Hostname: page-component-7479d7b7d-8zxtt Total loading time: 0 Render date: 2024-07-11T12:18:01.988Z Has data issue: false hasContentIssue false

Hectorite Complexes with Cu(II) and Fe(II)-1,10-Phenanthroline Chelates

Published online by Cambridge University Press:  01 July 2024

V. E. Berkheiser
Affiliation:
Dept. of Crop and Soil Sciences, Michigan State University, East Lansing, MI 48824, U.S.A.
M. M. Mortland
Affiliation:
Dept. of Crop and Soil Sciences, Michigan State University, East Lansing, MI 48824, U.S.A.

Abstract

Characteristics and properties of complexes of a smectite (hectorite) with 1,10-phenanthroline (phen) chelates with iron or copper were determined by a variety of physical and chemical measurements. The complex ions showed high selectivity for the hectorite surface. Basal spacings of 17.4 Å were produced by Fe(II) or Cu(II) analogues of M(phen)32+ hectorite. Adsorption of gases and vapors by the M(phen)32+ hectorite complex revealed large surface areas and reflected intrinsic characteristics of the complex ions. Lower surface areas were found for copper phen hectorite than iron phen hectorite probably because of the loss of a ligand from the Cu(II) ion. ESR spectra confirmed that appreciable Cu(II) existed as the bis-phen complex under certain conditions. An increase in the oxidation potential of the Fe(phen)32+-Fe(phen)33+ couple above that in pure solvent was noted when these complexes were supported by the mineral surface.

Type
Research Article
Copyright
Copyright © Clay Minerals Society 1977

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.)

Footnotes

*

Journal Article No. 7817. Michigan Agricultural Experiment Station. Work partially supported by National Science Foundation Grant No. MP S74-18201.

References

Adamson, A. W. (1967) Physical Chemistry of Surfaces, pp. 585589: Interscience, New York.Google Scholar
Allen, H. C., Kokoszka, G. F. and Inskeep, R. G. (1964) The electron paramagnetic resonance spectrum of some tris-complexes of Cu(II): J. Am. Chem. Soc. 86, 10231025.CrossRefGoogle Scholar
Berkheiser, V. and Mortland, M. M. (1975) Variability in exchange ion position in smectite: dependence on interlayer solvent: Clays & Clay Minerals 23, 404410.CrossRefGoogle Scholar
Blau, F. (1898) Uber neue organische metallverbindungen: Monatsh. 19, 647689.CrossRefGoogle Scholar
Bower, C. A. (1962) Adsorption of o-phenanthroline by clay minerals and soils: Soil Sci. 93, 192195.Google Scholar
Burchett, S. and Meloan, C. E. (1972) Infrared studies of water bound to some extracted phenanthroline and phenanthroline chelates: J. Inorg. Nucl. Chem. 14, 12071213.CrossRefGoogle Scholar
Clementz, D. M. and Mortland, M. M. (1974) Properties of reduced charge montmorillonite: tetra-alkylammonium ion exchange forms: Clays & Clay Minerals 22, 223229.CrossRefGoogle Scholar
Farmer, V. C. and Russell, J. D. (1967) Infrared absorption spectrometry in clay studies: Clays & Clay Minerals 15, 121142.CrossRefGoogle Scholar
Gast, R. G. and Mortland, M. M. (1971) Self-diffusion of alkylammonium ions in montmorillonite: J. Colloid Interface Sci. 37, 8092.CrossRefGoogle Scholar
Greenland, D. J. and Quirk, J. P. (1962) Adsorption of 1-n-alkyl pyridinium bromides by montmorillonite: Clays & Clay Minerals 9, 484499.CrossRefGoogle Scholar
Hall, J. R., Marchant, N. K. and Plowman, R. A. (1963) Coordination compounds of substituted, 1,10-phenanthrolines and related dipyridyls: Aust. J. Chem. 16, 3441.CrossRefGoogle Scholar
Hathaway, B. J., Hodgson, P. G. and Power, P. C. (1974) Single-crystal electronic and electron spin resonance spectra of three tris-chelate copper(II) complexes: Inorg. Chem. 13, 20092013.CrossRefGoogle Scholar
Hume, D. N. and Kolthoff, I. M. (1943) A revision of the oxidation potentials of the orthophenanthroline– and dipyridyl–ferrous complexes: J. Am. Chem. Soc. 65, 18951897.CrossRefGoogle Scholar
Inskeep, R. G. (1962) Infrared spectra of metal complexes below 600 cm–1: the spectra of the tris complexes of 1,10-phenanthroline and 2,2′-bipyridine with the transition metals iron(II) through zino(II): J. Inorg. Nucl. Chem. 24, 763776.CrossRefGoogle Scholar
James, B. R. and Williams, R. J. P. (1961) The oxidation–reduction potentials of some copper complexes: J. Chem. Soc. 20072019.CrossRefGoogle Scholar
Jensen, A., Basolo, F. and Neumann, H. M. (1958) Mechanism of racemization of complex ions. IV. Effect of added large ions upon the rates of dissociation and racemization of tris-(1,10-phenanthroline)-iron(II) ion: J. Am. Chem. Soc. 80, 23542358.CrossRefGoogle Scholar
Lagaly, G. and Weiss, A. (1975) The layer charge of smectite layer silicates: In Proc. International Clay Conf. (1975) (edited by Bailey, S. W.) , pp. 157172. Applied Publishing Ltd., Wilmette, IL.Google Scholar
Lawrie, D. C. (1961) A rapid method for the determination of approximate surface areas of clays: Soil Sci. 92, 188191.CrossRefGoogle Scholar
Mortland, M. M. and Berkheiser, V. E. (1976) Triethylene–diamine–clay complexes as matrices for adsorption and catalytic reactions: Clays & Clay Minerals 24, 6063.CrossRefGoogle Scholar
Schilt, A. A. (1969) Analytical Applications of 1,10-phenanthroline and Related Compounds: Pergamon Press, Oxford.Google Scholar
Schilt, A. A. and Taylor, R. C. (1959) Infrared spectra of 1,10-phenanthroline metal complexes in the rock salt region below 2000 cm–1: J. Inorg. Nucl. Chem. 9, 211221.CrossRefGoogle Scholar