Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-05-14T14:28:01.230Z Has data issue: false hasContentIssue false

Adsorption of Tyrosinase Onto Montmorillonite as Influenced by Hydroxyaluminum Coatings

Published online by Cambridge University Press:  28 February 2024

A. Naidja
Affiliation:
Department of Soil Science, University of Saskatchewan, 51 Campus Drive, Saskatoon, Saskatchewan S7N 5A8 Canada
A. Violante
Affiliation:
Dipartimento di Scienze Chimico-Agrarie, Università di Napoli, “Federico II,” Via Università 100, 80055 Portici, Napoli, Italy
P. M. Huang
Affiliation:
Department of Soil Science, University of Saskatchewan, 51 Campus Drive, Saskatoon, Saskatchewan S7N 5A8 Canada

Abstract

In soil environments, the surfaces of clay minerals are often coated with hydrolytic products of Al. However, limited information is available on the effect of hydroxyaluminum coatings on the interlayering of enzymes for montmorillonite. The objective of this study was to compare the adsorption of tyrosinase onto montmorillonite as influenced by levels of hydroxyaluminum coatings. Tyrosinase is one of the strongest catalysts in the transformation of phenolic compounds. Adsorption of tyrosinase onto Ca-montmorillonite (Ca-Mte) and different hydroxyaluminum-montmorillomte complexes (Al(OH)x-Mte), containing 1.0, 2.5 and 5.0 mmol coated Al/g clay, was studied both in the absence and in the presence of a phosphate buffer at pH 6.5 and 25°C. Except for Ca-Mte in the absence of phosphate where the adsorption isotherm was of C type (linear), the adsorption isotherms were of L type (Langmuir). More tyrosinase molecules were adsorbed onto Ca-Mte than onto the Al(OH)x-Mte complexes, both in the absence and in the presence of phosphate. This indicated the easy accessibility of the enzyme to the uncoated Ca-Mte surfaces. The presence of phosphate did not significantly affect the amount of tyrosinase adsorbed onto Ca-Mte, but substantially reduced the adsorption of tyrosinase onto Al(OH)x-Mte complexes. The higher the level of hydroxyaluminum coatings, the lower the amount of tyrosinase was adsorbed. Because of their affinity to the aluminous surfaces, phosphate ions evidently competed strongly with tyrosinase for Al(OH)x-Mte complexes adsorption sites. The intercalation of tyrosinase by Ca-Mte was indicated by the increased d-spacing of the complex as the amount of the enzyme adsorbed increased. The infrared spectra of tyrosinase-Ca-Mte complex showed that the amide II band of tyrosinase at 1540 cm-1 was practically unaffected by adsorption. The amide I band at 1654 cm-1 was shifted toward a higher frequency, indicating a slight perturbation in the protein conformation. This perturbation became more noticeable in the presence of Al(OH)x-Mte complexes. The data indicated that hydroxyaluminum coatings play an important role in retarding the adsorption of tyrosinase by montmorillonite, and phosphate effectively competes with tyrosinase for the adsorption sites on Al(OH)x-Mte complexes.

Type
Research Article
Copyright
Copyright © 1995, The Clay Minerals Society

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

Alexiades, C. A., and Jackson, M. L. 1965. Quantitative determination of vermiculite. Soil Sci. Soc. Amer. J. 29: 522529.Google Scholar
Alikhan, M. A., 1976. The tyrosinase system in the terrestrial isopod, Porcellio laevis Latreille (Porcellionidae, Isopoda). Comp. Biochem. Physiol. 54B: 3742.Google Scholar
Barnhisel, R. I., and Bertsch, P. M. 1989. Chlorites and hydroxy-interlayered vermiculite and smectite. In Minerals in Soil Environments, 2nd ed. Dixon, J. B., and Weed, S. B., eds. Madison, Wisconsin: Soil Science Society of America, 729788.Google Scholar
Boyd, S. A., and Mortland, M. M. 1990. Enzyme interactions with clays and clay-organic matter complexes. In Soil Biochemistry. 6, Bollag, J. M., and Stotzky, G., eds. New York: Marcel Dekker, 128.Google Scholar
Brunauer, S., Emmett, P. H., and Teller, E. 1938. Adsorption of gases in multimolecular layers. J. Amer. Chem. Soc. 60: 309319.CrossRefGoogle Scholar
Burns, R. G., 1978. Enzyme activity in soil: Some theoretical and practical considerations. In Soil Enzymes. Burns, R. G., ed. London: Academic Press, 295399.Google Scholar
Burns, R. G., 1990. Microorganisms, enzymes and soil colloids surfaces. In Soil Colloids and their Associations in Aggregates. De Boot, M. F., Hayes, M. H. B., and Her-billon, A., eds. New York: Plenum Press, 337361.Google Scholar
Chen, J. S., Wei, C., Rolle, R. S., Otwell, W. S., Balaban, M. O., and Marshall, M. R. 1991. Inhibitory effect of kojic acid on some plant and crustacean polyphenoloxidases. J. Agric. Food Chem. 39: 13961401.Google Scholar
Elliot, A., 1969. Infra-red Spectra and Structure of Organic Long-chain Polymers. London: Edward Arnold Ltd.Google Scholar
Eltantawy, I. M., and Arnold, P. W. 1973. Reappraisal of ethylene glycol mono-ethyl ether (EGME) method for surface area estimations of clays. J. Soil Sci. 24: 232238.CrossRefGoogle Scholar
Ferreiro, E. A., de Busseti, S. G., and Helmy, A. K. 1992. Effect of montmorillonite on phosphate sorption by hydrous Al-oxides. Geoderma. 55: 111118.Google Scholar
Fusi, P., Ristori, L., Calamai, L., and Stotzky, G. 1989. Adsorption and binding of protein on “Clean” (homoionic) and “Dirty” (coated with Fe oxyhydroxides) montmorillonite, illite and kaolinite. Soil Biol. Biochem. 21: 911920.Google Scholar
Garwood, G. A., Mortland, M. M., and Pinnavaia, T. J. 1983. Immobilization of glucose oxidase on montmorillonite clay and ionic modes of binding. J. Mol. Cat. 22: 153163.Google Scholar
Gianfreda, L., Rao, M. A., and Violante, A. 1991. Invertase (β-fructosidase): Effect of montmorillonite, Al-hydroxide and Al(OH)x-montmorillonite complex on activity and kinetic properties. Soil Biol. Biochem. 23: 581587.Google Scholar
Gianfreda, L., Rao, M. A., and Violante, A. 1992. Adsorption, activity and kinetic properties of urease on montmorillonite, aluminum hydroxide and Al(OH)x-montmorillonite complexes. Soil Biol. Biochem. 24: 5158.Google Scholar
Giles, C. H., McEwan, T. H., Nakhwa, S. N., and Smith, D. 1960. Studies in adsorption. XI. A system with classification of solution adsorption isotherms, and its use in diagnosis of adsorption mechanisms and in measurement of specific surface area of solids. J. Chem. Soc. 786: 39733993.CrossRefGoogle Scholar
Hsu, P. H., 1992. Reaction of OH-Al polymers with smectites and vermiculites. Clays & Clay Miner. 40: 300305.CrossRefGoogle Scholar
Janovitz-Klapp, A. H., Richard, F. C., Goupy, P. M., and Nicolas, J. J. 1990. Inhibition studies on apple polyphenol oxidase. J. Agric. Food. Chem. 38: 926931.Google Scholar
Kiss, S., Dragan-Bularda, M., and Radulescu, D. 1975. Biological significance of enzymes accumulated in soil. Adv. in Agron. 27: 2587.Google Scholar
Lowell, S., and Shields, J. E. 1991. Powder Surface Area and Porosity. 3rd ed. New York, NY: Chapman and Hall.Google Scholar
Mantsh, H. H., Casal, H. L., and Jones, R. N. 1986. Resolution enhancement of infrared spectra of biological systems. In Advances in Spectroscopy 13. Clark, R. J. H., and Hester, R. E., eds. New York: Wiley and Sons, 146.Google Scholar
McLaren, A. D., Peterson, G. H., and Barshad, I. 1958. The adsorption and reactions of enzymes and proteins on clay minerals: IV. Kaolinite and Montmorillonite. Soil Sci. Soc. Am. Proc. 22: 239244.Google Scholar
Morgan, H. W., and Corke, C. T. 1976. Adsorption, desorption, and activity of glucose oxidase on selected clay species. Can. J. Microbiol. 22: 684693.Google Scholar
Naidja, A., and Huang, P. M. 1994. Aspartic acid interaction with Ca-montmorillonite: Adsorption, desorption and thermal stability. Appl. Clay Sci. 9: 265281.Google Scholar
Quiquampoix, H., 1987. A stepwise approach to the understanding of extracellular enzyme activity in soil I. Effect of electrostatic interactions on the conformation of a β-D glucosidase adsorbed on different mineral surfaces. Biochimie 69: 753763.CrossRefGoogle Scholar
Sepelyak, R. J., Feldkamo, J. R., Moody, T. E., White, J. L., and Hem, S. 1984. Adsorption of pepsin by aluminum hydroxide I: Adsorption mechanism. J. Pharmaceutical Sci. 73: 15141517.Google Scholar
Sjoblad, R. D., and Bollag, J. M. 1981. Oxidative coupling of aromatic compounds by enzymes from soil microorganisms. In Soil Biochemistry. 5, Paul, E. A., and Ladd, J. N., eds. New York: Marcel Dekker, 113152.Google Scholar
Skujins, J., 1978. History of abiontic soil enzyme research. In Soil Enzymes. Burns, R. G., ed. London: Academic Press, 149.Google Scholar
Susi, H., Timasheff, S. N., and Stevens, L. 1967. Infrared spectra and protein conformation in aqueous solutions. The amide I band in H2O and D2O solutions. J. Biol. Chem. 242: 54605466.Google Scholar
Susi, H., and Byler, M. 1983. Protein structure by Fourier Transform infrared spectroscopy: Second derivative spectra. Biochem. Biophys. Res. Comm. 115: 391397.Google Scholar
Talibudeen, O., 1955. Complex formation between mont-morillonoid clays and amino-acids and proteins. Trans. Farad. Soc. 51: 582590.Google Scholar
Timasheff, S. N., Susi, H., and Stevens, L. 1967. Infrared spectra and protein conformation in aqueous solutions. II. Survey of globular proteins. J. Biol. Chem. 242: 54675473.Google Scholar
Tu, A. T., 1986. Peptide backbone conformation and microenvironment of protein side chains. In Advances in Spectroscopy. 13, Clark, R. J. H., and Hester, R. E., eds. New York: Wiley, 146.Google Scholar
Violante, A., Colombo, C., and Buondonno, A. 1991. Competitive adsorption of phosphate and oxalate by aluminum oxides. Soil Sci. Soc. Am. J. 55: 6570.Google Scholar
Violante, A., Dipartmento di Scienze Chimico-Agrarie, Università di Napoli “Frederico II,” Via Università 100, 80055 Portici, Napoli, Italy.Google Scholar
Weast, R. C., 1974. Handbook of Chemistry and Physics. 55th ed. Cleveland, Ohio: CRC Press.Google Scholar