Hostname: page-component-7bb8b95d7b-dvmhs Total loading time: 0 Render date: 2024-09-19T17:56:03.422Z Has data issue: false hasContentIssue false

Thermoplastic Deformation and Residual Stress Topography of 4H-SiC Wafers

Published online by Cambridge University Press:  15 March 2011

Robert. S. Okojie
Affiliation:
NASA Glenn Research Center, 21000 Brookpark Road, M/S 77-1, Cleveland, OH 44135, USA
Ming Zhang
Affiliation:
Department of Materials Science & Engineering, CWRU, Cleveland, OH 44106, USA
Get access

Abstract

We have measured thermoplastic deformation in as-received, single-side polished, 4H-SiC wafers and also residual stresses in homoepitaxially grown epilayers on wafers by radius curvature measurements. The wafers studied had n-type resistivities of 0.010-0.011 ω-cm and p-type resistivities of 4.42, 4.72, 9.57 ω-cm. In a first thermal excursion to 900 °C in vacuum, the bow height of the bare substrates in all cases decreased with temperature. Upon cooling down, however, the bow heights remained largely unchanged from their values at 900 °C. A second cyclic excursion to 900 °C did not yield any significant change in the curvature, thus indicating that the substrates had thermoplastically deformed in the first heating cycle. Epilayers having nitrogen doping between 5 × 1017 and 2 × 1019 cm−3 grown on the n- and p-type substrates resulted in compressive stresses ranging between 190 and 400 MPa in the epilayers. Transmission electron microscopy (TEM) examination of the n-type epilayer (with doping levels of 5 × 1017 cm−3 and 5 × 1018 cm−3) on the n-type substrate, revealed bands of stacking faults (SFs) confined within the epilayers after the bicrystals were further annealed at 1150°C in nitrogen for thirty minutes. These doping levels are approximately one and two orders of magnitude below the reported threshold value of 3 × 1019 cm−3 previously suggested for the onset generation of SFs in annealed n-type 4H-SiC epilayers. The calculated residual stresses in all the epilayers were above the critical stress for the motion of dislocations above 1000 °C in 4H-SiC. Thus the SFs that form by glide of pre-existing partial dislocations may actually be stress induced and occur across a much wider range of doping levels. Therefore, it is possible that a significant mechanism for formation of the stacking faults and 3C bands observed in thermally treated 4H-SiC wafer is stress relief via the generation and motion of new and pre-existing partial dislocations on the basal planes of 4H-SiC.

Type
Research Article
Copyright
Copyright © Materials Research Society 2004

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

[1] Tatsumi, M., Hosokawa, Y., Iwasaki, T., Toyoda, N., and Fujita, K., Mat. Sci. and Eng. B 28, 6571 (1994).Google Scholar
[2] Neudeck, P. G., in The VLSI Handbook, edited by Chen, W.-K. (CRC Press and IEEE Press, Boca Raton, Florida, 2000), p. 6.16.24.Google Scholar
[3] Müller, S. G., Glass, R. C., Hobgood, H. M., Tsvetkov, V. F., Brady, M., Henshall, D., Jenny, J. R., Malta, D., and Carter, C. H., J. Cryst. Growth 211, 325332 (2000).Google Scholar
[4] Tsvetkov, V. F., Henshall, D. N., Brady, M. F., Glass, R. C., and Carter, J. C. H., (Mater. Res. Soc. Proc. 512, Pittsburgh, PA 1998) pp. 8999.Google Scholar
[5] Hong, M. H., Samant, A. V., and Pirouz, P., “Stacking Fault Energy of 6H-SiC and 4H-SiC Single Crystals”, Phil. Mag. A, 80(4), 919935 (2000).Google Scholar
[6] Lendenmann, H., Dahlquist, F., Johansson, N., Bergman, J., Bleichner, H., and Ovren, C., in Performance and Reliability of High Power SiC Diodes, Nara Centennial Hall, Nara Japan, 2000 (Research and Development Association for Future Electron Devices), p. 125130.Google Scholar
[7] Okojie, R., Xhang, M., Pirouz, P., Tumakha, S., Jessen, G., and Brillson, L., Mater. Sci. Forum 389–393, 451454 (2002).Google Scholar
[8] Skromme, B. J., Palle, K., Poweleit, C. D., Bryant, L. R., Vetter, W. M., Dudley, M., Moore, K., and Gehoski, T., Mater. Sci. Forum, 389–393, p. 455 (2001).Google Scholar
[9] Lindefelt, U., Iwata, H., Öberg, S., and Briddon, P. R., Phys. Rev. B 67, 155204 (2003).Google Scholar
[10] Miao, M. S., Limpijumnong, S., and Lambrecht, W. R. L., Appl. Phys. Lett. 79, 43604362 (2001).Google Scholar
[11] Liu, J. Q., Chung, H. J., Kuhr, T., Li, Q., and Skowronski, M., Appl. Phys. Lett. 80, 21112113 (2002).Google Scholar
[12] Kuhr, T. A., Liu, J., Chung, H. J., Skowronski, M., and Szmulowic, F., J. Appl. Phys. 92, 58635871 (2002).Google Scholar
[13] Cree, in 4600 Silicon Drive, 27703 ed. (Durham, NC).Google Scholar
[14] Frontier Semiconductor, Inc, 1631 North First Street, San Jose, CA 95112 USA. (www.frontiersemi.com).Google Scholar
[15] Okojie, R. S., Xhang, M., and Pirouz, P., to be published in the Proceedings of the International Conference on Silicon Carbide and Related Materials, October 5-10, 2003, Lyon, France.Google Scholar
[16] Ohring, M., The Materials Science of Thin Films, Academic Press, Inc., New York. (1992).Google Scholar
[17] Ellison, A., Radamson, H., Tuominen, M., Milita, S., Hallin, C., Henry, A., Kordina, O., Tuomi, T., Yakimova, R., Madar, R., and Janzen, E., Diamond and Related Materials 6, 13691373 (1997).Google Scholar
[18] Hull, D. and Bacon, D. J., Introduction to Dislocations, Fourth ed. (Butterworth-Heinemann, Woburn, 2001).Google Scholar