Sei sulla pagina 1di 6

Early View publication on www.interscience.wiley.

com (issue and page numbers not yet assigned; citable using Digital Object Identifier DOI (www.doi.org))

Cryst. Res. Technol., 1 6 (2010) / DOI 10.1002/crat.201000022

Effect of indium content and rapid solidification on microhardness and micro-creep of Sn-Zn eutectic lead free solder alloy
R. M. Shalaby*
Metal Physics Lab., Physics Department, Faculty of Science, Mansoura University, Mansoura, Egypt Received 10 November 2009, revised 16 December 2009, accepted 22 December 2009 Published online 29 January 2010 Key words lead free solder, Sn-Zn-In, microhardness, micro-creep, rapid solidification, melting point. The Sn-Zn alloys have been considered as lead-free solders. In this paper, the effect of 0.0, 0.5, 1.0, 1.5 and 2.0 wt.% Indium as ternary additions on melting temperature, structure, microhardness and micro-creep of the Sn-9Zn lead-free solders were investigated. It is shown that the alloying additions of Indium to the Sn-Zn binary system result in a suppression of the melting point to 187.9 C. From x-ray diffraction analysis, a new intermetallic compound phase, designated -In3Sn is detected. The formation of an intermetallic compound phase causes a pronounced increase in the electrical resistivity and mechanical strength. Also, an interesting connection between dynamic Youngs modulus and the axial ratio (c/a) of the unit cell of the -Sn was found in which Youngs modulus increases with increasing the axial ratio (c/a). The ternary Sn-9Zn-xIn exhibits creep resistance superior to Sn-9Zn binary alloy. The better creep resistance of the ternary alloy is attributed to solid solution effect and precipitation of In3Sn in the Sn matrix. The addition of small amounts of In is found to refine the effective grain size and consequently, improves hardness. The 89%Sn-9%Zn-2%In alloy is a lead-free solder designed for possible drop-in replacement of Pb-Sn solders.
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Introduction

Sn-Zn alloys have been expected to be one of the best alternative choices for a Sn-Pb eutectic solder because their melting temperatures are close to that of a Sn-Pb eutectic alloy. Microhardness measurement technique is a very sensitive technique to detect structure changes of different soft solders at different temperatures. Usually microhardness testing is a non-destructive testing but it leaves a small pit in the structure. Micro hardness testing can be the easiest way to determine the mechanical properties of the different phases of the structure [1]. The rule of thumb is the higher the hardness the higher is the mechanical strength. For soldering technology it is very important to study the microhardness of the structure because in soldering many soft and hard phases form, which are hard and brittle, and induce some hardness in the structure. Although eutectic alloys have high mechanical strength and considerable amount of hardness, introduction of third element to eutectic alloy can be an interesting study to compare. Investigations on microhardness of Sn-Zn based lead-free solder alloys as replacement of Sn-Pb solder has been reported [2]. Effect of cooling rate on hardness was also studied and compared among these solder pastes. The results show that the Sn-9Zn is the most sensitive to cooling rate. Indentation creep of lead-free Sn9Zn and Sn8Zn3Bi solder alloys was studied by [3]. Creep behavior of Sn9%Zn and Sn8%Zn3%Bi solders together with Sn37% Pb, as the material for comparison, was studied by indentation tests at room-temperature (T > 0.6Tm). The addition of indium to the Sn-Zn binary system improves the wetting characteristics of the alloy and lowers the melting temperature [4]. Creep, stress relaxation, and plastic deformation in Sn-Ag and Sn-Zn eutectic solders has been studied by H.Mavoori et al [5]. They were studied two eutectic lead-free solders are investigated for their creep and stress relaxation behavior. The creep tests were done in load-control with initial stresses in the range of 10-22 MPa at two temperatures, 25 and 80 C. In recent years, some researchers have worked on the Sn-Zn system [6]. New lead
____________________

* Corresponding author: e-mail: doctorrizk2@yahoo.co.uk


2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

R. M. Shalaby: Microhardness and micro-creep of Sn-Zn eutectic lead free solder alloy

free, Sn-Zn-In solder alloys have been studied by McCormack et al. [7]. It is shown that the alloying additions of In to the Sn-Zn binary system result in a suppression of the melting point. Lead-tin (Pb-Sn) solders are the most prominent joining materials for the interconnection and packaging of modern electronics because of their unique combination of low cost and convenient material properties. However, in view of increasing environmental and health concerns, alternative solder alloy systems need to be considered. For electronic parts and devices, solder joints provide electrical conductivity and suitable mechanical strength [8,9]. So that the aim of the present work to study the effect of indium additions and rapid solidification on the structure, melting temperature, microhardness and micro-creep of Sn-Zn eutectic lead free solder alloy.

Experimental

Sample preparation Five alloys of compositions Sn-9 wt.%Zn, Sn-9 wt.%Zn-0.5 wt.%In, Sn-9 wt.%Zn1.0 wt.%In, Sn-9 wt.%Zn-1.5 wt.%In and Sn-9 wt.%Zn-2.0 wt.%In were prepared by melting pure Sn, Zn and In (purity >99.99 %). Required quantities of the used metals were weighed out and melted in a porcelain crucible using an induction furnace at 600 C. After the alloys were molten, the melt was thoroughly agitated to effect homogenization. Then by ejecting the molten alloys on a rotating copper wheel (2900 rpm) which corresponds to a linear speed of 31.4 m/sec of the melt-spinning technique. The resulting alloys have long ribbons form of about 100 m in thickness and 1 cm width (constant width). Sample characterization X-ray diffractometer (XRD, X pert PRO, PANalytical using CuK target with secondary monochromator, in Central Metallurgical R and D institute, Technical Service Department, ElTebbin, Helwan, Cairo) was used to identify the structure of all produced alloys. The melting temperature of these alloys was determined by Differential Scanning Calorimetry [(DSC- 16, Setaram, France) in Technical Service Department, El-Tebbin, Helwan, Cairo] with a heating rate of 10 K/min. The measurement of resistivity is carried out by the double bridge method. The dynamic youngs modulus of melt spun alloys was examined by a dynamic resonance method. The produced samples were tested in a Vickers microhardness tester, where a diamond pyramid indenter with square base is used and the Vickers hardness number is given by HV=0.185F/d2 , where F is the applied load in N and d, is the average diagonal length in mm. Each reading was an average of a least ten measurements on the surface of the specimens. Micro-creep measurements as described elsewhere [10] were also carried out using a Vickers hardness tester using a fixed load of 0.49 N for dwell time up to 90 s.
Table 1 The details of the XRD analysis.
System Phase designation -Sn Zn -Sn Zn -In3Sn -Sn Zn -In3Sn -Sn Zn -In3Sn -Sn Zn -In3Sn Crystal system Crystal size (nm) 381.83 304 238.27 241.8 210 238.27 162.46 171 276.81 247.78 171.72 212.53 234.04 93.59 Typical parameter at concentration given wt.% In a () c () c/a 5.842 3.1799 0.5443 5.842 4.68 5.848 3.1855 0.545

Sn91Zn9 Sn90.5Zn9In0.5

Body centered tetragonal Hexagonal Body centered tetragonal Hexagonal Tetragonal Body centered tetragonal Hexagonal Tetragonal Body centered tetragonal Hexagonal Tetragonal Body centered tetragonal Hexagonal Tetragonal

Sn90Zn9In1.0

4.587 0.9801 3.1849 0.5446

Sn90Zn9In1.5

4.68 4.589 0.9805 5.84 3.1868 0.5456 4.682 4.59 0.9803 5.834 3.1871 0.5462 4.682 4.588 0.9799

Sn90Zn9In2.0

Results and discussion

Structure Figure 1 shows the x-ray diffraction patterns for the prepared alloys. The pattern for the eutectic Sn-9Zn alloy indicates two phases, a body centered tetragonal -Sn matrix phase and a secondary phase of hexagonal Zn. For alloys containing indium, the same phases appeared and three peaks observed for
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

Cryst. Res. Technol. (2010)

intermetallic compound -In3Sn at 2 = 43.17, 70.13 and 86.47. The In3Sn compound has -tetragonal structure. There is no others peak characteristic of an In phase, which means a complete solubility of In in the Sn matrix. The variation of the axial ratio (c/a) for -Sn with the variation of In concentration are listed in table 1. The axial ratio increases to maximum value 0.5462 at 2 wt.% In. The crystalline size is determined from the XRD pattern by using Scherers equation Dhkl=0.9/hklcos, where Dhkl is the crystalline size, is the wavelength of Cu K =1.54056 , is the reflection angle and hkl is the full width at half maximum (FWHM)[11]. Addition of small amount of indium refines the effective grain size while retaining uniform distribution of In3Sn precipitates in the solidification microstructure. The details of the XRD analysis are listed in table 1.

Fig. 1 The x-ray diffraction (XRD) patterns of melt spun Sn-9Zn-In alloys.

Fig. 2 The differential scanning calorimetry (DSC) melting curves for melt-spun alloys.

Thermal properties Figure 2 shows the differential scanning calorimetry (DSC) exothermic peaks of Sn9Zn-xIn (x=0, 0.5, 1.0, 1.5 and 2 wt.%) solder alloys during heating with heating rate 10 K/min. From this figure the melting point Tm, enthalpy of fusion (Hf) and pasty range of these alloys are calculated and presented in table 2. Differential scanning calorimetry (DSC) of the Sn-9Zn-xIn alloys showed small melting point depressions in the ternary, Indium-containing alloys when compared to the melting temperature (198 C) of the binary Sn-9Zn eutectic alloy. The 1wt. % In alloy exhibits a melting point suppressed to 195.16 C. Further alloying modifications using up to 2 wt. % of Indium have been found to suppress the melting point to as low as ~ 187.9 C. This decrease can be attributed to the dissolution or precipitation of In phase as a lower melting point phase than that of the eutectic matrix. The pasty range which is the range between the solidus and liquidus temperatures can also be calculated. The results showed that the pasty range of the Sn-9Zn alloy is equal to zero, while other alloys exhibited an increase of the pasty range to 4.84 C. This increase indicated that the additions of Indium up to 2 wt.% to the Sn-9Zn eutectic alloy converted the eutectic alloy to hypoeutectic alloys.
Table 2 Thermal analysis. Alloy Melting point (C) Sn-9Zn 198.00 Sn-9Zn-0.5In 195.96 Sn-9Zn-1.0In 195.16 Sn-9Zn-1.5In 191.93 Sn-9Zn-2.0In 187.90
www.crt-journal.org

Pasty range (C) zero 4.06 4.84 4.84 4.03

Enthalpy (v.s/mg) 23.0030 20.3284 18.7834 20.1012 20.1090


2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

R. M. Shalaby: Microhardness and micro-creep of Sn-Zn eutectic lead free solder alloy

Microhardness Figure 3 shows the variation of Vickers hardness number with the applied load for Sn9Zn-xIn (x=0, 0.5, 1, 1.5 and 2 wt. %) solder alloys. It shows that, the Vickers hardness number of all alloys varies non-linearly with the applied load in the same manor. This type of non-linear behavior agrees with the results observed in [12-15] and known as the indentation size effect [16]. Figure 4 shows the variation of microhardness Hv with Indium content for different applied loads, 0.098, 0.245, 0.49, 0.98, 196 and 2.94 N. The hardness number increases with increasing Indium content up to 352.8 MPa at 2 wt.% In. The increase of hardness number can be attributed to presence of intermetallic compound -In3Sn and refinement the crystalline size by addition of Indium and effect of rapid solidification.

Fig. 3 Vickers micrihardness values versus the applied load of Sn-9Zn-In alloys.

Fig. 4 The variation of Vickers microhardness (Hv) with Indium content.

Fig. 5 Variation of Vickers microhardness with indentation time of Sn-9Zn-In alloys.

Fig. 6 The creep behavior of as-quenched melt-spun Sn-9Zn-In alloys.

Micro-creep Figure 5 shows the variation of Hv versus indentation time. It shows continuous decrease of Hv with time. Since the hardness is given by: Hv=0.102F/S MPa where F, is the applied load in N, and S is the surface area of the indentation mm2. Since the applied load is fixed at 0.098 N, the decrease of Hv is therefore due to the increase of area. The increase of area (strain) here is considered as the fractional change in area and is given by: Strain () =S/S0 where S is the increase in surface area and S0 is the original area taken at the lowest of the indentation time. The measurement of creep by this method is termedmicro-creep the term micro-creep comes from applying the stress to a very small area comparable to the area of a grain, and it is observed by microscope. Plotting the strain with indentation time, we obtain a typical creep curve, as shown in
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

Cryst. Res. Technol. (2010)

figure 6. The creep resistance of Sn-9Zn is higher than that of the Sn-9Zn-xIn alloys. It is known as the microcreep curve that differ than the ordinary creep curve in which it shows two of the three stages of the ordinary creep behavior. The third stage, which is the fracture stage, is not obtained here since it is impossible because this method is a compression method not tensile method and different from the tensile method in which the fracture stage is not occurred. The first stage records a fast increase of strain with time of the indentation starts from beginning up to 40 s of indentation time. The second stage represents a slow increase region in which the strain increases by lower rates for all alloys. The high creep resistance of indium containing alloy improved more than an order of magnitude. Strengthening is attributed to a substantial refinement of precipitates in the solidification microstructure consequently, significantly improved creep resistance. Also, the better creep resistance of the ternary alloy is attributed to solid solution effect and precipitation of In3Sn in the Sn matrix. Mechanical properties and electrical resistivity Table 3 shows the variation of dynamic Youngs modulus with indium content. It shows an increase of the youngs modulus to the 54 MPa at 2wt.% In. This observed results of dynamic Youngs modulus with Indium concentration can be explained in terms of the variation in the c/a ratio of -Sn. We noted that the Youngs modulus E increases with increasing the c/a ratio. The increase in the c/a ratio means the stretching of the unit cell along the c-axis, this modification in the shape of the unit cell of the -Sn matrix may result in an increase in the Youngs modulus. In microelectronic devices the solder serves as an electrical interconnection. The electrical resistivity () of as-quenched melt spun alloys is measured at room temperature and the results are shown in Table 3. The resistivity increases significantly with increasing indium content and attains the highest value of about 17.6x10-6 ohm.cm. Such a high electrical resistivity is considered to have been attained by the combination of several affects; the uniform distribution of fine precipitates the introduction of internal defects such as dislocations and the formation of the intermetallic compound In3Sn. The intermetallic compound In3Sn atoms act as scattering centers in -Sn matrix.
Table 3 Dynamic Youngs modulus and electrical resistivity. Alloy Dynamic Youngs modulus (MPa) Sn-9Zn 45 2 Sn-9Zn-0.5In 46 1.8 Sn-9Zn-1.0In 48 1.3 Sn-9Zn-1.5In 50 1.5 Sn-9Zn-2.0In 54 2 Electrical resistivity 10-6 .cm 14.0 0.5 14.8 0.6 15.7 0.3 16.8 0.5 17.6 0.4

Conclusions

I have investigated the properties of Sn-Zn-In solders and found the following: 1. The addition of In to Sn-9Zn eutectic alloy lead to the formation of new intermetallic compound phase In3Sn. 2. An interesting connection exists between Youngs modulus and the axial ratio (c/a) of the unit cell of the Sn, since it has been found that, by increasing the axial ratio, Youngs modulus increases and vice versa. 3. Addition of small amount of indium to Sn-9Zn eutectic solder refines the effective grain size. 4. The ternary alloy Sn-9Zn-2In has the most suitable properties required for solder applications as a replacement of Sn-Pb eutectic alloy; it has a lower melting point 187.9 C, which is very close to that of Sn-Pb solder alloy and a higher value of the Vickers microhardness, 352.84 MPa at 0.098 N, 5 s. 5. The ternary Sn-9Zn-2In lead free solders (alternative solder) used for electronics parts and devices because they have suitable electrical resistivity and mechanical strength. 6. The microcreep measurement from microhardness is a powerful tool in the determination the creep characteristics of alloys. Usually microhardness testing is a non-destructive testing. Also, the creep resistance of Sn-9Zn-xIn is higher than that of the Sn-9Zn is due to refinement of precipitates in the solidification microstructure.
Acknowledgments I express my deep gratitude and appreciation to Professor Mustafa Kamal, Head of Metal Physics Group, Physics Department, Faculty of Sience, Mansoura University, Mansoura, Egypt.

References
www.crt-journal.org
2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

6 [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16]

R. M. Shalaby: Microhardness and micro-creep of Sn-Zn eutectic lead free solder alloy I. Shohji, T. Nakamura, F. Mori, and S. Fujiuchi, J. Mater. Trans. 43, 1797 (2002). R. A. Islam, B. Y. Wu, M. O. Alam, Y. C. Cham, and W. Jillek, J. Alloys Comp. 392, 149 (2005). R. Mahmudi, A. R. Geranmayeh, H. Khanbareh, and N. Jahangiri, Mat. Des. 30, 574 (2009). M. McCormack and S. Jin, J. Electro. Mater. 23, 123 (1994). H. Mavoori, J. Chin, S. Vaynman, B. Moran, L. Keer, and M. Fine, J. Electro. Mater. 26, 123 (1997). C. M. L. Wu, M. L. Huang, Y. C. Chan, and J. K. L. Lai, J. Electro. Mater. 29, 1515 (2000). M. McCormack, S. Jin, and H. S. Chen, J. Electron. Mater. 23, 687 (1994). R. M. Shalaby, J. Alloys Comp. 480, 334 (2009). M. L. Huang, T. Loeher, A. Ostmann, and H. Reichl, Appl. Phys. Lett. 86, 181908 (2005). T. El-Ashram and R. M. Shalaby, Electron. Mater. 34, 212 (2005). B. D. Cullity, Elements of X-ray Diffraction, 2nd Edition (Addison-Wesely, 1978), p. 248. R. M. Shalaby, J. Mater. Electron. 16, 187 (2005). J. Kyo Kim, K. Do-Hyung, and P. Hwang, J. Mater. Sci. 35, 4185 (2000). J. Gong, H. Miao, Z. Zhao, and Z. Guan, Mater. Sci. Eng. A 303, 179 (2001). R. Tickoo, R. P. Tandon, K. K. Bam Zai, and P. N. Kotru, Mater. Chem. Phys. 80, 446 (2003). O. Uzun, T. Karaaslan, M. Gogebakan, and M. Keskin, Alloys Comp. 376, 149 (2004).

2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.crt-journal.org

Potrebbero piacerti anche