Hostname: page-component-848d4c4894-x24gv Total loading time: 0 Render date: 2024-05-08T15:20:19.720Z Has data issue: false hasContentIssue false

Pattern Formation on Silicon-on-Insulator

Published online by Cambridge University Press:  01 February 2011

Frank S. Flack
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Bin Yang
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Minghuang Huang
Affiliation:
University of Utah SaltLake City, UT 84112, U. S. A.
Matt Marcus
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Jason Simmons
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Olivia M. Castellini
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Mark A. Eriksson
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Feng Liu
Affiliation:
University of Utah SaltLake City, UT 84112, U. S. A.
Max G. Lagally
Affiliation:
Materials Research Science and Engineering Center University of Wisconsin Madison, WI 53706, U. S. A.
Get access

Abstract

The strain driven self-assembly of faceted Ge nanocrystals during epitaxy on Si(001) to form quantum dots (QDs) is by now well known. We have also recently provided an understanding of the thermodynamic driving force for directed assembly of QDs on bulk Si (extendable to other QD systems) based on local chemical potential and curvature of the surface. Silicon-on-insulator (SOI) produces unique new phenomena. The essential thermodynamic instability of the very thin crystalline layer (called the template layer) resting on an oxide can cause this layer, under appropriate conditions, to dewet, agglomerate, and self-organize into an array of Si nanocrystals. Using low-energy electron microscopy (LEEM), we observe this process and, with the help of first-principles total-energy calculations, we provide a quantitative understanding of this pattern formation. The Si nanocrystal pattern formation can be controlled by lithographic patterning of the SOI prior to the dewetting process. The resulting patterns of electrically isolated Si nanocrystals can in turn be used as a template for growth of nanostructures, such as carbon nanotubes (CNTs). Finally we show that this growth may be controlled by the flow dynamics of the feed gas across the substrate.

Type
Research Article
Copyright
Copyright © Materials Research Society 2005

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

REFERENCES

1. Kane, W.M., Spratt, J.P., and Hershinger, L.W., J. Appl. Phys. 37, 20852089 (1966).Google Scholar
2. Scharnhorst, P., Surf. Sci. 15, 380–6 (1969).Google Scholar
3. Hummel, R.E., DeHoff, R.T., Matts-Goho, S., and Goho, W.M., Thin Solid Films 78, 114 (1981).Google Scholar
4. Kwon, J.-Y., Yoon, T.-S., Kim, K.-B., and Min, S.-H., J. Appl. Phys. 93, 3270–8 (2003).Google Scholar
5. Srolovitz, D.J., Yang, W., and Goldiner, M.G. in Polycrystalline Thin Films: Structure, Texture, Properties, and Applications II, edited by Frost, H.J., Parker, M.A., Ross, C.A., and Holm, E.A., (Mater. Res. Soc. Symp. Proc. 403, Pittsburgh, Pa, 1996), pp. 313.Google Scholar
6. Woll, A.R., Moran, P., Rehder, E.M., Yang, B., Kuech, T.F., and Lagally, M.G. in Current Issues in Heteroepitaxial Growth - Stress Relaxation and Self Assembly, edited by Stach, E., Chason, E., Hull, R., and Bader, S., (Mater. Res. Soc. Symp. Proc. 696, Pittsburgh, Pa, 2002), pp. 119–24.Google Scholar
7. Yamamoto, S., Masuda, S., Yasufuku, H., Ueno, N., Harada, Y., Ichinokawa, T., Kato, M., and Sakai, Y., J. Appl. Phys. 82, 29542960 (1997).Google Scholar
8. Ding, Y., Yamamuro, S., Farrell, D., and Majetich, S.A., J. Appl. Phys. 93, 74117413 (2003).Google Scholar
9. Yamagata, K. and Yonehara, T., Appl. Phys. Lett. 61, 2557–9 (1992).Google Scholar
10. Ono, Y., Nagase, M., Tabe, M., and Takahashi, Y., Jpn. J. Appl. Phys, Pt 1 34, 1728–35 (1995).Google Scholar
11. Sugiyama, N., Tezuka, T., and Kurobe, A., J. Crys. Growth 192, 395401 (1998).Google Scholar
12. Legrand, B., Agache, V., Melin, T., Nys, J.P., Senez, V., and Stievenard, D., J. Appl. Phys. 91, 106–11 (2002).Google Scholar
13. Nuryadi, R., Ishikawa, Y., and Tabe, M., Appl. Surf. Sci. 159–160, 121–6 (2000).Google Scholar
14. Ishikawa, Y., Imai, Y., Ikeda, H., and Tabe, M., Appl. Phys. Lett. 83, 3162–4 (2003).Google Scholar
15. Eaglesham, D.J., White, A.E., Feldman, L.C., Moriya, N., and Jacobson, D.C., Phys. Rev. Lett. 70, 1643–6 (1993).Google Scholar
16. Bermond, J.M., Metois, J.J., Egea, X., and Floret, F., Suf. Sci. 330, 4860 (1995).Google Scholar
17. Brommer, K.D., Needels, M., Larson, B.E., and Joannopoulos, J.D., Phys. Rev. Lett. 68, 1355–8 (1992).Google Scholar
18. Dabrowski, J., Mussig, H.-J., and Wolff, G., Phys. Rev. Lett. 73, 1660–3 (1994).Google Scholar
19. Laracuente, A., Erwin, S.C., and Whitman, L.J., Phys. Rev. Lett. 81, 5177–80 (1998).Google Scholar
20. Yang, B., Zhang, P., Lagally, M.G., Lu, G.-H., Huang, M., and Liu, F., Submitted to Physical Review Letters.Google Scholar
21. Rayleigh, L., Proc. London Math. Soc. 10, 4 (1878).Google Scholar
22. Liu, F., Wu, F., and Lagally, M.G., Chem. Rev. 97, 10451061 (1997).Google Scholar
23. Wu, F. and Lagally, M.G., Phys. Rev. Lett., 75, 2534 (1995).Google Scholar
24. Franklin, N.R. and Dai, H., Adv. Mat. 12, 890894 (2000).Google Scholar
25. Kitajima, T., Liu, B., and Leone, S.R., Appl. Phys. Lett. 80, 497 (2002).Google Scholar
26. Homma, Y., Kobayashi, Y., Ogino, T., and Yamashita, T., Appl. Phys. Lett. 81, 2261 (2002).Google Scholar
27. Jung, Y.J., Homma, Y., Ogino, T., Kobayashi, Y., Takagi, D., Wei, B., Vajtai, R., and Ajayan, P.M., J. Phys. Chem. B 107, 68596864 (2003).Google Scholar
28. At atmospheric pressure, the flow velocity ν of the methane gas is calculated to be 0.33 cm/s (the methane mass flow rate is 400 sccm and the diameter d of the furnace tube is 2 inches). We estimate that the Reynolds number (Re) of methane flow in the furnace tube is smaller than 10 using the formula Re=ρνd/η, where ρ is the density of methane gas (ρ=178g/m3 at 900°C and 1 atm), and η is the gas viscosity (η=1.8×10-5 kg/m-s). Also, the mean free path, λ, of methane molecules is very small compared to the tube diameter during growth (λ=5×10-3 cm for N2 at room temperature and 1 Torr). The Knudsen number is much smaller than 0.01. Thus, the gas flow can be modeled by continuum theory.Google Scholar
29. Hu, Y., Werner, C., and Li, D., J. Fluids Eng. 125, 871879 (2003).Google Scholar