang et al., 2013).
The growth of Cu6Sn5 on the solder side mainly depends on
the accessibility of Cu atoms in the solder (Rao et al., 2010;
Peng et al., 2007). Because most of the Cu atoms in the bulk
solder have been consumed to form Cu6Sn5 particles in the
eutectic structure, the amount of free Cu atoms that can
diffuse to the solder/Cu6Sn5 interface is very small, which
limits the growth of Cu6Sn5 on the solder side (Gong et al.,
2009; Peng et al., 2007). Therefore, after some time, the
Cu3Sn layers are expanded on both sides, resulting in shifting
of the Cu/Cu3Sn interface toward the Cu pad and the Cu3Sn/
Cu6Sn5 interface toward the Cu6Sn5. This is the reason for
the slower growth of Cu6Sn5 than Cu3Sn at higher
temperatures (Zhang et al., 2012a; Rao et al., 2010; Peng
et al., 2007).
By the same reaction, in another study, the Cu6Sn5 was
found to be converted to Cu3Sn at the interface (Zhang et al.,
2012a). Thus, the amount of Cu atoms that could diffuse to
the interface of Cu6Sn5/solder was found to be greatly reduced
again. The growth of Cu6Sn5 on the solder side yet again
depends on the availability of Cu atoms in the solder, which
matches the explanation proposed by Liu et al., 2009. As for
during thermal aging, the Cu3Sn layer stretched on both sides
causing a thicker Cu3Sn layer (Zhang et al., 2012a). This
similarity was found by other researchers (Zhang et al., 2012a;
Peng et al., 2007). Figure 8 shows a simple interface between
the solder and substrate together with the IMC formed for an
SAC solder alloy with a Cu substrate.
Besides Cu6Sn5 and Cu3Sn IMC, the Ag3Sn IMC has also
been observed to be dispersed on the Sn matrix in several

2.3.2 Microstructures and interfacial properties of SAC with added


elements
The previous explanations only concerned the IMC and
microstructures of SAC solders. There are also several additions
of elements and particles that cause some alteration in the
interfacial and microstructural properties (Gao et al., 2010; Law
et al., 2006; Yu et al., 2004). Adding elements tends to provide a
better refined microstructure compared with the bare SAC (Law
et al., 2006; Anderson et al., 2001). Adding elements such as Co
(Anderson et al., 2001), La (Law et al., 2006; Yu et al., 2004), Ce
(Law et al., 2006) and Bi (Yu et al., 2004) refines the
microstructure of the solder. These elements refine the
microstructure by acting as barriers for the growth of an excessive
IMC layer that decreases the joint reliability (Law et al., 2006; Yu
et al., 2004; Anderson et al., 2001). This is mainly due to the
adsorption effect of these active elements (Yu et al., 2004;
Kotadia et al., 2012; Gain et al., 2011).
According to Chuang et al., 2012; adding Ti into
Sn-3.5Ag-0.5Cu (SAC) revealed that the grain size of Sn and
the width of the eutectic area became much finer. As further
additions were made, the grain size was almost half of that of
the Sn-3.5Ag-0.5Cu (SAC), as computed in Table III
(Chuang et al., 2012). While adding elements like Ce,
antimony or La, the grain size and the microstructure were
also refined, which can be explained by the theory of Gibbs
energy which is lower in the SAC modified with the addition
of elements (Gao et al., 2010). The addition of Zn,
meanwhile, reduces the size of large Ag3Sn plates and permits
a better microstructure in Sn-3.8Ag-0.7Cu (SAC) solder
joints, which is also detailed in Table III (Kotadia et al., 2012).
The addition of these elements tends to provide a better
interfacial layer by producing a thin and preferable
intermetallic layer (Kotadia et al., 2012; Mayappan and
Ahmad, 2010). For instance, the element Ce has high affinity
to the constituent of Sn which will result in a decreased driving
force for CuSn IMC formation (Law et al., 2006). Similarly,
adding In into Sn-0.3Ag-0.7Cu provides a uniform
distribution of the IMC and microstructure on the solder
(Kanlayasari et al., 2009).

Figure 8 Cu3Sn and Cu6Sn5 IMCs between solder and substrate

154

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Table III Microstructure of elements added to SAC


Specimen (addition)

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Sn-3.5Ag-0.5Cu (SAC)
SAC-0.25Ti
SAC-0.5Ti
SAC-1Ti
Sn-3.8Ag-0.7Cu (SAC)
SAC-0.5Zn
SAC-1Zn
SAC1.5Zn

Sn (m)

Average Ag3Sn (m)

Eutectic area (m)

References

24.8 5.9
13.2 4.2
7.2 2.1
4.8 4.8

73
57
45
23

6.8 2.8
4.5 1.2
2.1 0.8
1.2 0.4

(Chuang et al., 2012)

(Kotadia et al., 2012)

In another experiment, Co nanoparticles were added to


Sn-3.8Ag-0.7Cu (SAC), and the IMC layer appeared to have
a relatively flat morphology in the Co-containing solder
(Haseeb and Leng, 2011). In the same research, the thickness
of Cu3Sn which formed due to diffusion in SAC was larger as
compared with the Co-added SAC. Similarly, during
high-temperature exposure, the Co nanoparticles delayed the
growth of Cu3Sn (Haseeb and Leng, 2011; Chang et al.,
2011). Clearly, the presence of Co nanoparticles significantly
reduced the effective interdiffusion coefficient of Cu3Sn and
suppressed the growth of this IMC (Haseeb and Leng, 2011).
This explanation was supported by Lin et al., 2003.
Meanwhile, the addition of molybdenum (Mo)
nanoparticles contributed to a decrease in the overall IMC
thickness, as noted in Figure 11 (Haseeb et al., 2012). The
retardation of the growth of IMC thickness and scallop
diameter was mainly due to the effect of Mo nanoparticles
which acted as discrete particles by absorbing at the grain
boundaries of interfacial IMC scallops (Haseeb et al., 2012).
The blocking of channels for diffusion by these nanoparticles
impeded the movement of Cu atoms from the substrate to the
liquid solder (Haseeb et al., 2012; Tsao et al., 2012). This
particular characteristic directly reduced the thickness of the
IMC layer (Gao et al., 2010). Also, the adsorbed Mo
nanoparticles on the IMC surface acted as a barrier to the
coalescence of neighbouring scallops which caused a
reduction in the scallop diameter (Haseeb et al., 2012).
The presence of ZrO2 nanoparticles seems to inhibit the
formation of the IMC layer in Sn-3.5Ag-0.5Cu (SAC) solder
alloy (Gain et al., 2011). The second-phase ZrO2
nanoparticles were found to alter the driving force for the
growth of the IMC layer, as well as the diffusivity of the
elements involved in its growth (Gain et al., 2011). Similarly,
the relative relationship of the growth velocities between
crystalline directions of the IMC particles also varied and thus
reduced the IMC particle size, even after several reflows
(Figure 12) (Gain et al., 2011; Lin et al., 2003) (Table IV).
Addition of TiO2 nanoparticles (Tsao et al., 2012; Chang
et al., 2011) and Al2O3 (Tsao et al., 2010) also benefitted the
microstructure of the SAC solder by reducing the average
spacing between the IMC and the grain size (Table V). The
explanation is similar to the research by Gain et al., 2011. This
particular reduction in average size and grain size is important
in producing better joint reliability with well-defined
microstructure for improved strength (Chang et al., 2011;
Kotadia et al., 2012).

Furthermore, with a reduced driving force for the diffusion of


Cu due to the addition of elements, the Cu atomic flux moving
from the Cu side to the solder side decreases, lowering the
reaction for the formation of Cu3Sn (Gao et al., 2010).
Consequently, adding Ce into Sn-3.5Ag-0.7Cu (SAC) affects
the growth thickness of the IMC (Gao et al., 2010; Law et al.,
2006). Similarly, addition of elements such as iron and Ge can
suppress the growth of the intermetallic layer (Gao et al., 2010).
Meanwhile, Kotadia et al. (2012) revealed that the thickness of
the interfacial layer, after aging was done, was lower in the
Zn-containing Sn-3.8Ag-0.7Cu/0.5-1.5Zn (SAC) than in the
bare Sn-3.8Ag-0.7Cu (SAC). In the same research, Zn
suppressed the voids of Cu3Sn and produced a better joint. In
another research, the addition of Ni retarded growth of the
Cu3Sn by forming a better thin layer of (Cu, Ni)6Sn5, and this
IMC was said to be stable (Yoon et al., 2009).
Adding elements such as Ce and La was believed to provide
the vitamin that was able to suppress the growth of the
intermetallic layer (Gao et al., 2010). These particles do not
form a solid solution and act as a barrier for the growth of
thicker Cu3Sn which is found to be the main cause of the
whole thickening of the interfacial layer (Gao et al., 2010;
Zhang et al., 2012a). The data in Figure 9 show the effect of
adding Ce and La on the interfacial layer as aging time
increases (Gao et al., 2010).
2.3.3 Microstructures and interfacial properties of SAC with added
nanoparticles
Besides adding elements, studies have confirmed that the
addition of nanoparticles also gives benefits to the SAC
microstructure and interface (Chang et al., 2011; Tsao et al.,
2010; Gain et al., 2011; Fouzder et al., 2011). Nanoparticles
were found to produce a lower diffusion coefficient which
reduced the growth of thicker IMCs in the solder (Chang
et al., 2011). Besides that, these nanoparticles did not take
part in diffusion processes and later acted as a dispersion
strengthening mechanism, creating better microstructure, as
well as retarding the growth of total IMC layer (Chang et al.,
2011; Gain et al., 2011; Fouzder et al., 2011).
Addition of SrTiO3 nanoparticles into Sn-3.5Ag-0.5Cu
(SAC) promotes a high nucleation density in the eutectic
colonies during solidification (Fouzder et al., 2011). These
second-phase SrTiO3 nanoparticles and the fine IMC formed
due to their presence retard the growth rate of IMCs (Fouzder
et al., 2011; Haseeb et al., 2012). Figure 10 shows that, even
after several reflows, SAC with added SrTiO3 nanoparticles
produces a thinner IMC than the base SAC.
155

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Figure 9 IMC thickness vs aging time for SAC with different added
elements (Ce and La)

Figure 12 Effect of ZrO2 nanoparticles on SAC IMC thickness

Table IV IMC thickness of SAC with/without the addition of nanoparticles

Figure 10 IMC thickness vs number of reflows for both SAC and SAC
with added SrTiO3

Nanoparticles
TiO2
ZrO2
Mo

Thickness of IMC
(m)
Without
With
addition
addition
4.44
2.8
2.2

References
(Chang et al., 2011)
(Gain et al., 2011)
(Haseeb et al., 2012)

2.78
2.0
1.3

Table V Grain size of SAC and SAC with added nanoparticles

Specimen
Sn-3.5Ag-0.7Cu
(SAC)
SAC-0.5TiO2
Sn-3.5Ag-0.5Cu
(SAC)
SAC-0.25Al2O3
SAC-0.5 Al2O3
SAC-1 Al2O3

Figure 11 Influence of Mo nanoparticles in SAC on IMC thickness


with different reflows

Average
Grain size (m)
spacing (m)
Ag3Sn
0.54 0.32

0.63 0.31

0.28 0.18
2.55

0.19 0.08
0.85

0.74
0.27
0.23

0.74
0.27
0.23

References
(Tsao et al.,
2012)
(Tsao et al.,
2010)

crucial factor in maintaining the solders integrity and


performance (Shen and Chan, 2009; Ma and Suhling, 2009).
This property helps to provide better tensile and shear
strengths and hardness of the solder.
SAC containing the lowest Ag content produces less shear
strength compared with the high content of Ag, then again, over
addition causes degradation in the strength (Keller et al., 2011).
In the same research, the amount of 3 weight per cent Ag, which
is usually the content for the eutectic composition (Abtew and
Selvaduray, 2000; Law et al., 2006; Chang et al., 2006), produces
a high strength (Keller et al., 2011). Compared with SnPb solder,
SAC solder possesses better shear strength. Further adding Pb
into Sn will cause deterioration in the strength of the SnPb
solder, while it is different for the SAC solder, where only a slight
addition of Ag significantly enhances the strength of SAC (Gong
et al., 2009). In terms of ductility, the SnPb solders ductility
decreases severely with the addition of Pb; nonetheless, this never
exceeds the maximum elongation possessed by the SAC (Gong

The microstructures and interfacial properties are important


factors affecting the mechanical properties of SAC solder.
Hence, some knowledge of these properties will be beneficial for
better understanding of the mechanical properties.
2.4 Mechanical properties of SAC
Mechanical properties are related to the strength of a solder
alloy and are the most important aspect in the electronic
packaging industry (Whalley, 2004; Manko, 2001). In
producing a reliable solder alloy, strength is concerned as the
156

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

et al., 2009). This shows that the SAC solder actually provides a
better property than the SnPb solder. This observation was
equivalent to the research of Liu et al., 2009.
The tensile strength of SAC solder increases with increasing
Ag content; nevertheless, increasing the Ag content in
Sn(3.0/2.0/1.0)Ag 0.5Cu solder reduced the ductility (Che
et al., 2010; Gong et al., 2009). It can be seen that the elastic
modulus, yield stress and ultimate tensile strength (UTS)
increased (Figure 13) with increasing Ag content for all strain
rates (Che et al., 2010). Because the total elongation at the
fracture of a solder increases with increasing strain rate, it
shows that SAC with a larger content of Ag exhibits a larger
elastic modulus than for SAC with a lower Ag content (Che
et al., 2010). Also, at a large strain rate, the creep of a solder
could be fully developed, which means that at the high rate,
the solder should exhibit higher elastic modulus, yield stress
and UTS, and all these correlate with the higher Ag content as
studied by Che et al., 2010 (Keller et al., 2011).

Table VI Effect of Ti addition on the UTS and yield strength (YS) of SAC
Specimen
(addition)
Sn-3.5Ag-0.5Cu
(SAC)
SAC-0.25Ti
SAC-0.5Ti
SAC-1Ti

UTS
(MPa)

0.2 YS
(MPa)

References

47.2 4.8
50.1 4.6
53.2 5.2
57.6 3.8

43.1 3.6
46.3 4.2
52.1 4.5
54.5 4.7

(Chuang et al., 2012)

and finer distribution of IMC which contributes to blocking


the movement of dislocation along with reducing stress
concentration (Lin and Chuang, 2010; Chang et al., 2011;
Haseeb and Leng, 2011).
Kanlayasari et al. (2009) conducted research into the
addition of In into Sn-0.3Ag-0.7Cu (SAC) to test the
hardness and the tensile strength of the modified SAC. In this
research, it was shown that the hardness increased by about 81
per cent, while the tensile strength also increased by 79 per
cent after adding In. This was ascribed to the fine and uniform
distribution of the IMC that influenced the strength of the
solder by providing better solder joints (Kanlayasari et al.,
2009; Laurila et al., 2005). Figures 14 and 15 display, as In
was added, both the hardness and tensile strength increases
due to the explanation that been stated.

2.4.1 Mechanical properties of SAC with added elements


Researchers have come up with some factors that comprise the
main measures for producing a highly strengthened solder
(Humpston and Jacobson, 2004; Zhang et al., 2010b; Kikuchi
et al., 2001; Tsukamoto et al., 2010). Correspondingly, adding
some elements influences in providing better mechanical
properties for the SAC solder (Chang et al., 2011; Tsao et al.,
2010; Gain et al., 2011; Fouzder et al., 2011). Adding Ce and
La provides better shear strength than for the Sn-3.5Ag-0.7Cu
bulk (Gao et al., 2010; Law et al., 2006). Besides that, adding
Bi into Sn-3.5Ag-0.7Cu also produces higher shear strength
than the prior one (Gao et al., 2010; Law et al., 2006; Wiese
and Wolter, 2004). A better microstructure provides the
enhancement in higher yield strength and UTS, where these
criteria are seen by adding Ti into Sn-3.5Ag-0.5Cu (SAC)
(Chuang et al., 2012). Table VI shows the effect of Ti addition
on the strength.
In another study, adding Ce decreased the tensile strength
of the Sn3Ag 0.5Cu alloy (SAC) from 43.7 to 21.9 MPa,
which was concluded to be due to the stress concentration
caused by the coarse CeSn3 IMC in SAC with added Ce (Lin
and Chuang, 2010). However, in the same study, adding Zn
to the SAC Ce element tended to increase the tensile
strength from 21.9 to 65.3 MPa and 64.3 MPa for 0.2 and 0.5
weight per cent additions of Zn, respectively (Lin and
Chuang, 2010). The reason behind this is due to a uniform

2.4.2 Mechanical properties of SAC with added nanoparticles


Nanoparticle additions to Sn-Ag-Cu (SAC) refine the
microstructure and provide a better mechanical strength in the
solder alloy (Gain et al., 2011; Tsao et al., 2012). The particles
capability of acting as an obstacle for dislocation motion must be
Figure 14 Hardness of SAC with added In

Figure 13 Effect on UTS of increasing Ag content in SAC


Figure 15 Tensile strength of SAC with added In

157

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

the reason for the enhancement of the mechanical properties


(Gain et al., 2011). The particles must be sufficiently fine and
stable with respect to their size and interparticle spacing,
which relates to a proper and fine microstructure (Whalley,
2004; Duan et al., 2004). Mainly, this is achievable by
second-phase dispersion strengthening, where the solder joint
shows better strength than the unreinforced SAC solder
(Chang et al., 2011).
Surface absorption theory can be utilized to explain the
refinement of Ag3Sn particles in the SAC composite solder,
which is similar to the study of Tsao et al. (2010) (Chang et al.,
2011). The presence of TiO2 nanoparticles retarded the growth
and restrained drifting of the IMC layer by effectively acting as a
barrier for Sn diffusion toward the Cu substrate, thus hindering
the growth of the interfacial IMC layer. The result is shown in
Table IV and indicates that the SAC with added nanoparticles
created a thinner IMC, as per the explanation above (Haseeb and
Leng, 2011; Chang et al., 2011).
As for the ZrO2 addition, due the second-phase dispersion
strengthening of these nanoparticles, the Sn-3.5Ag-0.5Cu
(SAC) solder joints containing ZrO2 nanoparticles produced
higher strength than the unreinforced SAC solder joints, even
after several reflows were made (Figure 16) (Gain et al., 2011;
Lin et al., 2003). The improvement in strength was ascribed to
the presence and distribution of fine dispersoid particles that
influenced the deformation characteristics by inhibiting grain
boundary sliding and, at the same time, retarding the
dislocation movement (Chang et al., 2011).
There was an increase in microhardness resulting from the
addition of TiO2 nanoparticles to the solder matrix
Sn-3.5Ag-0.5Cu (SAC) (Chang et al., 2011). These reinforcing
TiO2 nanoparticles and the refined IMC of submicron Ag3Sn
due to the addition can be considered as the main reasons for the
improved microhardness Sn-3.5Ag-0.5Cu (SAC)/TiO2
composite solders (Tsao et al., 2010; Chang et al., 2011). As
active nanoparticles, TiO2 will accumulate at the interface of
Ag3Sn intermetallic particles and reduce the growth of coarser
Ag3Sn (Chang et al., 2011; Tsao et al., 2012). A similar increase
in hardness was established by adding Al2O3 to Sn-3.5Ag-0.5Cu
(SAC) (Tsao et al., 2012).

shown good benefits and are now widely used in the


electronics industry. This particular solder is perceived to offer
better properties in terms of melting, wettability,
microstructure and interfacial, as well as mechanical
properties in terms of tensile and shear strengths. All these
properties were discussed as crucial aspects in providing a
viable solder. This paper has reviewed all these aspects of this
particular SAC alloy together with the addition of elements
and nanoparticles into the solder. Many studies suggest that
adding elements and nanoparticles produces additional
benefits in the SAC solder system. Besides, some research was
conducted by synthesizing the SAC solders into nanoparticles
and this was also proven to possess better properties.
In terms of the melting temperature, the SAC solder seems to
possess a higher melting temperature than the SnPb solder.
However, the melting temperature of this solder falls within the
range of the average reflow temperature in the electronic
packaging industry and would not really affect the process of
device interconnection. An encouraging thing here is that this
melting temperature can actually be reduced by adding elements
such as Ti and Zn. Adding these elements tends to decrease the
melting temperature of the SAC solder alloy by about 3C.
Adding nanoparticles, meanwhile, tends to increase the melting
temperature slightly. Nonetheless, these increments will not
damage other devices due to a very slight increment only with no
drastic changes in the solidification temperature.
As for wettability concerns, the SAC solder gives a better
spreading area than the SnPb solder. This particular property is
important in producing a good joint between the solder and the
substrate so that the device is kept intact for a long service time.
Not only that, the SAC solder appears to produce a higher
wetting force and lower contact angle after being added with
elements such as Bi, Er, Ce and La. Even so, the addition of
some elements such as Mg, which is prone to oxidation, damages
the wetting properties. Further addition of nanoparticles also
produces better wettability between the substrate and the SAC
solders. These influences have been discussed and summarized.
As for the microstructural properties, the SAC solder
comprises a fine and well-defined microstructure, which directly
enhances the strength of the solder. After soldering, the SAC
solder contains few IMCs due to the diffusion activity between
the solder and the substrate. These IMCs are important to
produce a good bond between the solder and substrate for
long-term usage. The SAC solder produces a better IMC after
soldering and gives better joint integrity. However, in the long
term, these IMCs tend to become thicker and degrade the bond.
Hence, studies have proposed an alternative by adding elements
or nanoparticles to overcome this problem. Adding elements and
nanoparticles reduces the diffusion coefficient and thus has an
influence in producing a thinner IMC, which gives a better joint.
This shows that the SAC solder is adaptable via these additions
to produce better interfacial properties which are important for
the mechanical properties of the solder.
The mechanical properties are influenced by the
microstructure and interfacial properties of a solder. In this
particular SAC solder system, the eutectic SAC solders have
better mechanical properties by inducing higher shear and tensile
strengths than the other Pb-free solder alloy. Additions of
elements such as Ti and nanoparticles like ZrO2 also give an
improvement for the mechanical properties of the SAC solder.

3. Conclusion
Various Pb-free solder alloys that overcome the health and
environmental concerns caused by SnPb solders have been
studied and investigated. Among them, the SAC alloys have
Figure 16 Shear strength of ZrO2-reinforced SAC

158

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Therefore, the mechanical properties of the SAC solder seen to


have better and enhanced results for use in the electronic
industry and suggest that it can withstand stress and impact
during its service time.
This paper has reviewed all the properties of the Pb-free SAC
solder system from its development for overcoming
environmental problems to achieving and sustaining itself as a
viable candidate in the electronic packaging industry. The
Pb-free SAC solder can offer an alternative to all the drawbacks
that traditional SnPb solder possesses and also be involved in
upcoming new innovations for future needs. Although many
studies have been performed on this particular solder, not much
information has been gathered in a review of this sort to give a
better understanding of SAC solder alloys. This paper has
reviewed the importance of SAC solder in the electronic
packaging industry and provided information to enable better
knowledge of the alloy.

application, 35th International Electronics Manufacturing


Technology, Perak, Malaysia, p. 59.
Ervina, E.M.N. and Tan, S.Y. (2013), Wettability of molten
Sn-Zn-Bi solder on Cu substrate, Applied Mechanics and
Materials, Vol. 315, pp. 675-680.
Ervina, E.M.N., Singh, A. and Chuan, Y.T. (2013), A
review: influence of nano particles reinforced on solder
alloy, Soldering and Surface Mount Technology, Vol. 25
No. 4, pp. 229-241.
Fouzder, T., Shafiq, I., Chan, Y.C., Sharif, A. and
Yung, W.K.C. (2011), Influence of SrTiO3 nanoparticles
on the microstructure and shear strength of Sn-Ag-Cu
solder on Au/Ni metallized Cu pads, Journal of Alloys and
Compounds, Vol. 509 No. 5, pp. 1885-1892.
Gain, A.K., Chan, Y.C. and Yung, W.K.C. (2011), Effect of
additions of ZrO2 nano-particles on the microstructure and
shear strength of SnAgCu solder on Au/Ni metallized Cu
pads, Microelectronics Reliability, Vol. 51 No. 12,
pp. 2306-2313.
Gao, Y., Zou, C., Yang, B., Zhai, Q., Liu, J., Zhuravlev, E.
and Schick, C. (2009), Nanoparticles of SnAgCu lead-free
solder alloy with an equivalent melting temperature of SnPb
solder alloy, Journal of Alloys and Compounds, Vol. 484 Nos
1/2, pp. 777-781.
Gao, L., Xue, S., Zhang, L., Sheng, Z., Ji, F., Dai, W.,
Yu, S.L. and Zeng, G. (2010), Effect of alloying elements
on properties and microstructures SnAgCu solders,
Microelectronic Engineering, Vol. 87 No. 11, pp. 2025-2034.
Garcia, L.R., Osrio, W.R., Peixoto, L.C. and Garcia, A.
(2010), Mechanical properties of Sn-Zn lead-free solder
alloys based on the microstructure array, Materials
Characterization, Vol. 61 No. 2, pp. 212-220.
Gong, J., Liu, C., Conway, P.P. and Silberschmidt, V.V.
(2009), Initial formation of CuSn intermetallic
compounds between molten SnAgCu solder and Cu
substrate, Scripta Materialia, Vol. 60 No. 5, pp. 333-335.
Gong, J., Liu, C., Conway, P.P. and Silberschmidt, V.V.
(2008), Evolution of CuSn intermetallics between molten
SnAgCu solder and Cu substrate, Acta Materialia, Vol. 56
No. 16, pp. 4291-4297.
Gong, J., Liu, C., Conway, P.P. and Silberschmidt, V.V.
(2009), Formation of Sn dendrites and SnAg eutectics in a
SnAgCu solder, Scripta Materialia, Vol. 61 No. 7,
pp. 682-685.
Haseeb, A.S.M.A., Arafat, M.M. and Johan, M.R. (2012),
Stability of molybdenum nanoparticles in Sn3.8Ag 0.7Cu
solder during multiple reflow and their influence on interfacial
intermetallic compounds, Materials Characterization, Vol. 64,
pp. 27-35.
Haseeb, A.S.M.A. and Leng, T.S. (2011), Effects of Co
nanoparticle addition to Sn-3.8Ag-0.7Cu solder on
interfacial structure after reflow and ageing, Intermetallics,
Vol. 19 No. 5, pp. 707-712.
Humpston, G. and Jacobson, D.M. (2004), Principles of
Soldering, ASM International, Materials Park, OH,
pp. 1-47.
Islam, M.N., Chan, Y.C., Rizvi, M.J. and Jillek, W. (2005),
Investigations of interfacial reactions of SnZn based and
SnAgCu lead-free solder alloys as replacement for SnPb

References
Abtew, M. and Selvaduray, G. (2000), Lead-free solders in
microelectronics, Materials Science and Engineering, Vol. 27
Nos 5/6, pp. 19-141.
Anderson, I.E., Foley, J.C., Cook, B.A., Harringa, J.,
Terpstra, R.L. and Unal, O. (2001), Alloying effects in
near-eutectic Sn-Ag-Cu solder alloys for improved
microstructural stability, Journal of Electronic Materials,
Vol. 30 No 9, pp. 1050-1059.
Chuang, C.L., Tsao, L.C., Lin, H.K. and Feng, L.P. (2012),
Effects of small amount of active Ti element additions on
microstructure and property of Sn3.5Ag0.5Cu solder,
Materials Science and Engineering A, Vol. 558, pp. 478-484.
Chang, T.-C., Wang, J.-W., Wang, M.-C. and Hon, M.-H.
(2006), Solderability of Sn9Zn 0.5Ag1In lead-free
solder on Cu substrate: Part 1. Thermal properties,
microstructure, corrosion and oxidation resistance, Journal
of Alloys and Compounds, Vol. 422 Nos 1/2, pp. 239-243.
Chang, S.Y., Jain, C.C., Chuang, T.H., Feng, L.P. and
Tsao, L.C. (2011), Effect of addition of TiO2
nanoparticles on the microstructure, microhardness and
interfacial reactions of Sn3.5AgXCu solder, Materials and
Design, Vol. 32 No. 10, pp. 4720-4727.
Che, F.X., Zhu, W.H., Poh, E.S.W., Zhang, X.W. and
Zhang, X.R. (2010), The study of mechanical properties
of SnAgCu lead-free solders with different Ag contents
and Ni doping under different strain rates and
temperatures, Journal of Alloys and Compounds, Vol. 507
No. 1, pp. 215-224.
Duan, L.L., Yu, D.Q., Han, S.Q., Ma, H.T. and Wang, L.
(2004), Microstructural evolution of Sn9Zn3Bi solder/
Cu joint during long-term aging at 170C, Journal of Alloys
and Compounds, Vol. 381 Nos 1/2, pp. 202-207.
El-Daly, A.A. and Hammad, A.E. (2010), Elastic properties
and thermal behavior of SnZn based lead-free solder
alloys, Journal of Alloys and Compounds, Vol. 505 No. 3,
pp. 793-800.
Ervina, E.M.N. and Marini, A. (2012), A review of solder
evolution in electronic application, International Journal of
Engineering and Applied Sciences, Vol. 1 No. 1, pp. 1-10.
Ervina, E.M.N. and Singh, A. (2012), Characterization of
mechanical testing on lead free solder on electronic
159

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

solder, Journal of Alloys and Compounds, Vol. 400 Nos 1/2,


pp. 136-144.
Jeon, S.-J., Hyun, S., Lee, H.-J., Kim, J.-W., Ha, S.-S.,
Yoon, J.-W., Jung, S.-B. and Lee, H.-J. (2008),
Mechanical reliability evaluation of Sn-37Pb solder joint
using high speed lap-shear test, Microelectronic Engineering,
Vol. 85 No. 10, pp. 1967-1970.
Kanlayasiri, K., Mongkolwongrojn, M. and Ariga, T. (2009),
Influence of indium addition on characteristics on
Sn-03.Ag-0.7Cu solder alloy, Journals of Alloys and
Compounds, Vol. 485 Nos 1/2, pp. 225-230.
Kikuchi, S., Nishimura, M., Suetsugu, K., Ikari, T. and
Matsushige, K. (2001), Strength of bonding interface in
lead-free Sn alloy solders, Materials Science and Engineering
A, Vols. 319/321, pp. 475-479.
Kotadia, H.R., Mokhtari, O., Clode, M.P., Green, M.A. and
Mannan, S.H. (2012), Intermetallic compound growth
suppression at high temperature in SAC solders with Zn
addition on Cu and Ni-P substrate, Journal of Alloys and
Compounds, Vol. 511 No. 1, pp. 176-188.
Keller, J., Baither, D., Wilke, U. and Schmitz, G. (2011),
Mechanical properties Pb-free SnAg solder joints, Acta
Materialia , Vol. 59 No. 7, pp. 2731-2741.
Laurila, T., Vuorinen, V. and Kivilahti, J.K. (2005),
Interfacial reactions between lead-free solders and
common base materials, Materials Science and Engineering
R, Vol. 49 Nos 1/2, pp. 1-60.
Law, C.M.T., Wu, C.M.L., Yu, D.Q., Wang, L. and
Lai, J.K.L. (2006), Microstructure, solderability, and
growth of intermetallic compounds of Sn-Ag-Cu-RE
lead-free solder alloys, Journal of Electronic Materials,
Vol. 35 No. 1, pp. 89-93.
Lee, N.-C. (2002), Reflow Soldering and Trouble Shooting,
Newnes, Butterworth-Heinemann, UK.
Li, F.-Q. and Wang, C.Q. (2006), Influence of interfacial
reaction between molten SnAgCi solder droplet and Au/Ni/
Cu pad on IMC evolution, Transactions of Nonferrous
Metals Society China, Vol. 16 No. 1, pp. 18-22.
Lin, H.-J. and Chuang, T.-H. (2010), Effects of Ce and Zn
additions on the microstructure and mechanical properties
of Sn3Ag 0.5Cu solder joints, Journal of Alloys and
Compounds, Vol. 500 No. 2, pp. 167-174.
Lin, D.C., Liu, S., Guo, T.M., Wang, G.-X., Srivatsan, T.S.
and Petraroli, M. (2003), An investigation of nanoparticles
addition on solidification kinetics and microstructure
development of tin-lead solder, Materials Science and
Engineering A, Vol. 360 Nos 1/2, pp. 285-292.
Liu, P., Yao, P. and Liu, J. (2009), Effects of multiple reflows
on interfacial reaction and shear strength of SnAgCu and
SnPb solder joints with different PCB surface finishes,
Journal of Alloys and Compounds, Vol. 470, pp. 188-194.
Lu, S., Lou, F. and Wang, B.H. (2008), International
conference of Electronic Packaging Technology and High Density
Packaging., Shangai, China.
Ma, H. and Suhling, J.C. (2009), A review of mechanical
properties of lead-free solders for electronic packaging,
Journal of Materials Science, Vol. 44 No. 5, pp. 1141-1158.
Mahmudi, R., Geranmayeh, A.R., Noori, H., Jahangiri, N.
and Khanbareh, H. (2008), Effect of cooling rate on the
room-temperature impression creep of lead-free Sn9Zn

and Sn 8Zn3Bi solders, Materials Science and Engineering


A, Vol. 487 Nos 1/2, pp. 120-125.
Manko, H.H. (2001), Solders and Soldering Materials: Design
Production, Analysis for Reliable Bonding, 4th ed., McGraw
Hill, New York, NY.
Mayappan, R. and Ahmad, Z.A. (2010), Effect of Bi addition
on the activation energy for the growth of Cu5Zn8
intermetallic in the SnZn lead-free solder, Intermetallics,
Vol. 18 No. 4, pp. 1730-1735.
Mei, Z., Holder, H.A. and Vander Plas, H.A. (2006),
Low-temperature solders, Hewlett-Packard Journal,
Article 10.
Morando, C., Fornaro, O., Garbellini, O. and Palacio, H.
(2012), Microstructure evolution during the aging at
elevated temperature of Sn-Ag_cu solder alloys, Procedia
Materials Science, Vol. 1, pp. 80-86.
Nai, S.M.L., Wei, J. and Gupta, M. (2006), Influence of
ceramic reinforcements on the wettability and mechanical
properties of novel lead-free solder composites, Thin Solid
Films, Vol. 504 Nos 1/2 , pp. 401-404.
Naisbitt, G.K. (2006), Must System Solderability Testing Using
the Wetting Balance and Micro-Wetting Balance Methods,
Solderability Testing, Gen3 Systems Limited, Henkel
(Multicore Solders), Hampshire.
Noor, E.E.M., Sharif, N.M., Yew, C.K., Ariga, T.,
Ismail, A.B. and Hussain, Z. (2010), Wettability and
strength of InBiSn lead-free solder alloy on copper
substrate, Journal of Alloys and Compounds, Vol. 507 No. 1,
pp. 290-296.
Peng, W., Monlevade, E. and Marques, M.E. (2007), Effect
of thermal aging on the interfacial structure of SnAgCu
solder joints on Cu, Microelectronics Reliability, Vol. 47
No. 12, pp. 2161-2168.
Prabhu, K.N., Deshapande, P. and Satyanarayan (2012),
Effect of cooling rate during solidification of Sn-Zn
lead-free solder alloy on its microstructure, tensile strength
and ductile-brittle transition temperature, Materials Science
and Engineering A, Vol. 533, pp. 64-70.
Roubaud, P., Ng, G., Henshall, G., Bulwith, R., Herber, R.,
Prasad, S., Carson, F., Kamath, S. and Garcia, A. (2001),
Impact of intermetallic growth on the mechanical strength
of Pb-Free Sn-Ag BGA assemblies, paper presented at
APEX 2001, 16-18 January, San Diego, CA.
Satyanarayan and Prabhu, K.N. (2011), Reactive wetting,
evolution of interfacial and bulk IMCs and their effect on
mechanical properties of eutectic SnCu solder alloy,
Advances in Colloid and Interface Science, Vol. 166 Nos 1/2,
pp. 87-118.
Shalaby, R.M. (2010), Effect of rapid solidification on
mechanical properties of a lead free Sn3.5Ag solder,
Journal of Alloys and Compounds, Vol. 505 No. 1,
pp. 113-117.
Shang, J., Flury, M., Harsh, J.B. and Zollars, R.L. (2008),
Comparison of different methods to measure contact
angles of soil colloids, Journal of Colloid and Interface
Science, Vol. 328 No. 2, pp. 299-307.
Shen, J. and Chan, Y.C. (2009), Research advances in
nano-composite solders, Microelectronics Reliability, Vol. 49
No. 3, pp. 223-234.
160

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

Effect of alloying element and nanoparticle additions

Soldering & Surface Mount Technology

Ervina Efzan Mhd Noor and Amares Singh

Volume 26 Number 3 2014 147161

Sugunama, K. and Dekker, M. (2004), Lead free soldering in


electronics: Science and Technology and Environmental Impact,
Marcel Dekker Inc., University Osaka, Osaka.
Sundelin, J.J., Nurmi, S.T. and Lepisto, T.K. (2008),
Recrystallization behaviour of SnAgCu solder joints,
Materials Science and Engineering A, Vol. 474 No. 2,
pp. 201-207.
Tay, S.L., Haseeb, A.S.M.A., Johan, M.R., Munroe, P.R. and
Quadir, M.Z. (2013), Influence of Ni nanoparticle on the
morphology
and
growth
interfacial
intermetallic
compounds between Sn-3.8Ag-0.7Cu lead-free solder and
copper substrate, Intermetallics, Vol. 33, pp. 8-15.
Tsao, L.C., Huang, C.H., Chuang, C.H. and Chen, R.S.
(2012), Influence of TiO2 nanoparticles addition on the
microstructural and mechanical properties of Sn0.7Cu
nano-composite solder, Materials Science and Engineering:
A, Vol. 545, pp. 194-200.
Tsao, L.C., Chang, S.Y., Lee, C.I., Sun, W.H. and
Huang, C.H. (2010), Effects of nano-Al2O3 additions on
microstructure
development
and
hardness
of
Sn3.5Ag0.5Cu solder, Materials and Design, Vol. 31
No. 10, pp. 4831-4835.
Tsukamoto, H., Nishimura, T., Suenaga, S. and Nogita, K.
(2010), Shear and tensile impact strength of lead-free
solder ball grid arrays placed on Ni (P)/Au surface-finished
substrates, Materials Science and Engineering B, Vol. 171
Nos 1/3, pp. 162-171.
Wang, L., Yu, D.Q., Zhao, J. and Huang, M.L. (2002),
Improvement of wettability and tensile property in
Sn-Ag-RE lead free solder alloy, Materials Letters, Vol. 56
No. 6, pp. 1039-1042.
Whalley, D.C. (2004), A simplified reflow soldering process
model, Materials Processing Technology, Vol. 150 Nos 1/2,
pp. 1134-1144.
Wiese, S. and Wolter, K.-J. (2004), Microstructure and
creep behaviour of eutectic SnAg and SnAgCu solders,
Microelectronics Reliability, Vol. 44 No. 12, pp. 1923-1931.
Xu, D.X., Lei, Y.P., Xia, Z.D., Guo, F. and Shi, Y.W.
(2008), Experimental wettability study of lead-free solder
on Cu substrates using varying flux and temperature,
Journal of Electronic Materials, Vol. 37 No. 1, pp. 125-133.
Yang, S.C., Chang, C.C., Tsai, M.H. and Kao, C.R. (2010),
Effect of Cu concentration, solder volume, and
temperature on the reaction between SnAgCu solders and
Ni S.C, Journal of Alloys and Compounds, Vol. 499 No. 2,
pp. 149-153.
Yang, Y., Li, Y., Lu, H., Yu, C. and Chen, J. (2013),
Interdiffusion at the interface between Sn-based solders
and Cu substrate, Microelectronics Reliability, Vol. 53 No. 2,
pp. 327-333.
Yoon, J.-W. and Jung, S.-B. (2006), High temperature
reliability and interfacial reaction of eutectic Sn 0.7Cu/Ni
solder joints during isothermal aging, Microelectronics
Reliability, Vol. 46 Nos 5/6, pp. 905-914.

Yoon, J.-W., Noh, B.-I., Kim, B.-K., Shur, C.-C. and


Jung, S.-B. (2009), Wettability and interfacial reaction of
Sn-Ag-Cu/Cu and Sn-Ag-Ni/Cu solder joints, Journals of
Alloys and Compound, Vol. 486 Nos 1/2, pp. 142-147.
Yu, D.Q., Zhao, J. and Wang, L. (2004), Improvement on
the microstructure stability, mechanical and wetting
properties of SnAgCu lead-free solder with the addition
of rare earth elements, Journal of Alloys and Compounds,
Vol. 376 Nos 1/2, pp. 170-175.
Zhang, W., Zhong, Y. and Wang, C. (2012b), Effect of
diamond additions on wettability and distribution of
SnAgCu composite solder, Journal Materials Science and
Technology, Vol. 28 No. 7, pp. 661-665.
Zhang, Q.K., Zhu, Q.S., Zou, H.F. and Zang, Z.F. (2010b),
Fatigue fracture mechanisms of Cu/lead-free solder
interfaces, Materials Science A, Vol. 527 No. 6,
pp. 1367-1376.
Zhang, L., Xue, S.B., Zeng, G., Gao, L.L. and Ye, H.
(2012a), Interface reaction between SnAgCu/SnAgCuCe
solders and Cu substrate subjected to thermal cycling and
isothermal aging, Journal of Alloys and Compounds,
Vol. 510 No. 1, pp. 38-45.
Zhang, L., Xue, S.B., Gao, L.L., Sheng, Z., Ye, H.,
Xiao, Z.X., Zeng, G., Chen, Y. and Yu, S.L. (2010a),
Development of SnZn lead-free solders bearing alloying
elements, Journal of Materials Science: Materials in
Electronics, Vol. 21 No. 1, pp. 1-15.
Zhou, J., Sun, Y. and Xue, F. (2005), Properties of low
melting point SnZnBi solders, Journal of Alloys and
Compounds, Vol. 397 Nos 1/2, pp. 260-264.

Further reading
Chandra Rao, B.S.S., Weng, J., Shen, L., Lee, T.K. and
Zeng, K.Y. (2010), Morphology and mechanical
properties of intermetallic compounds in SnAgCu solder
joints, Microelectronic Engineering, Vol. 87 No. 11,
pp. 2416-2422.
Chen, H., Wang, C., Lie, M. and Tian, D. (2007), Influence
of thermal cycling on the microstructure and shear strength
of Sn3.5Ag0.75Cu and Sn63Pb37 solder joints on Au/Ni
metallization, Journal of Material and Science Technology,
Vol. 23 No. 1, pp. 68-72.
Lin, C.Y., Mohanty, U.S. and Chou, J.H. (2010), High
temperature
synthesis
of
Sn3.5Ag 0.5Zn
alloy
nanoparticles by chemical reduction method, Journal of
Alloys and Compounds, Vol. 501 No. 2, pp. 204-210.
Zou, C., Gao, Y., Yang, B. and Zhai, Q. (2010), Melting and
solidification properties of the nanoparticles of
Sn3.0Ag0.5Cu
lead-free
solder
alloy,
Materials
Characterization, Vol. 61 No. 4, pp. 474-480.

Corresponding author
Ervina Efzan Mhd Noor can be contacted at: ervina.noor@
mmu.edu.my and ervinaefzan@gmail.com

To purchase reprints of this article please e-mail: reprints@emeraldinsight.com


Or visit our web site for further details: www.emeraldinsight.com/reprints

161

This article has been cited by:

Downloaded by Universiti Malaysia Perlis At 18:31 23 November 2015 (PT)

1. Srivalli Chellvarajoo, M.Z. Abdullah. 2016. Microstructure and mechanical properties of Pb-free Sn3.0Ag0.5Cu solder pastes
added with NiO nanoparticles after reflow soldering process. Materials & Design 90, 499-507. [CrossRef]
2. Bakhtiar Ali. 2015. Advancement in microstructure and mechanical properties of lanthanum-doped tin-silver-copper lead free
solders by optimizing the lanthanum doping concentration. Soldering & Surface Mount Technology 27:2, 69-75. [Abstract] [Full
Text] [PDF]

You might also like