You are on page 1of 6

Journal of ELECTRONIC MATERIALS, Vol. 37, No.

6, 2008

Regular Issue Paper

DOI: 10.1007/s11664-007-0366-3
2008 TMS

Effect of SiC Nanoparticle Additions on Microstructure


and Microhardness of Sn-Ag-Cu Solder Alloy
PING LIU,1,2 PEI YAO,1,3,4 and JIM LIU2
1.College of Material Science & Engineering, Tianjin University, Tianjin 300072, P.R. China.
2.PDR-Asia, Motorola (China) Electronics Ltd., Tianjin 300457, P.R. China. 3.Institute of
Superconducting and Electronic Materials, University of Wollongong, Wollongong, NSW 2522,
Australia. 4.e-mail: pyao@tju.edu.cn

In this work, lead-free composite solders were produced by mechanically


mixing nominal 20 nm moissanite SiC particles with Sn-3.8Ag-0.7Cu solder
paste. The effects of the amount of SiC addition on the melting behavior,
microstructure, and microhardness of as-solidified composite solders were
systematically investigated. In comparison with solder without the addition of
SiC nanoparticles, the subgrain of b-Sn, the intermetallic compounds (IMCs)
average grain size and distance decreased significantly in the composite solder
matrix. This was possibly ascribed to the strong adsorption effect and high
surface free energy of the SiC nanoparticles. Our results showed that
0.05 wt.% addition of SiC nanoparticles could improve the microhardness by
44% compared with the noncomposite and that the average grain size and
distance changed from 0.5 lm to 0.2 lm and from 0.6 lm to 0.32 lm, respectively. The refined IMCs, acting as a strengthening phase in the solder matrix,
enhanced the microhardness of the composite solders, in good agreement with
the prediction of the classic theory of dispersion strengthening.
Key words: Lead free solder, intermetallic compound, microhardness,
microstructure

INTRODUCTION
Electronic devices such as computers, cellular
phones, and portable products have become thinner
and smaller while their functions are becoming
more complicated. Miniaturization of these electronic devices demands better solder-joint performance. An attractive and potentially available
method of enhancing solder joint is by adding reinforcements to solder alloys,1 to form a composite
solder. The reinforcing particles should have small
size and serve to suppress grain-boundary sliding,
large intermetallic compound formation, and grain
growth, thereby causing the stress in the solder
joints to be distribute uniformly. In this way the
solder joint could provide better reliability with

(Received April 11, 2007; accepted December 4, 2007;


published online February 23, 2008)

874

improved thermal stability of the microstructure


and a lower creep rate.
Guo et al. investigated the creep behavior of
15 vol.% Cu, Ni, Ag microparticle-doped composite
solders.2,3 In comparison with the noncomposite
Sn-3.5Ag solder, they found that the creep resistance was significantly improved for Cu, Ni particlereinforced composite single-lap solder joints at 25C,
65C, and 105C, but that there was no obvious
improvement for the Ag particle-reinforced solder
joint. Research by Yu et al.4 revealed that the
grains of the Sn-Ag-Cu lead-free solder were refined
by the addition of rare-earth elements and that the
precipitated IMCs were finer due to the adsorption
effect of active rare-earth elements. They also found
that the tensile strength of the solder increased as a
result of the addition of rare-earth elements. A new
improved solder was fabricated by incorporating
nanosized, nonreacting, noncoarsening oxide dispersoids of TiO2 into conventional Sn-37Pb solder by

Effect of SiC Nanoparticle Additions on Microstructure and Microhardness of Sn-Ag-Cu Solder Alloy

Mavoori et al.5 They developed a novel coating


approach of particles to favor the homogeneous
distribution of dispersoids, and these TiO2 nanoparticle composite solders were found to have much
greater creep resistance than the control sample of
eutectic Sn-80Au solder. This has great significance for the replacement of conventional Sn-80Au
solder with a high melting point (278C) for creepresistance applications such as optical or optoelectronic packaging. Researchers at IBM6 have
developed composite solders by blending Mo or Ta
powders with Pb-Sn (50:50) solder powders to obtain
a stronger bonding strength with a metal substrate.
Lin et al.7 have also investigated the microstructure
and hardness of composite Sn-Pb solders. Enhanced
hardness of the eutectic Sn-Pb solder was achieved
in their work by adding 2 wt.% of TiO2 nanopowders. Nai et al.810 successfully synthesized SnAgCu-based solder materials with different weight
proportions of multiwalled and single-walled carbon
nanotubes and ZrO2 nanoparticles via the powder
metallurgical technique. All of those composite solders exhibited improved mechanical strength while
maintaining an approximately constant melting
temperature, indicating that existing soldering
process could be used. Shen et al.11 explained the
enhancement of mechanical properties and refinement of Ag3Sn grain size in ZrO2 nanoparticles
composite Sn-3.5Ag solders using the classic theory
of dispersion strengthening and the adsorption
effect of nanoparticles, respectively.
The WagnerLifschitzSlezov equation for
Ostwald ripening:12
d3 t  d30 aDcc0 t;

875

EXPERIMENTAL PROCEDURES
SiC nanoparticle-reinforced composite solders
were prepared via mechanically incorporating
0.01 wt.%, 0.05 wt.%, and 0.2 wt.% of about 20 nm
SiC particles into Sn-3.8Ag-0.7Cu solder paste.
The composite solders are hereafter merely referred
to as SAC-xSiC, where x represents the weight
percentages. The lead-free Sn-3.8Ag-0.7Cu solder
paste, made up of 20 lm to 50 lm solder spheres
and 10 wt.% to 12 wt.% rosin mildly activated
(RMA) flux, was purchased from Henkel Loctite Co.
SiC nanoparticles were purchased from Kiln Nanometer Technology Development Co., Ltd. Figure 1a
shows a transmission electron microscope (TEM)
image of the SiC nanoparticles taken by a Tecnai G2
F20 (FEI Co.) operated at 200 kV. X-ray diffraction
(XRD, D/max 2500 V/PC) analysis was performed
to determine the phase composition of the SiC

(1)

where d(t) is the diameter of the particle at time t,


d0 is the initial size, a is a constant, D is the diffusion coefficient, c is the interfacial energy, and c0
is the solubility, shows that for maximum dimensional stability the dispersed particles should be
chosen to have minimum diffusivity, interfacial
energy, and solubility in the matrix, and that
nanoparticles reduce D within the matrix material
and the interfacial energy between the intermetallics and the matrix. SiC nanoparticles widely
used for reinforcement in metal matrix13,14 composite materials seem to match these requirements
well.
Due to the toxicity of Pb in traditional solders,
solder material has to become lead free. So far the
Sn-Ag-Cu solder family is regarded as the most
promising replacement for Sn-Pb eutectic solder. In
recent years, composite Sn-Ag-Cu lead-free solders
have attracted many researchers.1517 However, the
SiC reinforced lead-free solder has not been
reported. The objectives of the present work are to
investigate the effect of additions of SiC nanoparticle on the melting behavior, microstructure, and
microhardness of lead-free composite Sn-3.8
Ag-0.7Cu (wt.%) solders.

Fig. 1. (a) TEM image of the SiC morphology and size distribution.
(b) X-ray diffraction scan of the SiC nanoparticles.

876

Liu, Yao, and Liu

nanoparticles (Fig. 1b). The fabrication process of


the composite solder material is as following: dispersing SiC nanoparticles in ethanol (analytical
reagent grade) ultrasonically for 10 min, mechanically mixing the nanoparticles and solder paste in a
ceramic crucible for 30 min to promote homogeneous distribution of nanoparticles, and then
baking the composite solder at 80C to evaporate
excessive ethanol. The same procedure was used to
fabricate the noncomposite solder.
After the completion of the fabrication process, a
small amount (about 0.2 g) of the composite solder
paste was melted in a reflow oven with a standard
profile (Fig. 2), as used in the industrial application,
and then solidified into button-shaped solder specimens on an aluminum plate. After that, the specimens were mounted in the cold epoxy and ground
down to a 600-grit size using silicon carbide paper
under water cooling. The ground samples were
mechanically polished with 6 lm, 3 lm, and 1 lm
diamond pastes and 0.02 lm aluminum oxide suspension. For accurate observation of the microstructure of the solder bulk, the specimen was
selectively etched with a solution consisting of
5 vol.% HCl and 95 vol.% ethanol. An XL30 TMP
ESEM scanning electron microscope (FEI Co.) was
employed to examine the surface morphology.
The melting characteristics of the composite solders were investigated using a differential scanning
calorimeter (DSC, Perkin-Elmer Pyris 1). For each
alloy, the sample was scanned twice. The sample
was first scanned from ambient temperature up to
250C at a heating rate of 10C/min, followed by
cooling to ambient temperature naturally, then
scanned again up to 250C at the same heating rate.
The second scanning thermograph was used to
represent the melting behavior of the alloys.
The microhardness was determined by using a
EVERONE MH-6 microhardness tester to identify
the correlation between microstructure and
mechanical properties. Vickers hardness values
were obtained by the following equation:
Hv 1:854

F
d2

(2)

Fig. 2. The reflow profile with peak temperature: 237C; time above
217C: 70 s.

where F (units: kgf) is the applied load and d (units:


mm) is the average length of the indentation diagonals. The chosen indenting load and dwell time
were 10 g and 10 s, respectively. To ensure an
accurate result, ten different indents were performed on the polished surfaces of each composite
sample. The Vickers hardness value of each sample
was then taken as the average of the individual
measurements.
RESULTS AND DISCUSSIONS
Melting Characteristics
Figure 3 shows the DSC curves of the composite
solder specimens doped with different concentrations of SiC nanoparticles. Figure 3a exhibits the
DSC curve of the Sn-3.8Ag-0.7Cu without any
addition. Figure 3bd shows the DSC scans of
Sn-3.8Ag-0.7Cu solders reinforced with 0.01 wt.%,
0.05 wt.%, and 0.2 wt.% of SiC nanoparticles,
respectively. For each case, only one eutectic peak is
observed and the melting temperature of the composites is below that of the un-reinforced alloy,
which agrees well with previous results from other
rigid nano-reinforcements.9,16,18 A decrease in the
melting point of the composite solders is observed
with increasing amount of SiC. In the case of
0.2 wt.% addition of SiC, the well-defined endothermic peak shifts to 218.9C from 219.9C.
Therefore, the resultant composite solders can be
readily adopted using existing reflow conditions.
The reduction in melting point of the solder alloys is
possibly ascribable to an increase in the surface
instability rendered by the addition of SiC nanoparticles with a higher surface free energy. Also, the
size effect of the SiC nanoparticles can significantly
alter the grain boundary/interfacial characteristics
of the solders to induce this change in physical
properties.1921
Microstructure Evaluations
The SEM observation indicated that few defects
were present in the solidified composite solders. Figure 4 shows the microstructure of the
as-solidified Sn-3.8Ag-0.7Cu composite solders. The
EDS results indicated the existence of b-Sn dendrites, an Sn-rich phase, and Ag3Sn and Cu6Sn5
IMCs. The addition of SiC nanoparticles influences
the growth and size of the IMCs in the solder matrix. In comparison with the noncomposite solder,
the IMCs size in the composite solder alloys containing 0.01 wt.% and 0.05 wt.% SiC nanoparticles
is smaller. A similar result has been reported in the
literature.7 However, for the Sn-3.8Ag-0.7Cu composites, the IMCs size did not always decrease with
increasing amount of SiC, for example, the composite solder containing 0.2 wt.% and 0.01 wt.% SiC
nanoparticles have identical IMCs grain size. This
may be attributable to the agglomeration and segregation of SiC nanoparticles in the bulk solder
because the van der Waals forces cause the SiC

Effect of SiC Nanoparticle Additions on Microstructure and Microhardness of Sn-Ag-Cu Solder Alloy

877

Fig. 3. DSC scans for: (a) Sn-3.8Ag-0.7Cu solder, (b) SAC-0.01SiC, (c) SAC-0.05SiC, and (d) SAC-0.2SiC.

Fig. 4. SEM micrographs of microstructure: (a) Sn-3.8Ag-0.7Cu solder, (b) SAC-0.01SiC, (c) SAC-0.05SiC, and (d) SAC-0.2SiC.

nanoparticles to become entangled with each other


as they approach 0.2 wt.%, therefore the IMCs size
will not decrease more. The SEM investigations
revealed that both deep-etched composite specimens
and noncomposite solder also present a similar trend

in the microstructure of the b-Sn phase, as shown in


Fig. 5. The noncomposite solder exhibited a larger
average subgrain size than the composite solders.
In order to accurately determine the influence of
SiC nanoparticles on the morphology of IMCs, the

878

Liu, Yao, and Liu

Fig. 5. Subgrain microstructure of the b-Sn phase in: (a) noncomposite solder and (b) composite solders.

Table I. Comparison of the IMCs average size and the spacing between IMCs in composite solder matrices
(units: lm)
SAC-0.01SiC

SAC-0.05SiC

SAC-0.2SiC

0.45 0.08
0.6 0.1

0.3 0.03
0.38 0.04

0.24 0.02
0.32 0.03

0.32 0.02
0.39 0.04

Microhardness
Hardness is a key parameter and is often used to
evaluate the mechanical properties of metal materials. The hardness value provides an indication
of resistance to deformation, densification, and
cracking. As shown in Fig. 6, the microhardness varied from 16.2 0.4 kg/mm2 to 23.36 1.2 kg/mm2
and increased with increased SiC content, approaching a maximum value with 0.05% SiC addition.
The microhardness enhancement of these composite
solders is 30% to 44% compared with that of
the noncomposite solder. The apparent strengthening effect can be attributed to the reduction
of the spacing between IMCs. This result is

23.36
21.08

21.18

IMCs size and the spacing between the IMCs in the


as-solidified solder alloys were investigated statistically. In this procedure, 20 locations were chosen
at random in the samples and mean values were
calculated (Table I). It was found that the addition
of the SiC nanoparticles to the Sn-3.8Ag-0.7Cu solder reduced the average size of the IMCs and the
spacing between them significantly.
Generally, the plane of the IMCs with the
maximum surface tension grows fastest, while the
amount of adsorption of the surface-active material
in this plane is maximized. According to the theory of
adsorption of surface-active material,22 an increase
in adsorption elements decreases the surface energy
and therefore decreases the growth velocity of IMCs.
For composite solders, the size of the IMCs is larger
than the size of the SiC nanoparticles, thus the
experimental result that the presence of the SiC
nanoparticles refines the IMCs can be explained by
the adsorption of surface-active material (i.e., SiC
nanoparticles) at the IMCs grain boundaries.

M i c r o h a r d n e s s ( Kg /m m )

Average size
Average spacing

SAC

20
16.2

10

0
0

0.01

0.05

0.2

SiC (wt%)

Fig. 6. Bar graph showing the influence of SiC nanoparticle addition


on the microhardness of the composite solders.

consistent with studies on rigid nanoparticles composite solders.11,18


According to the dispersion strengthening
theory,22 IMC particles can enhance the strength of
the solder matrix. The stress acting on the particles
surface can be expressed by piling up of linear
dislocations, as shown in the following equation:
s ns0 ;

(3)

where s is the stress at the particle surface, n is the


number of piled-up dislocations, and s0 is the yield
stress of the solder alloy.
Meanwhile, the number of piled-up dislocation is
n

p1  tLs0
Gb

(4)

Effect of SiC Nanoparticle Additions on Microstructure and Microhardness of Sn-Ag-Cu Solder Alloy

where L is the average spacing between the second


particles, b is the Burgers vector, t is the Poissons
ratio, and G is the shear elastic modulus of the
substrate. Hence,
s

p1  tLs20
:
Gb

So the yield stress of the alloy becomes:


s
Gbs
s0
:
p1  tL

(5)

(6)

If we assume that s is the fracture stress of the


secondary particles and that t is constant, this
means that, as the average spacing, L, of the secondary particles (i.e., the IMCs) decreases, the yield
stress of the alloy increases. Therefore, finely
dispersed secondary particles can enhance the
mechanical properties. For the composite, the
amount of SiC nanoparticles is very small and
the dispersion strengthening effect is mainly contributed by the IMCs. The average spacing between
the IMCs is 0.6 lm in the noncomposite solder.
After the addition of 0.05 wt.% SiC, the IMCs are
obvious refined, with an average spacing of 0.24 lm,
that is L1/L2 = 2.5, leading to s2/s1 = 1.58 according
to Eq. 6. In the experimental results in Fig. 6, the
ratio of microhardness is Hv2/Hv1 = 23.36/
16.2 = 1.44, which is close to the theoretical
prediction from dispersion strengthening theory.
CONCLUSIONS
SiC nanoparticles in concentrations of 0.01 wt.%,
0.05 wt.%, and 0.2 wt.% were used to reinforce the
Sn-3.8Ag-0.7Cu composite lead-free solders produced via a mechanical blending method. The
melting temperature for the composite solders was
found to be lower than that of the noncomposite by
about 1C, indicating that existing soldering process
are still applicable. SEM observation of the microstructure of the composite solders revealed finer
IMCs in the solder matrix comparing to the noncomposite solder. It was also shown that there exist
some subgrains of the b-Sn phase with smaller size.
These two facts are attributed to the adsorption of
SiC nanoparticles with high surface free energy on
the grain surface during solidification. Microhardness was observed to increase with increasing SiC
content and approaches a maximum value of
23.36 kg/mm2 at 0.05 wt.% SiC addition. The
microhardness enhancement of these composite
solders is 30% to 44% relative to the hardness of the
noncomposite solder. This apparent strengthening
effect can be attributed to the reduction of the

879

spacing between the IMCs because the IMCs play


the role of a secondary reinforcing phase. The
enhancement of microhardness shows good correlation with the composites microstructure, and the
experimental results agree well with the theoretical
prediction from dispersion strengthening theory. In
this work, the presence of these nanoparticles
enhances the properties of the Sn-3.8Ag-0.7Cu solder alloy significantly. However, it is not possible to
determine a precise relationship at this time. The
effects for all concentrations are equal within the
limits of statistical significance.
ACKNOWLEDGEMENTS
The authors would like to acknowledge the
finance and project support of Physical and Digital
Realization Research (PDR-Asia) of Motorola, Inc.
The support of Wang Hui, Du Haiyan, and Xue Tao
with the SEM and XRD analyses is greatly appreciated.
REFERENCES
1. Fu. Guo, J. Mater. Sci. 18, 129 (2007).
2. F. Guo, J.P. Lucas, and K.N. Subramanian, J. Mater. Sci.
12, 27 (2001).
3. F. Guo, J.P. Lucas, K.N. Subramanian, and T.R. Bieler,
J. Electron. Mater. 30, 1222 (2001).
4. D.Q. Yu, J. Zhao, and L. Wang, J. Alloy Compd. 376, 170
(2004).
5. H. Mavoori and S. Jin, J. Electron. Mater. 27, 1216 (1998).
6. Composite Solders IBM Technical Disclosure Bull, 29,
1573 (1986).
7. D.C. Lin, G.X. Wang, T.S. Srivatsan, and M. Petraroli,
Mater. Lett. 57, 3193 (2003).
8. S.M.L. Nai, J. Wei, and M. Gupta, Mater. Sci. Eng. A 423,
166 (2006).
9. S.M.L. Nai, J. Wei, and M. Gupta, Solid State Phenom. 111,
59 (2006).
10. K. Mohna Kumar, V. Kripesh, Lu Shen, and A.A.O. Tray,
Thin Solid Film 504, 371 (2006).
11. J. Shen, Y.C. Liu, Y.J. Han, Y.M. Tian, and H.X. Gao, Mater.
Sci. Eng. A 441, 135 (2006).
12. I.M. Lifschitz and V.V. Slezov, Soviet Phys. JETP 35, 331
(1959).
13. R. Hambleton, H. Jones, and W.M. Rainforth, Mater. Sci.
Eng. A 304306, 524 (2001).
14. V.S. Aigbodion and S.B. Hassan, Mater. Sci. Eng. A 447, 355
(2007).
15. K.M. Kumar, V. Kripesh, and A.A.O. Tray, J. Alloys Compd.
(2006), doi:10.1016/j.jallcom.2006.10.123.
16. P. Yao, P. Liu, and J. Liu, J. Alloys Compd. (2007), doi:
10.1016/j.jallcom.2007.08.041.
17. I.E. Anderson, J. Mater. Sci.: Mater. Electron. 18, 55 (2007).
18. K.M. Kumar, V. Kripesh, and A.A.O. Tray, J. Alloys Compd.
(2007), doi:10.1016/j.jallcom.2007.01.045.
19. F.A. Lindemann, Phys. Z. 11, 609 (1910).
20. P.R. Couchman and C.K. Ryan, Philos. Mag. A 37, 327
(1978).
21. J.M. Ziman, Principles of the Theory of Solids (London:
Cambridge University Press, 1972).
22. Z.D. Xia, Z.G. Chen, Y.W. Shi, N. Mu, and N. Sun, J. Electron. Mater. 31, 564 (2002).

You might also like