Thermal Behavior of Hydroxymethylated Resorcinol (HMR)-Treated Maple Veneer

Authors

  • Jungil Son
  • William T. Y. Tze
  • Douglas J. Gardner

Keywords:

Hydroxymethylated resorcinol, dynamic mechanical thermal analysis, storage modulus, loss modulus, damping ratio, glass-transition temperature, drying time, solubility parameter, differential scanning calorimetry

Abstract

The objective of this research was to study the effect of hydroxymethylated resorcinol (HMR) treatment on the thermal and dynamic mechanical properties of maple veneer. The veneers were soaked in HMR solution for either 1, 15, or 30 min, and subsequently dried for either 1, 12, or 24 h at 20°C and 65% relative humidity. Dynamic mechanical thermal analysis (DMTA) tests were performed at a controlled heating rate of 5°C/min using a 3-point bending mode at an oscillatory frequency of 1Hz and an oscillating dynamic strain of 0.01%. Differential scanning calorimetry (DSC) was performed from -40 to 150°C at a heating rate of 10°C/min. Depending on the amount of drying, the storage moduli of wood can be unaltered or reduced as a response to HMR soaking time. Overall, there was no evidence that HMR treatments reinforce wood.

The lignin glass-transition temperature of HMR-treated maple veneer decreased with an increase in treatment time. The lowering of Tg by HMR treatments was confirmed by DSC results. Both DMTA and DSC data showed a glass-transition shift of wood hemicellulose that was subtle or none in responding to HMR treatments. HMR was theoretically determined to have a closer solubility parameter match (better compatibility) with lignin compared to the other wood cell-wall polymers (i.e., cellulose and hemicellulose). Based on these findings, HMR is postulated to act as a lignin plasticizer.

This study provides new insights into the interactions of HMR with wood and is expected to stimulate further investigations that lead to a better understanding of the wood bond durability enhancement of HMR treatment.

References

Backman, A. C., and K. A. H. Lindberg. 2004. Interaction between wood and polyvinyl acetate glue studied with dynamic mechanical analysis and scanning electron microscopy. J. Appl. Polymer Sci. 91:3009-3015.nFuruno, T., Y. Imamura, and H. Kajita. 2004. The modification of wood by treatment with low molecular weight phenol-formaldehyde resin: properties enhancement with neutralized phenolic-resin and resin penetration into wood cell walls. Wood Sci. Technol. 37:349-361.nHansen, C. M., and A. Björkman. 1998. The ultrastructure of wood from a solubility parameter point of view. Holzforschung 52:335-344.nHaygreen, J. G., and J. L. Bowyer. 1989. Forest Products and Wood Science—An Introduction. 2nd ed. Iowa State University Press, Ames, IA. 500 pp.nHensley, J., R. Lopez-Anido, and D. J. Gardner. 2000. Adhesive bonding of wood with vinyl ester resin. Pages 136-138 in G. L. Anderson, ed. Proc. 23rd Annual Meeting of the Adhesion Society. G. L. Anderson, eds. The Adhesion Society, Blacksburg, VA.nHoy, K. L. 1985. The Hoy tables of solubility parameters. Solvent and Coatings Materials Research and Development Department, Union Carbide Corporation, South Charleston, WV. 144 pp.nIrvine, G. M. 1984. The glass transitions of lignin and hemicellulose and their measurement by differential thermal analysis. Tappi J. 67(5):118-121.nKelley, S. S., T. G. Rials, and W. G. Glasser. 1987. Relaxation behaviour of the amorphous components of wood. J. Mat. Sci. 22:617-624.nKobayashi, S., A. Nakai, and H. Hamada. 2003. Suppression of microscopic damage in epoxy-based composites using flexible interphase. J. Mat. Sci. Letters 22: 1315-1318.nMarcinko, J. J., S. Devathala, P. L. Rinaldi, and S. Bao. 1998. Investigating the molecular and bulk dynamics of PMDI/Wood and UF/Wood composites. Forest Prod. J. 48(6):81-84.nOksman, K., and H. Lindberg. 1995. Interaction between wood and synthetic polymers. Holzforschung 49:249-254.nPark, B., B. Riedl, E. W. Hsu, and J. Shields. 1999. Differential scanning calorimetry of phenol-formaldehyde resins cure-accelerated by carbonates. Polymer 40: 1689-1699.nRodriguez, F. 1982. Principles of polymer systems. Hemisphere Publishing, New York, NY. 575 pp.nSears, J. K., and and J. R. Darby. 1982. The technology of plasticizers. John Wiley and Sons, New York, NY.nSon, J., and D. J. Gardner. 2004. Dimensional stability measurements of thin wood veneers using the Wilhelmy plate technique. Wood Fiber Sci. 36(1):98-106.nSuzuki, N., and H. Ishida. 2003. Effect of the miscibility between a silane and a sizing agent on silanol condensation. J. Appl. Polymer Sci. 87:589-598.nVick, C. B. 1995. Coupling agent improves durability of PRF bonds to CCA-treated southern pine. Forest Prod. J. 45(3):78-84.nVick, C. B. 1996. Hydroxymethylated resorcinol coupling agent for enhanced adhesion to wood. Pages 47-55, in A. W. Christiansen, and A. H. Conner, eds. Proc. of Wood Adhesives 1995; June 29-30, 1995; Portland, OR. Forest Prod. Soc., Madison, WI.nVick, C. B., K. Richter, B. H. River, and A. R. Fried. 1995. Hydroxymethylated resorcinol coupling agent for enhanced durability of bisphenol-A epoxy bonds to Sitka spruce. Wood Fiber Sci. 27(1):2-12.nVick, C. B., A. W. Christiansen, and E. A. Okkonen. 1998. Reactivity of hydroxyl-methylated resorcinol coupling agent as it affects durability of epoxy bonds to Douglas-fir. Wood Fiber Sci. 30(3):312-322.nVick, C. B., K. Richter, and B. H. River. 1994. Hydroxymethylated resorcinol coupling agent and method for bonding wood. U. S. Patent No. 5,543,487, January 19, 1994.n

Downloads

Published

2007-06-05

Issue

Section

Research Contributions