You are on page 1of 16

Journal of ELECTRONIC MATERIALS, Vol. 46, No.

3, 2017
DOI: 10.1007/s11664-016-5225-7
© 2017 The Minerals, Metals Materials Society

Effects of Solidification Thermal Parameters on Microstructure


and Mechanical Properties of Sn-Bi Solder Alloys

BISMARCK LUIZ SILVA,1 VÍTOR COVRE EVANGELISTA DA SILVA,2


AMAURI GARCIA,3 and JOSÉ EDUARDO SPINELLI 2,4

1.—Department of Materials Engineering, Federal University of Rio Grande do Norte, UFRN,


Natal, RN 59078-970, Brazil. 2.—Department of Materials Engineering, Federal University of São
Carlos, UFSCar, São Carlos, SP 13565-905, Brazil. 3.—Department of Manufacturing and
Materials Engineering, University of Campinas, UNICAMP, Campinas, SP 13083-860, Brazil.
4.—e-mail: spinelli@ufscar.br

Samples extracted along the length of directionally solidified (DS) castings of


three Sn-xBi alloys (x = 34 wt.%Bi, 52 wt.%Bi and 58 wt.%Bi) were first
evaluated metallographically and then subjected to scanning electron micro-
scopy and energy-dispersive x-ray spectroscopy analyses. The characteristic
length scale of both eutectic and dendritic phases forming the microstructure
were determined and correlated with solidification thermal parameters
(growth rate V, and cooling rate T˙). Tensile and Vickers hardness tests were
performed to allow strength and ductility to be discussed as a function of both
microstructure features and alloy solute content. The tertiary dendrite arm
spacings along the length of the DS Sn-52 wt.%Bi alloy casting are shown to be
lower than those obtained for the Sn-34 wt.%Bi alloy casting. The results of
mechanical tests show that, with the decrease in the alloy Bi content, both
tensile strength and hardness are improved. This is shown to be mainly at-
tributed to the higher density of Bi precipitates decorating the Sn-rich den-
drites, which are finer than the equivalent phase developed for the Sn-52 wt.%
Bi alloy. However, the ductility is shown to be significantly improved for
specimens associated with regions of more refined microstructure of the Sn-
52 wt.%Bi alloy DS casting. A microstructure combining much branched
dendrites, fine Bi particles within the β-Sn dendritic matrix and an important
proportion of very fine eutectic formed by alternate Bi-rich and Sn-rich phase,
seems to be conducive to this higher ductility. In this case, the fracture surface
is shown to be more finely broken with presence of dimples for this particular
condition, i.e., characteristic of a ductile fracture mode.

Key words: Sn-Bi alloys, solidification, microstructure, Bi precipitates,


tensile strength, fracture surface

INTRODUCTION render this class of alloys particularly suitable for


applications in the electronics industry, where low-
Sn-Bi solder alloys have recently received special
temperature soldering is required. Under such cir-
attention, since they are seen as potential lead-free
cumstances, control of as-soldered Sn-Bi
solder alternatives to traditional Sn-Pb alloys due to
microstructures becomes necessary, especially if
the following inherent characteristics: low melting
reliable final properties are envisaged.
temperature, good wettability, relative low cost and
Appropriate microstructural arrangements of Sn-
excellent creep resistance.1–5 Such characteristics
Bi solder joints may improve the mechanical re-
sponse. It is well known that the mechanical char-
acteristics of the Sn-rich phase are affected by the
(Received July 15, 2016; accepted December 9, 2016;
published online January 10, 2017) combination of solid solution and further precipita-

1754
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1755
Solder Alloys

tion of Bi particles.5 This is one of the metallurgical evolution of the Sn-Bi eutectic cannot be found in
aspects to be stressed in as-soldered microstruc- the literature. Gigliotti and Graham23 established a
tures of Sn-Bi alloys. The complexity of the dendritic zone of coupled growth in Sn-Bi alloys. Such a zone
network, the sizes of dendritic spacings, the mor- was shown to be located at the Bi-rich side of the
phology and length scale of the eutectic mixture are eutectic, which indicates a typical coupled zone in a
other features to be considered. It is worth noting faceted/non-faceted eutectic system. The
that the role of Bi precipitates on the final microstructures of such areas which nucleated in
mechanical properties, such as ductility and the coupled zone were shown to vary with the Bi
strength of near-eutectic Sn-Bi alloys, remains content. At the Sn-rich end of the zone, the
undetermined. The effect of characteristics of Bi microstructure was shown to be mainly irregular,
precipitates, for instance morphology and size, as a with some areas in which it was complex-regular.
function of both alloy solute content and solidifica- For compositions higher in Bi, the eutectic was
tion thermal parameters deserve deeper investiga- complex-regular, characterized by trifoils of Bi, with
tions. lamellar-like or fishbone branches.
In spite of the aforementioned positive charac- Over recent years, more investigations have fo-
teristics, the poor ductility is still the major problem cused on near-eutectic Sn-Bi solder alloys. Accord-
restraining the essential processing and industrial ing to Lai and Ye,24 non-eutectic Sn-Bi alloys have
applications of Sn-Bi alloys. This seems to be asso- better mechanical properties when compared with
ciated with the peculiar characteristics of the those of a eutectic alloy. Furthermore, these authors
microstructure of these alloys when compared with reported that the eutectic phase has lower
those of other Sn-based alloys. Instead of a Sn-rich mechanical strength when compared with that of
matrix with intermetallic compounds or precipitates the primary -Sn phase. Indeed, mechanical property
distributed in it similar to most Sn-based solders, a parameters of the two phases, calculated by the
large number of brittle Bi particles throughout the Abaqus software, resulted in lower strength for the
Sn-rich matrix characterizes the microstructure of eutectic phase. Lai and Ye24 reported that the ulti-
usual Sn-Bi solders. Sn-Bi eutectic or near-eutectic mate tensile strength (UTS) of Sn-10Bi and Sn-20Bi
alloys compositions with or without additions of alloys were slightly above 80 MPa, being higher
alloying elements are considered interesting possi- than the UTS of a Sn-58Bi alloy. The elongations of
bilities.6–9 both Sn-10Bi and Sn-20Bi alloys were shown to be
A number of studies in the literature10–14 have about 50%, which is also higher than the corre-
reported that the microstructure of the Bi-Sn sponding property of the Sn-58Bi alloy.
eutectic alloy is constituted by Bi-rich and Sn-rich It is known that the cooling rate experienced
lamellar phases, as well as by Bi precipitates within during liquid-to-solid transformation of a solder fil-
the Sn-rich phase. Chen et al.5 affirmed that the let directly affects the microstructure of Sn-based
lamellar structure in Sn-Bi alloys is the most solders, having a significant influence on the
important factor influencing ductility. Two defor- resulting mechanical behavior.25–29 For example,
mation mechanisms are proposed associated with Xiaowu and collaborators14 directionally solidified a
the strain rate sensitivity of these alloys, that is, a Sn-58Bi eutectic alloy using a Bridgman-type
phase boundary-related mechanism changes to a solidification furnace, and reported the evolution of
dislocation glide mechanism with the increase in lamellar microstructures consisting of Sn and Bi
strain rate. The latter phenomenon leads to highly phases. After proposing inter-relationships between
deteriorated ductility. The phase boundary mecha- interphase spacing (λ) and parameters of the pro-
nism is characterized by sliding between the two cess, such as the pulling rate (V), experimental
phases. Thus, the interaction between Sn-rich and equations relating Vickers hardness (HV) and the
Bi-rich phases within the eutectic mixture will interphase spacing were proposed. It was found that
strongly affect the deformation behavior of Sn-Bi the microhardness decreases with the increase in
alloys.5 the microstructural spacing. These types of experi-
Glazer et al.12,15–21 showed that Bi has significant mental correlations remain unknown when tran-
solid solubility (21.0 wt.%Bi) in Sn at the eutectic sient heat flow conditions during solidification of
temperature, thus allowing precipitation of Bi in the Sn-Bi eutectic-rich solder alloys are considered.
Sn-rich phase (solid-state precipitation) at lower Gusakova et al.30 compared the microstructural
temperatures. The eutectic microstructure that re- morphologies associated with the solidification of a
sults from the melt after slow cooling conditions has Sn-58 wt.%Bi eutectic alloy under slow (2 K/s) and
been reported as a quasi-regular microstructure, fast (105 K/s) cooling conditions. A morphological
whereas degenerate material at the boundaries of transition from lamellar to a microcrystalline pat-
the eutectic grains are reported for moderate cooling tern was reported to occur with an increase in
rates.15 Faster cooling rates are associated with cooling rate.
even more complex eutectics, which are the so- Since results on the microstructures of eutectic-
called complex-regular eutectics.22 Systematic rich Sn-Bi solder alloys and their dependences on
studies emphasizing the effects of cooling rate (T˙) solidification cooling rate and growth rate are scarce
and alloy solute content on the morphology and in the literature, this investigation has focused on
1756 Silva, da Silva, Garcia, and Spinelli

the effects of alloy Bi content (from 34 wt.% to 58 wt. pan), was used to acquire the images. The primary,
%) on microstructural features emphasizing den- tertiary (λ1, λ3) and secondary dendrite arm spac-
dritic and eutectic arrangements, which include ings (λ2) were measured on transverse and longitu-
dendritic and eutectic spacings and morphological dinal sections of the DS castings, respectively. The
characteristics of the Bi-rich phase. The triangle method was employed to measure λ1,
microstructural evolutions of Sn-Bi alloys solidified whereas λ2 and λ3 were measured by the intercept
under a transient heat flow regime will be investi- method, as reported by Gündüz and Çadirli34 (see
gated, using a water-cooled directional solidification Fig. 3a and b). The eutectic spacings (λc = λcoarse and
system. A comprehensive characterization through λf = λfine) were measured on the transverse sections
scanning electron microscopy (SEM) and optical by the intercept method, as shown in Fig. 3c. At
microscopy is envisaged with a view to permitting least 40 measurements were carried out for each
experimental growth laws relating microstructural selected position along the length of the DS cast-
spacings and solidification thermal parameters to be ings.
derived. In addition, hardness and tensile proper- Furthermore, detailed microstructural charac-
ties of Sn-34 wt.%Bi, 52 wt.%Bi and 58 wt.%Bi al- terization was performed using a scanning electron
loys will be investigated with the intention of microscope (SEM–EDS) FEI (Inspect S50L). Cross-
establishing correlations between ductility/strength section samples of the three Sn-Bi alloys were
values and typical length scales of the phases examined by SEM and elemental mapping was
forming the microstructure along the length of the performed to determine the relative distribution of
directionally solidified (DS) Sn-Bi alloys castings. the elements. The fracture surfaces of the tensile
samples were observed by a field emission gun–
EXPERIMENTAL PROCEDURE SEM–EDS Philips (XL30 FEG) under secondary
electron mode.
In order to promote vertical directional solidifi-
Transverse specimens for tensile tests were ex-
cation, an apparatus designed in such a way that
tracted from different positions along the length of
heat is directionally extracted by the bottom of the
the DS castings. These specimens were prepared
casting (see Fig. 1) was used. The transient direc-
according to specifications of the ASTM Standard E
tional solidification experiment results in a range of
8 M/04 and tested at a strain rate of about
as-solidified Sn-Bi microstructures, obtained under
3 × 103 s−1. Microhardness tests were performed on
significantly different cooling rates in a single
transverse sections of the samples of each alloy by
casting.31–33 The solidification setup used in the
using a test load of 500 g and a dwell time of 15 s. A
experiments allows heat to be extracted unidirec-
Shimadzu HMV-G20 model hardness tester was
tionally through a water-cooled bottom made of low
used. The adopted Vickers microhardness was the
carbon steel (SAE 1020), promoting vertical upward
average of at least 20 measurements on each sam-
directional solidification. A stainless steel split mold
ple, which were extracted from the following posi-
was used having an internal diameter of 60 mm, a
tions from the bottom of the casting: 5 mm, 10 mm,
height of 157 mm and a wall thickness of 5 mm. The
15 mm, 20 mm, 25 mm, 30 mm, 40 mm, 50 mm,
lateral inner mold surface was covered with a layer
60 mm, 70 mm, and 80 mm.
of alumina-silica insulating paste to minimize radial
heat losses. The bottom part of the mold was closed
RESULTS AND DISCUSSION
with a thin (3-mm-thick) carbon steel sheet. The
upward solidification experiments were carried out Directional solidification experiments under
with Sn-34 wt.%Bi, Sn-52 wt.%Bi and Sn-58 wt.%Bi transient heat flow regime, using a water-cooled
solder alloys, which are indicated in the Sn-Bi phase solidification setup, were run in order to acquire the
diagram of Fig. 2. Continuous temperature mea- evolutions of temperature versus time along the
surements in the casting were monitored during length of the Sn-Bi castings, as shown in Fig. 4. The
solidification via the output of a bank of fine type J experimental cooling curves refer to thermocouples
thermocouples sheathed in 1.5-mm outside diame- located at specific distances (P) from the cooled
ter stainless steel tubes, and positioned at different bottom of the castings. Further, such experimental
positions from the heat-extracting surface at the cooling curves and the liquidus (TL) and eutectic
bottom of the DS castings. All thermocouples were (TE) temperatures of the alloys of 172.2°C, 150.0°C,
connected by coaxial cables to a data logger inter- and 136.5°C for the Sn-34 wt.%Bi, Sn-52 wt.%Bi
faced with a computer and the temperature data and Sn-58 wt.%Bi alloys, respectively, have been
were acquired automatically. used in order to determine the solidification thermal
Selected transverse (perpendicular to the growth parameters: growth rate (V), cooling rate (T˙) and
direction) and longitudinal samples of the DS alloys thermal gradient (G = T˙/V).
castings were polished (solution of alumina 1 µm The cooling curves attained during transient
and water) and etched with a solution of 2 mL HCl, solidification allowed plots of position (P) versus
10 mL FeCl3 and 100 mL H2O applied during 5–15 s time (t), to be generated for all Sn-Bi chemistries
to reveal the microstructures. An optical image examined. The mentioned time (t) corresponds to
processing system, Olympus GX51 (Olympus, Ja- the instant that either the liquidus isotherms of the
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1757
Solder Alloys

Mold

Computer and
data acquision
soware
Temperature
controller of the
Direconal
Solidificaon furnace

Thermocouples
connected in a
data logger
Water flow
meter

Fig. 1. Vertical upward directional solidification casting assembly and its devices.

Sn-34 wt.%Bi and Sn-52 wt.%Bi alloys, or the The experimental evolution of cooling and growth
eutectic front of the Sn-58 wt.%Bi alloy passed by rates show that these thermal parameters decrease
each thermocouple placed along the length of the DS with progress of solidification. Microstructural
castings. parameters (dendritic and eutectic spacings) are
A numerical technique, based on the least squares expected to be reversely affected by this decrease, as
method, was used to fit mathematical power func- it will be seen below.
tions of the form P(t) = atb (a; b are constants) to Typical cooling rates during reflow procedures for
these experimental plots. The derivative of these solders in industrial practice remain in the range
functions with respect to time gave values for the of 3.0–10.0 °C/s.35 Directional solidification (DS)
growth rate (V). Moreover, the data acquisition experiments as those performed in the present re-
system permitted accurate determination of the search work allow the same order of magnitude of
slope of the experimental cooling curves. Hence, the cooling rates to be obtained. Final microstructure is
cooling rate was determined along the length of the a direct consequence of this solidification thermal
castings, by considering the thermal data recorded parameter. Hence, DS results have a correspon-
immediately after the passage of either the liquidus dence with the standard procedures that are used in
isotherm (hypoeutectic compositions) or the eutectic industry.
front (eutectic alloy) by each thermocouple. Representative transverse microstructures of the
The experimental points and tendencies of P × t, DS Sn-Bi alloys castings are shown in Fig. 6. Ar-
V × P, T˙ × P and G × P can be seen in Fig. 5. As the rows within the images highlight some of their
experimental evolutions for the three Sn-Bi alloys elements. The solidification cooling rates associated
were inserted jointly in each graph, it is possible to with the microstructures shown in Fig. 6 are 1.0°C/
carry out a direct comparison between them. All s and 0.3°C/s. Tertiary dendritic branches are
experimental scatters were fitted to power func- clearly seen for the hypoeutectic Sn-34 wt.%Bi and
tions, as indicated in Fig. 5. It can be observed that 52 wt.%Bi alloys (see Fig. 6a and b), while large Bi
the experimental V and T˙ values for the eutectic Sn- particles and “fishbone” eutectic can be observed for
58 wt.%Bi solder (dotted lines) are lower than those the eutectic Sn-58 wt.%Bi specimens, as shown in
found for the other Sn-Bi alloys. A single fit function Fig. 6c. The fishbone-like eutectic is observed in the
has been adopted for the experimental evolutions of DS Sn-Bi specimens as an isolated structure from
thermal gradient during progress of directional the lamellar eutectic. Similar microstructural as-
solidification of the Sn-34 wt.%Bi, Sn-52 wt.%Bi and pects have been reported by Gigliotti et al.23 who
Sn-58 wt.%Bi solder alloys. demonstrated the presence of trifoils of Bi, with
1758 Silva, da Silva, Garcia, and Spinelli

lamellar-like or fishbone branches for alloys of


higher Bi content.
Because the typical non-faceted/faceted growth of
the Sn-Bi eutectic phase, the coupled zone in this
34wt%Bi binary system has been reported to exhibit a skewed
52wt%Bi
shape,23 tending toward the Bi-side of the equilib-
58wt%Bi rium phase diagram. It seems that the primary Bi
phase may nucleate at given undercooling levels or
at given growth velocities (see Fig. 6c), and grow
combined with the eutectic phase.
By comparing the microstructures associated
with the two cooling rate levels, as exhibited in
Fig. 6, it can be seen that the length scale of the
microstructure increases with the decrease in cool-
ing rate, regardless the alloy composition examined.
A complex dendritic array has been identified for
the Sn-34 wt.%Bi and 52 wt.%Bi alloys with the
presence not only of primary dendrite stalks but
also secondary and tertiary dendritic branches in all
cases examined. As expected, the proportion of
eutectic increases with increasing Bi content. The
Fig. 2. Sn-Bi phase diagram highlighting the Sn-Bi alloys examined
in the present investigation.
eutectic mixture is formed by rhombohedral Bi-rich
and body-centered tetragonal Sn-rich phases. No

Fig. 3. Schematic representation of (a, c) transverse and (b) longitudinal sections of methods used to measure dendritic and interphase
spacings: intercept method for λc, λf, λ2 and λ3 and triangle method for λ1. L is the length of the line and n is the number of intercepted branches.
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1759
Solder Alloys

(a) 220 Bi and Sn phases can be seen characterized by light


200 Sn-34wt.%Bi Positions from metal/mold interface:
4mm 9mm 14mm
gray and dark gray regions, respectively. The
19mm 42mm microstructures become finer for higher cooling rates.
180
A large number of fine Bi particles have been formed
160 TL = 172.2°C during cooling, which can be noted particularly in the
Temperature (°C)

140 Sn-34 wt.%Bi alloy specimens through the inset SEM


images. It can be inferred that Bi atoms were firstly
120
dissolved in the Sn phase and secondly precipitated
100 during the cooling stage after solidification. The
80 microstructural spacings between these Bi precipi-
60
tates have been determined for the specimens asso-
ciated with cooling rates of 12°C/s, 1°C/s, and 0.3°C/s,
40 that is, 0.94 μm ± 0.32 μm, 1.45 μm ± 0.30 μm, and
20 2.01 μm ± 0.71 μm, respectively, for the Sn-34 wt.%Bi
0 100 200 300 400 500 600 700 800 alloy. Likewise, λ values of 1.91 μm ± 0.45 μm,
Time (s) 3.42 μm ± 0.40 μm and 2.46 μm ± 0.44 μm were
determined for the Sn-52 wt.%Bi alloy. It can be in-
220
(b) Sn-52wt%Bi Positions from metal/mold interface:
ferred that a higher density of Bi precipitates may
200 5mm 10mm 15mm characterize the Sn-rich dendritic areas in the
59mm 75mm
180 microstructure of the Sn-34 wt.%Bi alloy. Addition-
160 TL = 150°C
ally, the cooling rate affects the distribution of these
Bi particles, since the microstructural spacing in-
Temperature (°C)

140
creased with decrease in cooling rate. These particles
120 have either spherical or ellipsoidal morphologies, as
100 shown in the inset SEM images of Fig. 7.
The Sn-58 wt.%Bi alloy has been reported to have
80
a complex regular eutectic microstructure, in which
60 two types of regions can be observed: zones of a
40 regular repeating pattern and other zones of ran-
20
dom orientation.36 It is worth noting that the two
0 100 200 300 400 500 600 700 800 phases forming the eutectic mixture in the present
Time (s) study are arranged as alternated not-flat plates, as
can be seen in Fig. 7.
(c) 220
Figures 7 and 8 show that the Sn-Bi eutectic-rich
Sn-58wt.%Bi
alloys have two eutectic characteristic sizes: fine
200 Positions from metal/mold interface: (black squares in Fig. 7) and coarse. For Sn-Bi al-
4mm 9mm 14mm
180 19mm 42mm 55mm loys, the transition from lamellar (quasi-regular) to
160 a more complex-regular morphology depends
strongly on the solidification thermal parameters, i.
Temperature (°C)

TE = 136.5°C
140
e., more complex structures are associated with
120 higher cooling rates. It seems that fine eutectic
100 features can be established in certain sites, followed
80
by coarser structures in the surrounding areas, as a
consequence of local thermal instabilities.
60
Fishbone-like eutectic cells and trifoils of Bi can
40 be seen in Fig. 8. These structures have also been
20 reported by Mokhtari and Nishikawa.37 They
0 100 200 300 400 500 600 700 800 stressed that they are formed due to a shift toward
Time (s) the Bi-rich side of the coupled zone diagram, as
established for the Sn-Bi alloy system.23
Fig. 4. Temperature profiles registered during unsteady-state
directional solidification of (a) Sn-34 wt.%Bi, (b) Sn-52 wt.%Bi and (c) Furthermore, Bi particles within the -Sn den-
Sn-58 wt.%Bi solders. dritic matrix are indicated in Fig. 7 by yellow ar-
rows. The morphology of the Bi-rich phase (light
gray) in the eutectic regions remains mainly defined
by the paired growth of trailing and leading phases.
dendrites have been found from the bottom to the As a result, this phase assumes an elongated mor-
top of the DS Sn-58 wt.%Bi (eutectic) alloy casting. phology, as shown in Fig. 7. On the other hand, the
Typical SEM microstructures of the two hypoeu- presence of Bi particles in the -Sn phase is a con-
tectic Sn-Bi alloys are shown in Fig. 7 with a view to sequence of a process of precipitation during cooling
permitting details on the formation of the lamellar after the completion of solidification. In this case,
eutectic structure to be examined. Alternated layers of nucleation and growth are affected by a combination
1760 Silva, da Silva, Garcia, and Spinelli

(a) 90 (b) 1,6

Sn-34wt%Bi Sn-34wt%Bi
80 1,4 Sn-52wt%Bi
Sn-52wt%Bi
0.56 -0.8
P= 2.9 tL VL = 3.7 P
70
1,2
Sn-58wt%Bi

Growth rate, V (mm/s)


Sn-58wt%Bi
60 P= 0.56 tE
0.7
-0.41
VE = 0.3 P
Position (mm)

1,0
50
0,8
40
0,6
30

20 0,4

10 0,2

0
0,0
0 50 100 150 200 250 300 350 400 450 500 550
0 10 20 30 40 50 60 70 80 90
Time corresponding to eutectic/liquidus passage (s) Position (mm)

18
(c)
16 Sn-34wt%Bi
Sn-52wt%Bi
-1.96
14 TL= 345 (P)
Cooling rate, T (°C/s)

12 Sn-58wt%Bi
-1.96
T= 345 (P)
10

0
0 10 20 30 40 50 60 70 80 90
Position (mm)

18
(d)
16 Sn-34wt%Bi
Sn-52wt%Bi
Sn-58wt%Bi
14
Thermal Gradient, G (°C/mm)

-1.04
G = 63 (P)
12

10

0 10 20 30 40 50 60 70 80 90
Position (mm)
Fig. 5. Experimental plots obtained for the Sn-Bi alloys corresponding to: (a) position versus time at which the liquidus/eutectic front reaches
each thermocouple, (b) growth rate, (c) cooling rate and (d) thermal gradient as a function of position.
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1761
Solder Alloys

Fig. 6. Representative optical microstructures and details on the phases and their morphologies considering two distinct levels of cooling rate
(left and right images) for (a) Sn-34 wt.%Bi, (b) Sn-52 wt.%Bi and (c) Sn-58 wt.%Bi solder alloys.

of Bi solid solubility along the Sn-rich phase, which The eutectic mixture is the main microstructural
is affected by the cooling rate imposed under non- component forming the as soldered Sn-Bi joints,
equilibrium solidification and further precipita- taking into account alloy compositions potentially
tion/growth of Bi particles. able to replace Sn-Pb alloys, i.e., eutectic-rich Sn-Bi
Figure 9 shows the elemental SEM–EDS maps for alloys. SEM microstructures as those showed in
the Sn-52 wt.%Bi and 58 wt.%Bi alloys specimens Figs. 7 and 8 were examined through an image
corresponding to a cooling rate of 0.2°C/s. Sn con- analyzer, which allowed the experimental determi-
trast (in green) is higher in Sn-rich dendrites and in nation of the eutectic spacing along the length of the
the eutectic mixture. Bi (in red) also appears con- DS castings. Hence, the average λfine and λcoarse
centrated in the eutectic mixture, but with less values with their standard deviations are shown in
intensity within the dendrite of the Sn-52 wt.%Bi Fig. 10 for all Sn-Bi alloys examined. The values,
alloy. A coarse Bi-rich particle can be seen in the Sn- measured from the lamellar eutectic structure, have
58 wt.%Bi alloy specimen. A very low fraction of Sn been plotted against the growth rate (V). Consider-
spots (in green) can be observed inside this particle. ing both fine and coarse eutectic spacings, the
1762 Silva, da Silva, Garcia, and Spinelli

Fig. 7. SEM microstructures of (a) Sn-34 wt.%Bi and (b) Sn-52 wt.%Bi alloys considering cooling rates of 12˚C/s, 1˚C/s and 0.3˚C/s (top to
bottom, respectively). The insets emphasize the distribution of Bi precipitates within the dendritic matrix (Sn-rich phase) of the Sn-34 wt.%Bi
alloy. P is the position from the cooled surface of the DS casting.

experimental evolutions can be represented by the alloy is lower than those obtained for the other two
−1/2 exponent proposed by Jackson and Hunt.38 Sn-Bi alloys.
The experimental fittings encompass the follow- Additional experimental data from Ref. 14 have
ing λ ranges during unsteady-state growth of the been included in Fig. 10b to encompass lower values
eutectic mixture: from 1.7 μm to 3.9 μm (coarse) and of growth rate (V) in order check the range of
from 0.8 μm to 1.3 μm (fine) for the Sn-34 wt.%Bi applicability of the experimental power law derived
alloy; from 0.9 μm to 3.1 μm (coarse) and from in the present study for the Sn-58 wt.%Bi alloy. The
0.8 μm to 1.8 μm (fine) for the Sn-52 wt.%Bi alloy, experimental law, λfine = 0.40 × (V)−1/2, is able to
and from 2.2 μm to 3.1 μm (coarse) and from 1.0 μm roughly represent also the inserted points.
to 1.7 μm (fine) for the Sn-58 wt.%Bi alloy. Overall, Figures 11 shows the evolutions of primary (λ1),
if a single growth rate is considered, for example, tertiary (λ3) and secondary (λ2) dendritic arm spac-
0.2 mm/s, the eutectic spacing of the Sn-58 wt.%Bi ings as a function of both cooling rate (T˙) and
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1763
Solder Alloys

Fig. 8. Microstructures features observed in the Sn-58 wt.%Bi alloy. The presence of different sizes of eutectic arrangements (coarse and fine) is
emphasized for two different positions from the cooled surface of the DS casting: 15 and 40 mm.

growth rate (V) for the Sn-34 wt.%Bi and 52 wt.%Bi approaches for modeling dendritic and cellular
solder alloys. The points in the graphs represent the growths, uncertainties associated with different
average experimental microstructural spacing metallic systems and experimental growth condi-
along with its standard deviation. Linear relation- tions can arise.
ships are shown fitted to the points with a view to Although the primary dendritic spacing, λ1, of the
representing experimental power laws of dendritic Sn-34 wt.%Bi alloy is smaller than that determined
growth. for the Sn-52 wt.%Bi alloy, considering any cooling
The power function exponents −1/4 and −1/2 rate examined, λ3 of the Sn-34 wt.%Bi alloy is sig-
characterize the primary/tertiary and secondary nificantly higher when compared with the experi-
dendrite arm spacings variations with cooling rate mental values obtained for the Sn-52 wt.%Bi alloy.
and growth rate, respectively, for the Sn-52 wt.%Bi In spite of the decrease in λ2 with the increase in
alloy. It is worth noting that the Sn-52 wt.%Bi alloy growth rate, it seems that the magnitude of the
has a near-eutectic composition with 88% of eutectic secondary spacing is not significantly affected by
fraction according to Scheil’s equation.39 In this the increase in the alloy Bi content from 34 wt.%Bi
case, due to this high proportion of eutectic mixture, to 52 wt.%Bi.
the classic Jackson and Hunt growth law for Figure 12 depicts the experimental dependences
eutectics37 is expected to apply to this alloy. Thus, λ1 of Vickers hardness on the inverse square root of the
and λ3 may be related to V−1/2. The cooling rate (T˙) eutectic spacing, λfine. The mechanical strength will
was shown in the literature40,41 to be given for be significantly affected not only by dendritic spac-
transient solidification conditions by the following ings of low and high order but also by the eutectic
expression: constant × V2. If such expression is in- mixture and its length scale. Since the Sn-Bi alloys
serted into that proposed by Jackson and Hunt, λ1 examined in the present research are considered
may be related to the cooling rate according to a eutectic-rich compositions, the length scale of the
power function exponent of −1/4. This is in agree- eutectic mixture was chosen in order to parame-
ment with the experimental growth law derived in terize the experimental inter-relationships of hard-
the present study for the Sn-52 wt.%Bi solder alloy. ness, ultimate tensile strength (σu), yield strength
In the case of the Sn-34 wt.%Bi alloy, −0.55 and (σy) and elongation (δ) versus microstructural
−1.1 power laws represent the evolutions of λ1/λ3 spacings.
and λ2 with the cooling rate and the growth rate, A decrease in hardness with the increase in the
respectively. The −0.55 exponent is recognized as length scale of the microstructure can be observed
the most appropriate based on a number of experi- for the hypoeutectic Sn-34 wt.%Bi and 52 wt.%Bi
mental evidences in recent studies considering the alloys. However, an opposite experimental tendency
growth of primary dendrite trunks under transient is shown to occur for the eutectic Sn-58 wt.%Bi al-
solidification conditions.42,43 This kind of experi- loy. This indicates that hardness is negatively af-
mental law fits well to the experimental scatters of fected by the growth of trifoils and faceted particles
λ1 and λ3 obtained for the Sn-34 wt.%Bi alloy. It is of Bi, with fishbone branches combined with the
worth noting that the theoretical model of Bouchard eutectic mixture.
and Kirkaldy44 for growth of secondary dendritic Typical stress–strain curves are plotted consid-
arms predicts a −2/3 exponent for λ2 × V relation- ering tests performed with specimens extracted at
ships. Nevertheless, such a theoretical exponent P = 6 mm (black line) and P = 90 mm (gray line)
cannot represent the experimental scatters of both from the bottom of the DS castings (Fig. 13). Com-
alloys examined, considering λ2 as a function of paring the curves obtained for the Sn-34 wt.%Bi and
growth rate. It is essential to be aware that, despite the Sn-58 wt.%Bi alloys, it can be seen that, with
the robustness of the experimental and theoretical the increase in the alloy Bi content, both ultimate
1764 Silva, da Silva, Garcia, and Spinelli

Fig. 9. Elemental SEM–EDS maps obtained along the transverse specimens corresponding to a cooling rate of 0.2˚C/s for the DS: (a) Sn-52 wt.
%Bi and (b) Sn-58 wt.%Bi alloys castings.

tensile strength and elongation-to-fracture de- the eutectic composition does not reveal the pres-
crease. For the alloy with higher Bi content, the ence of Sn-rich areas.
necking stage seems to be shortened, which means The stress–strain curves obtained for the differ-
that the toughness has deteriorated. Such differ- ent positions in the Sn-52 wt.%Bi alloy DS casting
ences may be explained by the fact that both reveal an evident contrast between them, which can
strength and plasticity of the primary Sn-rich phase be explained with basis on microstructure features.
are recognized as being better than those of the In this case, a brittle behavior is associated with
eutectic phase.24 The Sn-34 wt.%Bi alloy has been P = 90 mm, with the corresponding stress–strain
characterized by a significant fraction of Sn-rich curves exhibiting low ductility. The coarse eutectic
dendrites (see Fig. 6) whilst the microstructure of lamellar structure may be the responsible for such
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1765
Solder Alloys

(a) function of the fine eutectic spacing along the length


of the DS castings. Most of the experimental scatters
λcoarse (μm)

have been fitted by Hall–Petch type correlations.


10
1 Considering the Sn-34 wt.%Bi and 52 wt.%Bi al-
loys, σu, σy and δ increase with decreasing λfine. The
tensile strengths, represented by σu and σy, of the
Coarse eutectic spacing,

Sn-34 wt.%Bi alloy casting are higher than the


corresponding values of the Sn-58 wt.%Bi alloy.
Sn-34wt%Bi
This is due to the combination of a higher proportion
0
λcoarse = 1.45 (V)
-1/2 2
- R =0.88 of areas of Sn-rich phase and a higher density of Bi
10
Sn-52wt%Bi precipitates inside those areas in the Sn-34 wt.%Bi
λcoarse = 1.1 (V)
-1/2 2
- R =0.93 alloy, as demonstrated by the smaller spacings be-
Sn-58wt%Bi tween the precipitates of this alloy when compared
-1/2 2
λcoarse = 0.75 (V) - R =0.81
with those determined for the Sn-52 wt.%Bi alloy. A
-2 -1 0
decrease in the eutectic spacing means a reduction
10 10 10
in size and more homogeneous distribution of the
Growth rate, V (mm/s) Bi-rich and Sn-rich phases throughout the
microstructure and this seems to enhance the
(b) mechanical strength of the mentioned alloys. In
Sn-52wt%Bi contrast, the values of elongation-to-fracture con-
sidering 1/λ1/2
fine > 0.9 (see Fig. 14c) are higher for
-1/2 2
λfine = 0.67 (V)
λfine (μm)

- R =0.79

Sn-34wt%Bi the Sn-52 wt.%Bi alloy when compared with those


Sn-58wt%Bi
1
10 λfine = 0.40 (V)
-1/2 2
- R =0.85
found for the Sn-34 wt.%Bi alloy.
As mentioned before, the best balance of strength
Fine eutectic spacing,

and ductility is that obtained with 1/λ1/2 fine = 1.23,


which represents samples extracted from the posi-
tion that is closest to the cooled bottom of the DS Sn-
52 wt.%Bi alloy casting. For values of 1/λ1/2 fine lower
0
10 than 0.85, a recovery in strength is noted by the
lines representing the results of the Sn-52 wt.%Bi
Results on Bridgman-type furnace
alloy (see Fig. 14a and b). As discussed in a previous
Sn-58wt%Bi [ref 31] article,38 this can be attributed to both an increase
-3 -2 -1 0 in the eutectic fraction from this point and a
10 10 10 10
homogeneous distribution of this eutectic mixture
Growth rate, V (mm/s)
due to the more complex dendritic network caused
Fig. 10. (a) Coarse and (b) fine lamellar spacing as a function of the by the presence of tertiary dendrite arms.
growth rate (V) for the Sn-Bi alloys. R2 is the coefficient of determi- Lower tensile properties are associated with the
nation.
Sn-58 wt.%Bi alloy casting, with an opposite ten-
dency when compared with the other alloys, con-
sidering σu and σy. Bi is considered a brittle phase
and this seems to contribute to this particular
behavior. In contrast, an arrangement of much behavior since the microstructure of the eutectic Sn-
branched dendrites with growth of high-ordered λ3 58 wt.%Bi alloy is constituted by a predominance of
arms, fine Bi phase precipitates within the β-Sn the eutectic mixture and by trifoils and faceted Bi
dendritic matrix, and an important proportion of particles. The ductility deteriorates when compared
very fine eutectic formed by alternate Bi-rich and with the corresponding results of the two other al-
Sn-rich phase (λfine = 0.8 μm and λcoarse = 0.9 μm) loys examined. It is worth noting that open and solid
characterizes the microstructure associated with circle values of strength representing the Sn-58 wt.
the relative position of 6 mm of the Sn-52 wt.%Bi %Bi and the Sn-52 wt.%Bi alloys samples, respec-
alloy DS casting. The combined reinforcement roles tively, remain very close regardless ofr the alloy Bi
of lower values of λfine, λcoarse and high density of Bi content for 1/λ1/2
fine < 0.8. This is due to the very high
precipitates inside the Sn-rich phase were shown to area fraction of coarse lamellar eutectic in the Sn-
induce beneficial effects on the tensile properties 52 wt.%Bi alloy samples (>85%), which approxi-
with both strength and ductility increasing when mates both alloys regarding their microstructures.
compared with the corresponding values obtained Figure 15 shows SEM fracture surfaces of tensile
for P = 90 mm, as shown in Fig. 13b. specimens of the Sn-34 wt.%Bi, Sn-52 wt.%Bi and
Figure 14 synthesizes the results of tensile prop- Sn-58 wt.%Bi alloys from positions in the DS cast-
erties of all the three Sn-Bi alloys experimentally ings corresponding to higher values of ductility, as
examined. The ultimate tensile strength (σu), yield shown in Fig. 14c. Two SEM images with different
tensile strength (σy) and elongation-to-fracture (δ) magnifications (×1000 and ×4000) are depicted for
are plotted with their standard deviations as a each alloy examined.
1766 Silva, da Silva, Garcia, and Spinelli

3
(a) (b) 10
3
10 Sn-34wt%Bi

Secondary dendritic spacing, λ2 (μm)


-1.1 2
Primary dendritic spacing, λ1 (μm)

λ2 = 8.8 (VL) - R =0.99

Sn-52wt%Bi
-1/2 2
2 λ2 = 15.4 (VL) - R =0.91
10
2
10

Sn-34wt%Bi 1
-0.55 2 10
1
10 λ1 = 64 (T) - R =0.96

Sn-52wt%Bi
-1/4 2
λ1 = 310 (T) - R =0.98

0
0
10 10
-1 0
-1
10 10
0 1
10 10 10
Growth rate, VL (mm/s)
Cooling rate, T (°C/s)

(c)
Sn-34wt%Bi
-0.55 2
λ3 = 17 (T) - R =0.99
Tertiary dendritic spacing, λ3 (μm)

Sn-52wt%Bi
-1/4 2
λ3 = 10.5 (T) - R =0.97

1
10

-1 0 1
10 10 10

Cooling rate, T (°C/s)


Fig. 11. (a) Primary (b) secondary and (c) tertiary dendrite arm spacing as a function of cooling rate and growth rate for the DS Sn-34 wt.%Bi and
52 wt.%Bi alloys castings.

The fracture surfaces for the Sn-34 wt.%Bi and 30


Sn-58 wt.%Bi alloys are characterized by the pres-
ence of faceted Bi particles as indicated by black 27
arrows. These cleavage facets may induce the
occurrence of fracture throughout such phase.
Vickers Hardness (HV)

24
Overall, a brittle or faceted fracture type may be
indicated in these cases, which means that the 21
failure mode is basically the same for such Sn-Bi
alloys. Despite the presence of brittle fracture as- 18
pects in the Sn-34 wt.%Bi alloy, the presence of
ductile fracture bulges can be noted on the surface, 15 Sn-34%Bi
as can be seen in Fig. 15a. Sn-52%Bi
Sn-58%Bi
In the case of the Sn-52 wt.%Bi alloy, a particular 12
behavior is observed as compared with the other
aforementioned fractures. A typical ductile fracture 9
0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5
mode is associated with the stress–strain curve of
1/2 -1/2
high toughness (black line in Fig. 13b). As shown in 1/(λfine) (μm)
Fig. 15b, the fracture surface appears to be very fi- Fig. 12. Evolution of Vickers hardness as a function of the inverse of
nely broken, showing a large number of dimples the square root of the fine eutectic spacing for Sn-Bi solder alloys.
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1767
Solder Alloys

(a) 70 (a) 70
65 Sn34wt.%Bi

Ultimate Tensile Strength (MPa)


68
60
66
55
50 64 0.5
σu = 51.5 + 7.1 (1/λfine )
45 62
Stress (MPa)

40 60
35 58
30 56
0.5
25 54 σu = 70 - 20.7 (1/λfine )
0.5
20 52 σu = 61.5 - 7.0 (1/λfine )
15 - Positions from 50
metal/mold interface (P): Sn-34wt%Bi
10 48
6mm 90mm Sn-52wt%Bi
5 Sn-58wt%Bi

σu,
46
0 44
0 3 6 9 12 15 18 21 24 27 30 0,4 0,5 0,6 0,7 0,8 0,9 1,0 1,1 1,2 1,3 1,4
Strain (%) 0.5 - 0.5
1/(λfine) ( µm )
(b) 70
(b) 60
65 Sn-34wt%Bi

Yield Tensile Strength (MPa)


Sn52wt.%Bi 58
60 Sn-52wt%Bi 0.5
σy = 43.2 + 8.3 (1/λfine )
55
56 Sn-58wt%Bi

50 54
45 52
Stress (MPa)

40 50
35 48 0.5
σy = 39.5 + 9.8 (1/λfine )
30
46
25
44 0.5
20 σy = 54.0 - 9.5 (1/λfine ) 0.5
σy = 62 - 21.2 (1/λfine )
15 42
- Positions from metal/mold interface (P):
10 40
σy,

6mm 90mm
5 38
0 36
0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 0,4 0,5 0,6 0,7 0,8 0,9 1,0 1,1 1,2 1,3 1,4
Strain (%) 0.5 - 0.5
1/(λfine) ( µm )
(c) 70
(c) 70
65 65 Sn-34wt%Bi
Sn58wt.%Bi
60 Sn-52wt%Bi
Elogantion-to-fracture (%)

60
Sn-58wt%Bi
55 55
50 50 0.5
δ = -43 + 75 (1/λfine )
45 45
Stress (MPa)

40 40
35 35 0.5
δ = 9.9 + 9.9 (1/λfine )
30 30
25 25
20 20
15 15
δ,

10 - Positions from metal/mold interface (P): 10 0.5


6mm 90mm δ = -5 + 20.5 (1/λfine )
5 5
0 0
0 3 6 9 12 15 18 21 24 27 30 0,4 0,5 0,6 0,7 0,8 0,9 1,0 1,1 1,2 1,3 1,4
0.5 - 0.5
Strain (%) 1/(λfine) ( µm )
Fig. 13. Typical stress–strain curves of specimens extracted from Fig. 14. (a) Ultimate tensile strength-σu, (b) yield tensile strength-σy
two regions; close to the cooled bottom of the DS castings and (c) elongation-to-fracture-δ as a function of fine (λfine) eutectic
(P = 6 mm) and close to the top of the castings (P = 90 mm): (a) Sn- spacing for the Sn-34 wt.%Bi, 52 wt.%Bi and 58 wt.%Bi alloys.
52 wt.%Bi, (b) Sn-52 wt.%Bi and (c) Sn-58 wt.%Bi alloys castings.

that may lead to enhancement of toughness. As a sliding of the interfaces between adjacent phases is
very fine eutectic structure is formed by alternate attained due to the increase in contact area between
Bi-rich and Sn-rich phases (λfine = 0.8 μm and the phases.45 Such a complex interaction may occur
λcoarse = 0.9 μm), an enhancement effect on the especially during necking, with a direct beneficial
1768 Silva, da Silva, Garcia, and Spinelli

Fig. 15. SEM fracture surfaces of the Sn-Bi alloys corresponding to tensile specimens extracted at different positions from the cooled surface of
the DS castings: (a) Sn-34 wt.%Bi; (b) Sn-52 wt.%Bi and (c) Sn-58 wt.%Bi.

effect to toughness as can be seen in the corre- despite the non-equilibrium solidification condi-
sponding stress–strain curve of Fig. 13b. tions, the eutectic DS Sn-58 wt.%Bi alloy casting
has no evidence of dendritic regions, being
CONCLUSIONS characterized by a complex regular eutectic
microstructure typified by coarse and fine re-
The following conclusions can be drawn from the
gions, fishbone-like eutectic cells and trifoils of
present experimental study:
Bi.
1. The DS hypoeutectic Sn-Bi alloys castings were 2. Experimental power–function growth laws were
shown to be characterized by a Sn-rich matrix shown to characterize the growth of primary,
having a complex dendritic array formed not only secondary and tertiary dendritic branches, as a
by primary dendrite stalks but also secondary function of solidification cooling rate and growth
and tertiary dendritic branches with a lamellar rate. The evolution of the lamellar eutectic
eutectic mixture in the interdendritic regions spacing versus the growth rate was shown to be
formed by rhombohedral Bi-rich and body-cen- represented by the −1/2 exponent proposed by the
tered tetragonal Sn-rich phases. In contrast, classical Jackson and Hunt eutectic growth law.
Effects of Solidification Thermal Parameters on Microstructure and Mechanical Properties of Sn-Bi 1769
Solder Alloys

3. A decrease in hardness was shown to occur with 13. T.H. Chuang and H.F. Wu, J. Electron. Mater. 40, 71 (2011).
the increase in the length scale of the eutectic 14. H. Xiaowu, L. Ke, and A. Fanrong, China Foundry 9, 360
(2012).
mixture considering the hypoeutectic Sn-34 wt.% 15. J. Glazer, Inter. Mater. Rev. 40, 65 (1995).
Bi and 52 wt.%Bi alloys. However, an opposite 16. E.P. Wood and K.L. Nimmo, J. Electron. Mater. 23, 709
experimental trend was derived for the eutectic (1994).
Sn-58 wt.%Bi alloy, indicating that hardness has 17. J.F. Li, S.H. Mannan, M.P. Clode, D.C. Whalley, and D.A.
Hutt, Acta Mater. 54, 2907 (2006).
been negatively affected by the growth of trifoils 18. J.W. Yoon, C.B. Lee, and S.B. Jung, Mater. Trans. A 43,
and faceted particles of Bi (with fishbone 1712 (2002).
branches) associated with the eutectic mixture. 19. K.W. Moon, W.J. Boettinger, U.R. Kattner, C.A. Handw-
4. Hall–Petch type equations have been proposed erker, and D.J. Lee, J. Electron. Mater. 30, 45 (2001).
relating σu; σy and δ with λfine. For the Sn-34 wt. 20. X.Y. Liu, H.T. Ma, J. Zhao, and L. Wang, Conference on
High Density Microsystem Design and Packaging and
%Bi and 52 wt.%Bi alloys, σu, σy and δ increase Component Failure Analysis, 1 (2005).
with decreasing λfine. A decrease in the eutectic 21. A. Torres, L. Hernández, and O. Domı́nguez, Mater. Sci.
spacing induced a more homogeneous distribu- Appl. 3, 355 (2012).
tion of the Bi-rich and Sn-rich phases throughout 22. M.N. Croker, R.S. Fidler, and R.W. Smith, Proc. R. Soc.
Lond. A 335, 15 (1973).
the microstructure, with beneficial effects on the 23. M.X.F. Gigliotti, L.F.P. Graham, and G.A. Colligan, Metall.
mechanical strength. The lowest tensile proper- Trans. 1, 1038 (1970).
ties were shown to be associated with the Sn- 24. Z. Lai and D. Ye, J. Mater. Sci. Mater. Electron. 27, 3182
58 wt.%Bi alloy with an opposite tendency when (2016).
compared with the corresponding results of the 25. M. McCormack, H.S. Chen, G.W. Kammlott, and S. Jin, J.
Electron. Mater. 26, 954 (1997).
hypoeutectic alloys. This particular behavior has 26. U. Böyük and N. Maraşli, Mater. Chem. Phys. 119, 442
been attributed to the microstructure of the (2010).
eutectic Sn-58 wt.%Bi alloy, which was shown 27. P.D. Pereira, J.E. Spinelli, and A. Garcia, Mater. Des. 45,
to be formed by predominance of eutectic mixture 377 (2013).
28. W.R. Osório, D.R. Leiva, L.C. Peixoto, L.R. Garcia, and A.
and by trifoils and faceted Bi particles. Garcia, J. Alloys Compd. 562, 194 (2013).
29. W.R. Osório, L.C. Peixoto, L.R. Garcia, N. Mangelinck-Noël,
ACKNOWLEDGEMENTS and A. Garcia, J. Alloys Compd. 572, 97 (2013).
30. O.V. Gusakova, V.G. Shepelevich, and L.P. Shcherbachen-
The authors acknowledge the financial support ko, J. Surf. Investig.-X-Ray 10, 146 (2016).
provided by FAPESP (São Paulo Research Founda- 31. A.P. Silva, J.E. Spinelli, and A. Garcia, J. Alloys Compd.
tion, Brazil: Grants 2015/11863-5 and 2013/08259- 475, 347 (2009).
3), CNPq and CAPES-COFECUB (Grant 857/15). 32. J.E. Spinelli, N. Cheung, P.R. Goulart, J.M.V. Quaresma,
and A. Garcia, Int. J. Therm. Sci. 51, 145 (2012).
33. P.R. Goulart, J.E. Spinelli, and N. Cheung, Mater. Chem.
REFERENCES Phys. 119, 272 (2010).
34. M. Gunduz and E. Çadirli, Mater. Sci. Eng. A 327, 167
1. M.H. Braga, J. Vizdal, A. Kroupa, J. Ferreira, D. Soares, and (2002).
L.F. Malheiros, CALPHAD 31, 468 (2007). 35. J.E. Spinelli, B. Silva, and A. Garcia, Mater. Des. 58, 482
2. R.M. Shalaby, Mater. Sci. Eng. A 560, 86 (2013). (2014).
3. Y. Goh, A.S.M.A. Haseeb, and M.F.M. Sabri, Electrochim. 36. L. Shen, P. Septiwerdani, and Z. Chen, Mater. Sci. Eng. A
Acta 90, 265 (2013). 558, 253 (2012).
4. P. Vianco, J. Rejent, and R. Grant, Mater. Trans. 45, 765 37. O. Mokhtari and H. Nishikawa, Mater. Sci. Eng. A 651, 831
(2004). (2016).
5. X. Chen, J. Zhou, F. Xue, and Y. Yao, Mater. Sci. Eng. A 662, 38. K.A. Jackson and J.D. Hunt, Trans. Metall. Soc. AIME 236,
251 (2016). 1129 (1966).
6. H. Sun, Q. Li, and Y.C. Chan, J. Mater. Sci. Mater. Electron. 39. B. Silva, H. Nguyen-Thi, G. Reinhart, N. Mangelinck-Nöel,
25, 4380 (2014). A. Garcia, and J.E. Spinelli, Mater. Charact. 107, 43 (2015).
7. L. Shen, Z.Y. Tan, and Z. Chen, Mater. Sci. Eng. A 561, 232 40. A. Garcia and M. Prates, Metall. Trans. B 9, 449 (1978).
(2013). 41. A. Garcia, T.W. Clyne, and M. Prates, Metall. Trans. B 10,
8. Y. Peng and K. Deng, J. Alloys Compd. 625, 44 (2015). 85 (1979).
9. R. Dai, S.G. Zhang, and J.G. Li, J. Electron. Mater. 40, 2458 42. L.R. Garcia, W.R. Osório, L.C. Peixoto, and A. Garcia, Ma-
(2011). ter. Charact. 61, 212 (2010).
10. Z. Mei and J.W. Morris Jr., J. Electron. Mater. 21, 599 43. J.E. Spinelli, B.L. Silva, and A. Garcia, J. Electron. Mater.
(1992). 43, 1347 (2014).
11. J.L.F. Goldstein and J.W. Morris Jr., J. Electron. Mater. 23, 44. D. Bouchard and J.S. Kirkaldy, Metall. Mater. Trans. B 28,
477 (1994). 651 (1997).
12. F. Hua, Z. Mei, and J. Glazer, Electronic Components; 45. C.H. Raeder, D. Mitlin, and R.W. Messler Jr, J. Mater. Sci.
Technology Conference, 48th IEEE, p. 277 (1998). 33, 4503 (1998).

You might also like