Вы находитесь на странице: 1из 5

Journal of Alloys and Compounds 484 (2009) 777781

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Nanoparticles of SnAgCu lead-free solder alloy with an equivalent melting


temperature of SnPb solder alloy
Yulai Gao a, , Changdong Zou a , Bin Yang a , Qijie Zhai a , Johan Liu b,c ,
Evgeny Zhuravlev d , Christoph Schick d
a

Shanghai Key Laboratory of Modern Metallurgy & Materials Processing, Shanghai University, Yanchang Road 149, Shanghai 200072, PR China
Bionano Systems Laboratory, Department of Microtechnology and Nanoscience, MC2, Chalmers University of Technology, SE - 412 96, Gteborg, Sweden
SMIT Center, School of Mechatronics and Automation, Shanghai University, Yanchang Road 149, Shanghai 200072, PR China
d
Institute of Physics, University of Rostock, Universittsplatz 3, 18051 Rostock, Germany
b
c

a r t i c l e

i n f o

Article history:
Received 6 April 2009
Received in revised form 4 May 2009
Accepted 9 May 2009
Available online 18 May 2009
Keywords:
Lead-free solder
Nanoparticles
Melting temperature depression
Arc technique
SnAgCu

a b s t r a c t
Nanoparticles were prepared with a consumable-electrode direct current arc (CDCA) technique. The
results showed that the calorimetric melting onset temperature of the nanoparticles of SnAgCu solder
alloy could be as low as 179 C, equivalent to that of the traditionally used SnPb eutectic alloy (183 C).
Moreover, the homogenous melting model (HMM) and GibbsThomson equation were employed to
theoretically estimate the size-dependent melting temperature of the as-prepared nanoparticles. The
structure and morphology of the nanoparticles were analyzed with a high-resolution transmission electron microscopy (HRTEM). The CDCA technique showed promising prospect in manufacturing large
amounts of nanoparticles with controlled shape, small size, narrow particle size distribution and nearly
oxide-free composition. This undoubtedly puts forward a novel feasible approach to manufacture high
quality lead-free solders for electronic products.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Electronic packaging is a manufacturing technology used for
electronic products. Packaging provides a medium for electronic
interconnections and mechanical support, and solder alloys provide
the electrical and mechanical connections between the die (chip)
and the bonding pads. The selection of materials for solder alloys
is, therefore, critical and plays an important role in solder joint reliability. The SnPb alloy has been the most widely used solder alloy
system as an interconnection material in the electronic packaging
industry. The Pb-containing alloys are reliable, well tested and quite
inexpensive, and SnPb alloys over the whole composition range
can be used as solders. There are, however, disadvantages with Pbcontaining solders. Apart from the undeniable toxicity of damaging
humans nervous system, Pb is also harmful to the environment
by causing ground water contamination. As a result, Pb-containing
solder alloys are either being banned or phased out from the electronic industry. Legislations, such as the Waste from Electrical and
Electronic Equipment (WEEE) and the Restriction of the use of Hazardous Substances (RoHS) have been implemented in the European

Corresponding author. Tel.: +86 21 56332144.


E-mail address: ylgao@shu.edu.cn (Y. Gao).
0925-8388/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2009.05.042

Union since 2006. Therefore, an urgent necessity exists for more


appropriate substitutes for the SnPb solders.
It is also important to ensure that the physical and mechanical properties of the substitutes are comparable or even superior
to those of SnPb solder. From a manufacturing point of view, the
melting temperature of the solder alloy is a crucial factor that has
to be taken into account in order to achieve high package quality.
A large variety of lead-free solders have already been developed,
mainly involving the SnCu, SnNi and SnAg systems [1]. These
new lead-free solders have been identied as the most promising
alternatives to the eutectic SnPb solder. However, the higher melting temperatures of these alloys, comparing to that of the eutectic
SnPb solder, limit their applications in the electronic industry.
For that matter, new methods have been explored to decrease
the melting temperature of solder alloys. In 1954, Takagi carried
out the rst experimental investigation dealing with the effect of
size on the melting of very small particles [2]. Utilizing electron
microscope, Sambles observed the lowering of the melting point
of the gold particles [3]. In 1976, Buffat also described that the
melting temperature could be depressed by decreasing the particle
size to nanometer scale [4]. Later, nanocalorimetry was employed
to characterize the melting temperature of a small number of
nanoparticles [5,6], and hot stage transmission electron microscopy
was used to observe the melting behavior of isolated nanoparticles

778

Y. Gao et al. / Journal of Alloys and Compounds 484 (2009) 777781

[7]. Wong et al. reported the size-dependent melting temperature


depression of Sn nanoparticles [8] by traditional differential scanning calorimetry (DSC). They also reported size-dependent melting
temperature depression of Sn3.5Ag nanoparticles synthesized by
a chemical reduction method [9]. However, the production rate of
nanoparticles was limited by this method. Hsaio and Duh manufactured and studied nanoparticles of Sn3.5AgxCu (x = 0.2, 0.5,
1.0), synthesized for lead-free solder applications [10]. They did not,
however, observe any obvious melting temperature depression in
their investigation.
When the particle size was further decreased until to innitesimal size, e.g. in which only tens or hundreds of atoms were
contained, an anomalous size dependence of the melting temperature was observed and it was deemed that not only the cluster size
but also structural features governed their correlations [11]. Nevertheless, no such correlation has been observed for the particle size
scale of interest for solder applications. Banhart et al. [12] observed
that the melting temperature of Sn and Pb nanocrystals might be
increased to a value even higher than the corresponding bulk alloy if
exterior shells exist. Embedded in some matrix, some nanoparticles
revealed even opposite changing tendency of the melting temperatures [1316], indicating that the exterior structure of nanoparticles
played a signicant role in their melting behavior.
Premelting, which occurs before the whole melting of the clusters or nanoparticles, is easily produced near crystal defects [17].
Except the large specic surface area, the presence of large interfaces in nanoparticles was another crucial factor to inuence the
melting temperature and should be taken into account [18]. Pusey
and co-workers [19,20] observed that premelting at a lower temperature was produced along grain boundaries before complete
melting of the crystalline colloids. Riegler and Kohler [21] also
thought that the interfacial properties would strongly affect the
phase transition behavior of the small particles due to their large
interfaces, and premelting would occur on surrounding interfaces.
However, little information was available about grain boundary premelting of metals due to their high temperature and the resultant
difculty for simultaneous observations. Fortunately, the recently
developed coherent X-ray diffraction imaging technique might be a
possible approach to study the premelting of nanocrystal interfaces
owing to its diffraction patterns and latest third-generation synchrotron radiation sources to obtain full three-dimensional images
[22].
In contrast to the lower melting temperature (183 C) of SnPb
eutectic solder alloy [23], therefore, the lead-free solder alloys usually cause a raised temperature of about 3040 C during electronic
assembling. And the increased temperature reduces the integrity,
reliability and functionality of printed wiring boards, components
and other attachment [24].
As a result, much effort has been put lately on nding possible
ways to solve some of the issues imposed by lead-free solders. Lin et
al. employed the supernatant process to obtain the Sn3.5Ag solder
alloy nanoparticles, yet no obvious melting temperature depression
was observed possibly due to the aggregation of those nanoparticles
[25]. Therefore, it remains an urgent necessity to conduct an indepth study of feasible processes to lower the melting temperature
of the currently used lead-free solder alloy systems.
2. Experimental procedures
The lead-free master alloy with a nominal composition of Sn3.0Ag0.5Cu (wt.%)
was prepared by an induction melting method. Fig. 1 shows the DSC melting curve
of the master alloy. The melting temperature was 217.8 C, which equals the equilibrium melting temperature. Integration of the peak yields 67 J/g for the heat of
fusion.
Several models described the size dependence of the melting temperature of
nano-sized particles. Generally it is assumed that the melting is initiated by a continuous vibrational lattice instability on the solid surface or in a solid-solid interface
[18]. For nano-scale particles with large surfaces, the instability becomes obvious

Fig. 1. DSC curve of the SnAgCu bulk alloy at the heating rate of 30 C/min.

and usually the nanoparticles will lose stability at lower temperature than their
bulk precursor. The size-dependent melting temperature of nanoparticles could be
estimated using the homogenous melting model (HMM) [4]. This model was established based on the assumption that the solid and liquid particles of the same mass
are in equilibrium with their common vapor. The free energy of a solid particle with
a radius r is modied by an extra term: 2V/r, where , V and r are the surface tension, mole volume and radius of the particle, respectively. In other words, increasing
the free energy of the particles causes depression of the melting temperature. The
size-dependent melting temperature of nanoparticles is then given by the following
equation:
Tm (r) =

bulk
Tm

bulk
+ 273.15)
2(Tm

Hfbulk s r

 2/3 

sv lv

s
l

(1)

where r represents the radius of a spherical particle and Tm (r) is its melting temperbulk
is the melting temperature of the bulk alloy, Hfbulk is
ature (in degree Celsius), Tm
the heat of fusion for the bulk alloy, s and l are the solid and liquid phase densities
of the bulk alloy, and  sv and  lv are the solidvapor and liquidvapor interfacial
energies, respectively.
The melting temperature depression for small crystals can also be evaluated by
the GibbsThomson equation [26]:
bulk

Tm (r) = Tm

bulk
+ 273.15)sl
2(Tm

(2)

Hfbulk s r

where  sl is the solidliquid interfacial energy. For many metals,  sl  sv  lv


[27]. As a matter of fact, if the value of s was deemed the same as that of
l , then  sl =  sv  lv [28]. That is to say, under these conditions the HMM and
GibbsThomson equations could be expressed by the same formula.
The melting temperature and the heat of fusion of the bulk master alloy were
obtained by the DSC measurement shown in Fig. 1. However, some necessary thermodynamic parameters of Sn3.0Ag0.5Cu alloy were unavailable, thus those of the
similar Sn3.5Ag (wt.%) system were used [23]. The data for calculation are listed in
Table 1, and the dependence of melting temperature on nano-sized particles was calculated by the HMM and GibbsThomson equation, as shown in Fig. 2. The solid curve
denotes the melting temperatures calculated by the HMM and GibbsThomson
equations, the upper and lower horizontal dashed lines correspond to the bulk melting temperatures of SnAgCu alloy and SnPb eutectic solder, respectively. According
to Fig. 2, particles of about 10 nm diameter should show a melting temperature
comparable to that of SnPb eutectic solder (183 C).
Nanoparticles were prepared from the master alloy with a recently developed
consumable-electrode direct current arc (CDCA) technique as shown schematically
in Fig. 3. Liquid parafn was chosen as a dielectric protection media. The anode
and cathode, with a diameter of 6 mm, were made from the master alloy by means
of a suction casting technique. The anode and the cathode electrodes were xed
inside the CDCA container and connected to a power supply. When the two electrodes get close enough, an arc discharge could be produced between them, causing a
local melting and breakdown of the solder material into nanoparticles. The obtained
particles were rinsed with chloroform for several times to remove the liquid paraf-

Table 1
Data used in the calculation of size-dependent melting temperatures [23].
bulk
Tm
( C)

Hfbulk (J/g)

s (g/cm3 )

 sv (J/m2 )

 lv (J/m2 )

 sl =  sv
 lv (J/m2 )

217.8

67

7.39

0.51

0.43

0.082

Y. Gao et al. / Journal of Alloys and Compounds 484 (2009) 777781

Fig. 2. Melting temperature dependence of SnAgCu nanoparticles on particle diameter.


n. Furthermore, the rinsed particles were centrifugally separated at 4000 rpm for
45 min to remove the possibly existed large particles.
The phases of the prepared nanoparticles were examined on a Rigaku D/Max2200 X-ray diffractometer (XRD). The morphology and size distribution of the
nanoparticles were revealed by means of JSM-6700F type eld-emission scanning
electron microscopy (FE-SEM). The melting temperatures of the bulk master alloy
and the nanoparticles were measured with PerkinElmer Pyris 1 differential scanning calorimeter (DSC) under the protection of high-purity nitrogen. The detailed
structure and morphology of the nanoparticles was further studied with JEM-2010F
type high-resolution transmission electron microscopy (HRTEM).

3. Results and discussion


Unlike the chemical reduction method applied by Wong and
co-workers [9,10], the CDCA technique is suitable for most conducting materials, and the manufacturing process could be continuous
allowing preparation of lead-free solder nanoparticles at high
production rates. During the preparation of the nanoparticles, discharge and solder alloy breakdown took place inside the dielectric
media, protecting the nanoparticles from serious oxidation. The
XRD pattern of the as-prepared nanoparticles is shown in Fig. 4. The
Sn, Ag3 Sn and Cu6 Sn5 phases were detected, implying that a good
alloying was produced during the particle preparation process. In
addition, no oxide traces were found by the XRD measurement,
indicating that the nanoparticles were not signicantly oxidized.
Arc current is one of the crucial parameters to control nanoparticle formation, especially the particle size and size distribution.
The morphology and sizes of the prepared nanoparticles at 20 A
arc current were obtained from SEM images, as shown in Fig. 5.
The nanoparticles were in spherical shape on the whole, but particle agglomeration was also detected. The size distribution of the

Fig. 3. Schematic of the CDCA setup for nanoparticles preparation (1) cathode, (2)
anode (connected to a high current and low voltage power source), (3) and (4)
bulk alloy electrodes, (5) arc discharge taking place between the electrodes, and
(6) dielectric coolant.

779

Fig. 4. XRD pattern of the nanoparticles prepared by CDCA technique under the
protection of liquid parafn.

nanoparticles was determined by image analysis with Image-Pro


plus software, and the size distribution histogram is shown in the
inset of Fig. 5. It was found that the size of most particles was in the
range 1560 nm with an average of about 30 nm.
The melting temperature of the aforementioned nanoparticles was determined with DSC at a heating rate of 30 C/min. To
reduce the possible experimental error, the DSC measurement was
repeated for three cycles, as shown in Fig. 6. It was found that the
melting peak was shifted to lower temperatures and broadened.
The peak onset temperature was decreased to about 197 C, 20 C
lower than that of the bulk master alloy.
Increasing the arc current to 50 A led to a signicantly narrower
size distribution of the nanoparticles, as shown in Fig. 7. The particle
size was concentrated in the range 2542 nm, again with an average
of about 30 nm.
The nanoparticles prepared with an arc current of 50 A were
measured in the DSC at heating rate 30 C/min for ve cycles,
presented in Fig. 8. The results show that the melting onset temperature is lowered below 190 C and remained nearly unchanged
even after repeated heatingcooling cycles. Due to the large specic
surface area of the nanoparticles, it was hard to avoid oxidation dur-

Fig. 5. SEM image of the nanoparticles of the SnAgCu alloy prepared by CDCA technique at 20 A arc current. The inset shows the size distribution histogram of the
nanoparticles.

780

Y. Gao et al. / Journal of Alloys and Compounds 484 (2009) 777781

Fig. 6. DSC curves for nanoparticles of SnAgCu alloy prepared with current value of
20 A heated at 30 C/min for three cycles. The vertical dashed lines correspond to the
expected melting temperatures according Eq. (1) for the particle diameters given on
top. Curves are displaced vertically for clarity.

ing preparation, storage and the measurement. That is to say, some


oxidation layer was generally present on the particle surface. As
a result, two opposite effects are expected during these repeated
DSC measurements. The rst effect was the further oxidation of the
nanoparticles, which would decrease the size of the metallic core
in the nanoparticles. Oxidation causes a continuous size reduction
of the metal core and therefore a further decreasing melting temperature. In parallel to the decreasing core size the amount of metal
undergoing the phase transition would decrease. This could explain
the slight decrease of the peak area in Fig. 8 with thermal cycling.
The second effect was the possible agglomeration of the nanoparticles during the melting in the former measurement cycles, which
would increase the real particle size. An increasing particle size
would lead to an increasing melting temperature at constant peak
area. According to the present repeated DSC measurements, it
seems that slight oxidation dominated the experiments. But the
results also revealed that the beginning of melting of the nanoparti-

Fig. 7. SEM image of the nanoparticles of SnAgCu alloy prepared by CDCA technique with arc current of 50 A. The inset shows the size distribution histogram of
the nanoparticles.

Fig. 8. DSC curves for nanoparticles of SnAgCu alloy prepared with current value of
50 A heated at 30 C/min for ve cycles. The vertical dashed lines correspond to the
expected melting temperatures according Eq. (1) for the particle diameters given on
top. Curves are displaced vertically for clarity.

cles of the SnAgCu solder alloy could be decreased to the equivalent


value of the traditionally used SnPb eutectic alloy, which was the
aim of the present study.
For the nanoparticles produced by the two different arc currents a broad melting range is observed. It spans from temperatures
comparable with the melting temperature of SnPb eutectic solder
alloy up to the bulk melting temperature of the SnAgCu alloy. In
Fig. 8 even a small peak at the equilibrium melting temperature is
present. According the particle size distribution one would expect a
maximum around 212 C in the DSC curves, which is not present in
Fig. 8. The maximum at about 205 C corresponds to a particle size
of about 20 nm. Also the observed peak onset temperature of about
195 C and 180 C in Figs. 6 and 8, respectively, cannot be explained
with the particle size distribution. Either the metallic core of the
particles is signicantly smaller than the outer size of the particles or there are inner structures present yielding a lower melting
temperature.
HRTEM was employed to detect the detailed structure and morphology of the nanoparticles. Fig. 9 shows the HRTEM image of
one as-prepared nanoparticle prepared under the arc current of
50 A. It was interesting to notice that several particles with different phase structures formed agglomerates. The XRD result shown
in Fig. 4 indicates that the nanoparticles had a structure of Ag3 Sn.
The HRTEM structure analysis also proved the presence of the Ag3 Sn
phase. In particular, the HRTEM image demonstrated that the single
nanoparticle observed by SEM was actually a polyparticle or a polycrystal. That is to say, the seeming single nanoparticle observed
by SEM were possibly agglomerated particles or crystals. Whether
the inner interfaces cause some premelting or the effective particle size is smaller than that seen in the SEM pictures is not known
yet. Both effects may explain the low melting temperatures seen
in Figs. 6 and 8, which are not completely in accordance with the
expected values from the size distributions shown in Figs. 5 and 7.
Another explanation of the low melting temperature could be
a reduction of the metallic core size of the nanoparticle due to
some oxide layer. The curves in Fig. 8 show a change in shape
as well as a reduction of the peak area. Changes in curve shape
could be explained by changes in the internal structure of the particles, e.g. reduction of internal surface area. But such processes

Y. Gao et al. / Journal of Alloys and Compounds 484 (2009) 777781

781

work does offer the direct evidences of the feasibility to decrease


the melting temperature of Sn3.0Ag0.5Cu (wt.%) sloder alloy to
the equivalent temperature of the traditional SnPb eutectic solder
alloy without signicant problems due to oxidation of the nanoparticles at ambient conditions. The CDCA technique, suitable for most
electrical conducting materials, showed high potential in manufacturing large amounts of nanoparticles with controlled shape, small
size, narrow particle size distribution and nearly oxide-free composition. This undoubtedly puts forward a novel feasible approach
to manufacture high-grade electronic products using lead-free solders.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (Grant No.50571057), AM Foundation of STCSM
(Grant No. 08520740500) and Robert Bosch Foundation (Grant No.
32.5.8003.0025.0/MA01). Liu acknowledges the Swedish National
Science Foundation support (Grant No. 621-2007-4660).
References

Fig. 9. HRTEM image of one as-prepared SnAgCu nanoparticle, arc current 50 A.

cannot explain the decrease in peak area. Oxidation seems to be a


more plausible explanation for the observed changes. The smallest particles resulting in the lowest melting temperatures may be
completely oxidized and disappear in the melting curves, as seen
in Fig. 8. At the same time less metal is present and the entire peak
decreases. Assuming an oxide layer of about 5 nm would explain the
shift in the peak maximum in Fig. 8 and seems to be reasonable.
4. Conclusions
Broad endothermic peaks were observed from the DSC measurements (Figs. 6 and 8) of nanoparticles produced from a
Sn3.0Ag0.5Cu (wt.%) master alloy with a consumable-electrode
direct current arc (CDCA) technique. A fraction of the produced
nanoparticles showed a melting temperature similar to that of
eutectic SnPb solder. Most interestingly, they partly survived several melting crystallization cycles, indicating that the oxidation of
the nanoparticles was continued slowly but the liquid particles did
not totally agglomerate even under a series of heating and cooling
measurements. The data indicate a broad size distribution of the
nanoparticles, which melt step by step. Nanoparticles not melted at
the anticipated low processing temperature may cause difculties
in the soldering process. Notwithstanding these limitations, this

[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]

T. Laurila, V. Vuorinen, J.K. Kivilahti, Mater. Sci. Eng. R 49 (2005) 1.


M. Takagi, J. Phys. Soc. Jpn. 9 (1954) 359.
J.R. Sambles, Proc. R. Soc. Lond. A 324 (1971) 339.
P. Buffat, J.-P. Borel, Phys. Rev. A 13 (1976) 2287.
S.L. Lai, J.Y. Guo, V. Petrova, G. Ramanath, L.H. Allen, Phys. Rev. Lett. 77 (1996)
99.
M. Zhang, M.Y. Efremov, F. Schiettekatte, E.A. Olson, A.T. Kwan, S.L. Lai, T.
Wisleder, J.E. Greene, L.H. Allen, Phys. Rev. B 62 (2000) 10548.
W.A. Jesser, R.Z. Shneck, W.W. Gile, Phys. Rev. B 69 (2004) 144121.
H. Jiang, K.-s. Moon, H. Dong, F. Hua, C.P. Wong, Chem. Phys. Lett. 429 (2006)
492.
H. Jiang, K.-s. Moon, F. Hua, C.P. Wong, Chem. Mater. 19 (2007) 4482.
L.-Y. Hsiao, J.-G. Duh, J. Electrochem. Soc. 152 (2005) J105.
A. Aguado, J.M. Lpez, Phys. Rev. Lett. 94 (2005) 233401.
F. Banhart, E. Hernndez, M. Terrones, Phys. Rev. Lett. 90 (2003) 185502.
R. Goswami, K. Chattopadhyay, Acta Mater. 52 (2004) 5503.
R. Goswami, K. Chattopadhyay, P.L. Ryder, Acta Mater. 46 (1998) 4257.
J. Zhong, L.H. Zhang, Z.H. Jin, M.L. Sui, K. Lu, Acta Mater. 49 (2001) 2897.
O.A. Yeshchenko, I.M. Dmitruk, A.A. Alexeenko, A.M. Dmytruk, Phys. Rev. B 75
(2007) 085434.
F.J. Bartis, Nature 276 (1978) 849.
R.W. Cahn, Nature 323 (1986) 668.
P.N. Pusey, Science 309 (2005) 1198.
A.M. Alsayed, M.F. Islam, J. Zhang, P.J. Collings, A.G. Yodh, Science 309 (2005).
H. Riegler, R. Kohler, Nature 3 (2007) 890.
M.A. Pfeifer, J.W. Williams, I.A. Vartanyants, R. Harder, I.K. Robinson, Nature 442
(2006) 63.
M. Abtew, G. Selvaduray, Mater. Sci. Eng. R 27 (2000) 95.
J. Chriastelov, M. Ozvold, J. Alloys Compd. 457 (2008) 323.
C.Y. Lin, J.H. Chou, Y.G. Lee, U.S. Mohanty, J. Alloys Compd. 470 (2009) 328.
J. Sun, S.L. Simon, Thermochim. Acta 463 (2007) 32.
Q.S. Mei, K. Lu, Prog. Mater. Sci. 52 (2007) 1175.
E.A. Olson, M.Y. Efremov, M. Zhang, Z. Zhang, L.H. Allen, J. Appl. Phys. 97 (2005)
034304.

Вам также может понравиться