10.07.2015 Views

Fabrication, microstructure, and mechanical properties of tin ...

Fabrication, microstructure, and mechanical properties of tin ...

Fabrication, microstructure, and mechanical properties of tin ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

5828 M.J. Burek et al. / Materials Science <strong>and</strong> Engineering A 528 (2011) 5822–5832aFlow stress at 5% strain (MPa)100010010~ 920 nm~ 560 nm~ 350 nm10 -5 10 -4 10 -3 10 -2 10 -1Engineering strain rate (s-1 )bEngineering strain rate (s -1 )10 -110 -210 -3~ 5.5410 -4~ 920 nm~ 560 nm10 -510 100 1000Flow stress at 5% strain (MPa)Fig. 10. (a) Log–log plot <strong>of</strong> engineering flow stress measured at ∼5% strain as a function <strong>of</strong> engineering strain rate for all <strong>tin</strong> nanopillars tested in this work. (b) Log–log plot<strong>of</strong> averaged strain rate as a function <strong>of</strong> average engineering flow stress measured at ∼5% strain for <strong>tin</strong> nanopillars. The error bars indicate one st<strong>and</strong>ard deviation.respectively. These specimens were compressed with the sameengineering strain rate <strong>of</strong> ∼0.001 s −1 . The stress–strain curves showthat the nanopillars exhibit strain bursts or s<strong>of</strong>tening events afterinitial yielding. Such strain bursts are believed to be associatedwith sudden catastrophic shear failures along crystallographic slipplanes.3.3. Post compression SEM analysisTo further underst<strong>and</strong> how small-scale <strong>tin</strong> structures behaveplastically, all nanopillar specimens were carefully inspected before<strong>and</strong> after compression using high resolution SEM. Typical electronmicrographs <strong>of</strong> <strong>tin</strong> nanopillars with average diameters <strong>of</strong>920 nm, 560 nm, <strong>and</strong> 350 nm before <strong>and</strong> after uniaxial compressionat strain rate <strong>of</strong> ∼0.001 s −1 are shown in Figs. 7–9, respectively.These figures show two dis<strong>tin</strong>ct plastic flow features –crystallographic shear <strong>of</strong>f-sets <strong>and</strong> material extrusion. Both deformationfeatures were observed for all three nanopillar sizes. Eventhough fine crystallographic slip lines are not apparent on thepost compression images, examples <strong>of</strong> gross deformation by shearalong crystallographic planes can be found in these SEM micrographs.This type <strong>of</strong> crystallographic shear behavior has beenreported in almost all deformed single-crystalline metallic nanopillars[14–25,27–31,38–41]. Another deformation characteristic thatcan be found in Figs. 7–9 is sidewall surface wrinkles <strong>and</strong> bulgesperpendicular to the loading direction. All compressed <strong>tin</strong> nanopillarsshowed this type <strong>of</strong> plastic deformation where the compressivestrain is compensated by material extrusions from the nanopillarsidewalls. The material extrusion type <strong>of</strong> deformation in <strong>tin</strong>nanopillars is not unexpected. At high homologous temperatures, alarge number <strong>of</strong> slip systems will be activated which allows deformationin a greater number <strong>of</strong> crystallographic slip planes to occursimultaneously. The consequence is that fewer discrete slip planesare observed in the post-compression nanopillar, <strong>and</strong> instead strainresembles extrusion like deformation.3.4. Influence <strong>of</strong> strain rate <strong>and</strong> nanopillar sizeEngineering flow stresses <strong>of</strong> <strong>tin</strong> nanopillars measured at 5.0%nominal strain are plotted as a function <strong>of</strong> strain rate in Fig. 10(a).The data in this plot includes results from nanopillar compressions<strong>of</strong> three different diameters: 350 nm, 560 nm, <strong>and</strong> 920 nm. Forthe two largest nanopillar sizes <strong>of</strong> 560 nm <strong>and</strong> 920 nm, the resultsclearly show that the <strong>tin</strong> flow stresses are sensitive to the deformationrate. The data reveals that the material strength is low whenthe structure is deformed slowly. In contrast, the yield strengthincreases when the specimen is compressed at a fast strain rate. Inorder to show a clear trend <strong>of</strong> strain rate sensitivity, the averageflow stress values <strong>of</strong> 560 nm <strong>and</strong> 920 nm diameter nanopillars areplotted in Fig. 10(b), where the error bars correspond to one st<strong>and</strong>arddeviation. Fig. 10(b) shows a noticeable increase in 920 nmdiameter <strong>tin</strong> nanopillar strength from ∼46 to 112 MPa when thestrain rate increases from 0.0001 s −1 to 0.01 s −1 . Similarly, for the560 nm diameter <strong>tin</strong> nanopillars, the data shows an increase from∼61 to 96 MPa for the same increase in strain rate.The general relationship between the uniaxial strain rate <strong>and</strong>creep stress has been well documented <strong>and</strong> is displayed in Eq. (1).To compare the <strong>tin</strong> nanopillar strain rate sensitivity with the bulksamples from the previous studies, the data from Fig. 10(b) arefitted with Eq. (1). The stress exponent extracted from this curvefit is approximately 5.54 ± 1.53, where the spread corresponds toone st<strong>and</strong>ard deviation. This value agrees with the single-crystalimpression creep results reported from Chu <strong>and</strong> Li [6], with exponentvalues measured near room temperatures in the range <strong>of</strong>4.4–5.0. The results indicate that the strain rate sensitivity <strong>of</strong> singlecrystal<strong>tin</strong> nanostructures is similar to bulk.Fig. 11 shows the <strong>tin</strong> nanopillar engineering flow stress resultsmeasured at 5.0% nominal strain plotted as a function <strong>of</strong> featuresize. These structures were compressed at a constant strain rate <strong>of</strong>Flow stress at 5% strain (MPa)1000100~ D -0.572σ-1strain rate ~ 0.001 s10100 1000Pillar diameter (nm)Fig. 11. Log–log plot <strong>of</strong> engineering flow stress measured at ∼5% strain as a function<strong>of</strong> pillar diameters. These samples were compressed at a ∼0.001 s −1 strain rate. Theerror bars indicate one st<strong>and</strong>ard deviation.


M.J. Burek et al. / Materials Science <strong>and</strong> Engineering A 528 (2011) 5822–5832 5831kling/extrusion are both observed in the post-compression SEMimage. This is consistent with the deformation mechanisms observationsin Figs. 7–9. The stress–strain data collected from uniaxialcompression <strong>of</strong> this <strong>tin</strong> nanopillar is plotted in Fig. 12(c). Thestress–strain behavior includes similar strain bursts to those illustratedpreviously in Fig. 6. It is important to note that this particularpillar was only characterized by SXRD after the compression test.No information was collected in the as-fabricated state.The SXRD analysis <strong>of</strong> the deformed <strong>tin</strong> nanopillar againidentified only one unique body-centered tetragonal (BCT) Lauediffraction pattern for the entire nanopillar volume, indica<strong>tin</strong>g thecompressed nanopillar is single-crystalline. The Laue diffractionpattern indicates the compressed <strong>tin</strong> nanopillar is BCT with theplane normal to the vertical axis <strong>of</strong> the pillar close to the <strong>tin</strong> (2 0 4)plane. Two individual Laue diffraction spots <strong>of</strong> (2 0 4) <strong>and</strong> (1 0 5)were extracted from this pattern <strong>and</strong> plotted in Fig. 12(a) <strong>and</strong> (b),respectively. The shapes <strong>of</strong> these two Laue spots are fairly regularwith circular geometry sugges<strong>tin</strong>g r<strong>and</strong>om distribution <strong>of</strong> dislocationsdue to the nanopillar deformation. Qualitatively, the Lauediffraction peaks <strong>of</strong> this deformed specimen are similar to the asfabricated<strong>tin</strong> nanopillar shown in Fig. 4(a) <strong>and</strong> (b).Qualitative analysis <strong>of</strong> the (2 0 4) Laue diffraction peak inFig. 13(a) were also conducted along the intensity traces at aparticular <strong>and</strong> along the 2 axis. This pr<strong>of</strong>ile was fitted withLorentzian curves as shown in Fig. 13(c). The FWHM value <strong>of</strong> thispeak broadening is approximately 0.458 ◦ for the compressed <strong>tin</strong>nanopillar. When compared with the peak broadening between asfabricated<strong>and</strong> post compression spots shown in Figs. 5(b) <strong>and</strong> 13(c),the results show that there is an angular width difference <strong>of</strong>∼0.019 ◦ . This variation is within the experimental error <strong>and</strong> resolution<strong>of</strong> the SXRD technique. The angular resolution in theSXRD experiments is limited by the charge-coupled device (CCD)camera pixel size which translates to resolution limit <strong>of</strong> ∼0.03 ◦in the Laue spot width measurement. No significant change inthe peak broadening is expected here from the small crystalsize, as well as the instrumentation before or after the deformation.It is then also reasonable to propose that there is nosignificant change in the peak broadening due to r<strong>and</strong>om dislocationdensity associated with the nanopillar deformation. TheLaue diffraction results discussed here suggest that the defect densityin this <strong>tin</strong> nanopillar after compression is very similar to theas-fabricated <strong>tin</strong> nanopillar described in Fig. 4. This may indicatethat the nucleation <strong>and</strong> multiplication rates <strong>of</strong> dislocations duringthe uniaxial compression process are <strong>of</strong>fset by the dislocationannihilation rate at the nanopillar surface. Thus, there is no significan<strong>tin</strong>crease or accumulation <strong>of</strong> dislocations in these small<strong>tin</strong> structures. Such a finding is similar to that reported earlier byBudiman et al. [47] in an ex situ study <strong>of</strong> single-crystalline goldnanopillars fabricated from by FIB milling, despite different crystalstructures.4. ConclusionsIn conclusion, we have developed fabrication <strong>and</strong> integrationtechniques to produce large grain BCT <strong>tin</strong> nanopillars with diametersas small as 70 nm. Tin nanopillar flow stress data measuredat different deformation rates indicate that these nanostructurespossess similar strain rate sensitivity to their bulk counterpart,where the strength <strong>of</strong> <strong>tin</strong> increases with deformation rate. Additionally,the strength <strong>of</strong> <strong>tin</strong> nanopillars was observed to increasewith reduced diameter. This flow stress size dependence appearsto have the same characteristics as other single-crystalline FCCmetals tested in uniaxial compression at the nanoscale. Microstructuralcharacterization by SXRD indicates that there is no drasticincrease <strong>of</strong> dislocation within the compressed <strong>tin</strong> nanopillars. Thissuggests that the rate <strong>of</strong> dislocation generated by the deformationprocess is <strong>of</strong>fset by annihilation at the nanopillar surfaceAcknowledgementsT.Y. Tsui thanks Canadian NERSC Discovery, NERSC ResearchTools <strong>and</strong> Instruments, <strong>and</strong> the Canada Foundation for Innovation(CFI) for the financial support <strong>of</strong> this research. The authorsgratefully acknowledge critical support <strong>and</strong> infrastructure providedfor this work by the Department <strong>of</strong> Energy (DOE), Office<strong>of</strong> Science, Office <strong>of</strong> Basic Energy Sciences. The Advanced LightSource is supported by the Director, Office <strong>of</strong> Science, Office <strong>of</strong>Basic Energy Sciences, Materials Sciences Division, <strong>of</strong> the U.S.Department <strong>of</strong> Energy under Contract No. DE-AC02-05CH11231 atLawrence Berkeley National Laboratory <strong>and</strong> University <strong>of</strong> California,Berkeley, California. The move <strong>of</strong> the micro-diffraction programfrom ALS beamline 7.3.3 onto to the ALS superbend source 12.3.2was enabled through the NSF grant #0416243. One <strong>of</strong> the authors(ASB) is supported by the Director, Los Alamos National Laboratory(LANL), under the Director’s Postdoctoral Research Fellowshipprogram (#20090513PRD2).References[1] M. Abtew, G. Selvaduray, Materials Science <strong>and</strong> Engineering R: Reports 27(2000) 95–141.[2] J. Glazer, International Materials Reviews 40 (1995) 65–93.[3] W.J. Plumbridge, Materials at High Temperatures 17 (2000) 381–387.[4] H. Mukaibo, T. Momma, M. Mohamedi, T. Osaka, Journal <strong>of</strong> the ElectrochemicalSociety 152 (2005) A560–A565.[5] F. Yang, J.C.M. Li, Journal <strong>of</strong> Materials Science: Materials in Electronics 18 (2007)191–210.[6] S.N.G. Chu, J.C.M. Li, Materials Science <strong>and</strong> Engineering 39 (1979) 1–10.[7] O.D. Sherby, Acta Metallurgica 10 (1962) 135–147.[8] R.E. Frenkel, O.D. Sherby, J.E. Dorn, Acta Metallurgica 3 (1955) 470–472.[9] J. Weertman, J.E. Breen, Journal <strong>of</strong> Applied Physics 27 (1956) 1189–1193.[10] J.E. Breen, J. Weertman, Transactions <strong>of</strong> the American Institute <strong>of</strong> Mining <strong>and</strong>Metallurgical Engineers 203 (1955) 1230.[11] F. Mohamed, K. Murty, J. Morris, Metallurgical <strong>and</strong> Materials Transactions B 4(1973) 935–940.[12] V. Raman, R. Berriche, Journal <strong>of</strong> Materials Research 7 (1992) 627–638.[13] M.J. Mayo, W.D. Nix, Acta Metallurgica 36 (1988) 2183–2192.[14] M.D. Uchic, D.M. Dimiduk, J.N. Flor<strong>and</strong>o, W.D. Nix, Science 305 (2004) 986–989.[15] M.D. Uchic, D.M. Dimiduk, Materials Science <strong>and</strong> Engineering A 400–401 (2005)268–278.[16] D.M. Dimiduk, M.D. Uchic, T.A. Parthasarathy, Acta Materialia 53 (2005)4065–4077.[17] C.P. Frick, B.G. Clark, S. Orso, A.S. Schneider, E. Arzt, Materials Science <strong>and</strong>Engineering A 489 (2008) 319–329.[18] Z.W. Shan, R.K. Mishra, S.A. Syed Asif, O.L. Warren, A.M. Minor, Nature Materials7 (2008) 115–119.[19] J.R. Greer, W.C. Oliver, W.D. Nix, Acta Materialia 53 (2005) 1821–1830.[20] C.A. Volkert, E.T. Lilleodden, Philosophical Magazine 86 (2006) 5567–5579.[21] J.R. Greer, W.D. Nix, Applied Physics A: Materials Science <strong>and</strong> Processing 80(2005) 1625–1629.[22] D. Kiener, C. Motz, T. Schöberl, M. Jenko, G. Dehm, Advanced Engineering Materials8 (2006) 1119–1125.[23] D. Kiener, W. Grosinger, G. Dehm, R. Pippan, Acta Materialia 56 (2008) 580–592.[24] A.T. Jennings, M.J. Burek, J.R. Greer, Physical Review Letters 104 (2010) 135503.[25] K.S. Ng, A.H.W. Ngan, Acta Materialia 56 (2008) 1712–1720.[26] H. Bei, S. Shim, E.P. George, M.K. Miller, E.G. Herbert, G.M. Pharr, Scripta Materialia57 (2007) 397–400.[27] H. Bei, S. Shim, G.M. Pharr, E.P. George, Acta Materialia 56 (2008) 4762–4770.[28] S. Brinckmann, J.-Y. Kim, J.R. Greer, Physical Review Letters 100 (2008)155502–155504.[29] J.-Y. Kim, J.R. Greer, Applied Physics Letters 93 (2008) 101913–101916.[30] J.-Y. Kim, J.R. Greer, Acta Materialia 57 (2009) 5245–5253.[31] J.-Y. Kim, D. Jang, J.R. Greer, Scripta Materialia 61 (2009) 300–303.[32] J.-Y. Kim, D. Jang, J.R. Greer, Acta Materialia (2009).[33] G. Lee, J.-Y. Kim, A.S. Budiman, N. Tamura, M. Kunz, K. Chen, M.J. Burek, J.R.Greer, T.Y. Tsui, Acta Materialia 58 (2010) 1361–1368.[34] S.M. Han, T. Bozorg-Grayeli, J.R. Groves, W.D. Nix, Scripta Materialia 63 (2010)1153–1156.[35] J. Greer, J.-Y. Kim, M.J. Burek, JOM 61 (2009) 19–25.[36] J.R. Greer, D. Jang, J.-Y. Kim, M.J. Burek, Advanced Functional Materials 19 (2009)2880–2886.[37] A.S. Schneider, B.G. Clark, C.P. Frick, P.A. Gruber, E. Arzt, Materials Science <strong>and</strong>Engineering A 508 (2009) 241–246.


5832 M.J. Burek et al. / Materials Science <strong>and</strong> Engineering A 528 (2011) 5822–5832[38] A.S. Schneider, D. Kaufmann, B.G. Clark, C.P. Frick, P.A. Gruber, R. Mönig, O. Kraft,E. Arzt, Physical Review Letters 103 (2009) 105501.[39] S.-W. Lee, S.M. Han, W.D. Nix, Acta Materialia 57 (2009) 4404–4415.[40] J.-Y. Kim, D. Jang, J.R. Greer, Acta Materialia 58 (2010) 2355–2363.[41] M.D. Uchic, P.A. Shade, D.M. Dimiduk, Annual Review <strong>of</strong> Materials Research 39(2009) 361–386.[42] M.J. Burek, J.R. Greer, Nano Letters 10 (2010) 69–76.[43] M. Tian, J. Wang, J. Snyder, J. Kurtz, Y. Liu, P. Schiffer, T.E. Mallouk, M.H.W. Chan,Applied Physics Letters 83 (2003) 1620–1622.[44] N. Tamura, A.A. MacDowell, R. Spolenak, B.C. Valek, J.C. Bravman, W.L. Brown,R.S. Celestre, H.A. Padmore, B.W. Batterman, J.R. Patel, Journal <strong>of</strong> SynchrotronRadiation 10 (2003) 137–143.[45] A.S. Budiman, Department <strong>of</strong> Materials Science <strong>and</strong> Engineering, Stanford,2008.[46] B.C. Valek, Stanford University, 2003.[47] A.S. Budiman, S.M. Han, J.R. Greer, N. Tamura, J.R. Patel, W.D. Nix, Acta Materialia56 (2008) 602–608.[48] G. Feng, A.S. Budiman, W.D. Nix, N. Tamura, J.R. Patel, Journal <strong>of</strong> Applied Physics104 (2008) 043501–043512.[49] A.S. Budiman, N. Li, Q. Wei, J.K. Baldwin, J. Xiong, H. Luo, D. Trugman, Q.X.Jia, N. Tamura, M. Kunz, K. Chen, A. Misra, Thin Solid Films 519 (2011)4137–4143.[50] Z. Budrovic, H. Van Swygenhoven, P.M. Derlet, S. Van Petegem, B. Schmitt,Science 304 (2004) 273–276.[51] J.R. Greer, C.R. Weinberger, W. Cai, Materials Science <strong>and</strong> Engineering A 493(2008) 21–25.[52] M. Nagasaka, Japanese Journal <strong>of</strong> Applied Physics 28 (1989) 446–452.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!