Hostname: page-component-7479d7b7d-c9gpj Total loading time: 0 Render date: 2024-07-13T05:03:35.737Z Has data issue: false hasContentIssue false

X-ray photoelectron spectroscopy of uv laser irradiated sapphire and alumina

Published online by Cambridge University Press:  03 March 2011

A.J. Pedraza
Affiliation:
Department of Materials Science and Engineering, The University of Tennessee, Knoxville, Tennessee 37996-2200
J.W. Park
Affiliation:
Department of Materials Science and Engineering, The University of Tennessee, Knoxville, Tennessee 37996-2200
H.M. Meyer III
Affiliation:
Martin Marietta Energy Systems, P.O. Box 2009, Oak Ridge, Tennessee 37831
D.N. Braski
Affiliation:
Oak Ridge National Laboratory, P.O. Box 2008, Oak Ridge, Tennessee 37831-6056
Get access

Abstract

X-ray photoelectron spectroscopy (XPS) was performed in as-received, thermally annealed, and laser-irradiated sapphire and alumina specimens in order to study the effects of the different treatments on surface chemistry and properties. Laser irradiations with a 308 nm wavelength laser were performed in air and in a reducing atmosphere consisting of a mixture of Ar and 4% of hydrogen. The atomic percentages of carbon, aluminum, and oxygen were measured in all the specimens. Particular attention was paid to the percentages of oxygen in the oxide and in a hydroxyl state. The XPS analyses clearly established that a very thin film of metallic aluminum is formed on the surface of both alumina and sapphire substrates when they are irradiated under a reducing atmosphere. However, the film is discontinuous because it is electrically insulating. Substrates irradiated in air have metallic aluminum only for fluences below 0.4 J/cm2. The valence band photoemission spectra of as-received, annealed, and laser-irradiated specimens were measured. In irradiated specimens, the width of the valence band spectra was found to decrease by ∼10%. One possible cause of this decrease is the generation of point defects during laser irradiation. Electroless copper deposition occurs on sapphire and alumina substrates if their surface has been activated by laser irradiation. The time required for copper deposition was monitored by measuring the electrical resistivity in the irradiated area while the substrates were immersed in an electroless bath. The kinetics of deposition on laser-activated substrates and the XPS results show that the presence of metallic aluminum accelerates the deposition process. However, the presence of aluminum is not the sole reason for laser activation in alumina. Very strong metal-ceramic bonding is produced after thermal annealing of samples having preirradiated substrates. This result is explained in terms of the excess oxygen that is present at the ceramic surface after irradiation.

Type
Articles
Copyright
Copyright © Materials Research Society 1994

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

1Esrom, H., in Chemical Perspectives of Microelectronic Materials II, edited by Interrante, L. V., Jensen, K. F., Dubois, L. H., and Gross, M. E. (Mater. Res. Soc. Symp. Proc. 204, Pittsburgh, PA, 1991), p. 457.Google Scholar
2Esrom, H., Zhang, J-Y., and Pedraza, A. J., in Photons and Low Energy Particles in Surface Processing, edited by Ashby, C. I. H., Brannon, J. H., and Pang, S. W. (Matet. Res. Soc. Symp. Proc. 236, Pittsburgh, PA, 1992), p. 383.Google Scholar
3Pedraza, A. J., Godbole, M. J., DeSilva, M. J., and Lowndes, D. H., in Laser Ablation in Materials Processing: Fundamentals and Applications, edited by Braren, B., Dubowski, J. J., and Norton, D. P. (Mater. Res. Soc. Symp. Proc. 285, Pittsburgh, PA, 1993), pp. 203-208.Google Scholar
4Pedraza, A. J., DeSilva, M. J., and Lowndes, D. H., unpublished.Google Scholar
5DeSilva, M.J., Pedraza, A. J., and Lowndes, D. H., J. Mater. Res. 9, 1019 (1994).CrossRefGoogle Scholar
6Chen, C. H. and McCann, M.P., Opt. Commun. 60, 296 (1986).CrossRefGoogle Scholar
7Ihlemann, J., Wolff, B., and Simmon, P., Appl. Phys. A 54, 363 (1992).CrossRefGoogle Scholar
8Webb, R. L., Jensen, C., Langford, S. C., and Dickinson, J. T., J. Appl. Phys. 74 (4), 2323 (1993).CrossRefGoogle Scholar
9Lowndes, D. H., DeSilva, M.J., Godbole, M. J., Pedraza, A. J., and Geohegan, D. B., in Laser Ablation in Materials Processing: Fundamentals and Applications, edited by Braren, B., Dubowski, J. J., and Norton, D. P. (Mater. Res. Soc. Symp. Proc. 285, Pittsburgh, PA, 1993), p. 191.Google Scholar
10Ohuchi, F. S., French, R. H., and Kasowsky, R. V., J. Appl. Phys. 62 (6), 2286 (1987).CrossRefGoogle Scholar
11Ohuchi, F. S. and Kohyama, M., J. Am. Ceram. Soc. 74 (6), 1163 (1991).CrossRefGoogle Scholar
12Pedraza, A. J., Park, J. W., DeSilva, M.J., Lowndes, D. H., Braski, D. N., and MeyeT, H.M. III, AIP Conf. Proc. 288, 329 (1993).CrossRefGoogle Scholar
13Dreyfus, R. W., Kelly, R., and Walkup, R. E., Appl. Phys. Lett. 49 (21), 1478 (1986).CrossRefGoogle Scholar
14French, R. H., Coble, R. L., Kasowsky, R. V., and Ohuchi, F. S., Physica B 150, 47 (1988).CrossRefGoogle Scholar