Vous êtes sur la page 1sur 12

Lead-free Solder

1. Introduction
The motivation for developing lead-free solders is to remove Pb from the electronic manufacturing and waste disposal processes, as its toxicity is well established. It is not practical to recycle lead as is done in other industries, because the amounts are small compared to the size of systems it is in. While shredded electronic waste is often put in landlls in the US, this waste is incinerated in more densely populated areas, which puts lead vapor into the air. Moreover, in the manufacturing environment, the vapor pressure of Pb is much higher than most other metals, posing a direct inhalation hazard to workers. In contrast, Sn has a much lower vapor pressure, and is somewhat less toxic than Pb. For this reason, European Union (EU) legislation passed the Waste from Electrical and Electronic Equipment (WEEE) and the Restriction of the Use of Certain Hazardous Substances in Electrical and Electronic Equipment (RoHS) directives in 2002. The four heavy metals (lead, cadmium, mercury, and hexavalent chromium) and the brominated ame retardants, polybrominated bypheneis (PBB) and polybrominated diphenyl ether (PBDE), were banned in new electronic equipment in the EU as of 1 July 2006 with some exemptions, for example, lead in high melting temperature solders (i.e., SnPb solder alloys containing more than 85% lead, or lead in solders for network infrastructure equipment for switching, signal transmission, as well as network management for telecommunication). However, this exemption is not permanent and will be reviewed. Exempted uses of Pb-bearing solder must have acceptable substitutes by 2014. This multinational decision has led to vigorous development of alternative solder alloys largely based upon Sn, and the differences in properties affect the entire technology stream of printed circuit boards and packages. As tinlead (SnPb) solder has been effective for thousands of years, there is a deep experience base which is lacking for Pb-free solders. Much of the existing electronic system infrastructure is based upon properties of PbSn solder. Solders perform effectively at 80% of its melting temperature, giving it a performance capability comparable to superalloys used in jet engines, but the performance demands on solder joints are continually increasing due to increasing current density arising from continuing miniaturization. SnPb solder has a two-phase microstructure with about a 2:1 volume ratio of Sn:Pb-rich phases. The very soft Pb-rich phase carries much of the cyclic deformation, making the alloy soft. In contrast, Sn-based solders have about 5%

small, hard, intermetallic strengthening particles, giving it completely different metallurgical characteristics than SnPb alloys, because all of the plastic deformation takes place in the Sn phase. As Sn is harder than Pb, Sn-based solders are stronger than SnPb solders, which stresses the surrounding components more than SnPb solders do. This is problematic because existing die and circuit board design is based on the softer properties of SnPb and its lower melting temperature. While a drop in replacement for SnPb has been sought, it has not been found. As Sn-based eutectic alloys have only a small amount of other elements, the eutectic temperatures are close to the 231 1C melting point of Sn; 220 1C is a typical eutectic temperature for Sn-based alloys. These differences from SnPb are signicant, making the experience base with SnPb solder nearly irrelevant; hence, coping with these differences has required a number of design adjustments throughout the electronic manufacturing infrastructure. Without any deep experience base, engineering design decisions require increased reliance on analytical modeling to predict product reliability. In the following sections, the fundamental information required for analytical modeling of Sn-based solders is described to show what must be incorporated into predictive models. In addition, challenges to solder technology that were not important when SnPb was common are examined, such as high current density and electromigration that comes with shrinking size scales, and problems with formation of tin whiskers, which form spontaneously and can cause shorts between circuit paths. Finally, the infrastructural requirements for manufacturing are discussed to illustrate how implementation of leadfree solders is accomplished.

2. Physical Properties of Tin


Tin is a group IV element that is transitional between the diamond cubic structure of Si and Ge, and the face-centered cubic structure of Pb. No other metal has a body-centered tetragonal structure crystal structure like b Sn, which is illustrated in the inset of Fig. 1(a). The tetrahedral bonding around the center atom is evident, and if extended to neighboring unit cells, it becomes apparent that the crystal has a squashed diamond cubic structure. The b (white) tin transforms to a (gray) tin below 13.2 1C, which has the diamond cubic structure. In practice, the presence of alloying elements stabilizes the b structure, but at low temperatures and long times, this transformation (called tin pest) is occasionally observed, which can cause internal fracture that causes a solid to become a pile of powder. Table 1 shows how Sn compares with several other important elements used in electronic systems. Note that the coefcient of thermal expansion (CTE) varies 1

Lead-free Solder
(a) 70 c 60 [001] c (101) (111) E, GPa, Rayne 61 CTE, ppm C1 (001) (100) a 40 c /a = 0.5456

50

30 [111] 20 [101]

10 a [100] 0 (b) Youngs modulus (GPa) Modulus 001 , c 110 111 E(001) / E(100) CTE c /a ratio or 80 70 60 50 40 CTE (ppm K1) or 30 20 CTE, a 10 0 50 101 100 , a CTE, c 6 5 4 3 2 1 10 20 30 40 50 [110] 60 70

50

100

150

200

Temperature (C)

Figure 1 (a) Anisotropic CTE and elastic modulus for tin and the body-centered tetragonal crystal structure; the magnitude is the distance from the origin. (b) Temperature dependence of CTE and elastic modulus. From Bieler T R, Jiang H, Lehman L P, Kirkpatrick T, Cotts E J, Nandagopal B 2008 Inuence of Sn grain size and orientation on the thermomechanical response and reliability of Pb-free solder joints. IEEE Trans. Compon. Packag. Technol. (CPMT) 31 (2), 37081; House D G, Vernon E V 1960 Determination of the elastic moduli of tin single crystals and their variation with temperature. Br. J. Appl. Phys. 11, 2549; and Deshpande V T, Sirdeshmukh D B 1962 Thermal expansion of tin in the betagamma transition region. Acta Cryst. 15, 294.

Lead-free Solder
Table 1
Comparative physical and mechanical properties of elements in electronic materials. Property CTE (ppm K ), B300 K Electrical resistivity (mO cm) Thermal conductivity (W m 1 K 1) T(melt) (1C) Entropy of melting ( J mol 1 K 1) Elastic modulus (GPa) Yield strength (MPa) Ultimate tensile strength (MPa)
1

Pb 29.1 20.8 33 327.5 8.3 1040 o5 18

Sn 1530 1320 63.2 232 14.0 2269 10 220

Ge 6.1 50 64 937.4 28.7 103155 150

Si 2.49 10 000 124 1412 30.0 131196 270

Al 24 2.7 210 660.4 11.2 6476 15 182

Cu 16.4 1.7 385 1083 9.6 67192 33 210

Ag 19.6 1.55 419 962 9.2 44120 140

Au 14.4 2.2 301 1064 9.7 43116 120

Where ranges are given for Sn, the rst number is for the a direction, the second the c direction.

by a factor of 2 with crystal direction (illustrated in Fig. 1(a)), unlike cubic metals that have an isotropic CTE. The CTE on the (001) plane is close to that of Cu (and the FR-4 polymer matrix glass ber composite circuit boards). Also, the elastic modulus varies by a factor of 3. These properties are strongly temperature dependent, as shown in Fig. 1(b); the CTE anisotropy ratio is nearly constant at B2 with temperature, but the elastic modulus anisotropy ratio varies strongly with temperature. Tin has lower electrical resistivity and higher thermal conductivity than Pb, which are benecial for solder joints, and particularly important for high-current-density applications. The entropy of melting (DSm DHm/Tm) is between the low values common to cubic metals and higher values for diamond cubic elements. This entropy change is important because much more energy must be released to convert amorphous liquid to an orderly crystal lattice. Consequently, a moving solidication front will generate a heat wave as it moves. In practice, solidication of Sn requires more undercooling than other metals, and in the small volumes of solder joints, 501 of undercooling is commonly observed. With such undercooling, a critical size nucleus will grow quickly, so joints often have only one or a few crystal orientations.

(a)

As-assembled, no aging Sn

Ag3Sn

SAC305 (b) Sn

5 m

150 C/1000h

Ag3Sn

SAC305

3. Role of Alloying Elements on Microstructure


As tin has relatively low solubility for alloying elements, the properties of the Sn-based solder matrix will be similar to the physical properties of tin; however, the mechanical properties of a joint are greatly affected by the spatial arrangement of secondphase precipitates that form due to the presence of alloying elements. The most common alloying elements are Ag and Cu. A wide variety of compositions near the ternary eutectic composition of Sn 3.7Ag0.9Cu have been explored, which melts at 217 1C. An alloy numbering system has been established; for example, SAC359 denotes the eutectic composition and SAC405 has more silver and less

5 m

Figure 2 After assembly, Ag3Sn particles in this SAC 305 (Sn3.0Ag0.5Cu) joint are clustered in the interdentritic spaces (a), and after aging, they coarsen (b), making the ow stress decrease with time.

copper than the eutectic, that is, 4.0Ag and 0.5Cu. The most popular alloy is SAC305, but there is growing interest in SAC alloys with lower Ag. The solidication microstructure is extremely sensitive to alloy composition, volume of the solder joint, and the heating and cooling history, which dramatically affect the size and morphology of 3

Lead-free Solder

Figure 3 Microstructures of Sn1Ag, SAC305 (Sn3.0Ag0.5Cu), and Sn0.1Cu solder joints illustrating a range of microstructure morphologies of the tin phase. Adapted from Lehman L P, Xing Y, Bieler T R, Cotts E J Cyclic twin nucleation in tin based solder alloys. Acta Mater. (submitted).

0.25 Fraction, (1 bins) 0.2 0.15 0.1 Sn1Ag SAC 305

100 Common [100]

30-50

001 0.05 0 0 15 30 45 60 75 90 100 m 7 Misorientation (deg) <60 twin >60

Figure 4 Misorientation histograms of Sn1Ag and SAC305 solder balls in Fig. 3. Pole gures and unit cell prisms illustrate beach ball texture for the SAC305 obtained from the EBSP map of the SAC305 ball. Colored lines indicate boundaries and their misorientations.

intermetallic phases as well as the microstructure of the b tin phase. After solidication, the most common microstructure is a divorced eutectic with primary Sn dendrites as the primary phase and an interdendritic ternary eutectic that forms between dendrites shown in the top image of Fig. 2. Figure 3 illustrates some extremes in observed Sn microstructures with different amounts of Ag and Cu. With no copper, the microstructure appears to be a ne-grained polycrystal, but with added copper, single crystals or tricrystals are commonly observed; sometimes, the tricrystal has the morphology of six segments like a beach ball. From electron backscatter pattern (EBSP) mapping, crystal orientations for both compositions show that only three crystal orientations are present, with B601 misorientations about a common [100] axis. When no silver is present, 4

three plates B601 apart often form but subsequent crystals have a variety of other orientations, and these joints are more like a polycrystal. Figure 4 shows that a misorientation histogram is rather similar for the Sn1Ag and SAC305, despite the very different microstructures. Both joints have a beach ball texture in pole gures, such as those shown from the EBSP map of the SAC305 joint in Fig. 3. Such pole gures are common regardless of whether or not the grain morphology appears like a beach ball. The major difference between Sn1Ag and SAC305 is that there are few low-angle boundaries in the SnAg, due to the small grain size, and many low-angle boundaries in the large grains in the SAC305. The preference for 601 misorientations is a signature of solidication twinning (two types occur, on {101} or {301} planes with 57.21 and 62.81, respectively).

Lead-free Solder
The boundaries of twin structures have a very low interfacial energy, which makes them highly preferred. Solidication twinning is also the basis for the apparently ne-grained microstructure shown for the Sn1Ag composition, which actually has three twinned orientations that interpenetrate each other. The heating and cooling history also causes much variability in observed microstructures. If cooling rates are slow and/or undercooling is large, then precipitation of large plates of Ag3Sn and/or rods of Cu6Sn5 can form (e.g., Fig. 5(a)). Upon remelting (reow), these large intermetallic phases require a long time to dissolve, such that these plates become sinks for alloying elements in solution, which locally alters the density of small particles of the Ag3Sn that form in the nal solidication of eutectic liquid (Fig. 5(b)). These large precipitates degrade mechanical properties. As the solidication path is complex, other alloying elements can signicantly modify the solidication path, and hence, the microstructure and properties. Reducing the amount of undercooling can be achieved by microalloying SAC with very small amounts of Cr, Ti, Ni, Co, Zn, Mn, Fe, Ge, Si, Bi, and/or In. Furthermore, Sn forms a number of other eutectic alloys with In, Bi, and Zn, which have lower melting points than SAC alloys (e.g., Sn8Zn3 Bi has 197 1C melting temperature). In and Bi are more expensive, and large amounts of Zn lead to corrosion problems, which make these eutectic systems less popular than SAC-based alloys. SnBi and SnIn eutectic alloys
60 Polyslip, less recovery 50
Shear stress (MPa)

Ag 3Sn Cu6Sn5

10 m

25 C0.001 s1 25 C0.1 s1 150 C0.001 s1 150 C0.1 s1 Sn3.5Ag on Cu

40 30 20 10 0

Figure 5 Morphology of intermetallic precipitates: (a) Sn etched away reveals shape of Ag3Sn plates and Cu6Sn5 rods. (b) Appearance of a small Ag3Sn plate within a b Sn dendrite, and surrounding small particles of Ag3Sn. (a) Reproduced with permission from Lu H Y, Balkan H, Ng KYS 2005 Solidliquid reactions: the effect of Cu content on SnAgCu interconnects. JOM 57 (6), 305, copyright The Minerals, Metals & Materials Society (TMS). (b) Reproduced with permission from Sigelko J, Choi S, Subramanian K N, Lucas J P, Bieler T R 1999 Effect of cooling rate on microstructure and mechanical properties of eutectic SnAg solder joints with and without intentionally incorporated Cu6Sn5 reinforcements. J. Electron. Mater. 28 (11), 11848, copyright The Minerals, Metals & Materials Society (TMS).

Slip, balanced by recovery


0 1 2 Simple shear 3 4

Figure 6 Single lap shear deformation of solder joints depends on temperatures and strain rate. Data from Rhee H, Subramanian K N, Lee A, Lee J G 2003 Mechanical characterization of Sn3.5Ag solder joints at various temperatures. Soldering and Surface Mount Technology 15 (3), 216.

Lead-free Solder
have even lower melting points that make these solders valuable when building complex systems, where remelting of prior solder joints is undesirable (e.g., within packages to prevent damage from volume expansion from melting where joints are constrained by a ller material). decreases the stress needed to cause dislocations to bypass the particlessee Fig. 2. Alloying affects both the as-fabricated strength and its decrease in strength with aging time. As the elastic and CTE properties are strong functions of temperature and crystal orientation, the stress evolution in a solder joint starts with the cooling process; as the joint reaches room temperature, signicant internal stresses are present, which will relax slowly, as Sn is above half of its melting temperature even at 40 1C. Hence, modeling of stressstrain behavior requires sophisticated models that can track solidication processes that dene the initial microstructure, internal stress, and their evolution during service.

4. Mechanical Properties of Solder Joints


Some mechanical properties of tin, in comparison with other materials used in packages are shown in Table 1. Clearly, Sn is much stronger than Pb, implying that lead-free solder joints are stronger, which is not always desired, as greater strength transfers higher loads to substrates and packages. Figure 6 illustrates how temperatures and strain rate affect the shear stressstrain properties of SAC alloys. At low temperatures, the yield strength is similar at different strain rates, but more work hardening occurs at higher strain rates, due to the lack of time for diffusion-assisted dislocation recovery (annihilation) processes. When recovery is possible, there is less work hardening, which reduces the ow stability, leading to heterogeneous strains and reduced strain to failure. At higher temperatures and lower rates, recovery occurs concurrently with slip, leading to almost no hardening at all. Figure 7 shows how the maximum shear strength depends on aging time at elevated temperature for several alloy compositions. Aging causes coarsening of the ne-scale precipitates, which increases the distance between particles that correspondingly
50

5. Interfacial Properties and Intermetallic Formation


When shear failures are studied, the fracture surface is commonly entirely in the tin, even though the crack propagates close to the interface. This is a consequence of the geometry of the solder joint, as maximum stress states usually develop close to the interface. However, at high strain rates and/or lower temperatures, the Sn becomes stronger than the intermetallic phase boundaries, and fracture occurs in the intermetallic layer that forms between the solder and the package or circuit board. Thus, the evolution of the interfacial intermetallic phases also affect the reliability of a solder joint. Figure 8 shows how aging leads to growth of the intermetallic layer

45
Maximum stress (MPa)

Sn3.5Ag Sn3.5Ag1.0Cu Sn3.9Ag0.6Cu Sn3.7Ag0.6Cu0.3Co

Sn3.0Ag0.5Cu Sn3.7Ag0.9Cu Sn3.7Ag0.7Cu0.2Fe

40

35

30

25

20 0 100 200 300 400 500 600 700 800 900 1000 1100 Hours at 150C

Figure 7 Effect of alloying and aging on maximum shear strength. With kind permission from Springer Science Business Media: J. Mater. Sci.: Mater. Electron., Development of SnAgCu and SnAgCuX alloys for Pb-free electronic solder applications, Vol. 18, 2007, pp. 5576, Anderson I E.

Lead-free Solder
(a) As-assembled, no aging

6. Shock (High Strain-rate Deformation)


For many portable electronic systems, drop testing is important, as the properties of solders are sensitive to strain rate (Fig. 6). At lower strain rates, dislocation motion and recovery (annihilation) are assisted by thermal vibration, which is a time-dependent statistical process. At higher rates, motion of dislocations is not assisted by thermal vibration, so the ow stress increases with strain rate, often leading to fracture at interfaces rather than through the solder. This undesired increase in strength of the solder can be mitigated by having fewer particles that strengthen the solder (aging coarsens the particles, but also thickens the brittle intermetallic layer). SAC 105 has a lower yield stress, and hence a lower strength in shock conditions, that makes this solder more attractive for applications that must survive shock (but this comes at the cost of a 10 1C higher melting temperature). Microalloying with Ni and Cr can improve drop resistance, as they slow intermetallic layer growth, which reduces the brittleness of the intermetallic layer. Figure 9 shows how high strain-rate properties are sensitive to alloying and hence microstructure and interfacial reactionssome alloy compositions show much greater variability than others.

Cu3Sn

Cu

Cu6Sn5 SAC305 (b)


5 m

150 C/1000 h

Cu3Sn

Cu

Cu6Sn5 SAC305 Ag3Sn (light gray)


5 m

7. High-current-density Issues
With decreasing sizes of electronic packages and systems, the current density increases in solder joints. Moreover, the under bump metallization (UBM) and solder bump contact areas become smaller; ipchip solder bumps are in the range of 50100 mm, and as solder balls need to carry currents of about 0.2 A, the current density reaches 104 A cm 2. Thus, electromigration is an increasingly important issue, as high DC currents provide an electron wind that moves Sn atoms preferentially in the direction of electron ow. Increased current density causes joule heating that increases the temperature and softens the solder. Voids commonly develop on the side from which electrons enter, and hillocks develop where electrons depart. However, the imposition of the electric eld puts a bias on the direction of probable vacancy (atom) jumps, which alters rate processes that depend on diffusion, such as creep deformation. Electric currents can cause changes in grain orientation to orient the c-axis with the direction of current ow; this driving force arises from the fact that the c/ a ratio of Sn self-diffusion and resistance is 1.53 and 0.38, respectively (resistance is higher and diffusion is slower in the c direction). Hence, grains with the c-axis aligned with the current ow will experience a driving force to rotate. Interstitial diffusivity is also highly anisotropic; Cu interstitial diffusion is 500 times faster along the c-axis than the a-axis (and Ni diffuses 70 000 times faster along the c-axis). These 7

Figure 8 (a) A joint from a ball grid array package made from SAC30 (Sn 3.0Ag) solder with SOP surface nish (solder on pad). (b) After aging, the Cu6Sn5 intermetallic layer grew thicker with time and temperature, and a Cu3Sn layer formed between the Cu6Sn5 and copper layer.

between solder and copper, and how a Cu3Sn layer develops between the Cu6Sn5 and the copper layers with increasing time. This interface is often problematic, as voids often develop in the Cu3Sn layer or interface. There are several ways that fracture can be facilitated due to reactions in the interfacial layer. The Sn solder and the substrate metallurgy form a diffusion couple, and the reaction layer develops quickly when the solder is molten. The rate of the reaction is highly sensitive to the composition of the solder and the substrate, which is partially dissolved by the molten Sn, which then alters the alloy composition near the interface. Interfacial reactions continue during service in the solid state, leading to complex composition gradients and continuing reactions at rates that depend on the temperature history and composition. There is much literature that examines the kinetics of phase evolution of interfacial reactions, and the outcomes are often specic to the elements present and the thermomechanical history.

Lead-free Solder
4000 Impact strength (J m2) 3500 3000 2500 2000 1500 1000 500 0 SAC + SAC + SAC + SAC + SAC + SAC + SAC + SAC + SAC + SAC SAC SAC SnSAC Si Cr Co (305) (379) (396) 3.5Ag (3610) Zn Mn Ni Ti Ge Fe (Eut) Composition [4] [3] [4] [3] [3] [3] [3] [3] [3] [3] [3] [3] [3] [3]

Figure 9 Impact strength of several solder alloys. With kind permission from Springer Science Business Media: J. Mater. Sci.: Mater. Electron., Development of SnAgCu and SnAgCuX alloys for Pb-free electronic solder applications, Vol. 18, 2007, pp. 5576, Anderson I E.

effects make tin grains with the c-axis aligned with current ow particularly prone to early damage by transport of underbump metal atoms in the direction of electron ow. Current crowding occurs in locations where the path length of mobile electrons is concentrated, leading to pileups of Sn or alloy atoms whose transport is blocked by the intermetallic boundary. This leads to hillock or whisker formation on the side where electrons leave the joint. Highfrequency RF currents cause electron ow to be on the surfaces due to the skin effect, so the magnitude of damage is smaller.

8. Tin Whisker Formation and Mitigation


Tin whiskers commonly form on substrates coated with tin. These are problematic, as their growth can cause shorts between current carrying lines. This problem is a consequence of Sn being in a state of compression, which results from copper diffusion into Sn, which causes a volume increase from the formation of intermetallic compounds. Grain boundaries provide fast paths that allow Sn atom ux to locations where the surface oxide is broken, so that an emerging whisker can relieve the pressure imposed by a growing intermetallic. The problem can be reduced or eliminated by preventing Cu diffusion into Sn, and managing the internal stress state so that the surface of the Sn remains in tension rather than compression.

9. Manufacturing and Technology Issues


As shown in Fig. 10, Pb-free solder materials are used in various parts of interconnections in electronic 8

packages. The rst-level interconnectis between the chip and the package substrate and lead-free solder bump material are used in the case of a Flip Chip BGA (FCBGA). Because under ll material is used in this level, thermal fatigue and the mechanical shock is a lesser concern, but electromigration is an important issue because of the higher function temperature at the chip and higher current density. The second-level interconnect is the solder joint between the package substrate and the printed circuit board (PCB) or printed wire board (PWB). Based on the packaging structure and form factors there are several families of packages, each of which have particular challenges with lead-free solder. For example, chip scale packages (CSP) and wafer level packages (WLP) have relatively small body size but have a high CTE mismatch, plastic ball grid array (PBGA) packages are one of the most popular package types which have a plastic overmold over the wire bonded chip, and there are FCBGA packages which have relatively large body size and usually have a heat spreader on top of the package. To assure their reliability, where reliability of the solder joint is the ability to function during service conditions for a specic period of time without exceeding acceptable failure levels, several driving forces which alter the structural stability of the package during time must be considered. One of the driving forces is the thermal expansion difference between the package and the PCB. Figure 11 shows orientation maps of three joints that illustrate the effect of CTE on internal stress history. As solder joints tend to have a random set of orientations, the average CTE denes the amount of displacement of the package from the board. As the red orientation has the c-axis nearly parallel to the substrate, its low

(a) Die size Printed circuit board (PCB) Solder ball (2nd level interconnect)

Package body size

Solder bump (1st level interconnect) Solder ball (2nd level interconnect) Underfill material

Plastic ball grid array (PBGA) Silicon chip die Thermal interface material (TIM)

(b)

Package substrate

Flip chip ball grid array (FCBGA)

Printed circuit board (PCB)

Figure 10 (a) Schematic diagram of a conventional plastic ball grid array (PBGA) package and (b) a ip chip ball grid array (FCBGA) package.

Lead-free Solder

+ 100 m Crack 0 c PWB

Temp c PWB

c-axis direction Time

Figure 11 Three orientation maps showing orientation of c-axes in three joints, one of which cracked after 2500 thermal cycles. The plot shows how stress state varies for three single crystals with different crystal orientations (Bieler et al. 2008).
99 Weibull SnAgCu/SnAgCu W2 RRXSRM MED F = 24 / S = 1 SnAgCuBi/SnAgCu 50 Cumulative percent W2 RRXSRM MED F = 23 / S = 2 SnPb/SnPb W2 RRXSRM MED F = 50 / S = 0

90

10

Early failures

1 100 1000 Cycles 10 000

Figure 12 Weibull plot shows statistics of failure; points far to the left of the linear regression are identied early failures. The cause of early failures is not yet identied. Reproduced with permission from Hillman D, Wilcoxon R 2006 JCAA/JG-PP lead free solder testing for high reliability applications. As originally published in the 2006 SMTA International Conference Proceedings. Surface Mount Technology Association, Edina, MN, pp. 83660.

CTE perpendicular to the board will cause the joint to be in tension at high temperature and compression at low temperature. The purple orientation, which has the c-axis perpendicular to the board, will have 10

the opposite stressstrain history. In joints with the green orientation, or in joints with multiple orientations, the CTE will more likely be closer to the average value, and hence have smaller stresses

Lead-free Solder
perpendicular to the board. The red orientation has been found to be more likely to crack during thermal cycling, but the blue orientations are more likely to fail under high current conditions. Thus, it is desirable to avoid these two extreme orientations, but to date, there is no obvious way to control the crystal orientations during solidication. The rate of cycling is important, as accelerated thermal cycles (or thermal shock tests) do not allow stress relaxation to occur, and hence, the magnitude of plastic cyclic strain is lower. It is well established that the number of cycles to failure decreases as hold times for a thermal cycle are increased, implying that damage develops during hold times. Mechanical shock and vibration are especially important for Pb-free solder materials. These loading conditions affect the stability of the solder joint individually or as a system, where degradation and ultimately failure is accelerated. To estimate the overall lifetime of the solder joint, there are several statistical approaches. Failures in solder joints, in particular due to constant degradation do not happen simultaneously and are usually well-distributed over time. One of the common methods of estimating the expected lifetime of solder joints statistically uses the Weibull distribution. Two parameters are used in this method, the slope (a measure of the degree of spread of the distribution) and the other is some intercept value (typically N (63.2%))the characteristic life of the Weibull distribution. On Weibull distribution graphs, using the two dening parameters, measured data will ideally plot as a straight line, frequently simplifying data analysis. Figure 12 shows an example of such an analysis for a particular kind of package, where early failures are suggested (points outside the range of the linear data). This example also shows that the data are not linear, implying that there are different failure modes for early failure and later failure. There is consensus that Pb-free solder joints have a longer thermal cycling lifetime than SnPb solder joints. However, the factors of grain orientation, the complex nature of Sn as the base material, and the combination of thermal and mechanical cyclic deformation make assessment and prediction challenging. For a product having Pb-free material as an interconnect, it is recommended to follow the Design for Reliability (DfR) guideline which applies to certain products to achieve a good, in other words, an expected lifetime at the end-use condition. Even though there are time-dependent factors such as creep during function, mechanical fatigue, and Sn grain orientation effects, accelerated tests are available to give some indication of the relative lifetimes of the solder joint or the product by using carefully selected acceleration factors. The standardized test methods are listed in several documents, such as the IPCJEDEC9701 A Performance test methods and qualication requirements for surface mount solder attachments, and for mechanical testing such as monotonic bend testing, IPC-JEDEC9702 is a good reference.

10. Implications of Pb-free Technology on Die and Board Design


The anisotropic properties of Sn along with different chemical behaviors will provide engineering and design challenges for the foreseeable future in electronic systems. As the industry copes with the anisotropic properties of Sn-based solders, nextgeneration board and die fabrication design decisions will be affected, so that they can accommodate the stronger properties of Sn-based solders. See also: Electrical and Electronic Connectors: Materials and Technology; Electronic Packages: Quality and Reliability; Electronic Packaging Materials: Properties and Selection; Electronic Packaging: Thermal, Mechanical, and Environmental Durability; Optoelectronic Packaging: Solder Assembly.

Bibliography
Anderson I E 2007 Development of SnAgCu and SnAg CuX alloys for Pb-free electronic solder applications. J. Mater. Sci.: Mater. Electron. 18, 5576 Brandes E A, Brook G B (eds.) 1997 Smithells Metal Reference Book, 7th edn. Butterworth, London Bieler T R, Jiang H, Lehman L P, Kirkpatrick T, Cotts E J, Nandagopal B 2008 Inuence of Sn grain size and orientation on the thermomechanical response and reliability of Pb-free solder joints. IEEE Trans. Compon. Packag. Technol. (CPMT) 31 (2), 37081 Borgesen P, Bieler T, Lehman L P, Cotts E J 2007 Pb-free solder: new materials considerations for microelectronics processing. MRS Bull 32, 3605 Chawla N, Sidhu R S 2007 Microstructure-based modeling of deformation in Sn-rich (Pb-free) solder alloys. J. Mater. Sci.: Mater. Electron. 18 (13), 17589 Deshpande V T, Sirdeshmukh D B 1962 Thermal expansion of tin in the betagamma transition region. Acta Cryst 15, 294 Dyson B F, Anthony T R, Turnbull D 1967 Interstitial diffusion of copper in tin. J. Appl. Phys. 38 (8), 3408 Henderson D W, Woods J J, Gosselin T A, Bartelo J, King D E, Kohonen T M, Korhonen M A, Lehman L P, Cotts E J, Noyan I C, Kang S K, Lauro P, Shih D-Y, Goldsmith C, Puttlitz K J 2004 The microstructure of Sn in near eutectic SnAgCu alloy solder joints and its role in thermomechanical fatigue. J. Mater. R. 19 (6), 160812 Hillman D, Wilcoxon R 2006 JCAA/JG-PP lead free solder testing for high reliability applications. In: SMTA International Conference Proceedings on CD. Surface Mount Technology Association, Edina, MN, pp. 836860 House D G, Vernon E V 1960 Determination of the elastic moduli of tin single crystals and their variation with temperature. Br. J. Appl. Phys. 11, 2549 Kinney C, Morris J W, Lee T K, Liu K C, Xue J, Towne D 2009 The inuence of an imposed current on the creep of Sn AgCu solder. J. Electron. Mater. 38 (2), 2216

11

Lead-free Solder
Lehman L P, Athavale S N, Fullem T Z, Giamis A C, Kinyanjui R K, Lowenstein M, Mather K, Patel R, Rae D, Wang J, Xing Y, Zavalij L, Borgesen P, Cotts E J 2004 Growth of Sn and intermetallic compounds in SnAgCu solder. J. Electron. Mater. 33 (12), 142939 Lehman L P, Xing Y, Bieler T R, Cotts E J Cyclic twin nucleation in tin based solder alloys. Acta Mater. (submitted). Lin Y W, Lai Y-S, Lin Y L, Tu C-T, Kao C R 2008 Tin whisker growth induced by high electron current density. J. Electron. Mater. 37 (1), 1722 Lu H Y, Balkan H, Ng K Y S 2005 Solidliquid reactions: the effect of Cu content on SnAgCu interconnects. JOM 57 (6), 305 Lu M, Shih D-Y, Lauro P, Goldsmith C, Henderson D W 2008 Effect of Sn grain orientation on electromigration degradation mechanism in high Sn-based Pb-free solders. Appl. Phys. Lett. 92, 211909 MatWeb: Material Property Data. http://www.matweb.com (accessed August 2009) Plumbridge W J 2007 Tin pest in electronics?. Circuit World 33 (1), 914 Rayne J A, Chandrasekhar B S 1960 Elastic constant of beta tin from 42 K to 300 K. Phys. Rev. 120, 1658 Rhee H, Subramanian K N, Lee A, Lee J G 2003 Mechanical characterization of Sn3.5Ag solder joints at various temperatures. Solder. Surf. Mt. Tech. 15 (3), 216 Riddington J R, Sahota M K 2003 Mechanical properties of lead alloys in compression. J. Mater. Civ. Eng. July/August, 3238 Rosenberg A, Winegard W C 1954 The rate of growth of dendrites in supercooled tin. Acta Metall 2, 342 Shen Y-L, Chawla N, Ege E S, Deng X 2005 Deformation analysis of lap-shear testing of solder joints. Acta Mater 53, 263342 Sigelko J, Choi S, Subramanian K N, Lucas J P, Bieler T R 1999 Effect of cooling rate on microstructure and mechanical properties of eutectic SnAg solder joints with and without intentionally incorporated Cu6Sn5 reinforcements. J. Electron. Mater. 28 (11), 11848 Telang A U, Bieler T R, Crimp M A 2006 Grain boundary sliding on near-7, 14, and 221 special boundaries during thermomechanical cycling in surface-mount leadfree solder joint specimens. Mater. Sci. Eng., A 421 (1/2), 2234 Tu K N, Li J C M 2005 Spontaneous whisker growth on leadfree solder nishes. Mater. Sci. Eng., A 409 (1/2), 1319 WebElements: the periodic table on the web. http://www. webelements.com (accessed August 2009) Wu A T, Gusak A M, Tu K N, Kao C R 2005 Electromigration-induced grain rotation in anisotropic conducting beta tin. Appl. Phys. Lett. 86, 241902

T. R. Bieler and Tae-kyu Lee

Copyright r 2010 Elsevier Ltd. All rights reserved. No part of this publication may be reproduced, stored in any retrieval system or transmitted in any form or by any means: electronic, electrostatic, magnetic tape, mechanical, photocopying, recording or otherwise, without permission in writing from the publishers. Encyclopedia of Materials: Science and Technology ISBN: 978-0-0804-3152-9 pp. 112 12

Vous aimerez peut-être aussi