Anda di halaman 1dari 15

phys. stat. sol. (a) 198, No. 1, 76 90 (2003) / DOI 10.1002/pssa.

200306592

Study of the mechanical and thermal properties of Sn5 wt% Sb solder alloy at two annealing temperatures
M. M. EL-Bahay1, M. E. EL Mossalamy1, M. Mahdy1, and A. A. Bahgat*
Department of Physics, Faculty of Science, AL-Azhar Univ., Nasr City 11884, Cairo, Egypt, (1) Girls Branch Received 10 June 2002, revised 2 January 2003, accepted 21 February 2003 Published online 22 May 2003 PACS 62.20.Hg; 64.70.Kb, 81.70.Pg Sn5 wt% Sb alloy is one of the materials considered for replacing lead-containing alloys for soldering in electronic packaging. Differential thermal analysis (DTA) and specific heat of the sample were studied. Wetting contact angle measurements of the alloy on different substrates were carried out at high temperature. Microhardness tests as a function of temperature were performed to calculate the effective activation energy of the solder alloy Sn Sb and compared with the pure elements Sn and Sb. Isothermal creep curves for alloy samples were obtained under different constant applied stresses at different working temperatures ranging from 463 K to 503 K, followed by annealing the samples at two different temperatures before and above the threshold value (Tm/2). The transient creep parameters and the values of the stress exponent n were calculated for the two annealing temperatures. Microstructure examinations of the as-cast alloy at room temperature and after the two annealing treatments with the effect of the cold work deformation and creep test on the structure change and properties of Sn Sb alloys were reported. The stress rupture test was also measured.

Introduction

Considerable research and development efforts are underway towards the development of lead free solder alloys. The choice of alloy used depends upon factors such as the strength and corrosion resistance. Most of these solders are tin-containing binary and ternary alloys [1]. The use of tin alloys for joining metals is often referred to as soft soldering. Low levels of antimony addition are known to improve the mechanical properties of tin and to inhibit the allotropic transformation. The antimonial tin solder (Sn95Sb5) alloy with a near peritectic composition has the advantages of good room-temperature creep resistance [2, 3] and mechanical strength, although its melting point is high 518 K compared to that of the popular lead tin eutectic 456 K. This solder has a very high room-temperature ductility and is used in electrical equipment, copper tubing, cooling coils for refrigerators, and avoids lead for service in contact with foodstuffs [4]. It is also adequate for consideration in prototype solder joint development in microelectronics packaging [5, 6]. This presentation reports some physical analysis of the binary Sn 5 wt% Sb solder. The specific heat and the differential thermal analysis (DTA) were studied in the temperature range 320 to 550 K. Wetting contact angle due to different metal substrates and the microhardness as a function of temperature were measured. Because creep is one of the most common and important micromechanical deformation mechanisms in solder joints, we performed creep tests on samples annealed at two different temperatures, 326 2 K
*

Corresponding author: e-mail: alaabahgat@hotmail.com, Fax: +02 27 33 948

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 0031-8965/03/19807-0076 $ 17.50+.50/0

phys. stat. sol. (a) 198, No. 1 (2003)

77

and 486 2 K. The test was done on samples with step change in the initial stress ranging from 3.95 to 5.86 MPa, under different test temperatures from 463 to 503 K. The presentation analyzes the experimental data of the transient and the steady-state mechanisms during constant-stress creep tests at two annealing temperatures. Also, the stress rupture was calculated for the different conditions of the creep tests. Metallographic examination was performed on the as-cast alloy at RT and on the (343 2 K) and (486 2 K) annealtreated specimens. The structure changes of the alloy were reported before and after the creep test.

Experimental details

The material investigated, Sn 5 wt% Sb, solder alloy was prepared from spec-pure elements 99.99%. It was formed by melting the constituents in a Pyrex tube under a fluxing agent (Colophony) to prevent the sample oxidation in air. The melting was carried out on a benzene flame for more than 15 min while the tube was shaken to ensure homogenization of the melt. After solidification the tube was broken to obtain the alloy. Excess flux was removed by soaking the ingot in pure carbon tetrachloride (CCl4) for a sufficient time. The density of the sample was measured by the displacement method using carbon tetrachloride (CCl4) as the immersion liquid. Measurement of wetting angle of metallic drops on smooth surfaces was evaluated photographically with a steel sphere as a reference. A home-made computer-assisted differential thermal analyzer, DTA [7], was used to define the melting point of the alloy as well as to calculate the heat of fusion. On the other hand the specific heat of Sn 5%Sb sample was measured applying Newton cooling curves with Al2O3 as the reference material, the data were collected via a computer. Microhardness measurement [8] was performed on the sample using a Vickers Shimadzu microhardness tester. Grinding and mechanical polishing were necessary to obtain a polished smooth and flat parallel surface before indentation testing. The applied load was 50 g for 5 s and seven readings of different indentations were taken at room temperature to obtain the mean value and the standard deviation was calculated. Additional modification to this system was made so that microhardness as a function of temperature can be evaluated with high precision. The bulk samples were divided into two groups, and isochronally annealed for one hour at 343 2 K and 486 2 K, respectively, in a furnace with an accuracy of 2 K. Creep specimens of Sn 5% Sb were swaged in the form of wires (diameter 0.07 cm) with cold work of approximately 94.5%. The wires were then divided into two groups and were annealed for one hour at 326 2 K (0.635Tm) and 486 2 K (0.947Tm), respectively. Following heat treatment the samples were allowed to slowly cool to room temperature at a rate of 4 102 K s1. Tensile creep tests were carried out using a home-made computerassisted creep machine in the temperatures range up to 503 K, with an accuracy of 2 K under constantstress conditions. An iron rod that is coupled to the bottom of the weight tray, with its vertical movement within the core of a linear transformer, was used to evaluate the strain. This is connected to an acvoltmeter fed to a PC computer via an A/D converter. An overall resolution of 40 m was achieved, which corresponds to a strain of the order 0.002 and is sufficient for ductile samples. All samples (bulk, wires before and after creep test) were polished, etched in a solution consisting of (3 ml HNO3 + 97 ml alcohol) for about 7 s and washed by distilled water. Microphotographical examination of the structure changes of the specimens after different heat treatments, cold work and the creep test was done. Metallography was performed using an optical microscope (Carl Zeiss Jena) at different magnifications.

Results and discussion

An X-ray diffraction pattern of the powdered sample of the Sn 5 wt% Sb alloy was recorded using a Philips (1390) X-ray diffractometer. The diffraction pattern shows the existence of sharp peaks, which

78

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy Fig. 1 Microstructure of as-cast Sn 5 wt% Sb alloy, a) at RT, b) after annealing at 343 2 K for 1 h, c) after annealing at 486 2 K for 1 h.

are attributed to the crystalline nature of the sample according to the JCPDS X-ray diffraction file. The value of the bulk density as determined in the present work was 7.3 g cm3, from which the porosity is determined to be 0.5%, proving the efficiency of the present preparation procedure. The microstructure of the as-cast Sn Sb alloy at room temperature, and after two different annealing temperatures 343 K and 486 K with magnification (500) are shown in Fig. 1ac. According to the SnSb phase diagram [9] considered from the multiphase alloys whose chemical composition does not correspond exactly to the eutectic concentration, a precipitated phase can segregate along with the eutectic. For instance in the alloys whose concentration is within the temperature interval between the liquidus and solidus lines [10], a precipitated phase ( phase) usually solidifies. This phase is in the form of large grains that are linked to the peculiarities of the growth of the precipitated phase as solidification of the alloy proceeds within the temperature interval between the liquidus and solidus as shown in Fig. 1a. The structure of the bulk samples after different annealing temperatures (343 2 K and 486 2 K) are shown in (Fig. 1b and c). An appreciable increase of temperature causes the formation of large grains in the alloy. By comparing the micrographs it can be seen that the alloy heated to a higher temperature has coarser grains, a reduced plasticity, and higher strength. 3.1 Thermal analysis The calculated enthalpy, H, from the DTA thermogram released in the transformation process [11] is 37.216 kJ/kg. On the other hand, the specific heat curves versus temperature together with the temperature time cooling curve for Sn 5% Sb are shown in Fig. 2(a and b). The solidus and liquidus points are 506 1 and 516 1 K from the DTA and the specific heat measurements respectively, i.e. the same as in Refs. [4, 9]. The result of the specific heat at a temperature of 344 K is 219 J/kg K. For any alloy to be worthwhile as a solder for the electronics industry it must possess certain specific qualities like a melting range. It must have a liquidus temperature (above which it is completely molten) that is sufficiently low so that electronic components and boards are not damaged during soldering. In practice this means that it must be usable at 533 K, which is the maximum exposure temperature limit for the majority of electronic components. By the same argument it must have a solidus temperature (below which it is completely

phys. stat. sol. (a) 198, No. 1 (2003)

79 Fig. 2 a) Specific heat versus temperature for Sn 5 wt% Sb alloy and b) temperature/time cooling curve.

solid) that is sufficiently high so that during service the solder joints do not lose their mechanical strength. The melting point of antimonial tin solder alloy is 516 K, provides a useful compromise between these two criteria. A property of importance to the manufacturing or product engineer is wettability. Indeed, wettability, defined as the tendency for liquid metal to spread on a solid surface, is the precursor of solderability, which describes the solders ability to form an actual joint on a circuit board. Wettability is quantitatively assessed by the contact angle formed at the solder substrateflux triple point. Compared with a contact angle of 17 4 for Pb Sn eutectic, the contact angle of Sn 5% Sb as obtained is higher 43 4, but is adequate for consideration in prototype development [5, 6]. In the present study, a fluxing agent (Colophonic acid) was used for coating the liquid solution during the preparation of the molten alloys, resembling the real industrial application condition. The measured contact angles from the wetting behavior of the prepared binary molten alloy on different substrate (Cu, CuZn30, glass, and pure Al metal, probably Al2O3 coated) at a temperature of 573 K after 60 s elapsed time are 40, 41, 118, and 115, respectively. The glass substrate was chosen to resemble a dielectric one. 3.2 Microhardness measurements Hardness measurements during heating (hot hardness measurements) are important when studying various grades of heat-stable and heat-resistance structural materials. The factors that may affect these measurements are: the local cooling caused by a colder indenter may cause varieties in the measured hardness and thus affect the accuracy of measurements. Appreciable errors may also be due to the oxidation of the specimen surface, which occurs if heating is done in a furnace without a protective atmosphere, the oxidation being more pronounced with increasing temperature [10]. From the relation between the microhardness and the heating temperature of the Sn5 wt% Sb alloy and pure metals Sn and Sb, it is seen that by increasing the alloy temperature the microhardness decreases exponentially, meaning that heating can reduce wear resistance by decreasing mechanical properties due to softening. An Arrhenius equation describes the obtained results quite well, and which can be given in the form: HV(T) = HV0 e(E/kT) , (1)

CP (J / kg.K)

80
Sb Sn-5 wt% Sb Sn

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy Fig. 3 Variation of the Vickers hardness number as a function of (1/T) according to Arrhenius equation for pure metal Sn, Sb, and Sn 5 wt% Sb alloy.

6.4 6.2 6.0

where HV(T) is the hardness under thermal condition, HV0 is the hardness of the initial 5.0 state, k is the Boltzmann constant and E may be considered to describe the effective activa4.5 tion energy associated with thermal softening of the material. As a measure of the microhardness, a plot of the logarithm of the hard4.0 ness versus the reciprocal of the absolute temperature would be a straight line as shown in 3.5 Fig. 3, according to Eq. (1). The effective 2.7 3.0 3.3 activation energy E, of the Sn, Sb, and SnSb alloy were calculated accordingly. Table 1, 1/T (103 K1) summarizes the microhardness at RT and the effective activation energy for the pure metal Sn, Sb, and the Sn 5% Sb alloy. From the table, the effective activation energy may seem to be rather low. This may originate from a combination of complex plastic processes occurring in the surface layer and can not be simply related to single elementary processes. The measurement of hardness, especially microhardness, is a usual method to characterize the mechanical properties of solid state surfaces. It is a measure of a materials resistance to localized plastic deformation. A technique to strengthen and harden metals is alloying with impurity atoms that go into either substitutional or interstitial positions; this is called solid-solution strengthening. Therefore alloys are stronger than pure metals because impurity atoms that go into solid solution ordinarily impose lattice strains on the surrounding host atoms, in addition to the precipitation of a phase. Lattice strain field interactions between dislocations and these impurity atoms result, and, consequently, dislocation movement is restricted. Meyers and Chawla [12] point out that the work hardening introduced in the surface by polishing may be important at lighter loads. Dieter [13] argues that for lighter loads the impression is considerably affected by the small amount of elastic recovery, i.e. that is normally negligible for higher loads. Braunovic [14] proposed that the mechanism for load dependence of indentation hardness is associated with the movement of dislocations introduced by the indenter. The generation of dislocations at a contact surface can be visualized as arising from a punching mechanism. Pinning and blocking of dislocation may impede this movement by lattice defects such as grain boundaries, vacancies, solute atoms, and precipitation of the new phase ( phase). 3.3 Creep results The creep behavior observed for an initially annealed alloy at 326 2 K (i.e. 0.635Tm) and 486 2 K (i.e. 0.947Tm) for one hour were of the normal type, i.e. a rapid strain jump (instantaneous strain) folTable 1 Room-temperature Vickers hardness number and the effective activation energy of the pure metals and the SnSb alloy.

ln Hv

Sb HV at RT (N/mm ) effective activation energy (kJ/mol)


2

Sn 76.7 4 4.37 0.05

Sn5 wt% Sb 158.7 5 4.29 0.05

582.8 19 4.55 0.05

phys. stat. sol. (a) 198, No. 1 (2003)


5.1 4.7

81 Fig. 4 Isothermal creep curves for Sn 5 wt% Sb alloy annealed at 486 K for 1 h, under different stresses and test temperatures.
3. 3. 9
4.3

0.2 0.1

lowed by decreasing strain rate with time (primary creep) followed by an extended period of constant strain rate 0.0 0 2000 4000 6000 8000 10 000 creep (secondary creep) with tertiary creep, see Fig. 4 for example. It is usual to obtain at each test temperature Time (s) (463, 473, 483, 493, and 503 K) respectively a family of creep curves, each member produced by a different initial 0.4 stress (3.9, 4.3, 4.7, 5.1, 5.4, and 5.8 MPa). The creep data at 463, 483, and 503 K for example, with initially an0.3 nealed alloys (486 K for 1 h) at the different stresses are shown in Fig. 4. The creep curves have the same typical 0.2 behavior for both pure metals and many class M creep alloys [15]. When specimens initially annealed at 326 K 0.1 are strained at a temperature below the threshold (Tm/2; Tm T = 483 K = melting temperature in K), continuous strain hardening 0.0 0 1000 2000 3000 4000 occurs. This strain hardening systematically decreases as the temperature of strain exceeds a threshold value. Such Time (s) behavior also occurs for the samples annealed at the high temperature (486 K = 0.947Tm) when the ductility at test 0.5 temperatures of 463 and 483 K is small. It is in the range 0.4 of 25% to 45% for the different stresses, while that for specimens annealed at the low temperature (326 K = 0.3 0.635Tm) reaches 40% to 80% under the same deformation 1 5. conditions. At the test temperature of 503 K, the behaviors 0.2 of the two groups are different, although the ductility is nearly the same, the creep time was lower for specimens 0.1 T = 503 K treated at a low annealing temperature of 326 K than for 0.0 those at high annealing temperature. For high-temperature 0 200 400 600 800 strain, the recovery is activated as much by internal stresses in the alloy as by thermal agitation alone, or the Time (s) rise in temperature accentuates dynamic recovery as much as pure thermal recovery. The representative plot of the strain rate versus time for the Sn Sb alloy shown in Fig. 5 is an indication of the three different creep stages (transient, steady-state, and tertiary). This curve illustrates the large change in creep rate that occurs during the creep test. Since the stress and temperature are constant, this variation in strain rates (creep rates) is the result of changes in the internal structure of the material with creep strain and time. The following mechanisms are considered to determine the creep mechanism: dislocation glide, non-conservative motion of dislocation, grain-boundary sliding, stress-directed diffusion of vacancies and intercrystalline void nucleation and growth [16].
T = 463 K
5.1 4.7
5.8 MP a 5.4

4.3

4.3

0.3

5.8 M Pa 5.4

0.4

Pa

5.8 M

3.3.1 Transient creep characteristics The transient creep deformation [17] is determined by the movement of mobile dislocations and ends in their effective locking at obstacles or is originally considered as being due to the exhaustion of dislocation sources. Mott [18] proposed the theory of thermally activated dislocation motion taking into account

5.4

4.7

3.

82
2.4 1.6
. 102

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy
A B C

= 5.8 MPa T = 503 K

Fig. 5 Example of creep rate curve for Sn 5 wt% Sb alloy showing different stages or intervals of the creep behavior. A represents the transient-state interval, B represents the steady-state interval and C represents the tertiary-state interval.

0.8 0

20

40 t (s)

60

80

the increase in dislocation density and occurrence of a certain degree of work hardening during the transient creep deformation. Because solder alloys are used at high homologous temperature (T/Tm = 0.9) where Tm is the melting point of a solder, and stresses (105 < /E < 103), where E is Youngs modulus, dislocation creep is frequently assumed to determine the deformation kinetics [16]. The transient creep could be represented by the empirical equation [19], a
2.5

tr = t ntr ,

(2)

3.0 3.5 4.0 T = 503 K 483 473 463 2.0 3.0 4.0 ln (t) 5.0 6.0

4.5

1.5 2.0 2.5

where and ntr are the transient creep parameters, t is the creep time (in s). The values of tr were extracted from the experimental creep curves within the interval presented by, A, shown in Fig. 5. Figure 6a and b and Fig. 7a and b show the transient creep time transient strain [ln (t) versus ln tr] curves for the Sn 5 wt% Sb specimen under different test temperatures and different stresses at the annealing temperatures 326 2 K and 486 2 K, respectively. The rate-controlling mechanism is determined by the transient creep parameters and ntr, which are given by Eq. (2). The parameter ntr was calculated from the slopes of the lines shown in Fig. 6 and Fig. 7. Values ranging from 0.57 to 0.72 and 0.37 to 0.53 for the two different annealing temperatures, respectively, were evaluated and shown in Fig. 8a and Fig. 9a. From these figures, ntr, was also found to increase

ln tr ln tr

3.0 3.5 4.0 4.5 3.0 3.5 4.0 4.5 ln (t) T = 503 K 493 473 463 5.0 5.5

Fig. 6 Relation between ln (t) versus ln tr for Sn 5 wt% Sb alloy annealed at 326 2 K and tested at various temperatures and stresses of a) 5.1 MPa and b) 5.8 MPa.

phys. stat. sol. (a) 198, No. 1 (2003)

83

Fig. 7 Relation between ln (t) versus ln tr for Sn 5 wt% Sb alloy annealed at 486 2 K and tested at various temperatures and stresses of a) 4.7 MPa and b) 5.1 MPa. T = 503 K, 483 K, 473 K, 463 K.

Fig. 8 Dependence of a) the exponent ntr and b) the parameter , on the test temperature at different applied stresses for the Sn 5 wt% Sb alloy annealed at temperature 326 2 K. 5.8 MPa, 4.7 MPa.

with increasing the test temperature. The value of on the other hand changes with increasing test temperatures as shown in Fig. 8b and Fig. 9b, while having values ranging from 24 104 to 118 104 and 16 104 to 89 104 sn, for the two different annealing temperatures respectively. By performing creep tests for the alloy it is expected that in the low-temperature region, Sn-rich phase and Sb-rich phase coarsen and the solubility of Sb in Sn increases with increasing test temperature [20]. The increase in solubility may contribute to the observed increase in the transient strain rate at higher temperatures. As shown from the phase diagram and from the DTA and specific heat measurements (Fig. 2) the solidus and liquidus points are 506 1 and 516 1 K, respectively. A deviation from the usual behavior ln tr versus ln t, and the parameters and ntr at the test temperature of 503 K and annealing temperature of 326 K can be seen (Fig. 6a and b) and (Fig. 8a and b). This may be related to the overall time interval of this region (only from t = 20.0 to 55.0 s or less). While the test temperature is almost equal to the solidus point and the microstructure and grain size shown in Fig. 1b may not have the necessary time interval to grow as in Fig. 1c (annealing at 486 K) where this deviation is not observed (Fig. 7 and Fig. 9). 3.3.2 The steady-state creep rate The steady-state creep time can be determined from Fig. 5. The discussion below will compare and analyze the available data for the Pb-free alloy Sn Sb on the following: 1. Stress- and temperature-dependence of creep on the stress exponent (n) parameter. 2. Activation energy of the steady-state creep.

84
0.55 a 0.50

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy

Fig. 9 Dependence of a) the exponent ntr and b) the parameter , on the test temperature at different applied stresses for the Sn 5 wt% Sb alloy annealed at temperature 486 2 K. 5.1 MPa, 4.7 MPa

ntr

0.45

0.40

0.35 460 470 480 490 500 510 5.1 MPa 4.7 Temperature (K) 100 b 80 60
104

40

20

0 460 470 480 490 500 510 Temperature (K)

A commonly used equation by design engineers for stress and temperature dependence of creep strain rate, especially in the electronics industry, is the Norton power law [21],
s = A n e Q / kT ,

(3)

where s is the steady-state strain rate, A is a constant, is the stress, n is the stress exponent, Q is the activation energy, k is the Boltzmanns constant and T is the absolute temperature. The stress exponent, n, determines the type of deformation mechanism (viscous glide, screw dislocation, climb mechanism, etc.). The transition from viscous glide (stress exponent n 3) to dislocation climb (stress exponent n 4) confirms the theoretical prediction by Bird et al. [22], in which they concluded that the steady-state creep rate of a metal deforming at temperature above 0.5Tm through any diffusion-controlled mechanism is described by a power-law relation [23],
s = BDV Gb kT G
n

(4)

where G is the shear modulus, b is the Burgers vector and B is a constant.

phys. stat. sol. (a) 198, No. 1 (2003)

85

Although many mechanisms can be operable at one time, usually the one that exhibits the highest creep rate becomes dominant over certain ranges of /G and T. Since the diffusion coefficient DV in Eq. (4) can be described by
DV = D0 e Q / kT .

(5)

Equation (3) can be written as


s = C n Q / kT . e T

(6)

It is to be noted that the form of Eq. (6) differs from Eq. (3) by a factor of 1/T. Of course, G, decreases with heating imparting a temperature dependence to C. The T term in the exponent is much more important than the pre-exponent 1/T and the small decrease in the modulus on heating. Both the Norton and Dorn power laws Eq. (3) and Eq. (6) gave good fits to the steady-state flow stress versus strain rate. The creep data are well described by both equations [24]. The double-log plot of minimum creep rate versus stress at different test temperatures, are shown in Fig. 10a and b for Sn5 wt% Sb alloy for the two different annealing temperatures respectively. From the figures the values of the stress exponent, n, are in the range 3.2 to 5.9 as determined for the two anneal2.5

n = 5.9

3.0
. log (s)
n = 3.6 n = 4.05 n = 4.8 n = 5.7

3.5

4.0 4.5 T = 503 K 0.60 0.65 0.70 0.75 log [ (MPa)] 0.80 493 K 483 K 2.5 b 3.0 3.5
. log (s)
n = 3.2 n = 5.5 n = 5.2 n = 4.6

473 K 463 K

4.0 4.5 5.0 0.60 0.65 0.70 0.75 log [(MPa)]

n = 4.6

Fig. 10 Double-log plot of minimum creep rate s versus stress at different test temperatures for Sn 5 wt% Sb alloy a) annealed at 326 2 K and b) annealed at 486 2 K, respectively.

0.80

86

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy Table 2 The stress exponent n for Sn Sb alloy at different annealing and test temperatures as obtained from Fig. 10.

annealing temperature 326 2 K

test temperature 463 473 483 493 503 463 473 483 493 503

exponent n (present) 0.1 5.7 4.8 4.05 3.6 5.9 4.6 4.6 5.2 5.5 3.2

exponent n [reference] 3.2 [3] 7.0 [3] 3.0 [1] 5.0 [1]

486 2 K

6 7
. ln (s)

5.8 MPa 5.4 5.1 4.7 4.3 3.9

8 9 10 11 a 2.00 2.05 2.10 1/T (103 K1) 2.15

7 8
. ln (s)

5.8 MPa 5.4 5.1 4.7 4.3 3.9

ing temperatures. Table 2 shows the present experimental data of the, n, exponent in comparison with previous studies. These values for n suggest that the deformation is associated with viscous flow at grain boundaries due to localized slip within the grain [25]. Moreover, Sherby and Burke [15] observed that, for small grain size, grain-boundary shearing is expected when viscous glide and dislocation-climb mechanisms are operating. On the other hand, the activation energy for creep (i.e. the probability that a climb event occurs) depends on the applied stress and on the structure. The activation energy of Sn5 wt% Sb alloy could be calculated from the slopes of the lines in Fig. 11a and b at the two different annealing temperatures, respectively. From these lines the values of the activation energy are given in Table 3. The activation energy of the steady-state creep is not far from that of self-diffusion 100 kJ/mol [26]. Figure 12a shows the microstructure of Sn 5% Sb alloy after cold deformation. The

9 10 11 b 2.00 2.05 2.10 1/T (103 K1) 2.15

Fig. 11 Relation between ln ( s ) and 1/T (K) for Sn 5 wt% Sb alloy at different applied stresses a) annealed at 326 2 K and b) annealed at 486 2 K.

phys. stat. sol. (a) 198, No. 1 (2003) Table 3 The activation energy from the steady state at different stresses for the Sn Sb alloy.

87

annealing temperature 326 2 K

stress (MPa) 5.8 5.4 5.1 4.7 4.3 3.9 5.8 5.4 5.1 4.7 4.3 3.9

activation energy (kJ/mol) 0.2 105.4 123.8 129.5 108.8 123.6 119.3 107.4 108.2 115.2 135.8 115.9 120.0

average activation energy (kJ/mol) 118.4

486 2 K

117.1

structures of the alloy after the low annealing temperature 326 K, Fig. 12b, have nearly the same grain size. After increasing the annealing temperature to 486 K as shown in Fig. 12c, or after recrystallization it is evident that the shape and size of the grains have changed and grown. Under the same conditions the alloy that has formed a fine-grained structure by (primary) recrystallization will experience pronounced grain growth if it is heated at a higher temperature. The mechanism involved the rapid migration of the boundaries of a few of the primary recrystallized grains, with the result that the majority of the primary grains are consumed and very large secondary grains are created. This process is shown in Fig. 1c and Fig. 12c for comparison. It is evident that in this alloy the secondary grains are more perfect than the initial recrystallized grains. Other names for these phenomena are discontinuous grain growth and abnormal grain growth [27]. As shown in Fig. 13a, the micrograph after low-temperature annealing at 326 K and after

Fig. 12 Microstructure of Sn 5 wt% Sb a) after cold work at RT, b) before creep test after annealing at 326 2 K for 1 h and c) before creep test after annealing at 486 2 K for 1 h.

88

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy Fig. 13 Microstructure of Sn 5 wt% Sb a) after creep test under 5.4 MPa and 463 K test temperature and annealing at 326 2 K for 1 h. b) after creep test under 5.1 MPa and 463 K test temperature after annealing at 486 2 K for 1 h.

the creep test at stress of 5.47 MPa differs in grain size from that in Fig. 13b, where deformation shows considerable coarsening of the grains [28]. The density decrease of the SnSb intermetallic particles within the matrix is due to grain-boundary segregation and/or dissolution of the SnSb intermetallic. But at the higher annealing temperature of 486 K, the deformation that occurred due to the creep test made changes in the shape of grains as shown in Fig. 13b. 3.3.3 The stress-rupture test The creep test measures the dimensional changes, which occur from elevated temperature exposure, while the stress rupture test measures the effect of temperature on the long-time load-bearing characteristics. In the creep test the total strain is often less than 0.5%, while in the stress-rupture test the total strain may be around 50%. The mode of failure for the alloy is predominantly intergranular [28]. The basic information obtained from the stress-rupture test is the time to cause failure at a given nominal stress for a constant temperature. The stress-rupture time curves for Sn5 wt% Sb alloy are shown in Fig. 14a and b at 326 2 K and 486 2 K annealing temperatures respectively. At low annealing temperature (Fig. 14a) the straight lines obtained from the log (stress) log (time) curves for all test temperatures have nearly equal slopes. At higher annealing temperatures (Fig. 14b) we obtain different slopes at high test temperatures 503, 493, and 483 K respectively, while at lower test temperatures of 463 and 473 K, the curves have two different slopes where the annealing temperature in this case is greater than the test creep temperature. Changes in the slope of the stress-rupture line as shown in Fig. 14a and b, are due to structural changes occurring in the material, e.g., changes in intergranular fracture, recrystallization, and grain growth or other structural changes.

Conclusions

Some physical and creep properties at two different annealing temperatures of the solder alloy Sn 5 wt% Sb, were studied experimentally and discussed in this paper. The main conclusions are as follows: 1. The thermal analysis includes differential thermal analysis, DTA, the latent heat of fusion and specific heat were measured. The solidus and liquidus points are 506 K and 516 K, respectively, this means

phys. stat. sol. (a) 198, No. 1 (2003)

89 Fig. 14 Log stress-log rupture time relation for annealed Sn 5 wt% Sb alloy at a) 326 2 K and b) 486 2 K.

0.80

0.75
log [(MPa)]

0.70

0.65

0.60 1.5 2.0 2.5 3.0 3.5 log [t (s)] 4.0 4.5 T = 503 K 493 483 b 473 463

0.80 0.75 0.70 0.65 0.60 0.55 2.0 2.5 3.0 3.5 log [t (s)] 4.0

log [(MPa)]

4.5

that they are sufficiently high so that during service the solder joints do not lose their mechanical strength. 2. The wetting behavior of the alloy developed in this study was evaluated at different substrates. On CuZn30 and pure Cu at 300 C, the contact angles are 41 and 40, respectively. While on Al and glass substrates they are 115 and 118, respectively. This means that the alloys melt could be used for soldering or coating brass objects, while for Al or glass they could be used in non-wettability applications. 3. Microhardness measurements on the pure Sn, Sb, and SnSb alloy were measured at room temperature. The low levels of antimony addition increase the microhardness of tin from 76.7 to 158.7 N/mm2. The effective activation energy was calculated from an Arrhenius plot as ln HV versus 1/T for the pure metals and the Sn Sb alloy where their values are in the range 4.29 to 4.55 kJ/mol. 4. The degree to which the three creep stages are readily distinguishable depends strongly on the initial applied stress and temperature. A maximum ductility of 90% was obtained at a low annealing temperature of 328 K for the test temperature of 463 K. 5. The parameters and ntr are determined from the transient creep stage. The values of the exponent ntr and parameter change with increasing temperature. From the steady-state stage, the values of the

90

M. M. EL-Bahay et al.: Mechanical and thermal properties of Sn5 wt% Sb solder alloy

stress exponent n between 3.2 and 5.9 depend on the applied stress. The n exponent values are governed by dislocation glide at n = 3 and at n = 4 the dislocation climb is predominant. 6. The microstructure of the as-cast Sn 5% Sb alloy represents a precipitated phase ( phase) and usually solidifies in the form of large grains within the temperature interval between the liquidus and solidus, while the alloy heated to a higher temperatures has coarser grains. 7. The stress rupture times have a regular trend in the case of low annealing temperature. For the higher annealing temperature of 486 K, the results have a different behavior because the applied annealing temperature was greater than the lower test temperature.
Acknowledgement The authors would like to express their gratitude to the referee whose critical comments made the presentation more consistent.

References
[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] S. Jin, J. Met. 45, 13 (1993). D. Hanson and E. J. Sanford, J. Inst. Met. 62, 215 (1938). J. Hwang, in Electronics Packaging and Interconnection Handbook (McGraw-Hill Book Co., New York, 1991). S. L. Hoyt, in Metals Properties (McGraw-Hill Book Com., New York, 1954). P. T. Vianco and D. R. Frear, J. Met. 45, 14 (1993). J. S. Hwang, Solder Paste in Electronics Packaging (Van Nostrand Reinhold, New York, 1989). A. A. Bahgat and H. EL-Bahnasawy, 3rd Internat. Scientific Conf., Faculty of Science, Al-Azhar University, 1998. W. D. Callister, Jr., Materials Science and Engineering An Introduction, 3rd ed. (John Wiley & Sons, Inc., New York, 1994), p. 130. T. B. Massalski et al., Binary Alloy Phase Diagrams, Vol. 2 (American Society for Metals, Ohio, USA, 1987), p. 2018. Yu. A. Geller and A. G.Rakhshtadt, Science of Materials (Mir Publishers Moscow, 1977). M. A. Mousa Imran, N. S. Saxena, D. Bhandari, and M. Husain, phys. stat. sol. (a) 181, 357 (2000). M. A. Meyers and K. M. Chawla, Mechanical Metallurgy Principles and Applications (Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1984). G. E. Dieter, Mechanical Metallurgy (McGraw-Hill Book Co., New York, 1986), p. 335. M. Braunovic, in: Proc. Internat. Symp. Sci. Hardness Testing and Its Research Applications, Detroit, USA, October, 1971. O. D. Sherby and P. M. Burke, Prog. Mater. Sci. 13, 325 (1967). J. Cadek, Creep in Metallic Materials (Elsevier, New York, 1988). V. I. Igoshev and J. I. Kleiman, J. Electron. Mater. 29, 244 (2000). N. F. Mott and F. R. N. Nabarro. Proc. Bristol Conf. Strength of Solids (Phys. Soc., London, 1948). N. F. Mott, Philos. Mag. 44, 742 (1953). J. Friedel, Dislocations (Pergamon Press, London, 1964). M. T. Mostafa, phys. stat. sol. (a) 163, 39 (1997). G. S. Al-Ganainy et al., phys. stat. sol. (a) 165, 185 (1998). J. H. Lau, Thermal Stress and Strain in Microelectronics Packaging (Van Nostrand Reinhold, New York, 1993). J. E. Bird, A. K. Mukherjee, and J. E. Dorn, Correlations between High-Temperature Creep Behavior and Structure, in: Quantitative Relation between Properties and Microstructure (Israel University Press, 1969), pp. 255342. E. G. Dieter, Mechanical Metallurgy (McGraw-Hill Book Co., New York, 1988). H. Mavoori, J. Chin, S. Vaynman, B. Moran, L. Keer, and M. Fine, J. Electron. Mater. 26, 783 (1997). W. Betteridge and A. W. Franklin, J. Inst. Met. 80, 147 (1951). J. Weertman, J. Appl. Phys. 26, 1213 (1955). A. H. Cottrell, An Introduction to Metallurgy (The ELBS Press, London, 1974), p. 150. G. A. Guy and J. J. Hren, Elements of Physical Metallurgy (Addision-Wesley Publishing Co., New York, 1974). R. K. Mahidhara, S. M. Sastry, Iwona Turlik, and K. L. Murty, Scr. Met. Mater. 31, 1145 (1994).

Anda mungkin juga menyukai