doi:10.1016/j.actamat.2005.07 - nasa · 2008. 4. 8. · title: doi:10.1016/j.actamat.2005.07.016...

18
Whisker and Hillock formation on Sn, Sn–Cu and Sn–Pb electrodeposits W.J. Boettinger * , C.E. Johnson, L.A. Bendersky, K.-W. Moon, M.E. Williams, G.R. Stafford Materials Science and Engineering Laboratory, Metallurgy Division, NIST, Mailstop 8555, 100 Bureau Dr., Gaithersburg, MD 20899, United States Received 3 March 2005; received in revised form 20 July 2005; accepted 20 July 2005 Available online 8 September 2005 Abstract High purity bright Sn, Sn–Cu and Sn–Pb layers, 3, 7 and 16 lm thick were electrodeposited on phosphor bronze cantilever beams in a rotating disk apparatus. Beam deflection measurements within 15 min of plating proved that all electrodeposits had in-plane compressive stress. In several days, the surfaces of the Sn–Cu deposits, which have the highest compressive stress, develop 50 lm contorted hillocks and 200 lm whiskers, pure Sn deposits develop 20 lm compact conical hillocks, and Sn–Pb deposits, which have the lowest compressive stress, remain unchanged. The differences between the initial compressive stresses for each alloy and pure Sn is due to the rapid precipitation of Cu 6 Sn 5 or Pb particles, respectively, within supersaturated Sn grains produced by electrodepos- ition. Over longer time, analysis of beam deflection measurements indicates that the compressive stress is augmented by the forma- tion of Cu 6 Sn 5 on the bronze/Sn interface, while creep of the electrodeposit tends to decrease the compressive stress. Uniform creep occurs for Sn–Pb because it has an equi-axed grain structure. Localized creep in the form of hillocks and whiskers occurs for Sn and Sn–Cu because both have columnar structures. Compact hillocks form for the Sn deposits because the columnar grain boundaries are mobile. Contorted hillocks and whiskers form for the Sn–Cu deposits because the columnar grain boundary motion is impeded. Published by Elsevier Ltd on behalf of Acta Materialia Inc. Keywords: Whiskers; Electroplating; Creep; Grain boundary diffusion; Solder 1. Introduction Sn whiskers have been an industrial concern and an interesting problem for many years. They are known to cause short circuits in fine pitch pretinned electrical components. In contrast to many whisker growth pro- cesses, Sn whiskers grow by the addition of material at their base not at their tip; i.e., they grow out of the sub- strate [1]. They can grow from as-formed electrodepos- its, vapor deposited material [2] and intentionally deformed coatings of Sn [3]. Similar whiskers are ob- served in Cd, In and Zn. Whiskers appear to be a local response to the existence of residual stress and compres- sive residual stress is usually considered a precondition for whisker growth [3]. Annealing or melting (reflow in solder terminology) usually mitigates the growth, although subsequent bending of leads can re-establish compressive stress. In 1966, Pb additions of a few per- cent to Sn electroplate were found to greatly reduce the tendency to form whiskers [4] and interest in the sub- ject waned. Recently, interest in Pb-free surface finishes for green manufacturing of electronic components has reopened this dormant field and an annotated bibliogra- phy has been prepared [5]. Questions remain as to the dominant source of stress and the precise growth mech- anism. The final goal is the development of a practical mitigation strategy for electronic components during electroplating, storage and/or service. 1359-6454/$30.00 Published by Elsevier Ltd on behalf of Acta Materialia Inc. doi:10.1016/j.actamat.2005.07.016 * Corresponding author. Tel.: +1 301 975 6160; fax: +1 301 975 4553. E-mail addresses: [email protected], wboettinger@nist. gov (W.J. Boettinger). Acta Materialia 53 (2005) 5033–5050 www.actamat-journals.com

Upload: others

Post on 03-Jan-2021

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Acta Materialia 53 (2005) 5033–5050

www.actamat-journals.com

Whisker and Hillock formation on Sn, Sn–Cu and Sn–Pbelectrodeposits

W.J. Boettinger *, C.E. Johnson, L.A. Bendersky, K.-W. Moon,M.E. Williams, G.R. Stafford

Materials Science and Engineering Laboratory, Metallurgy Division, NIST, Mailstop 8555, 100 Bureau Dr., Gaithersburg, MD 20899, United States

Received 3 March 2005; received in revised form 20 July 2005; accepted 20 July 2005Available online 8 September 2005

Abstract

High purity bright Sn, Sn–Cu and Sn–Pb layers, 3, 7 and 16 lm thick were electrodeposited on phosphor bronze cantilever beamsin a rotating disk apparatus. Beam deflection measurements within 15 min of plating proved that all electrodeposits had in-planecompressive stress. In several days, the surfaces of the Sn–Cu deposits, which have the highest compressive stress, develop 50 lmcontorted hillocks and 200 lm whiskers, pure Sn deposits develop 20 lm compact conical hillocks, and Sn–Pb deposits, which havethe lowest compressive stress, remain unchanged. The differences between the initial compressive stresses for each alloy and pure Snis due to the rapid precipitation of Cu6Sn5 or Pb particles, respectively, within supersaturated Sn grains produced by electrodepos-ition. Over longer time, analysis of beam deflection measurements indicates that the compressive stress is augmented by the forma-tion of Cu6Sn5 on the bronze/Sn interface, while creep of the electrodeposit tends to decrease the compressive stress. Uniform creepoccurs for Sn–Pb because it has an equi-axed grain structure. Localized creep in the form of hillocks and whiskers occurs for Sn andSn–Cu because both have columnar structures. Compact hillocks form for the Sn deposits because the columnar grain boundariesare mobile. Contorted hillocks and whiskers form for the Sn–Cu deposits because the columnar grain boundary motion is impeded.Published by Elsevier Ltd on behalf of Acta Materialia Inc.

Keywords: Whiskers; Electroplating; Creep; Grain boundary diffusion; Solder

1. Introduction

Sn whiskers have been an industrial concern and aninteresting problem for many years. They are knownto cause short circuits in fine pitch pretinned electricalcomponents. In contrast to many whisker growth pro-cesses, Sn whiskers grow by the addition of material attheir base not at their tip; i.e., they grow out of the sub-strate [1]. They can grow from as-formed electrodepos-its, vapor deposited material [2] and intentionallydeformed coatings of Sn [3]. Similar whiskers are ob-served in Cd, In and Zn. Whiskers appear to be a local

1359-6454/$30.00 Published by Elsevier Ltd on behalf of Acta Materialia In

doi:10.1016/j.actamat.2005.07.016

* Corresponding author. Tel.: +1 301 975 6160; fax: +1 301 9754553.

E-mail addresses: [email protected], [email protected] (W.J. Boettinger).

response to the existence of residual stress and compres-sive residual stress is usually considered a preconditionfor whisker growth [3]. Annealing or melting (reflow insolder terminology) usually mitigates the growth,although subsequent bending of leads can re-establishcompressive stress. In 1966, Pb additions of a few per-cent to Sn electroplate were found to greatly reducethe tendency to form whiskers [4] and interest in the sub-ject waned. Recently, interest in Pb-free surface finishesfor green manufacturing of electronic components hasreopened this dormant field and an annotated bibliogra-phy has been prepared [5]. Questions remain as to thedominant source of stress and the precise growth mech-anism. The final goal is the development of a practicalmitigation strategy for electronic components duringelectroplating, storage and/or service.

c.

Page 2: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

5034 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

As with Sn whiskers, compressive residual stress alsoplays a role in hillock growth on the surface of thin me-tal films made by vapor deposition. Hillock growth is of-ten treated as a localized diffusional creep/grainboundary sliding phenomenon that relieves the com-pressive stress (see, for example [6–8]). The compressivestress is typically generated by the differential thermalcontraction of the deposit and substrate during coolingfrom deposition temperature. On the other hand Sn elec-troplating is performed at room temperature; thus,residual stresses cannot have a thermal origin. One ofthe most commonly discussed sources of compressivestress in Sn electrodeposits is intermetallic compound(IMC) formation due to the reaction of Sn with Cu inthe substrate metal [9–11]. Despite the differences be-tween hillock growth on vapor deposited metals andSn whisker growth from electrodeposits, localized creepis a potential mechanism for both.

In a previous paper [12], it was shown that filamen-tary whisker defects were not observed on bright pureSn electrodeposits if high purity (18 MX cm) water wasused to prepare the commercial methanesulfonate elec-trolyte. Intentional Cu additions as an impurity to theelectrolyte in the range from 0.5 · 10�3 to 25 ·10�3 mol/L did however cause whiskers and hillocks toform. But IMC forms on the interface between the de-posit and the substrate for both pure Sn and Sn–Cudeposits. Thus, we questioned whether IMC formationon the interface between the substrate and the depositwas the major stress source for whisker formation, espe-cially because electroplating itself often produces signif-icant residual stress in deposits [13].

Using deflection measurements of plated cantileverbeams, the present paper seeks to determine whetherthe in-plane residual stress in Sn–Cu electrodeposits iscompressive and greater than that for pure Sn electrode-posits. Sn–Pb alloys were included in the study becauseof the known whisker mitigation effect of Pb. The paperalso determines the relative amounts of stress generatedby the electrodeposition process, the alloy additions andhow the reaction at the deposit/substrate boundarychanges the deposit stress with time. The paper reportsthe propensity for whisker/hillock formation and theinterior microstructure of the electrodeposits and com-pares this tendency with the stress and plastic strain rateof the deposits. Finally, a mechanism based on localizedcreep of columnar grain structures is suggested for thelocalized surface disturbance in response to the com-pressive stress.

2. Reaction of Sn with Cu

The formation of the layer of Cu6Sn5 by reaction be-tween Cu and a layer of Sn is a relatively slow processcontrolled by diffusion. Onishi and Fujibuchi [14] mea-

sured the intermetallic growth rates in diffusion couplesbetween Cu and Sn between 109 and 220 �C. The thick-ness of the total intermetallic layer (Cu6Sn5 + Cu3Sn) isgiven by

dIMC ¼ffiffiffiffiffiBt

pð1Þ

with B = 6.23 · 10�6exp(Q/RT) cm2/s and Q = 57.7 kJ/mol. At 298 K, the extrapolated value of B is4.7 · 10�16 cm2/s. A square root of time dependence isvalid as long as the intermetallic layer is much thinnerthan the Sn and Cu layers. Alternately, Tu and Thomp-son [15] have measured the growth rate of intermetallicat room temperature. They observed only Cu6Sn5 andobtained

dIMC ¼ B0t ð2Þwith B 0 = 4 · 10�12 cm/s for IMC thicknesses up to300 nm. This linear form implies that interface attach-ment kinetics is dominant for room temperature IMCgrowth, not diffusion, at least in its initial stage. Finally,in unpublished work [16] using Sn plating on a Cu metal-lized quartz crystal resonator and electrochemical strip-ping after different hold times, the intermetallic wasfound to grow as

ffiffiffiffiffiBt

pat room temperature with

B = 1.76 · 10�15 cm2/s. Below, we presentmeasurementsof residual stress in deposits 15 min after plating. Theabove considerations indicate that less than 20 nm ofintermetallic can form during this time and would havenegligible effect on the stress in 3–16 lm thick deposits.On the other hand, the above suggests that between 1.2and 2.3 lm may form in a year (3 · 107 s) and the stressof the deposit may be influenced at longer times.

For pure Sn in contact with a Cu substrate, we do notexpect, nor has it ever been observed to our knowledge,that discrete (isolated) intermetallic particles form on Sngrain boundaries due to rapid diffusion of Cu from thesubstrate. The supersaturation required to nucleate a dis-crete IMC particle could only occur during the short per-iod of time before a continuous layer of the same IMCcoats the Cu/Sn interface. After that time, formationand/or further growth of discrete IMC particles on grainboundaries is not possible. Similarly, we have not ob-served rapid penetration of IMC along Sn grain bound-aries. However, as shown below, when Cu2+ is presentin the electrolyte, discrete IMC particles can form withinthe Sn grains and along grain boundaries by solid stateprecipitation from a Cu supersaturated Sn solid solutionformed by electrochemical co-deposition. The distinctionis important because, as argued below, the stress-freestrains caused by the two processes differ in sign.

3. Experimental method

Using a metal shear, coupons, 2.5 cm square, werecut from 152 lm thick, half hard, rolled phosphor

Page 3: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5035

bronze with nominal mass fractions of Cu–5%Sn–(0.03–0.35)%P. The edges were deburred with a jeweler�s fileand the surface to be plated was polished with 0.3 lmAl2O3. Using the geometry shown in Fig. 1, cantileverbeams, 2 mm · 20 mm, and a supporting frame werechemically etched from the coupons using acid resistanttape and 60% nitric acid. It was found necessary to em-ploy beams perpendicular to the sheet rolling directionto obtain reproducible beam deflections after plating.This may lead to a non-isotropic biaxial response ofthe beams, but this factor will be ignored.

Prior to electroplating, all substrates were immersed in25% sulfuric acid solution for 5 s to remove oxide andwere rinsed in deionized water. Bright electrodeposits ofSn, Sn–Cu and Sn–Pb alloys were electrodeposited froma commercial methanesulfonate electrolyte containing340 mmol/L Sn2+. The electrolyte was preparedwith highpurity water with a resistivity of 18.3 MX cm. The canti-lever samples were attached with plater�s tape to a diskelectrode assembly and rotated during plating at 100 rota-tions per min to provide reproducible hydrodynamicconditions. This rotation speed creates a uniform hydro-dynamic boundary layer approximately 40 lm thickalong the electrode surface [17]. All coupon surfaces ex-cept the cantilever were masked. Plating was performedat a constant current density of 60 mA/cm2 in 1 L of solu-tion at 25 ± 0.5 �C to various average thicknesses be-tween 3 and 16 lm (determined by weight gain afterplating). The anode was a 99.999% pure Sn sheet. Theplating efficiency was between 98% and 100%. Thus,hydrogen evolution is not thought to affect the results ofthis study. The coupons were removed from the bath

Fig. 1. Drawing of a cantilever beam coupon with plated region shownin gray. The three frame positions used for tip position measurementare indicated. Dimensions are in mm.

while still electrified. At this current density, the platingrate is approximately 0.05 lm/s and 320 s is required toform a 16 lm deposit. For the Sn–Cu deposits the Cu2+

concentration in the electrolyte was 15.0 mmol/L by theaddition of copper methanesulfonate [Cu(CH3SO3)2].An examination of the alloy-electrolyte compositions re-ported in [12] indicates that the co-deposition of copper isdiffusion-limited at this Cu2+ concentration and currentdensity. For the Sn–Pb deposits the Pb2+ concentrationin the electrolyte was 9.2 mmol/L by the addition of leadmethanesulfonate [Pb(CH3SO3)2].

After electrodeposition, the plating tape was dissolvedwith acetone and the samples were rinsed with deionizedwater. For each coupon, the position of the beam tip withrespect to the frame was measured before plating, within15 min after plating and then at various times up to2 · 106 s (20 days). The height of the beam tipwith respectto the frame was measured at three positions as shown inFig. 1 by the changes in focus position in an opticalmicro-scope. These measurements were done with the couponhorizontal in the two orientations with respect to gravity.Correction was made for the increased thickness of thebeamdue to plating.The position of the beamat each timewas taken as the mean of the six measurements. Thechange in position from the initial position (prior to plat-ing) is defined as the deflection. A negative deflectionmeans that the electrodeposit surface is convex indicatingthat the deposit wants to expand (positive stress-freestrain), the constraint of the underlying substrate puttingthe deposit in biaxial in-plane compression.

The plated surfaces were examined without prepara-tion for the presence of whiskers and/or hillocks usingoptical and scanning electron microscopy (SEM). Se-lected samples were prepared for optical and SEMmetallographic cross-sectional examination by mount-ing in epoxy and using standard polishing procedures.Sn, Sn–Cu and Sn–Pb samples plated on amorphouscarbon were prepared for chemical analysis and fortransmission electron microcopy (TEM) by cold stageprecision ion milling.

Tensile tests with strain gauges were conducted on the150 lm thick phosphor bronze substrate material.Young�s modulus and Poisson�s ratio were determinedto be 134 ± 1.6 GPa and 0.386 ± 0.02, respectively.Microhardness measurements were performed oncross-sections of the 16 lm thick electrodeposits. AKnoop indenter with 1 g load was employed and ori-ented with the long axis parallel to the plating surface.Using the approximation that the yield stress is 1/3 ofthe microhardness, yield strength values of 44 ± 2,64 ± 4 and 44 ± 4 MPa were obtained for the pure Sn,Sn–Cu and Sn–Pb deposits, respectively. The higheryield stress of the Cu–Sn deposits is likely due to thepresence of the hard intermetallic particles. All mea-sured stresses in the deposits are below these yield stressvalues.

Page 4: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 2. Low, intermediate and high magnification SEM micrographsof the electrodeposit surface of a 16 lm thick pure Sn electrodeposit ona cantilever beam showing bimodal size distribution of conicalhillocks.

5036 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

4. Results

4.1. Surface microstructure of deposits

On pure Sn deposits, compact conical hillocksapproximately 2 lm high · 2 lm wide appear within900 s of plating. These appear to be grains that have ri-sen from the surface. Within two days, approximately10% of the small hillocks grow while the others remainthe same size. This process produces the bimodal surfacestructure shown in Fig. 2. Grain boundary groves arealso apparent on the deposit surface. Multiple electronbackscatter diffraction patterns taken from the smalland large conical hillocks show most to be single grains.Patterns taken from 26 conical hillocks surface showedno obvious preferred orientation. The single grain struc-ture of the conical hillocks was confirmed by an SEMimage of a focused ion beam (FIB) cross-section(Fig. 3) on a 16 lm thick Sn sample [18]. We note a sig-nificant change in the as-plated columnar grain structureof the deposit under the hillock. Significant grainboundary motion has occurred in conjunction with thedevelopment of the hillock. It is important to note thatthe continuous white contrast on the Sn columnar grainboundaries is not IMC formed by diffusion up the grainboundaries. It is an artifact of the FIB procedure (seeFig. 9) as is the darker contrast in the hillock grainabove the deposit surface.

Filamentary whiskers and hillocks were evident onthe Sn–Cu deposits within 2 days of plating. Fig. 4shows SEM views of a 16 lm thick Sn–Cu alloy depositsurface. The hillocks are not conical and have a con-torted appearance. They are also much larger (50 lmwide · 50 lm wide) than the conical hillocks in the pureSn deposits. Whiskers appear to spew forth fromapproximately 10% of these features. However, thewhiskers actually appear first followed by continuedaccumulation of Sn at the base. The grain boundarieson the undisturbed part of the surface are difficult toidentify because of a fine dispersion of particles andhave been sketched in. These particles are Cu6Sn5 thathave coarsened over several months from the initiallyfine precipitates seen by TEM that formed within theSn grains shortly after plating (as described below). Toobtain a sense of the time scale during which the surfacedisturbances grow, Fig. 5 shows a sequence of SEMphotographs of a feature that grew in volume with time.Contorted hillocks and whisker filaments occur on thesame sample (Fig. 4). Both are shown in Fig. 6, whichare SEM images of FIB cross-sections [18] at two posi-tions on a single 16 lm thick Sn–Cu sample. Note thatthe as-plated grains are columnar. The contorted hillockshape forms because of the activity of more than onegrain under the disturbance and the IMC hindered grainboundary motion under the hillock. Grain boundarypinning appears to have been very effective under the fil-

amentary whisker where no grain boundary motion hasoccurred in the deposit under the filament.

Fig. 7 shows SEM views of a 16 lm thick Sn–Pb elec-trodeposit surface 85 days after plating. The surface isfree of hillocks and whiskers. Only grain boundarygrooves are evident. The backscattered SEM imageshows that coarsened Pb particles exist as separategrains mixed with Sn grains. A FIB cross-section(Fig. 8) [18] shows that the grain structure of the Sn–Pbelectrodeposit is not columnar. Compared to the Sn

Page 5: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 3. SEM image of FIBed sample showing small and large hillockson a 16 lm thick pure Sn electrodeposit. IMC and the bronze are seenat the bottom [18]. The white contrast on the Sn columnar grainboundaries is an artifact of the FIB procedure as is the darker contrastin the hillock grain above the deposit surface.

Fig. 4. Low, intermediate and high magnification SEM micrographsof the electrodeposit surface of a 16 lm thick Sn–Cu electrodeposit ona cantilever beam (photo taken after 176 days (1.5 · 107 s)). The finelines center and lower right in the top figure are long whiskers. A fewSn grain boundary locations are indicated in the bottom figure.

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5037

and Sn–Cu deposits, the Sn–Pb deposits have manymore grain boundaries parallel to the top surface. It isknown that the co-deposition of Pb significantly reducesany crystallographic texture in the Sn by preventingcolumnar growth [19].

Examination of the surfaces of the various 3 and7 lm thick deposits showed the same structures notedabove. However, after cantilever measurements werestopped, re-examination of the surfaces after 2.5 yearsshowed that while the 7 and 16 lm thick deposits sur-faces remained the same, the 3 lm thick Sn deposithad developed a very low density of conical hillocksand long filamentary whiskers and the 3 lm thick depos-it of Sn–Pb had developed a low density of short 20 lmlong whiskers. It is possible that excessive handling ofthe cantilevers led to these changes for the thin deposits.

4.2. Internal deposit microstructure, composition and

microhardness

SEM micrographs of metallographic cross-sections ofthe 16 lm (average) deposits of the three alloys areshown in Fig. 9. The columnar grain structure of theSn and Sn–Cu deposits is confirmed. These samples wereapproximately one-year-old (3 · 107 s) when sectioned.

Page 6: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 5. Time sequence of the formation of an eruption on a Sn–Cu 16 lm thick cantilever sample. Prior to 2 days (1.5 · 105 s) no surface upheaval isobserved.

Fig. 6. SEM images of FIBed 16 lm thick Sn–Cu electrodeposit:(a) location showing a contorted eruption; (b) location showing afilamentary whisker on the same sample [18].

5038 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

In all three micrographs, a 1.5–2.5 lm thick scallopedintermetallic Cu6Sn5 layer is seen on the interface be-tween the deposit and the substrate. A much thinnerCu3Sn layer is also present but not visible. The thicknessof the Cu6Sn5 layer is in general agreement with that de-scribed in Section 1. In the Sn–Cu deposits intermetallicCu6Sn5 is also seen distributed throughout the deposit,primarily on Sn grain boundaries. No IMC is seen onthe grain boundaries of the pure Sn or Sn–Pb samples.We again note the presence of many transverse grainboundaries in the Sn–Pb deposit compared to the pureSn and Sn–Cu deposits. The Pb phase location is not re-vealed in these micrographs.

Determining the deposit over-all composition proveddifficult due to the low solute levels, small sample vol-umes and the two-phase nature of the deposits. Severalmethods were used: inductively coupled plasma spec-trometry analysis (ICP), energy dispersive spectroscopy(EDS) in the SEM from area scans (10 lm in the platingdirection by 20 lm) of cross-sectioned 16 lm thickdeposits, and the area fraction of Cu6Sn5 (excluding thaton the bronze/Sn interface) for the Sn–Cu deposit. TheCu concentration ranged between 1.4% and 3.7% massfraction Cu depending on the method used. The areafraction of Pb in the Sn–Pb sample could not be mea-sured due to polishing difficulties. The other measure-ment methods gave Pb concentrations between 1.0%and 3.5% mass fraction Pb. Because the deposition con-ditions were tightly controlled, we do not believe theseranges represents true variation in sample composition.The expected Cu concentration, based on diffusion-limited Cu2+ and a diffusion coefficient of5 · 10�6 cm2/s [20], is 3.3% mass fraction Cu. We willset the compositions to be 3% mass fraction Cu and2% mass fraction Pb for the analysis below.

Page 7: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 7. Low, medium and high magnification SEM micrographs of theelectrodeposit surface of a 16 lm thick Sn–Pb electrodeposit on acantilever beam after 85 days (7.3 · 106 s). The highest magnificationSEM micrograph employed back scattered electron contrast andclearly shows four isolated Pb grains.

Fig. 8. SEM images of FIBed 16 lm thick Sn–Pb electrodeposit. Thegrain structure is not columnar [18].

Fig. 9. Cross-sections of 16 lm thick (nominal) electrodeposits onphosphor bronze: (a) pure Sn; (b) Sn–Cu; (c) Sn–Pb approximately oneyear after plating.

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5039

TEM examination (Fig. 10) of a Cu-containing de-posit on glassy carbon one day after deposition showsfine intermetallic precipitates within the Sn grains anda few on the grain boundaries. It is well known that elec-trodeposition is capable of producing alloy deposits thatare super-saturated; i.e., they contain more alloyingaddition than the phase diagram permits at the temper-ature of deposition [21]. It is likely that the fine IMCparticles in Fig. 10 form by precipitation from solid

solution quite rapidly after plating (within 103 s) andevidently coarsen over several months when they appearmostly along the Sn grain boundary as seen in Fig. 9.Similarly, Sheng et al. [22] found many grain boundary

Page 8: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 10. TEM (a) bright and (b) dark field views of Sn–Cu depositapproximately 24 h after electroplating. Cu6Sn5 intermetallic particleswere identified by SAD in the interior of the Sn grains and on the grainboundaries. The dark field images show that the internal particlesshare a common orientation.

Fig. 11. TEM of Sn–Pb electrodeposit one month after plating: (a) Pbparticles are approximately 50 nm tetragonally distorted cuboidswithin the Sn grains and (b) larger Pb particles on the Sn grainboundaries.

5040 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

precipitates of Cu6Sn5 in aged Sn–Cu electrodeposits ofeutectic composition on Cu leadframes but very few insamples with a pure Sn finish. We note that supersatura-tion of Sn with up to 2.2% mass fraction Cu has beenreported using rapid solidification and immediate dif-fraction measurements at liquid nitrogen temperature[23]. Thus, similar supersaturation by electrodepositionis likely.

TEM examination of a Pb-containing sample one dayafter deposition on glassy carbon showed extremely fine(10 nm) spherical precipitates of Pb. One month afterplating (Fig. 11), the Pb particles are approximately50 nm tetragonally distorted cuboids within the grainsand larger particles on the Sn grain boundaries. Table 1

summarizes the observed surface defects and the internalmicrostructures of the deposits.

4.3. Hillock volume

Measuring the total amount of material pushed upabove the deposit surface during whisker growth is dif-ficult due to the irregularity of the shapes. The regularityof the conical hillocks observed on the 16 lm thick pureSn electrodeposits provides an opportunity for such ameasurement. The distribution of projected areas (ontothe plated surface) of the hillock population was measured

Page 9: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Table 1Summary of surface defects and microstructure of the deposits

Sn Sn–Cu Sn–Pb

Surface defects Conical hillocks Contorted hillocks and filamentary whiskers NoneSn grain structure Columnar, mobile Columnar, immobile Equi-axedInitial alloy structure – Fine Cu6Sn5 Fine PbAged alloy structure – Coarse Cu6Sn5 on grain boundaries Sparse equi-axed Pb grains

Cu6Sn5 forms slowly on the deposit/substrate boundary for all.

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5041

by quantitative metallography 7.86 · 106 s after plating.Each hillock was assumed to be a cone with a radius de-rived from its measured projected area and with an apexsemi-angle (estimated from FIB cross-sections) of 45�for the small hillocks and 58� for the large hillocks ofthe bimodal distribution. The total hillock volume perunit plated surface area is 0.012 lm. We estimate thatabout a third of this volume develops within 1 h of plat-ing. Dividing by the deposit thickness, 16 lm, an aver-age volumetric plastic strain, DV/V, of 8 · 10�4 wasdetermined. An estimate of the in-plane biaxial stressthat was relieved by this hillock formation [24] is

Dr ¼ E0f

2

DVV

; ð3Þ

where E0f is the biaxial modulus of Sn (78 GPa, see

Table 2). The compressive stress reduction is 30 MPa.This number is of the same order as, but about twice,the stress relaxation measured over the 7.86 · 106 s bythe cantilever beam measurements.

4.4. Cantilever beam deflection measurements

4.4.1. Initial deflectionAfter release from the rotating assembly, the deposit

side of all cantilever beams was immediately convex; i.e.,the deflection is negative. This indicates that the depositwants to expand with respect to the beam and is thus ina state of compressive stress due to the constraint of thesubstrate. First, we examine the average stress in the de-posit determined from the earliest deflection measure-ment at approximately 103 s shown in Fig. 12. Asdescribed above, the intermetallic on the bronze/Sn

Table 2Parameters for Stoney�s equation and other mechanical parameters

E0S = substrate biaxial modulus = ES/(1 � mS) = 216 GPa

(ES = 134 GPa, mS = 0.386)dS = substrate thickness = 152 lmdf = film thickness = 3, 7, 16 lmL = beam length = 20 mmd = deflection of beam from position prior to platingK = curvature of beam = 2d/L2

E0f = Sn biaxial modulus = 78 GPa (ES = 50 GPa, mS = 0.357 [26])

E0IMC = Cu6Sn5 biaxial modulus = 122 GPa (ES = 86 GPa, mS = 0.3[27])

interface is quite thin (20 nm) at this time and can be ne-glected. To treat the 16 lm thick electrodeposit, the first-order correction to Stoney�s equation is required. We as-sume that the bi-axial in-plane stress in the deposit be-fore it is released from the rotating disk is uniform.The stress before release to first order depends on themodulus of the deposit. However, the stress in the de-posit after it is released, rf, does not depend on the mod-ulus of the deposit and is given by

rf ¼E0S

6

d2S

d f

K 1� d f

dS

� �¼ E0

S

3

d2S

d f

d

L21� d f

dS

� �; ð4Þ

where the notation is defined in Table 2. Brenner andSenderoff [25] summarize various approximations forthick deposits.

Fig. 13 shows a plot of the initial compressive stress(approximately 15 min after plating) vs. deposit thick-ness for the three deposit compositions using the param-eter values in Table 2. All stresses are compressive. Foreach alloy, the initial stress becomes more compressivewith increasing thickness. Because the diffusion fieldsin the electrolyte, and thereby the deposit alloy compo-sitions, reach steady-state within the first 0.1 lm ofdeposition due to the rotation of the electrode, the ob-served stress-thickness relationship cannot be attributedto a concentration gradient in the deposition direction.The reason for the increase in the magnitude of the com-pressive stress with deposit thickness is unclear.

Regardless of deposit thickness, the Sn–Cu depositshave the highest compressive stress, the pure Sn hasintermediate compressive stress, and the Sn–Pb deposithas the lowest compressive stress. A positive correlationexists between the magnitude of the compressive stressand the degree of disruption of the surface morphology;viz., the Sn–Cu samples have whiskers and large con-torted hillocks, the pure Sn has only small hillocksand the Sn–Pb alloys are free of hillocks and whiskers.

4.4.2. Deflection vs. time

As shown in Fig. 12, the beams return toward theirinitial positions with many of the samples actually pass-ing through the initial position. Use of Eq. (4) would im-ply that the deposit stress begins compressive andbecomes tensile with time. However this is not necessar-ily correct. To determine the average stress and thestrain rate in the Sn and Sn alloy deposits from the

Page 10: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 12. (a) Deflection vs. time for 16 lm thick deposits. Lines are power law fits. (b) Deflection vs. time curves for 7 lm thick deposits. Lines arepower law fits. (c) Deflection vs. time curves for 3 lm thick deposits. Lines are power law fits.

5042 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

deflection data, we need to perform a ‘‘two layer Stoneyanalysis’’, one layer for the deposit and one for thegrowing IMC. The details of a model are given in theAppendix and the analysis of the time dependence willbe deferred to the discussion.

1 The measured density agrees with the measured lattice parametersof the B82 (P63/mmc) structure of Cu6Sn5 if one considers the vacant2d Wycoff sites required to achieve the correct phase composition.

5. Discussion

5.1. Molar volume of Sn–Cu and Sn–Pb alloys

Before we can analyze the reasons for the variousstates of stress in the deposit, it is necessary to considerthe molar volume (volume per mole of atoms) for Sn–Cu and Sn–Pb alloys as summarized by Fig. 14. For

Sn–Cu, crystallographic data was used for the pure ele-ments and the Cu-rich phases. For Cu6Sn5, Fields et al.[27] measured a density of 8.28 g/cm3 that converts to10.6 cm3/mol1 (of atoms) using the composition of0.44 atomic fraction Sn given in [28]. In the same man-ner, the measured density of Cu3Sn, 8.90 g/cm3, con-verts to 8.6 cm3/mol. The equilibrium solubility of Cuin solid Sn is very small: 6 · 10�3 mass fraction Cu at227 �C [29] and 10�7 mass fraction Cu at room temper-ature (thermodynamic calculation [30]). As noted above,however, there is evidence from Fig. 10 that a supersat-

Page 11: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 14. Volume per mole of atoms for phases for the Sn–Cu and Sn–Pbsystems. Dashed lines show supersaturated solid solutions. Thevertical arrows show the volume change for the various metallurgicalprocesses.

Fig. 13. Initial stress (approximately 15 min after plating) as afunction of plating thickness.

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5043

urated solid solution of Cu in Sn is formed by the elec-trodeposition process. Because no lattice parameter datawas reported in [23], we require an estimate of the molarvolume of Sn supersaturated with Cu, Sn(Cu). Given thework on fast diffusion2 of Cu in Sn, the presumptionthat the Cu is dissolved interstitially in Sn [31,32], andassuming that the partial molar volume of interstitialCu is zero, the volume per mole of atoms of supersatu-rated Sn(Cu) is shown in Fig. 14 and is given by

V SnðCuÞM ¼ 16:28ð1� XCuÞ; ð5Þ

where XCu it the mole fraction of Cu. These consider-ations would imply that the lattice parameter of Sn isunchanged by the interstitial Cu. The lattice parametersfor the Sn–Pb system are summarized in Ref. [33]. Spe-cifically, Sn rich Sn(Pb) solid solution up to 0.012 massfraction Pb were measured at 183 �C in [34]. After con-version from kX units to SI units and changing the valueof the intercept to get the accepted value for pure Sn atroom temperature, the results are

V SnðPbÞM ¼ 16:28þ 3:52X Pb; ð6Þ

where XPb it the mole fraction of Pb. This expression isplotted in Fig. 14.

We extract a few numbers from Fig. 14 for use below.First, the formation of a layer of Cu6Sn5 at an interfacebetween Sn and Cu is considered. The volumetric strainfor this process can be read directly from the graph ordetermined analytically. It is the difference between the

2 We note that Ni also exhibits fast diffusion in Sn [35] due to itsinterstitial nature and would be expected to exhibit the same molarvolume behavior.

molar volume of the Cu6Sn5 and the molar volume ofan unalloyed mixture of pure Cu and pure Sn of thesame average composition as the intermetallic. The frac-tional change is �0.056. In other words, the IMC takesup 5.6% less volume than a rule of mixture combinationof the Sn and Cu reactants from which the IMC forms.

From Fig. 14, we can also extract the volume changewhen precipitation of Cu6Sn5 occurs from a supersatu-rated Sn(Cu) solid solution of mass fraction 3% Cu. Itis the difference between the molar volume of the super-saturated solution and the volume of a equilibrium mix-ture of Sn and Cu6Sn5 of the same average composition

Page 12: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

5044 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

as the supersaturated solution. For a deposit of 3% massfraction Cu (5.5% atom fraction), the fractional changein volume is +0.022. That is, the volume of a supersat-urated solution of Sn(Cu) expands by 2.2% when ittransforms to an equilibrium mixture of Sn and Cu6Sn5phases. Finally, the volume change of material thatundergoes precipitation of Pb from a supersaturatedSn(Pb) solution can be obtained from the graph. For adeposit of 2% mass fraction Pb (1.16% atom fraction),the fractional change in volume is �0.0013. That is,the volume of a supersaturated solution of Sn(Pb)shrinks by 0.13% when it transforms to an equilibriummixture of Sn and Pb phases.

5.2. Initial stress

Understanding the causes of growth stress in thinfilms formed by electrodeposition (and by other meth-ods such as condensation from the vapor [36]) are ac-tive areas of research. We consider here only thedifference between the stress in the pure Sn, and thosein the alloy deposits. For a given thickness, we assumethat the alloy deposits are formed with the samegrowth stress as pure Sn and that the precipitationprocess in the alloys shortly after deposition attemptsto alter the deposit volume and hence change thestress prior to the first measurement. Table 3 givesthe experimentally determined deviation of the stress-free strain values above or below that determinedfor the pure Sn. Also given is 1/3 of the change inthe deposit volumetric strain expected from the changein molar volumes for precipitation from supersatu-rated solid solutions described in Section 5.1.

If the model worked perfectly, the numbers in eachrow of the last two columns should be the same. Similarresults are obtained for the 3 and 7 lm thick electrode-posits, except the experimental increments get smaller asthe thickness decreases. This explanation produces thecorrect trends for the alloys if not the correct numericalvalues; viz., that the addition of Cu increases the state ofcompression over pure Sn, and that Pb causes the stressto become less compressive. The quantitative errors maybe due to the difficulties in measuring the deposit con-centrations and/or the assumption that interstitial Cu

Table 3Effect of alloying addition on initial stress of 16 lm deposits

Deposit material Measured initialstress (MPa)

Correspondingstress-free strain

Sn �15.5 2.3 · 10�4

Sn–2%Pb �4.0 0.6 · 10�4

Sn–3%Cu �36.5 5.6 · 10�4

Note from the Appendix that e0f ¼ eTf � rfE0f

¼ � 4

3

d

L2dS �

rfE0f

.

has zero partial molar volume. A positive value woulddecrease the proposed amount of expansion due toCu6Sn5 precipitation listed in the last column of Table 3and improve the agreement.

5.3. Time dependence of beam deflection – average creepresponse

With time the precipitates coarsen, a process assumedto not alter the deposit stress. To account for the stressinduced in the deposit due to the IMC formation on theinterface between the deposit and the substrate requiresthe use of the model and equations in the Appendix.This analysis necessitates the choice of a value for thestress in the IMC layer, rIMC (more properly the stressfree strain, e0IMC). The most negative choice for the inter-metallic stress free strain is obtained using 1/3 the valueof the volumetric strain for planar IMC formation frompure Sn and Cu; viz., e0IMC ¼ �0:02 (see Section 5.1).Using the modulus for the IMC in Table 2, rIMC �+4.8 GPa. This stress value is an upper bound becausethe IMC may not swell isotropically during growth;e.g., it might expand more in the growth direction to re-duce the strain energy. A second estimate is obtainedexperimentally from in situ wafer curvature measure-ments after the deposition of 100 nm of Sn onto a Cucoated glass cantilever [16]. After 20 nm of IMC hasformed, a beam deflection consistent with a tensile stressin the IMC of +1.2 GPa was obtained. A more appro-priate estimate for thicker IMC layers comes from thefollowing experiments. For three pure Sn cantilever elec-trodeposits, 16 lm thick and aged for approximately3 · 107 s, the Sn was removed by dissolution in concen-trated HCl. This procedure leaves the approximately2.5 lm thick IMC behind. The deflections of the beams(compared to the original unplated position) were then+7, +30 and +37 lm; i.e., toward the deposit side. Fromthese experiments, we conclude that rIMC is between+11 and +62 MPa (e0IMC ¼ �10�4 and �5 · 10�4). Leeand Lee [11] performed similar stripping experimentsbut unfortunately did not report the IMC stress. Dueto the uncertainty in the proper value of e0IMC, we willcompute the stress in the electrodeposit using the deflec-tion data and several assumed values of e0IMC.

Deviation of stress-freestrain from that forpure Sn

1/3 of volume changedue to ppt. of Cu6Sn5or Pb from Cu or Pbsupersaturated Sn

– –�1.7 · 10�4 �4 · 10�4

3.3 · 10�4 7 · 10�3

Page 13: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 15. Deposit stress vs. time extracted from deflection data for pureSn, 7 lm thick samples. Curves are shown from top to bottom for fourdifferent values of the intermetallic stress-free strain, e0IMC: 0,�1 · 10�3, �2 · 10�3, �3 · 10�3.

Fig. 16. Strain vs. time histories. The total deposit strain is propor-tional to the measured deflection. Elastic strain and stress free + plasticstrain are extracted from deflection data using the model for twoassumed values of intermetallic stress-free strain: (a) 0 and (b)�3 · 10�3. Note that in (b) the elastic strain, which is proportionalto the stress, is compressive for all times even though the deflectionchanges sign.

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5045

Using the measured deflections and the model andequations in the Appendix, we plot in Fig. 15 the depositstress, rf, as a function of time for the 7 lm thick pureSn deposit for several values of e0IMC. With e0IMC ¼ 0,the deposit stress changes sign from compression to ten-sion at the same time that the beam deflection changessign; the deflection is proportional to rfdf + rIMCdIMC,(Eqs. (A1) and (A3)). For a value e0IMC ¼ �3� 10�3,the stress remains compressive for all time despite thefact that the deflection changes sign. Thus, the idea pre-sented by several authors that IMC formation on theinterface between the substrate and the deposit com-presses the remaining Sn is confirmed. Fig. 16 showsthe different parts of the deposit strain for two of thecases shown in Fig. 15. The total deposit strain, eTf , isproportional to the deflection (Eqs. (A1) and (A4))and is positive if the deflection is negative. The relativesizes of the elastic and plastic parts derived from themodel, eelf and eplf , depend on the assumed value ofe0IMC. The elastic strain is always proportional to thestress.

The plastic strain rate is determined by differentiationof the plastic strain and a plot of log(�strain rate) vs.log(�stress) can be constructed. Such plots constitutethe experimentally obtained average compression creepresponse of the deposits if we can assume that steady-state creep behavior applies at each instant. Reassur-ingly, use of a positive value of e0IMC (all measurementsand estimates yield negative values) leads to unrealisticplastic response; the sign of the stress and strain rateare not the same. Extremely negative values of e0IMC

(e.g., �2 · 10�2) lead to equally unphysical creep curveswhere the stress exponent increases with decreasingcompressive stress. These considerations place boundson the value of e0IMC.

The creep curves for the 16 lm thick deposits areshown in Fig. 17 for three values of e0IMC, 0, �0.001,�0.002 that span the full range of reasonable valuesfor the 16 lm deposits. This corresponds to IMC stressbetween +0 and +244 MPa. One slides down the creepcurve as time progresses. For large compressive stress(early times, thin IMC), the curves are independent ofe0IMC and exhibit stress exponents of approximately 3,6, and 15 for the Sn–Pb, Sn and Sn–Cu, respectively.We also note that Sn–Pb creeps faster than pure Sn,which in turn creeps faster than Sn–Cu. For smaller val-ues of compressive stress (later time, thicker IMC), thecreep curves show considerable dependence on e0IMC.The curves can be used to show just how much the de-posit compression is increased by the intermetallic bydrawing a horizontal line. Overall this analysis suggeststhat IMC formation on the Cu/Sn interface does notgreatly increase the compressive stress over the leveldue to the initial plating stress in these experiments.

Page 14: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

Fig. 17. log(�plastic strain rate) vs. log(�stress) for 16 lm thickelectrodeposits derived from deflection measurements. For eachmaterial: bottom curve, no correction for layer intermetallic stress;center curve, corrected for layer intermetallic stress, eIMC

0 ¼ �0:001(stress = 125 MPa tensile); top curve, eIMC

0 ¼ �0:002 (stress = 250MPa tensile). The prediction of [39] for whisker growth rate (convertedto average strain rate in the in-plane direction) is shown by dashed line.

Fig. 18. Region of a deposit near its surface that contains a small grainwith interior oblique faces surrounded by columnar grains under bi-axial compressive stress. The three-dimensional path of Sn atoms torelieve the stress is along columnar grain faces (arrows) within thedeposit, through the triple lines, and across the oblique (black) faces.Accretion of Sn on these faces forces the whisker above the depositsurface.

5046 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

5.4. Localized creep as a mechanism for whisker/hillock

formation

We presume that the creep curve for Sn–Pb shown inFig. 17 represents uniform creep alone because no whis-kers, cones or hillocks form. The electrodeposit wouldswell slightly and uniformly in thickness as the compres-sive stress relaxes. For incompressible plasticity, theplastic strain (strain rate) normal to the electrodepositsurface equals minus two times the in-plane strain(strain rate). On the other hand, the creep curves forSn and Sn–Cu include the non-uniform creep that pro-duces the surface disruption. In other words the hillockand whisker formation is part (or all) of the compressioncreep response of these materials when constrained by asubstrate. We note that the creep rates of the pure Snand Sn–Pb deposits are much slower than those re-ported for bulk samples of similar compositions withthe same grain size [37,38].

As mentioned in Section 1, diffusional creep has beenemployed to explain hillock growth in thin films. Fol-lowing this approach, Tu [10] and Hutchinson et al.[39] have proposed models for whisker growth velocitybased on an imposed constant in-plane compressivestress and grain boundary diffusion of Sn to the whiskerbase. Grain boundary diffusion is expected to be thedominant self diffusion mechanism in Sn because thehomologous temperature for Sn corresponding to roomtemperature is 0.6. With whiskers a distance 2b apartand with diameter 2a, both models arrive at a tip growthvelocity v given by

v ¼ �‘2DgbX

RTa2 lnðb=aÞ

� �rf ; ð7Þ

where X is the molar volume of Sn, Dgb a grain bound-ary diffusion coefficient, RT is the Boltzmann factor anda signed stress, rf, has been used. A stress gradient in thedeposit with cylindrical symmetry under each whisker isassumed to drive the flux of Sn toward that whisker. Thelength differs in the two models significantly dependingon the geometrical details. The model of Hutchinsonet al., which uses 2 ‘Dgb/g, the deposit thickness, and re-places Dgb with Dgb

2kg , where k and g are the effective

thickness of a grain boundary and the grain size, respec-tively, leads to the correct order of magnitude of whiskergrowth rate using the literature accepted values of the Sngrain boundary diffusion coefficient [40]. From thesemodels, one can compute the plastic strain rate normalto the deposit if it were averaged over the entire deposit

Page 15: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5047

surface as va2/b2df; giving an in-plane creep rate for anincompressible deposit of �va2/2b2df or

_ef ¼‘

d f

XDgb

RTb2 lnðbaÞrf . ð8Þ

We note that this prediction is linear in stress and wouldproduce a creep rate with slope of unity as superimposedin Fig. 17, where we have used ‘ = df, kDgb = 6 · 10�22

m3/s [40], g = 3 lm, X = 1.7 · 10�5 m3/mol, a = 2 lmand b = 100 lm.

It is interesting to note, despite the approximations ofthis whisker growth model, that the average strain ratederived from this model is roughly tangential to thecreep curves for the pure Sn, lies below the creep curvesfor Sn–Pb and cuts through the creep curve for Sn–Cu.As a mechanism of creep, whisker growth would then beleast active for Sn–Pb, be marginal for pure Sn and oc-cur readily for Sn–Cu. For Sn–Cu, creep likely occurs bya mechanism with high exponent before whisker creepbecomes dominant at lower compressive stress. Thismay be related to the observation that whisker forma-tion appears to have an ‘‘incubation time’’.

The above calculation for whisker growth rate needsimprovement. A detailed model is beyond the scope of

Fig. 19. Due to the absence of transverse grain boundaries in the columnarrelieved by hillock or whisker growth, respectively, depending on whether thstructure of Sn–Pb, uniform creep is possible and there is no need to form s

the present paper, but it would be a special case of Coblecreep [41]. The diffusion potential (chemical potential ofSn atoms minus chemical potential of vacancies) is high-est for grain faces with the most compressive normalstress. This would correspond to grain faces with nor-mals lying in the plane of the deposit due to the bi-axialin-plane stress state; i.e., the columnar grain faces. Obli-que grain faces would have lower diffusion potential andgrain faces parallel to the deposit surface would be least.(The surface of the deposit would also have the least dif-fusion potential, but we presume surface diffusion is sup-pressed by native oxide.) This difference in diffusionpotential causes a flux of Sn along the normal grainfaces to oblique faces. If many oblique or parallel facesare located within the deposit (as with the Sn–Pb depos-its), creep is relatively uniform and no hillocks/whiskerswould occur. Averaged over many grains this uniformcreep process leads to a slight thickening of the depositand a reduction in the in-plane stress without whiskerformation.

If oblique or parallel faces occur rarely, as in thecolumnar structures of the Sn and Sn–Cu samples, thediffusion flux is forced toward the free surface. Any ob-lique grain face near, but under, the deposit surface

grain structures of (a) Sn or (b) Sn–Cu, in-plane compressive stress ise grain boundaries are (a) mobile or (b) pinned. In the more equi-axedurface protrusions.

Page 16: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

5048 W.J. Boettinger et al. / Acta Mate

would be ripe for accumulation of Sn and pushing of thegrain upward to produce surface features, such as hill-ocks and whiskers. As with any diffusional creep pro-cess, grain boundary sliding and accommodation mustoccur simultaneously [42]. We note however that Shenget al. [22] observe whisker growth from matte Sn andSn–Cu deposits even though the grain structure is notcolumnar. Thus, avoiding columnar grain structuresshould be considered only part of a general whisker mit-igation strategy.

A proper model for whisker growth from a columnargrain structure must assess the grain boundary diffusionpath through a three-dimensional network of columnargrain faces connecting to oblique grain faces near the de-posit surface, for example at a small surface grain asshown in Fig. 18. Whisker growth would only be possi-ble at a few small surface grains formed at the end of thedeposition process or possibly by recrystallization afterplating. Fig. 19 shows different situations for the re-sponse of a small surface grain depending on the generalgrain shape and mobility of the grain boundaries in theelectrodeposits. Fig. 19(a) shows the case for pure Sndeposits with a columnar grain structure in which theactive surface grain can broaden laterally by grainboundary migration as it is pushed upward. The local-ized creep is thus spread over a larger surface area; neverforming a filamentary whisker. In a film under stress, theseeds of a model for this lateral grain boundary motionand surface upheaval on the trailing side can be found in[43]. Many hillocks appear to have a small flat top attheir center indicating the original location of the hillockgrain prior to the migration of its boundaries. For verythin deposits, hillock base broadening may be haltedwhen the hillock grain bottom has impinged on the layerIMC. If sufficient compressive stress remains, the hillockmay then grow in a more whisker-like manner. Such aprocess may be important for deposits with thicknessthe same as the lateral grain size.

In Fig. 19(b) is depicted the situation for the Sn–Cu deposits where the grain boundaries are pinnedor impeded by IMC particles. These particles areformed by precipitation from Sn supersaturated solidsolution formed by co-deposition of Sn and Cu fromthe electrolyte. Hillock grain base broadening is hin-dered and the upward motion of material is concen-trated into contorted hillocks or whiskers. Contortedhillocks occur when several surface grains with obli-que internal faces in close proximity are present;hence, the multi-grained nature seen in Fig. 6. Whis-kers occur when a single surface grain with immobileoblique internal faces persists in time for the Sn fluxto continually accrete. Fig. 19(c) shows the case ofthe non-columnar structure of Sn–Pb where manyinternal grain boundaries are available for accretionof Sn producing a more uniform thickening of the de-posit to relieve compressive stress.

6. Conclusions

1. Plating on cantilever beam samples shows that com-

rialia 53 (2005) 5033–5050

pressive stresses are present in high purity bright Snelectrodeposits within 15 min after plating. Interme-tallic growth along the Sn/Cu interface should beinsignificant (6 nm thickness) over this short time atroom temperature. Thus, compressive stresses areintrinsic to the plating process under the conditionsof our experiments.

2. Co-deposition of Sn with Cu or Sn with Pb in theamount of a few mass percent alters the compressivestress compared to pure Sn by the time of first mea-surement. Larger compressive stress is generated inthe electrodeposit with Cu addition to the electrolyte;smaller compressive stress is generated in the depositwith Pb addition to the electrolyte. The increase/decrease in compressive stress for the alloys is dueto volume changes of the deposit caused by the solidstate precipitation of Cu6Sn5/Pb particles by metal-lurgical precipitation from the supersaturated Sn–Cu/Sn–Pb solid solution formed by the platingprocess.

3. Cu addition does not alter the columnar grainstructure found in pure Sn deposits, whereas thePb addition causes an equi-axed grain structure todevelop.

4. The analysis of the time dependence of the measuredbeam deflections indicates that the additional com-pressive stress due to intermetallic formation on thedeposit/beam interface over long time periods is ofthe same order of magnitude as the plating stress. Itis not the sole source of compressive stress.

5. Whiskers and contorted hillocks are observed on theSn–Cu deposit surfaces, conical hillocks are observedon the pure Sn deposits and no surface disturbancesare found on the Sn–Pb deposits.

6. We propose that the compressive stress in the depos-its relaxes by Coble creep (grain boundary transportof Sn and vacancies). For the equi-axed structure ofthe Sn–Pb deposits, transport between interior grainfaces with different normal stress permits uniformswelling of the deposit thickness and no localized sur-face disturbance. For the columnar structures of theSn and Sn–Cu deposits, the Sn flows toward the freesurface. In the absence of surface transport due tonative oxide, the Sn accumulation is localized to obli-que grain faces near but under the free surface thatforces the grain upward.

7. The lack of lateral grain boundary mobility of thecolumnar structure in the Sn–Cu deposits caused byprecipitate pinning is thought to promote whiskerformation by maintaining a laterally small grainwhere Sn flux accumulates. Rather benign (electroni-cally) hillock formation occurs in pure Sn due toextensive lateral grain boundary migration.

Page 17: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050 5049

8. A suggested whisker mitigation strategy based on thiswork would be to avoid impurities in the electrolytesuch as Cu that are fast diffusers and hence occupyinterstitial sites in the Sn lattice. Electrolytes contain-ing solutes occupying substitutional sites in the Snlattice and also whose precipitation from supersatu-rated solid solution tends to reduce the volume ofthe deposit should be beneficial.

9. A second suggestion is to employ plating conditionsand/or additives to the electrolyte that break up thecolumnar structure and promote the formation ofgrain boundaries parallel to the deposit surface. Suchboundaries would allow a more uniform creep ofelectrodeposits to relieve compressive stress.

Acknowledgments

The authors are indebted to R. Parke and S. Claggettfor the TEM sample preparation, to R. Fields for thephosphor bronze mechanical measurements, U.R. Katt-ner for thermodynamic calculations and to Daniel Josellfor many fruitful discussions regarding beam deflection.Thanks are also in order to G. Galyon and M. Palmerfor the FIB pictures, which greatly clarified the micro-structures and assisted this research.

Appendix. Time-dependent deflection of a beam with two

thin films

To estimate the stresses in each layer, which cannotbe individually determined from the beam deflection,we develop a model to describe the time-dependentdeflection of the beam. The deflection is due to threecauses: (1) an initial residual stress in the deposit dueto plating and fast precipitation processes, that is as-sumed independent of time; (2) stress due to the growthof a layer of IMC that forms between the Sn and Cusubstrate that has a different volume than the Sn andCu from which it formed; (3) relaxation due to creep.The model assumes that the creep is uniform. For thismodel, we will employ the thin deposit approximation,to avoid unnecessary complication in the exposition.

Let ri, di, ei, E0i be the stress, thickness, strain and

modulus in layer i, where i = IMC or f for the planarIMC layer and electrodeposited film, respectively. Letthe beam curvature be given by

KðtÞ ¼ 1

rðtÞ ¼2dðtÞL2

; ðA1Þ

where d(t) > 0 if the electrodeposit surface is concave(deposit on top, curved up). To treat the fact that theIMC layer thickens with time and the Sn layer thins withtime, we use:

dIMCðtÞ ¼ffiffiffiffiffiBt

p;

d f ¼ d0f � sdIMCðtÞ;

dS ¼ d0S � ð1� sÞdIMC;

ðA2Þ

where s � 5/11. For dIMC and df � dS, balance of mo-ments and forces for the substrate and the two layersyields

rIMCðtÞdIMCðtÞ þ rfðtÞd fðtÞ ¼E0S

6ðdSðtÞÞ2KðtÞ. ðA3Þ

With the thin approximation, the total strain in bothlayers (measured from a flat beam) is the same and is gi-ven by

eTIMC ¼ eTf ¼ � 2

3dSðtÞKðtÞ. ðA4Þ

The total strain in the IMC layer will be assumed to bethe sum the stress-free strain due to its volume changefrom the reactants and its stress-dependent elastic strain(no creep)

eTIMC ¼ e0IMC þ eelIMC ðA5Þwith

rIMC ¼ E0IMCe

elIMC ¼ E0

IMC½eTIMC � e0IMC�. ðA6ÞThe total strain in the Sn deposit will be assumed to bethe sum of the stress free strain due to the electrodepos-ition process (plus the fast precipitation processes in thealloys), stress-dependent elastic and plastic strains

eTf ¼ e0f þ eelf þ eplf ðA7Þwith

rf ¼ E0fe

elf ¼ E0

f ½eTf � e0f � eplf � ðA8Þand

_eplf ¼ f ðrfÞ; ðA9Þwhere the dot denotes a time derivative. The functionf(rf) could represent power law creep for example ofthe form

_eplf ¼ Krnf . ðA10Þ

For choices of a parabolic growth constant B, and a va-lue for e0IMC, one can extract rf(t) and the quantity½e0f þ eplf ðtÞ� or its time derivative _eplf ðtÞ from the measureddata, d(t). On can then plot log strain rate vs. log stressfor the electrodeposit to estimate the function, Eq. (A9).Because the stress and strain rates are negative (com-pression creep), absolute values are used for Fig. 17.

Alternately, one can convert these equations to a sys-tem of coupled ordinary differential equations for d(t),rf(t), rIMC(t) and eplf ðtÞ for a given creep law and theother materials parameters. Numerical solution showshow the electrodeposit stress is affected by the growingIMC. In general, the stress in the film becomes slightly

Page 18: doi:10.1016/j.actamat.2005.07 - NASA · 2008. 4. 8. · Title: doi:10.1016/j.actamat.2005.07.016 Created Date: 10/12/2005 9:57:08 AM

5050 W.J. Boettinger et al. / Acta Materialia 53 (2005) 5033–5050

more compressive due to the IMC growth. The amountis sensitive to the how positive is the chosen value ofe0IMC.

References

[1] Furuta N, Hamamura K. Jpn J Appl Phys 1969;8:1404.[2] Chason E, Kabakian L, Kumar S. Materials for space applica-

tions. In: Chipara M, Edwards DL, Benson RS, Phillips S, editors.Mater Res Soc Symp Proc, vol. 851, Warrendale, PA; 2005. p.NN5.17.

[3] Fisher RM, Darken LS, Carroll KG. Acta Metall 1954;2:368.[4] Arnold SM. Plating 1966;53:96.[5] Galyon GT. Annotated Tin Whisker Bibliography, National

Electronics Manufacturing Initiative Inc. (NEMI), Herdon, VA,July; 2003. Available from: www.nemi.org.

[6] Chadhari P. J Appl Phys 1974;45:4339.[7] Jackson MS, Li C-Y. Acta Metall 1982;30:1993.[8] Doerner MF, Nix WD. CRC Crit Rev Solid State Mater Sci

1988;14:225.[9] Tu KN. Acta Metall 1973;21:347.[10] Tu KN. Phys Rev B-Cond Matter 1994;49:2030.[11] Lee BZ, Lee DN. Acta Mater 1998;46:3701.[12] Moon K-W, Williams ME, Johnson CE, Stafford GR, Handwer-

ker CA, Boettinger WJ. In: Hanada S, Zhong Z, Nam SW, WrightRN, editors. Proceedings of the fourth pacific rim internationalconference on advanced materials and processing. Sendai,Japan: The Japanese Institute of Metals; 2001. p. 1115–8.

[13] Weil R. Plating 1971;58:137.[14] Onishi M, Fujibuchi H. Trans J Inst Metal 1975;16:539.[15] Tu KN, Thompson RD. Acta Metall 1982;30:947.[16] Stafford GR, Bertocci U, Guyer JE, NIST; 2004 [unpublished

research].[17] Bard AJ, Faulkner LR. Electrochemical methods: fundamentals

and applications. New York, NY: Wiley; 1980. p. 288.[18] Galyon GT, Palmer M, IBM; 2004, unpublished images of NIST

samples.

[19] Petersson I, Ahlberg E. J Electroanal Chem 2000;485:166.[20] Tindall GW, Bruckenstein S. Anal Chem 1968;40:1402.[21] Brenner A. Electrodeposition of alloys, vol. 1. New York,

NY: Academic Press; 1963. p. 195.[22] Sheng GTT, Hu CF, Choi WJ, Tu KN, Bong YY, Nguyen L. J

Appl Phys 2002;92:64.[23] Kane RH, Giessen BC, Grant NJ. Acta Metall 1966;14:605.[24] Kim D-K, Nix WD, Deal MD, Plummer JD. J Mater Res

2000;15:1709.[25] Brenner A, Senderoff S. J Res NBS 1949;42:105.[26] Smithells CJ. Metal reference book. 4th ed. New York, NY: Ple-

num Press; 1967.[27] Fields RJ, Low III SR, Lucey Jr GK. In: Cieslak MJ, Perepezko

JH, Kang S, Glicksman ME, editors. Metal science of join-ing. Warrendale, PA: The Minerals, Metals & Materials Society;1992. p. 165–73.

[28] Saunders N, Miodownik AP. Bull Alloy Phase Diagram1990;11:278.

[29] Homer CE, Plummer H. J Inst Metal 1939;64:169.[30] Kattner UR. NIST, Gaithersburg, MD; 2002 [unpublished

research].[31] Dyson BF, Anthony TR, Turnbull D. J Appl Phys 1967;38:3408.[32] Warburton WK, Turnbull D. Thin Solid Films 1975;25:71.[33] Karakaya I, Thompson WT. Bull Alloy Phase Diagram

1988;9:144.[34] Lee JA, Raynor GV. Proc Phys Soc 1954;67:737.[35] Yeh DC, Huntington HB. Phys Rev Lett 1984;53:1469.[36] Floro JA, Chason E, Cammarata RC, Srolovitz D. MRS Bull

2002;27:19.[37] Schneibel JH, Hazeldine PM. J Mater Sci 1983;18:562.[38] McCabe RJ, Fine ME. Metall Mater Trans 2002;33A:1531.[39] Hutchinson B, Oliver J, Nylen M, Hagstrom J. Mater Sci Forum

2004;467–470:465.[40] Kaur I, Gust W. Handbook of grain boundary and interface

boundary diffusion data, vol. 2. Stuttgart: Ziegler Press; 1989. p.1277.

[41] Coble RL. J Appl Phys 1963;34:1679.[42] Raj R, Ashby MF. Metall Trans 1971;2:1113.[43] Genin FY. J Appl Phys 1995;77:5130.