Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-06-22T13:04:19.607Z Has data issue: false hasContentIssue false

Precursor and processing effects on BaPbO3 formation kinetics

Published online by Cambridge University Press:  01 March 2006

Jun-ichi Tani
Affiliation:
Department of Inorganic Chemistry, Osaka Municipal Technical Research Institute, Joto-ku, Osaka 536-8553, Japan
Guerman Popov
Affiliation:
Department of Materials Science and Engineering, Rutgers, The State University of New Jersey, Piscataway, New Jersey 08854
Paul R. Mort
Affiliation:
Procter & Gamble Co., Cincinnati, Ohio 45217
Richard E. Riman*
Affiliation:
Department of Materials Science and Engineering, Rutgers, The State University of New Jersey, Piscataway, New Jersey 08854
*
a) Address all correspondence to this author. e-mail: riman@rci.rutgers.edu
Get access

Abstract

The synthesis of BaPbO3 from a wide range of mixtures containing metalorganic precursors, nitrate precursors, lead oxides, barium oxide and peroxide was investigated, and the kinetics was analyzed using the Johnson–Mehl–Avrami (JMA) equation. It was found that Ba and Pb stearate soaps and Pb oxalate that were used as metalorganic precursors formed BaCO3 and PbO or Pb3O4 after firing at 440 °C. The formation rate of BaPbO3 from a metalorganic precursor system is not higher than that from the conventional BaCO3–PbO system and does not depend on mixing methods or the kinds of metalorganic precursors but instead on the synthesis atmosphere. In the case of the BaCO3–PbO system, the Avrami exponent (n) is ∼1, indicating that the reaction is controlled by the phase-boundary-contraction interface reaction. For the BaO2–PbO2 system, n has two values ∼1 and ∼0.3, depending on the reaction temperature and time, indicating that the reaction is either controlled by the phase-boundary-contraction interface reaction or diffusion-controlled reaction. In the Ba nitrate–Pb nitrate system, phase-pure BaPbO3 is obtained at 550 °C, which is 250 °C lower than in the case of the BaCO3–PbO system. The value of n for the nitrate system is ∼1.5, indicating that the reaction is controlled by a three-dimensional (3D) diffusion-controlled nucleation mechanism. In the BaO–PbO system, the formation of BaPbO3 started at 350 °C by an exothermic reaction and the content of BaPbO3 in the product was ∼40 wt%, which is independent of reaction temperature as well as time in the temperature range of 350–500 °C.

Type
Articles
Copyright
Copyright © Materials Research Society 2006

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.Yasukawa, M., Murayama, N.: A promising oxide material for high-temperature thermoelectric energy conversion: Ba1− x Srx PbO3 solid solution system. Mater. Sci. Eng. B 54, 64 (1998).Google Scholar
2.Sleight, A.W., Gillson, J.L., Bierstedt, P.E.: High-temperature superconductivity in the BaPb1−xBixO3 systems. Solid State Commun. 17, 27 (1975).Google Scholar
3.Cava, R.J., Batlogg, B., Espinosa, G.P., Ramirez, A.P., Krajewski, J.J.Peck, W.F. Jr.Rupp, L.W. Jr.Cooper, A.S.: Superconductivity at 3.5 K in BaPb0.75Sb0.25O3: Why is Tc so low? Nature 339, 291 (1989).CrossRefGoogle Scholar
4.Nitta, T., Nagase, K., Hayakawa, S., Iida, Y.: Formation and properties of barium metaplumbate. J. Am. Ceram. Soc. 48, 642 (1965).CrossRefGoogle Scholar
5.Vidyasagar, K., Gopalakrishnan, J., Rao, C.N.R.: Synthesis of complex metal oxides using hydroxide, cyanide, and nitrate solid solution precursors. J. Solid State Chem. 58, 29 (1985).Google Scholar
6.Yamanaka, A., Maruyama, T., Atake, T., Saito, Y.: Preparation of BaPbO3 from coprecipitated barium-lead oxalate. Thermochemica Acta 115, 207 (1987).CrossRefGoogle Scholar
7.Chang, M.C., Wu, J.M., Cheng, S.Y., Chen, S.Y.: Reaction kinetics and mechanism of BaPbO3 formation. Mater. Chem. Phys. 65, 57 (2000).CrossRefGoogle Scholar
8.Mort, P.R., Riman, R.E.: Determination of homogeneity scale in ordered and partially ordered mixtures. Powder Technol. 82, 93 (1995).Google Scholar
9.Mort, P.R., Riman, R.E.: Reactive multicomponent powder mixtures prepared by microencapsulation: Pb(Mg1/3Nb2/3)O3 synthesis. J. Am. Ceram. Soc. 75, 1581 (1992).Google Scholar
10.Chung, F.H.: Quantitative interpretation of x-ray diffraction patterns of mixtures. I. Matrix-flushing method for quantitative multicomponent analysis. J. Appl. Crystallogr. 7, 519 (1974).CrossRefGoogle Scholar
11.Lide, D.R.: Handbook of Chemistry and Physics, 83th ed. (CCR Press, New York, 2002), pp. 444, 4–65.Google Scholar
12.Ikushima, H., Hayakawa, S.: Electrical properties of BaPbO3 ceramics. Solid State Electron 9, 921 (1966).CrossRefGoogle Scholar
13.Walker, P.L. Jr.Rusinko, F. Jr.Austin, L.G. Gas reactions of carbon, in Advances in Catalysis and Related Subjects, XI, edited by Eley, D.D., Selwood, P.W., and Weizs, P.B. (Academic Press Inc., New York, 1959), pp. 133221.Google Scholar
14.Wilson, O.C. Jr.Riman, R.E.: Morphology control of lead carboxylate powders via anionic substitutional effects. J. Colloid Interface Sci. 167, 358 (1994).Google Scholar
15.Shaikh, A.S., Vest, G.M.: Kinetics of BaTiO3 and PbTiO3 formation from metalorganic precursors. J. Am. Ceram. Soc. 69, 682 (1986).CrossRefGoogle Scholar
16.Wang, X., Zhao, C., Wang, Z., Wu, F., Zhao, M.: Synthesis of BaTiO3 nanocrystals by stearic acid gel method. J. Alloy Compd. 204, 33 (1994).CrossRefGoogle Scholar
17.Barin, I.: Thermochemical Data of Pure Substances. (Weinheim, Germany, New York, 1989).Google Scholar
18.Hancock, J.D., Sharp, J.H.: Method of comparing solid-state kinetic data and its application to the decomposition of kaolinite, brucite, and BaCO3. J. Am. Ceram. Soc. 55, 74 (1972).Google Scholar
19.Avrami, M.: Kinetics of phase change, I: General theory. J. Chem. Phys. 7, 121103 (1939).CrossRefGoogle Scholar
20.Avrami, M.: Kinetics of phase change, II: Transformation-time relations for random distribution of nuclei. J. Chem. Phys. 8, 212 (1940).Google Scholar
21.Avrami, M.: Kinetics of phase change, III: Granulation, phase change, and microstructure. J. Chem. Phys. 9, 177 (1941).Google Scholar
22.Sun, C-L., Wang, H-W., Chang, M-C., Lin, M-S., Chen, S-Y.: Characterization of BaPbO3 and Ba(Pb1−xBix)O3 thin films. Mater. Chem. Phys. 78, 507 (2002).Google Scholar
23.Hulbert, S.F.: Models for solid-state reactions in powdered compacts: A review. Br. Ceram. Soc. J. 6, 11 (1967).Google Scholar